This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Convergence rates of the splitting scheme for parabolic linear stochastic Cauchy problems

Sonja Cox     Jan van Neerven S.G.Cox@tudelft.nl, J.M.A.M.vanNeerven@tudelft.nl, Delft Institute of Applied Mathematics; Delft University of Technology; P.O. Box 5031; 2600 GA Delft; The Netherlands. The second named author gratefully acknowledges support by VICI subsidy 639.033.604 of the Netherlands Organisation for Scientific Research (NWO).
Abstract

We study the splitting scheme associated with the linear stochastic Cauchy problem

dU(t)\displaystyle dU(t) =AU(t)dt+dW(t),t[0,T],\displaystyle=AU(t)\,dt+dW(t),\quad t\in[0,T],
U(0)\displaystyle U(0) =x,\displaystyle=x,

Here AA is the generator of an analytic C0C_{0}-semigroup S={S(t)}t0S=\{S(t)\}_{t\ge 0} on a Banach space EE and W={W(t)}t0W=\{W(t)\}_{t\ge 0} is a Brownian motion with values in a fractional domain space EβE_{\beta} associated with AA. We prove that if α,β,γ,θ0\alpha,\beta,\gamma,\theta\ge 0 are such that γ+θ<1\gamma+\theta<1 and (αβ+θ)++γ<12(\alpha-\beta+\theta)^{+}+\gamma<\frac{1}{2}, then the approximate solutions U(n)U^{(n)} converge to the solution UU in the space Cγ([0,T];Eα)C^{\gamma}([0,T];E_{\alpha}), both in LpL^{p}-means and almost surely, with rate 1/nθ1/n^{\theta}.

keywords:
Splitting scheme, stochastic evolution equations, analytic semigroups, γ\gamma-radonifying operators, γ\gamma-boundedness, stochastic convolutions, Lie-Trotter product formula.
AMS:
Primary: 35R15, 60H15; Secondary: 47D06, 60J35

1 Introduction

We are concerned with the convergence of the splitting scheme for the stochastic linear Cauchy problem

{dU(t)=AU(t)dt+dW(t),t[0,T],U(0)=x,\left\{\begin{aligned} dU(t)&=AU(t)\,dt+dW(t),\quad t\in[0,T],\\ U(0)&=x,\end{aligned}\right. (SCPx)

were AA is the generator of a C0C_{0}-semigroup S={S(t)}t0S=\{S(t)\}_{t\ge 0} on a real Banach space EE, W={W(t)}t0W=\{W(t)\}_{t\ge 0} is an EE-valued Brownian motion on a probability space (Ω,)(\Omega,{\mathbb{P}}), and xEx\in E is an initial value which is kept fixed throughout the paper. The concept of the scheme is to alternately add an increment of the Brownian motion WW and run the semigroup SS on a time interval of equal length. Taking time steps Δt(n):=T/n\Delta t^{(n)}:=T/n and writing tj(n):=jT/nt_{j}^{(n)}:=jT/n and ΔWj(n):=W(tj(n))W(tj1(n))\Delta W_{j}^{(n)}:=W(t_{j}^{(n)})-W(t_{j-1}^{(n)}), this generates a finite sequence {Ux(n)(tj(n))}j=0n\{U_{x}^{(n)}(t_{j}^{(n)})\}_{j=0}^{n} defined by

Ux(n)(t0(n))\displaystyle U_{x}^{(n)}(t_{0}^{(n)}) :=x,\displaystyle=x,
Ux(n)(tj(n))\displaystyle U_{x}^{(n)}(t_{j}^{(n)}) :=S(Δt(n))(Ux(n)(tj1(n))+ΔWj(n)),j=1,,n.\displaystyle=S(\Delta t^{(n)})(U_{x}^{(n)}(t_{j-1}^{(n)})+\Delta W_{j}^{(n)}),\quad j=1,\dots,n.

We have the explicit formula

Ux(n)(tj(n))=S(tj(n))x+i=1jS(tji+1(n))ΔWi(n),j=0,,n.U_{x}^{(n)}(t_{j}^{(n)})=S(t_{j}^{(n)})x+\sum_{i=1}^{j}S(t_{j-i+1}^{(n)})\Delta W_{i}^{(n)},\quad j=0,\dots,n.

Assuming the existence of a unique solution UxU_{x} of the problem (SCPx) (see Proposition 7 below), we may ask for conditions ensuring the convergence of Ux(n)(T)U_{x}^{(n)}(T) to Ux(T)U_{x}(T) in Lp(Ω;E)L^{p}(\Omega;E) for some (all) 1p<1\le p<\infty or even in EE almost surely. In order to describe our approach we start by noting that each Ux(n)(tj(n))U_{x}^{(n)}(t_{j}^{(n)}) can be represented as a stochastic integral of the discretised function

S(n)(t):=j=0n1Ij(n)S(tj(n)),t[0,T],S^{(n)}(t):=\sum_{j=0}^{n}1_{I_{j}^{(n)}}\otimes S(t_{j}^{(n)}),\quad t\in[0,T],

where I0(n)={0}I_{0}^{(n)}=\{0\} and Ij(n)=(tj1(n),tj(n)]I_{j}^{(n)}=(t_{j-1}^{(n)},t_{j}^{(n)}] for j=1,,nj=1,\dots,n. Indeed, defining the stochastic integral of a step function in the obvious way, we have

Ux(n)(tj(n))=S(n)(tj(n))x+0tj(n)S(n)(tj(n)s)𝑑W(s),j=0,,n.U_{x}^{(n)}(t_{j}^{(n)})=S^{(n)}(t_{j}^{(n)})x+\int_{0}^{t_{j}^{(n)}}S^{(n)}(t_{j}^{(n)}-s)\,dW(s),\quad j=0,\dots,n. (1)

On the other hand, the exact solution of (SCPx), if it exists, is given by the stochastic convolution integral

Ux(t):=S(t)x+0tS(ts)𝑑W(s),t[0,T].U_{x}(t):=S(t)x+\int_{0}^{t}S(t-s)\,dW(s),\quad t\in[0,T]. (2)

For the precise definition of the stochastic integral we refer to Section 3. Comparing (1) and (2) we see that the problem of convergence of the splitting scheme is really a problem of convergence of ‘Riemann sums’ for stochastic integrals. Let us henceforth put

Ux(n)(t):=S(n)(t)x+0tS(n)(ts)𝑑W(s),t[0,T].U_{x}^{(n)}(t):=S^{(n)}(t)x+\int_{0}^{t}S^{(n)}(t-s)\,dW(s),\quad t\in[0,T].

The second formula interpolates the data in the identity (1) in a way that makes them easily accessible with continuous time techniques; other possible interpolations, such as piecewise linear interpolation, do not have this advantage. Needless to say, in Theorems 1 and 2 below we are primarily interested in what happens at the time points t=tj(n)t=t_{j}^{(n)}. From S(n)(tj(n))x=S(tj(n))xS^{(n)}(t_{j}^{(n)})x=S(t_{j}^{(n)})x we see that

Ux(n)(tj(n))Ux(tj(n))=U0(n)(tj(n))U0(tj(n))U_{x}^{(n)}(t_{j}^{(n)})-U_{x}(t_{j}^{(n)})=U_{0}^{(n)}(t_{j}^{(n)})-U_{0}(t_{j}^{(n)})

for all xEx\in E, and therefore it suffices to analyse convergence of the splitting scheme with initial value 0. In what follows, in order to simplify notations we shall write U(t):=U0(t)U(t):=U_{0}(t) and U(n)(t):=U0(n)(t)U^{(n)}(t):=U_{0}^{(n)}(t).

Our first result extends and simplifies previous work by Kühnemund and the second-named author [21, Theorems 4.3 and 5.2].

Theorem 1.

Each of the conditions (a) and (b) below guarantees that the problem

{dU(t)=AU(t)dt+dW(t),t[0,T],U(0)=0,\left\{\begin{aligned} dU(t)&=AU(t)\,dt+dW(t),\quad t\in[0,T],\\ U(0)&=0,\end{aligned}\right. (SCP0)

admits a unique solution U={U(t)}t[0,T]U=\{U(t)\}_{t\in[0,T]} which satisfies

limn(supt[0,T]𝔼U(n)(t)U(t)p)=0\lim_{n\to\infty}\Big{(}\sup_{t\in[0,T]}{\mathbb{E}}\|U^{(n)}(t)-U(t)\|^{p}\Big{)}=0

for all 1p<1\le p<\infty:

  1. (a)

    EE has type 22;

  2. (b)

    SS restricts to a C0C_{0}-semigroup on the reproducing kernel Hilbert space associated with WW.

The class of spaces satisfying condition (a) includes all Hilbert spaces and the spaces Lp(μ)L^{p}(\mu) for 2p<2\le p<\infty. It follows from the results in [26] that condition (b) is satisfied if the transition semigroup associated with the solution process is analytic.

The main result of this article, Theorem 2 below, concerns actual convergence rates for the splitting scheme in the case that the semigroup SS is analytic on EE. The convergence is considered in suitable Hölder norms in space and time, with explicit bounds for the convergence rate.

We denote by EαE_{\alpha} the fractional power space of exponent α0\alpha\ge 0 associated with AA (see Section 4 for more details).

Theorem 2.

Suppose that the semigroup SS is analytic on EE and that WW is a Brownian motion in EβE_{\beta} for some β0\beta\ge 0. Then the problem (SCP0) admits a unique solution U={U(t)}t[0,T]U=\{U(t)\}_{t\in[0,T]}, and for all α,γ,θ0\alpha,\gamma,\theta\ge 0 such that γ+θ<1\gamma+\theta<1 and (αβ+θ)++γ<12(\alpha-\beta+\theta)^{+}+\gamma<\frac{1}{2} one has the estimate

(𝔼U(n)UCγ([0,T];Eα)p)1p1nθ,1p<,\displaystyle\big{(}{\mathbb{E}}\big{\|}U^{(n)}-U\big{\|}^{p}_{C^{\gamma}([0,T];E_{\alpha})}\big{)}^{\frac{1}{p}}\lesssim\frac{1}{n^{\theta}},\quad 1\le p<\infty,

with implied constant independent of n1n\ge 1.

By a Borel-Cantelli argument, this result implies the almost sure convergence of U(n)U^{(n)} to UU in Cγ([0,T];Eα)C^{\gamma}([0,T];E_{\alpha}) with the same rates.

The proof of Theorem 2 heavily relies on the theory of γ\gamma-radonifying operators and γ\gamma-boundedness techniques. Standard techniques from stochastic analysis which are commonly used in connection with the problems considered here, such as Itô’s formula and the Burkholder-Davis-Gundy inequalities, are unavailable in the present general framework (unless one makes additional assumptions on EE, such as martingale type 22 or the UMD property). We also cannot use factorisation techniques (as introduced by Da Prato, Kwapień and Zabczyk [7]), the reason being that the semigroup property on which this technique relies fails for the discretised semigroup S(n)S^{(n)}.

Example 3.

Theorem 2 may be applied to second order elliptic operators of the form

Af(x)=i,j=1daij(x)ijf(x)+i=1dbi(x)if(x)+c(x)f(x).Af(x)=\sum_{i,j=1}^{d}a_{ij}(x)\partial_{ij}f(x)+\sum_{i=1}^{d}b_{i}(x)\partial_{i}f(x)+c(x)f(x).

Under minor regularity assumptions on the coefficients aij=ajia_{ij}=a_{ji}, bib_{i} and cc, such operators generate analytic semigroups on E=Lq(d)E=L^{q}({\mathbb{R}}^{d}) with 1<p<1<p<\infty (see [25, Chapter 3]) and one has Eα=H2α,q(d)E_{\alpha}=H^{2\alpha,q}({\mathbb{R}}^{d}) for all 0<α<120<\alpha<\frac{1}{2}. Applying Theorem 2 (with β=0\beta=0), we obtain convergence of the splitting scheme in the space Cγ([0,T];H2α,q(d))C^{\gamma}([0,T];H^{2\alpha,q}({\mathbb{R}}^{d})) for any γ0\gamma\ge 0 such that 0<α+γ<120<\alpha+\gamma<\frac{1}{2}. By the Sobolev embedding H2α,q(d)C02αd/q(d)\smash{H^{2\alpha,q}({\mathbb{R}}^{d})\hookrightarrow C_{0}^{2\alpha-{d}/q}({\mathbb{R}}^{d})} [39, Section 2.8], this implies the convergence of the splitting scheme in the mixed Hölder space Cγ([0,T];C02αd/q(d))C^{\gamma}([0,T];C_{0}^{2\alpha-{d}/q}({\mathbb{R}}^{d})). As a consequence, we obtain convergence in the mixed Hölder space

Cγ([0,T];C02δ(d)),γ,δ0,γ+δ<12,C^{\gamma}([0,T];C_{0}^{2\delta}({\mathbb{R}}^{d})),\quad\gamma,\delta\ge 0,\ \ \gamma+\delta<\tfrac{1}{2},

with rate 1/nθ1/n^{\theta} for any θ<12γδ\theta<\tfrac{1}{2}-\gamma-\delta; this rate improves when the noise is more regular. Similar results can be obtained for elliptic operators on smooth domains DdD\subseteq{\mathbb{R}}^{d} subject to various types of boundary conditions (as long as they generate an analytic semigroup on Lq(D)L^{q}(D)).

For semi-linear (Stratonovich type) SPDEs governed by second order elliptic operators on d{\mathbb{R}}^{d} and driven by multiplicative noise, convergence in E=L2(d)E=L^{2}({\mathbb{R}}^{d}) of splitting schemes like the one considered here has been proved by various authors [2, 3, 12, 14, 27]. Using techniques from PDE and stochastic analysis it is shown by Gyöngy and Krylov [14] that, with respect to the norm of E=L2(d)E=L^{2}({\mathbb{R}}^{d}), for finite-dimensional noise and with sufficiently smooth coefficients one obtains the maximal estimate

(𝔼supt[0,T]U(n)UL2(d)p)1p1n,1p<.\big{(}{\mathbb{E}}\sup_{t\in[0,T]}\|U^{(n)}-U\|_{L^{2}({\mathbb{R}}^{d})}^{p}\big{)}^{\frac{1}{p}}\lesssim\frac{1}{n},\quad 1\le p<\infty.

Our result is valid in the full scale of spaces Lq(d)L^{q}({\mathbb{R}}^{d}) and infinite-dimensional noise, with a rate which (for smooth enough noise) is only slightly worse that 1/n1/n and is independent of 1q<1\le q<\infty. More precisely, for β12+α\beta\ge\frac{1}{2}+\alpha and taking γ=0\gamma=0 we obtain uniform convergence with rate 1/nθ1/n^{\theta} for any 0θ<10\le\theta<1. In addition to that we obtain Hölder regularity in both space and time. On the other hand, as we already mentioned, Gyöngy and Krylov [14] consider the semi-linear case and multiplicative noise.

The next example shows that by working in suitable fractional extrapolation spaces (this technique is explained in [10]; see also [4, 5]), the assumption that WW is a Brownian motion can be weakened to WW being a cylindrical Brownian motion (see, e.g., [30, 32] for the definition).

Example 4.

The stochastic heat equation on the unit interval [0,1][0,1] with Dirichlet boundary conditions driven by space-time white noise can be put into the present framework by taking for EE the extrapolation space FρF_{\rho} with F=Lq(0,1)F=L^{q}(0,1) and ρ<14\rho<-\frac{1}{4}. As we shall explain in Example 22, this entails the convergence of the splitting scheme in the mixed Hölder space

Cγ([0,T];C02δ[0,1]),γ,δ0,γ+δ<14,C^{\gamma}([0,T];C_{0}^{2\delta}[0,1]),\quad\gamma,\delta\ge 0,\ \ \gamma+\delta<\tfrac{1}{4},

with rate 1/nθ1/n^{\theta} for any θ<14γδ\theta<\frac{1}{4}-\gamma-\delta.

It is shown in [8] that any approximation scheme for a one-dimensional stochastic heat equation with additive space-time white noise which incorporates the contributions of the noise only by means of the terms ΔWk\Delta W_{k}, k=1,,nk=1,\ldots,n, cannot have a convergence rate better than 1/n141/n^{\frac{1}{4}}. This shows that the exponent 14\frac{1}{4} in Example 4 is the best possible.

The field of numerical approximation of stochastic partial differential equations (SPDEs) is a very active one; an up-to-date overview of the available results can be found in [19]. In [13] convergence rates are considered for various approximations schemes in space and time of a quasi-linear parabolic SPDE driven by white noise. The authors obtain a convergence rate 1/n141/n^{\frac{1}{4}} in LpL^{p} for an implicit Euler scheme. In [36] convergence in probability is proved (without rates) for the same SPDE with state-dependent dispersion. Rates for path-wise convergence are given for quasi-linear parabolic SPDEs in [15, 18, 24], albeit only for colored noise. It seems likely that the methods of this paper can be extended to the implicit Euler scheme and to semi-linear problems with multiplicative noise; we plan to address such extensions in a future paper.

The paper is organised as follows. Section 2 presents some preliminary material about spaces of γ\gamma-radonifying operators. The proofs of Theorems 1 and 2 are presented in Sections 3 (Theorems 9, 10) and 4 (Theorem 19), respectively.

It is known that each of the conditions in Theorems 1 and 2 implies that the solution process UU has continuous trajectories. In the final Section 5 we present an example which shows that without any additional assumptions on the space EE and/or the semigroup SS the splitting scheme may fail to converge even if a solution UU with continuous trajectories exists.

2 Preliminaries

Let {γj}j1\{\gamma_{j}\}_{j\ge 1} be a sequence of independent standard Gaussian random variables on a probability space (Ω,)(\Omega,{\mathbb{P}}), let \mathscr{H} be a real Hilbert space (later we shall take =L2(0,T;H)\mathscr{H}=L^{2}(0,T;H), where HH is another real Hilbert space) and EE a real Banach space. A bounded operator RR from \mathscr{H} to EE is called γ\gamma-summing if

Rγ(,E)2:=suph𝔼j=1kγjRhj2,\|R\|_{\gamma_{\infty}(\mathscr{H},E)}^{2}:=\sup_{h}{\mathbb{E}}\Big{\|}\sum_{j=1}^{k}\gamma_{j}Rh_{j}\Big{\|}^{2},

is finite, where the supremum is taken over all finite orthonormal systems h={hj}j=1kh=\{h_{j}\}_{j=1}^{k} in \mathscr{H}. It can be shown that γ(,E)\|\cdot\|_{\gamma_{\infty}(\mathscr{H},E)} is indeed a norm which turns the space of γ\gamma-summing operators into a Banach space.

The space γ(,E)\gamma(\mathscr{H},E) of γ\gamma-radonifying operators is defined to be the closure of the finite rank operators under the norm γ\|\cdot\|_{\gamma_{\infty}}; it is a closed subspace of γ(,E)\gamma_{\infty}(\mathscr{H},E). A celebrated result of Kwapień and Hoffmann-Jørgensen [17, 23] implies that if EE does not contain a closed subspace isomorphic to c0c_{0} then γ(,E)=γ(,E)\gamma(\mathscr{H},E)=\gamma_{\infty}(\mathscr{H},E).

Since convergence in γ(,E)\gamma(\mathscr{H},E) implies convergence in (,E)\mathscr{L}(\mathscr{H},E), every operator Rγ(,E)R\in\gamma(\mathscr{H},E), being the operator norm limit of a sequence of finite rank operators from \mathscr{H} to EE, is compact.

If \mathscr{H} is separable with orthonormal basis {hj}j1\{h_{j}\}_{j\ge 1}, then an operator R:ER:\mathscr{H}\to E is γ\gamma-radonifying if and only if the Gaussian sum j1γjRhj\sum_{j\ge 1}\gamma_{j}Rh_{j} converges in L2(Ω;E)L^{2}(\Omega;E), and in this situation we have

Rγ(,E)2=𝔼j1γjRhj2.\|R\|_{\gamma(\mathscr{H},E)}^{2}={\mathbb{E}}\Big{\|}\sum_{j\ge 1}\gamma_{j}Rh_{j}\Big{\|}^{2}.

The general case may be reduced to the separable case by observing that for any Rγ(,E)R\in\gamma(\mathscr{H},E) there exists a separable closed subspace R\mathscr{H}_{R} of \mathscr{H} such that RR vanishes on the orthogonal complement R\mathscr{H}_{R}^{\perp}.

If Rγ(H,E)R\in\gamma(H,E) is given and {hj}j1\{h_{j}\}_{j\ge 1} is an orthonormal basis for R\mathscr{H}_{R}, the sum j1γjRhj\sum_{j\ge 1}\gamma_{j}Rh_{j} defines a centred EE-valued Gaussian random variable. Its distribution μ\mu is a centred Gaussian Radon measure on EE whose covariance operator equals RRRR^{\ast}. We will refer to μ\mu as the Gaussian measure associated with RR. In the reverse direction, if YY is a centred EE-valued Gaussian random variable with reproducing kernel Hilbert space \mathscr{H}, then \mathscr{H} is separable, the natural inclusion mapping i:Ei:\mathscr{H}\hookrightarrow E is γ\gamma-radonifying, and we have

iγ(,E)2=𝔼Y2.\|i\|_{\gamma(\mathscr{H},E)}^{2}={\mathbb{E}}\|Y\|^{2}.

Below we shall need the following simple continuity result.

Proposition 5.

Let (X,d)(X,d) be a metric space and let V:X(E,F)V:X\to\mathscr{L}(E,F) be strongly continuous. Then for all Rγ(,E)R\in\gamma(\mathscr{H},E) the function VR:Xγ(,F)VR:X\to\gamma(\mathscr{H},F),

(VR)(ξ):=V(ξ)R,ξX,(VR)(\xi):=V(\xi)R,\qquad\xi\in X,

is continuous.

Proof.

Suppose first that RR is a finite rank operator, say R=j=1khjxjR=\sum_{j=1}^{k}h_{j}\otimes x_{j} with {hj}j=1k\{h_{j}\}_{j=1}^{k}\in\mathscr{H} orthonormal and {xj}j=1k\{x_{j}\}_{j=1}^{k} a sequence in EE. Suppose that limnξn=ξ\lim_{n\to\infty}\xi_{n}=\xi in XX. Then

limnV(ξn)RV(ξ)Rγ(,F)2=limn𝔼j=1kγj(V(ξn)V(ξ))xj2=0.\lim_{n\to\infty}\|V(\xi_{n})R-V(\xi)R\|_{\gamma(\mathscr{H},F)}^{2}=\lim_{n\to\infty}{\mathbb{E}}\Big{\|}\sum_{j=1}^{k}\gamma_{j}(V(\xi_{n})-V(\xi))x_{j}\Big{\|}^{2}=0.

The general case follows from the density of the finite rank operators in γ(,E)\gamma(\mathscr{H},E) and the norm estimate V(ξ)Rγ(,F)V(ξ)Rγ(,E)\|V(\xi)R\|_{\gamma(\mathscr{H},F)}\le\|V(\xi)\|\|R\|_{\gamma(\mathscr{H},E)}. ∎

3 Proof of Theorem 1

We start with a brief discussion of stochastic integrals of operator-valued functions. Let HH be a Hilbert space and fix T>0T>0. An HH-cylindrical Brownian motion, indexed by [0,T][0,T] and defined on a probability space (Ω,,)(\Omega,\mathscr{F},{\mathbb{P}}), is a mapping WH:L2(0,T;H)L2(Ω)W_{H}:L^{2}(0,T;H)\rightarrow L^{2}(\Omega) with the following properties:

  • for all hL2(0,T;H)h\in L^{2}(0,T;H) the random variable WH(h)W_{H}(h) is Gaussian;

  • for all h1,h2L2(0,T;H)h_{1},h_{2}\in L^{2}(0,T;H) we have 𝔼WH(h1)WH(h2)=h1,h2{\mathbb{E}}W_{H}(h_{1})W_{H}(h_{2})=\langle h_{1},h_{2}\rangle.

Formally, an HH-cylindrical Brownian motion can be thought of as a standard Brownian motion taking values in the Hilbert space HH. One easily checks that WHW_{H} is linear and that for all h1,,hnL2(0,T;H)h_{1},\ldots,h_{n}\in L^{2}(0,T;H) the random variables WH(h1),,WH(hn)W_{H}(h_{1}),\ldots,W_{H}(h_{n}) are jointly Gaussian. These random variables are independent if and only if h1,,hnh_{1},\ldots,h_{n} are orthogonal in HH. For further details see [28, Section 3].

A finite rank step function is function of the form n=1N1(an,bn]Bn\sum_{n=1}^{N}1_{(a_{n},b_{n}]}\otimes B_{n} where each operator Bn:HEB_{n}:H\to E is of finite rank. The stochastic integral with respect to WHW_{H} of such a function is defined by setting

0T1(a,b](hx)𝑑WH:=WH(1(a,b]h)x\int_{0}^{T}1_{(a,b]}\otimes(h\otimes x)\,dW_{H}:=W_{H}(1_{(a,b]}\otimes h)\otimes x

and this definition is extended by linearity. A function Ψ:(0,T)(H,E)\Psi:(0,T)\to\mathscr{L}(H,E) is said to be stochastically integrable with respect to WHW_{H} if there exists a sequence of finite rank step functions Ψn:(0,T)(H,E)\Psi_{n}:(0,T)\to\mathscr{L}(H,E) such that:

  1. (i)

    for all hHh\in H we have limnΨnh=Ψh\lim_{n\to\infty}\Psi_{n}h=\Psi h in measure on (0,T)(0,T);

  2. (ii)

    the limit Y:=limn0TΨn𝑑WHY:=\lim_{n\to\infty}\int_{0}^{T}\Psi_{n}\,dW_{H} exists in probability.

In this situation we write

Y=0TΨ𝑑WHY=\int_{0}^{T}\Psi\,dW_{H}

and call YY the stochastic integral of Ψ\Psi with respect to WHW_{H}.

As was shown in [32], for finite rank step functions Ψ\Psi one has the isometry

𝔼0TΨ𝑑WH2=RΨγ(L2(0,T;H),E)2,{\mathbb{E}}\Big{\|}\int_{0}^{T}\Psi\,dW_{H}\Big{\|}^{2}=\|R_{\Psi}\|_{\gamma(L^{2}(0,T;H),E)}^{2}, (3)

where RΨ:L2(0,T;H)ER_{\Psi}:L^{2}(0,T;H)\to E is the bounded operator represented by Ψ\Psi, i.e.,

RΨf=0TΨ(t)f(t)𝑑t,fL2(0,T;H).R_{\Psi}f=\int_{0}^{T}\Psi(t)f(t)\,dt,\quad f\in L^{2}(0,T;H). (4)

As a consequence, a function Ψ:(0,T)(H,E)\Psi:(0,T)\to\mathscr{L}(H,E) is stochastically integrable on (0,T)(0,T) with respect to WHW_{H} if and only if ΨxL2(0,T;H)\Psi^{\ast}x^{\ast}\in L^{2}(0,T;H) for all xEx^{\ast}\in E^{\ast} and there exists an operator RΨγ(L2(0,T;H),E)R_{\Psi}\in\gamma(L^{2}(0,T;H),E) such that

RΨf,x=0T[f(t),Ψ(t)x]𝑑t,xE.\langle R_{\Psi}f,x^{\ast}\rangle=\int_{0}^{T}[f(t),\Psi^{\ast}(t)x^{\ast}]\,dt,\quad x^{\ast}\in E^{\ast}.

The isometry (3) extends to this situation. The following simple observation [10, Lemma 2.1] will be used frequently:

Proposition 6.

For all gL2(0,T)g\in L^{2}(0,T) and Rγ(H,E)R\in\gamma(H,E) the function gR:tg(t)RgR:t\mapsto g(t)R belongs to γ(L2(0,T;H),E)\gamma(L^{2}(0,T;H),E) and we have

gRγ(L2(0,T;H),E)=gL2(0,T)Rγ(H,E).\|gR\|_{\gamma(L^{2}(0,T;H),E)}=\|g\|_{L^{2}(0,T)}\|R\|_{\gamma(H,E)}.

For the remainder of this section we fix an EE-valued Brownian motion W={W(t)}t0W=\{W(t)\}_{t\ge 0} and T>0T>0. Let HH be the reproducing kernel Hilbert space associated with the Gaussian random variable W(1)W(1) and let i:HEi:H\hookrightarrow E be the natural inclusion mapping. Then WW induces an HH-cylindrical Brownian motion WHW_{H} by putting

WH(fix):=0TfdW,x,fL2(0,T),xE.\displaystyle W_{H}(f\otimes i^{\ast}x^{\ast})=\int_{0}^{T}f\,d\langle W,x^{*}\rangle,\quad f\in L^{2}(0,T),\ x^{*}\in E^{*}. (5)

This motivates us to call a function Ψ:(0,T)(E)\Psi:(0,T)\to\mathscr{L}(E) stochastically integrable with respect to WW if the function Ψi:(0,T)(H,E)\Psi\circ i:(0,T)\to\mathscr{L}(H,E) is stochastically integrable with respect to WHW_{H}, in which case we put

0TΨ𝑑W:=0T(Ψi)𝑑WH.\int_{0}^{T}\Psi\,dW:=\int_{0}^{T}(\Psi\circ i)\,dW_{H}.

It is easy to check that for all S(E)S\in\mathscr{L}(E) the indicator function 1(a,b]S1_{(a,b]}\otimes S is stochastically integrable with respect to WW and

0T1(a,b]S𝑑W=S(W(b)W(a)).\int_{0}^{T}1_{(a,b]}\otimes S\,dW=S(W(b)-W(a)).

This shows that the definition is consistent with (1) and (2).

Now let S={S(t)}t0S=\{S(t)\}_{t\ge 0} denote a C0C_{0}-semigroup of bounded linear operators on EE, with generator AA. We will be interested in the case where the function to be integrated against WHW_{H} is one of the following:

Φ(t):=S(t)i,Φ(n)(t):=j=1n1Ij(n)(t)[S(tj(n))i],t(0,T).\Phi(t):=S(t)\circ i,\quad\Phi^{(n)}(t):=\sum_{j=1}^{n}1_{I_{j}^{(n)}}(t)\otimes[S(t_{j}^{(n)})\circ i],\quad t\in(0,T).

We may define bounded operators RΦ(n)R_{\Phi^{(n)}} and RΦR_{\Phi} from L2(0,T;H)L^{2}(0,T;H) to EE by the formula (4). Being associated with γ(H,E)\gamma(H,E)-valued step functions, the operators RΦ(n)R_{\Phi^{(n)}} belong to γ(L2(0,T;H),E)\gamma(L^{2}(0,T;H),E) by Proposition 6. Concerning the question whether the operator RΦR_{\Phi} is in γ(L2(0,T;H),E)\gamma(L^{2}(0,T;H),E) we have the following result [32, Theorem 7.1].

Proposition 7.

Let Φ(t)=S(t)i\Phi(t)=S(t)\circ i. The following assertions are equivalent:

  1. (i)

    the operator RΦR_{\Phi} belongs to γ(L2(0,T;H),E)\gamma(L^{2}(0,T;H),E);

  2. (ii)

    the function Φ\Phi is stochastically integrable on (0,T)(0,T) with respect to WHW_{H};

  3. (iii)

    for some (all) xEx\in E the problem (SCPx) admits a unique solution UxU_{x}.

In this situation, for all xEx\in E and t[0,T]t\in[0,T] we have

Ux(t)=S(t)x+0tS(ts)𝑑W(s)=S(t)x+0tΦ(ts)𝑑WH(s)U_{x}(t)=S(t)x+\int_{0}^{t}S(t-s)\,dW(s)=S(t)x+\int_{0}^{t}\Phi(t-s)\,dW_{H}(s)

almost surely.

In [32] an example is presented showing even for rank one Brownian motions WW in EE the equivalent conditions need not always be satisfied for all C0C_{0}-semigroups SS on EE. The conditions are satisfied, however, if one of the following additional conditions holds:

  1. (a)

    EE is a type 2 Banach space,

  2. (b)

    SS restricts to a C0C_{0}-semigroup on HH,

  3. (c)

    SS is an analytic C0C_{0}-semigroup on EE.

We refer to [10, 32] for the easy proofs.

We are now in a position to state the main result of this section. We use the notations introduced above, and let μ\mu and μ(n)\mu^{(n)} denote the Gaussian measures on EE associated with the operators RΦR_{\Phi} and RΦ(n)R_{\Phi^{(n)}}, respectively.

Theorem 8.

Suppose that the equivalent conditions of Proposition 7 are satisfied. The following assertions are equivalent:

  1. 1.

    limnUx(n)(T)=Ux(T)\lim_{n\to\infty}U_{x}^{(n)}(T)=U_{x}(T) in Lp(Ω;E)L^{p}(\Omega;E) for some (all) xEx\in E and some (all) 1p<1\le p<\infty;

  2. 2.

    limnRΦ(n)=RΦ\lim_{n\to\infty}R_{\Phi^{(n)}}=R_{\Phi} in γ(L2(0,T;H),E)\gamma(L^{2}(0,T;H),E);

  3. 3.

    limnμ(n)=μ\lim_{n\to\infty}\mu^{(n)}=\mu weakly.

In this situation we have limnUx(n)(t)=Ux(t)\lim_{n\to\infty}U_{x}^{(n)}(t)=U_{x}(t) in Lp(Ω;E)L^{p}(\Omega;E) for all xEx\in E, t[0,T]t\in[0,T], and 1p<1\le p<\infty, and in fact we have

sup0tT𝔼Ux(n)(t)Ux(t)psup0tTS(n)(t)xS(t)x+𝔼U(n)(T)U(T)p,\sup_{0\le t\le T}{\mathbb{E}}\|U_{x}^{(n)}(t)-U_{x}(t)\|^{p}\le\sup_{0\le t\le T}\|S^{(n)}(t)x-S(t)x\|+{\mathbb{E}}\|U^{(n)}(T)-U(T)\|^{p},

where, as before, U(n)=U0(n)U^{(n)}=U_{0}^{(n)} and U=U0U=U_{0} correspond to the initial value 0.

Proof.

We begin by proving the equivalence of (1), (2), (3). Clearly it suffices to consider the initial value x=0x=0.

For a given 1p<1\le p<\infty, a sequence of EE-valued centred Gaussian random variables converges in Lp(Ω;E)L^{p}(\Omega;E) if and only if it converges in probability in EE. Therefore, if (1) holds for some 1p<1\le p<\infty, then it holds for all 1p<1\le p<\infty.

Taking p=2p=2 in (1) the equivalence (1)\Leftrightarrow(2) follows from the identity (3) and the representations (1) and (2).

Next we claim that limnRΦ(n)x=RΦx\lim_{n\to\infty}R_{\Phi^{(n)}}^{\ast}x^{\ast}=R_{\Phi}^{\ast}x^{\ast} in L2(0,T;H)L^{2}(0,T;H) for all xEx^{\ast}\in E^{\ast}. Once we have shown this, the equivalence (2)\Leftrightarrow(3) follows from [16, Theorem 3.1] (or by using the argument of [34, page 18ff]). To prove the claim we fix xEx^{\ast}\in E^{\ast} and note that in L2(0,T;H)L^{2}(0,T;H) we have

RΦx=iS()x,RΦ(n)x=j=1n1Ij(n)()iS(tj(n))x.R_{\Phi}^{\ast}x^{\ast}=i^{\ast}S^{\ast}(\cdot)x^{\ast},\quad R_{\Phi^{(n)}}^{\ast}x^{\ast}=\sum_{j=1}^{n}1_{I_{j}^{(n)}}(\cdot)\otimes i^{\ast}S^{\ast}(t_{j}^{(n)})x^{\ast}.

The inclusion mapping i:HEi:H\hookrightarrow E is γ\gamma-radonifying and hence compact. As a consequence, the weak-continuity of tS(t)xt\mapsto S^{\ast}(t)x^{\ast} implies that tiS(t)x=Φ(t)xt\mapsto i^{\ast}S^{\ast}(t)x^{\ast}=\Phi^{\ast}(t)x^{\ast} is continuous on [0,T][0,T]. It follows that limnRΦ(n)()x=RΦ()x\lim_{n\to\infty}R_{\Phi^{(n)}}^{\ast}(\cdot)x^{\ast}=R_{\Phi}^{\ast}(\cdot)x^{\ast} in L(0,T;H)L^{\infty}(0,T;H), and hence in L2(0,T;H)L^{2}(0,T;H).

The final assertion an immediate consequence of (1), (2), and covariance domination [32, Corollary 4.4]. ∎

The assertions (1), (2), (3) are equivalent to the validity of a Lie–Trotter product formula for the Ornstein-Uhlenbeck semigroup 𝒫={𝒫(t)}t0{\mathscr{P}}=\{\mathscr{P}(t)\}_{t\ge 0} associated with the problem (SCPx), which is defined on the space Cb(E)C_{\rm b}(E) of all bounded real-valued continuous functions on EE by the formula

𝒫(t)f(x)=𝔼f(Ux(t)),xE,t0,\mathscr{P}(t)f(x)={\mathbb{E}}f(U_{x}(t)),\qquad x\in E,\ t\ge 0,

where UxU_{x} is the solution of (SCPx). In order to explain the precise result, let us denote by 𝒮={𝒮(t)}t0\mathscr{S}=\{{\mathscr{S}}(t)\}_{t\ge 0} and 𝒯={𝒯(t)}t0\mathscr{T}=\{{\mathscr{T}}(t)\}_{t\ge 0} the semigroups on Cb(E)C_{\rm b}(E) corresponding to the drift term and the diffusion term in (SCPx). Thus,

𝒮(t)f(x)=f(S(t)x),𝒯(t)f(x)=𝔼f(x+W(t)),t0,xE.\begin{aligned} {\mathscr{S}}(t)f(x)&=f(S(t)x),\\ {\mathscr{T}}(t)f(x)&={\mathbb{E}}f(x+W(t)),\end{aligned}\qquad t\ge 0,\ x\in E.

Each of the semigroups 𝒫\mathscr{P}, 𝒮\mathscr{S} and 𝒯\mathscr{T} is jointly continuous in tt and xx, uniformly on [0,T]×K[0,T]\times K for all compact sets KEK\subseteq E. It was shown in [21] that if condition (3) of Theorem 8 holds, then for all fCb(E)f\in C_{\rm b}(E) we have the Lie–Trotter product formula

𝒫(t)f(x)=limn[𝒯(t/n)𝒮(t/n)]nf(x){\mathscr{P}}(t)f(x)=\lim_{n\rightarrow\infty}\big{[}{\mathscr{T}}({t}/{n}){\mathscr{S}}({t}/{n})\big{]}^{n}f(x) (6)

with convergence uniformly on [0,T]×K[0,T]\times K for all compact sets KEK\subseteq E. Conversely it follows from the proof of this result that (6) with x=0x=0 implies condition (3) of Theorem 8. In the same paper it was shown that (6) holds if at least one of the next two conditions is satisfied:

  1. (a)

    EE is isomorphic to a Hilbert space;

  2. (b)

    SS restricts to a C0C_{0}-semigroup on HH.

Thus, either of these conditions implies the convergence limnUx(n)(t)=Ux(t)\lim_{n\to\infty}U_{x}^{(n)}(t)=U_{x}(t) in Lp(Ω;E)L^{p}(\Omega;E) for all xEx\in E and t[0,T]t\in[0,T] of the splitting scheme. The proofs in [21] are rather involved. A simple proof for case (b) has been subsequently obtained by Johanna Tikanmäki (personal communication). In Theorems 9 and 10 below we shall give simple proofs for both cases (a) and (b), based on the Proposition 5 and an elementary convergence result for γ\gamma-radonifying operators from [30], respectively. Moreover, case (a) is extended to Banach spaces with type 22. Recall that a Banach space is said to have type 1p21\le p\le 2 if there exists a constant C0C\ge 0 such that for all finite choices x1,,xkEx_{1},\dots,x_{k}\in E we have

(𝔼j=1kγjxj2)12C(j=1kxjp)1p.\Big{(}{\mathbb{E}}\Big{\|}\sum_{j=1}^{k}\gamma_{j}x_{j}\Big{\|}^{2}\Big{)}^{\frac{1}{2}}\le C\Big{(}\sum_{j=1}^{k}\|x_{j}\|^{p}\Big{)}^{\frac{1}{p}}.

Hilbert spaces have type 22 and LpL^{p}-spaces (1p<1\le p<\infty) have type min{p,2}\min\{p,2\}. We refer to [1] for more information.

Theorem 9.

If EE has type 22, then the equivalent conditions of Proposition 7 and Theorem 8 hold for every C0C_{0}-semigroup SS on EE. As a consequence we have

limnsupt[0,T]𝔼U(n)(t)U(t)p=0,1p<.\lim_{n\to\infty}\sup_{t\in[0,T]}{\mathbb{E}}\|U^{(n)}(t)-U(t)\|^{p}=0,\quad 1\le p<\infty.
Proof.

By Proposition 5 we have ΦC([0,T];γ(H,E))\Phi\in C([0,T];\gamma(H,E)). This clearly implies that limnΦ(n)=Φ\lim_{n\to\infty}\Phi^{(n)}=\Phi in L(0,T;γ(H,E))L^{\infty}(0,T;\gamma(H,E)), and hence in L2(0,T;γ(H,E))L^{2}(0,T;\gamma(H,E)). Since EE has type 22, by [33, Lemma 6.1] the mapping ΨRΨ\Psi\mapsto R_{\Psi} defines a continuous inclusion L2(0,T;γ(H,E))γ(L2(0,T;H),E)L^{2}(0,T;\gamma(H,E))\hookrightarrow\gamma(L^{2}(0,T;H),E). It follows that limnRΦ(n)=RΦ\lim_{n\to\infty}R_{\Phi^{(n)}}=R_{\Phi} in γ(L2(0,T;H),E)\gamma(L^{2}(0,T;H),E). ∎

Theorem 10.

If SS restricts to a C0C_{0}-semigroup on HH, then the equivalent conditions of Proposition 7 and Theorem 8 hold. As a consequence we have

limnsupt[0,T]𝔼U(n)(t)U(t)p=0,1p<.\lim_{n\to\infty}\sup_{t\in[0,T]}{\mathbb{E}}\|U^{(n)}(t)-U(t)\|^{p}=0,\quad 1\le p<\infty.
Proof.

Let SHS_{H} denote the restricted semigroup on HH. From the identity S(t)i=iSH(t)S(t)\circ i=i\circ S_{H}(t) we have RΦ=iTR_{\Phi}=i\circ T and RΦ(n)=iT(n)R_{\Phi^{(n)}}=i\circ T^{(n)}, where TT and T(n)T^{(n)} are the bounded operators from L2(0,T;H)L^{2}(0,T;H) to HH defined by

Tf:=0TSH(t)f(t)𝑑t,T(n)f=0Tj=1n1Ij(n)(t)SH(tj(n))f(t)dt.Tf:=\int_{0}^{T}S_{H}(t)f(t)\,dt,\quad T^{(n)}f=\int_{0}^{T}\sum_{j=1}^{n}1_{I_{j}^{(n)}}(t)S_{H}(t_{j}^{(n)})f(t)\,dt.

Since limn(T(n))h=Th\lim_{n\to\infty}(T^{(n)})^{*}h=T^{*}h for all hHh\in H by the strong continuity of the adjoint semigroup SHS_{H}^{\ast} (see [37]), it follows from [30, Proposition 2.4] that limnRΦ(n)=RΦ\lim_{n\to\infty}R_{\Phi^{(n)}}=R_{\Phi} in γ(L2(0,T;H),E)\gamma(L^{2}(0,T;H),E). ∎

4 Proof of Theorem 2

In this section we shall prove convergence of the splitting scheme under the assumption that the C0C_{0}-semigroup generated by AA is analytic; no assumptions on the space EE are made. In this situation we are also able to give explicit rates of convergence in suitable interpolation spaces.

We begin with a minor extension of a result due to Kalton and Weis [20]. It enables us to check whether certain (H,E)\mathscr{L}(H,E)–valued functions define operators belonging to γ(L2(0,T;H),E)\gamma(L^{2}(0,T;H),E). We refer to [28, Section 13] for a detailed proof.

Proposition 11.

Let Φ:(a,b)γ(H,E)\Phi:(a,b)\to\gamma(H,E) be continuously differentiable with

ab(sa)12Φ(s)γ(H,E)𝑑s<.\int_{a}^{b}(s-a)^{\frac{1}{2}}\|{\Phi^{\prime}(s)}\|_{\gamma(H,E)}\,ds<\infty.

Define RΦ:L2(a,b;H)ER_{\Phi}:L^{2}(a,b;H)\to E by

RΦf:=abΦ(t)f(t)𝑑t.R_{\Phi}f:=\int_{a}^{b}\Phi(t)f(t)\,dt.

Then RΦγ(L2(a,b;H),E)R_{\Phi}\in\gamma(L^{2}(a,b;H),E) and

RΦγ(L2(a,b;H),E)(ba)12Φ(b)γ(H,E)+ab(sa)12Φ(s)γ(H,E)𝑑s.\|R_{\Phi}\|_{\gamma(L^{2}(a,b;H),E)}\le(b-a)^{\frac{1}{2}}\|{\Phi(b)}\|_{\gamma(H,E)}+\int_{a}^{b}(s-a)^{\frac{1}{2}}\|\Phi^{\prime}(s)\|_{\gamma(H,E)}\,ds.

For α0\alpha\ge 0 and large enough ww\in{\mathbb{R}} we define

Eα:=𝒟((wA)α),E_{\alpha}:=\mathscr{D}((w-A)^{\alpha}),

which is known to be independent of the choice of ww. It is a Banach space with respect to the norm xEα:=(wA)αx\|x\|_{E_{\alpha}}:=\|(w-A)^{\alpha}x\|. This norm depends of course on ww, but any two such norms are mutually equivalent. In what follows we consider ww to be fixed.

We shall also need the extrapolation spaces EαE_{-\alpha}, defined for α>0\alpha>0 as the closure of EE with respect to the norm xEα:=(wA)αx\|x\|_{E_{-\alpha}}:=\|(w-A)^{-\alpha}x\|. It follows readily from the definitions that for any two α,β\alpha,\beta\in{\mathbb{R}} the operator (wA)α(w-A)^{\alpha} defines an isomorphism from EβE_{\beta} onto EβαE_{\beta-\alpha}.

In the next two remarks we fix α,β0\alpha,\beta\ge 0 and iγ(H,Eβ)i\in\gamma(H,E_{\beta}), and suppose that SS is an analytic C0C_{0}-semigroup on EE with generator AA.

Remark 12.

By [35, Theorem 2.6.13(c)] one has, for any θ0\theta\ge 0,

S(t)(E,Eθ)tθ\displaystyle\|S(t)\|_{\mathscr{L}(E,E_{\theta})}\lesssim t^{-\theta} (7)

with implied constant independent of t[0,T]t\in[0,T]. From this and the ideal property for γ\gamma-radonifying operators we obtain the following estimate for Φ(t):=S(t)i\Phi(t):=S(t)\circ i:

Φ(t)γ(H,Eα)\displaystyle\|\Phi^{\prime}(t)\|_{\gamma(H,E_{\alpha})} AS(t)(Eβ,Eα)iγ(H,Eβ)\displaystyle\le\|AS(t)\|_{\mathscr{L}(E_{\beta},E_{\alpha})}\|i\|_{\gamma(H,E_{\beta})}
=S(t)(E,Eα+1β)iγ(H,Eβ)t(α+1β)+iγ(H,Eβ)\displaystyle=\|S(t)\|_{\mathscr{L}(E,E_{\alpha+1-\beta})}\|i\|_{\gamma(H,E_{\beta})}\lesssim t^{-(\alpha+1-\beta)^{+}}\|i\|_{\gamma(H,E_{\beta})}

where r+:=max{0,r}r^{+}:=\max\{0,r\} for rr\in{\mathbb{R}}; the implied constant is independent of t[0,T]t\in[0,T] and iγ(H,Eβ)i\in\gamma(H,E_{\beta}). If αβ<12\alpha-\beta<\frac{1}{2}, it then follows from Proposition 11 that

RΦγ(L2(0,t;H),Eα)tmin{12α+β,32}iγ(H,Eβ),\|R_{\Phi}\|_{\gamma(L^{2}(0,t;H),E_{\alpha})}\lesssim t^{\min\{\frac{1}{2}-\alpha+\beta,\frac{3}{2}\}}\|i\|_{\gamma(H,E_{\beta})},

with implied constant independent of t[0,T]t\in[0,T] and iγ(H,Eβ)i\in\gamma(H,E_{\beta}). In particular, taking α=β=0\alpha=\beta=0 we see that the equivalent conditions of Proposition 7 hold.

Remark 13.

Suppose that δ[0,12)\delta\in[0,\tfrac{1}{2}). Identifying operator-valued functions with the integral operators they induce, we have

ssδS(ts)iγ(L2(0,t;H),Eα)\displaystyle\|s\mapsto s^{-\delta}S(t-s)i\|_{\gamma(L^{2}(0,t;H),E_{\alpha})}
ssδS(ts)iγ(L2(0,t2;H),Eα)+s(ts)δS(s)iγ(L2(0,t2;H),Eα).\displaystyle\qquad\le\|s\mapsto s^{-\delta}S(t-s)i\|_{\gamma(L^{2}(0,\frac{t}{2};H),E_{\alpha})}+\|s\mapsto(t-s)^{-\delta}S(s)i\|_{\gamma(L^{2}(0,\frac{t}{2};H),E_{\alpha})}.

Applying Proposition 11 to both terms on the right-hand side, if αβ<12\alpha-\beta<\frac{1}{2} it follows that

[ssδS(ts)i]γ(L2(0,t;H),Eα)[s\mapsto s^{-\delta}S(t-s)i]\in\gamma(L^{2}(0,t;H),E_{\alpha})

for all t[0,T]t\in[0,T].

We need to introduce the following terminology. Let EE and FF be Banach spaces. A family of operators (E,F)\mathscr{R}\subseteq\mathscr{L}(E,F) is called γ\gamma-bounded if there exists a finite constant C0C\ge 0 such that for all finite choices R1,,RNR_{1},\dots,R_{N}\in\mathscr{R} and vectors x1,,xNEx_{1},\dots,x_{N}\in E we have

𝔼n=1NγnRnxn2C2𝔼n=1Nγnxn2.{\mathbb{E}}\Big{\|}\sum_{n=1}^{N}\gamma_{n}R_{n}x_{n}\Big{\|}^{2}\le C^{2}{\mathbb{E}}\Big{\|}\sum_{n=1}^{N}\gamma_{n}x_{n}\Big{\|}^{2}.

The least admissible constant CC is called the γ\gamma-bound of \mathscr{R}, notation γ()\gamma(\mathscr{R}). We refer to [6, 9, 22, 40] for examples and more information. In these references the related notion of RR-boundedness is discussed; this notion is obtained by replacing the Gaussian random variables by Rademacher variables in the above definition. Any RR-bounded set is also γ\gamma-bounded, and the two notions are equivalent if EE has finite cotype.

We continue with a multiplier result, also due to Kalton and Weis [20]. We refer to [28, Section 5] for a detailed proof.

Proposition 14.

Suppose that EE and FF are Banach spaces and M:(0,T)(E,F)M:(0,T)\to\mathscr{L}(E,F) is a strongly measurable function (in the sense that tM(t)xt\mapsto M(t)x is strongly measurable for every xEx\in E) with γ\gamma-bounded range ={M(t):t(0,T)}\mathscr{M}=\{M(t):\ t\in(0,T)\}. Then for every finite rank simple function Φ:(0,T)γ(H,E)\Phi:(0,T)\to\gamma(H,E) the operator RMΦR_{M\Phi} belongs to γ(L2(0,T;H),F)\gamma_{\infty}(L^{2}(0,T;H),F) and

RMΦγ(L2(0,T;H),F)γ()RΦγ(L2(0,T;H),E).\|R_{M\Phi}\|_{\gamma_{\infty}(L^{2}(0,T;H),F)}\le\gamma(\mathscr{M})\,\|R_{\Phi}\|_{\gamma(L^{2}(0,T;H),E)}.

As a result, the map M~:RΦRMΦ\widetilde{M}:R_{\Phi}\mapsto R_{M\Phi} has a unique extension to a bounded operator

M~:γ(L2(0,T;H),E)γ(L2(0,T;H),F)\widetilde{M}:\gamma(L^{2}(0,T;H),E)\to\gamma_{\infty}(L^{2}(0,T;H),F)

of norm M~γ()\|\widetilde{M}\|\le\gamma(\mathscr{M}).

In the applications of this result below it will usually be possible to check that actually we have RMΦγ(L2(0,T;H),F)R_{M\Phi}\in\gamma(L^{2}(0,T;H),F).

We will also need the following sufficient condition for γ\gamma-boundedness, which is a variation of a result of Weis [40, Proposition 2.5].

Proposition 15.

Let EE and FF be Banach spaces, and let f:(0,T)(E,F)f:(0,T)\to\mathscr{L}(E,F) be an function such that for all xEx\in E the function tf(t)xt\mapsto f(t)x is continuously differentiable with integrable derivative. Then the set :={f(t):t(0,T)}\mathscr{F}:=\{f(t):\ t\in(0,T)\} is γ\gamma-bounded in (E,F)\mathscr{L}(E,F) and

γ()f(0+)+f1.\gamma(\mathscr{F})\le\|f(0+)\|+\|f^{\prime}\|_{1}.

Here is a simple application:

Lemma 16.

Let the C0C_{0}-semigroup SS be analytic on EE.

  1. (1)

    For all 0α<δ0\le\alpha<\delta and t(0,T]t\in(0,T] the set 𝒮α,δ,t={sδS(s):s[0,t]}\mathscr{S}_{\alpha,\delta,t}=\{s^{\delta}S(s):\ s\in[0,t]\} is γ\gamma-bounded in (E,Eα)\mathscr{L}(E,E_{\alpha}) and we have

    γ(𝒮α,δ,t)tδα,t(0,T],\gamma(\mathscr{S}_{\alpha,\delta,t})\lesssim t^{\delta-\alpha},\quad t\in(0,T],

    with implied constant independent of t(0,T]t\in(0,T].

  2. (2)

    For all 0<α10<\alpha\le 1 the set 𝒯α,t={S(s)I:s[0,t]}\mathscr{T}_{\alpha,t}=\{S(s)-I:\ s\in[0,t]\} is γ\gamma-bounded in (Eα,E)\mathscr{L}(E_{\alpha},E) and we have

    γ(𝒯α,t)tα,t[0,T],\gamma(\mathscr{T}_{\alpha,t})\lesssim t^{\alpha},\quad t\in[0,T],

    with implied constant independent of t[0,T]t\in[0,T].

Proof.

For the proof of (1) we refer to [10] or [29, Lemma 10.17]. To prove (2) it will be shown that for any fixed and large enough ww\in{\mathbb{R}} the set

𝒯α,tw:={ewsS(s)I:s[0,t]}\mathscr{T}_{\alpha,t}^{w}:=\{e^{-ws}S(s)-I:\ s\in[0,t]\}

is γ\gamma-bounded in (Eα,E)\mathscr{L}(E_{\alpha},E) with γ\gamma-bound tα\lesssim t^{\alpha}. From this we deduce that {S(s):s[0,t]}\{S(s):\ s\in[0,t]\} is γ\gamma-bounded in (Eα,E)\mathscr{L}(E_{\alpha},E) with γ\gamma-bound 1\lesssim 1. In view of the identity

S(s)I=(ewsS(s)I)+(1ews)S(s)S(s)-I=(e^{-ws}S(s)-I)+(1-e^{-ws})S(s)

and noting that 1ewss1-e^{-ws}\lesssim s, this will prove the assertion of the lemma.

For all xEx\in E and 0st0\le s\le t,

ewsS(s)xx=0sewr(Aw)S(r)x𝑑r.e^{-ws}S(s)x-x=\int_{0}^{s}e^{-wr}(A-w)S(r)x\,dr.

By (7) and Proposition 15 the set 𝒯α,tw\mathscr{T}_{\alpha,t}^{w} is γ\gamma-bounded in (Eα,E)\mathscr{L}(E_{\alpha},E) and γ(𝒯α,tw)0tsα1𝑑stα.\gamma(\mathscr{T}_{\alpha,t}^{w})\lesssim\int_{0}^{t}s^{\alpha-1}\,ds\lesssim{t^{\alpha}}.

We shall again write U=U0U=U_{0} and U(n)=U0(n)U^{(n)}=U_{0}^{(n)} for the solution of (SCP0) and its approximations by the splitting scheme.

Theorem 17.

Assume that the semigroup SS is analytic on EE and that WW is a Brownian motion in EβE_{\beta} for some β0\beta\ge 0. Then the equivalent conditions of Proposition 7 and Theorem 8 hold. Moreover, for all α0\alpha\ge 0 and 0θ10\le\theta\le 1 such that αβ+θ<12\alpha-\beta+\theta<\frac{1}{2}, and all t[0,T]t\in[0,T] we have

RΦ(n)RΦγ(L2(0,t;H),Eα)nθt12(αβ+θ)+.\displaystyle\|R_{\Phi^{(n)}}-R_{\Phi}\|_{\gamma(L^{2}(0,t;H),E_{\alpha})}\lesssim n^{-\theta}t^{\frac{1}{2}-(\alpha-\beta+\theta)^{+}}. (8)

with implied constant independent of n1n\ge 1 and t[0,T]t\in[0,T]. As a consequence, for all 1p<1\le p<\infty the solution UU of (SCP0) satisfies

(𝔼U(n)(t)U(t)Eαp)1pnθt12(αβ+θ)+\big{(}{\mathbb{E}}\|U^{(n)}(t)-U(t)\|_{E_{\alpha}}^{p}\big{)}^{\frac{1}{p}}\lesssim n^{-\theta}t^{\frac{1}{2}-(\alpha-\beta+\theta)^{+}} (9)

with implied constant independent of n1n\ge 1 and t[0,T]t\in[0,T].

Proof.

The estimate (9) follows from (8) via Theorem 8.

By rescaling time we may assume that T=1T=1. Let α,β,θ\alpha,\beta,\theta be as indicated. We begin by noting that the embedding i:HEi:H\hookrightarrow E associated with WW belongs to γ(H,Eβ)\gamma(H,E_{\beta}).

Pick (αβ+θ)+<δ<12(\alpha-\beta+\theta)^{+}<\delta<\frac{1}{2}. Note that for 0<sT0<s\le T we have S(n)(s)=S(n1ns)S^{(n)}(s)=S(n^{-1}\lceil ns\rceil) and sn1nss\le n^{-1}\lceil ns\rceil, so one can write, for all n1n\ge 1,

Φ(n)(s)Φ(s)=sδS(s)(S(n1nss)I)sδi.\Phi^{(n)}(s)-\Phi(s)=s^{\delta}S(s)\circ\big{(}S(n^{-1}\lceil ns\rceil-s)-I\big{)}\circ s^{-\delta}i. (10)

Fix t(0,1]t\in(0,1]. By the first part of Lemma 16 the set

𝒮δ={sδS(s):s[0,t]}\mathscr{S}_{\delta}=\{s^{\delta}S(s):\ s\in[0,t]\}

is γ\gamma-bounded in (E,E(αβ+θ)+)\mathscr{L}(E,E_{(\alpha-\beta+\theta)^{+}}) (hence in (E,Eαβ+θ)\mathscr{L}(E,E_{\alpha-\beta+\theta}), hence in (Eβθ,Eα)\mathscr{L}(E_{\beta-\theta},E_{\alpha}), with the same upper bounds for the γ\gamma-bounds, because S(t)S(t) commutes with the fractional powers of AA) and we have

γ(𝒮δ)tδ(αβ+θ)+.\displaystyle\gamma(\mathscr{S}_{\delta})\lesssim t^{\delta-(\alpha-\beta+\theta)^{+}}. (11)

By the second part of Lemma 16 the set

𝒯θ,1n={S(s)I:s[0,n1]}\mathscr{T}_{\theta,\frac{1}{n}}=\{S(s)-I:\ s\in[0,n^{-1}]\}

is γ\gamma-bounded in (Eθ,E)\mathscr{L}(E_{\theta},E) (and hence in (Eβ,Eβθ)\mathscr{L}(E_{\beta},E_{\beta-\theta}), with the same estimate for the γ\gamma-boundedness constant), and we have

γ(𝒯θ,1n)nθ.\displaystyle\gamma(\mathscr{T}_{\theta,\frac{1}{n}})\lesssim n^{-\theta}. (12)

Using (10), Remark 13, Proposition 14, the identity

Rssδiγ(L2(0,t;H),Eβ)\displaystyle\|R_{s\mapsto s^{-\delta}i}\|_{\gamma(L^{2}(0,t;H),E_{\beta})} =ssδL2(0,t)iγ(H,Eβ)t12δiγ(H,Eβ)\displaystyle=\|s\mapsto s^{-\delta}\|_{L^{2}(0,t)}\|i\|_{\gamma(H,E_{\beta})}\eqsim t^{\frac{1}{2}-\delta}\|i\|_{\gamma(H,E_{\beta})} (13)

together with the estimates (11), and (12), and noting that n1nssn1n^{-1}\lceil ns\rceil-s\le n^{-1}, one obtains

RΦ(n)RΦγ(L2(0,t;H),Eα)\displaystyle\|R_{\Phi^{(n)}}-R_{\Phi}\|_{\gamma(L^{2}(0,t;H),E_{\alpha})} γ(𝒮δ)γ(𝒯θ,1n)Rssδiγ(L2(0,t;H),Eβ)\displaystyle\le\gamma(\mathscr{S}_{\delta})\gamma(\mathscr{T}_{\theta,\frac{1}{n}})\|R_{s\mapsto s^{-\delta}i}\|_{\gamma(L^{2}(0,t;H),E_{\beta})}
nθt12(αβ+θ)+iγ(H,Eβ).\displaystyle\lesssim n^{-\theta}t^{\frac{1}{2}-(\alpha-\beta+\theta)^{+}}\|i\|_{\gamma(H,E_{\beta})}.

Remark 18.

The condition αβ+θ<12\alpha-\beta+\theta<\frac{1}{2} implies, in view of the restriction 0θ10\le\theta\le 1, that αβ<12\alpha-\beta<\frac{1}{2}. For αβ<12\alpha-\beta<-\frac{1}{2}, Theorem 17 gives a rate of convergence of order n1n^{-1}, whereas for 12αβ<12-\frac{1}{2}\le\alpha-\beta<\frac{1}{2} we obtain the rate nθn^{-\theta} for any 0θ<12α+β0\le\theta<\frac{1}{2}-\alpha+\beta. For 12<αβ<12-\frac{1}{2}<\alpha-\beta<\frac{1}{2} one can in fact obtain a slightly better rate at the final time TT, namely (lnlnn)/n(12α+β)(\ln\ln n)/n^{(\frac{1}{2}-\alpha+\beta)}. More precisely, for n3n\ge 3 we have

RΦ(n)RΦγ(L2(0,T;H),Eα)lnlnnn12α+β\displaystyle\|R_{\Phi^{(n)}}-R_{\Phi}\|_{\gamma(L^{2}(0,T;H),E_{\alpha})}\lesssim\frac{\ln\ln n}{n^{\frac{1}{2}-\alpha+\beta}} (14)

with constants independent of n1n\ge 1.

Once again observe that by scaling we may (and do) assume that T=1T=1. In order to prove (14) we first give an estimate for a given time interval [a,b][a,b] where 0<a<b10<a<b\le 1. In that case, for δ>αβ+1\delta>\alpha-\beta+1 one has

RΦ(n)RΦγ(L2(a,b;H),Eα)n1a12δbδαβ+1\|R_{\Phi^{(n)}}-R_{\Phi}\|_{\gamma(L^{2}(a,b;H),E_{\alpha})}\lesssim n^{-1}a^{\frac{1}{2}-\delta}b^{\delta-\alpha-\beta+1} (15)

with implied constant independent of n1n\ge 1 and 0<a<b10<a<b\le 1. The proof of (15) is similar to that of (8), the main difference being that we no longer need to ensure the square integrability of ssδs\mapsto s^{-\delta} near s=0s=0 in (13). The details are as follows. Fix n1n\ge 1 and 0<a<b10<a<b\le 1 and pick an arbitrary δ>αβ+1\delta>\alpha-\beta+1. Then,

Rssδiγ(L2(a,b;H),Eβ)=ssδL2(a,b)iγ(H,Eβ)a12δiγ(H,Eβ),\displaystyle\|R_{s\mapsto s^{-\delta}i}\|_{\gamma(L^{2}(a,b;H),E_{\beta})}=\|s\mapsto s^{-\delta}\|_{L^{2}(a,b)}\|i\|_{\gamma(H,E_{\beta})}\lesssim a^{\frac{1}{2}-\delta}\|i\|_{\gamma(H,E_{\beta})}, (16)

with implied constant independent of a(0,1]a\in(0,1] and b(a,1]b\in(a,1]; the last inequality uses that δ12\delta\ge\frac{1}{2}. As in the proof of Theorem 17 the set 𝒯1n:={S(s)I:s[0,n1]}\mathscr{T}_{\frac{1}{n}}:=\{S(s)-I:\ s\in[0,n^{-1}]\} is γ\gamma-bounded in (Eβ,Eβ1)\mathscr{L}(E_{\beta},E_{\beta-1}), with γ\gamma-bound

γ(𝒯1n)n1.\displaystyle\gamma(\mathscr{T}_{\frac{1}{n}})\lesssim n^{-1}. (17)

Finally, since δ>αβ+1\delta>\alpha-\beta+1, as in the proof of Theorem 17 the set 𝒮δ={sδS(s):s[a,b]}\mathscr{S}_{\delta}=\{s^{\delta}S(s):\ s\in[a,b]\} is γ\gamma-bounded in (Eβ1,Eα)\mathscr{L}(E_{\beta-1},E_{\alpha}) with

γ(𝒮δ)bδαβ+1.\displaystyle\gamma(\mathscr{S}_{\delta})\lesssim b^{\delta-\alpha-\beta+1}. (18)

Combining (16), (17), and (18) we obtain

RΦ(n)RΦγ(L2(a,b;H),Eα)n1a12δbδαβ+1iγ(H,Eβ).\displaystyle\|R_{\Phi^{(n)}}-R_{\Phi}\|_{\gamma(L^{2}(a,b;H),E_{\alpha})}\lesssim n^{-1}a^{\frac{1}{2}-\delta}b^{\delta-\alpha-\beta+1}\|i\|_{\gamma(H,E_{\beta})}.

Returning to the proof of estimate (14) we fix an integer n3n\ge 3. Because βα<12\beta-\alpha<\frac{1}{2} one can pick δ>0\delta>0 such that 1+αβ<δ32+2(αβ)1+\alpha-\beta<\delta\le\frac{3}{2}+2(\alpha-\beta). For j=0,1,j=0,1,\dots define aj:=n1+2ja_{j}:=n^{-1+2^{-j}}. Note that a0=1a_{0}=1 and limjaj=n1\lim_{j\to\infty}a_{j}=n^{-1}. If in (15) we take a=aja=a_{j} and b=aj1b=a_{j-1} we obtain the estimate

RΦ(n)RΦγ(L2(aj,aj1;H),Eα)\displaystyle\|R_{\Phi^{(n)}}-R_{\Phi}\|_{\gamma(L^{2}(a_{j},a_{j-1};H),E_{\alpha})}
n1+(1+2j)(12δ)+(1+21j)(δα+β1)iγ(H,Eβ)\displaystyle\qquad\lesssim n^{-1+(-1+2^{-j})(\frac{1}{2}-\delta)+(-1+2^{1-j})(\delta-\alpha+\beta-1)}\|i\|_{\gamma(H,E_{\beta})}
=n12+αβ+2j(δ322(αβ))iγ(H,Eβ)\displaystyle\qquad=n^{-\frac{1}{2}+\alpha-\beta+2^{-j}(\delta-\frac{3}{2}-2(\alpha-\beta))}\|i\|_{\gamma(H,E_{\beta})}
n12+αβiγ(H,Eβ),\displaystyle\qquad\le n^{-\frac{1}{2}+\alpha-\beta}\|i\|_{\gamma(H,E_{\beta})},

where the last inequality used that δ32+2(αβ)\delta\le\frac{3}{2}+2(\alpha-\beta). Set kn=(lnlnn)/ln2k_{n}=\lceil(\ln\ln n)/\ln 2\rceil, so that aknen1a_{k_{n}}\le en^{-1}. Using this estimate for akna_{k_{n}}, from Theorem 17 we obtain, for any choice of 0θ<12α+β0\le\theta<\frac{1}{2}-\alpha+\beta (which then satisfies θ<1\theta<1),

RΦ(n)RΦγ(L2(0,akn;H),Eα)akn12α+βθnθiγ(H,Eβ)n12+αβiγ(H,Eβ).\|R_{\Phi^{(n)}}-R_{\Phi}\|_{\gamma(L^{2}(0,a_{k_{n}};H),E_{\alpha})}\lesssim a_{k_{n}}^{\frac{1}{2}-\alpha+\beta-\theta}n^{-\theta}\|i\|_{\gamma(H,E_{\beta})}\lesssim n^{-\frac{1}{2}+\alpha-\beta}\|i\|_{\gamma(H,E_{\beta})}.

Combining the above one gets

RΦ(n)RΦγ(L2(0,1;H),Eα)\displaystyle\|R_{\Phi^{(n)}}-R_{\Phi}\|_{\gamma(L^{2}(0,1;H),E_{\alpha})}
RΦ(n)RΦγ(L2(0,akn;H),Eα)+j=1kNRΦ(n)RΦγ(L2(aj,aj1;H),Eα)\displaystyle\qquad\lesssim\|R_{\Phi^{(n)}}-R_{\Phi}\|_{\gamma(L^{2}(0,a_{k_{n}};H),E_{\alpha})}+\sum_{j=1}^{k_{N}}\|R_{\Phi^{(n)}}-R_{\Phi}\|_{\gamma(L^{2}(a_{j},a_{j-1};H),E_{\alpha})}
(1+lnlnn)n12+αβiγ(H,Eβ).\displaystyle\qquad\lesssim(1+\ln\ln n)n^{-\frac{1}{2}+\alpha-\beta}\|i\|_{\gamma(H,E_{\beta})}.

This gives the estimate (14).

Under the assumptions that SS is analytic on EE and WW is a Brownian motion on EE, the solution UU of (SCP0) has a version with trajectories in Cγ([0,T];Eα)C^{\gamma}([0,T];E_{\alpha}) for any α,γ0\alpha,\gamma\ge 0 such that α+γ<12\alpha+\gamma<\frac{1}{2} [10]. The main result of this paper asserts that also the approximating processes U(n)U^{(n)} have trajectories in Cγ([0,T];Eα)C^{\gamma}([0,T];E_{\alpha}) and that the splitting scheme converges with respect to the Cγ([0,T];Eα)C^{\gamma}([0,T];E_{\alpha})-norm, with a convergence rate depending on α\alpha and γ\gamma and the smoothness of the noise.

Theorem 19.

Let SS be analytic on EE and suppose that WW is a Brownian motion in EβE_{\beta} for some β0\beta\ge 0. If α,θ,γ0\alpha,\theta,\gamma\ge 0 satisfy θ+γ<1\theta+\gamma<1 and (αβ+θ)++γ<12(\alpha-\beta+\theta)^{+}+\gamma<\frac{1}{2}, then for all 1p<1\le p<\infty the solution UU of (SCP0) satisfies

(𝔼U(n)UCγ([0,T],Eα)p)1pnθ,\displaystyle\big{(}{\mathbb{E}}\|U^{(n)}-U\|_{C^{\gamma}([0,T],E_{\alpha})}^{p}\big{)}^{\frac{1}{p}}\lesssim n^{-\theta},

with implied constant independent of n1n\ge 1.

Proof.

By scaling we may assume T=1T=1. Put V(n):=U(n)UV^{(n)}:=U^{(n)}-U. Let α,β,γ\alpha,\beta,\gamma and θ\theta be as indicated. Without loss of generality we assume that γ>0\gamma>0. The main step in the proof is the following claim.

Claim 20.

There exists a constant CC such that for all n1n\ge 1, all 0s<t10\le s<t\le 1 satisfying ts<12nt-s<\frac{1}{2n} we have

(𝔼V(n)(t)V(n)(s)Eα2)12Cnθ(ts)γ.\big{(}{\mathbb{E}}\|V^{(n)}(t)-V^{(n)}(s)\|_{E_{\alpha}}^{2}\big{)}^{\frac{1}{2}}\le Cn^{-\theta}(t-s)^{\gamma}.
Proof.

Fix n1n\ge 1 and 0s<t10\le s<t\le 1 such that ts<12nt-s<\frac{1}{2n}. Clearly,

(𝔼V(n)(t)V(n)(s)Eα2)12\displaystyle\ \big{(}{\mathbb{E}}\|V^{(n)}(t)-V^{(n)}(s)\|_{E_{\alpha}}^{2}\big{)}^{\frac{1}{2}} (𝔼stΦ(tr)Φ(n)(tr)dW(r)2)12\displaystyle\le\Big{(}{\mathbb{E}}\Big{\|}\int_{s}^{t}\Phi(t-r)-\Phi^{(n)}(t-r)dW(r)\Big{\|}^{2}\Big{)}^{\frac{1}{2}} (19)
+(𝔼0sΦ(tr)Φ(sr)dW(r)2)12\displaystyle\qquad+\Big{(}{\mathbb{E}}\Big{\|}\int_{0}^{s}\Phi(t-r)-\Phi(s-r)dW(r)\Big{\|}^{2}\Big{)}^{\frac{1}{2}}
+(𝔼0sΦ(n)(tr)Φ(n)(sr)dW(r)2)12.\displaystyle\qquad+\Big{(}{\mathbb{E}}\Big{\|}\int_{0}^{s}\Phi^{(n)}(t-r)-\Phi^{(n)}(s-r)dW(r)\Big{\|}^{2}\Big{)}^{\frac{1}{2}}.

For the first term we note that by (3) (and the remark following it) and (8) one has

(𝔼stΦ(n)(tr)Φ(tr)dWH(r)Eα2)12\displaystyle\Big{(}{\mathbb{E}}\Big{\|}\int_{s}^{t}\Phi^{(n)}(t-r)-\Phi(t-r)\,dW_{H}(r)\Big{\|}^{2}_{E_{\alpha}}\Big{)}^{\frac{1}{2}} (20)
=(𝔼0tsΦ(n)(r)Φ(r)dWH(r)Eα2)12\displaystyle\qquad=\Big{(}{\mathbb{E}}\Big{\|}\int_{0}^{t-s}\Phi^{(n)}(r)-\Phi(r)\,dW_{H}(r)\Big{\|}^{2}_{E_{\alpha}}\Big{)}^{\frac{1}{2}}
nθ(ts)12(αβ+θ)+iγ(H,Eβ)\displaystyle\qquad\lesssim n^{-\theta}(t-s)^{\frac{1}{2}-(\alpha-\beta+\theta)^{+}}\|i\|_{\gamma(H,E_{\beta})}
nθ(ts)γiγ(H,Eβ).\displaystyle\qquad\le n^{-\theta}(t-s)^{\gamma}\|i\|_{\gamma(H,E_{\beta})}.

The estimate for the second term is extracted from arguments in [31]; see also [29, Theorem 10.19]. Fix η>0\eta>0 such that (αβ+θ)++γ<η<12(\alpha-\beta+\theta)^{+}+\gamma<\eta<\frac{1}{2}. Then the set {tηS(t):t(0,T)}\{t^{\eta}S(t):\ t\in(0,T)\} is γ\gamma-bounded in (E,E(αβ+θ)++γ)\mathscr{L}(E,E_{(\alpha-\beta+\theta)^{+}+\gamma}) (hence in (E,Eαβ+θ+γ)\mathscr{L}(E,E_{\alpha-\beta+\theta+\gamma}), hence in (Eβθγ,Eα)\mathscr{L}(E_{\beta-\theta-\gamma},E_{\alpha})) by the first part of Lemma 16, and therefore

(𝔼0sΦ(tr)Φ(sr)dWH(r)Eα2)12\displaystyle\Big{(}{\mathbb{E}}\Big{\|}\int_{0}^{s}\Phi(t-r)-\Phi(s-r)\,dW_{H}(r)\Big{\|}_{E_{\alpha}}^{2}\Big{)}^{\frac{1}{2}} (21)
=(𝔼0s[(sr)ηS(sr)][(sr)η(S(ts)I)i]𝑑WH(r)Eα2)12\displaystyle\qquad=\Big{(}{\mathbb{E}}\Big{\|}\int_{0}^{s}[(s-r)^{\eta}S(s-r)]\circ[(s-r)^{-\eta}(S(t-s)-I)\circ i]\,dW_{H}(r)\Big{\|}_{E_{\alpha}}^{2}\Big{)}^{\frac{1}{2}}
(𝔼0s(sr)η(S(ts)I)i)dWH(r)Eβθγ2)12\displaystyle\qquad\lesssim\Big{(}{\mathbb{E}}\Big{\|}\int_{0}^{s}(s-r)^{-\eta}(S(t-s)-I)\circ i)\,dW_{H}(r)\Big{\|}_{E_{\beta-\theta-\gamma}}^{2}\Big{)}^{\frac{1}{2}}
=(0s(sr)2η𝑑r)12(S(ts)I)iγ(H,Eβθγ)\displaystyle\qquad=\Big{(}\int_{0}^{s}(s-r)^{-2\eta}\,dr\Big{)}^{\frac{1}{2}}\|(S(t-s)-I)\circ i\|_{\gamma(H,E_{\beta-\theta-\gamma})}
S(ts)I(Eβ,Eβθγ)iγ(H,Eβ)\displaystyle\qquad\lesssim\|S(t-s)-I\|_{\mathscr{L}(E_{\beta},E_{\beta-\theta-\gamma})}\|i\|_{\gamma(H,E_{\beta})}
S(ts)I(Eγ+θ,E)iγ(H,Eβ)\displaystyle\qquad\eqsim\|S(t-s)-I\|_{\mathscr{L}(E_{\gamma+\theta},E)}\|i\|_{\gamma(H,E_{\beta})}
(ts)γ+θiγ(H,Eβ)\displaystyle\qquad\lesssim(t-s)^{\gamma+\theta}\|i\|_{\gamma(H,E_{\beta})}
nθ(ts)γiγ(H,Eβ).\displaystyle\qquad\lesssim n^{-\theta}(t-s)^{\gamma}\|i\|_{\gamma(H,E_{\beta})}.

To estimate the third term on the right-hand side of (19), we first define sets B0B_{0} and B1B_{1} by

B0:=\displaystyle B_{0}= {r(0,s):S(n)(tr)=S(n)(sr)}\displaystyle\,\{r\in(0,s):\ S^{(n)}(t-r)=S^{(n)}(s-r)\}
=\displaystyle= {r(0,s):n(tr)=n(sr)},\displaystyle\,\{r\in(0,s):\ \lceil n(t-r)\rceil=\lceil n(s-r)\rceil\},
B1:=\displaystyle B_{1}= {r(0,s):S(n)(tr)=S(n1)S(n)(sr)}\displaystyle\,\{r\in(0,s):\ S^{(n)}(t-r)=S(n^{-1})S^{(n)}(s-r)\}
=\displaystyle= {r(0,s):n(tr)=n(sr)+1}.\displaystyle\,\{r\in(0,s):\ \lceil n(t-r)\rceil=\lceil n(s-r)\rceil+1\}.

Both equalities follow from the identity S(n)(u)=S(n1nu)S^{(n)}(u)=S(n^{-1}\lceil nu\rceil) for u(0,T)u\in(0,T). By definition of B0B_{0} and B1B_{1} one has

(𝔼0sΦ(n)(tr)Φ(n)(sr)dWH(r)Eα2)12\displaystyle\Big{(}{\mathbb{E}}\Big{\|}\int_{0}^{s}\Phi^{(n)}(t-r)-\Phi^{(n)}(s-r)\,dW_{H}(r)\Big{\|}^{2}_{E_{\alpha}}\Big{)}^{\frac{1}{2}} (22)
(𝔼B0Φ(n)(tr)Φ(n)(sr)dWH(r)Eα2)12\displaystyle\qquad\le\Big{(}{\mathbb{E}}\Big{\|}\int_{B_{0}}\Phi^{(n)}(t-r)-\Phi^{(n)}(s-r)\,dW_{H}(r)\Big{\|}^{2}_{E_{\alpha}}\Big{)}^{\frac{1}{2}}
+(𝔼B1Φ(n)(tr)Φ(n)(sr)dWH(r)Eα2)12\displaystyle\qquad\qquad+\Big{(}{\mathbb{E}}\Big{\|}\int_{B_{1}}\Phi^{(n)}(t-r)-\Phi^{(n)}(s-r)\,dW_{H}(r)\Big{\|}^{2}_{E_{\alpha}}\Big{)}^{\frac{1}{2}}
=(𝔼B1S(n)(sr)(S(n1)I)i𝑑WH(r)Eα2)12,\displaystyle\qquad=\Big{(}{\mathbb{E}}\Big{\|}\int_{B_{1}}S^{(n)}(s-r)(S(n^{-1})-I)i\,dW_{H}(r)\Big{\|}^{2}_{E_{\alpha}}\Big{)}^{\frac{1}{2}},

noting that the integrand of the integral over B0B_{0} vanishes.

Set δ:=θ+γ\delta:=\theta+\gamma. To estimate the right-hand side, observe that from αβ+δ<12\alpha-\beta+\delta<\frac{1}{2} we may pick η>0\eta>0 such that αβ+δ<η<12\alpha-\beta+\delta<\eta<\frac{1}{2}. Using the identity S(n)(u)=S(n1nu)S^{(n)}(u)=S(n^{-1}\lceil nu\rceil) and applying Proposition 14 and part (1) of Lemma 16, and then using the estimate S(u)I(Eδ,E)uδ\|S(u)-I\|_{\mathscr{L}(E_{\delta},E)}\lesssim u^{\delta} and Proposition 6, we obtain

(𝔼B1S(n)(sr)(S(n1)I)i𝑑WH(r)Eα2)12\displaystyle\Big{(}{\mathbb{E}}\Big{\|}\int_{B_{1}}S^{(n)}(s-r)(S(n^{-1})-I)i\,dW_{H}(r)\Big{\|}^{2}_{E_{\alpha}}\Big{)}^{\frac{1}{2}} (23)
(𝔼B1(n1n(sr))ηS(n1n(sr))\displaystyle\quad\eqsim\Big{(}{\mathbb{E}}\Big{\|}\int_{B_{1}}(n^{-1}\lceil n(s-r)\rceil)^{\eta}S(n^{-1}\lceil n(s-r)\rceil)
×(n1n(sr))η(S(n1)I)idWH(r)Eα2)12\displaystyle\quad\quad\times(n^{-1}\lceil n(s-r)\rceil)^{-\eta}(S(n^{-1})-I)i\,dW_{H}(r)\Big{\|}^{2}_{E_{\alpha}}\Big{)}^{\frac{1}{2}}
(𝔼B1(n1n(sr))η(S(n1)I)i𝑑WH(r)Eβδ2)12\displaystyle\quad\lesssim\Big{(}{\mathbb{E}}\Big{\|}\int_{B_{1}}(n^{-1}\lceil n(s-r)\rceil)^{-\eta}(S(n^{-1})-I)i\,dW_{H}(r)\Big{\|}^{2}_{E_{\beta-\delta}}\Big{)}^{\frac{1}{2}}
nδ(s)ηL2(B1)iγ(H,Eβ).\displaystyle\quad\lesssim n^{-\delta}\|(s-\cdot)^{-\eta}\|_{L^{2}(B_{1})}\|i\|_{\gamma(H,E_{\beta})}.

In order to estimate the L2(B1)L^{2}(B_{1})-norm of the function fs(r):=(sr)ηf_{s}(r):=(s-r)^{-\eta} we note that B1j=1nB1(j)B_{1}\subseteq\bigcup_{j=1}^{n}B_{1}^{(j)}, where

B1(j)\displaystyle B_{1}^{(j)} ={r(0,s):srjn1<tr}\displaystyle=\{r\in(0,s):\ s-r\le jn^{-1}<t-r\}
={r(0,s):jn1t+s<srjn1}.\displaystyle=\{r\in(0,s):\ jn^{-1}-t+s<s-r\le jn^{-1}\}.

From this it is easy to see that |B1(j)|ts|B_{1}^{(j)}|\le t-s and that for rB1(j)r\in B_{1}^{(j)} one has

(sr)2η(jn1t+s)2ηn2η(j12)2η(s-r)^{-2\eta}\le(jn^{-1}-t+s)^{-2\eta}\le n^{2\eta}(j-\tfrac{1}{2})^{-2\eta}

(the latter inequality following from ts<1/2nt-s<1/2n), and therefore

fsL2(B1)2=B1|fs(r)|2𝑑rn2η|B1(j)|j=1n1(j12)2ηn(ts).\|f_{s}\|_{L^{2}(B_{1})}^{2}=\int_{B_{1}}|f_{s}(r)|^{2}\,dr\le n^{2\eta}|B_{1}^{(j)}|\sum_{j=1}^{n}\frac{1}{(j-\tfrac{1}{2})^{2\eta}}\lesssim n(t-s).

As a consequence,

fsL2(B1)n12(ts)12=n12(ts)12γ(ts)γnγ(ts)γ.\|f_{s}\|_{L^{2}(B_{1})}\lesssim n^{\frac{1}{2}}(t-s)^{\frac{1}{2}}=n^{\frac{1}{2}}(t-s)^{\frac{1}{2}-\gamma}(t-s)^{\gamma}\lesssim n^{\gamma}(t-s)^{\gamma}. (24)

Combining the estimates (23) and (24) and estimating the non-negative powers of ss by 11 we find

(𝔼B1S(n)(sr)(S(n)(ts)I)i𝑑WH(r)Eα2)12nθ(ts)γiγ(H,Eβ).\Big{(}{\mathbb{E}}\Big{\|}\int_{B_{1}}S^{(n)}(s-r)(S^{(n)}(t-s)-I)i\,dW_{H}(r)\Big{\|}^{2}_{E_{\alpha}}\Big{)}^{\frac{1}{2}}\\ \lesssim n^{-\theta}(t-s)^{\gamma}\|i\|_{\gamma(H,E_{\beta})}. (25)

Claim 20 now follows by combining (19), (20), (21), (22) and (25). ∎

We are now ready to finish the proof of the theorem. By the triangle inequality and Theorem 17, for all 0s,t10\le s,t\le 1 we have

(𝔼V(n)(t)V(n)(s)Eα2)12\displaystyle\big{(}{\mathbb{E}}\|V^{(n)}(t)-V^{(n)}(s)\|_{E_{\alpha}}^{2}\big{)}^{\frac{1}{2}} (𝔼U(n)(t)U(t)Eα2)12+(𝔼U(n)(s)U(s)Eα2)12\displaystyle\le\big{(}{\mathbb{E}}\|U^{(n)}(t)-U(t)\|_{E_{\alpha}}^{2}\big{)}^{\frac{1}{2}}+\big{(}{\mathbb{E}}\|U^{(n)}(s)-U(s)\|_{E_{\alpha}}^{2}\big{)}^{\frac{1}{2}}
nδiγ(H,Eβ).\displaystyle\lesssim n^{-\delta}\|i\|_{\gamma(H,E_{\beta})}.

Hence if ts(2n)1t-s\ge(2n)^{-1} one has

(𝔼V(n)(t)V(n)(s)Eα2)12nδiγ(H,Eβ)nθ(ts)γiγ(H,Eβ).\big{(}{\mathbb{E}}\|V^{(n)}(t)-V^{(n)}(s)\|_{E_{\alpha}}^{2}\big{)}^{\frac{1}{2}}\lesssim n^{-\delta}\|i\|_{\gamma(H,E_{\beta})}\lesssim n^{-\theta}(t-s)^{\gamma}\|i\|_{\gamma(H,E_{\beta})}. (26)

The random variables V(n)(t)V^{(n)}(t) being Gaussian, from the claim and (26) combined with the Kahane-Khintchine inequalities we deduce that for all 1q<1\le q<\infty and 0s<t10\le s<t\le 1 one has

(𝔼V(n)(t)V(n)(s)Eαq)1qnθ(ts)γiγ(H,Eβ).\displaystyle\big{(}{\mathbb{E}}\|V^{(n)}(t)-V^{(n)}(s)\|^{q}_{E_{\alpha}}\big{)}^{\frac{1}{q}}\lesssim n^{-\theta}(t-s)^{\gamma}\|i\|_{\gamma(H,E_{\beta})}. (27)

Now fix any 0<γ<γ0<\gamma^{\prime}<\gamma and take 1/γ<q<1/\gamma^{\prime}<q<\infty. Then by (27) and the Kolmogorov-Chentsov criterion with LqL^{q}-moments (see [11, Theorem 5]),

U(n)ULq(Ω;Cγ1q([0,T];Eα))U(n)UCγ([0,T];Lq(Ω;Eα))nθiγ(H,Eβ).\|U^{(n)}-U\|_{L^{q}(\Omega;C^{\gamma^{\prime}-\frac{1}{q}}([0,T];E_{\alpha}))}\lesssim\|U^{(n)}-U\|_{C^{\gamma}([0,T];L^{q}(\Omega;E_{\alpha}))}\lesssim n^{-\theta}\|i\|_{\gamma(H,E_{\beta})}.

This inequality shows that for all 0<γ¯<γ0<\bar{\gamma}<\gamma we have

U(n)ULq(Ω;Cγ¯([0,T];Eα))nθiγ(H,Eβ)\|U^{(n)}-U\|_{L^{q}(\Omega;C^{\bar{\gamma}}([0,T];E_{\alpha}))}\lesssim n^{-\theta}\|i\|_{\gamma(H,E_{\beta})}

for all sufficiently large 1q<1\le q<\infty. It is clear that once we know this, this inequality extends to all values 1q<1\le q<\infty. This completes the proof of the theorem (with γ¯\bar{\gamma} instead of γ\gamma, which obviously suffices). ∎

Corollary 21.

Suppose that SS is analytic on EE and that WW is a Brownian motion in EβE_{\beta} for some β0\beta\ge 0. Let α,γ,θ0\alpha,\gamma,\theta\ge 0 satisfy θ+γ<1\theta+\gamma<1 and (αβ+θ)++γ<12(\alpha-\beta+\theta)^{+}+\gamma<\frac{1}{2}. Then for almost all ωΩ\omega\in\Omega there exists a constant C(ω)C(\omega) such that the solution UU of (SCP0) satisfies

Ux(n)(,ω)Ux(,ω)Cγ([0,T];Eα)C(ω)nθ for all n=1,2,\|U_{x}^{(n)}(\cdot,\omega)-U_{x}(\cdot,\omega)\|_{C^{\gamma}([0,T];E_{\alpha})}\le\frac{C(\omega)}{n^{\theta}}\ \hbox{ for all }\ n=1,2,\dots
Proof.

Set

Ωn:={ωΩ:U(n)(,ω)U(,ω)Cγ([0,T];Eα)>1nθ}.\Omega_{n}:=\Big{\{}\omega\in\Omega:\ \|U^{(n)}(\cdot,\omega)-U(\cdot,\omega)\|_{C^{\gamma}([0,T];E_{\alpha})}>\frac{1}{n^{\theta}}\Big{\}}.

Pick θ¯>θ\bar{\theta}>\theta in such a way that 0αβ+γ+θ¯<120\le\alpha-\beta+\gamma+\bar{\theta}<\frac{1}{2} and let p1p\ge 1 be so large that (θ¯θ)p>1(\bar{\theta}-\theta)p>1. By Theorem 19, applied with θ¯\bar{\theta} instead of θ\theta, and Chebyshev’s inequality,

(Ωn)nθp𝔼U(n)(,ω)U(,ω)Cγ([0,T];Eα)pCpn(θ¯θ)p{\mathbb{P}}(\Omega_{n})\le n^{\theta p}{\mathbb{E}}\|U^{(n)}(\cdot,\omega)-U(\cdot,\omega)\|_{C^{\gamma}([0,T];E_{\alpha})}^{p}\le\frac{C^{p}}{n^{(\bar{\theta}-\theta)p}}

with constant CC independent of nn. By the choice of pp we have n1(Ωn)<\sum_{n\ge 1}{\mathbb{P}}(\Omega_{n})<\infty, and therefore by the Borel-Cantelli lemma

({ωΩ:ωΩn infinitely often})=0.{\mathbb{P}}(\{\omega\in\Omega:\ \omega\in\Omega_{n}\hbox{ infinitely often}\})=0.

For the ωΩ\omega\in\Omega belonging to this set we have

C(ω):=supn1nθU(n)(,ω)U(,ω)Cγ([0,T];Eα)<.C(\omega):=\sup_{n\ge 1}n^{\theta}\|U^{(n)}(\cdot,\omega)-U(\cdot,\omega)\|_{C^{\gamma}([0,T];E_{\alpha})}<\infty.

We conclude this section with an application of our results to the stochastic heat equation on the unit interval driven by space-time white noise. This example is merely included as a demonstration how such equations can be handled in the present framework. We don’t strive for the greatest possible generality. For instance, as in [5, 10] the Laplace operator can be replaced by more general second order elliptic operators.

Example 22.

Consider the following stochastic partial differential equation driven by space-time white noise ww:

{ut(t,x)=Δu(t,x)+wt(t,x),x[0,1],t[0,T],u(0,x)=0,x[0,1],u(t,0)=u(t,1)=0,t[0,T].\left\{\begin{aligned} \frac{\partial{u}}{\partial t}(t,x)&=\Delta u(t,x)+\frac{\partial w}{\partial t}(t,x),\quad&x\in[0,1],\ t\in[0,T],\\ u(0,x)&=0,\quad&x\in[0,1],\\ u(t,0)&=u(t,1)=0,\quad&t\in[0,T].\end{aligned}\right. (28)

Following the approach of [10] we put H:=L2(0,1)H:=L^{2}(0,1) and F:=Lq(0,1)F:=L^{q}(0,1), where the exponent q2q\ge 2 is to be chosen later on. In order to formulate the problem (28) as an abstract stochastic evolution equation of the form

{dU(t)=AU(t)dt+dW(t),t[0,T],U(0)=0,\left\{\begin{aligned} dU(t)&=AU(t)\,dt+dW(t),\quad t\in[0,T],\\ U(0)&=0,\end{aligned}\right. (29)

where WW is a Brownian motion with values in a suitable Banach space EE, we fix an arbitrary real number ρ<14\rho<-\frac{1}{4}, to be chosen in a moment, and let E:=FρE:=F_{\rho} denote the extrapolation space of order ρ-\rho associated with the Dirichlet Laplacian in FF. It is shown in [10] (see also [5, Lemma 6.5]) that the identity operator on HH extends to a γ\gamma-radonifying embedding from HH into EE. As a result, the HH-cylindrical Brownian motion WHW_{H} canonically associated with ww (see (5)) may be identified with a Brownian motion WW in EE. Furthermore the extrapolated Dirichlet Laplacian, henceforth denoted by AA, satisfies the assumptions of Theorem 19 in EE.

Let UU be the solution of (29) in EE. By definition, we shall regard UU as the solution of (28). Suppose now that we are given real numbers γ,δ,θ0\gamma,\delta,\theta\ge 0 satisfy

γ+δ+θ<14.\gamma+\delta+\theta<\tfrac{1}{4}.

This ensures that one can choose α0\alpha\ge 0 and ρ<14\rho<-\frac{1}{4} in such a way that α+ρ>δ\alpha+\rho>\delta and α+γ+θ<12\alpha+\gamma+\theta<\frac{1}{2}. By Theorem 19 (with β=0\beta=0), for all 1p<1\le p<\infty the splitting scheme associated with problem (29) satisfies

(𝔼U(n)UCγ([0,T],Eα)p)1pnθ.\displaystyle\big{(}{\mathbb{E}}\|U^{(n)}-U\|_{C^{\gamma}([0,T],E_{\alpha})}^{p}\big{)}^{\frac{1}{p}}\lesssim n^{-\theta}.

Putting η:=α+ρ\eta:=\alpha+\rho we have Eα=(Fρ)α=FηE_{\alpha}=(F_{\rho})_{\alpha}=F_{\eta}, and this space embeds into FF since η>δ0\eta>\delta\ge 0.

Choose q2q\ge 2 so large that 2δ+1q<2η2\delta+\frac{1}{q}<2\eta. We have

Fη=H02η,q(0,1)={fH2η,q(0,1):f(0)=f(1)=0}F_{\eta}=H_{0}^{2\eta,q}(0,1)=\{f\in H^{2\eta,q}(0,1):\ f(0)=f(1)=0\}

with equivalent norms. By the Sobolev embedding theorem,

H2η,q(0,1)C2δ[0,1]H^{2\eta,q}(0,1)\hookrightarrow C^{2\delta}[0,1]

with continuous inclusion. Here C2δ[0,1]C^{2\delta}[0,1] is the space of all Hölder continuous functions f:[0,1]f:[0,1]\to{\mathbb{R}} of exponent 2δ2\delta. We denote C02δ[0,1]={fC2δ[0,1]:f(0)=f(1)=0}C_{0}^{2\delta}[0,1]=\{f\in C^{2\delta}[0,1]:f(0)=f(1)=0\}. Putting things together we obtain a continuous inclusion

FηC02δ[0,1].F_{\eta}\hookrightarrow C_{0}^{2\delta}[0,1].

We have proved the following theorem (cf. Example 4).

Theorem 23.

For all 0δ<140\le\delta<\frac{1}{4} the stochastic heat equation (28) admits a solution UU in C02δ[0,1]C_{0}^{2\delta}[0,1], and for all γ,θ0\gamma,\theta\ge 0 satisfying γ+δ+θ<14\gamma+\delta+\theta<\frac{1}{4} we have

(𝔼U(n)UCγ([0,T],C02δ[0,1])p)1pnθ.\displaystyle\big{(}{\mathbb{E}}\|U^{(n)}-U\|_{C^{\gamma}([0,T],C_{0}^{2\delta}[0,1])}^{p}\big{)}^{\frac{1}{p}}\lesssim n^{-\theta}.

By Corollary 21, we also obtain almost sure convergence with respect to the norm of Cγ([0,T],C02δ[0,1])C^{\gamma}([0,T],C_{0}^{2\delta}[0,1]) with rate 1/nθ1/n^{\theta}.

5 A counterexample for convergence

We shall now present an example of a C0C_{0}-semigroup SS on a Banach space EE and an EE-valued Brownian motion WW such that the problem (SCP0) admits a solution with continuous trajectories whilst the associated splitting scheme fails to converge. Although the actual construction is somewhat involved, the semigroup in this example is a translation semigroup on a suitable vector-valued Lebesgue space. Such semigroups occur naturally in the context of stochastic delay equations.

We take E=Lq(0,1;p)E=L^{q}(0,1;\ell^{p}), with 1p<21\le p<2 and q2q\ge 2, and consider the EE-valued Brownian motion Wf=wfW_{f}=w\otimes f, where ww is a standard real-valued Brownian motion and fEf\in E is a fixed element. With this notation a function Ψ:(0,1)(E)\Psi:(0,1)\to\mathscr{L}(E) is stochastically integrable with respect to WfW_{f} if and only if Ψf:(0,1)E\Psi f:(0,1)\to E is is stochastically integrable with respect to ww, in which case we have

01Ψ𝑑Wf=01Ψf𝑑w.\int_{0}^{1}\Psi\,dW_{f}=\int_{0}^{1}\Psi f\,dw.

Let 1p<21\le p<2 and u>2pu>\frac{2}{p} be fixed. For k=1,2,k=1,2,\ldots and j=0,2k11j=0,\ldots 2^{k-1}-1 define the intervals Ik,j=(2j+12k,2j+12k+2uk]I_{k,j}=(\frac{2j+1}{2^{k}},\frac{2j+1}{2^{k}}+2^{-uk}]. As in particular u>1u>1, for all k=1,2,k=1,2,\dots the intervals Ik,iI_{k,i} and Ik,jI_{k,j} are disjoint for iji\neq j. Let 0<r<1p20<r<1-\frac{p}{2} and denote the basic sequence of unit vectors in p\ell^{p} by {en}n1\{e_{n}\}_{n\ge 1}. Inspired by [38, Example 3.2] we define fL(;p)f\in L^{\infty}(\mathbb{R};\ell^{p}) by

f(t):=k=1j=02k112rpk1Ik,j(t)e2k1+j.\displaystyle f(t)=\sum_{k=1}^{\infty}\sum_{j=0}^{2^{k-1}-1}2^{-\frac{r}{p}k}1_{I_{k,j}}(t)e_{2^{k-1}+j}.

Observe that f(t)=0f(t)=0 for t(0,1)t\in\mathbb{R}\setminus(0,1) and ff is well-defined: because Ik,jI_{k,j} and Ik,iI_{k,i} are disjoint for iji\neq j one has, for any t(0,1)t\in(0,1),

f(t)ppk=12rk<.\displaystyle\|f(t)\|^{p}_{\ell^{p}}\le\sum_{k=1}^{\infty}2^{-rk}<\infty.

For a given interval I=(a,b]I=(a,b], 0a<b<0\le a<b<\infty, we write ΔwI:=w(b)w(a)\Delta w_{I}:=w(b)-w(a).

Claim 24.

The function ff is stochastically integrable on (0,1)(0,1) and

01f(t)𝑑w(t)\displaystyle\int_{0}^{1}f(t)\,dw(t) =n=101f(t),enen𝑑w(t)\displaystyle=\sum_{n=1}^{\infty}\int_{0}^{1}\langle{f(t)},{e^{*}_{n}}\rangle e_{n}\,dw(t) (30)
=k=1j=02k112rpkΔwIk,je2k1+j,\displaystyle=\sum_{k=1}^{\infty}\sum_{j=0}^{2^{k-1}-1}2^{-\frac{r}{p}k}\Delta w_{I_{k,j}}e_{2^{k-1}+j},

where {en}n1\{e_{n}^{*}\}_{n\ge 1} is the basic sequence of unit vectors in p\ell^{p^{\prime}}, 1p+1p=1\frac{1}{p}+\frac{1}{p^{\prime}}=1.

Proof.

We shall deduce this from [32, Theorem 2.3, (3)(1)(3)\Rightarrow(1)]. Define the p\ell^{p}-valued Gaussian random variable

X:=k=1j=02k112rpkΔwIk,je2k1+j.X:=\sum_{k=1}^{\infty}\sum_{j=0}^{2^{k-1}-1}2^{-\frac{r}{p}k}\Delta w_{I_{k,j}}e_{2^{k-1}+j}.

This sum converges absolutely in Lp(Ω;p)L^{p}(\Omega;\ell^{p}). Indeed, let γ\gamma denote a standard Gaussian random variable. Then by Fubini’s theorem one has

𝔼k=1j=02k112rpkΔwIk,je2k1+jpp\displaystyle{\mathbb{E}}\Big{\|}\sum_{k=1}^{\infty}\sum_{j=0}^{2^{k-1}-1}2^{-\frac{r}{p}k}\Delta w_{I_{k,j}}e_{2^{k-1}+j}\Big{\|}^{p}_{\ell^{p}} =k=1j=02k112rk2u2pk𝔼|γ|p\displaystyle=\sum_{k=1}^{\infty}\sum_{j=0}^{2^{k-1}-1}2^{-rk}2^{-\frac{u}{2}pk}{\mathbb{E}}|\gamma|^{p}
=k=12k(1ru2p)1𝔼|γ|p<.\displaystyle=\sum_{k=1}^{\infty}2^{k(1-r-\frac{u}{2}p)-1}{\mathbb{E}}|\gamma|^{p}<\infty.

By the Kahane-Khintchine inequalities, the sum defining XX converges absolutely in Lq(Ω;p)L^{q}(\Omega;\ell^{p}) for all 1q<1\le q<\infty.

For any linear combination a=n=1Nanenpa^{*}=\sum_{n=1}^{N}a_{n}e^{*}_{n}\in\ell^{p^{\prime}} one easily checks that

X,a=01f(t),a𝑑w(t).\displaystyle\langle{X},{a^{*}}\rangle=\int_{0}^{1}\langle{f(t)},{a^{*}}\rangle\,dw(t).

Hence by [32, Theorem 2.3], ff is stochastically integrable and (30) holds. ∎

By similar reasoning (or an application of [32, Corollary 2.7]), for all ss\in\mathbb{R} the function tf(t+s)t\mapsto f(t+s) is stochastically integrable on (0,1)(0,1) and

01f(t+s)𝑑w(t)=n=101f(t+s),enen𝑑w(t).\int_{0}^{1}f(t+s)\,dw(t)=\sum_{n=1}^{\infty}\int_{0}^{1}\langle{f(t+s)},e_{n}^{\ast}\rangle e_{n}\,dw(t).

Let q1q\ge 1 and let {S(t)}t\{S(t)\}_{t\in\mathbb{R}} be the left-shift group on Lq(;p)L^{q}(\mathbb{R};\ell^{p}) defined by

(S(t)g)(s)\displaystyle(S(t)g)(s) =g(t+s),s,t,gLq(;p).\displaystyle=g(t+s),\quad s,t\in{\mathbb{R}},\ g\in L^{q}({\mathbb{R}};\ell^{p}).
Claim 25.

For any q1q\ge 1 the Lq(;p)L^{q}(\mathbb{R};\ell^{p})-valued function tS(t)ft\mapsto S(t)f is stochastically integrable on (0,1)(0,1) and

(01S(t)f𝑑w(t))(s)\displaystyle\Big{(}\int_{0}^{1}S(t)f\,dw(t)\Big{)}(s) =01f(t+s)𝑑w(t)\displaystyle=\int_{0}^{1}f(t+s)\,dw(t)

for almost all ss\in\mathbb{R} almost surely.

Proof.

For s(1,1)s\not\in(-1,1) the function tf(t+s)t\mapsto f(t+s) is identically 0 on (0,1)(0,1), and for s(1,1)s\in(-1,1) we have

𝔼01f(t+s)𝑑w(t)pq𝔼01f(t)𝑑w(t)pq.{\mathbb{E}}\Big{\|}\int_{0}^{1}f(t+s)\,dw(t)\Big{\|}_{\ell^{p}}^{q}\le{\mathbb{E}}\Big{\|}\int_{0}^{1}f(t)\,dw(t)\Big{\|}_{\ell^{p}}^{q}.

As a consequence, Lq(Ω;p)L^{q}(\Omega;\ell^{p})-valued function s01f(s+t)𝑑w(t)s\mapsto\int_{0}^{1}f(s+t)\,dw(t) defines an element of Lq(;Lq(Ω;p))L^{q}({\mathbb{R}};L^{q}(\Omega;\ell^{p})). Under the natural isometry Lq(;Lq(Ω;p))Lq(Ω;Lq(;p))L^{q}({\mathbb{R}};L^{q}(\Omega;\ell^{p}))\simeq L^{q}(\Omega;L^{q}({\mathbb{R}};\ell^{p})) we may identify this function with an element YLq(Ω;Lq(;p))Y\in L^{q}(\Omega;L^{q}({\mathbb{R}};\ell^{p})). To establish the claim, with an appeal to [32, Theorem 2.3] it suffices to check that for all apa^{*}\in\ell^{p^{\prime}} and Borel sets A()A\in\mathscr{B}(\mathbb{R}) we have

01S(t)f,1Aa𝑑w(t)=Y,1Aa.\int_{0}^{1}\langle{S(t)f},1_{A}\otimes{a^{*}}\rangle\,dw(t)=\langle Y,1_{A}\otimes a^{\ast}\rangle.

By writing out both sides, this identity is seen to be an immediate consequence of the stochastic Fubini theorem (see, e.g., [32, Theorem 3.3]). ∎

In the same way one sees that for t0t\ge 0 the stochastic integrals 0tS(s)f𝑑w(s)\int_{0}^{t}S(-s)f\,dw(s) are well-defined. Because the process t0tS(s)f𝑑w(s)t\mapsto\int_{0}^{t}S(-s)f\,dw(s) is a martingale having a continuous version by Doob’s maximal inequality, we also know that the convolution process

U(t):=0tS(ts)f𝑑w(s)=S(t)0tS(s)f𝑑w(s)U(t):=\int_{0}^{t}S(t-s)f\,dw(s)=S(t)\int_{0}^{t}S(-s)f\,dw(s)

has a continuous version. However, as we shall see, the splitting scheme for UU fails to converge.

For n1n\ge 1 define

S(n)\displaystyle S^{(n)} :=k=1n1(k1n,kn]S(kn).\displaystyle=\sum_{k=1}^{n}1_{(\frac{k-1}{n},\frac{k}{n}]}\otimes S\Big{(}\frac{k}{n}\Big{)}.

Observe that for any s,ts,t\in\mathbb{R}

(S(n)(t)f)(s)\displaystyle(S^{(n)}(t)f)(s) =k=1n1(k1n,kn](t)f(kn+s).\displaystyle=\sum_{k=1}^{n}1_{(\frac{k-1}{n},\frac{k}{n}]}(t)f\Big{(}\frac{k}{n}+s\Big{)}. (31)

Similarly to the above one checks that

(01S(n)(t)f𝑑w(t))(s)\displaystyle\Big{(}\int_{0}^{1}S^{(n)}(t)f\,dw(t)\Big{)}(s) =k=1nf(kn+s)[w(kn)w(k1n)]\displaystyle=\sum_{k=1}^{n}f\Big{(}\frac{k}{n}+s\Big{)}\Big{[}w\Big{(}\frac{k}{n}\Big{)}-w\Big{(}\frac{k-1}{n}\Big{)}\Big{]}

for almost all ss\in{\mathbb{R}} almost surely.

The clue to this example is that for nn fixed and s(0,2un]s\in(0,2^{-un}] the function t(S(2n)(t)f)(s)t\mapsto(S^{(2^{n})}(t)f)(s) always ‘picks up’ the values of ff at the left parts of the dyadic intervals where ff is defined to be non-zero. Thus for these values of ss the function t(S(2n)(t)f)(s)t\mapsto(S^{(2^{n})}(t)f)(s) it is nowhere zero and its stochastic integral blows up as nn\rightarrow\infty. We shall make this precise. Our aim is to prove that for certain values of q>2q>2 (to be determined later on) one has

𝔼01S(2n)(t)f𝑑w(t)Lq(;p)pas n.\displaystyle{\mathbb{E}}\Big{\|}\int_{0}^{1}S^{(2^{n})}(t)f\,dw(t)\Big{\|}^{p}_{L^{q}(\mathbb{R};\ell^{p})}\rightarrow\infty\quad\textrm{as }n\rightarrow\infty. (32)

By Minkowski’s inequality we have, for any n1n\ge 1 and qpq\ge p,

[𝔼01S(2n)(t)f𝑑w(t)Lq(;p)p]1p\displaystyle\Big{[}{\mathbb{E}}\Big{\|}\int_{0}^{1}S^{(2^{n})}(t)f\,dw(t)\Big{\|}^{p}_{L^{q}(\mathbb{R};\ell^{p})}\Big{]}^{\frac{1}{p}}
[(𝔼(01S(2n)(t)f𝑑w(t))(s)pp)qp𝑑s]1q.\displaystyle\qquad\ge\Big{[}\int_{\mathbb{R}}\Big{(}{\mathbb{E}}\Big{\|}\Big{(}\int_{0}^{1}S^{(2^{n})}(t)f\,dw(t)\Big{)}(s)\Big{\|}_{\ell^{p}}^{p}\Big{)}^{\frac{q}{p}}\,ds\Big{]}^{\frac{1}{q}}.

Now fix n1n\ge 1. For any 1kn1\le k\le n and any j=0,,2k11j=0,\ldots,2^{k-1}-1 there exists a unique 1ik,j2n11\le i_{k,j}\le 2^{n}-1 such that ik,j2n=2j+12k\frac{i_{k,j}}{2^{n}}=\frac{2j+1}{2^{k}}. Now observe that by definition of ff one has for s(0,2un]s\in(0,2^{-un}] that

f(ik,j2n+s),e2k1+j=2krp.\displaystyle\Big{\langle}f\big{(}\frac{i_{k,j}}{2^{n}}+s\big{)},e^{*}_{2^{k-1}+j}\Big{\rangle}=2^{-k\frac{r}{p}}.

Using this and representation (31) one obtains that for s(0,2un]s\in(0,2^{-un}], 1kn1\le k\le n, j=0,,2k11j=0,\ldots,2^{k-1}-1, and any t(ik,j12n,ik,j2n]=:Ik,jnt\in\big{(}\frac{i_{k,j}-1}{2^{n}},\frac{i_{k,j}}{2^{n}}\big{]}=:I_{k,j}^{n},

(S(2n)(t)f)(s),e2k1+j=2krp.\displaystyle\big{\langle}{(S^{(2^{n})}(t)f)(s)},e^{*}_{2^{k-1}+j}\rangle=2^{-k\frac{r}{p}}. (33)

To prove (32), we now estimate

(𝔼(01S(2n)(t)f𝑑w(t))(s)pp)qp𝑑s\displaystyle\int_{\mathbb{R}}\Big{(}{\mathbb{E}}\Big{\|}\Big{(}\int_{0}^{1}S^{(2^{n})}(t)f\,dw(t)\Big{)}(s)\Big{\|}_{\ell^{p}}^{p}\Big{)}^{\frac{q}{p}}\,ds
02un(k=1nj=02k11𝔼|01(S(2n)(t)f)(s),e2k1+j𝑑w(t)|p)qp𝑑s\displaystyle\qquad\ge\int_{0}^{2^{-un}}\Big{(}\sum_{k=1}^{n}\sum_{j=0}^{2^{k-1}-1}{\mathbb{E}}\Big{|}\int_{0}^{1}\big{\langle}{(S^{(2^{n})}(t)f)(s)},{e^{*}_{2^{k-1}+j}}\big{\rangle}\,dw(t)\Big{|}^{p}\Big{)}^{\frac{q}{p}}\,ds
02un(k=1nj=02k112kr𝔼|ΔwIk,jn|p)qp𝑑s\displaystyle\qquad\ge\int_{0}^{2^{-un}}\Big{(}\sum_{k=1}^{n}\sum_{j=0}^{2^{k-1}-1}2^{-kr}{\mathbb{E}}|\Delta w_{I_{k,j}^{n}}|^{p}\Big{)}^{\frac{q}{p}}\,ds
=02un(k=1n2k12kr2np2𝔼|γ|p)qp𝑑s\displaystyle\qquad=\int_{0}^{2^{-un}}\Big{(}\sum_{k=1}^{n}2^{k-1}2^{-kr}2^{-n\frac{p}{2}}{\mathbb{E}}|\gamma|^{p}\Big{)}^{\frac{q}{p}}\,ds
2un12n(1rp2)qp(𝔼|γ|p)qp,\displaystyle\qquad\ge 2^{-un-1}2^{n(1-r-\frac{p}{2})\frac{q}{p}}({\mathbb{E}}|\gamma|^{p})^{\frac{q}{p}},

where in the second inequality we use (33) and γ\gamma denotes a standard Gaussian random variable. Thus if u+(1rp2)qp>0-u+(1-r-\frac{p}{2})\frac{q}{p}>0, that is, if q>up/(1rp2)q>up/(1-r-\frac{p}{2}) (recall that r<1p2r<1-\frac{p}{2}), the left-hand side expression diverges as nn\rightarrow\infty.

Acknowledgments

We thank Ben Goldys, Arnulf Jentzen, Markus Kunze, and Mark Veraar for helpful comments and for suggesting several improvements. We also thank the anonymous referee for pointing out various references.

References

  • [1] F. Albiac and N.J. Kalton. “Topics in Banach Space Theory”, volume 233 of Graduate Texts in Mathematics. Springer, New York, 2006.
  • [2] A. Bensoussan, R. Glowinski, and A. Răşcanu. Approximation of the Zakai equation by the splitting up method. SIAM J. Control Optim., 28(6):1420–1431, 1990.
  • [3] A. Bensoussan, R. Glowinski, and A. Răşcanu. Approximation of some stochastic differential equations by the splitting up method. Appl. Math. Optim., 25(1):81–106, 1992.
  • [4] Z. Brzeźniak. Stochastic partial differential equations in M-type 22 Banach spaces. Potential Anal., 4(1):1–45, 1995.
  • [5] Z. Brzeźniak. On stochastic convolution in Banach spaces and applications. Stochastics Stochastics Rep., 61(3-4):245–295, 1997.
  • [6] P. Clément, B. de Pagter, F.A. Sukochev, and H. Witvliet. Schauder decompositions and multiplier theorems. Studia Math., 138(2):135–163, 2000.
  • [7] G. Da Prato, S. Kwapień, and J. Zabczyk. Regularity of solutions of linear stochastic equations in Hilbert spaces. Stochastics, 23(1):1–23, 1987.
  • [8] A.M. Davie and J.G. Gaines. Convergence of numerical schemes for the solution of parabolic stochastic partial differential equations. Math. Comp., 70(233):121–134 (electronic), 2001.
  • [9] R. Denk, M. Hieber, and J. Prüss. RR-boundedness, Fourier multipliers and problems of elliptic and parabolic type. Mem. Amer. Math. Soc., 166(788), 2003.
  • [10] J. Dettweiler, J.M.A.M. van Neerven, and L. Weis. Space-time regularity of solutions of the parabolic stochastic Cauchy problem. Stoch. Anal. Appl., 24(4):843–869, 2006.
  • [11] D. Feyel and A. de La Pradelle. On fractional Brownian processes. Potential Anal., 10(3):273–288, 1999.
  • [12] P. Florchinger and F. Le Gland. Time-discretization of the Zakai equation for diffusion processes observed in correlated noise. Stochastics Stochastics Rep., 35(4):233–256, 1991.
  • [13] I. Gyöngy. Lattice approximations for stochastic quasi-linear parabolic partial differential equations driven by space-time white noise. II. Potential Anal., 11(1):1–37, 1999.
  • [14] I. Gyöngy and N. Krylov. On the splitting-up method and stochastic partial differential equations. Ann. Probab., 31(2):564–591, 2003.
  • [15] I. Gyöngy and A. Millet. Rate of convergence of space time approximations for stochastic evolution equations. Potential Anal., 30(1):29–64, 2009.
  • [16] B.H. Haak and J.M.A.M. van Neerven. Uniformly γ\gamma-radonifying families of operators and the linear stochastic Cauchy problem in Banach spaces. ArXiv:math/0611724.
  • [17] J. Hoffmann-Jørgensen. Sums of independent Banach space valued random variables. Studia Math., 52:159–186, 1974.
  • [18] A. Jentzen. Pathwise numerical approximation of SPDEs with additive noise under non-global Lipschitz coefficients. Potential Anal., 31(4):375–404, 2009.
  • [19] A. Jentzen and P.E. Kloeden. The numerical approximation of stochastic partial differential equations. Milan Journal of Mathematics, 7(1):205–244, 2009.
  • [20] N.J. Kalton and L. Weis. The HH^{\infty}-calculus and square function estimates. In preparation.
  • [21] F. Kühnemund and J.M.A.M. van Neerven. A Lie-Trotter product formula for Ornstein-Uhlenbeck semigroups in infinite dimensions. J. Evol. Equ., 4(1):53–73, 2004.
  • [22] P.C. Kunstmann and L. Weis. Maximal LpL_{p}-regularity for parabolic equations, Fourier multiplier theorems and HH^{\infty}-functional calculus. In In: “Functional Analytic Methods for Evolution Equations”, volume 1855 of Lecture Notes in Math., pages 65–311. Springer, Berlin, 2004.
  • [23] S. Kwapień. On Banach spaces containing c0c_{0}. volume 52, pages 187–188. 1974. A supplement to the paper by J. Hoffmann-Jørgensen: “Sums of independent Banach space valued random variables” (Studia Math. 52 (1974), 159–186).
  • [24] G.J. Lord and J. Rougemont. A numerical scheme for stochastic PDEs with Gevrey regularity. IMA J. Numer. Anal., 24(4):587–604, 2004.
  • [25] A. Lunardi. “Analytic Semigroups and Optimal Regularity in Parabolic Problems”. volume 16 of Progr. Nonlin. Diff. Eq. Appl., Birkhäuser Verlag, Basel, 1995.
  • [26] J. Maas and J.M.A.M. van Neerven. Boundedness of Riesz transforms for elliptic operators on abstract Wiener spaces. J. Funct. Anal., 257(8):2410–2475, 2009.
  • [27] N. Nagase. Remarks on nonlinear stochastic partial differential equations: an application of the splitting-up method. SIAM J. Control Optim., 33(6):1716–1730, 1995.
  • [28] J.M.A.M. van Neerven. γ\gamma-Radonifying operators – a survey. To appear in Proc. CMA. ArXiv:math/0911.3788.
  • [29] J.M.A.M. van Neerven. “Stochastic Evolution Equations”. Lecture Notes of the 11th Internet Seminar, TU Delft, OpenCourseWare, http://ocw.tudelft.nl, 2008.
  • [30] J.M.A.M. van Neerven, M.C. Veraar, and L. Weis. Stochastic integration in UMD Banach spaces. Annals Probab., 35:1438–1478, 2007.
  • [31] J.M.A.M. van Neerven, M.C. Veraar, and L. Weis. Stochastic evolution equations in UMD Banach spaces. J. Funct. Anal., 255(4):940–993, 2008.
  • [32] J.M.A.M. van Neerven and L. Weis. Stochastic integration of functions with values in a Banach space. Studia Math., 166(2):131–170, 2005.
  • [33] J.M.A.M. van Neerven and L. Weis. Weak limits and integrals of Gaussian covariances in Banach spaces. Probab. Math. Statist., 25(1):55–74, 2005.
  • [34] A.L. Neidhardt. “Stochastic Integrals in 22-Uniformly Smooth Banach Spaces”. PhD thesis, University of Wisconsin, 1978.
  • [35] A. Pazy. “Semigroups of Linear Operators and Applications to Partial Differential Equations”, volume 44 of Applied Mathematical Sciences. Springer-Verlag, New York, 1983.
  • [36] R. Pettersson and M. Signahl. Numerical approximation for a white noise driven SPDE with locally bounded drift. Potential Anal., 22(4):375–393, 2005.
  • [37] R.S. Phillips. The adjoint semi-group. Pacific J. Math., 5:269–283, 1955.
  • [38] J. Rosiński and Z. Suchanecki. On the space of vector-valued functions integrable with respect to the white noise. Colloq. Math., 43(1):183–201 (1981), 1980.
  • [39] H. Triebel. “Interpolation Theory, Function Spaces, Differential Operators”, volume 18 of North-Holland Mathematical Library. North-Holland Publishing Co., Amsterdam, 1978.
  • [40] L. Weis. Operator-valued Fourier multiplier theorems and maximal LpL_{p}-regularity. Math. Ann., 319(4):735–758, 2001.