arXiv:1602.08858v1 [math.PR] 29 Feb 2016 osdr o every for consider, admvral ihzr xetto n aineoe sa As one. variance and expectation zero with variable random eeae yteBona oinadcmltdb ulsets. null by completed and motion Brownian the by generated Let ae ntedsrt ienie( sequ noise appropriate time by discrete the on Skorokhod based the motion and Brownian the derivative, to respect with h i fteppri oprovide to is paper the of aim The i.e. ebifl ics u ancnegnerslsi slightly a in results time convergence discrete Malliavin main for the our principle of discuss invariance briefly distribution of We kind the some on as depend regarded not do limits hr ( where as 1991 Date e od n phrases. and words Key oiin lr-cn eiaie naineprinciple. invariance derivative, Clark-Ocone position, ξ (1) easm htteapoiaigsequence approximating the that assume We . B pi ,2018. 3, April : ahmtc ujc Classification. Subject ( = otnoscascecet of coefficients chaos continuous nteBrolisaeb rigorous by Wi space the Bernoulli on the calculus Malliavin on between analogies formal known eiaiet hi otnoscutrat.Mroe,gi Moreover, counterparts. d continuous and their integral, to Skorokhod derivative discrete a derivative, Malliavin esrberno variable random measurable be hc di ho eopsto ntrso discrete of terms in decomposition chaos to a respect admit which ables ξ rvd eesr n ucetcniin o ekadstro and weak for conditions sufficient and necessary provide Abstract. admwlso h aepoaiiysaecnegn to converging space probability same the on walks random ho decomposition: Chaos os ( noise ftewienoise white the of utpeWee nerl st rjc admvariable random a project to is Wiener multiple i n ) B i ∈ t ) N t ≥ ξ sasqec fidpnetrno aibe hc aethe have which variables random independent of sequence a is 0 i n B ) eaBona oino rbblt pc (Ω space probability a on motion Brownian a be i n ∈ Suppose edrv eesr n ucetcniin o strong for conditions sufficient and necessary derive we , n N alai acls togapoiain tcatcint stochastic approximation, strong calculus, Malliavin ∈ ycnieigtedsrt iefunctions time discrete the considering by ICEIIGMLIVNCALCULUS MALLIAVIN DISCRETIZING N HITA EDRADPTRPARCZEWSKI PETER AND BENDER CHRISTIAN B = B ˙ ∀ t 1 saBona oinand motion Brownian a is { t 1 · · · , f ≥ h ersi dabhn h ho eopsto ntrso terms in decomposition chaos the behind idea heuristic The 2 X n,k X B . . . , lim : 0 ˙ ξ L t ( k X i i ovrec ftedsrt ho offiinsof coefficients chaos discrete the of convergence via n B i 2 00,6H5 60F25. 60H05, 60H07, hsie a emd iooswt epc otediscrete the to respect with rigorous made be can idea This . ) 1 } apoiainrslsfrsm ai prtr fMalliav of operators basic some for results -approximation nteseilcs fbnr os,orrslsspotthe support results our noise, binary of case special the In . t i n i . . . , L admwl approximation walk random a , ∈ 1. 2 N := n cnegneresults. -convergence ttrsotta nalorapoiainrsls the results, approximation our all in that out turns It . →∞ Introduction k √ = ) 1 n B t X n ⌊ B i nt =1 n 1 = k k/ ⌋ uha h ho eopsto,teMalliavin the decomposition, chaos the as such ! ξ B B 2 i n B E n t t , n   sa prxmtn euneo rescaled of sequence approximating an is npoaiiy (1) probability. in X ovre to converges ≥ e eune( sequence a ven srt nlge fteClark-Ocone the of analogues iscrete j Y =1 0 k ng , nrsaeadMlivncalculus Malliavin and space ener calculus. B ξ nomlway: informal ne fapoiaigoperators approximating of ences utpeWee nerl with integrals Wiener multiple L i n onws npoaiiy We probability. in pointwise j 2 Suppose   os,hneorrslscnbe can results our hence noise, cnegneo discretized a of -convergence , iceecutratof counterpart discrete L X F 2 gas -rnfr,casdecom- chaos S-transform, egrals, cnegnet a to -convergence P , B ∈ X L onws nprobability, in pointwise ,weethe where ), n ξ 2 frno vari- random of ) (Ω sasquare-integrable a is , F aedistribution same P , X n nproducts on ) othe to σ σ ( B -field )- B F we in is f 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 (f )n N converges strongly in X ∈ L2([0, )k) to the kth chaos coefficient of X, for every k N (Example 35). This is a simple∞ consequence of a general Wiener chaos limit theorem∈ (Theorem 29), which provides equivalent conditions for the strong L2(Ω, , P )-convergence of a sequence of n n F 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 (1982)] on convergence in distribution of discrete multiple Wiener integrals to strong L2(Ω, , P )-convergence (in our setting, i.e. when the limiting multiple Wiener integral is drivenF 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 [Gzyl (2006)] and the references therein. Instead we introduce the n discretized Malliavin derivative at time j N with respect to the noise (ξi )i N by ∈ ∈ n n n Dj X = √nE[ξj X (ξi )i N j ], | ∈ \{ } which is the gradient of the best approximation in L2(Ω, , P ) of X as a linear function in ξn with σ(ξn, i N j )-measurable coefficients. TheoremF 13 below implies that, j i ∈ \ { } if (Xn) converges weakly in L2(Ω, , P ) to X and the sequence of discretized Malliavin n n F 2 derivatives (D X )n N converges weakly in L (Ω [0, )), then X belongs to the n ∈ × ∞ domain of the⌈ continuous·⌉ 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 ex- pected 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 17 and 36. (3) : Defining the discrete Skorokhod integral as the adjoint operator to the discretized Malliavin derivative leads to

M n n n n n ξi δ (Z ) := lim E[Zi (ξj )j 1,...,M i ] , M | ∈{ }\{ } √n →∞ Xi=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 [Nualart and Pardoux (1988)]. Analogous results for the ‘closedness across the dis- cretization levels’ as in the case of the discretized Malliavin derivative and sufficient con- ditions for strong L2(Ω, , P )-convergence of a sequence of discrete Skorokhod integrals are provided in TheoremsF 9, 19 and 37. When restricted to predictable integrands, the convergence results for the Skorokhod integral give rise to necessary and sufficient condi- tions for strong and weak L2(Ω, , P )-convergence of a sequence of discrete Itˆointegrals (Theorem 21). This result can beF applied to study different discretization schemes for the generalized Clark-Ocone derivative (which provides the integrand in the predictable representation of a square-integrable as Itˆointegral with respect to the Brownian motion B). In this respect, Theorems 24 and 26 below complement related results in the literature such as [Briand et al. (2002), Le˜ao and Ohashi (2013)] and the references therein. We note that related classical limit theorems for stochastic integrals (with adapted integrands) [Jakubowski et al. (1989),Kurtz and Protter (1991)] and for multiple Wiener integrals [Surgailis (1982), Avram and Taqqu (1986), Avram (1988)], or robustness results for martingale representations [Jacod et al. (2000),Briand et al. (2002)] are usually obtained in the framework of (or using techniques of) convergence in distribution (on the Skorokhod space). DISCRETIZING 3

In contrast, we exploit that strong and weak convergence in L2(Ω, , P ) can be character- ized in terms of the S-transform, which is an important tool in whiteF noise analysis, see e.g. [Kuo (1996), Janson (1997), Holden et al. (2010)], and corresponds to taking expectation under suitable changes of measure. We introduce a discrete version of the S-transform in terms n 2 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 theF continuous S-transform (Theorem 1). With this observation at hand, all our convergence results can be obtained in a surprisingly 2 n simple way by computing suitable L (Ω,σ(ξi )i N, P )-inner products and their limits as n tends to infinity. However, all these results can be seen∈ as strong and weak invariance principles for Malliavin calculus. 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, , P )-convergence and the convergence of the discrete S-transform to the continuous one. EquivalentF 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. 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. Finally, in Section 7, we consider the special case of binary noise, in which discrete Malliavin calculus is very well studied, see e.g. the monograph by [Privault (2009)]. We explain that the statement of our convergence results can be simplified in this case and demonstrate by a toy example how to apply the results numerically in a Monte Carlo framework.

2. Weak and strong L2-convergence via discrete S-transforms In this section, we study strong and weak L2(Ω, , P )-convergence of a sequence (Xn) of random n n n F variables, where X is := σ(ξi , i N)-measurable, to an -measurable X. As a main result of this section (TheoremF 1), we provide∈ an equivalent criterionF for this convergence, which only requires to compute a family of L2(Ω, n, P )-inner products (hence, expectations which involve Fn 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 j ξi , ξ1,ξ2,... independent and with the same distribution as ξ, i=1 !j N X ∈ 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 (2) nt t 0 ⌊ ⌋ ≥ nt   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. [M¨orters and PeresP (2010), 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, that the mapping 1 B can be extended to a continuous linear mapping from L2([0, )) (0,t] 7→ t ∞ 4 C. BENDER AND P. PARCZEWSKI to L2(Ω, , P ), which is known as the Wiener integral. We denote the Wiener integral of a function fF L2([0, )) by I(f). The discrete Wiener integral is given by ∈ ∞ 1 ∞ In(f n) := f n(i)ξn. √n i Xi=1 Here, the discrete time function f n is a member of

2 n n 2 1 ∞ n 2 Ln(N) := f : N R : f L2 (N) := (f (i)) < , ( → k k n n ∞) Xi=1 which obviously ensures that the In(f n) converges (strongly) in L2(Ω, n, P ). The Wick exponential is, by definition, the stochastic exponential of a WienerF integral I(f), i.e.,

∞ 2 exp⋄(I(f)) := exp I(f) 1/2 f (s)ds . −  Z0  Hence, its discrete counterpart, the discrete Wick exponential, is given by

∞ n n n 1 n n exp⋄ (I (f )) := 1+ f (i)ξ . (3) √n i Yi=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( f 2 N ) < . (4) ≤ k kLn( ) ∞ Notice also that ∞ n n n n n n n n n exp⋄ (I (f )) = 1+ exp⋄ (I (f 1[1,i])) exp⋄ (I (f 1[1,i 1])) − − i=1 X  ∞ n n n n n ξi = 1+ f (i)exp⋄ (I (f 1[1,i 1])) , (5) − √n Xi=1 which is the discrete counterpart of the Dol´eans-Dade equation. We finally recall that, for every X L2(Ω, , P ) and f L2([0, )), the S-transform is defined as ∈ F ∈ ∞ (SX)(f) := E[X exp⋄(I(f))]. n 2 n n 2 Analogously, for every X L (Ω, , P ) and f Ln(N), we introduce the discrete S- transform as ∈ F ∈ n n n n n n n (S X )(f ) := E[X exp⋄ (I (f ))]. We emphasize that the S-transform is a powerful tool in the analysis, see, e.g., [Kuo (1996)], and has been succesfully applied in the theory of stochastic partial differential equations, see [Holden et al. (2010)]. To the best of our knowledge the discrete S-transform has, however, not been studied in the literature. Let us next denote by the set of step functions on left half-open intervals, i.e., functions of the form E m g(x)= a 1 (x), m N, a , b , c R. j (bj ,cj] ∈ j j j ∈ Xj=1 2 As the set of Wick exponentials of step functions exp⋄(I(g)), g is total in L (Ω, , P ), see e.g. [Janson (1997), Corollary 3.40], every L2(Ω, { , P )-random variable∈ E} is uniquely determinedF by its S-transform. More precisely, if for X,Y F L2(Ω, , P ), (SX)(g) = (SY )(g) for every g , then X = Y P -almost surely. We define the∈ discretizationF of a step function g as ∈E ∈E gˇn = (ˇgn(1), gˇn(2),...) := (g(1/n), g(2/n),...) , and notice that gˇn : g L2 (N) { ∈E}⊂ n is the dense subspace of discrete time functions with finite support. DISCRETIZING MALLIAVIN CALCULUS 5

The convergence results of integral and derivative operators in this paper rely on the following characterization of L2(Ω, , P )-convergence in terms of convergence of the discrete S-transform to the continuous S-transform.F Theorem 1. Suppose X, Xn L2(Ω, , P ) for every n N, with Xn being n-measurable. Then the following assertions are∈ equivalentF as n tends to∈ infinity: F (i) Xn X strongly (resp. weakly) in L2(Ω, , P ). (ii) (SnX→n)(ˇgn) (SX)(g) for every g , andF additionally E[(Xn)2] E[X2] in the case → ∈E n 2 → of strong convergence (resp. supn N E[(X ) ] < in the case of weak convergence). ∈ ∞ Moreover, in the case of strong convergence, (i) is also equivalent to

n n n n n 2 (iii) (X , exp⋄ (I (ˇg ))) (X, exp⋄(I(g))) in distribution for every g , and ((X ) )n N is uniformly integrable.→ ∈ E ∈ Remark 2. Note, that X L2(Ω, , P ) is, of course, not determined by its univariate distri- ∈ F I(g) bution, but it is uniquely determined by all the bivariate distributions of (X, e⋄ ), g , in view of the injectivity of the S-transform. This observation motivates that the characteri∈zation E of strong L2(Ω, , P )-convergence via convergence in distribution in item (iii) of Theorem 1 can hold. F In view of Lemma 4 below, the proof of Theorem 1 can be reduced to the following strong L2-convergence result for (discrete) Wick exponentials. Proposition 3. Suppose g . Then, we have strongly in L2(Ω, , P ), as n tends to infinity: ∈E F n n n exp⋄ (I (ˇg )) 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 weak convergence results for stochastic differential equations, see, e.g., [Avram (1988), Kurtz and Protter (1991)] and the references therein. For sake of completeness, we here provide an elementary proof. Proof. Let m g = a 1 . j (bj ,cj ] ∈E Xj=1 We denote by C,N constants in N such that g is bounded by C and has support in [0,N]. Decomposing

n n n 2 E (exp⋄(I(g)) exp⋄ (I (ˇg ))) − h 2 n n i n n n n 2 = E (exp⋄(I(g))) 2E [exp⋄ (I (ˇg )) exp⋄(I(g))] + E (exp⋄ (I (ˇg ))) , − it suffices to showh i h i

n n n 2 2 (i) limn E (exp⋄ (I (ˇg ))) = E (exp⋄(I(g))) , →∞ n n n (ii) exp⋄ (I (ˇgh )) exp⋄(I(g))i in probability,h i → because under (i) the integrand in the second term on the right-hand side is uniformly integrable. (i) Due to p< q r p

Nn n 2 1 n 2 ∞ 2 (ˇg (i)) = gˇ ( n ) L2([0, )) g(s) ds. n k ⌈ ·⌉ k ∞ → 0 Xi=1 Z 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 ∈

Nn Nn n n n 2 1 n n 2 1 n 2 E (exp⋄ (I (ˇg ))) = E (1 + gˇ (i)ξ ) = 1+ (ˇg (i)) √n i n i=1   i=1     Y Y ∞ 2 2 exp g(s) ds = E (exp⋄(I(g))) . → 0 Z  h i (ii) In order to treat the large jumps of Bn and the small ones separately, we consider

n,1 n n,2 n ξ := ξ 1 √n , ξ := ξ 1 √n , i i ξn i i ξn > {| i |≤ 2C } {| i | 2C } cp. also [Sottinen (2001)]. Then,

Nn Nn n n n 1 n n,1 1 n n,2 n,1 n,2 exp⋄ (I (ˇg )) = 1+ gˇ (i)ξ 1+ gˇ (i)ξ =: E E √n i √n i · Yi=1   Yi=1   n We note that, for every ǫ> 0, by the independence of (ξi )i N, ∈ ξn P ( ξ >ǫ√n )Nn Nn P sup | i | >ǫ = 1 1 {| | } 0, (8) √n − − Nn → (i=1,...,Nn )!   because, by square-integrability of ξ, P ( ξ >ǫ√n )n 0, see, e.g., [Shiryaev (1996), p. 208]. Hence, for every ǫ> 0, {| | } →

n n,2 n,2 ξi P ( E 1 ǫ ) P ( sup ξi > 0 )= P sup | | > 1/(2C) 0, {| − |≥ } ≤ {i=1,...,Nn | | } (i=1,...,Nn √n )! →

n,2 i.e., (E )n N converges to 1 in probability. By construction,∈ each factor in En,1 is larger than 1/2. Applying a Taylor expansion to the logarithm, thus, yields

Nn Nn ξn,1 1 (ξn,1)2 log En,1 = gˇn(i) i (ˇgn(i))2 i + R √n − 2 n n Xi=1 Xi=1 with a remainder term satisfying

n Nn n,1 2 8C ξj n 2 (ξi ) Rn sup | | (ˇg (i)) . | |≤ 3 j=1,...,Nn √n ! n Xi=1 It, thus, suffices to show n,1 Nn n ξi (iii) i=1 gˇ (i) √n I(g) in probability, →n,1 2 Nn 2 (ξ ) (iv) P (ˇgn(i)) i ∞ g(s)2ds in probability. i=1 n → 0 Indeed,P by (8), the remainder termR then vanishes in probability as n tends to infinity, and, thus, 1 En,1 exp I(g) ∞ g(s)2ds in probability. → − 2  Z0  DISCRETIZING MALLIAVIN CALCULUS 7

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 (6),

Nn n m m n ξi n n lim gˇ (i) = lim aj B B = aj Bc Bb = I(g), in probability. n √n n cj − bj j − j →∞ i=1 →∞ j=1 j=1 X X   X  1 nt n 2 Finally, by the , n i⌊=1⌋(ξi ) converges to t in probability for every t 0, and, hence, by (6), ≥ P Nn n 2 m n 2 (ξi ) 2 ∞ 2 lim (ˇg (i)) = a (cj bj)= g(s) ds, in probability. n j →∞ n − 0 Xi=1 Xj=1 Z 

The following simple lemma from functional analysis turns out to be useful. Lemma 4. 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: (i) xn x strongly (resp. weakly) in H. (ii) x →,xa x,xa for every a A, and additionally x x in the case of h n niH → h iH ∈ k nkH → k kH 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 [Yosida (1995), Theorem V.1.1] in (i) or by assumption∈ k (ii).k Thus, for every a A, by the strong convergence a a ∈ of (xn) to x , a a a a a a xn,xn H xn,x H = xn,xn x H sup xm H xn x H 0. (9) |h i − h i | |h − i |≤ m N k k k − k → ∈ a Let us treat the case of weak convergence: If (i) holds, the term xn,x H in (9) converges a a h i to x,x H , and then so does xn,xn H , which implies (ii). Conversely, if (ii) holds, the first h i a h a i a term xn,xn H in (9) tends to x,x H , and then so does xn,x H , which yields (i) in view of [Yosidah (1995),i Theorem V.1.3].h Thei case of strong convergenceh isi an immediate consequence, as, in a Hilbert space, strong convergence is equivalent to weak convergence and convergence of the norms [Yosida (1995), Theorem V.1.8]. 

We are now in the position to prove Theorem 1.

Proof of Theorem 1. ‘(i) (ii)’: Proposition 3 and Lemma 4 apply immediately in view of the definition of the (discrete)⇔S-transform, and as the set of Wick exponentials of step functions 2 exp⋄(I(g)), g is total in L (Ω, , P ). ‘({ i) with strong∈ convergence E} (iii)’:F This is is a direct consequence of Proposition 3 and the 2 ⇒ assumed strong L (Ω, , P )-convergence of (Xn). ‘(iii) (ii) with strongF convergence’: By (iii) and the continuous mapping theorem, the se- ⇒ n n n n quence (X exp⋄ (I (ˇg ))) converges in distribution to X exp⋄(I(g)). Moreover, this sequence is n 2 n n n 2 uniformly integrable, because so are the sequences ( X ) by assumption and ( exp⋄ (I (ˇg )) ) by Proposition 3. Hence, | | | |

n n n n n n n (S X )(ˇg )= E[X exp⋄ (I (ˇg )] E[X exp⋄ I(g)] = (SX)(g). → Moreover, thanks to the uniform integrability of ((Xn)2) and the convergence in distribution Xn d X, we have E[(Xn)2] E[X2]. This completes the proof of (ii) with strong convergence. → → 

We close this section with an example. 8 C. BENDER AND P. PARCZEWSKI

Example 5. (i) In this example, we provide a simple proof, that, for every X L2(Ω, , P ), Xn := E[X n] converges to X strongly in L2(Ω, , P ). Indeed, by Proposition∈ 3, forF every g , |F F ∈E n n n n n n n n n n (S X )(ˇg )= E [E[X ]exp⋄ (I (ˇg ))] = E [X exp⋄ (I (ˇg ))] E [X exp⋄(I(g))] = (SX)(g). |F → As E[(Xn)2] E[X2], Theorem 1 implies weak L2(Ω, , P )-convergence of (Xn) to X. The same theorem≤ finally yields strong L2(Ω, , P )-convergence,F since, by the already established weak convergence, F E[(Xn)2]= E [E [Xn n] X]= E[XnX] E[X2]. | F → We note that this result can alternatively be derived by the uniform integrability of ((Xn)2) via the concept of convergence of filtrations making use of [Coquet et al. (2001), Proposition 2]. n n n (ii) Denote by ( t)t 0 the augmented Brownian filtration and let i = σ(ξ1 ,...,ξi ). We 2 F ≥ F n assume X L (Ω, T , P ). Then, one can always approximate X by a sequence (XT ) strongly 2 ∈ F n n n in L (Ω, , P ), where XT is measurable with respect to nT . Indeed, take any sequence (X ) F F⌊ ⌋ of n-measurable random variables which converges strongly in L2(Ω, , P ) to X, and define nF n n F XT = E[X nT ]. Then, for every g , by Proposition 3, |F⌊ ⌋ ∈E nT ⌊ ⌋ n n n n n 1 n n n n (S X )(ˇg ) = E X 1+ g(i/n)ξ = E X exp⋄ (I ((g1ˇ ) )) T  √n i  (0,T ] Yi=1   h i E X exp⋄(I(g1 )) = E [XE[exp⋄(I(g)) ]] = (SX) (g) . → (0,T ] |FT Moreover,   n 2 n 2 sup E[(XT ) ] sup E[(X ) ] < . n N ≤ n N ∞ ∈ ∈ n 2 2 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,F because F E[(Xn)2]= E[XnX]+ E[Xn(Xn X)] E[X2]. T T T − → 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. [Janson (1997), Theorem 16.46, Theorem 16.50]. 2 2 Definition 6. Z L (Ω [0, )) := L (Ω [0, ), ([0, )), P λ[0, )) is said to belong ∈ × ∞ × ∞ F⊗B ∞ ⊗ ∞ to the domain D(δ) of the Skorokhod integral, if there is an X L2(Ω, , P ) such that for every g ∈ F ∈E ∞ (SX)(g)= (SZt)(g)g(t)dt. Z0 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

2 n 2 n n 2 1 ∞ n 2 Ln(Ω N) := Z : N L (Ω, , P ), Z L2 (Ω N) := E[(Zi ) ] < . × ( → F k k n × n ∞) Xi=1 Moreover, we recall the definitions n := σ(ξn, j N), n := σ(ξn,...,ξ ), F j ∈ FM 1 M DISCRETIZING MALLIAVIN CALCULUS 9 and introduce the shorthand notations n n n n i := σ(ξj , j N i ), M, i := σ(ξj , j 1,...,M i ). F− ∈ \ { } F − ∈ { } \ { } n 2 n Definition 7. We say, Z Ln(Ω N) belongs to the domain D(δ ) of the discrete Skorokhod integral, if ∈ × M n n n n n ξi δ (Z ) := lim E[Zi M, i] . (10) M |F − √n →∞ Xi=1 exists strongly in L2(Ω, , P ). If this is the case, δn(Zn) is called the discrete Skorokhod integral of Zn. F We note that, by the independence of E[Zn n ] and ξn, each summand on the right-hand i |FM, i i side of (10) is indeed a member of L2(Ω, , P ). Moreover,− the martingale convergence theorem implies that, for every Zn L2 (Ω N) andF N N, Zn1 D(δn) and ∈ n × ∈ [1,N] ∈ N n n n n n ξi δ (Z 1[1,N])= E[Zi i] . (11) |F− √n Xi=1 2 2 Hence, the discrete Skorokhod integral is densely defined from Ln(Ω N) to L (Ω, , P ). We will show in Proposition 14 below that it is a closed operator. × F Remark 8. This definition of the discrete Skorokhod integral closely resembles the following Riemann-sum approximation of the Skorokhod integral by [Nualart and Pardoux (1988)], who show that under appropriate conditions on Z, n 1 − (i+1)/n δ(Z1[0,1]) = lim E n Zsds (Bs,B1 Br)0 s i/n (i+1)/n r 1 B(i+1)/n Bi/n n " i/n − ≤ ≤ ≤ ≤ ≤ # − →∞ i=0 Z X  strongly in L2(Ω, , P ). F As a first main result of this section we are going to show the following weak approximation theorem for Skorokhod integrals. n n n Theorem 9. Suppose Z D(δ ) for every n N, and (Z n )n N converges to Z weakly in ∈ ∈ ⌈ ·⌉ ∈ L2(Ω [0, )). Then, the following assertions are equivalent: × ∞ n n 2 (i) supn N E[ δ (Z ) ] < . (ii) Z D∈ (δ)|and (δn|(Zn))∞converges to δ(Z) weakly in L2(Ω, , P ) as n tends to infinity. ∈ F As a first tool for the proof we state the discrete S-transform of a discrete Skorokhod integral. Proposition 10. Suppose Zn D(δn). Then, for every g , ∈ ∈E n n n n 1 ∞ n n n n (S δ (Z )) (ˇg )= (S Zi )(ˇg 1N i )ˇg (i). n \{ } Xi=1 This result is a special case of the more general Proposition 14 below, to which we refer the reader for the proof. The second tool for the proof of Theorem 9 is the following variant of Theorem 1 for stochastic processes. 2 n n 2 Theorem 11. Suppose Z L (Ω [0, )), (Z )n N satisfies Z Ln(Ω N) for every n N. Then the following assertions∈ are equivalent× ∞ as n tends∈ to infinity:∈ × ∈ (i) (Zn ) converges strongly (resp. weakly) to Z in L2(Ω [0, )). n × ∞ (ii) For⌈ every·⌉ g, h ∈E 1 ∞ n n n ˇn ∞ (S Zi )(ˇg )h (i) (SZs)(g)h(s)ds. n → 0 Xi=1 Z 10 C. BENDER AND P. PARCZEWSKI

n 2 2 and, additionally, E[ 0∞(Z ns ) ds] E[ 0∞ Zs ds] in the case of strong convergence n 2⌈ ⌉ → (resp. supn N E[ 0∞(Z ns ) ds] < in the case of weak convergence). ∈ R⌈ ⌉ ∞ R 1 n n n ˇn 1 n n n ˇn Moreover, in (ii), n i∞=1R (S Zi )(ˇg )h (i) can be replaced by n i∞=1(S Zi )(ˇg 1N i )h (i). \{ } Proof. We wish to applyP Lemma 4 in order to prove the equivalenceP of (i) and (ii). As L2(Ω [0, )) = L2(Ω, , P ) L2([0, )) (with the tensor product in the sense of Hilbert spaces),× ∞ F ⊗ ∞ 2 the set exp⋄(I(g))h; g, h is total in L (Ω [0, )). In view of Proposition 3 and (7), n {n n n ∈ E} × ∞ 2 (exp⋄ (I (ˇg ))hˇ ( n ))n N converges to exp⋄(I(g))h strongly in L (Ω [0, )) for every g, h . As ⌈ ·⌉ ∈ × ∞ ∈ E 1 ∞ n n n ˇn n n n n ˇn (S Zi )(ˇg )h (i)= Z n , e⋄ (I (ˇg ))h ( n ) , n ⌈ ·⌉ ⌈ ·⌉ L2(Ω [0, )) Xi=1 D E × ∞ Lemma 4 applies indeed. We finally note, that the ‘Moreover’-part of the assertion is an immediate consequence of the Cauchy-Schwarz inequality and the estimate

n n n n n n 2 E exp⋄ (I (ˇg )) exp⋄ (I (ˇg 1N i )) − \{ } h n n n 2 n n  i 2 = E exp⋄ (I (ˇg 1N i )) E gˇ (i)ξi /√n \{ } h n 2 2 i h i exp( gˇ L2 (N))sup g(j) /n 0,  ≤ k k n j N | | → ∈ making use of (4) in the last line.  We are now ready to give the proof of Theorem 9. Proof of Theorem 9. As the implication ‘(ii) (i)’ is trivial, we only have to show the converse implication. To this end, note first that, by Proposition⇒ 10 and Theorem 11, for every g , ∈E n n n n ∞ lim (S δ (Z ))(ˇg )= (SZt)(g)g(t)dt. (12) n →∞ Z0 n n As the sequence (δ (Z ))n N is norm bounded by (i), it has a weakly convergent subsequence [Yosida (1995), Theorem V.2.1].∈ We denote its limit by X. Then, applying Theorem 1 and (12) along the subsequence, we obtain, for every g , ∈E ∞ (SX)(g)= (SZt)(g)g(t)dt. (13) Z0 Hence, by Definition 6, Z D(δ) and δ(Z) = X. Finally, by Theorem 1 and (12)–(13), weak 2 ∈ n n L (Ω, , P )-convergence of (δ (Z ))n N to δ(Z) holds along the whole sequence, and not only alongF 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 12. A random variable X L2(Ω, , P ) is said to belong to the domain D1,2 of the Malliavin derivative, if there is a stochastic∈ processF Z L2(Ω [0, )) such that for every g, h , ∈ × ∞ ∈E ∞ ∞ (SZ )(g)h(s)ds = E X exp⋄(I(g)) I(h) g(s)h(s)ds . s − Z0   Z0  In this case, Z is unique and DX := Z is called the Malliavin derivative X. For every X L2(Ω, , P ) we define the discretized Malliavin derivative of X at j N with n∈ F ∈ respect to (ξi )i N by ∈ n n n Dj X := √nE[ξj X j]. |F− DISCRETIZING MALLIAVIN CALCULUS 11

n 2 2 We observe that, for fixed j, Dj is a continuous linear operator from L (Ω, , P ) to L (Ω, , P ), because by H¨older’s inequality for conditional expectations and the independenceF of the familyF n (ξi )i N, ∈ n 2 2 n n 2 n 2 n Dj X nE[X j] E[(ξj ) j]= nE[X j]. | | ≤ |F− |F− |F− 1,2 We say that X belongs to the domain Dn of the discretized Malliavin derivative, if the process n n 2 n D X := (Di X)i N is a member of Ln(Ω N). 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, , P ) to 2 F Ln(Ω N). × n In the following theorem and in the remainder of the paper we use the convention Z0 = 0 for Zn L2 (Ω N). ∈ n × n 2 n 1,2 Theorem 13. Suppose (X )n N converges to X weakly in L (Ω, , P ) and X Dn for every n N. Then, the following are∈ equivalent: F ∈ ∈ 1 n n 2 (i) supn N n i∞=1 E[(Di X ) ] < . ∈ 1,2 n n ∞ 2 (ii) X D and (D n X )n N converges to DX weakly in L (Ω [0, )). ∈ P ⌈ ·⌉ ∈ × ∞ The proof is prepared by two propositions. The first one contains the duality relation between the discrete Skorokhod integral and discretized Malliavin derivative. Proposition 14. 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 ∞ E [ZnDnX]= E[δn(Zn)X]. n i i Xi=1 n n n We emphasize that, choosing X = exp⋄ (I (ˇg )), g , in Proposition 14, we obtain the assertion of Proposition 10. Indeed, we only have to note∈ Ethat, for every f n L2 (N), ∈ n n n n n n n n n Di exp⋄ (I (f )) = f (i)exp⋄ (I (f 1N i )). \{ } n n 1,2 Proof. Suppose first, that Z D(δ ) and X Dn . Then, for every M N, and i N, ∈ ∈ ∈ ∈ n n 2 n 2 E √nE[ξi X M, i] E Di X . |F − ≤ | | Hence, by the martingale convergenceh theorem and i dominatedh convergence,i

1 ∞ n n n 2 lim E √nE[ξi X M, i] Di X = 0. M n |F − − →∞ i=1 X h i Consequently, M 1 ∞ n n 1 n n n E [Zi Di X] = lim E Zi √nE[ξi X M, i] n M n |F − →∞ Xi=1 Xi=1 M   1 n n n = lim E Xξi E[Zi M, i] M √n |F − →∞ i=1 X   M n n n ξi n n = lim E X E[Zi M, i] = E[Xδ (Z )]. M |F − √n →∞ " # Xi=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, , P ) such that for every X Dn , ∈ F ∈ 1 ∞ E [ZnDnX]= E[Y nX]. (14) n i i Xi=1 12 C. BENDER AND P. PARCZEWSKI

1,2 n 1,2 We first note that, by construction, X Dn if and only E[X ] Dn , and, if this is the case, both random variables have the same∈ discretized Malliavin|F derivative.∈ Hence, applying the duality relation (14), with X and E[X n], we observe that, Y n = E[Y n n]. Now suppose that 2 n 1,2 |Fn n n |F n X L (Ω, , P ). Then X Dn , D X = √nE[ξ X ] for every i M, and D X = 0 ∈ FM ∈ i i |FM, i ≤ i for i > M. Hence, (14) and the same manipulations as above− imply

M n n n n ξi E[Y X]= E X E[Zi M, i] , " |F − √n# Xi=1 i.e. M n n n n n ξi E[Y M ]= E[Zi M, i] . |F |F − √n Xi=1 n n 2 By the martingale convergence theorem, (E[Y M ])M N converges strongly in L (Ω, , P ) to E[Y n n]= Y n. Hence, Zn D(δn) and δn(Zn)=|F Y n.∈ Finally, closedness is a general propertyF of adjoint|F operators, see, e.g.,∈ [Yosida (1995), p. 196].  The next proposition is a consequence of the weak convergence result for discrete Skorokhod integrals in Theorem 9. Proposition 15. For every g, h , ∈E

n n n n n ∞ lim δ (exp⋄ (I (ˇg )) hˇ ) = exp⋄(I(g)) I(h) g(s)h(s)ds n − →∞  Z0  strongly in L2(Ω, , P ). F n n n n n n Proof. Notice first that, for fixed n N, exp⋄ (I (ˇg )) hˇ D(δ ), because hˇ (i) vanishes, if i is sufficiently large. A direct computation,∈ making use of (11∈ ), shows

∞ ξn n n n n ˇn n n n ˇn i δ (exp⋄ (I (ˇg )) h )= exp⋄ (I (ˇg 1N i ))h (i) . \{ } √n Xi=1 n For i = j we obtain, by independence of (ξk )k N, 6 ∈

n n n n n n n n n 1 n 1 n 2 1 E exp⋄ (I (ˇg 1N i )) exp⋄ (I (ˇg 1N j ))ξi ξj =g ˇ (i) gˇ (j) 1 +g ˇ (k) . \{ } \{ } √n √n n k N i,j     ∈ Y\{ } Combining this with an analogous calculation for the case i = j yields

∞ n n n n n 2 1 n n n n n 2 1 E δ (exp⋄ (I (ˇg )) hˇ ) = hˇ (i)ˇg (i)hˇ (j)ˇg (j) 1 +g ˇ (k) n2 n i,j=1,i=j k N i,j   h i X6 ∈ Y\{ }

1 ∞ 1 + hˇn(i)2 1 +g ˇn(k)2 . n n i=1 k N i   X ∈Y\{ } As g and h are bounded with compact support, it is straightforward to check in view of (7) that 2 2 n n n n n 2 R ∞ g(s) ds ∞ ∞ 2 lim E δ (exp⋄ (I (ˇg )) hˇ ) = e 0 h(s)g(s)ds + h(s) ds . (15) n →∞ Z0  Z0 ! h i n n n n n 2 Thus, (δ (exp⋄ (I (ˇg )) hˇ ))n N converges to δ(exp⋄(I(g)) h) weakly in L (Ω, , P ) by Theorem 9. The identity ∈ F ∞ δ(exp⋄(I(g)) h) = exp⋄(I(g)) I(h) g(s)h(s)ds −  Z0  can either be derived by a direct computation making use of the S-transform definition of the Skorokhod integral (Definition 6) or alternatively is a simple consequence of [Nualart (2006), DISCRETIZING MALLIAVIN CALCULUS 13

Proposition 1.3.3] in conjunction with Definition 1.2.1 in the same reference. Applying the Cameron-Martin shift [Janson (1997), Theorem 14.1] twice, we observe

2 2 ∞ R ∞ g(s) ds 2 E exp⋄(I(g)) I(h) g(s)h(s)ds = e 0 E exp⋄(I(g))I(h) − "  Z0  # 2   2 ∞ = eR0∞ g(s) dsE I(h)+ g(s)h(s)ds " Z0  # 2 2 ∞ ∞ = eR0∞ g(s) ds h(s)g(s)ds + h(s)2ds . Z0  Z0 ! Thanks to (15), this turns weak into strong convergence. 

The proof of Theorem 13 is now analogous to that of Theorem 9.

1 n n 2 n n 2 Proof of Theorem 13. ‘(ii) (i)’ is obvious, since n i∞=1 E[(Di X ) ] = 0∞ E[(D ns X ) ]ds. ⇒ ⌈ ⌉ n n n n ˇn ‘(i) (ii)’: Notice first that, for every g, h , by PropositionP 14 with ZR = exp⋄ (I (ˇg )) h and⇒ Proposition 15, ∈E

∞ 1 n n n n n n n n n n n lim (S D X )(ˇg )hˇ (i) = lim E[X δ (exp⋄ (I (ˇg ))hˇ )] n i n →∞ n →∞ Xi=1 ∞ = E X exp⋄(I(g)) I(h) g(s)h(s)ds , (16) −   Z0  n 2 n n since (X ) converges to X weakly in L (Ω, , P ). As the sequence (D n X )n N is norm F ⌈ ·⌉ ∈ bounded in L2(Ω [0, )) by (i), it has a weakly convergent subsequence. We denote its limit by Z. Applying (16)× and∞ Theorem 11 along this subsequence, we conclude

∞ ∞ (SZ )(g)h(s)ds = E X exp⋄(I(g)) I(h) g(s)h(s)ds . (17) s − Z0   Z0  Hence, X D1,2 and DX = Z by Definition 12. Finally, applying (16)–(17) and Theorem 11 ∈ n n 2 along the whole sequence (D X )n N, shows that this sequence converges weakly in L (Ω n ∈ × [0, )) to DX. ⌈ ·⌉  ∞ 1,2 In order to check the assumptions of Theorem 9, we consider the space Ln , which consists of n 2 n 1,2 processes Z L (Ω N) such that Z Dn for every i N and ∈ n × i ∈ ∈ 1 ∞ E DnZn 2 < . (18) n2 | j i | ∞ i,j=1,i=j X6   1,2 n n 1,2 Proposition 16. For every n N, Ln D(δ ) and, for Z Ln , ∈ ⊂ ∈ n n n ∞ n n ξi 2 δ (Z ) = E[Zi i] , (strong L (Ω, , P )-convergence), (19) |F− √n F Xi=1 1 ∞ 1 ∞ E n n 2 E E n n 2 E n n n n (δ (Z )) = Zi i + 2 (Di Zj )(Dj Zi ) . (20) n |F− n i=1 i,j=1,i=j h i X h   i X6   In particular, in the context of Theorem 9, assertion (i) is equivalent to 1 E n n n n (i’) supn N n2 i,j∞=1,i=j (Di Zj )(Dj Zi ) < , ∈ 6 ∞ h n 1,2 i if we additionallyP assume that Z Ln for every n N. ∈ ∈ 14 C. BENDER AND P. PARCZEWSKI

Proof. Fix N < N N. Then, 1 2 ∈ 2 N2 n N2 n n ξi 1 n n 2 n 2 E E[Zi i] = E E[Zi i] (ξi )  |F− √n  n |F− iX=N1 iX=N1 h i  N2   1 n n n n n n + E E[Zi i]E[Zj j]ξi ξj = (I)N1,N2 + (II)N1,N2 . n |F− |F− i,j=N1,i=j X 6   n By the independence of the discrete-time noise (ξi )i N and as the conditional expectation has norm 1, ∈ N 1 n n 2 1 ∞ n n 2 (I)1,N = E E[Zi i] E E[Zi i] < ,N , (21) n |F− → n |F− ∞ →∞ i=1 i=1 X   X   and (I) 0 as N ,N tend to infinity. In order to treat (II) , we first note that for N1,N2 → 1 2 N1,N2 any random variable Xn L1(Ω, n, P ) and i = j N, by Fubini’s theorem, ∈ F 6 ∈ n n n n n n E E X i j = E E X j i . (22) |F− F− |F− F− Hence, for i = j N,         6 ∈ n n n n n n n n n n n n E E[Zi i]E[Zj j]ξi ξj = E E[Zi ξj i]E[Zj ξi j] |F− |F− |F− |F− n n n n n n n n n n n n = E E E[Zi ξj i] j Zj ξi = E E E[Zi ξj j] i Zj ξi |F− F− |F− F−   n n n n n  n  1  n n n n   = E E[Zi ξj j]E[Zj ξi i] = E (Di Zj )(Dj Zi ) . |F− |F− n Consequently, by Young’s inequality,   

n n n n n n 1 n n 2 1 n n 2 n E E[Zi i]E[Zj j]ξi ξj E (Di Zj ) + E (Dj Zi ) . |F− |F− ≤ 2 2 1,2       The Ln -assumption, thus, ensures that

1 ∞ n n n n lim (II)1,N = E (Di Zj )(Dj Zi ) < N n2 ∞ →∞ i,j=1,i=j X6   n n and (II)N1,N2 0 as N1,N2 tend to infinity. Hence, by (11), the sequence (δ (Z 1[1,N]))N N is → ∈ Cauchy in L2(Ω, , P ). By the closedness of the discrete Skorokhod integral, Zn D(δn) and 1,2 F n ∈ we obtain Ln D(δ ), (19) and (20). We finally suppose that the assumptions of Theorem 9 ⊂ n 1,2 are in force and that Z Ln for every n N. Then, ∈ ∈ 1 ∞ n n 2 sup E E[Zi i] < , n N n |F− ∞ ∈ Xi=1   n because of the assumed weak convergence of the sequence (Z )n N. Thus, the sequence n ∈ n n 2 ⌈ ·⌉ (δ (Z ))n N is norm bounded in L (Ω, , P ), if and only if (i’) holds.  ∈ F As a consequence of the previous proposition, we obtain the following strong L2(Ω, , P )- convergence results to the Malliavin derivative. F n 2 Theorem 17. Suppose (X )n N converges to X strongly in L (Ω, , P ). Moreover assume that n 2,2 ∈ F X Dn for every n N, i.e. ∈ ∈ 1 ∞ 1 ∞ E (DnX)2 + E (DnDnX)2 < . n i n2 j i ∞ i=1 i,j=1,i=j X   X6   Then, the following assertions are equivalent: 1 E n 2 1 E n n 2 (i) supn N n i∞=1 (Di X) + n2 i,j∞=1,i=j (Dj Di X) < . ∈ 6 ∞  P   P h i DISCRETIZING MALLIAVIN CALCULUS 15

1,2 n n 2 (ii) X D , DX D(δ), (D X )n N converges to DX strongly in L (Ω [0, )), and n ∈ n∈ n n ∈ ⌈ ·⌉ 2 × ∞ (δ (D X ))n N converges to δ(DX) weakly in L (Ω, , P ). ∈ F Remark 18. Recall that L = δ D is the infinitesimal generator of the Ornstein-Uhlenbeck semigroup, see [Nualart (2006),− Section◦ 1.4], and is sometimes called Ornstein-Uhlenbeck op- erator (cf. also [Janson (1997), 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 2,2 n 1,2 Proof. Let Z = D X . Then, X Dn implies Z Ln . Note that, for i = j, by (22), i i ∈ ∈ 6 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 13 and Theorem 9 in conjunction with Proposition 16, assertion (i) is equivalent to 1,2 n n 2 (ii’) X D , DX D(δ), (D X )n N converges to DX weakly in L (Ω [0, )), and n ∈ n∈ n n ∈ ⌈ ·⌉ 2 × ∞ (δ (D X ))n N converges to δ(DX) weakly in L (Ω, , P ). ∈ F n n So we only need to show that under (ii’) the convergence of (D n X )n N to DX holds true in ⌈ ·⌉ ∈ the strong topology. However, by the duality relation in Proposition 14, the weak L2(Ω, , P )- n n n 2 n F convergence of (δ (D X ))n N and the strong L (Ω, , P )-convergence of (X )n N, ∈ F ∈ ∞ E[(Dn Xn)2]dt = E[δn(DnXn)Xn] E[δ(DX)X]= ∞ E[(D X)2]dt, n → t Z0 ⌈ ·⌉ Z0 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 19. Suppose (Z )n N converges strongly to Z in L (Ω [0, )) and assume that n ∈ × ∞ n 2,2 ⌈ ·⌉ Z Ln , i.e., for every n N, ∈ ∈ 1 ∞ 1 ∞ E DnZn 2 + E DnDnZn 2 < . n2 | i j | n3 | i j k | ∞ i,j=1,i=j i,j,k=1, i,j,k =3 X6   X|{ }|   Then the following assertions are equivalent: 1 E n n n n (i) supn N n2 i,j∞=1,i=j (Di Zj )(Dj Zi ) < and ∈ 6 ∞  P h i 1 ∞ 1 E E n n n 2 E n n n n n n sup 2 [Di Zj j] + 3 (Di Dj Zk )(Dk Dj Zi ) < . n N n |F− n  ∞ ∈ i,j=1,i=j i,j,k=1, i,j,k =3 X6   X|{ }|    1,2 n n 2  (ii) Z D(δ), δ(Z) D , (δ (Z ))n N converges to δ(Z) strongly in L (Ω, , P ) and n∈ n n ∈ ∈ 2 F (D n δ (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. [Nualart (2006), Proposition 1.3.8].

n 1,2 n n n Proposition 20. Suppose Z Ln . Then (Di Z )1N i D(δ ) for every i N, and ∈ \{ } ∈ ∈ n n n n n n n n Di δ (Z )= E[Zi i]+ δ (Di Z 1N i ). |F− \{ } n Proof. By (19) and the continuity of Di , n n n n n n n n ξi ∞ n n n ξj Di δ (Z )= Di E[Zi i] + Di E[Zj j] , |F− √n |F− √n   j=1,j=i   X6 16 C. BENDER AND P. PARCZEWSKI

(including the strong convergence of the series on the right-hand side in L2(Ω, , P )). By (22), for i = j, F 6 n n n n n n n n n n n n n n n E[ξi E[Zj j]ξj i]= E[E[ξi Zj j] i]ξj = E[E[ξi Zj i] j]ξj . |F− |F− |F− |F− |F− |F− Moreover, n 2 n n n n n E[(ξi ) E[Zi i] i]= E[Zi i]. |F− |F− |F− Hence, n n n n n n ∞ n n n ξj Di δ (Z )= E[Zi i]+ E[Di Zj j] , |F− |F− √n j=1,j=i X6 and the closedness of the discrete Skorokhod integral concludes.  2,2 n n Proof of Theorem 19. The Ln -assumption guarantees that, for every i N, (Di Z )1N i 1,2 n 2 ∈ \{ } ∈ Ln . As (Z )n N is norm bounded in L (Ω [0, )) by the assumed strong convergence to n ∈ × ∞ Z, we observe⌈ ·⌉ in view of Propositions 16 and 20 that (i) is equivalent to n n 2 1 n n n 2 (i’) supn N E[ δ (Z ) ] < and supn N n i∞=1 E[ Di δ (Z ) ] < . ∈ | | ∞ ∈ | | ∞ Thanks to Theorems 9 and 13, assertion (i’) is equivalent to P 1,2 n n 2 (ii’) Z D(δ), δ(Z) D , (δ (Z ))n N converges weakly to δ(Z) in L (Ω, , P ), and n∈ n n ∈ ∈ 2 F (D δ (Z ))n N converges to Dδ(Z) weakly in L (Ω [0, )). ∈ n × ∞ n n n Due to the strong convergence of (Z )n N to Z and the weak convergence of (D δ (Z ))n N n ∈ n ∈ to D(δ(Z)), the continuous time duality⌈ ·⌉ between Skorokhod integral and Malliavin⌈ ·⌉ derivative and its discrete time counterpart in Proposition 14 imply

n n 2 ∞ n n n n ∞ 2 δ (Z ) 2 = E[Z D δ (Z )]ds E[Z D δ(Z)]ds = δ(Z) 2 . k kL (Ω, ,P ) ns ns → s s k kL (Ω, ,P ) F Z0 ⌈ ⌉ ⌈ ⌉ Z0 F 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). ∈ 

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, , P ) with respect to the Brownian motion B. n 2F n n Suppose Z Ln(Ω N) is predictable with respect to ( i )i N, i.e., for every i N, Zi is ∈ × n n n F ∈ ∈ measurable with respect to i 1 = σ(ξ1 ,...,ξi 1). Then, F − − ∞ ξn δn(Zn)= Zn i =: ZndBn, i √n Xi=1 Z which means that the discrete Skorokhod integral reduces to the discrete Itˆointegral. Analo- 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 ∈ × ∞ R ( t)t [0, ), see, e.g. [Janson (1997), Theorem 7.41]. F ∈ ∞ In this case of predictable integrands, the approximation theorem for Skorokhod integrals (The- orem 9) can be improved as follows. Theorem 21. Suppose Z L2(Ω [0, )) is predictable with respect to the augmented Brownian ∈ × ∞ n 2 n filtration ( t)t [0, ), and, for every n N, Z Ln(Ω N) is predictable with respect to ( i )i N. Then, theF following∈ ∞ are equivalent: ∈ ∈ × F ∈ n 2 (i) (Z n )n N converges to Z strongly (resp. weakly) in L (Ω [0, )). ⌈ ·⌉ ∈ × ∞ DISCRETIZING MALLIAVIN CALCULUS 17

n n (ii) The sequence of discrete Itˆointegrals Z dB n N converges strongly (resp. weakly) ∈ in L2(Ω, , P ) to ∞ Z dB . F 0 s s R  Remark 22. We note that,R 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 [Kurtz and Protter (1991)] immediately imply the following result in our setting: Suppose that Z is predictable with respect to the Brownian filtration and its paths are right-continuous with left limits. Moreover, assume that n n n Z is predictable with respect to ( i )i N and (Z ) converges to Z uniformly on compacts F ∈ 1+n( ) in probability. Then, ⌊ · ⌋ n ⌊ ·⌋ n 1 n · lim Zi ξi = Zs dBs, n − →∞ √n 0 Xi=1 Z uniformly on compacts in probability. In contrast, our Theorem 21 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 21 will make use of the following proposition. Proposition 23. Suppose g, h . Then, strongly in L2(Ω [0, )), ∈E × ∞ n n n ˇn lim exp⋄ (I (ˇg 1[1, n 1]))h ( n ) = 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, ∈

∞ 2 n n n ˇn E exp⋄ (I (ˇg 1[1, nt 1]))h ( nt ) exp⋄(I(g1(0,t]))h(t) dt ⌈ ⌉− ⌈ ⌉ − Z0 M h  i n n n 2 2 2 E exp⋄ (I (ˇg 1[1, nt ])) exp⋄(I(g1(0,t])) h(t) dt ≤ ⌊ ⌋ − Z0 h  i ∞ 2 n n n ˇn 2 +2 E exp⋄ (I (ˇg 1[1, nt 1])) h ( nt ) h(t) dt, 0 ⌈ ⌉− | ⌈ ⌉ − | Z h i since nt 1= nt for Lebesgue almost every t 0. As, by (4) ⌈ ⌉− ⌊ ⌋ ≥ n n n 2 n 2 sup E exp⋄ (I (ˇg 1[1, nt 1])) sup exp( gˇ ( n ) L2([0, ))) < , (23) n N, t [0, ) ⌈ ⌉− ≤ n N k ⌈ ·⌉ k ∞ ∞ ∈ ∈ ∞ h  i ∈ the second term goes to zero by (7). Moreover, by the boundedness of h, the first one tends to zero by the dominated convergence theorem, since, for every t [0, ), by Proposition 3, ∈ ∞ n n n 2 lim E exp⋄ (I ((ˇg 1[1, nt ]))) exp⋄(I(g1(0,t])) = 0. n ⌊ ⌋ →∞ − h  i  Proof of Theorem 21. ‘(i) (ii)’: By the isometry for discrete Itˆointegrals, we have ⇒ 2 2 n n ∞ n 1 n ∞ n 2 E Z dB = E Zi ξi = E Z ns ds. (24) " #  √n  0 | ⌈ ⌉| Z  i=1 Z h i X n  2  Hence, if (Z n )n N converges to Z weakly in L (Ω, , P ), then the left-hand side in (24) is ⌈ ·⌉ ∈ F bounded in n N, and so Theorem 9 implies the asserted weak L2(Ω, , P ) convergence of ∈ n F the sequence of discrete Itˆointegrals to 0∞ ZsdBs. If(Z n )n N converges to Z strongly in ⌈ ·⌉ ∈ L2(Ω, , P ), then, by (24) and the continuous time Itˆoisometry, F R 2 2 n n ∞ 2 ∞ lim E Z dB = E Zs ds = E ZsdBs , n | | →∞ "Z  # Z0 "Z0  #   18 C. BENDER AND P. PARCZEWSKI which turns the weak L2(Ω, , P )-convergence of the sequence of discrete Itˆointegrals into strong L2(Ω, , P )-convergence.F F ‘(ii) (i)’: We first assume that the sequence of discrete Itˆointegrals converges weakly in L2(Ω⇒, , P ) to the continuous time Itˆointegral. By the implication ‘(i) (ii)’ (which we have alreadyF proved) and Proposition 23, we obtain, for every g, h , ⇒ ∈E ∞ 1 n n n ˇn n ∞ lim exp⋄ (I (ˇg 1[1,i 1]))h (i) ξi = exp⋄(I(g1(0,s])) h(s) dBs (25) n − √ →∞ n 0 Xi=1 Z strongly in L2(Ω, , P ). As Zn is predictable and F n n n n n n n E[exp⋄ (I (ˇg )) i 1] = exp⋄ (I (ˇg 1[1,i 1])), |F − − we get, for every g, h , by the discrete Itˆoisometry, ∈E 1 ∞ 1 ∞ n n n ˇn n n n n n ˇn (S Zi )(ˇg )h (i) = E[E[Zi i 1]exp⋄ (I (ˇg ))h (i)] n n |F − Xi=1 Xi=1 1 ∞ n n n n ˇn = E[Zi exp⋄ (I (ˇg 1[1,i 1]))h (i)] n − Xi=1 ∞ 1 ∞ 1 n n n n n ˇn n = E Zi ξi exp⋄ (I (ˇg 1[1,i 1]))h (i) ξi . " √n ! − √n !# Xi=1 Xi=1 The assumed weak L2(Ω, , P )-convergence of the sequence of discrete Itˆointegrals and the strong L2(Ω, , P )-convergenceF in (25) now imply F 1 ∞ n n n ˇn ∞ ∞ lim (S Z )(ˇg )h (i)= E ZsdBs exp⋄(I(g1(0,s])) h(s) dBs . n i →∞ n 0 0 Xi=1 Z Z  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,

1 ∞ n n n n ∞ lim (S Z )(ˇg )hˇ (i)= (SZs)(g)h(s)ds, g, h . n i →∞ n 0 ∈E Xi=1 Z We can now apply Theorem 11. As ∞ E Zn 2 ds is bounded in n by (24) and by the 0 | ns | 2 ⌈ ⌉ assumed weak L (Ω, , P )-convergenceR of theh discretei Itˆointegrals, the latter Theorem im- n F 2 plies that (Z n )n N converges to Z weakly in L (Ω [0, )). If we instead assume strong ⌈ ·⌉ ∈ × ∞ L2(Ω, , P )-convergence of the sequence of the discrete Itˆointegrals, a straightforward applica- tion ofF the isometries for discrete and continuous-time Itˆo integrals 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, , P ) 2 ∈ F there is a unique ( t)t [0, )- X L (Ω [0, )) such that F ∈ ∞ ∇ ∈ × ∞ X = E[X]+ ∞ XdB . (26) ∇s s Z0 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’s isometry the operator : L2(Ω, , P ) L2(Ω [0, )) is continuous≥ with norm∈ 1. ∇ F → × ∞ Except in the case of binary noise, the discrete time approximation B(n) of the Brownian motion n B does not satisfy the discrete time predictable representation property with respect to ( i )i N. F ∈ DISCRETIZING MALLIAVIN CALCULUS 19

Nonetheless one can consider the discrete time predictable projection of the discretized Malliavin derivative n n n n n 2 i X := E[Di X i 1]= √nE[ξi X i 1], X L (Ω, , P ), i N, ∇ |F − |F − ∈ F ∈ 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., [Briand et al. (2002), Zhang (2004),Geiss et al. (2012)]. The operator n 2 2 n : L (Ω, , P ) Ln(Ω N), X ( i X)i N ∇ F → × 7→ ∇ ∈ is continuous with norm one. Indeed, introducing the shorthand notation E [ ]= E[ n] and n,i · ·|Fi noting that the martingale (En,i[X])i N is, for fixed n N, uniformly integrable, and, thus, converges almost surely to E[X n], as∈ i tends to infinity,∈ one gets, by H¨older’s and Jensen’s inequality, |F

1 ∞ n 2 ∞ n 2 E √n En,i 1 [ξi X] = E (En,i 1 [ξi (En,i[X] En,i 1[X])]) n − − − − i=1 i=1 X h  i X h i ∞ n 2 2 E En,i 1 (ξi ) En,i 1 (En,i[X] En,i 1[X]) ≤ − − − − i=1 X h   h ii ∞ 2 2 n 2 2 = E (En,i[X]) (En,i 1[X]) = E (E[X ]) E [X] " − − # |F − Xi=1   h i E (X)2 E [X]2 . ≤ − We now denoteh by i

n := a + ZndBn; a R,Zn L2 (Ω N) predictable P ∈ ∈ n ×  Z  the closed subspace in L2(Ω, , P ), which admits a discrete time predictable integral represen- F 2 n n 2 tation. Note that, for every X L (Ω, , P ), a R, and ( i )i N-predictable Z Ln(Ω N), by the discrete Itˆoisometry, ∈ F ∈ F ∈ ∈ ×

n n 1 ∞ n n n E X a + Z dB = aE[X]+ E[Xξi E[Zi i 1]] √n |F −   Z  Xi=1 1 ∞ n n n n n n n = aE[X]+ E Zi √nE[Xξi i 1] = E E[X]+ XdB a + Z dB . n |F − ∇ i=1  Z  Z  X   Hence, n n π n X = E[X]+ XdB , (27) P ∇ Z where, for any closed subspace in L2(Ω, , P ), π denotes the orthogonal projection on . Our first approximation result forA the Clark-OconeF A derivative now reads as follows: A n 2 2 Theorem 24. Suppose (X )n N is a sequence in L (Ω, , P ) and X L (Ω, , P ). Then, the following are equivalent, as n tends∈ to infinity: F ∈ F n n 2 (i) (π n X E[X ])n N converges to X E[X] strongly (weakly) in L (Ω, , P ). Pn n− ∈ − 2 F (ii) ( n X )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, , P ). ∈ F Proof. Recall that by (26) and (27)

X E[X] = ∞ X dB , − ∇ s s Z0 20 C. BENDER AND P. PARCZEWSKI

n n n n n π n X E[X ] = X dB . P − ∇ Z n n n Hence, Theorem 21 provides the equivalence of (i) and (ii). As, for every g , exp⋄ (I (ˇg )) n by (5), the sufficient condition is a consequence of the following lemma.∈E ∈ P Lemma 25. Suppose that n, n N, are closed subspaces of L2(Ω, , P ) such that for every n N, A ∈ F ∈ n n n exp⋄ (I (ˇg )), g . { ∈E}⊂An 2 n n Then, strong (weak) L (Ω, , P )-convergence of (X )n N to X implies that (π n X )n N con- verges to X strongly (weakly)F in L2(Ω, , P ) as well. ∈ A ∈ F Proof. As, for every g , ∈E n n n n n n n n n n n n E[(π n X )exp⋄ (I (ˇg ))] = E[X π n (exp⋄ (I (ˇg )))] = E[X exp⋄ (I (ˇg ))], A A n n n n n n we obtain that (S X )(ˇg ) = (S (π n X ))(ˇg ). In the case of weak convergence, Theorem 1 now immediately applies, because A

n 2 n 2 E (π n X ) E (X ) . A ≤ In the case of strong convergence, weh also makei use ofh Theoremi 1, and note that by the already n established weak convergence of (π n X )n N and H¨older’s inequality, A ∈ n 2 n n n 2 E (π n X ) = E [X(π n X )] + E [(X X)(π n X )] E X , n . A A − A → →∞ h i    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 n the strong closure in L2(Ω, , P ) of the linear span of H F Ξn := Ξn := ξn, A N, A < , A i ⊆ | | ∞ ( i A ) Y∈ and emphasize that n = L2(Ω, n, P ), if and only if the noise distribution of ξ is binary. As Ξn consists of an orthonormalH basisF of n, every Xn n has a unique expansion in terms of this Hilbert space basis, which is calledH the Walsh decomposition∈H of Xn, n n n X = XAΞA, (28) A < | X| ∞ where Xn = E[XnΞn ] satisfies (Xn)2 < . The expectation and L2(Ω, , P )-inner A A A < A ∞ F product can be computed in terms| of| the∞ Walsh decomposition via E[Xn]= Xn and P ∅ E [XnY n]= XnY n, Xn,Y n n, (29) A A ∈H A < | X| ∞ cp. [Holden et al. (1992)]. A direct computation shows that the Walsh decomposition of a discrete Wick exponential is given by

n n n A /2 n n exp⋄ (I (f )) = n−| | f (i) ΞA. (30) A < i A ! | X| ∞ Y∈ In view of the M¨obius inversion formula [Aigner (2007), Theorem 5.5], we obtain, for every finite subset B of N, n B /2 B C n n Ξ = n| | ( 1)| |−| | exp⋄ (I (1 )). (31) B − C C B X⊆ Hence, the set exp n (In(ˇgn)), g is total in n. { ⋄ ∈ E} H DISCRETIZING MALLIAVIN CALCULUS 21

We now consider the finite-dimensional subspaces n := span Ξn , A 1,...,i , Hi { A ⊂ { }} and introduce, as a second approximation of the generalized Clark-Ocone derivative, the operator n 2 2 n ¯ : L (Ω, , P ) L (Ω N), X (π n ( X)) . n i 1 i i N ∇ F → × 7→ H − ∇ ∈ Notice that n n ¯ X = √nπ n (ξ X), i i 1 i ∇ H − n 2 if ξi X L (Ω, , P ). We are∈ now goingF to show the following variant of Theorem 24.

n 2 2 Theorem 26. Suppose (X )n N is a sequence in L (Ω, , P ) and X L (Ω, , P ). Then, the following are equivalent, as n tends∈ to infinity: F ∈ F n n 2 (i) (π n X E[X ])n N converges to X E[X] strongly (weakly) in L (Ω, , P ). ¯Hn n− ∈ − 2 F (ii) ( n X )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, , P ). ∈ F The proof is based on the simple observation that n n, i.e., for every Xn n, H ⊂ P ∈H ∞ 1 Xn = E[Xn]+ nXn ξn. (32) ∇i √n i Xi=1 n n n n In order to show this, we recall that exp⋄ (I (ˇg )), g is total in . Thus, by continuity of the discretized Clark-Ocone derivative{ and by the discre∈ E} te Itˆoisometry,H it suffices to show (32) in the case Xn = exp n (In(fˇn)) for f . A direct computation shows, ⋄ ∈E n n n n n n n n i exp⋄ (I (f )) = f (i)exp⋄ (I (f 1[1,i 1])), (33) ∇ − which in view of (5) completes the proof of (32).

Proof of Theorem 26. We first note that, for every X L2(Ω, , P ), ∈ F E[π n X] = E[X], (34) H ¯ n n i X = i (π n X). (35) ∇ ∇ H Indeed, as n n π n X = E[X]+ E[XΞA]ΞA, H 1 A < ≤|X| ∞ n n Eq. (34) is obvious. In order to prove (35), we recall first that i (π n X) i 1 (by (33) and continuity of the discretized Clark-Ocone derivative) and then∇ noteH that,∈ for H − every A 1,...,i 1 , ⊂ { − } n n n n n E ΞAE ξi X i 1 = E[ΞA i X]= E[ΞA i π n (X)] | F − ∪{ } ∪{ } H    n n n n 1 n = E ΞAE ξi π n (X) i 1 = E ΞA i (π n X) . H | F − √n∇ H      In particular, by (32), (34), and (35)

n n π n X = E[X]+ ¯ XdB , (36) H ∇ Z which is the analogue of (27). The proof of Theorem 24 can now be repeated verbatim with n replaced by n. P H We close this section with two remarks. 22 C. BENDER AND P. PARCZEWSKI

Remark 27. In view of Lemma 25 and the inclusion n n we observe that, for any sequence n 2 H ⊂ P (X )n N in L (Ω, , P ), ∈ F 2 lim Xn = X strongly (weakly) in L (Ω, , P ) n →∞ F 2 lim π n Xn = X strongly (weakly) in L (Ω, , P ) n P ⇒ →∞ F 2 lim π n Xn = X strongly (weakly) in L (Ω, , P ). n H ⇒ →∞ F In particular, by Theorems 24 and 26, if the sequence of discretized Clark-Ocone derivatives n n 2 ( X )n N converges to X strongly (weakly) in L (Ω [0, )), then so does the sequence ∇ n ∈ ∇ × ∞ ⌈ ·⌉ ¯ n n of modified discretized Clark-Ocone derivatives ( n X )n N . ∇⌈ ·⌉ ∈ Remark 28. The following result can be derived from [Briand et al. (2002), Theorem 5 and the examples in Section 5] under the additional assumption that E[ ξ 2+ǫ] < for some ǫ> 0 and on n | | 2 ∞ a finite time horizon: Strong convergence of (X )n N to X in L (Ω, , P ) implies convergence of the sequence of discretized Clark-Ocone derivatives∈ as stated inF (ii) of Theorem 24. Our n Theorem 26 additionally shows that the conditional expectations E[ i 1] in the definition of the discretized Clark-Ocone derivative can be replaced by the projection·|F on− the finite dimensional n n 2 subspace i , i.e., if (X )n N converges to X strongly in L (Ω, , P ), then H ∈ F n n √nπ n (ξ (τ (X )) X nt 1 nt n H⌈ ⌉− ⌈ ⌉ t [0, ) →∇   ∈ ∞ 2 strongly in L (Ω [0, )), where τn denotes the truncation at n. We also note that,× in∞ view of (36), ±

(π n X) (π n X) i i 1 ¯ H − H − iX = 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 the generalized Clark-Ocone derivative which is derived in [Le˜ao× and∞ Ohashi (2013), Corollary 4.1], but for the case of symmetric binary noise only.

5. Strong L2-approximation of the chaos decomposition In this section, we apply Theorem 1 in order to characterize strong L2(Ω, , P )-convergence of a sequence (Xn) (where Xn can be represented via multiple Wiener integralsF 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, , P ) has a unique Wiener chaos decomposition in terms of multiple Wiener integrals ∈ F

∞ k k X = I (fX ), (37) Xk=0 k 2 k 2 k where fX L ([0, ) ), see e.g. [Nualart (2006), Theorem 1.1.2]. Here, we denote by L ([0, ) ) the Hilbert∈ space of∞ square-integrable functions with respect to the k-dimensional Lebesgue∞ mea- sure and byfL2([0, )k) the subspace of functions in L2([0, )k) which are symmetric in the ∞ ∞ k variables. We apply the standard convention L2([0, )0) = L2([0, )0) = R, I0(f 0) = f 0, ∞ ∞ and recall that,f for k 1 and f k L2([0, )k), the multiple Wiener integral can be defined as iterated Itˆointegral: ≥ ∈ ∞ f

tfk t2 k k ∞ k I (f )= k! f (t1,...,tk)dBt1 dBtk 1 dBtk . · · · · · · − Z0 Z0 Z0 DISCRETIZING MALLIAVIN CALCULUS 23

The Itˆoisometry therefore immediately implies the following well-konwn Wiener-Itˆoisometry for multiple Wiener integrals,

k k k′ k′ k k E 2 k [I (f ) I (g )] = δk,k′ k! f , g L ([0, ) (38) h i ∞ for functions f k L2([0, )k) and gk′ L2([0, )k′ ). The main theorem∈ of this∞ section now∈ reads as∞ follows: f f n 2 Theorem 29 (Wiener chaos limit theorem). Suppose (X )n N is a sequence in L (Ω, , P ). Then the following assertions are equivalent as n tends to infinity:∈ F n 2 (i) The sequence (π n X ) converges strongly in L (Ω, , P ). H F n,k (ii) For every k N0, the sequence (f n )n N, defined via ∈ X ∈ n nk/2 n n,k E X Ξ d , nu1 ,..., nu N = k, k! nu1 ,..., nuk k fXn (u1, . . . , uk) := {⌈ ⌉ ⌈ ⌉} |{⌈ ⌉ ⌈ ⌉} ∩ | (39) ( h 0, i otherwise, d is strongly convergent in L2([0, )k) and ∞ ∞ n,k 2 lim lim sup k! fXn L2([0, )k) = 0. (40) m n k k ∞ →∞ →∞ k=m X d n ∞ k k In this case, the limit X of (π n X )n N has the Wiener chaos decomposition X = I (fX ) H ∈ k=0 k n,k 2 k P with f = lim f n in L ([0, ) ). X n X →∞ ∞ We recall that, byd Remark 27, the strong L2(Ω, , P )-convergence of (Xn) to X is a sufficient condition for the strong approximation of the chaosF coefficients of X as stated in (ii) of the above theorem. Before proving Theorem 29, we briefly discuss this result. To this end, we first recall the relation between Walsh decomposition and discrete chaos decomposition. The discrete multiple Wiener integrals are defined analogously to the continuous setting, see e.g. [Privault (2009), Section 1.3]. For all k,n N we consider the Hilbert space ∈ 2 L2 (Nk) := f n,k : Nk R : f n,k(i ,...,i ) < n  → 1 k ∞ (i1,...,i ) Nk  Xk ∈    endowed with the inner product  n,k n,k k n,k n,k f , g 2 Nk := n− f (i1,...,ik)g (i1,...,ik). h iLn( ) (i1,...,i ) Nk Xk ∈ 2 k The closed subspace of symmetric functions in Ln(N ) which vanish on the diagonal part ∂ := (i ,...,i ) Nk : i ,...,i < k k 1 k ∈ |{ 1 k}| n o 2 k is denoted by Ln(N ). n,k 2 k Then, for k N, the discrete multiple Wiener integral of f Ln(N ) with respect to the random walk∈Bfn is defined as ∈ n,k n,k k/2 n,k f n I (f ) := n− k! f (i ,...,i )Ξ . 1 k i1,...,ik k { } (i1,...,i ) N , i1<

n,k n,k n,k′ n,k′ n,k n,k E[I (f )I (g )] = δk,k k! f , g 2 k (41) f ′ h iLn(N ) 24 C. BENDER AND P. PARCZEWSKI

n,k 2 k n,k′ 2 k′ for f Ln(N ), g Ln(N ) and possibly different orders k, k′ N. As in the continuous ∈ ∈ n,0 ∈ 2 0 time setting, we apply the convention that I is the identity on Ln(N ) := R, and refer to [Privaultf (2009), Sectionf 1.3] for further properties of such discrete multiple Wiener integrals. We now fix Xn n. In view of the Walsh decomposition Xn = f E[XnΞn ]Ξn , we ∈ H A < A A observe that the discrete analog of the Wiener chaos decomposition | | ∞ P k/2 n ∞ k/2 n n n ∞ n,k n,k X = n− k! X Ξ = I (f n ), (42) k! i1,...,ik i1,...,ik X k { } { } k=0 (i1,...,i ) N ,i1<

Remark 30. Convergence of discrete multiple Wiener integrals to continuous multiple Wiener integrals was studied in [Surgailis (1982)] 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 n,k n,k ∈ of discrete multiple Wiener integrals (I (f ))n N converges in distribution to the multiple ∈ k k n,k k 2 k Wiener integral I (f ), if (f )n N converges to f strongly in L ([0, ) ). Our result lifts this convergence in distribution to strong∈ L2(Ω, , P )-convergence and adds∞ the converse: d F L2(Ω, , P )- lim In,k(f n,k)= Ik(f k) L2([0, )k)- lim f n,k = f k. n n F →∞ ⇔ ∞ →∞ 2 n,k n,k We note that the L (Ω, , P )-convergence of the sequence (I (f )) evend implies convergence in Lp(Ω, , P ) for p >F2, if E[ ξ r] < for some r>p. Indeed, in this case, the sequence ( In,k(f n,kF) p) is uniformly integrable| | by the∞ hypercontractivity inequality of [Krakowiak and Szulga (1986)] in| the variant| of [Bai and Taqqu (2014), Proposition 5.2]. The following elementary corollary of Theorem 29 generalizes Proposition 3. It makes use of the fact that the chaos decompositions of (discrete) Wick exponentials are given, for all f L2([0, )), f n L2 (N), by ∈ ∞ ∈ n ∞ ∞ k 1 k n n n n,k 1 n k exp⋄(I(f)) = I ( f ⊗ ), exp⋄ (I (f )) = I ( (f )⊗ 1∂c ). (45) k! k! k Xk=0 Xk=0 For a proof of the continuous case see e.g. [Janson (1997), Theorem 3.21, Theorem 7.26]. The statement of the discrete case is a direct consequence of (30). DISCRETIZING MALLIAVIN CALCULUS 25

2 n n 2 Corollary 31. Suppose f L ([0, )) and (f ) is a sequence with f Ln(N) for every n N. Then, as n tends to infinity∈ (in the∞ sense of strong convergence), ∈ ∈ f n f in L2([0, )) In(f n) I(f) in L2(Ω, , P ) → ∞ ⇔ → F n n n 2 exp⋄ (I (f )) exp⋄(I(f)) in L (Ω, , P ). c ⇔ → F Proof. In view of Theorem 29 and (45), we only have to show that f n 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. → ∞ c≥  n,k n,k 2 k Lemma 32. (i) Fix k N0. Suppose (f )n N is a sequence such that f Ln(N ) for every ∈ ∈ ∈ n,k k 2 k \n,k n N and (f ) converges to some f strongly in L ([0, ) ). Then, the sequence (f 1∂c ) ∈ ∞ k converges to f k strongly in L2([0, )k) as well. ∞ dn n 2 n (ii) Suppose (f )n N is a sequence such that f Ln(N) for every n N and (f ) converges to ∈ ∈ ∈ 2 \n k n\k some f strongly in L ([0, )). Then, for every k 2, the sequences ((f ) ) and ((f ) 1∂c ) ∞ ≥ ⊗ ⊗ k converge to f k strongly in L2([0, )k). c ⊗ ∞ Proof. (i) We decompose,

\n,k k \n,k n,k n,k k f 1∂c f L2([0, )k) f 1∂c f L2([0, )k) + f f L2([0, )k). k k − k ∞ ≤ k k − k ∞ k − k ∞ The second term goes to zero by assumption. The first one equals d d 1/2 n,k 21 f ( nu1 ,..., nuk ) nu1 ,..., nuk

lim 1 nu1 ,..., nu

k n,k n,k k (SI (f ))(g)= f (x1,...,xk)g⊗ (x1,...,xk)dx1 dxk. [0, )k · · · Z ∞ d d 26 C. BENDER AND P. PARCZEWSKI

Hence, by the Cauchy-Schwarz inequality, we conclude (SnIn,k(f n,k))(ˇgn) (SIk(f n,k))(g) −

n,k d\n k k = f (x1,...,xk) (ˇg )⊗ g⊗ (x1,...,xk)dx1 dxk [0, )k − · · · Z ∞   d 1/2 m,k k \n k sup f 2 g (ˇg ) 2 k , Ln(N) ⊗ ⊗ L ([0, ) ) ≤ m N k k k − k ∞  ∈  which tends to zero for n by Lemma 32.  →∞ Corollary 34. Suppose g . Then, for every k N, ∈E ∈ n,k n k k k I ((ˇg )⊗ 1∂c ) I (g⊗ ) k → strongly in L2(Ω, , P ). F Proof. We check item (ii) in Theorem 1. To this end, we decompose, for every g, h , ∈E n n,k n k n k k (S I ((ˇg )⊗ 1∂c ))(hˇ ) (SI (g⊗ ))(h) k − n n,k n k n k n\k (S I ((ˇg )⊗ 1∂c ))(hˇ ) (SI ((ˇg ) 1∂ c ))(h) ≤ k − ⊗ k k n\k k k + (SI ((ˇg ) 1∂c ))(h) (SI (g⊗ ))(h) . ⊗ k −

The first term on the righthand side tends to zero by Proposition 33. The second one equals, by the isometry for multiple Wiener integrals,

k n\k k h⊗ (x) (ˇg ) 1 c g⊗ (x)dx ⊗ ∂k [0, )k − Z ∞   and goes to zero by Lemma 32. Consequently, n n,k n k n k k lim (S I ((ˇg )⊗ 1∂c ))(hˇ ) = (SI (g⊗ ))(h) n k →∞ n,k n k 2 k k 2 for all k N0 and g, h . For h = g, this implies E[I ((ˇg )⊗ 1∂c ) ] E[I (g⊗ ) ] by ∈ ∈ E 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 29. n 2 Proof of Theorem 29. ‘(i) (ii)’: We denote the limit of (π n X )n N in L (Ω, , P ) by X and recall that ⇒ H ∈ F n n n n ∞ n,k n,k π n X = E[X ΞA]ΞA = I (f n ), H X A < k=0 | X| ∞ X n,k with fXn as defined in (43). Throughout the proof we omit the subscripts from the coefficients n ∞ n,k n,k ∞ k k of the chaos decompositions and write π n X = I (f ) and X = I (f ). Thanks to H k=0 k=0 Corollary 34 and the orthogonality of (discrete) multipleP Wiener integralsP of different orders, we obtain, for every k N , ∈ 0 n n,k n,k n 1 n n,k n k 1 k k k k (S I (f ))(ˇg )= E[π n (X )I ((ˇg )⊗ 1∂c )] E[XI (g⊗ )] = (SI (f ))(g). k! H k → k! n,k n,k 2 n 2 The estimate supn N E[(I (f )) ] supn N E[(π n X ) ] < now yields, in view of The- 2 ∈ ≤ ∈n,k n,kH ∞ k k n orem 1, weak L (Ω, , P )-convergence of (I (f ))n N towards I (f ). As π n X X strongly in L2(Ω, , PF), we thus obtain ∈ H → F n,k n,k 2 n,k n,k n k k k k 2 E[(I (f )) ]= E[I (f )π n X ] E[I (f )X]= E[(I (f )) ]. (47) H → DISCRETIZING MALLIAVIN CALCULUS 27

n,k n,k k k 2 Hence, I (f ) I (f ) strongly in L (Ω, , P ) for all k N0 by Theorem 1. Due to the isometries (38) and→ (41), this implies F ∈

n,k 2 n,k n,k 2 k k 2 k 2 k! f L2([0, )k) = I (f ) L2(Ω, ,P ) I (f ) L2(Ω, ,P ) = k! f L2([0, )k). (48) k k ∞ k k F → k k F k k ∞ Moreover, for every g , we obtain d ∈E k n,k k k n,k k k g⊗ , f f L2([0, )k) = (SI (f ))(g) (SI (f ))(g) h − i ∞ − k n,k n n,k n,k n n,k n,k n n n k k = (SI (f ))(g) (S I (f ))(ˇg ) + E I (f )exp⋄ (I (ˇg )) I (f )exp⋄(I(g)) d − d − 0,  h i → d 2 n,k n,k k k by Propositions 3 and 33, and the L (Ω, , P )-convergence of (I (f ))n N to I (f ). Since F ∈ the set g k, g is total in L2([0, )k), we may conclude that (f n,k) converges weakly in { ⊗ ∈ E} ∞ L2([0, )k) to f k by [Yosida (1995), Theorem V.1.3]. Finally, (48) turns this weak convergence into strong∞ L2([0, )k)-convergence.f In particular, the kth coefficient ind the chaos decomposition ∞ off the limiting random variable X is the strong L2([0, )k)-limit of (f n,k), as asserted. It remains to show (40). However, by (47) and the isometries∞ for (discrete) multiple Wiener integrals, d

∞ ∞ n,k 2 n,k n,k 2 lim k! f 2 k = lim I (f ) 2 n L ([0, ) ) n L (Ω, ,P ) →∞ k k ∞ →∞ k k F kX=m kX=m d m 1 n 2 − n,k n,k 2 = lim π n X 2 I (f ) 2 n H L (Ω, ,P ) L (Ω, ,P ) →∞ k k F − k k F ! Xk=0 m 1 2 − k k 2 = X L2(Ω, ,P ) I (f ) L2(Ω, ,P ) 0 k k F − k k F → Xk=0 as m tends to infinity. n,k ‘(ii) (i)’: In order to lighten the notation, we again denote the function f n from (43) by ⇒ X f n,k. Assuming (ii), the strong L2([0, )k)-limit of f n,k exists and will be denoted f k. We first n,k n,k k∞ k 2 show that (I (f )) converges to I (f ) strongly in L (Ω, , P ) for all k N0 by means of Theorem 1. To this end, we observe that, for every gd , F ∈ ∈E (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)  → d d by Proposition 33 and the isometry for continuous multiple Wiener integrals. Moreover, again, by the isometries for discrete and continuous multiple Wiener integrals,

n,k n,k 2 n,k 2 n,k 2 k 2 k k 2 E (I (f )) = k! f L2 (Nk) = k! f L2([0, )k) k! f L2([0, )k) = E (I (f )) . k k n k k ∞ → k k ∞ h i 2 n,k n,kh k i k So, Theorem 1 applies indeed. With thedL (Ω, , P )-convergence of I (f ) to I (f ) at hand, we can now decompose, for every m N, F ∈ 2 n ∞ k k lim sup E π n X I (f ) n  H −  →∞ k=0 X  m 1 m 1  − k k − n,k n,k 2 ∞ k k 2 3limsup I (f ) I (f ) L2(Ω, ,P ) + I (f ) L2(Ω, ,P ) ≤ n k − k F k k F →∞ Xk=0 Xk=0 kX=m ∞ n,k n,k 2 + I (f ) L2(Ω, ,P ) k k F ! kX=m 28 C. BENDER AND P. PARCZEWSKI

∞ ∞ k 2 n,k 2 = 3 k! f L2([0, )k) + 3limsup k! f L2([0, )k). (49) k k ∞ n k k ∞ kX=m →∞ kX=m By Fatou’s lemma, d

∞ ∞ ∞ k 2 n,k 2 n,k 2 k! f 2 k = k! lim f 2 k lim inf k! f 2 k . L ([0, ) ) n L ([0, ) ) n L ([0, ) ) k k ∞ →∞ k k ∞ ≤ →∞ k k ∞ kX=m kX=m kX=m n Hence, letting m tend to infinity in (49), wed observe, thanks to (40), thatd (π n X ) converges strongly in L2(Ω, , P ). H  F We close this section with an example. Example 35. Fix X L2(Ω, , P ). Theorem 29 with Xn = X for every n N, implies that the chaos coefficients f∈k , k NF , of X are given as the strong L2([0, )k)-limit∈ of X ∈ 0 ∞ k n n 1 B nu B( nu 1) n,k E ⌈ l⌉ − ⌈ l⌉− 1 f (u1, . . . , uk) := X nu1 ,..., nuk N =k . k! " 1/n !# {|{⌈ ⌉ ⌈ ⌉}∩ | } Yl=1 This formulad can be further simplified when X is T -measurable. Then, one can show, analo- F 2 gously to Example 5 (ii), that the sequence (π n X) converges to X strongly in L (Ω, , P ). nT H⌊ ⌋ F Applying Theorem 29 with the latter sequence, shows that the chaos coefficients f k , k N , are X ∈ 0 the strong L2([0, )k)-limit of ∞ k n n 1 B nu B( nu 1) n,k E ⌈ l⌉ − ⌈ l⌉− 1 f (u1, . . . , uk) := X nu1 ,..., nuk 1,..., nT =k . k! " 1/n !# {|{⌈ ⌉ ⌈ ⌉}∩{ ⌊ ⌋}| } Yl=1 d n,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 oned way to give a rigorous meaning of the heuristic formula k k 1 ˙ fX (u1, . . . , uk)= E X Bul , k! " !# Yl=1 where B˙ is white noise, which is called Wiener’s intuitive recipe in [Cutland and Ng (1991)]. 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

k 1 ∆bt1 ∆btk 2 f (◦t ,...,◦ t )= ◦E x(b) , t T = j∆t, 0 j < N , X 1 k k! ∆t · · · ∆t l ∈ { ≤ }    √ T 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 29) 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 n, i.e. on the subspace of random variables which admit a discrete chaos decomposition in termsH 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. DISCRETIZING MALLIAVIN CALCULUS 29

n 2 Theorem 36. Suppose (X )n N converges strongly in L (Ω, , P ) to X and, for every n N, n 1,2 ∈ F ∈ π n X Dn . Then the following are equivalent: H ∈ ∞ n,k 2 n,k (i) lim lim sup kk! f 2 k = 0 (with f as defined in (39)). m Xn L ([0, ) ) Xn →∞ n k=m k k ∞ 1→∞,2 n n 2 (ii) X D andP the sequenced (D (π n X ))n dN converges to DX strongly in L (Ω ∈ n H ∈ × [0, )) as n tends to infinity. ⌈ ·⌉ ∞ Note first, that by continuity of Dn for a fixed time i N, we get i ∈ n n n n n n n n n Di (π n X )= E[X ΞA]Di ΞA = √n E[X ΞA]ΞA i H \{ } A < A < ; i A | X| ∞ | | X∞ ∈ n n n = √n E[X ΞB i ]ΞB. ∪{ } B < ; i/B | | X∞ ∈ By the relation (42)–(43) between Walsh decomposition and discrete chaos decomposition, this identity can be reformulated as

n n ∞ n,k 1 n,k Di (π n X )= kI − (f n ( , i)). (50) H X · Xk=1 Hence, the isometry for discrete multiple Wiener integrals (41) implies

1 ∞ n n 2 ∞ n,k 2 E Di (π n X ) = kk! fXn L2([0, )k), (51) n | H | k k ∞ i=1 h i k=1 X X d i.e.,

n 1,2 ∞ n,k 2 π n X Dn kk! f n 2 k < . (52) H ∈ ⇔ k X kL ([0, ) ) ∞ k=1 ∞ X d This is in line with the characterization of the continuous Malliavin derivative in terms of the chaos decomposition, see e.g. [Nualart (2006)], which we show to be equivalent to Definition 12 in the Appendix:

1,2 ∞ k 2 X D kk! fX L2([0, )k) < , (53) ∈ ⇔ k k ∞ ∞ Xk=1 and, if this is the case,

∞ n,k 1 k ∞ 2 ∞ k 2 DtX = kI − (fX ( ,t)), a.e. t 0, E[(DtX) ]dt = kk! fX L2([0, )k). (54) · ≥ 0 k k ∞ Xk=1 Z Xk=1 After these considerations on the connection between (discretized) Malliavin derivative and (discrete) chaos decomposition, the proof of Theorem 36 turns out to be rather straightforward.

Proof of Theorem 36. By Theorem 29 (in conjunction with Remark 27), we observe that, for n,k k 2 k every k N0, (f )n N converges to fX strongly in L ([0, ) ). Hence, by (51), (53), and (54), ∈ Xn ∈ ∞

d ∞ n,k 2 ∞ k 2 (i) lim kk! fXn L2([0, )k) = kk! fX L2([0, )k) < ⇔ n k k ∞ k k ∞ ∞ →∞ k=1 k=1 X d X 1,2 1 ∞ n n 2 ∞ 2 X D and lim E Di (π n X ) = E[(DtX) ]dt. n H ⇔ ∈ →∞ n | | 0 Xi=1 h i Z Hence, the asserted equivalence is a direct consequence of Theorem 13.  30 C. BENDER AND P. PARCZEWSKI

We now wish to derive an analogous strong approximation result for the Skorokhod integral, which requires some additional notation. For every Zn L2 (Ω N) and k N , we denote ∈ n × ∈ 0 nk/2 n n n,k n,k E Z Ξ , i ,...,i N = k k! i i1,...,ik 1 k fZn (i1,...,ik, i) := fZn (i1,...,ik)= { } |{ }∩ | i ( h 0, i otherwise. n n Then, with π n Z := (π n Zi )i N, H H ∈

∞ n,k 2 n 2 k! fZn L2 (Nk+1) = π n Z L2 (Ω N) < , k k n k H k n × ∞ Xk=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 F (y ,...,y )= F (y ,...,y ), 1 k k! π(1) π(k) π X e n,k where the sum runs over the group of permutations of 1,...,k . With this notation, f n 1∂c { } Z k+1 2 k+1 is an element of Ln(N ). We can now state: e f n 2 n n Theorem 37. Suppose that, for every n N, Z Ln(Ω N) and π n Z D(δ ). Moreover, n ∈ ∈2 × H ∈ assume that (Z )n N converges to Z strongly in L (Ω [0, )). Then, the following assertions n ∈ × ∞ are equivalent: ⌈ ·⌉ ∞ n,k 1 c 2 (i) lim lim sup k! f n − 1∂ 2 k = 0. m Z k Ln(N ) →∞ n k=m k k →∞ n n 2 (ii) Z D(δ) andP(δ (eπ n Z )) converges to δ(Z) strongly in L (Ω, , P ) as n tends to infinity.∈ H F As a preparation of the proof we note that, for every M N, ∈ M n M n n n ξi n n n n ξi E[π n Zi M, i] = E[Zi ΞA] E[ΞA M, i] H |F − √n |F − √n i=1 i=1 A < X X | X| ∞ M M 1/2 n n n 1/2 n n n = n− 1 i/A E[Zi ΞA]ΞA i = n− E[Zi ΞB i ]ΞB { ∈ } ∪{ } \{ } i=1 A 1,...,M k=1 B 1...,M , B =k i B X ⊂{X } X ⊂{ X} | | X∈ M k 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,...,i ) N , i1<

n n ∞ n,k 1 2 π n Z D(δ ) k! f n − 1∂c 2 Nk < , (55) H ∈ ⇔ k Z k kLn( ) ∞ Xk=1 and, if this is the case, e

n n ∞ n,k n,k 1 δ (π n Z )= I (f n − 1∂c ), (56) H Z k Xk=1 n,0 N e i.e., fδn(π n Zn) = 0 and, for every k , H ∈ n,k n,k 1 f = f n − 1 c . δn(π n Zn) Z ∂k H e DISCRETIZING MALLIAVIN CALCULUS 31

For the proof of Theorem 37, we also provide the following variant of Theorem 29, ‘(i) (ii)’, for stochastic processes. ⇒ n 2 n Proposition 38. Suppose Z Ln(Ω N) for every n N and (Z n ) converges strongly 2 ∈ × ∈ ⌈k ·⌉ 2 k+1 in L (Ω [0, )) to Z as n tends to infinity. Define the functions fZ L ([0, ) ) via k × ∞ k ∈ ∞ fZ(t1,...,tk+1) := fZ (t1,...,tk). Then, for every k N0, as n tends to infinity, tk+1 ∈ fn,k fk Zn → Z strongly in L2([0, )k+1). ∞ d Proof. The proof largely follows the arguments in the proof of Theorem 29. We spell it out for sake of completeness. Let g, h . Then, by the isometry for (discrete) multiple Wiener integrals, Corollary 34, and (7), ∈ E

k 1 ∞ n,k n ⊗ ˇn n,k n k ˇn f n , (ˇg ) h = f n , (ˇg )⊗ 1∂c h (i) Z Zi k 2 Nk ⊗ L2([0, )k+1) n Ln( )   ∞ Xi=1 D E d d c 1 ∞ n n,k n k ˇn 1 ∞ n n,k n k ˇn = E[(π n Zi )I ((ˇg )⊗ 1∂c )]h (i)= E[Zi I ((ˇg )⊗ 1∂c )]h (i) n H k n k Xi=1 Xi=1 ∞ k k k k E[ZsI (g⊗ )]h(s)ds = f , g⊗ h . (57) Z 2 k+1 → 0 ⊗ L ([0, ) ) Z D E ∞ As 2 2 2 n,k ∞ n,k n,k n sup f n = sup E I (f n ) ds sup Z < , (58) Z Z ns n 2 n N L2([0, )k+1) n N 0 ⌈ ⌉ ≤ n N ⌈ ·⌉ L (Ω [0, )) ∞ ∈ ∞ ∈ Z   ∈ × ∞ k d n ⊗ n k 2 k+1 k (ˇg ) hˇ g⊗ h strongly in L ([0, ) ) by (7), and the set g⊗ h : g, h is total in⊗ the closed→ subspace⊗ of functions in∞L2([0, )k+1), which are symmetric{ ⊗ in the∈ first E} k ∞ d c n,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 2 k+1 → L ([0, ) ) →∞ L ([0, ) ) ∞ ∞ As d

1 ∞ n n,k n k ˇn 1 ∞ n n,k n,k n ˇn E[Zi I ((ˇg )⊗ 1∂c )]h (i) = (S I (fZn ))(ˇg )h (i), n k n i Xi=1 Xi=1 ∞ E k k ∞ k k [ZsI (g⊗ )]h(s)ds = (SI (fZs ))(g)h(s)ds, Z0 Z0 n,k n,k k k we may derive from (57)–(58) and Theorem 11, that I (f n ) converges to I (f ) weakly in Z n Z ⌈ ·⌉ · L2(Ω [0, )). Thus, × ∞ 2 n,k ∞ n,k n,k ∞ n,k n,k n f n = E I (f n )Zs ds + E I (f n )(Z Zs) ds Z Z ns Z ns ns L2([0, )k+1) 0 ⌈ ⌉ 0 ⌈ ⌉ ⌈ ⌉ − ∞ Z h i Z h i 2 d ∞ k k k E I (f )Zs ds = f . Zs Z 2 k+1 → 0 L ([0, ) ) Z h i ∞ 

Proof of Theorem 37. By the linearity of the embedding operator ( ), Minkowski inequality, · Proposition 38, and Lemma 32, we obtain, for every k N0, ∈ c n,k\ k n,k^ k 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 − 2 k+1 − 2 k+1 ≤ − 2 k+1 → L ([0, ) ) L ([0, ) ) L ([0, ) ) ∞ ∞ ∞ e e b e

32 C. BENDER AND P. PARCZEWSKI as n tends to infinity. Thus, due to Theorem 29 and (56), n n 2 (i) (δ (π n Z ))n N converges strongly in L (Ω, , P ). ⇔ H ∈ F Now, the implication ‘(ii) (i)’ is obvious, while the converse implication is a consequence of Theorem 9. ⇒  Remark 39. As a by-product of the proof of Theorem 37, we recover, thanks to Theorem 29, the well-known chaos decomposition of the Skorokhod integral as

∞ k k 1 δ(Z)= I (fZ− ). Xk=1 e 7. Binary noise In this section, 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. [Holden et al. (1992)], [Leitz-Martini (2000)], [Privault (2009)], and the references therein, usually with the aim to ex- plain the main ideas of Malliavin calculus by discussing the analogous operators in the simple toy setting. Note first that L2(Ω, n, P ) equals n in the binary case (and in this case only) by Fi Hi observing that both spaces have dimension 2i in this case. Hence, L2(Ω, n, P ) coincides with n F n for binary noise, and we can drop the orthogonal projections π n on in the statement ofH all previous results. In particular, every random variable Xn LH2(Ω, Hn, P ) then admits a chaos decomposition in terms of discrete multiple Wiener integrals,∈ and theF 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 [Privault (2009)]. In the binary case, the representations for the discrete Mallivain derivative and the discrete Skorokhod integral can be simplified considerably. Suppose Xn L2(Ω, n, P ). Then, there is n n n ∈ F a measurable map FXn : R∞ R such that X = FXn (ξ1 ,ξ2 ,...). A direct computation shows that, for every i N, → ∈ n n n n n Di X = √nE[ξi FXn (ξ1 ,ξ2 ,...) i] |F− √nb n n n n n n = FXn (ξ1 ,...,ξi 1, b, ξi+1,...) FXn (ξ1 ,...,ξi 1, 1/b, ξi+1,...) , (59) b2 + 1 − − − − n 2  hence, the Malliavin derivative becomes a difference operator. Moreover, for Z Ln(Ω N) and N N, the discrete Skorokhod integral can be rewritten as ∈ × ∈ N N ξn 1 δn(Zn1 )= Zn i (ξn)2DnZn, [1,N] i √n − n i i i Xi=1 Xi=1 n which can either be derived from [Privault (2009), Proposition 1.8.3] or by expanding Zi in its Walsh decomposition and noting that, for every finite subset A N, ⊂ n n 2 n n n n √nΞA i (ξi ) , i A n 2 n n ΞA E[ΞA i] √nξi = \{ } ∈ = (ξi ) Di ΞA. − |F− 0, i / A  ∈   Hence, for Zn L2 (Ω N) and N N, ∈ n × ∈ N n n n n n ξi δ (Z 1 )= FZn (ξ ,ξ ,...) [1,N] i 1 2 √n Xi=1 DISCRETIZING MALLIAVIN CALCULUS 33

n 2 (ξi ) b n n n n n n FZn (ξ1 ,...,ξi 1, b, ξi+1,...) FZn (ξ1 ,...,ξi 1, 1/b, ξi+1,...) . − √n(b2 + 1) i − − i − − (60) 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 √ n n ξi = n Bτi Bτi 1 , − − where, in the binary case,  

n n n b 1 τ := 0 , τ := inf s τ : B B n , , (61) 0 i i 1 s τi 1 − ≥ − − − ∈ √n b√n    n and the Brownian motion at the first-passage times τi can be simulated by the acceptance- rejection of [Burq and Jones (2008)]. We close this paper by a toy example which illustrates how to numerically compute Skorokhod integrals by our approximation results. Example 40. 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 Zi = sign(1/2 i/n) BnBn i (1 i/n) 1[1,n 1](i), i N, − − − − − ∈ n and note that (Z nt ) converges to Ztfor almost every t  0 in probability by (1). Hence, by ⌈ ⌉ ≥ n uniform integrability and dominated convergence, it is easy to check that (Z n )n N converges ⌈ ·⌉ ∈ to Z strongly in L2(Ω [0, )). We next observe that in the discrete chaos decomposition of × ∞

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, , P )-approximation as the number of time steps increases. F 34 C. BENDER AND P. PARCZEWSKI

δn(Zn), all the coefficient functions f n,k for k 4 vanish, because Zn is a polynomial of δn(Zn) ≥ i degree 2 in Bn. Hence, the tail condition in Theorem 37 is trivially satisfied and, consequently, n n 2 n (δ (Z ))n N converges to δ(Z) strongly in L (Ω, , P ). We now suppose that B is constructed via the Skorokhod∈ embedding (61) and simulate,F 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 ) 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 (60) 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, , P )- convergence takes place at the expected rate− of 1/2. F

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 12) and the more classical characterization in terms of the chaos decomposition, see (53)–(54). Proposition 41. Suppose X = Ik(f k ) L2(Ω, , P ). Then, the following are equivalent: k X ∈ F (i) There is a Z L2(Ω [0, )) such that for every g, h , P ∈ × ∞ ∈E ∞ ∞ (SZ )(g)h(s)ds = E X exp⋄(I(g)) I(h) g(s)h(s)ds . s − Z0   Z0  ∞ k 2 (ii) kk! fX L2([0, )k) < . k=1 k k ∞ ∞ P ∞ n,k 1 k If this is the case, then Zt = kI − (fX ( ,t)) for almost every t 0. k=1 · ≥ Proof. We first note that, forP every f, g , ∈E ∞ ∞ 1 k (^k 1) exp⋄(I(g)) I(h) g(s)h(s)ds = I ((g⊗ − h)), (62) − 0 (k 1)! ⊗  Z  Xk=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 ∞ k k 1 kfX (x)(g⊗ − h)(x) dx [0, )k ⊗ k=1 Z ∞ X 1/2 1/2 ∞ k 2 ∞ k 2(k 1) 2 − k! fX L2([0, )k) g L2 ([0, )) h L2([0, )) < . (63) ≤ k k ∞ ! (k 1)!k k ∞ k k ∞ ! ∞ Xk=1 Xk=1 − Hence, Fubini’s theorem implies

∞ k k 1 ∞ ∞ k k 1 kfX (x)(g⊗ − h)(x)dx = kfX (x,t)g⊗ − (x) h(t)dt, k ⊗ k 1 [0, ) 0 [0, ) − ! Xk=1 Z ∞ Z Xk=1 Z ∞ i.e., by (62) and the isometry for multiple Wiener integrals,

∞ ∞ ∞ k k 1 E X exp⋄(I(g)) I(h) g(s)h(s)ds = kfX (x,t)g⊗ − (x) h(t)dt − 0 0 [0, )k 1   Z  Z k=1 Z ∞ − ! X (64) DISCRETIZING MALLIAVIN CALCULUS 35 for every g, h . ∈E ‘(i) (ii)’: Assuming (i) and noting that (64) holds for every g, h , we observe that for every⇒g , α R, and Lebesgue-almost every s [0, ), ∈ E ∈E ∈ ∈ ∞ ∞ k 1 k 1 (k 1) ∞ k 1 k (k 1) α − fZ− ( ), g⊗ − L2([0, )k 1) = (SZs)(αg)= α − kfX ( ,s), g⊗ − L2([0, )k 1 . h s · i ∞ − h · i ∞ − Xk=1 Xk=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 is total in { ⊗ ∈ E} L2([0, )k), we obtain, for every k 1 and almost every s [0, ), ∞ ≥ ∈ ∞ k k 1 kf ( ,s)= f − . (65) f X · Zs Therefore, the isometry for multiple Wiener-Itˆointegrals implies

∞ k 2 ∞ 2 kk! fX L2([0, )k) = E[ Zs ]ds < . (66) k k ∞ 0 | | ∞ Xk=1 Z ∞ n,k 1 k ‘(ii) (i)’: Define Zt = kI − (fX ( ,t)). Assuming (ii), we observe by the first identity in ⇒ k=1 · (66) that Z belongs to L2P(Ω [0, )). By the isometry for multiple Wiener integrals and the chaos decomposition of a Wick× exponential∞ we get, for every g, h . ∈E ∞ ∞ ∞ k k 1 (SZs)(g)h(s)ds = kfX (x,t)g⊗ − (x)dx h(t)dt. k 1 0 0 [0, ) − ! Z Z Xk=1 Z ∞ Hence, (64) concludes. 

References [Aigner (2007)] Aigner, M. A course in Enumeration. Berlin, Heidelberg: Springer (2007). [Avram (1988)] Avram, F. Weak convergence of the variations, iterated integrals and Dolans-Dade exponentials of sequences of . Ann. Probab. 16 (1), 246–250 (1988). [Avram and Taqqu (1986)] Avram, F. and Taqqu, M. Symmetric polynomials of random variables attracted to an infinitely divisible law. Probab. Theory Relat. Fields 71 (4), 491–500 (1986). [Bai and Taqqu (2014)] Bai, S. and Taqqu, M. Generalized Hermite processes, discrete chaos and limit theorems. Stochastic Processes Appl. 124 (4), 1710–1739 (2014). [Briand et al. (2002)] Briand, P. and Delyon, B. and M´emin, J. On the robustness of backward stochastic differ- ential equations. Stochastic Processes Appl. 97, 229–253, (2002). [Burq and Jones (2008)] Burq, Z. A. and Jones, O. D., Simulation of Brownian motion at first-passage times. Math. Comput. Simulation 77, 64–71, (2008). [Coquet et al. (2001)] Coquet, F., M´emin, J., and Slominski, L. On weak convergence of filtrations. S´eminaire de probabilit´es 35, 306–328, (2001). [Cutland and Ng (1991)] Cutland, N. and Ng, S. On homogeneous chaos. Math. Proc. Cambridge Philos. Soc. 110 (2), 353–363, (1991). [Geiss et al. (2012)] Geiss, C. and Geiss, S. and Gobet, E. Generalized fractional smoothness and Lp-variation of BSDEs with non-Lipschitz terminal condition. Stochastic Process. Appl. 122 (5) 2078–2116 (2012). [Gzyl (2006)] Gzyl, H. An expos´eon discrete Wiener chaos expansions. Bol. Asoc. Mat. Venez. 13 (1), 3–27, (2006). [Holden et al. (1992)] Holden, H. and Lindstrøm, T. and Øksendal, B. and Ubøe, J. Discrete Wick products. Stochastic analysis and related topics (Oslo, 1992), Stochastics Monogr., 8, Gordon and Breach, Montreux, 123–148, (1993). [Holden et al. (2010)] Holden H. and Øksendal, B. and Ubøe, J. and Zhang, T. Stochastic Partial Differential Equations. A Modeling, White Noise Functional Approach. Second Edition New York: Springer (2010). [Jacod et al. (2000)] Jacod, J. and M´el´eard, S. and Protter, P. Explicit form and robustness of martingale repre- sentations. Ann. Probab. 28 (4), 1747–1780 (2000). [Jakubowski et al. (1989)] Jakubowski, A. and M´emin, J. and Pag`es, G. Convergence en loi des suites d’int´egrales stochastiques sur l’espace D1 de Skorokhod. Probab. Theory Related Fields 81 (1), 111–137 (1989). [Janson (1997)] Janson, S. Gaussian Hilbert Spaces., Cambridge: Cambridge University Press (1997). 36 C. BENDER AND P. PARCZEWSKI

[Krakowiak and Szulga (1986)] Krakowiak, W. and Szulga, J. Random multilinear forms. Ann. Probab. 14 (3), 955–973, (1986). [Kuo (1996)] Kuo, H.-H. White Noise Distribution Theory. Probability and Stochastics Series. Boca Raton, FL: CRC Press (1996). [Kurtz and Protter (1991)] Kurtz, T. G. and Protter, P. Weak limit theorems for stochastic integrals and sto- chastic differential equations. Ann. Probab. 19 (3), 1035–1080 (1991). [Le˜ao and Ohashi (2013)] Le˜ao, D. and Ohashi, A. Weak approximation for Wiener functionals. Ann. Appl. Probab. 23 (4), 1660–1691 (2013). [Leitz-Martini (2000)] Leitz-Martini, M. A discrete Clark-Ocone formula. Maphysto Research Report No 29 (2000). [M¨orters and Peres (2010)] M¨orters, P. and Peres, Y. Brownian motion Cambridge University Press, Cambridge, (2010). [Nualart (2006)] Nualart, D. The Malliavin Calculus and Related Topics. Second Edition. Probability and its Applications (New York). Springer (2006). [Nualart and Pardoux (1988)] Nualart, D. and Pardoux, E. with anticipating integrands. Probab. Theory Related Fields 78 (4), 535–581 (1988) [Privault (2009)] Privault, N. Stochastic Analysis in Discrete and Continuous Settings. Lecture Notes in Mathe- matics 1982. Berlin: Springer (2009). [Shiryaev (1996)] Shiryaev, A. N. Probability. Second edition. Graduate Texts in Mathematics 95. Berlin: Springer (1996). [Sottinen (2001)] Sottinen, T. Fractional Brownian motion, random walks and binary market models. Finance Stoch. 5, 343–355 (2001). [Surgailis (1982)] Surgailis, D. Domains of attraction of self-similar multiple integrals. Lithuanian Math. J. 22 (3), 327–340, (1982). [Yosida (1995)] Yosida, K. Functional analysis. Reprint of the sixth (1980) edition. Classics in Mathematics. Berlin. Springer (1995). [Zhang (2004)] Zhang, J. A numerical scheme for BSDEs. Ann. Appl. Probab. 14 (1), 459–488 (2004).

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]