On the convergence analysis of the inexact linearly implicit Euler scheme for a class of SPDEs ∗ P.A. Cioica, S. Dahlke, N. D¨ohring, U. Friedrich, S. Kinzel, 5 F. Lindner, T. Raasch, K. Ritter, R.L. Schilling 1 0 2 n a Abstract This paper is concerned with the adaptive numerical J treatment of stochastic partial differential equations. Our method of 3 choice is Rothe’s method. We use the implicit Euler scheme for the 2 time discretization. Consequently, in each step, an elliptic equation ] with randomright-handside has to be solved.In practice,this cannot R be performed exactly, so that efficient numerical methods are needed. P Well-establishedadaptivewaveletorfinite-elementschemes,whichare h. guaranteed to converge with optimal order, suggest themselves. We t investigate how the errors corresponding to the adaptive spatial dis- a cretization propagate in time, and we show how in each time step the m tolerances have to be chosen such that the resulting perturbed dis- [ cretization scheme realizes the same order of convergence as the one 1 with exact evaluations of the elliptic subproblems. v 6 MSC 2010: Primary: 60H15, 60H35; secondary: 65M22. 1 Key words: Stochastic evolution equation, stochastic partial differential 8 5 equation,Eulerscheme,Rothe’smethod,adaptivenumericalalgorithm,con- 0 vergence analysis. . 1 0 5 1 1 Introduction : v Xi Thispaperisconcernedwiththenumericaltreatmentofstochasticevolution equations of the form r a du(t) = Au(t)+f(u(t)) dt+B(u(t))dW(t), u(0) = u , (1) 0 (cid:0) (cid:1) on the time interval [0,T] in a real and separable Hilbert space U. Here, A : D(A) U U is a densely defined, strictly negative definite, self- ⊂ → adjoint,linearoperatorsuchthatzerobelongstotheresolventsetandthein- verseA−1 is compact on U.Theforcingterms f : D(( A)̺) D(( A)̺−σ) − → − ∗ThisworkhasbeensupportedbytheDeutscheForschungsgemeinschaft(DFG,grants DA 360/12-2, DA 360/13-2, DA 360/20-1, RI 599/4-2, SCHI 419/5-2) and a doctoral scholarship of thePhilipps-Universit¨at Marburg. 1 and B : D(( A)̺) (ℓ ,D(( A)̺−β)) are Lipschitz continuous maps for 2 − → L − suitable constants ̺, σ and β; and finally, W = (W(t)) is a cylindrical t∈[0,T] Wiener process on the sequence space ℓ = ℓ (N). In practical applications, 2 2 evolution equations of the form (1) are abstract formulations of stochastic partial differential equations (SPDEs, for short): Usually A is a differential operator, f a linear or nonlinear forcing term and B(u(t))dW(t) describes additive or multiplicative noise. They are models, e.g., for reaction diffusion processes corrupted by noise, which are frequently used for the mathemati- cal description of biological, chemical and physical processes. Details on the equation, the operators A, the forcing terms f and B and the initial condi- tion u are given in Section 2. Usually, the exact solution of (1) cannot be 0 computed explicitly, so thatnumericalschemesfor theconstructiveapproxi- mation of thesolutions areneeded.For stochastic parabolic equations, there aretwoprincipallydifferentapproaches:thevertical method of linesand the horizontal method of lines. The former starts with an approximation first in space and then in time. We refer to [23–25,27] for detailed information. The latter starts with a discretization first in time and then in space; it is alsoknownas Rothe’smethod.Inthestochastic setting,ithasbeen studied, e.g., in [6,22]. These references are indicative and by no means complete. Very often,thevertical method of linesis preferredsince,at firstsight,it seemstobealittlebitsimpler.Indeed,afterthedisretizationinspaceisper- formed, just an ordinary finite dimensional stochastic differential equation (SDE, for short) in time direction has to be solved, and there exists a huge amount of approaches for the numerical treatment of SDEs. However, there are also certain drawbacks.In many applications, the utilization of adaptive strategies to increase efficiency is almost unavoidable. In the context of the vertical method of lines, the combination with spatial adaptivity is at least not straightforward. In contrast, for the horizontal method of lines, the fol- lowing natural approach suggests itself. Using Rothe’s method, the SPDE canbeinterpretedasanabstractCauchyproblem,i.e.,asastochasticdiffer- ential equation in some suitable function spaces. Then, in time direction we might use an SDE-solver with step size control. This solver must be based on an implicit discretization scheme since the equation under consideration is usually stiff. Consequently, in each time step, an elliptic equation with random right-hand side has to be solved. To this end, as a second level of adaptivity, adaptive numerical schemes that are well-established for deter- ministic equations, can be used. We refer to [9,10,15] for suitable wavelet methods,andto[1–3,5,17–20,28,36,37]forthefiniteelementcase.Asbefore, these lists are not complete. Althoughthiscombinationwithadaptivestrategies isnatural,themath- ematical analysis of the resulting schemes seems to be still in its infancy. In the stochastic setting, Rothe’s method with exact evaluation of the elliptic subproblems, has been considered, e.g., in [6,22], and explicit convergence rates have been established, e.g., in [12,13,26,35]. First results concern- 2 ing the combination with adaptive space discretization methods based on wavelets have been shown in [31]. Even for the deterministic case, not many results concerning a rigorous convergence and complexity analysis of the overall scheme seem to be avail- able. To our best knowledge, the most far reaching achievements have been obtained in [7]. In this paper, it has been clarified how the tolerances for the elliptic subproblems in each time step have to be tuned so that the overall (perturbed) discretization scheme realizes the same order of convergence (in time direction) as the unperturbed one. Moreover, based on concepts from approximation theory and function space theory, respectively, a complexity analysis of the overallschemehas been derived.Itis theaim of this paper to generalize the analysis presentedin [7] to SPDEs of the form (1). We mainly consider the case of the implicit Euler scheme, and we concentrate on the convergence analysis. To our best knowledge, no result in this direction has been reported yet. Complexity estimates are beyond the scope of this work and will be presented in a forthcoming paper. For reader’s convenience,let us brieflyrecall the basicapproachof [7]for the deterministic case, confined to the implicit Euler scheme. As a typical example, let us consider the deterministic heat equation u′(t) = ∆u(t)+f(t,u(t)) on , t (0,T], O ∈ u = 0 on ∂ , t (0,T], O ∈ u(0) = u on , 0 O where Rd, d 1, denotes a bounded Lipschitz domain. We discretize O ⊂ ≥ thisequationbymeansofalinearlyimplicitEulerschemewithuniformtime steps. Let K N be the number of subdivisions of the time interval [0,T]. ∈ The step size will be denoted by τ := T/K, and the k-th point in time is denoted by t := τk, k 0,...,K . The linearly implicit Euler scheme, k ∈ { } starting at u , is given by 0 u u k+1 k − = ∆u +f(t ,u ), k+1 k k τ i.e., (I τ∆)u = u +τf(t ,u ), (2) k+1 k k k − for k = 0,...,K 1. If we assume that the elliptic problem − L v := (I τ∆)v = g on , v = 0, τ ∂O − O | can be solved exactly, then one step of the scheme (2) can be written as u = L−1R (u ), (3) k+1 τ τ,k k where R (w) := w+τf(t ,w) τ,k k 3 and L is a boundedly invertible operator between suitable Hilbert spaces. τ In practice, the elliptic problems in (3) cannot be evaluated exactly. Instead,weemploya‘blackbox’numericalscheme,whichforanyprescribed tolerance ε > 0 yields an approximation [v] of v := L−1R (w), where w is ε τ τ,k an element of a suitable Hilbert space, i.e., v [v] ε, ε k − k ≤ for a proper norm . What we have in mind are applications of adaptive k·k waveletsolvers, whichare guaranteed to convergewith optimal order, as de- veloped, e.g., in [9], combined with efficient evaluations of the nonlinearities f as they can be found, e.g., in [11,16,30]. In [7] we have investigated how the error propagates within the linearly implicit Euler scheme and how the tolerances ε in each time step have to be chosen, such that we obtain the k same order of convergence as in the case of exact evaluation of the ellip- tic problems. We have shown that the tolerances depend on the Lipschitz Lip constants C of the operators τ,j,k E = (L−1R ) (L−1R ) (L−1R ), τ,j,k τ τ,k−1 ◦ τ τ,k−2 ◦···◦ τ τ,j with 1 j k K, K N, via ≤ ≤ ≤ ∈ k−1 Lip u(t ) u˜ u(t ) u + C ε , k k − kk≤ k k − kk τ,j+1,k j j=0 X where u˜ is the solution to the inexactly evaluated Euler scheme at time t . k k Now let us come back to SPDEs of the form (1). Once again, for the (adaptive) numerical treatment of (1) we consider for K N and τ := T/K ∈ the linearly implicit Euler scheme u =(I τA)−1 u +τf(u )+√τB(u )χ , k+1 k k k k − (4) (cid:0) k = 0,...,K 1(cid:1), ) − with 1 χ := χK := W tK W tK , k k √τ k+1 − k (cid:0) (cid:0) (cid:1) (cid:0) (cid:1)(cid:1) where t := τk, k = 0,...,K. If we set k R (w) := w+τf(w)+√τB(w)χ , k = 0,...,K 1, τ,k k − L−1w := (I τA)−1w, k = 1,...,K, τ − the operators being defined between suitable Hilbert spaces and , the k k H G scheme (4) can again be rewritten as u = L−1R (u ), k = 0,...,K 1. (5) k+1 τ τ,k k − 4 We refer to Section 3 for a precise formulation of this scheme. Once again the elliptic problems in (5) cannot be evaluated exactly. Similar to the deterministic setting, we assume that we have at hand a ‘black box’ numerical scheme, which for any required w approximates v := (I τA)−1 w+f(w)+B(w)χ k − with a prescribed tolerance ε > 0.(cid:0)What we have in mi(cid:1)nd are applications of some deterministic solver for elliptic equations to individual realizations, e.g., an optimal adaptive wavelet solver as developed in [9], combined with proper evaluations of the nonlinearities f and B, see, e.g., [11,16,30], and an adequate truncation of the noise. It is the aim of this paper to inves- tigate how the error propagates within the linearly implicit inexact Euler scheme for SPDEs (cf. Proposition 4.3) and how the tolerances ε in each k time step have to be chosen, such that we obtain the same order of conver- gence (in time direction) for the inexact scheme as for its exact counterpart (cf. Theorem 4.2). Concerning the setting, we follow [35] and impose rather restrictive con- ditions on the different parts of Eq. (1). This allows us to focus on our main goal, i.e., the analysis of the error of the inexact counterpart of the Euler scheme(4), without spendingtoo muchtime on explainingdetails regarding the underlying setting, cf. Remark 2.12. Compared with [35] we allow the spatial regularity of the whole setting to be ‘shifted’ in terms of the addi- tional parameter ̺. In concrete applications to parabolic SPDEs, this will lead to estimates of the discretization error in terms of the numerically im- portant energy norm, cf. Example 2.11, provided that the initial condition u and the forcing terms f and B are sufficiently regular. 0 A different approach has been presented in [31], where additive noise is considered, a splitting method is applied, and adaptivity is only used for the deterministic part of the equation. We remark that the use of spatially adaptive schemes is useful especially for stochastic equations, where singu- larities appear naturally near the boundary due to the irregular behaviour of the noise, cf. [8] and the references therein. We choose the following outline. In Section 2 we present the setting and some examples of equations that fit into this setting. In Section 3 we show how to reformulate the linearly implicit Euler scheme as an abstract Rothe scheme and derive convergence rates under the assumption that we can evaluate the subproblems (5) exactly. We drop this assumption in Section 4 and focus on how to choose the tolerances for each subproblem, such that we can achieve the same order of convergence. 2 Setting In this section we describe the underlyingsetting in detail. It coincides with the one in [35] (‘shifted’ by ̺ 0). Furthermore we define the solution ≥ 5 concept under consideration and give some examples of equations, which fit into this setting. We start with assumptions on the linear operator in Eq. (1). Assumption 2.1. The operator A : D(A) U U is linear, densely de- ⊂ → fined,strictlynegativedefiniteandself-adjoint.Zerobelongstotheresolvent set of A and the inverse A−1 : U U is compact. There exists an α > 0 → such that ( A)−α is a trace class operator on U. − To simplify notation, the separable real Hilbert space U is always as- sumed to be infinite-dimensional. Under the assumption above, it follows that A enjoys a spectral decomposition of the form Av = λ v,e e , v D(A), (6) j j U j h i ∈ j∈N X where (ej)j∈N is an orthonormal basis of U consisting of eigenvectors of A with strictly negative eigenvalues (λj)j∈N such that 0 > λ λ ... λ , j . (7) 1 2 j ≥ ≥ ≥ → −∞ → ∞ For s 0 we set ≥ ∞ D(( A)s) := v U : ( λ )s v,e 2 < , (8) j j U − ∈ − h i ∞ n Xj=1(cid:12) (cid:12) o ( A)sv := ( λ )s v,e(cid:12) e , v D(cid:12)(( A)s), (9) j j U j − − h i ∈ − j∈N X so that D(( A)s), endowed with the norm := ( A)s , is a D((−A)s) U − k·k k − ·k Hilbert space; by construction this norm is equivalent to the graph norm of ( A)s. − For s < 0wedefineD(( A)s)as thecompletion of U withrespect to the − norm , defined on U by v 2 := ( λ )s v,e 2. k·kD((−A)s) k kD((−A)s) j∈N − j h jiU Thus, D(( A)s) can be considered as a space of formal sums − P (cid:12) (cid:12) (cid:12) (cid:12) v = v(j)e , such that ( λ )sv(j) 2 < j j − ∞ j∈N j∈N X X(cid:12) (cid:12) (cid:12) (cid:12) with coefficients v(j) R.Generalizing (9)in theobvious way,weobtain op- ∈ erators ( A)s, s R, which map D(( A)r) isometrically onto D(( A)r−s) − ∈ − − for all r R. ∈ The trace class condition in Assumption 2.1 can now be reformulated as the requirement that there exists an α > 0 such that Tr( A)−α = ( λ )−α < . (10) j − − ∞ j∈N X 6 Note that any linear operator with a spectral decomposition as in (6) and eigenvalues as in (7) and (10) fulfills Assumption 2.1. Let us consider a prime example of such an operator. Throughout this paper, we write L ( ) 2 O for the space of quadratically Lebesgue-integrable real-valued functions on a Borel-measurable subset of Rd. Furthermore, (U ;U ) stands for the 1 2 O L space of bounded linear operators between two Hilbert spaces U and U . If 1 2 the Hilbert spaces coincide, we simply write (U ) instead of (U ;U ). 1 1 1 L L Example 2.2. Let be a bounded open subset of Rd, set U := L ( ) and 2 O O let A:= ∆D be the Dirichlet-Laplacian on , i.e., O O ∆D : D(∆D) L ( ) L ( ) O O ⊆ 2 O → 2 O with domain d ∂2 D(∆D) = u H1( ) :∆u:= u L ( ) , O ∈ 0 O ∂x2 ∈ 2 O n Xi=1 i o where H1( ) stands for the completion in the L ( )-Sobolev space of or- 0 O 2 O der one of the set ∞( ) of infinitely differentiable functions with com- C0 O pact support in . Note that this definition of the domain of the Dirichlet- O Laplacian is consistent with the definition of D(( ∆D)s) for s = 1 in (8), − O see, e.g., [32, Remark 1.13] for details. This linear operator fulfills Assump- tion 2.1 for all α > d/2: It is well-known that it is densely defined, self- adjoint, and strictly negative definite. Furthermore it possesses a compact inverse(∆D)−1 :L ( ) L ( ),see,e.g.,[21].Moreover,Weyl’s lawstates O 2 O → 2 O that λ j2/d, j N, j − ≍ ∈ see [4], implying that (10) holds for all α > d/2. Next we state the assumptions on the forcing terms f and B. Assumption 2.3. For certain smoothness parameters 1 α ̺ 0, σ < 1 and β < − (11) ≥ 2 (α as in Assumption 2.1), we have f : D(( A)̺) D(( A)̺−σ), − → − B : D(( A)̺) (ℓ ;D(( A)̺−β)). 2 − → L − Furthermore, f and B are globally Lipschitz continuous, that is, there exist positive constants CLip and CLip such that for all v, w D(( A)̺), f B ∈ − Lip kf(v)−f(w)kD((−A)̺−σ) ≤ Cf kv−wkD((−A)̺), and Lip kB(v)−B(w)kL(ℓ2;D((−A)̺−β)) ≤ CB kv−wkD((−A)̺). 7 Remark 2.4. (i) The parameters σ and β in Assumption 2.3 are allowed to be negative. (ii) Assumption 2.3 follows the lines of [35] (‘shifted’ by ̺ 0). The linear ≥ growth conditions (3.5) and (3.7) therein follow from the (global) Lipschitz continuity of the mappings f and B. Finally, we describe the noise and the initial condition in Eq. (1). For the notion of a normal filtration we refer to [34]. Assumption 2.5. The noise W = (W(t)) is a cylindrical Wiener t∈[0,T] process on ℓ with respect to a normal filtration (F ) . The underlying 2 t t∈[0,T] probability space (Ω,F,P) is complete. For ̺ as in Assumption 2.3, the initial condition u in Eq. (1) satisfies 0 u L (Ω,F ,P;D(( A)̺)). 0 2 0 ∈ − Inthispaperweconsideramildsolutionconcept.Tothisendlet(etA) t≥0 be the strongly continuous semigroup of contractions on U generated by A. Definition 2.6. A mild solution to Eq. (1) (in D(( A)̺)) is a predictable − process u :Ω [0,T] D(( A)̺) with × → − sup E u(t) 2 < , (12) k kD((−A)̺) ∞ t∈[0,T] such that for every t [0,T] the equality ∈ t t u(t) = etAu + e(t−s)Af(u(s))ds+ e(t−s)AB(u(s))dW(s) (13) 0 Z0 Z0 holds P-almost surely in D(( A)̺). − Remark 2.7. (i) Let u : Ω [0,T] D(( A)̺) be a predictable process × → − fulfilling (12). Then, the first integral in (13) is meant to be a D(( A)̺)- − valued Bochner integral for P-almost every ω Ω; the second integral is a ∈ D(( A)̺)-valued stochastic integral as defined, e.g., in [14,34]. Both inte- − grals exist due to (12) and Assumptions 2.1 and 2.3. For example, consider- ing the stochastic integral in (13), we know that it exists as an element of L (Ω,F ,P;D(( A)̺)) if the integral 2 t − t E e(t−s)AB(u(s)) 2 ds (14) Z0 LHS(ℓ2;D((−A)̺)) (cid:13) (cid:13) is finite, where (ℓ ;(cid:13)D(( A)̺)) deno(cid:13)tes the space of Hilbert-Schmidt op- HS 2 L − erators from ℓ to D(( A)̺). The integrand in (14) can be estimated from 2 − above by Tr( A)−α ( A)β+α/2e(t−s)A 2 E ( A)−βB(u(s)) 2 , − − L(D((−A)̺)) − L(ℓ2;D((−A)̺)) and we hav(cid:13)e (cid:13) (cid:13) (cid:13) (cid:13) (cid:13) (cid:13) (cid:13) ( A)β+α/2e(t−s)A 2 C(t s)−(2β+α) − L(D((−A)̺)) ≤ − (cid:13) (cid:13) (cid:13) (cid:13) 8 for 2β +α [0,1). For 2β +α < 0 we have ( λ )2β+α ( λ )2β+α for all j 1 ∈ − ≤ − j N , which yields 0 ∈ ( A)β+α/2e(t−s)A 2 C. − L(D((−A)̺)) ≤ Moreover,bytheglo(cid:13)(cid:13)balLipschitzcontin(cid:13)(cid:13)uityofthemappingB : D(( A)̺) − → (ℓ ;D(( A)̺−β)), 2 L − E ( A)−βB(u(s)) 2 C 1+ sup E u(r) 2 . − L(ℓ2;D((−A)̺)) ≤ r∈[0,T] k kD((−A)̺) (cid:13) (cid:13) (cid:16) (cid:17) (cid:13) (cid:13) Thus, the stochastic integral in (13) is well-defined. (ii) For the case ̺ = 0 existence and uniqueness of a mild solution to Eq. (1) has been stated in [35, Proposition 3.1]. The proof consists of a modification of the proof of Theorem 7.4 in [14]—a contraction argument in L ([0,T];L (Ω;U)). For the general case ̺ 0 existence and uniqueness ∞ 2 ≥ can be proved analogously, see [29, Theorem 5.1]. Alternatively, the case ̺ > 0 can be traced back to the case ̺ = 0 as described in the proof of Proposition 2.8 below. Proposition2.8. LetAssumptions2.1,2.3and2.5befulfilled.Then,Eq.(1) has a unique (up to modifications) mild solution in D(( A)̺). − Proof. If Assumptions 2.1, 2.3 and 2.5 are fulfilled for ̺ = 0, Eq. (1) fits into the setting of [35] (the Hilbert space U is denoted by H there). By Proposition 3.1 therein thereexists a uniquemildsolution uto Eq.(1). Now suppose that Assumptions 2.1, 2.3 and 2.5 hold for some ̺ > 0. Set Uˆ := D(( A)̺), D(Aˆ):= D(( A)̺+1) − − and consider the unbounded operator Aˆ on Uˆ given by Aˆ:D(Aˆ) Uˆ Uˆ, v Aˆv := Av. ⊂ → 7→ Note that Aˆ fulfills Assumption 2.1 with A, D(A) and U replaced by Aˆ, D(Aˆ) and Uˆ, respectively. Defining the spaces D(( Aˆ)s) analogously to the − spacesD(( A)s),wehaveD(( A)̺+s) = D(( Aˆ)s),s R,sothatAssump- − − − ∈ tions 2.3 and 2.5 can be reformulated with ̺, D(( A)̺), D(( A)̺−σ) and − − D(( A)̺−β) replaced by ̺ˆ:= 0, D(( Aˆ)̺ˆ), D(( Aˆ)̺ˆ−σ) and D(( Aˆ)̺ˆ−β), − − − − respectively. Thus, the equation du(t) = Aˆu(t)+f(u(t)) dt+B(u(t))dW(t), u(0) = u , (15) 0 fitsintotheset(cid:0)tingof[35](nowU(cid:1)ˆ correspondstothespaceH there),sothat, by [35, Proposition 3.1], there exists a unique mild solution u to Eq. (15). Since the operators etA (U) and etAˆ (Uˆ) coincide on Uˆ U, it is ∈ L ∈ L ⊂ clear that any mild solution to Eq. (15) is a mild solution to Eq. (1) and vice versa. 9 Remark 2.9. If the initial condition u belongs to L (Ω,F ,P;D(( A)̺)) 0 p 0 − L (Ω,F ,P;D(( A)̺)) for some p > 2, the solution u even satisfies 2 0 s⊂up E u(t) p − < . This is a consequence of the Burkholder- t∈[0,T] k kD((−A)̺) ∞ Davis-Gundy inequality, cf. [14, Theorem 7.4] or [35, Proposition 3.1]. Anal- ogous improvements are valid for the estimates in Propositions 3.2 and 4.3 below. We finish this section with concrete examples for stochastic PDEs that fit into our setting. Example 2.10. Let be an open and bounded subset of Rd, U := L ( ), 2 O O andletA = ∆D betheDirichlet-Laplacianon asdescribedinExample2.2. O O We consider examples for stochastic PDEs in dimension d= 1 and d 2. ≥ First, let R1 be one-dimensional and consider the problem O ⊂ du(t,x) = ∆ u(t,x)dt+g(u(t,x))dt+h(u(t,x)) dW (t,x), x 1 (t,x) [0,T] , ∈ ×O (16) u(t,x) = 0, (t,x) [0,T] ∂ , ∈ × O u(0,x) = u (x), x , 0 ∈ O where u L ( ), g : R R and h : R R are globally Lipschitz 0 2 ∈ O → → continuous, and W = (W (t)) is a Wiener process (with respect to a 1 1 t∈[0,T] normal filtration on a complete probability space) whose Cameron–Martin space is some space of functions on that is continuously embedded in O L ( ), e.g., W is a Wiener process with Cameron–Martin space Hs( ) ∞ 1 O O for some s > 1/2. Let (ψk)k∈N be an arbitrary orthonormal basis of the Cameron–Martin space of W and define f and B as the Nemytskii type 1 operators f(v)(x) := g(v(x)), v L ( ), x , 2 ∈ O ∈ O B(v)a (x) := h(v(x)) akψk(x), v L2( ), a = (ak)k∈N ℓ2, x . ∈ O ∈ ∈ O k∈N (cid:0) (cid:1) X (17) Then, Eq. (1) is an abstract version of problem (16), and the mappings f and B are globally Lipschitz continuous (and thus linearly growing) from D(( A)0) = L ( ) to L ( ) and from D(( A)0) to (ℓ ;L ( )), respec- 2 2 2 2 − O O − L O tively. For B this follows from the estimate B(v )a B(v )a = h(v ) h(v ) a ψ k 1 − 2 kL2(O) (cid:13)(cid:13)(cid:0) 1 − 2 (cid:1)kX∈N k k(cid:13)(cid:13)L2(O) (cid:13)h(v ) h(v ) (cid:13)a ψ ≤ k 1 − 2 kL2(O)(cid:13)kX∈N k k(cid:13)L∞(O) (cid:13) (cid:13) ≤ Ckv1−v2kL2(O)kakℓ2(cid:13), (cid:13) where the last step is dueto the Lipschitz property of h and the assumption that the Cameron–Martin space of W is continuously embedded in L ( ). 1 ∞ O 10