DISCRETIZING MALLIAVIN

CHRISTIAN BENDER AND PETER PARCZEWSKI

Abstract. Suppose B is a Brownian motion and Bn is an approximating sequence of rescaled random walks on the same converging to B pointwise in probability. We provide necessary and sufficient conditions for weak and strong L2-convergence of a discretized Malliavin derivative, a discrete Skorokhod , and discrete analogues of the Clark-Ocone derivative to their continuous counterparts. Moreover, given a sequence (Xn) of random vari- ables which admit a chaos decomposition in terms of discrete multiple Wiener with respect to Bn, we derive necessary and sufficient conditions for strong L2-convergence to a σ(B)- measurable X via convergence of the discrete chaos coefficients of Xn to the continuous chaos coefficients of X. In the special case of binary noise, our results support the known formal analogies between on the Wiener space and Malliavin calculus on the Bernoulli space by rigorous L2-convergence results. MSC: 60H07, 60H05, 60F25 Keywords: Malliavin calculus, strong approximation, stochastic integrals, S-transform, chaos decomposition, invariance principle

1. Introduction

Let B = (Bt)t≥0 be a Brownian motion on a probability space (Ω, F,P ), where the σ-field F is generated by the Brownian motion and completed by null sets. Suppose ξ is a square-integrable random variable with zero expectation and variance one. As a discrete counterpart of B we consider, for every n ∈ N = {1, 2,...}, a approximation bntc 1 X Bn := √ ξn , t ≥ 0, t n i i=1 n where (ξi )i∈N is a sequence of independent random variables which have the same distribution as ξ. We assume that the approximating sequence Bn converges to B pointwise in probability, i.e. n ∀ t ≥ 0 : lim B = Bt in probability. (1) n→∞ t The aim of the paper is to provide L2-approximation results for some basic operators of Malliavin calculus with respect to the Brownian motion B such as the chaos decomposition, the Malliavin derivative, and the Skorokhod integral by appropriate sequences of approximating operators n based on the discrete time noise (ξi )i∈N. It turns out that in all our approximation results, the limits do not depend on the distribution of the discrete time noise, hence our results can be regarded as some kind of invariance principle for Malliavin calculus. We briefly discuss our main convergence results in a slightly informal way: (1) Chaos decomposition: The heuristic idea behind the chaos decomposition in terms of multiple Wiener integrals is to project a random variable X ∈ L2(Ω, F,P ) on products ˙ ··· ˙ of the Bt1 Btk . This idea can be made rigorous with respect to the discrete n noise (ξ ) ∈N by considering the discrete time functions i i   nk/2 Yk f n,k(i , . . . i ) = E X ξn  X 1 k k! ij j=1

Date: April 6, 2016. 1 2 C. BENDER AND P. PARCZEWSKI

k for pairwise distinct (i1, . . . ik) ∈ N . Our results show that, after a natural embedding n,k as step functions into continuous time, the sequence (fX )n∈N converges strongly in L2([0, ∞)k) to the kth chaos coefficient of X, for every k ∈ N (Example 34). This is a simple consequence of a general Wiener chaos limit theorem (Theorem 28), which provides equivalent conditions for the strong L2(Ω, F,P )-convergence of a sequence of n n random variables (X )n∈N (with each X admitting a chaos decomposition via multiple n Wiener integrals with respect to the discrete time noise (ξi )i∈N) in terms of the chaos coefficient functions. As a corollary, this Wiener chaos limit theorem lifts a classical result by Surgailis [28] on convergence in distribution of discrete multiple Wiener integrals to strong L2(Ω, F,P )-convergence and adds the converse implication (in our setting, i.e. when the limiting multiple Wiener integral is driven by a Brownian motion). (2) Malliavin derivative: With our weak moment assumptions on the discrete time noise, we cannot define a discrete Malliavin derivative in terms of a polynomial chaos as in the survey paper by [11] and the references therein. Instead we think of the discretized n Malliavin derivative at time j ∈ N with respect to the noise (ξ ) ∈N as √ i i n E n | n Dj X = n [ξj X (ξi )i∈N\{j}], which is the gradient of the best approximation in L2(Ω, F,P ) of X as a linear func- n n ∈ N \{ } tion in ξj with σ(ξi , i j )-measurable coefficients. In the case of binary noise, this definition coincides with the standard notion of the Malliavin derivative on the Bernoulli space, see, e.g., [25]. Theorem 12 below implies that, if (Xn) converges weakly 2 F n n in L (Ω, ,P ) to X and the sequence of discretized Malliavin derivatives (Ddn·eX )n∈N converges weakly in L2(Ω × [0, ∞)), then X belongs to the domain of the continu- ous Malliavin derivative and the continuous Malliavin derivative appears as the weak L2(Ω×[0, ∞))-limit. As the Malliavin derivative is a closed, but discontinuous , this is the best type of approximation result which can be expected when discretizing the Malliavin derivative. Sufficient conditions for the strong convergence of a sequence of discretized Malliavin derivatives, which can be checked in terms of the discrete-time approximations, are presented in Theorems 16 and 35. (3) Skorokhod integral: Defining the discrete Skorokhod integral as the adjoint operator to the discretized Malliavin derivative leads to XM n n n E n| n √ξi δ (Z ) := lim [Zi (ξj )j∈{1,...,M}\{i}] , M→∞ n i=1 for a suitable class of discrete time processes Zn, which is in line with the Riemann- sum approximation for Skorokhod integrals in terms of the driving Brownian motion in [24]. Analogous results for the ‘closedness across the discretization levels’ as in the case of the discretized Malliavin derivative and sufficient conditions for strong L2(Ω, F,P )- convergence of a sequence of discrete Skorokhod integrals are provided in Theorems 8, 18 and 36. When restricted to predictable integrands, the convergence results for the Skorokhod integral give rise to necessary and sufficient conditions for strong and weak L2(Ω, F,P )-convergence of a sequence of discrete Itˆointegrals (Theorem 20). This result can be applied to study different discretization schemes for the generalized Clark-Ocone derivative (which provides the integrand in the predictable representation of a square- integrable random variable as Itˆointegral with respect to the Brownian motion B). In this respect, Theorems 23 and 25 below complement related results in the literature such as [6, 20] and the references therein. We note that related classical limit theorems for stochastic integrals (with adapted integrands) [15, 19] and for multiple Wiener integrals [2, 3, 28], or robustness results for martingale representations [6, 14] are usually obtained in the framework of (or using tech- niques of) convergence in distribution (on the Skorokhod space). In contrast, we exploit that strong and weak convergence in L2(Ω, F,P ) can be characterized in terms of the S-transform, DISCRETIZING MALLIAVIN CALCULUS 3 which is an important tool in white noise analysis, see e.g. [13,16,18], and corresponds to taking expectation under suitable changes of measure. We introduce a discrete time version of the S- n 2 F transform in terms of the noise (ξi )i∈N and show that strong and weak L (Ω, ,P )-convergence can be equivalently expressed via convergence of the discrete S-transform to the continuous S-transform (Theorem 1). With this observation at hand, all our convergence results can be 2 n obtained in a surprisingly simple way by computing suitable L (Ω, σ(ξi )i∈N,P )-inner products and their limits as n tends to infinity. The paper is organized as follows: In Section 2, we introduce the discrete S-transform and discuss the connections between weak (and strong) L2(Ω, F,P )-convergence and the convergence of the discrete S-transform to the continuous one. Equivalent conditions for the weak L2- convergence of sequences of discretized Malliavin derivatives and discrete Skorokhod integrals to their continuous counterparts are derived in Section 3. By combining these weak L2-convergence results with the duality between discrete Skorokhod integral and discretized Malliavin derivative, we also identify sufficient conditions for the strong L2-convergence which can be checked solely in terms of the discrete time approximations. We are not aware of any such convergence results for general discrete time noise distributions in the literature. In Section 4, we specialize to the nonanticipating case and prove limit theorems for discrete Itˆointegrals and discretized Clark- Ocone derivatives. The strong L2-Wiener chaos limit theorem is presented in Section 5, and is applied in order to provide equivalent conditions for the strong L2-convergence of sequences of discretized Malliavin derivatives and discrete Skorokhod integrals in terms of tail conditions of the discrete chaos coefficients in Section 6. In the final Section 7, we first the consider the special case of binary noise (Subsection 7.1), in which discrete Malliavin calculus is very well studied, see e.g. the monograph [25]. We explain how our convergence results can be stated in a simplified way in this case and demonstrate by a toy example how to apply the results numerically in a Monte Carlo framework. In Subsection 7.2, we finally show, how our main convergence results can be translated into Donsker type theorems on convergence in distribution, when the discrete time noise is not necessarily assumed to be embedded into the driving Brownian motion. Two auxiliary results on the S-transform characterization of the Malliavin derivative and on the connection between strong L2-convergence and convergence in distribution are postponed to the Appendix.

2. Weak and strong L2-convergence via discrete S-transforms In this section, we study strong and weak L2(Ω, F,P )-convergence of a sequence (Xn) of random n F n n ∈ N F variables, where X is := σ(ξi , i )-measurable, to an -measurable X. As a main result of this section (Theorem 1), we provide an equivalent criterion for this convergence, which only requires to compute a family of L2(Ω, F n,P )-inner products (hence, expectations which involve n functionals of the discrete time noise (ξi )i∈N only) and their limits as n tends to infinity. Before doing so, let us recall that Bn can be constructed via a Skorokhod embedding of the random walk ! Xj ξi , ξ1, ξ2,... independent and with the same distribution as ξ, i=1 ∈N j √ into the rescaled Brownian motion ( nBt/n)t≥0. In this way, one obtains, for every n ∈ N, a n sequence of stopping times (τi )i∈N0 with respect to the augmentation of the filtration generated by B such that   n B := Bτ n bntc t≥0 bPntc 1 √ ≥ has the same distribution as ( n ξi)t 0 and converges to B uniformly on compacts in prob- i=1 ability (see e.g. [22, Lemma 5.24 (b)]). We now introduce the S-transform simultaneously in the continuous time setting and the discrete time setting, which turns ou to be the key tool for the proofs of our limit theorems. Recall, 4 C. BENDER AND P. PARCZEWSKI

2 that the mapping 1(0,t] 7→ Bt can be extended to a continuous linear mapping from L ([0, ∞)) to L2(Ω, F,P ), which is known as the Wiener integral. We denote the Wiener integral of a function f ∈ L2([0, ∞)) by I(f). The discrete Wiener integral is given by ∞ 1 X In(f n) := √ f n(i)ξn. n i i=1 Here, the discrete time function f n is a member of ( ) X∞ 2 N n N → R k nk2 1 n 2 ∞ Ln( ) := f : : f 2 N := (f (i)) < , Ln( ) n i=1 which obviously ensures that the series In(f n) converges (strongly) in L2(Ω, F n,P ). The Wick exponential is, by definition, the stochastic exponential of a Wiener integral I(f), i.e.,  Z ∞  exp(I(f)) := exp I(f) − 1/2 f 2(s)ds . 0 Hence, its discrete counterpart, the discrete Wick exponential, is given by Y∞    1 exp n (In(f n)) := 1 + √ f n(i)ξn . n i i=1 In particular, by Fatou’s lemma and the estimate 1 + x ≤ exp(x),

n n n 2 n 2 E[(exp (I (f ))) ] ≤ exp(kf k 2 N ) < ∞. (2) Ln( ) Notice also that, by the martingale convergence theorem, X∞        n n n E n n n | n − E n n n | n exp (I (f )) = 1 + exp (I (f )) (ξj )j≤i exp (I (f )) (ξj )j≤i−1 i=1 X∞ n n  n n ξi = 1 + f (i) exp n (I (f 1 − ))√ , (3) [1,i 1] n i=1 which is the discrete counterpart of the Dol´eans-Dadeequation. We finally recall that, for every X ∈ L2(Ω, F,P ) and f ∈ L2([0, ∞)), the S-transform is defined as (SX)(f) := E[X exp(I(f))]. n ∈ 2 F n n ∈ 2 N Analogously, for every X L (Ω, ,P ) and f Ln( ), we introduce the discrete S- transform as  (SnXn)(f n) := E[Xn exp n (In(f n))]. We emphasize that the S-transform is a powerful tool in the white noise analysis, see, e.g., [18], and has been successfully applied in the theory of stochastic partial differential equations, see [13]. To the best of our knowledge the discrete S-transform has, however, not been studied in the literature. Let us next denote by E the set of step functions on left half-open intervals, i.e., functions of the form Xm ∈ N ∈ R g(x) = aj1(bj ,cj ](x), m , aj, bj, cj . j=1 As the set of Wick exponentials of step functions {exp(I(g)), g ∈ E} is total in L2(Ω, F,P ), see e.g. [16, Corollary 3.40], every L2(Ω, F,P )-random variable is uniquely determined by its S-transform. More precisely, if for X,Y ∈ L2(Ω, F,P ), (SX)(g) = (SY )(g) for every g ∈ E, then X = YP -almost surely. We define the discretization of a step function g ∈ E as gˇn = (ˇgn(1), gˇn(2),...) := (g(1/n), g(2/n),...) , DISCRETIZING MALLIAVIN CALCULUS 5 and notice that { n ∈ E} ⊂ 2 N gˇ : g Ln( ) is the dense subspace of discrete time functions with finite support. The convergence results of integral and derivative operators in this paper rely on the following characterization of L2(Ω, F,P )-convergence in terms of convergence of the discrete S-transform to the continuous S-transform. Theorem 1. Suppose X,Xn ∈ L2(Ω, F,P ) for every n ∈ N, with Xn being F n-measurable. Then the following assertions are equivalent as n tends to infinity: (i) Xn → X strongly (resp. weakly) in L2(Ω, F,P ). (ii) (SnXn)(ˇgn) → (SX)(g) for every g ∈ E, and additionally E[(Xn)2] → E[X2] in the case E n 2 ∞ of strong convergence (resp. supn∈N [(X ) ] < in the case of weak convergence). In view of Lemma 3 below, the proof of Theorem 1 can be reduced to the following strong L2-convergence result for (discrete) Wick exponentials. Proposition 2. Suppose g ∈ E. Then, we have strongly in L2(Ω, F,P ), as n tends to infinity:   exp n (In(ˇgn)) → exp (I(g)). These type of convergence results for stochastic exponentials are somewhat standard and can be obtained in a much more general context by applying results on convergence in distribution for stochastic differential equations, see, e.g., [2, 19] and the references therein. For sake of completeness, we here provide an elementary proof. Proof. Let Xm ∈ E g = aj1(bj ,cj ] . j=1 We denote by C,N constants in N such that g is bounded by C and has support in [0,N]. It suffices to show h i h i n n n 2  2 (i) limn→∞ E (exp (I (ˇg ))) = E (exp (I(g))) ,   (ii) exp n (In(ˇgn)) → exp (I(g)) in probability.

(i) Due to p < dqe ≤ r ⇔ bpc < q ≤ brc for all p, q, r ∈ R, we obtain for every t ∈ (0, ∞), Xm n d e gˇ ( nt ) = aj1(bbj nc/n,bcj nc/n](t). (4) j=1 Hence,   √ Xm n   1 kg − gˇ (dn·e)k 2 ∞ ≤ 2 |a | √ → 0, (5) L ([0, )) j n j=1 and in particular, Z XNn 1 ∞ (ˇgn(i))2 = kgˇn(dn·e)k2 → g(s)2ds. n L2([0,∞)) i=1 0 n Thus, by the independence of the centered random variables (ξi )i∈N with unit variance and taking the boundedness of g into account, we get       YNn YNn  1 1 E (exp n (In(ˇgn)))2 = E (1 + √ gˇn(i)ξn)2 = 1 + (ˇgn(i))2 n i n i=1 i=1 Z ∞  h i → exp g(s)2ds = E (exp(I(g)))2 . 0 6 C. BENDER AND P. PARCZEWSKI

(ii) In order to treat the large jumps of Bn and the small ones separately, we consider

n,1 n √ n,2 n √ ξi := ξi 1{| n|≤ n }, ξi := ξi 1{| n| n }, ξi 2C ξi > 2C cp. also [27]. Then,     YNn YNn  1 1 exp n (In(ˇgn)) = 1 + √ gˇn(i)ξn,1 1 + √ gˇn(i)ξn,2 =: En,1 · En,2. n i n i i=1 i=1 n We note that, for every  > 0, by the independence of (ξi )i∈N, ( )!  √  |ξn| P ({|ξ| >  n})Nn Nn P sup √i >  = 1 − 1 − → 0, (6) i=1,...,Nn n Nn √ because, by square-integrability of ξ, P ({|ξ| >  n})n → 0, see, e.g., [26, p. 208]. Hence, for every  > 0, ( )! | n| {| n,2 − | ≥ } ≤ { | n,2| } √ξi → P ( E 1  ) P ( sup ξi > 0 ) = P sup > 1/(2C) 0, i=1,...,Nn i=1,...,Nn n n,2 n,1 i.e., (E )n∈N converges to 1 in probability. By construction, each factor in E is larger than 1/2. Applying a Taylor expansion to the logarithm, thus, yields XNn ξn,1 1 XNn (ξn,1)2 log En,1 = gˇn(i) √i − (ˇgn(i))2 i + R n 2 n n i=1 i=1 with a remainder term satisfying ! 8C |ξn| XNn (ξn,1)2 |R | ≤ sup √j (ˇgn(i))2 i . n 3 n n j=1,...,Nn i=1 It, thus, suffices to show P ξn,1 Nn n √i → (iii) i=1 gˇ (i) I(g) in probability, n R P (ξn,1)2 ∞ Nn n 2 i → 2 (iv) i=1 (ˇg (i)) n 0 g(s) ds in probability. Indeed, by (6), the remainder term then vanishes in probability as n tends to infinity, and, thus,  Z  1 ∞ En,1 → exp I(g) − g(s)2ds in probability. 2 0 The same argument, which was applied for the convergence of En,2, shows that we can (and n,1 n shall) replace ξi by ξi in (iii) and (iv). However, by (1) and (4), XNn n Xm   Xm  n ξi n n lim gˇ (i)√ = lim aj B − B = aj Bc − Bb = I(g) n→∞ n n→∞ cj bj j j i=1 j=1 j=1 P 1 bntc n 2 in probability. Finally, by the , n i=1 (ξi ) converges to t in probability for every t ≥ 0, and, hence, by (4), Z XNn n 2 Xm ∞ n 2 (ξi ) 2 2 lim (ˇg (i)) = a (cj − bj) = g(s) ds, in probability. n→∞ n j i=1 j=1 0  The following lemma from functional analysis turns out to be useful. Lemma 3. Suppose H is a , A is an arbitrary index set, {xa, a ∈ A} is total in H, ∈ a a and, for every a A, (xn)n∈N is a sequence in H which converges strongly in H to x . Then, the following are equivalent, as n tends to infinity: DISCRETIZING MALLIAVIN CALCULUS 7

(i) xn → x strongly (resp. weakly) in H. h a i → h ai ∈ k k → k k (ii) xn, xn H x, x H for every a A, and additionally xn H x H in the case of k k ∞ strong convergence (resp. supn∈N xn H < in the case of weak convergence). k k Proof. Firstly, we observe that supn∈N xn H is finite, either by weak convergence [29, Theorem ∈ a V.1.1] in (i) or by assumption (ii). Thus, for every a A, by the strong convergence of (xn) to xa, |h a i − h ai | |h a − ai | ≤ k k k a − ak → xn, xn H xn, x H = xn, xn x H sup xm H xn x H 0. (7) m∈N a Let us treat the case of weak convergence: If (i) holds, the term hxn, x iH in (7) converges h ai h a i to x, x H , and then so does xn, xn H , which implies (ii). Conversely, if (ii) holds, the first h a i h ai h ai term xn, xn H in (7) tends to x, x H , and then so does xn, x H , which yields (i) in view of [29, Theorem V.1.3]. The case of strong convergence is an immediate consequence, as, in a Hilbert space, strong convergence is equivalent to weak convergence and convergence of the norms [29, Theorem V.1.8].  In view of the definition of the (discrete) S-transform, and as the set of Wick exponentials of step functions {exp(I(g)), g ∈ E} is total in L2(Ω, F,P ), Theorem 1 now turns out to be a direct consequence of Proposition 2 and Lemma 3 with H = L2(Ω, F,P ). We close this section with an example. Example 4. (i) In this example, we provide a simple proof, that, for every X ∈ L2(Ω, F,P ), Xn := E[X|F n] converges to X strongly in L2(Ω, F,P ). Indeed, by Proposition 2, for every g ∈ E,  (SnXn)(ˇgn) = E [E[X|F n] exp n (In(ˇgn))]   = E [X exp n (In(ˇgn))] → E [X exp (I(g))] = (SX)(g). As E[(Xn)2] ≤ E[X2], Theorem 1 implies weak L2(Ω, F,P )-convergence of (Xn) to X. The same theorem finally yields strong L2(Ω, F,P )-convergence, since, by the already established weak convergence, E[(Xn)2] = E [E [Xn| F n] X] = E[XnX] → E[X2]. We note that this result can also be derived by the uniform integrability of ((Xn)2) via the concept of convergence of filtrations making use of [8, Proposition 2]. F F n n n (ii) Denote by ( t)t≥0 the augmented Brownian filtration and let i = σ(ξ1 , . . . , ξi ). We ∈ 2 F n assume X L (Ω, T ,P ). Then, one can always approximate X by a sequence (XT ) strongly 2 F n F n n in L (Ω, ,P ), where XT is measurable with respect to bnT c. Indeed, take any sequence (X ) of F n-measurable random variables which converges strongly in L2(Ω, F,P ) to X, and define n E n|F n ∈ E XT = [X bnT c]. Then, for every g , by Proposition 2,   bnT c   Y 1 (SnXn)(ˇgn) = E Xn 1 + √ g(i/n)ξn  T n i i=1 h i n   n n n ˇ  = E X exp (I ((g1(0,T ]) )) → E X exp (I(g1(0,T ]))  = E [XE[exp (I(g))|FT ]] = (SX)(g) . Moreover, E n 2 ≤ E n 2 ∞ sup [(XT ) ] sup [(X ) ] < . n∈N n∈N n 2 F 2 F Hence, (XT ) converges weakly in L (Ω, ,P ) to X by Theorem 1. Then, strong L (Ω, ,P )- convergence follows by Theorem 1 as well, because E n 2 E n E n n − → E 2 [(XT ) ] = [XT X] + [XT (X X)] [X ]. 8 C. BENDER AND P. PARCZEWSKI

3. Weak L2-approximation of the Skorokhod integral and the Malliavin derivative In this section, we first discuss weak L2-approximations of the Skorokhod integral and the Malliavin derivative via appropriate discrete-time counterparts. We then show how to lift these results from weak convergence to strong convergence via duality under appropriate conditions which can be formulated in terms of the discrete-time approximations. While most presentations of Malliavin calculus first introduce the Malliavin derivative and then define the Skorokhod integral as adjoint operator of the Malliavin derivative, we shall here employ the following equivalent characterization of the Skorokhod integral in terms of the S-transform, cp. [16, Theorem 16.46, Theorem 16.50]. 2 2 Definition 5. Z ∈ L (Ω × [0, ∞)) := L (Ω × [0, ∞), F ⊗ B([0, ∞)),P ⊗ λ[0,∞)) is said to belong to the domain D(δ) of the Skorokhod integral, if there is an X ∈ L2(Ω, F,P ) such that for every g ∈ E Z ∞ (SX)(g) = (SZt)(g)g(t)dt. 0 In this case, X is uniquely determined and δ(Z) := X is called the Skorokhod integral of Z. For the discrete-time approximation we first introduce the space ( ) X∞ 2 × N n N → 2 F n k nk2 1 E n 2 ∞ Ln(Ω ) := Z : L (Ω, ,P ), Z 2 ×N := [(Zi ) ] < . Ln(Ω ) n i=1 Moreover, we recall the definitions F n n ∈ N F n n := σ(ξj , j ), M := σ(ξ1 , . . . , ξM ), and introduce the shorthand notations F n n ∈ N \{ } F n n ∈ { }\{ } −i := σ(ξj , j i ), M,−i := σ(ξj , j 1,...,M i ). n ∈ 2 × N n Definition 6. We say, Z Ln(Ω ) belongs to the domain D(δ ) of the discrete Skorokhod integral, if XM n n n E n|F n √ξi δ (Z ) := lim [Zi M,−i] . (8) M→∞ n i=1 exists strongly in L2(Ω, F,P ). If this is the case, δn(Zn) is called the discrete Skorokhod integral of Zn. E n|F n n We note that, by the independence of [Zi M,−i] and ξi , each summand on the right-hand side of (8) is indeed a member of L2(Ω, F,P ). Moreover, the martingale convergence theorem n ∈ 2 × N ∈ N n ∈ n implies that, for every Z Ln(Ω ) and N , Z 1[1,N] D(δ ) and

XN n n n n n ξi δ (Z 1 ) = E[Z |F− ]√ . (9) [1,N] i i n i=1 2 × N 2 F Hence, the discrete Skorokhod integral is densely defined from Ln(Ω ) to L (Ω, ,P ). We will show in Proposition 13 below that it is a closed operator. Remark 7. This definition of the discrete Skorokhod integral closely resembles the following Riemann-sum approximation of the Skorokhod integral by [24], who show that under appropriate conditions on Z, " # Z i+1 n  E n Zsds (Bs,B1 − Br) ≤ ≤ i ≤ i+1 ≤ ≤ B(i+1)/n − Bi/n i 0 s n n r 1 n 2 converges strongly in L (Ω, F,P ) to δ(Z1[0,1]). DISCRETIZING MALLIAVIN CALCULUS 9

As a first main result of this section we are going to show the following weak approximation theorem for Skorokhod integrals. n ∈ n ∈ N n Theorem 8. Suppose Z D(δ ) for every n , and (Zdn·e)n∈N converges to Z weakly in L2(Ω × [0, ∞)). Then, the following assertions are equivalent: E | n n |2 ∞ (i) supn∈N [ δ (Z ) ] < . (ii) Z ∈ D(δ) and (δn(Zn)) converges to δ(Z) weakly in L2(Ω, F,P ) as n tends to infinity. As a first tool for the proof we state the discrete S-transform of a discrete Skorokhod integral. Proposition 9. Suppose Zn ∈ D(δn). Then, for every g ∈ E, X∞ n n n n 1 n n n n (S δ (Z )) (ˇg ) = (S Z )(ˇg 1N\{ })ˇg (i). n i i i=1 This result is a special case of the more general Proposition 13 below, to which we refer the reader for the proof. The second tool for the proof of Theorem 8 is the following variant of Theorem 1 for stochastic processes. ∈ 2 × ∞ n n ∈ 2 × N ∈ N Theorem 10. Suppose Z L (Ω [0, )), (Z )n∈N satisfies Z Ln(Ω ) for every n . Then the following assertions are equivalent as n tends to infinity: n 2 × ∞ (i) (Zdn·e) converges strongly (resp. weakly) to Z in L (Ω [0, )). (ii) For every g, h ∈ E ∞ Z 1 X ∞ (SnZn)(ˇgn)hˇn(i) → (SZ )(g)h(s)ds. n i s i=1 0 R ∞ R ∞ and, additionally, E[ (Zn )2ds] → E[ Z2ds] in the case of strong convergence R 0 dnse 0 s E ∞ n 2 ∞ (resp. supn∈N [ (Zd e) ds] < in the case of weak convergence). P 0 ns P 1 ∞ n n n ˇn 1 ∞ n n n ˇn Here, n i=1(S Zi )(ˇg )h (i) can be replaced by n i=1(S Zi )(ˇg 1N\{i})h (i) in (ii). Proof. We wish to apply Lemma 3 in order to prove the equivalence of (i) and (ii). As L2(Ω × [0, ∞)) = L2(Ω, F,P ) ⊗ L2([0, ∞)) (with the tensor product in the sense of Hilbert spaces), the set {exp(I(g))h; g, h ∈ E} is total in L2(Ω × [0, ∞)). By Proposition 2 and n n n ˇn  2 (5), (exp (I (ˇg ))h (dn·e))n∈N converges to exp (I(g))h strongly in L (Ω × [0, ∞)) for every g, h ∈ E. As X∞ D E 1  n n n n n ˇn n nI (ˇg )ˇn d ·e (S Zi )(ˇg )h (i) = Zdn·e, e h ( n ) , n L2(Ω×[0,∞)) i=1 Lemma 3 applies indeed. We finally note, that the modified assertion is an immediate conse- quence of the Cauchy-Schwarz inequality and the estimate h  i h  i h  i   2  2 √ 2 E n n n − n n n E n n n E n n exp (I (ˇg )) exp (I (ˇg 1N\{i})) = exp (I (ˇg 1N\{i})) gˇ (i)ξi / n ≤ k nk2 | |2 → exp( gˇ L2 (N)) sup g(j) /n 0, n j∈N making use of (2) in the last line.  We are now ready to give the proof of Theorem 8. Proof of Theorem 8. As the implication ‘(ii) ⇒ (i)’ is trivial, we only have to show the converse implication. To this end, note first that, by Proposition 9 and Theorem 10, for every g ∈ E, Z ∞ lim (Snδn(Zn))(ˇgn) = (SZ )(g)g(t)dt. (10) →∞ t n 0 10 C. BENDER AND P. PARCZEWSKI

n n As the sequence (δ (Z ))n∈N is norm bounded by (i), it has a weakly convergent subsequence [29, Theorem V.2.1]. We denote its limit by X. Then, applying Theorem 1 and (10) along the subsequence, we obtain, for every g ∈ E, Z ∞ (SX)(g) = (SZt)(g)g(t)dt. (11) 0 Hence, by Definition 5, Z ∈ D(δ) and δ(Z) = X. Finally, by Theorem 1 and (10)–(11), weak 2 n n L (Ω, F,P )-convergence of (δ (Z ))n∈N to δ(Z) holds along the whole sequence, and not only along the subsequence.  We now turn to the weak approximation of the Malliavin derivative. Again, we apply a definition in terms of the S-transform, which we show to be equivalent to the more classical one in terms of the chaos decomposition in the Appendix. Definition 11. A random variable X ∈ L2(Ω, F,P ) is said to belong to the domain D1,2 of the Malliavin derivative, if there is a Z ∈ L2(Ω × [0, ∞)) such that for every g, h ∈ E, Z ∞   Z ∞   (SZs)(g)h(s)ds = E X exp (I(g)) I(h) − g(s)h(s)ds . 0 0 In this case, Z is unique and DX := Z is called the Malliavin derivative X. For every X ∈ L2(Ω, F,P ) we define the discretized Malliavin derivative of X at j ∈ N with n respect to (ξi )i∈N by √ n E n |F n Dj X := n [ξj X −j]. n 2 F 2 F We note that, for fixed j, Dj is a continuous linear operator from L (Ω, ,P ) to L (Ω, ,P ), because by H¨older’sinequality for conditional expectations and the independence of the family n (ξi )i∈N, | n |2 ≤ E 2|F n E n 2|F n E 2|F n Dj X n [X −j] [(ξj ) −j] = n [X −j]. 1,2 We say that X belongs to the domain Dn of the discretized Malliavin derivative, if the process n n 2 ×N n D X := (Di X)i∈N is a member of Ln(Ω ). In this case D X is called the discretized Malliavin n n derivative of X with respect to (ξi )i∈N. As Dj is continuous for fixed j, it is easy to check that the discretized Malliavin derivative is a densely defined closed operator from L2(Ω, F,P ) to 2 × N Ln(Ω ). n In the following theorem and in the remainder of the paper we use the convention Z0 = 0 for n ∈ 2 × N Z Ln(Ω ). n 2 n 1,2 Theorem 12. Suppose (X )n∈N converges to X weakly in L (Ω, F,P ) and X ∈ Dn for every ∈ N n . Then, theP following are equivalent: 1 ∞ E n n 2 ∞ (i) supn∈N n i=1 [(Di X ) ] < . ∈ D1,2 n n 2 × ∞ (ii) X and (Ddn·eX )n∈N converges to DX weakly in L (Ω [0, )). The proof is prepared by two propositions. The first one contains the duality relation between the discrete Skorokhod integral and discretized Malliavin derivative. Proposition 13. For every n ∈ N, the discrete Skorokhod integral is the adjoint operator of n 1,2 the discretized Malliavin derivative. In particular, δ is closed and, for every X ∈ Dn and Zn ∈ D(δn), ∞ 1 X E [ZnDnX] = E[δn(Zn)X]. n i i i=1  We emphasize that, choosing X = exp n (In(ˇgn)), g ∈ E, in Proposition 13, we obtain the n ∈ 2 N assertion of Proposition 9. Indeed, for every f Ln( ), n n n n n n n n Di exp (I (f )) = f (i) exp (I (f 1N\{i})). DISCRETIZING MALLIAVIN CALCULUS 11

n n 1,2 Proof. Suppose first, that Z ∈ D(δ ) and X ∈ Dn . Then, for every M ∈ N, and i ∈ N, h √ i h i E E n |F n 2 ≤ E | n |2 n [ξi X M,−i] Di X . Hence, by the martingale convergence theorem and dominated convergence, X∞ h √ i 1 E E n |F n − n 2 lim n [ξi X M,−i] Di X = 0. M→∞ n i=1 Consequently, X∞ XM  √  XM   1 E n n 1 E n E n |F n √1 E nE n|F n [Zi Di X] = lim Zi n [ξi X M,−i] = lim Xξi [Zi M,−i] n M→∞ n M→∞ n i=1 "i=1 # i=1 XM n E E n|F n √ξi E n n = lim X [Zi M,−i] = [Xδ (Z )]. M→∞ n i=1 Conversely, suppose that Zn is in the domain of the adjoint operator of the discretized Malliavin n 2 1,2 derivative, i.e., there is an Y ∈ L (Ω, F,P ) such that for every X ∈ Dn , ∞ 1 X E [ZnDnX] = E[Y nX]. (12) n i i i=1 1,2 n 1,2 We first note that, by construction, X ∈ Dn if and only E[X|F ] ∈ Dn , and, if this is the case, both random variables have the same discretized Malliavin derivative. Hence, applying E |F n n E n|F n the duality relation (12), with X and [X √], we obtain Y = [Y ]. Now suppose that ∈ 2 F n ∈ D1,2 n E n |F n ≤ n X L (Ω, M ,P ). Then X n , Di X = n [ξi X M,−i] for every i M, and Di X = 0 for i > M. Hence, (12) and the same manipulations as above imply " # XM n n n n ξi E[Y X] = E X E[Z |F − ]√ , i M, i n i=1 i.e. XM n n n n n ξi E[Y |F ] = E[Z |F − ]√ . M i M, i n i=1 E n|F n 2 F By the martingale convergence theorem, ( [Y M ])M∈N converges strongly in L (Ω, ,P ) to E[Y n|F n] = Y n. Hence, Zn ∈ D(δn) and δn(Zn) = Y n. Now, closedness is a general property of adjoint operators, see [29, p. 196].  The next proposition is a consequence of the weak convergence result for discrete Skorokhod integrals in Theorem 8. Proposition 14. For every g, h ∈ E,  Z ∞    lim δn(exp n (In(ˇgn)) hˇn) = exp (I(g)) I(h) − g(s)h(s)ds →∞ n 0 strongly in L2(Ω, F,P ).  Proof. Notice first that, for fixed n ∈ N, exp n (In(ˇgn)) hˇn ∈ D(δn), because hˇn(i) vanishes, if i is sufficiently large. A direct computation, making use of (9), shows X∞ n n  n n n  n n n ξi δ (exp n (I (ˇg )) hˇ ) = exp n (I (ˇg 1N\{ }))hˇ (i)√ . i n i=1 n For i =6 j we obtain, by independence of (ξ ) ∈N, k k     Y  n n  n n n n n 1 n 1 n 2 1 E exp n (I (ˇg 1N\{ })) exp n (I (ˇg 1N\{ }))ξ ξ =g ˇ (i)√ gˇ (j)√ 1 +g ˇ (k) . i j i j n n n k∈N\{i,j} 12 C. BENDER AND P. PARCZEWSKI

Combining this with an analogous calculation for the case i = j yields   h i X∞ Y  2 1 1 E δn(exp n (In(ˇgn)) hˇn) = hˇn(i)ˇgn(i)hˇn(j)ˇgn(j) 1 +g ˇn(k)2 n2 n i,j=1, i=6 j k∈N\{i,j} ∞   1 X Y 1 + hˇn(i)2 1 +g ˇn(k)2 . n n i=1 k∈N\{i} As g and h are bounded with compact support, it is straightforward to check in view of (5) that ! h i Z ∞  Z ∞ R ∞ 2 n  n n n 2 g(s)2ds 2 lim E δ (exp n (I (ˇg )) hˇ ) = e 0 h(s)g(s)ds + h(s) ds . (13) →∞ n 0 0

n n n n ˇn  2 Thus, (δ (exp (I (ˇg )) h ))n∈N converges to δ(exp (I(g)) h) weakly in L (Ω, F,P ) by Theorem 8. The identity  Z ∞  δ(exp(I(g)) h) = exp(I(g)) I(h) − g(s)h(s)ds 0 can either be derived by a direct computation making use of the S-transform definition of the Skorokhod integral (Definition 5) or alternatively is a simple consequence of [23, Proposition 1.3.3] in conjunction with Definition 1.2.1 in the same reference. Applying the Cameron-Martin shift [16, Theorem 14.1] twice, we observe " #   Z ∞  h i 2 R ∞ I(g) g(s)2ds I(g) 2 E e I(h) − g(s)h(s)ds = e 0 E e I(h) 0 " #  Z ∞  R ∞ 2 g(s)2ds = e 0 E I(h) + g(s)h(s)ds 0 ! Z ∞  Z ∞ R ∞ 2 g(s)2ds 2 = e 0 h(s)g(s)ds + h(s) ds . 0 0 Thanks to (13), this turns weak into strong convergence.  The proof of Theorem 12 is now analogous to that of Theorem 8. Proof of Theorem 12. ‘(ii) ⇒ (i)’ is obvious, since ∞ Z 1 X ∞ E[(DnXn)2] = E[(Dn Xn)2]ds. n i dnse i=1 0  ‘(i) ⇒ (ii)’: Notice first that, for every g, h ∈ E, by Proposition 13 with Zn = exp n (In(ˇgn)) hˇn and Proposition 14, X∞ 1  lim (SnDnXn)(ˇgn)hˇn(i) = lim E[Xnδn(exp n (In(ˇgn))hˇn)] n→∞ n i n→∞ i=1   Z ∞  = E X exp(I(g)) I(h) − g(s)h(s)ds , (14) 0 n 2 F n n since (X ) converges to X weakly in L (Ω, ,P ). The sequence (Ddn·eX )n∈N is norm bounded in L2(Ω × [0, ∞)) by (i), and, hence, it has a weakly convergent subsequence. We denote its limit by Z. Applying (14) and Theorem 10 along this subsequence, we conclude Z ∞   Z ∞   (SZs)(g)h(s)ds = E X exp (I(g)) I(h) − g(s)h(s)ds . (15) 0 0 Hence, X ∈ D1,2 and DX = Z by Definition 11. Finally, applying (14)–(15) and Theorem 10 n n 2 × along the whole sequence (Ddn·eX )n∈N, shows that this sequence converges weakly in L (Ω [0, ∞)) to DX.  DISCRETIZING MALLIAVIN CALCULUS 13

1,2 In order to check the assumptions of Theorem 8, we consider the space Ln , which consists of n ∈ 2 × N n ∈ D1,2 ∈ N processes Z Ln(Ω ) such that Zi n for every i and ∞ 1 X   E |DnZn|2 < ∞. n2 j i i,j=1, i=6 j

1,2 n n 1,2 Proposition 15. For every n ∈ N, Ln ⊂ D(δ ) and, for Z ∈ Ln , X∞ n n n n n ξi 2 δ (Z ) = E[Z |F− ]√ , (strong L (Ω, F,P )-convergence), (16) i i n i=1 h i ∞ h i ∞ 1 X   1 X   E (δn(Zn))2 = E E Zn|F n 2 + E (DnZn)(DnZn) . (17) n i −i n2 i j j i i=1 i,j=1, i=6 j In particular, in the context of Theorem 8, assertion (i) is equivalent to P h i 1 ∞ E n n n n ∞ (i’) supn∈N n2 i,j=1, i=6 j (Di Zj )(Dj Zi ) < , n 1,2 if we additionally assume that Z ∈ Ln for every n ∈ N.

Proof. Fix N1 < N2 ∈ N. Then,    2 XN2 n XN2 h i  n n ξi   1 n n 2 n 2 E E[Z |F− ]√ = E E[Z |F− ] (ξ ) i i n n i i i i=N1 i=N1 N 1 X2   + E E[Zn|F n ]E[Zn|F n ]ξnξn n i −i j −j i j i,j=N1, i=6 j

= (I)N1,N2 + (II)N1,N2 . n By the independence of the discrete-time noise (ξi )i∈N and as the conditional expectation has norm 1, we obtain as N tends to infinity, ∞ 1 XN   1 X   (I) = E E[Zn|F n ]2 → E E[Zn|F n ]2 < ∞, 1,N n i −i n i −i i=1 i=1 → and (I)N1,N2 0 as N1,N2 tend to infinity. In order to treat (II)N1,N2 , we first note that for any random variable Xn ∈ L1(Ω, F n,P ) and i =6 j ∈ N, by Fubini’s theorem,         E E n|F n F n E E n|F n F n X −i −j = X −j −i . (18) Hence, for i =6 j ∈ N,     E E[Zn|F n ]E[Zn|F n ]ξnξn = E E[Znξn|F n ]E[Znξn|F n ]   i −i j −j i j   i j −i j i −j  E E E n n|F n F n n n E E E n n|F n F n n n = [Zi ξj −i] −j Zj ξi = [Zi ξj −j] −i Zj ξi   1   = E E[Znξn|F n ]E[Znξn|F n ] = E (DnZn)(DnZn) . i j −j j i −i n i j j i Consequently, by Young’s inequality,   1   1   n E E[Zn|F n ]E[Zn|F n ]ξnξn ≤ E (DnZn)2 + E (DnZn)2 . i −i j −j i j 2 i j 2 j i 1,2 The Ln -assumption, thus, ensures that X∞   1 E n n n n ∞ lim (II)1,N = (Di Zj )(Dj Zi ) < N→∞ n2 i,j=1,i=6 j → n n and (II)N1,N2 0 as N1,N2 tend to infinity. Hence, by (9), the sequence (δ (Z 1[1,N]))N∈N is Cauchy in L2(Ω, F,P ). By the closedness of the discrete Skorokhod integral, Zn ∈ D(δn) and 14 C. BENDER AND P. PARCZEWSKI

1,2 n we obtain Ln ⊂ D(δ ), (16) and (17). We finally suppose that the assumptions of Theorem 8 n 1,2 are in force and that Z ∈ Ln for every n ∈ N. Then, X∞   1 E E n|F n 2 ∞ sup [Zi −i] < , ∈N n n i=1 n because of the assumed weak convergence of the sequence (Zdn·e)n∈N. Thus, the sequence n n 2 (δ (Z ))n∈N is norm bounded in L (Ω, F,P ), if and only if (i’) holds.  As a consequence of the previous proposition, we obtain the following strong L2(Ω, F,P )- convergence results to the Malliavin derivative. n 2 Theorem 16. Suppose (X )n∈N converges to X strongly in L (Ω, F,P ). Moreover assume that n 2,2 X ∈ Dn for every n ∈ N, i.e. ∞ ∞ 1 X   1 X   E (DnXn)2 + E (DnDnXn)2 < ∞. n i n2 j i i=1 i,j=1, i=6 j Then, the following assertions are equivalent:  P   P h i 1 ∞ E n n 2 1 ∞ E n n n 2 ∞ (i) supn∈N n i=1 (Di X ) + n2 i,j=1, i=6 j (Dj Di X ) < . ∈ D1,2 ∈ n n 2 × ∞ (ii) X , DX D(δ), (Ddn·eX )n∈N converges to DX strongly in L (Ω [0, )), and n n n 2 (δ (D X ))n∈N converges to δ(DX) weakly in L (Ω, F,P ). Remark 17. Recall that L = −δ ◦ D is the infinitesimal generator of the Ornstein-Uhlenbeck semigroup, see [23, Section 1.4], and is sometimes called Ornstein-Uhlenbeck operator (cf. also [16, Example 4.7]). So the previous theorem provides, at the same time, sufficient conditions for the strong convergence to the Malliavin derivative and the weak convergence to the Ornstein- Uhlenbeck operator. n n n n ∈ D2,2 n ∈ L1,2 6 Proof. Let Zi = Di X . Then, X n implies Z n . Note that, for i = j, by (18), n n n n n n n n Dj Zi = Dj Di X = Di Dj X = Di Zj , n n n n n n 2 i.e. (Dj Zi )(Di Zj ) = (Dj Di X) . Hence, by Theorem 12 and Theorem 8 in conjunction with Proposition 15, assertion (i) is equivalent to ∈ D1,2 ∈ n n 2 × ∞ (ii’) X , DX D(δ), (Ddn·eX )n∈N converges to DX weakly in L (Ω [0, )), and n n n 2 (δ (D X ))n∈N converges to δ(DX) weakly in L (Ω, F,P ). n n So we only need to show that under (ii’) the convergence of (Ddn·eX )n∈N to DX holds true in the strong topology. However, by the duality relation in Proposition 13, the weak L2(Ω, F,P )- n n n 2 n convergence of (δ (D X ))n∈N and the strong L (Ω, F,P )-convergence of (X )n∈N, Z ∞ Z ∞ E n n 2 E n n n n → E E 2 [(Ddn·eX ) ]dt = [δ (D X )X ] [δ(DX)X] = [(DtX) ]dt, 0 0 making use of the continuous time duality between Skorokhod integral and Malliavin derivative in the last step.  The analogous result for the Skorokhod integral reads as follows. n 2 × ∞ Theorem 18. Suppose (Zdn·e)n∈N converges strongly to Z in L (Ω [0, )) and assume that n 2,2 Z ∈ Ln , i.e., for every n ∈ N, ∞ ∞ 1 X   1 X   E |DnZn|2 + E |DnDnZn|2 < ∞. n2 i j n3 i j k i,j=1, i=6 j i,j,k=1, |{i,j,k}|=3 Then the following assertions are equivalent: DISCRETIZING MALLIAVIN CALCULUS 15  P h i 1 ∞ E n n n n ∞ (i) supn∈N n2 i,j=1, i=6 j (Di Zj )(Dj Zi ) < and   X∞   X∞    1 E E n n|F n 2 1 E n n n n n n  ∞ sup 2 [Di Zj −j] + 3 (Di Dj Zk )(Dk Dj Zi ) < . n∈N n n i,j=1, i=6 j i,j,k=1, |{i,j,k}|=3

1,2 n n 2 (ii) Z ∈ D(δ), δ(Z) ∈ D , (δ (Z ))n∈N converges to δ(Z) strongly in L (Ω, F,P ) and n n n 2 × ∞ (Ddn·eδ (Z ))n∈N converges to Dδ(Z) weakly in L (Ω [0, )). As a preparation we explain how to compute the dicretized Malliavin derivative of a discrete Skorokhod integral, which is analogous to the continuous-time situation, cp. e.g. [23, Proposition 1.3.8].

n ∈ L1,2 n n ∈ n ∈ N Proposition 19. Suppose Z n . Then (Di Z )1N\{i} D(δ ) for every i , and n n n E n|F n n n n Di δ (Z ) = [Zi −i] + δ (Di Z 1N\{i}).

n Proof. By (16) and the continuity of Di ,     n X∞ n n n n n n n ξi n n n ξj D δ (Z ) = D E[Z |F− ]√ + D E[Z |F− ]√ , i i i i n i j j n j=1, j=6 i

(including the strong convergence of the series on the right-hand side in L2(Ω, F,P )). By (18), for i =6 j, E nE n|F n n|F n E E n n|F n |F n n E E n n|F n |F n n [ξi [Zj −j]ξj −i] = [ [ξi Zj −j] −i]ξj = [ [ξi Zj −i] −j]ξj . Moreover, E n 2E n|F n |F n E n|F n [(ξi ) [Zi −i] −i] = [Zi −i]. Hence, X∞ n n n n n n n n n ξj D δ (Z ) = E[Z |F− ] + E[D Z |F− ]√ , i i i i j j n j=1, j=6 i and the closedness of the discrete Skorokhod integral concludes. 

L2,2 ∈ N n n ∈ Proof of Theorem 18. The n -assumption guarantees that, for every i ,(Di Z )1N\{i} L1,2 n 2 × ∞ n . As (Zdn·e)n∈N is norm bounded in L (Ω [0, )) by the assumed strong convergence to Z, we observe in view of Propositions 15 and 19 that (i) is equivalent to P E | n n |2 ∞ 1 ∞ E | n n n |2 ∞ (i’) supn∈N [ δ (Z ) ] < and supn∈N n i=1 [ Di δ (Z ) ] < . Thanks to Theorems 8 and 12, assertion (i’) is equivalent to 1,2 n n 2 (ii’) Z ∈ D(δ), δ(Z) ∈ D ,(δ (Z ))n∈N converges weakly to δ(Z) in L (Ω, F,P ), and n n n 2 (D δ (Z ))n∈N converges to Dδ(Z) weakly in L (Ω × [0, ∞)). n n n n Due to the strong convergence of (Zdn·e)n∈N to Z and the weak convergence of (Ddn·eδ (Z ))n∈N to D(δ(Z)), the continuous time duality between Skorokhod integral and Malliavin derivative and its discrete time counterpart in Proposition 13 imply Z ∞ Z ∞ k n n k2 E n n n n → E k k2 δ (Z ) L2(Ω,F,P ) = [ZdnseDdnseδ (Z )]ds [ZsDsδ(Z)]ds = δ(Z) L2(Ω,F,P ). 0 0 n n Hence we obtain the convergence of (δ (Z ))n∈N to δ(Z) in the strong topology, i.e., assertion (ii’) is equivalent to assertion (ii).  16 C. BENDER AND P. PARCZEWSKI

4. Strong and weak L2-approximation of the Itoˆ integral and the Clark-Ocone derivative In this section, we first specialize the approximation result for the Skorokhod integral to pre- dictable integrands. In this way, we obtain necessary and sufficient conditions for strong and 2 n weak L -convergence of discrete Itˆointegrals with respect to the noise (ξi )i∈N to Itˆointegrals with respect to the Brownian motion B. Then, we discuss strong and weak L2-approximations to the Clark-Ocone derivative, which provides the predictable integral representation of a random variable in L2(Ω, F,P ) with respect to the Brownian motion B. n ∈ 2 × N F n ∈ N n Suppose Z Ln(Ω ) is predictable with respect to ( i )i∈N, i.e., for every i , Zi is F n n n measurable with respect to i−1 = σ(ξ1 , . . . , ξi−1). Then, ∞ Z X ξn δn(Zn) = Zn √i =: ZndBn, i n i=1 which means that the discrete Skorokhod integral reduces to the discrete Itˆointegral. Analo- R ∞ gously, the Skorokhod integral δ(Z) is well-known to coincide with the Itˆointegral 0 ZsdBs, when Z ∈ L2(Ω × [0, ∞)) is predictable with respect to the augmented Brownian filtration (Ft)t∈[0,∞), see, e.g. [16, Theorem 7.41]. In this case of predictable integrands, the approxima- tion theorem for Skorokhod integrals (Theorem 8) can be improved as follows. Theorem 20. Suppose Z ∈ L2(Ω×[0, ∞)) is predictable with respect to the augmented Brownian F ∈ N n ∈ 2 ×N F n filtration ( t)t∈[0,∞), and, for every n , Z Ln(Ω ) is predictable with respect to ( i )i∈N. Then, the following are equivalent: n 2 (i) (Z ) ∈N converges to Z strongly (resp. weakly) in L (Ω × [0, ∞)). dn·e n R  n n (ii) The sequence of discreteR Itˆointegrals Z dB n∈N converges strongly (resp. weakly) 2 F ∞ in L (Ω, ,P ) to 0 ZsdBs. Remark 21. We note that, in order to study convergence of Itˆointegrals (with respect to different filtrations), techniques of convergence in distribution on the Skorokhod space of right- continuous functions with left limits are classically applied. E.g., the results by [19] immediately imply the following result in our setting: Suppose that Z is predictable with respect to the Brow- nian filtration and its paths are right-continuous with left limits. Moreover, assume that Zn F n n is predictable with respect to ( i )i∈N and (Zb1+n(·)c) converges to Z uniformly on compacts in probability. Then, Xbn·c Z · n 1 n lim Z √ ξ = Zs−dBs, n→∞ i n i i=1 0 uniformly on compacts in probability. In contrast, our Theorem 20 provides an L2-theory and, in particular, includes the converse implication, namely that convergence of the discrete Itˆo integrals implies convergence of the integrands. The proof of Theorem 20 will make use of the following proposition. Proposition 22. Suppose g, h ∈ E. Then, strongly in L2(Ω × [0, ∞)),

n n n ˇn  lim exp (I (ˇg 1[1,dn·e−1]))h (dn·e) = exp (I(g1(0,·]))h(·). n→∞ Proof. Recall that the support of h is contained in [0,M] for some M ∈ N. Hence, we can decompose, Z ∞ h i 2 n n n ˇn  E exp (I (ˇg 1[1,dnte−1]))h (dnte) − exp (I(g1(0,t]))h(t) dt 0Z M h  i n n n  2 2 ≤ 2 E exp (I (ˇg 1[1,bntc])) − exp (I(g1(0,t])) h(t) dt 0 DISCRETIZING MALLIAVIN CALCULUS 17 Z ∞ h  i 2 n n n ˇn 2 +2 E exp (I (ˇg 1[1,dnte−1])) |h (dnte) − h(t)| dt, 0 since dnte − 1 = bntc for Lebesgue almost every t ≥ 0. As, by (2) h  i  2 kgˇn(dn·e)k2 n n n L2([0,∞)) sup E exp (I (ˇg 1[1,dnte−1])) ≤ sup e < ∞, n∈N, t∈[0,∞) n∈N the second term goes to zero by (5). Moreover, by the boundedness of h, the first one tends to zero by the dominated convergence theorem, since, for every t ∈ [0, ∞), by Proposition 2, h  i n n n  2 lim E exp (I ((ˇg 1[1,bntc]))) − exp (I(g1(0,t])) = 0. n→∞ 

Proof of Theorem 20. ‘(i) ⇒ (ii)’: By the isometry for discrete Itˆointegrals, we have " #   Z  ∞ 2 Z h i 2 X 1 ∞ E ZndBn = E  Zn √ ξn  = E |Zn |2 ds. (19) i n i dnse i=1 0 n 2 F Hence, if (Zdn·e)n∈N converges to Z weakly in L (Ω, ,P ), then the left-hand side in (19) is bounded in n ∈ N, and so Theorem 8 impliesR the asserted weak L2(Ω, F,P ) convergence of ∞ n the sequence of discrete Itˆointegrals to 0 ZsdBs. If (Zdn·e)n∈N converges to Z strongly in L2(Ω, F,P ), then, by (19) and the continuous time Itˆoisometry, "Z  # Z "Z  # 2 ∞   ∞ 2 lim E ZndBn = E |Z |2 ds = E Z dB , →∞ s s s n 0 0 which turns the weak L2(Ω, F,P )-convergence of the sequence of discrete Itˆointegrals into strong L2(Ω, F,P )-convergence. ‘(ii) ⇒ (i)’: We first assume that the sequence of discrete Itˆointegrals converges weakly in L2(Ω, F,P ) to the continuous time Itˆointegral. By the implication ‘(i) ⇒ (ii)’ (which we have already proved) and Proposition 22, we obtain, for every g, h ∈ E, X∞ Z ∞  1  n n n ˇn √ n lim exp (I (ˇg 1[1,i−1]))h (i) ξ = exp (I(g1(0,s])) h(s) dBs (20) n→∞ n i i=1 0 strongly in L2(Ω, F,P ). As Zn is predictable, we get, for every g, h ∈ E, by the discrete Itˆo isometry, X∞ X∞ 1 1  (SnZn)(ˇgn)hˇn(i) = E[E[Zn|F n ] exp n (In(ˇgn))hˇn(i)] n i n i i−1 i=1 i=1 X∞ 1 n  n n n = E[Z exp n (I (ˇg 1 − ))hˇ (i)] n i [1,i 1] "i=1 ! !# X∞ X∞ n 1 n  n n n 1 n = E Z √ ξ exp n (I (ˇg 1 − ))hˇ (i)√ ξ . i n i [1,i 1] n i i=1 i=1 The assumed weak L2(Ω, F,P )-convergence of the sequence of discrete Itˆointegrals and the strong L2(Ω, F,P )-convergence in (20) now imply     X∞ Z ∞ Z ∞ 1 n n n n I(g1 ) lim (S Z )(ˇg )hˇ (i) = E ZsdBs e (0,s] h(s) dBs . n→∞ n i i=1 0 0 18 C. BENDER AND P. PARCZEWSKI

 As (exp (I(g1(0,s])))s∈[0,∞) is a uniformly integrable martingale and Z is predictable, we obtain, by the Itˆoisometry and the definition of the S-transform, X∞ Z ∞ 1 n n n n lim (S Z )(ˇg )hˇ (i) = (SZs)(g)h(s)ds, g, h ∈ E. n→∞ n i i=1 0 R h i ∞ E | n |2 We can now apply Theorem 10. As 0 Zdnse ds is bounded in n by (19) and by the assumed weak L2(Ω, F,P )-convergence of the discrete Itˆointegrals, the latter Theorem im- n 2 × ∞ plies that (Zdn·e)n∈N converges to Z weakly in L (Ω [0, )). If we instead assume strong L2(Ω, F,P )-convergence of the sequence of the discrete Itˆointegrals, a straightforward applica- tion of the isometries for discrete and continuous-time Itˆointegrals turns the weak L2(Ω×[0, ∞))- convergence again into strong convergence. 

We now turn to the Clark-Ocone derivative. Recall that a Brownian motion has the predictable representation property with respect to its natural filtration, i.e., for every X ∈ L2(Ω, F,P ) 2 there is a unique (Ft)t∈[0,∞)- ∇X ∈ L (Ω × [0, ∞)) such that Z ∞ X = E[X] + ∇sXdBs. (21) 0

We refer to ∇X as the generalized Clark-Ocone derivative and recall that (∇tX)t≥0 is the 1,2 predictable projection of the Malliavin derivative (DtX)t≥0, if X ∈ D . By Itˆo’sisometry the operator ∇ : L2(Ω, F,P ) → L2(Ω × [0, ∞)) is continuous with norm 1. Except in the case of binary noise, the discrete time approximation B(n) of the Brownian motion F n B does not satisfy the discrete time predictable representation property with respect to ( i )i∈N. Nonetheless one can consider the discrete time predictable projection of the discretized Malliavin derivative √ ∇n E n |F n E n |F n ∈ 2 F ∈ N i X := [Di X i−1] = n [ξi X i−1],X L (Ω, ,P ), i , ∇n as discretization of the generalized Clark-Ocone derivative. We refer to ( i X)i∈N as discretized Clark-Ocone derivative of X and note that it has been extensively studied in the context of discretization of backward stochastic differential equations, see, e.g., [6, 10, 30]. The operator ∇n 2 F → 2 × N 7→ ∇n : L (Ω, ,P ) Ln(Ω ),X ( i X)i∈N E · E ·|F n is continuous with norm one. Indeed, introducing the shorthand notation n,i[ ] = [ i ] and noting that the martingale (En,i[X])i∈N is, for fixed n ∈ N, uniformly integrable, and, thus, converges almost surely to E[X|F n], as i tends to infinity, one gets, by H¨older’sand Jensen’s inequality, X∞ h√  i X∞ h i 1 n 2 n 2 E n E − [ξ X] = E (E − [ξ (E [X] − E − [X])]) n n,i 1 i n,i 1 i n,i n,i 1 i=1 i=1 X∞ h   h ii ≤ E E n 2 E E − E 2 n,i−1 (ξi ) n,i−1 ( n,i[X] n,i−1[X]) i=1" # X∞   h i 2 2 n 2 2 = E (En,i[X]) − (En,i−1[X]) = E (E[X|F ]) − E [X] h i=1 i ≤ E (X)2 − E [X]2 .

We now denote by  Z  Pn n n ∈ R n ∈ 2 × N := a + Z dB ; a ,Z Ln(Ω ) predictable DISCRETIZING MALLIAVIN CALCULUS 19 the closed subspace in L2(Ω, F,P ), which admits a discrete time predictable integral represen- ∈ 2 F ∈ R F n n ∈ 2 × N tation. Note that, for every X L (Ω, ,P ), a , and ( i )i∈N-predictable Z Ln(Ω ), by the discrete Itˆoisometry,   Z  X∞ n n 1 n n n E X a + Z dB = aE[X] + √ E[Xξ E[Z |F − ]] n i i i 1 i=1 ∞ 1 X  √  = aE[X] + E Zn nE[Xξn|F n ] n i i i−1  i=1Z   Z  = E E[X] + ∇nXdBn a + ZndBn .

Hence, Z n n πPn X = E[X] + ∇ XdB , (22)

2 where, for any closed subspace A in L (Ω, F,P ), πA denotes the orthogonal projection on A. Our first approximation result for the Clark-Ocone derivative now reads as follows:

n 2 2 Theorem 23. Suppose (X )n∈N is a sequence in L (Ω, F,P ) and X ∈ L (Ω, F,P ). Then, the following are equivalent, as n tends to infinity: n n 2 (i) (πPn X − E[X ])n∈N converges to X − E[X] strongly (weakly) in L (Ω, F,P ). ∇n n ∇ 2 × ∞ (ii) ( dn·eX )n∈N converges to X strongly (weakly) in L (Ω [0, )). n 2 A sufficient condition for (i), (ii) is that (X )n∈N converges to X strongly (weakly) in L (Ω, F,P ). Proof. Recall that by (21) and (22) Z ∞ X − E[X] = ∇XsdBs, Z0 n n n n n πPn X − E[X ] = ∇ X dB .

 Hence, Theorem 20 provides the equivalence of (i) and (ii). As, for every g ∈ E, exp n (In(ˇgn)) ∈ Pn by (3), the sufficient condition is a consequence of the following lemma.  Lemma 24. Suppose that An, n ∈ N, are closed subspaces of L2(Ω, F,P ) such that for every n ∈ N, n n n {exp (I (ˇg )), g ∈ E} ⊂ An. 2 n n Then, strong (weak) L (Ω, F,P )-convergence of (X )n∈N to X implies that (πAn X )n∈N con- verges to X strongly (weakly) in L2(Ω, F,P ) as well. Proof. As, for every g ∈ E, n  n n n  n n n  n n E[(πAn X ) exp n (I (ˇg ))] = E[X πAn (exp n (I (ˇg )))] = E[X exp n (I (ˇg ))],

n n n n n n we obtain that (S X )(ˇg ) = (S (πAn X ))(ˇg ). In the case of weak convergence, Theorem 1 now immediately applies, because h i h i n 2 n 2 E (πAn X ) ≤ E (X ) .

In the case of strong convergence, we also make use of Theorem 1, and note that by the already n established weak convergence of (πAn X )n∈N and H¨older’sinequality, as n tends to infinity, h i   n 2 n n n 2 E (πAn X ) = E [X(πAn X )] + E [(X − X)(πAn X )] → E X .  20 C. BENDER AND P. PARCZEWSKI

We shall finally discuss an alternative approximation of the generalized Clark-Ocone derivative, which involves orthogonal projections on appropriate finite-dimensional subspaces. To this end, we denote by Hn the strong closure in L2(Ω, F,P ) of the linear span of ( ) Y n n n ⊆ N | | ∞ Ξ := ΞA := ξi ,A , A < , i∈A and emphasize that Hn = L2(Ω, F n,P ), if and only if the noise distribution of ξ is binary. As Ξn consists of an orthonormal basis of Hn, every Xn ∈ Hn has a unique expansion in terms of this Hilbert space basis, which is called the Walsh decomposition of Xn, X n n n X = XAΞA, (23) |A|<∞ P n E n n n 2 ∞ 2 F where XA = [X ΞA] satisfies |A|<∞(XA) < . The expectation and L (Ω, ,P )-inner n n product can be computed in terms of the Walsh decomposition via E[X ] = X∅ and X E n n n n n n ∈ Hn [X Y ] = XAYA ,X ,Y , |A|<∞ cp. [12]. A direct computation shows that the Walsh decomposition of a discrete Wick expo- nential is given by ! X Y n n n −|A|/2 n n exp (I (f )) = n f (i) ΞA. (24) |A|<∞ i∈A In view of the M¨obiusinversion formula [1, Theorem 5.5], we obtain, for every finite subset B of N, X | | | |−| |  n B /2 − B C n n ΞB = n ( 1) exp (I (1C )). C⊆B  Hence, the set {exp n (In(ˇgn)), g ∈ E} is total in Hn. We now consider the finite-dimensional subspaces Hn { n ⊂ { }} i := span ΞA,A 1, . . . , i , and introduce, as a second approximation of the generalized Clark-Ocone derivative, the operator n 2 2 n ∇¯ : L (Ω, F,P ) → L (Ω × N),X 7→ (πHn (∇ X)) ∈N. n i−1 i i Notice that √ n n ∇¯ X = n πHn (ξ X), i i−1 i n ∈ 2 F if ξi X L (Ω, ,P ). We are now going to show the following variant of Theorem 23. n 2 2 Theorem 25. Suppose (X )n∈N is a sequence in L (Ω, F,P ) and X ∈ L (Ω, F,P ). Then, the following are equivalent, as n tends to infinity: n n 2 (i) (πHn X − E[X ])n∈N converges to X − E[X] strongly (weakly) in L (Ω, F,P ). ∇¯ n n ∇ 2 × ∞ (ii) ( dn·eX )n∈N converges to X strongly (weakly) in L (Ω [0, )). n 2 A sufficient condition for (i), (ii) is that (X )n∈N converges to X strongly (weakly) in L (Ω, F,P ). The proof is based on the simple observation that Hn ⊂ Pn, i.e., for every Xn ∈ Hn, ∞ X 1 Xn = E[Xn] + ∇nXn √ ξn. (25) i n i i=1  In order to show this, we recall that {exp n (In(ˇgn)), g ∈ E} is total in Hn. Thus, by continuity of the discretized Clark-Ocone derivative and by the discrete Itˆoisometry, it suffices to show  (25) in the case Xn = exp n (In(fˇn)) for f ∈ E. A direct computation shows,   ∇n n n n n n n n i exp (I (f )) = f (i) exp (I (f 1[1,i−1])), (26) DISCRETIZING MALLIAVIN CALCULUS 21 which in view of (3) completes the proof of (25). Proof of Theorem 25. We first note that, for every X ∈ L2(Ω, F,P ),

E[πHn X] = E[X], (27) ∇¯ n ∇n i X = i (πHn X). (28) Indeed, as X E E n n πHn X = [X] + [XΞA]ΞA, 1≤|A|<∞ ∇n ∈ Hn Eq. (27) is obvious. In order to prove (28), we recall first that i (πHn X) i−1 (by (26) and continuity of the discretized Clark-Ocone derivative) and then note that, for every A ⊂ {1, . . . , i − 1},    n n n n n E Ξ E ξ X| F − = E[Ξ X] = E[Ξ πHn (X)] A i i 1 A∪{i}  A∪{i}     n n n n 1 n = E Ξ E ξ πHn (X)| F − = E Ξ √ ∇ (πHn X) . A i i 1 A n i In particular, by (25), (27), and (28) Z n n πHn X = E[X] + ∇¯ XdB , (29) which is the analogue of (22). The proof of Theorem 23 can now be repeated verbatim with Pn replaced by Hn.  We close this section with two remarks. Remark 26. In view of Lemma 24 and the inclusion Hn ⊂ Pn we observe that, for any sequence n 2 (X )n∈N in L (Ω, F,P ), 2 lim Xn = X strongly (weakly) in L (Ω, F,P ) n→∞ 2 ⇒ lim πPn Xn = X strongly (weakly) in L (Ω, F,P ) n→∞ 2 ⇒ lim πHn Xn = X strongly (weakly) in L (Ω, F,P ). n→∞ In particular, by Theorems 23 and 25, if the sequence of discretized Clark-Ocone derivatives ∇n n ∇ 2 × ∞ ( dn·eX )n∈N converges to X strongly (weakly) in L (Ω [0, )), then so does the sequence ∇¯ n n of modified discretized Clark-Ocone derivatives ( dn·eX )n∈N . Remark 27. The following result can be derived from [6, Theorem 5 and the examples in Section 5] under the additional assumption that E[|ξ|2+] < ∞ for some  > 0 and on a finite n 2 time horizon: Strong convergence of (X )n∈N to X in L (Ω, F,P ) implies convergence of the sequence of discretized Clark-Ocone derivatives as stated in (ii) of Theorem 23. Our Theorem 25 E ·|F n additionally shows that the conditional expectations [ i−1] in the definition of the discretized Hn Clark-Ocone derivative can be replaced by the projection on the finite dimensional subspace i , n 2 i.e., if (X )n∈N converges to X strongly in L (Ω, F,P ), then √  n n → ∇ n πHn (ξdnte (τn(X )) X dnte−1 t∈[0,∞) 2 strongly in L (Ω × [0, ∞)), where τn denotes the truncation at n. We also note that, in view of (29),

(πHn X) − (πHn X) ∇¯ X = i i−1 i n − n Bi Bi−1 n ξi n can be rewritten as difference operator (where we apply the convention n = 1 when ξ vanishes). ξi i This representation shows the close relation to the weak L2(Ω × [0, ∞))-approximation result for 22 C. BENDER AND P. PARCZEWSKI the generalized Clark-Ocone derivative which is derived in [20, Corollary 4.1] for the case of binary noise.

5. Strong L2-approximation of the chaos decomposition In this section, we apply Theorem 1 in order to characterize strong L2(Ω, F,P )-convergence of a sequence (Xn) (where Xn can be represented via multiple Wiener integrals with respect to n the discrete time noise (ξi )i∈N) via convergence of the coefficient functions of such a discrete chaos decomposition. Recall first, that every X ∈ L2(Ω, F,P ) has a unique Wiener chaos decomposition in terms of multiple Wiener integrals X∞ k k X = I (fX ), (30) k=0

k ∈ f2 ∞ k 2 ∞ k where fX L ([0, ) ), see e.g. [23, Theorem 1.1.2]. Here, we denote by L ([0, ) ) the Hilbert space of square-integrable functions with respect to the k-dimensional Lebesgue measure and by f L2([0, ∞)k) the subspace of functions in L2([0, ∞)k) which are symmetric in the k variables. We f apply the standard convention L2([0, ∞)0) = L2([0, ∞)0) = R, I0(f 0) = f 0, and recall that, for f k ≥ 1 and f k ∈ L2([0, ∞)k), the multiple Wiener integral can be defined as iterated Itˆointegral: Z Z Z ∞ tk t2 k k ··· k ··· I (f ) = k! f (t1, . . . , tk)dBt1 dBtk−1 dBtk . 0 0 0 The Itˆoisometry therefore immediately implies the following well-konwn Wiener-Itˆoisometry for multiple Wiener integrals, 0 0 E k k k k 0 h k ki [I (f ) I (g )] = δk,k k! f , g L2([0,∞)k (31) f 0 f 0 for functions f k ∈ L2([0, ∞)k) and gk ∈ L2([0, ∞)k ). The main theorem of this section now reads as follows:

n 2 Theorem 28 (Wiener chaos limit theorem). Suppose (X )n∈N is a sequence in L (Ω, F,P ). Then the following assertions are equivalent as n tends to infinity: n 2 (i) The sequence (πHn X ) converges strongly in L (Ω, F,P ). dn,k (ii) For every k ∈ N0, the sequence (f n )n∈N, defined via " X # d k/2 n,k E n n n f n (u1, . . . , uk) := X Ξ{d e d e} 1{|{dnu e,...,dnu e}∩N|=k}. (32) X k! nu1 ,..., nuk 1 k

is strongly convergent in L2([0, ∞)k) and ∞ X d k n,kk2 lim lim sup k! fXn L2([0,∞)k) = 0. (33) m→∞ n→∞ k=m P∞ n k k In this case, the limit X of (πHn X )n∈N has the Wiener chaos decomposition X = I (fX ) k=0 k dn,k 2 k with f = lim f n in L ([0, ∞) ). X n→∞ X We recall that, by Remark 26, the strong L2(Ω, F,P )-convergence of (Xn) to X is a sufficient condition for the strong approximation of the chaos coefficients of X as stated in (ii) of the above theorem. Before proving Theorem 28, we briefly discuss this result. To this end, we first recall the relation between Walsh decomposition and discrete chaos decomposition. The discrete multiple Wiener DISCRETIZING MALLIAVIN CALCULUS 23 integrals are defined analogously to the continuous setting, see e.g. [25, Section 1.3]. For all k, n ∈ N we consider the Hilbert space       X 2 2 Nk n,k Nk → R n,k ∞ Ln( ) := f : : f (i1, . . . , ik) <  k (i1,...,ik)∈N endowed with the inner product X n,k n,k −k n,k n,k hf , g i 2 Nk := n f (i , . . . , i )g (i , . . . , i ). Ln( ) 1 k 1 k k (i1,...,ik)∈N The closed subspace of symmetric functions in L2 (Nk) which vanish on the diagonal part n n o k ∂k := (i1, . . . , ik) ∈ N : |{i1, . . . , ik}| < k

f2 Nk is denoted by Ln( ). ∈ N n,k ∈ f2 Nk Then, for k , the discrete multiple Wiener integral of f Ln( ) with respect to the random walk Bn is defined as X In,k(f n,k) := n−k/2k! f n,k(i , . . . , i )Ξn . 1 k {i1,...,ik} k (i1,...,ik)∈N , i1<···

Remark 29. Convergence of discrete multiple Wiener integrals to continuous multiple Wiener integrals was studied in [28] as a main tool for proving noncentral limit theorems. The results in Section 4 of the latter reference imply that, for every k ∈ N0, the sequence of discrete mul- n,k n,k tiple Wiener integrals (I (f ))n∈N converges in distribution to the multiple Wiener integral k k dn,k k 2 k I (f ), if (f )n∈N converges to f strongly in L ([0, ∞) ). Our result lifts this convergence in distribution to strong L2(Ω, F,P )-convergence and, more importantly, adds the converse: d L2(Ω, F,P )- lim In,k(f n,k) = Ik(f k) ⇔ L2([0, ∞)k)- lim f n,k = f k. n→∞ n→∞ We note that the L2(Ω, F,P )-convergence of the sequence (In,k(f n,k)) even implies convergence in Lp(Ω, F,P ) for p > 2, if E[|ξ|r] < ∞ for some r > p. Indeed, in this case, the sequence (|In,k(f n,k)|p) is uniformly integrable by the hypercontractivity inequality of [17] in the variant of [4, Proposition 5.2]. The following elementary corollary of Theorem 28 generalizes Proposition 2. It makes use of the fact that the chaos decompositions of (discrete) Wick exponentials are given, for all ∈ 2 ∞ n ∈ 2 N f L ([0, )), f Ln( ), by X∞ X∞ I(f) k 1 ⊗k n n n n,k 1 n ⊗k e = I ( f ), exp (I (f )) = I ( (f ) 1∂c ). (37) k! k! k k=0 k=0 For a proof of the continuous case see e.g. [16, Theorem 3.21, Theorem 7.26]. The statement of the discrete case is a direct consequence of (24). ∈ 2 ∞ n n ∈ 2 N ∈ N Corollary 30. Suppose f L ([0, )) and (f ) is a sequence with f Ln( ) for every n . Then, as n tends to infinity (in the sense of strong convergence), fcn → f in L2([0, ∞)) ⇔ In(f n) → I(f) in L2(Ω, F,P )   ⇔ exp n (In(f n)) → exp (I(f)) in L2(Ω, F,P ).

Proof. In view of Theorem 28 and (37), we only have to show that fcn → f strongly in L2([0, ∞)) \⊗ ⊗k 2 k implies that (f n) k1 c → f strongly in L ([0, ∞) ), for every k ≥ 2. This is a consequence ∂k of the following lemma.  ∈ N n,k n,k ∈ 2 Nk Lemma 31. (i) Fix k 0. Suppose (f )n∈N is a sequence such that f Ln( ) for every d k 2 k \ n ∈ N and (f n,k) converges to some f strongly in L ([0, ∞) ). Then, the sequence (f n,k1 c ) ∂k converges to f k strongly in L2([0, ∞)k) as well. n n ∈ 2 N ∈ N cn (ii) Suppose (f )n∈N is a sequence such that f Ln( ) for every n and (f ) converges to 2 \⊗ \⊗ some f strongly in L ([0, ∞)). Then, for every k ≥ 2, the sequences ((f n) k) and ((f n) k1 c ) ∂k converge to f ⊗k strongly in L2([0, ∞)k). Proof. (i) We decompose,

\n,k k \n,k dn,k dn,k k kf 1 c − f k 2 ∞ k ≤ kf 1 c − f k 2 ∞ k + kf − f k 2 ∞ k . ∂k L ([0, ) ) ∂k L ([0, ) ) L ([0, ) ) The second term goes to zero by assumption. The first one equals ! Z 1/2 | n,k d e d e |2 f ( nu1 ,..., nuk ) 1{|{dnu1e,...,dnuke}|

lim 1{|{dnu e,...,dnu e}|

(ii) As tensor powers commute with discretization and embedding, i.e. ⊗ ⊗k \ (ˇgn) k = ((g)ˇ⊗k)n, hcn = (hn)⊗k ∈ N ∈ E n ∈ 2 N for all k , g , h Ln( ), and as the tensor product is continuous, we observe inductively \ ⊗ that (f n)⊗k → f k strongly in L2([0, ∞)k). Then, for the second sequence, part (i) applies.  We now start to prepare the proof of Theorem 28. n,k n,k Proposition 32. Let k ∈ N0. Then, for all g ∈ E and sequences (f )n∈N such that f ∈ f2 k n,k L (N ) and sup ∈N kf kL2 (N) < ∞, n n n d lim (SnIn,k(f n,k))(ˇgn) − (SIk(f n,k))(g) = 0. n→∞ Proof. First note that, by (34), (37), and as f n,k vanishes on the diagonal ∂ , h i k  ⊗ n n,k n,k n E n,k n,k n n n h n,k n ki (S I (f ))(ˇg ) = I (f ) exp (I (ˇg )) = f , (ˇg ) L2 (Nk) Z n dn,k \n ⊗k = f (x1, . . . , xk)(ˇg ) (x1, . . . , xk)dx1 ··· dxk. [0,∞)k Analogously, making use of the Wiener-Itˆoisometry for the continuous chaos decomposition (31) instead of (34), we get Z k dn,k dn,k ⊗k (SI (f ))(g) = f (x1, . . . , xk)g (x1, . . . , xk)dx1 ··· dxk. [0,∞)k Hence, by the Cauchy-Schwarz inequality, we conclude

d (SnIn,k(f n,k))(ˇgn) − (SIk(f n,k))(g)

Z  

dn,k \n ⊗k − ⊗k ··· = f (x1, . . . , xk) (ˇg ) g (x1, . . . , xk)dx1 dxk [0,∞)k   1/2 m,k ⊗k \n ⊗k ≤ sup kf k 2 N kg − (ˇg ) k 2 k , Ln( ) L ([0,∞) ) m∈N which tends to zero for n → ∞ by Lemma 31.  Corollary 33. Suppose g ∈ E. Then, for every k ∈ N, n,k n ⊗k k ⊗k I ((ˇg ) 1 c ) → I (g ) ∂k strongly in L2(Ω, F,P ). Proof. We check item (ii) in Theorem 1. To this end, we decompose, for every g, h ∈ E,

n n,k n ⊗k ˇn k ⊗k (S I ((ˇg ) 1∂c ))(h ) − (SI (g ))(h) k

n n,k n ⊗k ˇn k n\⊗k ≤ (S I ((ˇg ) 1∂c ))(h ) − (SI ((ˇg ) 1∂c ))(h) k k k \⊗ k ⊗k + (SI ((ˇgn) k1 c ))(h) − (SI (g ))(h) . ∂k The first term on the righthand side tends to zero by Proposition 32. The second one equals, by the isometry for multiple Wiener integrals, Z   ⊗k \⊗ ⊗k h (x) (ˇgn) k1 c − g (x)dx ∂k [0,∞)k and goes to zero by Lemma 31. Consequently, n n,k n ⊗k ˇn k ⊗k lim (S I ((ˇg ) 1∂c ))(h ) = (SI (g ))(h) n→∞ k 26 C. BENDER AND P. PARCZEWSKI

n,k n ⊗k 2 k ⊗k 2 for all k ∈ N and g, h ∈ E. For h = g, this implies E[I ((ˇg ) 1 c ) ] → E[I (g ) ] by 0 ∂k the orthogonality of (discrete) multiple Wiener integrals of different orders. Thus, Theorem 1 applies. 

We are now in the position to give the proof of Theorem 28.

n 2 Proof of Theorem 28. ‘(i) ⇒ (ii)’: We denote the limit of (πHn X )n∈N in L (Ω, F,P ) by X and recall that X X∞ n E n n n n,k n,k πHn X = [X ΞA]ΞA = I (fXn ), |A|<∞ k=0 n,k with f n as defined in (36). Throughout the proof we omit the subscripts from the coeffi- X P∞ P∞ n n,k n,k k k cients and write πHn X = I (f ) and X = I (f ). Thanks to Corollary 33 and the k=0 k=0 orthogonality of (discrete) multiple Wiener integrals of different orders, we obtain, for every k ∈ N0,

n n,k n,k n 1 n n,k n ⊗k 1 k ⊗k k k (S I (f ))(ˇg ) = E[πHn (X )I ((ˇg ) 1∂c )] → E[XI (g )](SI (f ))(g). k! k k! E n,k n,k 2 ≤ E n 2 ∞ The estimate supn∈N [(I (f )) ] supn∈N [(πHn X ) ] < now yields, in view of The- 2 n,k n,k k k n orem 1, weak L (Ω, F,P )-convergence of (I (f ))n∈N towards I (f ). As πHn X → X strongly in L2(Ω, F,P ), we thus obtain n,k n,k 2 n,k n,k n k k k k 2 E[(I (f )) ] = E[I (f )πHn X ] → E[I (f )X] = E[(I (f )) ]. (38)

n,k n,k k k 2 Hence, I (f ) → I (f ) strongly in L (Ω, F,P ) for all k ∈ N0 by Theorem 1. Moreover, for every g ∈ E, we obtain

⊗k dn,k k k dn,k k k hg , f − f i 2 k = (SI (f ))(g) − (SI (f ))(g)  L ([0,∞) )  d = (SIk(f n,k))(g) − (SnIn,k(f n,k))(ˇgn) h i   + E In,k(f n,k) exp n (In(ˇgn)) − Ik(f k) exp (I(g)) → 0,

2 n,k n,k k k by Propositions 2 and 32, and the L (Ω, F,P )-convergence of (I (f ))n∈N to I (f ). Since ⊗ f d the set {g k, g ∈ E} is total in L2([0, ∞)k), we may conclude that (f n,k) converges weakly in f L2([0, ∞)k) to f k by [29, Theorem V.1.3]. Finally, (38) and the isometry for discrete multiple Wiener integrals turn this weak convergence into strong L2([0, ∞)k)-convergence. In particular, the kth coefficient in the chaos decomposition of the limiting random variable X is the strong d L2([0, ∞)k)-limit of (f n,k), as asserted. It remains to show (33). However, by (38) and the isometries for (discrete) multiple Wiener integrals, X∞ X∞ k dn,kk2 k n,k n,k k2 lim k! f 2 ∞ k = lim I (f ) 2 F n→∞ L ([0, ) ) n→∞ L (Ω, ,P ) k=m k=m ! mX−1 n 2 n,k n,k 2 = lim kπHn X k 2 F − kI (f )k 2 F n→∞ L (Ω, ,P ) L (Ω, ,P ) k=0 mX−1 k k2 − k k k k2 → = X L2(Ω,F,P ) I (f ) L2(Ω,F,P ) 0 k=0 as m tends to infinity. ⇒ n,k ‘(ii) (i)’: In order to lighten the notation, we again denote the function fXn from (36) by d f n,k. Assuming (ii), the strong L2([0, ∞)k)-limit of f n,k exists and will be denoted f k. We first DISCRETIZING MALLIAVIN CALCULUS 27

n,k n,k k k 2 show that (I (f )) converges to I (f ) strongly in L (Ω, F,P ) for all k ∈ N0 by means of Theorem 1. To this end, we observe that, for every g ∈ E,   d d (Sn In,k(f n,k))(ˇgn) = (Sn In,k(f n,k))(ˇgn) − (SIk(f n,k))(g) + (SIk(f n,k))(g) → (SIk(f k))(g) by Proposition 32 and the isometry for continuous multiple Wiener integrals. Moreover, again, by the isometries for discrete and continuous multiple Wiener integrals, h i h i E n,k n,k 2 k n,kk2 k dn,kk2 → k kk2 E k k 2 (I (f )) = k! f 2 Nk = k! f 2 ∞ k k! f 2 ∞ k = (I (f )) . Ln( ) L ([0, ) ) L ([0, ) ) So, Theorem 1 applies indeed. With the L2(Ω, F,P )-convergence of In,k(f n,k) to Ik(f k) at hand, we can now decompose, for every m ∈ N,   X∞ 2  n k k  lim sup E πHn X − I (f ) n→∞ k=0

mX−1 mX−1 ≤ k k k − n,k n,k k2 3 lim sup I (f ) I (f ) L2(Ω,F,P ) n→∞ k=0 k=0 ! X∞ X∞ k k k k2 k n,k n,k k2 + I (f ) L2(Ω,F,P ) + I (f ) L2(Ω,F,P ) k=m k=m X∞ X∞ k kk2 k dn,kk2 = 3 k! f L2([0,∞)k) + 3 lim sup k! f L2([0,∞)k). (39) n→∞ k=m k=m

dn,k k By Fatou’s lemma and the strong convergence of (f )n∈N to f , X∞ X∞ k kk2 ≤ k dn,kk2 k! f 2 ∞ k lim inf k! f 2 ∞ k . L ([0, ) ) n→∞ L ([0, ) ) k=m k=m n Hence, letting m tend to infinity in (39), we observe, thanks to (33), that (πHn X ) converges strongly in L2(Ω, F,P ).  We close this section with an example. Example 34. Fix X ∈ L2(Ω, F,P ). Theorem 28 with Xn = X for every n ∈ N, implies that k 2 k the chaos coefficients f , k ∈ N0, of X are given as the strong L ([0, ∞) )-limit of X " !# Yk Bn − Bn d 1 dnule (dnule−1) f n,k(u , . . . , u ) := E X 1{|{d e d e}∩N| }. 1 k k! 1/n nu1 ,..., nuk =k l=1

This formula can be further simplified when X is FT -measurable. Then, one can show, analo- 2 gously to Example 4 (ii), that the sequence (πHn X) converges to X strongly in L (Ω, F,P ). bnT c k ∈ N Applying Theorem 28 with the latter sequence, shows that the chaos coefficients fX , k 0, are the strong L2([0, ∞)k)-limit of " !# Yk Bn − Bn d 1 dnule (dnule−1) f n,k(u , . . . , u ) := E X 1{|{d e d e}∩{ b c}| }. 1 k k! 1/n nu1 ,..., nuk 1,..., nT =k l=1

dn,k In this case, for each fixed n ∈ N, only finitely many of the functions f , k ∈ N0, are not constant zero, and these are simple functions with finitely many steps sizes only. These two approximation formulas for the chaos coefficients of X are one way to give a rigorous meaning of the heuristic formula " !# 1 Yk f k (u , . . . , u ) = E X B˙ , X 1 k k! ul l=1 28 C. BENDER AND P. PARCZEWSKI where B˙ is white noise, which is called Wiener’s intuitive recipe in [9]. The latter paper provides another rigorous meaning to Wiener’s recipe via nonstandard analysis, which is closely related to our approximation formulas in the special case of symmetric Bernoulli noise. The authors show that    1 ∆b ∆b f k (◦t ,..., ◦t ) = ◦E x(b) t1 ··· tk , X 1 k k! ∆t ∆t √ P ∈ { ≤ 2} ∈ tl T = j∆t, 0 j < N , where N is infinite, ∆t = 1/N, bt(ω) = ∆t s

6. Strong L2-approximation of the Skorokhod integral and the Malliavin derivative In this section, we apply the Wiener chaos limit theorem (Theorem 28) in order to prove strong L2-approximation results for the Skorokhod integral and the Malliavin derivative. For the con- struction of the approximating sequences we compose the discrete Skorokhod integral and the discretized Malliavin derivative with the orthogonal projection on Hn, i.e. on the subspace of random variables which admit a discrete chaos decomposition in terms of multiple integrals with n respect to the discrete time noise (ξi )i∈N. We first treat the Malliavin derivative and aim at proving the following result. n 2 Theorem 35. Suppose (X )n∈N converges strongly in L (Ω, F,P ) to X and, for every n ∈ N, n 1,2 πHn X ∈ Dn . Then the following are equivalent: P∞ d d (i) lim lim sup kk!kf n,kk2 = 0 (with f n,k as defined in (32)). →∞ Xn L2([0,∞)k) Xn m n→∞ k=m ∈ D1,2 n n 2 × (ii) X and the sequence (Ddn·e(πHn X ))n∈N converges to DX strongly in L (Ω [0, ∞)) as n tends to infinity. n ∈ N Note first, that by continuity of Di for a fixed time i , we get X √ X n n E n n n n E n n n Di (πHn X ) = [X ΞA]Di ΞA = n [X ΞA]ΞA\{i} |A|<∞ |A|<∞; i∈A √ X E n n n = n [X ΞB∪{i}]ΞB. |B|<∞; i/∈B By the relation (35)–(36) between Walsh decomposition and discrete chaos decomposition, this identity can be reformulated as X∞ n n n,k−1 n,k · Di (πHn X ) = kI (fXn ( , i)). k=1 Hence, the isometry for discrete multiple Wiener integrals (34) implies X∞ h i X∞ 1 n n 2 dn,k 2 E |D (πHn X )| = kk!kf k , (40) n i Xn L2([0,∞)k) i=1 k=1 i.e., ∞ X d n ∈ D1,2 ⇔ k n,kk2 ∞ πHn X n kk! fXn L2([0,∞)k) < . k=1 This is in line with the characterization of the continuous Malliavin derivative in terms of the chaos decomposition, see e.g. [23], which we show to be equivalent to Definition 11 in the Appendix: X∞ ∈ D1,2 ⇔ k k k2 ∞ X kk! fX L2([0,∞)k) < , (41) k=1 DISCRETIZING MALLIAVIN CALCULUS 29 and, if this is the case, X∞ n,k−1 k · ≥ DtX = kI (fX ( , t)), a.e. t 0, (42) k=1 Z ∞ X∞ E 2 k k k2 [(DtX) ]dt = kk! fX L2([0,∞)k). 0 k=1 After these considerations on the connection between (discretized) Malliavin derivative and (discrete) chaos decomposition, the proof of Theorem 35 turns out to be rather straightforward.

Proof of Theorem 35. By Theorem 28 (in conjunction with Remark 26), we observe that, for dn,k k 2 k every k ∈ N ,(f ) ∈N converges to f strongly in L ([0, ∞) ). Hence, by (40), (41), and (42), 0 Xn n X ∞ ∞ X d X ⇔ k n,kk2 k k k2 ∞ (i) lim kk! f n 2 ∞ k = kk! f 2 ∞ k < n→∞ X L ([0, ) ) X L ([0, ) ) k=1 k=1 X∞ h i Z ∞ 1,2 1 n n 2 2 ⇔ X ∈ D and lim E |D (πHn X )| = E[(DtX) ]dt. n→∞ n i i=1 0 Hence, the asserted equivalence is a direct consequence of Theorem 12. 

We now wish to derive an analogous strong approximation result for the Skorokhod integral, n 2 which requires some additional notation. For every Z ∈ L (Ω × N) and k ∈ N0, we denote ( h n i nk/2 n n n,k n,k E Z Ξ , |{i1, . . . , ik} ∩ N| = k k! i {i1,...,ik} fZn (i1, . . . , ik, i) := fZn (i1, . . . , ik) = i 0, otherwise. n n Then, with πHn Z := (πHn Zi )i∈N, X∞ k n,kk2 k nk2 ∞ k! f n 2 Nk+1 = πHn Z 2 ×N < , Z Ln( ) Ln(Ω ) k=0 n,k but fZn is symmetric in the first k variables only and does not, in general, vanish on the diagonal. For a function F in k variables, we denote its symmetrization by 1 X Fe(y , . . . , y ) = F (y , . . . , y ), 1 k k! π(1) π(k) π en,k where the sum runs over the group of permutations of {1, . . . , k}. With this notation, f n 1 c Z ∂k+1 f2 Nk+1 is an element of Ln( ). We can now state: ∈ N n ∈ 2 × N n ∈ n Theorem 36. Suppose that, for every n , Z Ln(Ω ) and πHn Z D(δ ). Moreover, n 2 × ∞ assume that (Zdn·e)n∈N converges to Z strongly in L (Ω [0, )). Then, the following assertions are equivalent: P∞ en,k−1 2 (i) lim lim sup k!kf n 1 c k = 0. →∞ Z ∂k L2 (Nk) m n→∞ k=m n n n 2 (ii) Z ∈ D(δ) and (δ (πHn Z )) converges to δ(Z) strongly in L (Ω, F,P ) as n tends to infinity. As a preparation of the proof we note that, for every M ∈ N,

XM n XM X n n n ξi n n n n ξi E[πHn Z |F − ]√ = E[Z Ξ ] E[Ξ |F − ]√ i M, i n i A A M, i n i=1 i=1 |A|<∞ 30 C. BENDER AND P. PARCZEWSKI

XM X −1/2 E n n n = n 1{i/∈A} [Zi ΞA]ΞA∪{i} i=1 A⊂{1,...,M} XM X X −1/2 E n n n = n [Zi ΞB\{i}]ΞB k=1 B⊂{1...,M},|B|=k i∈B XM X Xk −1/2 ⊗k 1 n n n = n k 1 (i1, . . . , ik) E[Z Ξ{ }\{ }]Ξ{ } [1,M] k ij i1,...,ik ij i1,...,ik k k=1 (i1,...,ik)∈N , j=1 i1<···

dn,k k variables, we conclude again that fZn converges weakly to fZ in this subspace. Hence, it only remains to argue that

d 2 2 n,k → k → ∞ f n f , n . Z Z 2 ∞ k+1 L2([0,∞)k+1) L ([0, ) ) As X∞ X∞ 1 E n n,k n ⊗k ˇn 1 n n,k n,k n ˇn [Zi I ((ˇg ) 1∂c )]h (i) = (S I (fZn ))(ˇg )h (i), n k n i i=1 i=1 Z ∞ Z ∞ E k ⊗k k k [ZsI (g )]h(s)ds = (SI (fZs ))(g)h(s)ds, 0 0 n,k n,k k k we may derive from (45)–(46) and Theorem 10, that I (f n ) converges to I (f · ) weakly in Zdn·e Z L2(Ω × [0, ∞)). Thus, Z Z 2 ∞ h i ∞ h i dn,k n,k n,k n,k n,k n f n = E I (f n )Zs ds + E I (f n )(Zd e − Zs) ds Z Zd e Zd e ns L2([0,∞)k+1) 0 ns 0 ns Z h i ∞ 2 → E k k k I (fZs )Zs ds = fZ . 0 L2([0,∞)k+1)  c Proof of Theorem 36. By the linearity of the embedding operator (·), Minkowski inequality, Proposition 37, and Lemma 31, we obtain, for every k ∈ N , 0

en,k\ ek bn,k^ ek n,k\ k f n 1 c − f = f n 1 c − f ≤ f n 1 c − f → 0 Z ∂k+1 Z Z ∂k+1 Z Z ∂k+1 Z L2([0,∞)k+1) L2([0,∞)k+1) L2([0,∞)k+1) as n tends to infinity. Thus, due to Theorem 28 and (44), ⇔ n n 2 F (i) (δ (πHn Z ))n∈N converges strongly in L (Ω, ,P ). Now, the implication ‘(ii) ⇒ (i)’ is obvious, while the converse implication is a consequence of Theorem 8.  Remark 38. As a by-product of the proof of Theorem 36, we recover, thanks to Theorem 28, the well-known chaos decomposition of the Skorokhod integral as X∞ k ek−1 δ(Z) = I (fZ ). k=1 7. Examples In this section, we first specialize the previous results to the case of binary noise in discrete time. Then, we explain how to translate the main results of this paper into the language of convergence in distribution without imposing the condition that the discrete-time noise is embedded into the driving Brownian motion. 7.1. Binary noise. In this subsection, we specialize to the case of binary noise, i.e., we suppose that, for some constant b > 0, b2 1 P ({ξ = −1/b}) = ,P ({ξ = b}) = . b2 + 1 b2 + 1 We illustrate, that in this binary case, our approximation formulas for the Malliavin derivative and the Skorokhod integral give rise to a straightforward numerical implementation. We recall first that Malliavin calculus on the Bernoulli space is well-studied, see, e.g. [12], [21], [25], and the references therein, usually with the aim to explain the main ideas of Malliavin calcu- 2 F n lus by discussing the analogous operators in the simple toy setting. Note first that L (Ω, i ,P ) 32 C. BENDER AND P. PARCZEWSKI

Hn equals i in the binary case (and in this case only) by observing that both spaces have dimen- sion 2i in this case. Hence, L2(Ω, F n,P ) coincides with Hn for binary noise, and we can drop n the orthogonal projections πHn on H in the statement of all previous results. In particular, every random variable Xn ∈ L2(Ω, F n,P ) then admits a chaos decomposition in terms of dis- crete multiple Wiener integrals, and the representations of the discretized Malliavin derivative and the discrete Skorokhod integral in terms of the discrete chaos in Section 6 show that these operators coincide with the Malliavin derivative and the Skorokhod integral on the Bernoulli space, see [25]. In the binary case, the representations for the discrete Malliavin derivative and the discrete Skorokhod integral can be simplified considerably. Suppose Xn ∈ L2(Ω, F n,P ). Then, there is R∞ → R n n n a measurable map FXn : such that X = FXn (ξ1 , ξ2 ,...). A direct computation shows that, for every i ∈ N, √ n n n n n D X = nE[ξ F n (ξ , ξ ,...)|F ] i √ i X 1 2 −i  nb n n n n n n = F n (ξ , . . . , ξ , b, ξ ,...) − F n (ξ , . . . , ξ , −1/b, ξ ,...) , b2 + 1 X 1 i−1 i+1 X 1 i−1 i+1 n ∈ 2 × N hence, the Malliavin derivative becomes a difference operator. Moreover, for Z Ln(Ω ) and N ∈ N, the discrete Skorokhod integral can be rewritten as XN ξn 1 XN δn(Zn1 ) = Zn √i − (ξn)2DnZn, [1,N] i n n i i i i=1 i=1 n which can either be derived from [25, Proposition 1.8.3] or by expanding Zi in its Walsh de- composition and noting that, for every finite subset A ⊂ N,  √   √ nΞn (ξn)2, i ∈ A Ξn − E[Ξn |F n ] nξn = A\{i} i = (ξn)2DnΞn . A A −i i 0, i∈ / A i i A n ∈ 2 × N ∈ N Hence, for Z Ln(Ω ) and N , XN n n n n n ξi δ (Z 1 ) = FZn (ξ , ξ ,...)√ [1,N] i 1 2 n i=1 XN n 2  (ξi ) b n n n n n n − √ FZn (ξ , . . . , ξ − , b, ξ ,...)FZn (ξ , . . . , ξ − , −1/b, ξ ,...) . (47) n(b2 + 1) i 1 i 1 i+1 i 1 i 1 i+1 i=1 n Recall that the discrete noise (ξi )i∈N, can be constructed from the underlying Brownian motion (Bt)t∈[0,∞) via a Skorokhod embedding as √   n ξ = n B n − B n , i τi τi−1 where, in the binary case,    − n n n b 1 τ := 0 , τ := inf s ≥ τ − : Bs − Bτ n ∈ √ , √ , (48) 0 i i 1 i−1 n b n n and the Brownian motion at the first-passage times τi can be simulated by the acceptance- rejection algorithm of Burq&Jones [7]. We close this paper by a toy example which illustrates how to numerically compute Skorokhod integrals by our approximation results. Example 39. In this example, we approximate the Skorokhod integral δ(Z) for the process

Zt = sign(1/2 − t)(B1B1−t − (1 − t)))1[0,1](t), t ≥ 0, where we choose the sign-function to be rightcontinuous at 0. For the discrete time approximation we consider  n − n n − − ∈ N Zi = sign(1/2 i/n) Bn Bn−i (1 i/n) 1[1,n−1](i), i , DISCRETIZING MALLIAVIN CALCULUS 33

n ≥ and note that (Zdnte) converges to Zt for almost every t 0 in probability by (1). Hence, by n uniform integrability and dominated convergence, it is easy to check that (Zdn·e)n∈N converges to Z strongly in L2(Ω × [0, ∞)). We next observe that in the discrete chaos decomposition of n n n,k ≥ n δ (Z ), all the coefficient functions fδn(Zn) for k 4 vanish, because Zi is a polynomial of degree 2 in Bn. Hence, the tail condition in Theorem 36 is trivially satisfied and, consequently, n n 2 n (δ (Z ))n∈N converges to δ(Z) strongly in L (Ω, F,P ). We now suppose that B is constructed via the Skorokhod embedding (48) and simulate, for n = 4, 8,..., 215, 10000 independent copies n,l n n n (B )l=1,...,10000 of B by the Burq&Jones algorithm. The correponding realizations of δ (Z )

0 10

−1 10

−2 10

1 2 3 4 10 10 10 10 Figure 1. Log-log plot of the simulated strong L2(Ω, F,P )-approximation as the number of time steps increases.

n n and δ(Z) along the lth trajectory of the underlying Brownian motion are denoted δl (Z ) and δl(Z), l = 1,... 10000, respectively. For the discrete Skorokhod integral we implement formula (47) with N = n, while for the continuous Skorokhod integral we exploit that it can be computed analytically and equals B δ(Z) = B B2 − 1 − B . 1 1/2 2 1/2 Figure 1 shows, in the case of symmetric binary noise (b = 1), a log-log-plot of the empirical mean | n n − |2 (indicated by crosses) of δl (Z ) δl(Z) , l = 1,..., 10000, and the corresponding (asymptotical) 95%-confidence bounds (indicated by dots) as the number of time steps n increases. A linear regression (solid line) exhibits a slope of −0.5036 and, thus, indicates that strong L2(Ω, F,P )- convergence takes place at the expected rate of 1/2. 7.2. Donsker type theorems. In this subsection, we explain how the main results of the paper can be translated into Donsker type results on convergence in distribution, without assuming that the discrete time noise is embedded into the driving Brownian motion B. To this end, we construct, without loss of generality, the discrete time noise as coordinate mapping on the R∞ B∞ ∞ ∞ canonical space ( , ,Pξ ), where Pξ is the distribution of ξ and Pξ its countable product ∞ · E · measure. We denote expectation with respect to Pξ by E[ ], while [ ] still denotes expectation on (Ω, F,P ), i.e. on the probability space which carries the driving BrownianP motion B. We 2 R∞ × N N → 2 R∞ B∞ ∞ ∞ 2 ∞ write L ( ) for the set of mappings z : L ( , ,Pξ ) such that i=1 E[zi ] < . 34 C. BENDER AND P. PARCZEWSKI

∈ N 7→ n We also note that, for every n , the map F F ((ξi )i∈N) provides a one-to-one corre- 2 R∞ B∞ ∞ 2 F n spondence between L ( , ,Pξ ) and L (Ω, ,P ) thanks to the Doob-Dynkin lemma. The Pb c R∞ B∞ ∞ n √1 nt rescaled random walk approximation to B on ( , ,Pξ ) is given by bt (x) = n j=1 xi, n n n and, hence, Bt = bt ((ξi )i∈N). The translation of the strong L2-approximation results derived in the previous sections into results on convergence in distribution relies on the following key lemma. Here, D([0, ∞)) de- notes the space of rightcontinuous functions with left limits from [0, ∞) to R endowed with the Skorokhod topology. n 2 R∞ B∞ ∞ ∈ 2 F Lemma 40. a) Suppose (F )n∈N is a sequence in L ( , ,Pξ ) and X L (Ω, ,P ). Then, the following are equivalent: n n 2 F (i) The sequence (F ((ξi )i∈N))n∈N converges to X strongly in L (Ω, ,P ). n n (ii) The sequence (F , b )n∈N converges in distribution to (X,B) in R × D([0, ∞)) and the | n|2 ∞ sequence ( F )n∈N is uniformly integrable (with respect to Pξ ). n 2 ∞ 2 b) Suppose (z )n∈N is a sequence in L (R × N) and Z ∈ L (Ω × [0, ∞)). Then, the following are equivalent: n n 2 × ∞ (i) The sequence (zdn·e((ξi )i∈N))n∈N converges to Z strongly in L (Ω [0, )). n n 2 (ii) The sequence (z , b ) ∈N converges in distribution to (Z,B) in L ([0, ∞)) × D([0, ∞)) dn·e Pn −1 n n 2 ∞ and the sequence (n i=1(zi ) )n∈N is uniformly integrable (with respect to Pξ ). Proof. We first prove b). We wish to apply Theorem 44 in Appendix B with E = D([0, ∞)), 2 n n n H = L ([0, ∞)), L = B, and L = B . To this end, we first note that the sequence (Bbntc/n−B ) converges to zero in finite-dimensional distributions by (1). It is easy to check by standard criteria such as [5, Theorem 15.6] that this sequence is tight in D([0, ∞)). Hence, this sequence converges to zero in distribution in D([0, ∞)). Thus, by the (generalized) continuous mapping theorem [5, | − n| Theorem 5.5], supt∈[0,K] Bbntc/n Bt converges to zero in distribution (and consequently in probability) for every K ∈ N. Then, by continuity of B,(Bn) converges to B uniformly on compacts in probability. In particular, the Skorokhod distance between Bn and B converges to zero in probability. Moreover, L2(Ω × [0, ∞)) = L2(Ω, F,P ; L2([0, ∞))) by [16, Appendix C]. Therefore, Theorem 44 is indeed applicable and yields the equivalence of (i) and n n n 2 (ii’) The sequence (z ((ξ ) ∈N),B ) ∈N converges in distribution to (Z,B) in L ([0, ∞)) × dn·e j j nP D ∞ −1 n | n n |2 ([0, )) and the sequence (n i=1 zi ((ξj )j∈N) )n∈N is uniformly integrable (with respect to P ). n n n n n ∞ Since (zdn·e((ξj )j∈N),B ) has the same distribution under P as (zdn·e, b ) has under Pξ , the equivalence of (ii) and (ii’) is immediate. The proof of a) is the very same, but with H = R.  As an example, we now explain how to reformulate two of the main convergence results for the Malliavin derivative (Theorems 12 and 16) in the present setting. We leave it to the reader to rewrite Theorem 35 and the corresponding theorems for the Skorokhod integral in the obvious way. For F ∈ L2(R∞, B∞,P ∞) and i ∈ N, we denote ξ Z √ Dn i F := n xiF (x1, . . . , xi,...)Pξ(dxi), R Dn n n n whence, ( i F )((ξj )j∈N) = Di F ((ξj )j∈N). n 2 R∞ B∞ ∞ | n|2 Theorem 41. Suppose (F )n∈N is a sequence in L ( , ,Pξ ) such that ( F )n∈N is uni- n n formly integrable and (F , b )n∈N converges in distribution to (X,B) in R × D([0, ∞)) for some X ∈ L2(Ω, F,P ). Then: P −1 ∞ Dn n 2 ∈ D1,2 a) If the sequence (n i=1( i F ) )n∈N is uniformly integrable, then X and the fol- lowing are equivalentP as n tends to infinity: R −1 ∞ Dn n 2 ∞ E 2 (i) (n i=1 E[( i F ) ])n∈N converges to 0 [(DsX) ]ds. DISCRETIZING MALLIAVIN CALCULUS 35

∈ 2 × ∞ Dn n n (ii) There is a Z L (Ω [0, )) such that ( dn·eF , b )n∈N converges in distribution to (Z,B) in L2([0, ∞)) × D([0, ∞)). Dn n n 2 ∞ × D ∞ (iii) ( dn·eF , b )n∈N converges in distribution to (DX,B) in L ([0, )) ([0, )). b) If   X∞   X∞    1 Dn n 2 1 DnDn n 2  ∞ sup E ( i F ) + E ( j i F ) < , ∈N n n2 n i=1 i,j=1, i=6 j P −1 ∞ Dn n 2 Dn n n then (n i=1( i F ) )n∈N is uniformly integrable and ( dn·eF , b )n∈N converges in distribu- tion to (DX,B) in L2([0, ∞)) × D([0, ∞)). Remark 42. The general assumptions of the previous theorem are, in particular, satisfied, when X = g(B) and F n = g(bn), n ∈ N, for some bounded and continuous function g : D([0, ∞)) → R. n n n Proof of Theorem 41. We write X := F ((ξi )i∈N). Note first, that the assumptions guarantee, n 2 in view of Lemma 40 a), that (X )n∈N converges to X strongly in L (Ω, F,P ). a) Uniform integrability implies norm-boundedness. Hence, X∞   X∞   1 E n n 2 1 Dn n 2 ∞ sup (Di X ) = sup E ( i F ) < . ∈N n ∈N n n i=1 n i=1 ∈ D1,2 n n 2 × ∞ Thus, by Theorem 12, X and (Ddn·eX ) converges to DX weakly in L (Ω [0, )). ‘(i)⇒ (iii)’: As ∞ 1 X kDn Xnk2 = E[(DnF n)2], (49) dn·e L2(Ω×[0,∞)) n i i=1 2 × ∞ n n condition (i) turns weak L (Ω [0, ))- convergence of (Ddn·eX ) to DX into strong convergence. Hence, Lemma 40 b) concludes. ⇒ n n ‘(iii) (i)’: Assuming (iii), we obtain from Lemma 40 b), that (Ddn·eX ) converges to DX strongly in L2(Ω × [0, ∞)). Hence, the sequence of L2(Ω × [0, ∞))-norms converges as well, which thanks to (49) implies (i). ⇒ n n 2 × ∞ ‘(ii) (iii)’: If (ii) holds, (Ddn·eX ) converges to Z strongly in L (Ω [0, )) by Lemma 40 b). The strong L2(Ω × [0, ∞))-limit Z must, of course, coincide with the weak L2(Ω × [0, ∞))-limit DX, which establishes (iii). ‘(iii)⇒ (ii)’: obvious. b) This assertion is an immediate consequence of Lemma 40 and Theorem 16, ‘(i) ⇒ (ii)’. 

Appendix A. S-transform characterization of the Malliavin derivative In this appendix, we prove the equivalence between the definition of the Malliavin derivative in terms of the S-transform (Definition 11) and the more classical characterization in terms of the chaos decomposition, see (41)–(42). P k k ∈ 2 F Proposition 43. Suppose X = k I (fX ) L (Ω, ,P ). Then, the following are equivalent: (i) There is a stochastic process Z ∈ L2(Ω × [0, ∞)) such that, for every g, h ∈ E, Z ∞   Z ∞   (SZs)(g)h(s)ds = E X exp (I(g)) I(h) − g(s)h(s)ds . 0 0 P∞ k k k2 ∞ (ii) kk! fX L2([0,∞)k) < . k=1 P∞ n,k−1 k · ≥ If this is the case, then Zt = kI (fX ( , t)) for almost every t 0. k=1 36 C. BENDER AND P. PARCZEWSKI

Proof. We first note that, for every f, g ∈ E,   Z ∞ X∞  1 ^ e I(g) I(h) − g(s)h(s)ds = Ik((g⊗(k−1) ⊗ h)), (50) (k − 1)! 0 k=1 which can be verified by computing the S-transform of both sides. By the Cauchy-Schwarz inequality, we obtain for every f, g ∈ E, ∞ Z X k ⊗k−1⊗ kfX (x)(g h)(x) dx ∞ k k=1 [0, ) ! 1 ! 1 X∞ 2 X∞ 2 k − ≤ k!kf k k2 kgk2(k 1) khk2 < ∞. X L2([0,∞)k) (k − 1)! L2([0,∞)) L2([0,∞)) k=1 k=1 Hence, Fubini’s theorem implies ! X∞ Z Z ∞ X∞ Z k ⊗k−1 ⊗ k ⊗k−1 kfX (x)(g h)(x)dx = kfX (x, t)g (x) h(t)dt, ∞ k ∞ k−1 k=1 [0, ) 0 k=1 [0, ) i.e., by (50) and the isometry for multiple Wiener integrals,    ! Z ∞ Z ∞ X∞ Z E  − k ⊗k−1 X exp (I(g)) I(h) g(s)h(s)ds = kfX (x, t)g (x) h(t)dt ∞ k−1 0 0 k=1 [0, ) (51) for every g, h ∈ E. ‘(i) ⇒ (ii)’: Assuming (i) and noting that (51) holds for every g, h ∈ E, we observe that for every g ∈ E, α ∈ R, and Lebesgue-almost every s ∈ [0, ∞), X∞ X∞ k−1 k−1 ⊗(k−1) k−1 k ⊗(k−1) α hf (·), g i 2 k−1 = (SZ )(αg) = α hkf (·, s), g i 2 k−1 . Zs L ([0,∞) ) s X L ([0,∞) k=1 k=1 (Note, that the Lebesgue null set can be chosen independent of g, α. Indeed, one can first take α ∈ Q and step functions g with rational step sizes and interval limits, and then pass to the limit). Comparing the coefficients in the power series and noting that {g⊗k, g ∈ E} is total in f L2([0, ∞)k), we obtain, for every k ≥ 1 and almost every s ∈ [0, ∞),

kf k (·, s) = f k−1. X Zs Therefore, the isometry for multiple Wiener-Itˆointegrals implies X∞ Z ∞ k k k2 E | |2 ∞ kk! fX L2([0,∞)k) = [ Zs ]ds < . (52) k=1 0 P∞ ⇒ n,k−1 k · ‘(ii) (i)’: Define Zt = kI (fX ( , t)). Assuming (ii), we observe by the first identity in k=1 (52) that Z belongs to L2(Ω × [0, ∞)). By the isometry for multiple Wiener integrals and the chaos decomposition of a Wick exponential we get, for every g, h ∈ E. ! Z ∞ Z ∞ X∞ Z k ⊗k−1 (SZs)(g)h(s)ds = kfX (x, t)g (x)dx h(t)dt. ∞ k−1 0 0 k=1 [0, ) Hence, (51) concludes.  DISCRETIZING MALLIAVIN CALCULUS 37

Appendix B. On the connection between strong L2-convergence and convergence in distribution In this appendix, we prove, that, under suitable conditions, strong L2-convergence is completely specified by convergence in distribution of appropriate joint distributions in conjunction with uniform integrability. This is the essential tool for the results in Section 7.2. We make use of the following notations and assumptions:

• (Ω, F,P ) is a probability space, (E, dE) is a separable metric space, BE denotes the Borel σ-field on E and (H, k · kH) is a separable real Hilbert space. • L : (Ω, F) → (E, BE) is a measurable map. • Ln : (Ω, F) → (E, BE) for all n ∈ N, such that Ln is σ(L)-measurable and the sequence (Ln)n∈N satisfies dE(L, Ln) → 0 in probability as n tends to infinity. • L2(Ω, σ(L),P ; H) is the space of H-valued, σ(L)-measurable and square-integrable ran- dom variables (i.e X : (Ω, σ(L)) → H measurable, where H is endowed with its Borel 2 σ-field, and satisfies E[kXkH] < ∞). The convergence in distribution is denoted by ⇒.

2 Theorem 44. Suppose the notations and assumptions above, X ∈ L (Ω, σ(L),P ; H) and (Xn)n∈N 2 satisfies Xn ∈ L (Ω, σ(L),P ; H) for all n ∈ N. Then the following assertions are equivalent as n tends to infinity: 2 (i) Xn → X strongly in L (Ω, σ(L),P ; H). 2 (ii) (Xn,Ln) ⇒ (X,L) in H × E and (kXnkH)n∈N is uniformly integrable.

Proof. ‘(i) ⇒ (ii)’: Due to the assumptions on the sequence (Ln)n∈N and (i), we have 2 kXn − XkH + dE(Ln,L) → 0 in probability as n tends to infinity. This immediately implies the convergence in distribution in (ii). Moreover, by the Cauchy-Schwarz inequality, we observe for n → ∞, 2 2 2 1/2 2 1/2 E[|kXnkH − kXkH|] = E[|hXn − X,Xn + Xi|] ≤ E[|kXn − XkH] E[|kXn + XkH] → 0. 2 This gives the uniform integrability of the sequence (kXnkH)n∈N. ‘(ii) ⇒ (i)’: We firstly observe by the continuous mapping theorem, as n tends to infinity, 2 2 kXnkH ⇒ kXkH. Hence, the uniform integrability implies 2 2 E[kXnkH] → E[kXkH]. (53) Now we fix a uniformly continuous and bounded (u.c.b., for shorthand) map f : E → R and h ∈ H. The continuous mapping theorem gives that

f(Ln)hXn, hiH ⇒ f(L)hX, hiH. Moreover, by the de la Vall´ee-Poussin criterion, the inequality 2 2 2 2 sup E[|f(Ln)hXn, hiH| ] ≤ (sup |f(e)| )khkH sup E[kXnkH] < ∞ n∈N e∈E n∈N implies the uniform integrability of (hXn, f(Ln)hiH)n∈N. Thus we conclude

E[hXn, f(Ln)hiH] → E[hX, f(L)hiH]. (54) Hence, (53) and (54) equal the expressions in Lemma 3 (ii) for the Hilbert space L2(Ω, σ(L),P ; H) and it suffices to verify the assumptions in Lemma 3, i.e. that (iii) The set {f(L)h : f is u.c.b., h ∈ H} is total in L2(Ω, σ(L),P ; H). 2 (iv) For all u.c.b. f and h ∈ H: f(Ln)h → f(L)h strongly in L (Ω, σ(L),P ; H) as n tends to infinity. 38 C. BENDER AND P. PARCZEWSKI

‘(iii)’: As L2(Ω, σ(L),P ; H) is (a realization of) L2(Ω, σ(L),P )⊗H (tensor product in the sense of Hilbert spaces, cf. [16, Appendix E]), it suffices to show that the set {f(L): f is u.c.b.} is dense in L2(Ω, σ(L),P ). Let X = F (L) ∈ L2(Ω, σ(L),P ) and assume that E[Xf(L)] = 0 for all u.c.b. f. Then Z

F (x)f(x)PL(dx) = 0, E and, as in [5, Theorem 1.3], this implies Z

F (x)PL(dx) = 0, (55) A for every closed set A ⊂ E. Thanks to the regularity of probability measures on metric spaces, this implies (55) for all A ∈ BE. Hence, F = 0 PL-a.s., and therefore X = F (L) = 0 P -a.s. Thus we conclude (iii). ‘(iv)’: Due to dE(Ln,L) → 0 in probability and the continuous mapping theorem, we obtain f(Ln) → f(L) in probability. Hence, the dominated convergence theorem implies 2 2 2 E[kf(Ln)h − f(L)hkH] = E[|f(Ln) − f(L)| ]khkH → 0 as n tends to infinity. 

References

[1] Aigner, M. (2007). A course in Enumeration. Springer, Berlin. MR 2339282 [2] Avram, F. (1988). Weak convergence of the variations, iterated integrals and Dol´eans-Dadeexponentials of sequences of . Ann. Probab. 16 246–250. MR 0920268 [3] Avram, F. and Taqqu, M. (1986). Symmetric polynomials of random variables attracted to an infinitely divisible law. Probab. Theory Relat. Fields 71 491–500. MR 0833266 [4] Bai, S. and Taqqu, M. (2014). Generalized Hermite processes, discrete chaos and limit theorems. Stochastic Processes Appl. 124 1710–1739. MR 3163219 [5] Billingsley, P. (1968). Convergence of Probability Measures. John Wiley and Sons, Inc., New York-London- Sydney-Toronto. [6] Briand, P., Delyon, B. and Memin,´ J. (2002). On the robustness of backward stochastic differential equa- tions. Stochastic Processes Appl. 97 229–253. MR 1875334 [7] Burq, Z. A. and Jones, O. D. (2008). Simulation of Brownian motion at first-passage times. Math. Comput. Simulation 77 64–71. MR 2388251 [8] Coquet, F., Memin,´ J. and Slominski, L. (2001). On weak convergence of filtrations. S´eminaire de proba- bilit´esXXXV. Lecture Notes in Math. 1755 Springer, Berlin. 306–328. MR 1837294 [9] Cutland, N. and Ng, S. (1991). On homogeneous chaos. Math. Proc. Cambridge Philos. Soc. 110 353–363. MR 1113433 [10] Geiss, C., Geiss, S. and Gobet, E. (2012). Generalized fractional smoothness and Lp-variation of BSDEs with non-Lipschitz terminal condition. Stochastic Process. Appl. 122 2078–2116. MR 2921973 [11] Gzyl, H. (2006). An expos´e on discrete Wiener chaos expansions. Bol. Asoc. Mat. Venez. 13 3–27. MR 2267626 [12] Holden, H., Lindstrøm, T., Øksendal, B. and Ubøe, J. (1993). Discrete Wick products. Stochastic analysis and related topics. Stochastics Monogr. 8, Gordon and Breach, Montreux, 123–148. MR 1268009 [13] Holden H., Øksendal, B., Ubøe, J. and Zhang, T. (2010). Stochastic Partial Differential Equations. A Modeling, White Noise Functional Approach, 2nd ed. Springer, New York. MR 2571742 [14] Jacod, J., Mel´ eard,´ S. and Protter, P. (2000). Explicit form and robustness of martingale representa- tions. Ann. Probab. 28 1747–1780. MR 1813842 [15] Jakubowski, A., Memin,´ J. and Pages,` G. (1989). Convergence en loi des suites d’int´egralesstochastiques sur l’espace D1 de Skorokhod. Probab. Theory Related Fields 81 111–137. MR 0981569 [16] Janson, S. (1997). Gaussian Hilbert Spaces. Cambridge University Press, Cambridge. MR 1474726 [17] Krakowiak, W. and Szulga, J. (1986). Random multilinear forms. Ann. Probab. 14 955–973. MR 0841596 [18] Kuo, H.-H. (1996). White Noise Distribution Theory. Probability and Stochastics Series. CRC Press, Boca Raton. MR 1387829 [19] Kurtz, T. G. and Protter, P. (1991). Weak limit theorems for stochastic integrals and stochastic differ- ential equations. Ann. Probab. 19 1035–1080. MR 1112406 [20] Leao,˜ D. and Ohashi, A. (2013). Weak approximation for Wiener functionals. Ann. Appl. Probab. 23 1660–1691. MR 3098445 [21] Leitz-Martini, M. (2000). A discrete Clark-Ocone formula. Maphysto Research Report No. 29. DISCRETIZING MALLIAVIN CALCULUS 39

[22] Morters,¨ P. and Peres, Y. (2010). Brownian motion. Cambridge University Press, Cambridge. MR 2604525 [23] Nualart, D. (2006). The Malliavin Calculus and Related Topics, 2nd ed. Springer, New York. MR 2200233 [24] Nualart, D. and Pardoux, E. (1988). with anticipating integrands. Probab. Theory Related Fields. 78 535–581. MR 0950346 [25] Privault, N. (2009). Stochastic Analysis in Discrete and Continuous Settings. Lecture Notes in 1982. Springer, Berlin. MR 2531026 [26] Shiryaev, A. N. (1996). Probability, 2nd ed. Springer, New York. MR 1368405 [27] Sottinen, T. (2001). Fractional Brownian motion, random walks and binary market models. Finance Stoch. 5 343–355. MR 1849425 [28] Surgailis, D. (1982). Domains of attraction of self-similar multiple integrals. Lithuanian Math. J. 22 185– 201. MR 0684472 [29] Yosida, K. (1995). Functional analysis. Reprint of the sixth (1980) edition. Classics in Mathematics. Springer, Berlin. MR 1336382 [30] Zhang, J. (2004). A numerical scheme for BSDEs. Ann. Appl. Probab. 14 459–488. MR 2023027 —————— Saarland University, Department of Mathematics PO Box 151150, D-66041 Saarbrucken,¨ Germany, University of Mannheim, Institute of Mathematics A5,6, D-68131 Mannheim, Germany. E-mail address: [email protected], [email protected]