Convergence rates of the splitting scheme for parabolic linear stochastic Cauchy problems
Abstract
We study the splitting scheme associated with the linear stochastic Cauchy problem
Here is the generator of an analytic -semigroup on a Banach space and is a Brownian motion with values in a fractional domain space associated with . We prove that if are such that and , then the approximate solutions converge to the solution in the space , both in -means and almost surely, with rate .
keywords:
Splitting scheme, stochastic evolution equations, analytic semigroups, -radonifying operators, -boundedness, stochastic convolutions, Lie-Trotter product formula.AMS:
Primary: 35R15, 60H15; Secondary: 47D06, 60J351 Introduction
We are concerned with the convergence of the splitting scheme for the stochastic linear Cauchy problem
(SCPx) |
were is the generator of a -semigroup on a real Banach space , is an -valued Brownian motion on a probability space , and 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 and run the semigroup on a time interval of equal length. Taking time steps and writing and , this generates a finite sequence defined by
We have the explicit formula
Assuming the existence of a unique solution of the problem (SCPx) (see Proposition 7 below), we may ask for conditions ensuring the convergence of to in for some (all) or even in almost surely. In order to describe our approach we start by noting that each can be represented as a stochastic integral of the discretised function
where and for . Indeed, defining the stochastic integral of a step function in the obvious way, we have
(1) |
On the other hand, the exact solution of (SCPx), if it exists, is given by the stochastic convolution integral
(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
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 . From we see that
for all , and therefore it suffices to analyse convergence of the splitting scheme with initial value . In what follows, in order to simplify notations we shall write and .
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
(SCP0) |
admits a unique solution which satisfies
for all :
-
(a)
has type ;
-
(b)
restricts to a -semigroup on the reproducing kernel Hilbert space associated with .
The class of spaces satisfying condition (a) includes all Hilbert spaces and the spaces for . 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 is analytic on . The convergence is considered in suitable Hölder norms in space and time, with explicit bounds for the convergence rate.
We denote by the fractional power space of exponent associated with (see Section 4 for more details).
Theorem 2.
Suppose that the semigroup is analytic on and that is a Brownian motion in for some . Then the problem (SCP0) admits a unique solution , and for all such that and one has the estimate
with implied constant independent of .
By a Borel-Cantelli argument, this result implies the almost sure convergence of to in with the same rates.
The proof of Theorem 2 heavily relies on the theory of -radonifying operators and -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 , such as martingale type 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 .
Example 3.
Theorem 2 may be applied to second order elliptic operators of the form
Under minor regularity assumptions on the coefficients , and , such operators generate analytic semigroups on with (see [25, Chapter 3]) and one has for all . Applying Theorem 2 (with ), we obtain convergence of the splitting scheme in the space for any such that . By the Sobolev embedding [39, Section 2.8], this implies the convergence of the splitting scheme in the mixed Hölder space . As a consequence, we obtain convergence in the mixed Hölder space
with rate for any ; this rate improves when the noise is more regular. Similar results can be obtained for elliptic operators on smooth domains subject to various types of boundary conditions (as long as they generate an analytic semigroup on ).
For semi-linear (Stratonovich type) SPDEs governed by second order elliptic operators on and driven by multiplicative noise, convergence in 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 , for finite-dimensional noise and with sufficiently smooth coefficients one obtains the maximal estimate
Our result is valid in the full scale of spaces and infinite-dimensional noise, with a rate which (for smooth enough noise) is only slightly worse that and is independent of . More precisely, for and taking we obtain uniform convergence with rate for any . 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 is a Brownian motion can be weakened to being a cylindrical Brownian motion (see, e.g., [30, 32] for the definition).
Example 4.
The stochastic heat equation on the unit interval with Dirichlet boundary conditions driven by space-time white noise can be put into the present framework by taking for the extrapolation space with and . As we shall explain in Example 22, this entails the convergence of the splitting scheme in the mixed Hölder space
with rate for any .
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 , , cannot have a convergence rate better than . This shows that the exponent 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 in 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 -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 has continuous trajectories. In the final Section 5 we present an example which shows that without any additional assumptions on the space and/or the semigroup the splitting scheme may fail to converge even if a solution with continuous trajectories exists.
2 Preliminaries
Let be a sequence of independent standard Gaussian random variables on a probability space , let be a real Hilbert space (later we shall take , where is another real Hilbert space) and a real Banach space. A bounded operator from to is called -summing if
is finite, where the supremum is taken over all finite orthonormal systems in . It can be shown that is indeed a norm which turns the space of -summing operators into a Banach space.
The space of -radonifying operators is defined to be the closure of the finite rank operators under the norm ; it is a closed subspace of . A celebrated result of Kwapień and Hoffmann-Jørgensen [17, 23] implies that if does not contain a closed subspace isomorphic to then .
Since convergence in implies convergence in , every operator , being the operator norm limit of a sequence of finite rank operators from to , is compact.
If is separable with orthonormal basis , then an operator is -radonifying if and only if the Gaussian sum converges in , and in this situation we have
The general case may be reduced to the separable case by observing that for any there exists a separable closed subspace of such that vanishes on the orthogonal complement .
If is given and is an orthonormal basis for , the sum defines a centred -valued Gaussian random variable. Its distribution is a centred Gaussian Radon measure on whose covariance operator equals . We will refer to as the Gaussian measure associated with . In the reverse direction, if is a centred -valued Gaussian random variable with reproducing kernel Hilbert space , then is separable, the natural inclusion mapping is -radonifying, and we have
Below we shall need the following simple continuity result.
Proposition 5.
Let be a metric space and let be strongly continuous. Then for all the function ,
is continuous.
Proof.
Suppose first that is a finite rank operator, say with orthonormal and a sequence in . Suppose that in . Then
The general case follows from the density of the finite rank operators in and the norm estimate . ∎
3 Proof of Theorem 1
We start with a brief discussion of stochastic integrals of operator-valued functions. Let be a Hilbert space and fix . An -cylindrical Brownian motion, indexed by and defined on a probability space , is a mapping with the following properties:
-
•
for all the random variable is Gaussian;
-
•
for all we have .
Formally, an -cylindrical Brownian motion can be thought of as a standard Brownian motion taking values in the Hilbert space . One easily checks that is linear and that for all the random variables are jointly Gaussian. These random variables are independent if and only if are orthogonal in . For further details see [28, Section 3].
A finite rank step function is function of the form where each operator is of finite rank. The stochastic integral with respect to of such a function is defined by setting
and this definition is extended by linearity. A function is said to be stochastically integrable with respect to if there exists a sequence of finite rank step functions such that:
-
(i)
for all we have in measure on ;
-
(ii)
the limit exists in probability.
In this situation we write
and call the stochastic integral of with respect to .
As was shown in [32], for finite rank step functions one has the isometry
(3) |
where is the bounded operator represented by , i.e.,
(4) |
As a consequence, a function is stochastically integrable on with respect to if and only if for all and there exists an operator such that
The isometry (3) extends to this situation. The following simple observation [10, Lemma 2.1] will be used frequently:
Proposition 6.
For all and the function belongs to and we have
For the remainder of this section we fix an -valued Brownian motion and . Let be the reproducing kernel Hilbert space associated with the Gaussian random variable and let be the natural inclusion mapping. Then induces an -cylindrical Brownian motion by putting
(5) |
This motivates us to call a function stochastically integrable with respect to if the function is stochastically integrable with respect to , in which case we put
It is easy to check that for all the indicator function is stochastically integrable with respect to and
This shows that the definition is consistent with (1) and (2).
Now let denote a -semigroup of bounded linear operators on , with generator . We will be interested in the case where the function to be integrated against is one of the following:
We may define bounded operators and from to by the formula (4). Being associated with -valued step functions, the operators belong to by Proposition 6. Concerning the question whether the operator is in we have the following result [32, Theorem 7.1].
Proposition 7.
Let . The following assertions are equivalent:
-
(i)
the operator belongs to ;
-
(ii)
the function is stochastically integrable on with respect to ;
-
(iii)
for some (all) the problem (SCPx) admits a unique solution .
In this situation, for all and we have
almost surely.
In [32] an example is presented showing even for rank one Brownian motions in the equivalent conditions need not always be satisfied for all -semigroups on . The conditions are satisfied, however, if one of the following additional conditions holds:
-
(a)
is a type 2 Banach space,
-
(b)
restricts to a -semigroup on ,
-
(c)
is an analytic -semigroup on .
We are now in a position to state the main result of this section. We use the notations introduced above, and let and denote the Gaussian measures on associated with the operators and , respectively.
Theorem 8.
Suppose that the equivalent conditions of Proposition 7 are satisfied. The following assertions are equivalent:
-
1.
in for some (all) and some (all) ;
-
2.
in ;
-
3.
weakly.
In this situation we have in for all , , and , and in fact we have
where, as before, and correspond to the initial value .
Proof.
We begin by proving the equivalence of (1), (2), (3). Clearly it suffices to consider the initial value .
For a given , a sequence of -valued centred Gaussian random variables converges in if and only if it converges in probability in . Therefore, if (1) holds for some , then it holds for all .
Taking in (1) the equivalence (1)(2) follows from the identity (3) and the representations (1) and (2).
Next we claim that in for all . Once we have shown this, the equivalence (2)(3) follows from [16, Theorem 3.1] (or by using the argument of [34, page 18ff]). To prove the claim we fix and note that in we have
The inclusion mapping is -radonifying and hence compact. As a consequence, the weak∗-continuity of implies that is continuous on . It follows that in , and hence in .
The assertions (1), (2), (3) are equivalent to the validity of a Lie–Trotter product formula for the Ornstein-Uhlenbeck semigroup associated with the problem (SCPx), which is defined on the space of all bounded real-valued continuous functions on by the formula
where is the solution of (SCPx). In order to explain the precise result, let us denote by and the semigroups on corresponding to the drift term and the diffusion term in (SCPx). Thus,
Each of the semigroups , and is jointly continuous in and , uniformly on for all compact sets . It was shown in [21] that if condition (3) of Theorem 8 holds, then for all we have the Lie–Trotter product formula
(6) |
with convergence uniformly on for all compact sets . Conversely it follows from the proof of this result that (6) with 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:
-
(a)
is isomorphic to a Hilbert space;
-
(b)
restricts to a -semigroup on .
Thus, either of these conditions implies the convergence in for all and 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 -radonifying operators from [30], respectively. Moreover, case (a) is extended to Banach spaces with type . Recall that a Banach space is said to have type if there exists a constant such that for all finite choices we have
Hilbert spaces have type and -spaces () have type . We refer to [1] for more information.
Theorem 9.
Proof.
Theorem 10.
4 Proof of Theorem 2
In this section we shall prove convergence of the splitting scheme under the assumption that the -semigroup generated by is analytic; no assumptions on the space 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 –valued functions define operators belonging to . We refer to [28, Section 13] for a detailed proof.
Proposition 11.
Let be continuously differentiable with
Define by
Then and
For and large enough we define
which is known to be independent of the choice of . It is a Banach space with respect to the norm . This norm depends of course on , but any two such norms are mutually equivalent. In what follows we consider to be fixed.
We shall also need the extrapolation spaces , defined for as the closure of with respect to the norm . It follows readily from the definitions that for any two the operator defines an isomorphism from onto .
In the next two remarks we fix and , and suppose that is an analytic -semigroup on with generator .
Remark 12.
By [35, Theorem 2.6.13(c)] one has, for any ,
(7) |
with implied constant independent of . From this and the ideal property for -radonifying operators we obtain the following estimate for :
where for ; the implied constant is independent of and . If , it then follows from Proposition 11 that
with implied constant independent of and . In particular, taking we see that the equivalent conditions of Proposition 7 hold.
Remark 13.
Suppose that . Identifying operator-valued functions with the integral operators they induce, we have
Applying Proposition 11 to both terms on the right-hand side, if it follows that
for all .
We need to introduce the following terminology. Let and be Banach spaces. A family of operators is called -bounded if there exists a finite constant such that for all finite choices and vectors we have
The least admissible constant is called the -bound of , notation . We refer to [6, 9, 22, 40] for examples and more information. In these references the related notion of -boundedness is discussed; this notion is obtained by replacing the Gaussian random variables by Rademacher variables in the above definition. Any -bounded set is also -bounded, and the two notions are equivalent if 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 and are Banach spaces and is a strongly measurable function (in the sense that is strongly measurable for every ) with -bounded range . Then for every finite rank simple function the operator belongs to and
As a result, the map has a unique extension to a bounded operator
of norm .
In the applications of this result below it will usually be possible to check that actually we have .
We will also need the following sufficient condition for -boundedness, which is a variation of a result of Weis [40, Proposition 2.5].
Proposition 15.
Let and be Banach spaces, and let be an function such that for all the function is continuously differentiable with integrable derivative. Then the set is -bounded in and
Here is a simple application:
Lemma 16.
Let the -semigroup be analytic on .
-
(1)
For all and the set is -bounded in and we have
with implied constant independent of .
-
(2)
For all the set is -bounded in and we have
with implied constant independent of .
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 the set
is -bounded in with -bound . From this we deduce that is -bounded in with -bound . In view of the identity
and noting that , this will prove the assertion of the lemma.
We shall again write and for the solution of (SCP0) and its approximations by the splitting scheme.
Theorem 17.
Assume that the semigroup is analytic on and that is a Brownian motion in for some . Then the equivalent conditions of Proposition 7 and Theorem 8 hold. Moreover, for all and such that , and all we have
(8) |
with implied constant independent of and . As a consequence, for all the solution of (SCP0) satisfies
(9) |
with implied constant independent of and .
Proof.
By rescaling time we may assume that . Let be as indicated. We begin by noting that the embedding associated with belongs to .
Pick . Note that for we have and , so one can write, for all ,
(10) |
Fix . By the first part of Lemma 16 the set
is -bounded in (hence in , hence in , with the same upper bounds for the -bounds, because commutes with the fractional powers of ) and we have
(11) |
By the second part of Lemma 16 the set
is -bounded in (and hence in , with the same estimate for the -boundedness constant), and we have
(12) |
Using (10), Remark 13, Proposition 14, the identity
(13) |
together with the estimates (11), and (12), and noting that , one obtains
∎
Remark 18.
The condition implies, in view of the restriction , that . For , Theorem 17 gives a rate of convergence of order , whereas for we obtain the rate for any . For one can in fact obtain a slightly better rate at the final time , namely . More precisely, for we have
(14) |
with constants independent of .
Once again observe that by scaling we may (and do) assume that . In order to prove (14) we first give an estimate for a given time interval where . In that case, for one has
(15) |
with implied constant independent of and . 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 near in (13). The details are as follows. Fix and and pick an arbitrary . Then,
(16) |
with implied constant independent of and ; the last inequality uses that . As in the proof of Theorem 17 the set is -bounded in , with -bound
(17) |
Finally, since , as in the proof of Theorem 17 the set is -bounded in with
(18) |
Combining (16), (17), and (18) we obtain
Returning to the proof of estimate (14) we fix an integer . Because one can pick such that . For define . Note that and . If in (15) we take and we obtain the estimate
where the last inequality used that . Set , so that . Using this estimate for , from Theorem 17 we obtain, for any choice of (which then satisfies ),
Combining the above one gets
This gives the estimate (14).
Under the assumptions that is analytic on and is a Brownian motion on , the solution of (SCP0) has a version with trajectories in for any such that [10]. The main result of this paper asserts that also the approximating processes have trajectories in and that the splitting scheme converges with respect to the -norm, with a convergence rate depending on and and the smoothness of the noise.
Theorem 19.
Let be analytic on and suppose that is a Brownian motion in for some . If satisfy and , then for all the solution of (SCP0) satisfies
with implied constant independent of .
Proof.
By scaling we may assume . Put . Let and be as indicated. Without loss of generality we assume that . The main step in the proof is the following claim.
Claim 20.
There exists a constant such that for all , all satisfying we have
Proof.
Fix and such that . Clearly,
(19) | ||||
For the first term we note that by (3) (and the remark following it) and (8) one has
(20) | ||||
The estimate for the second term is extracted from arguments in [31]; see also [29, Theorem 10.19]. Fix such that . Then the set is -bounded in (hence in , hence in ) by the first part of Lemma 16, and therefore
(21) | ||||
To estimate the third term on the right-hand side of (19), we first define sets and by
Both equalities follow from the identity for . By definition of and one has
(22) | ||||
noting that the integrand of the integral over vanishes.
Set . To estimate the right-hand side, observe that from we may pick such that . Using the identity and applying Proposition 14 and part (1) of Lemma 16, and then using the estimate and Proposition 6, we obtain
(23) | ||||
In order to estimate the -norm of the function we note that , where
From this it is easy to see that and that for one has
(the latter inequality following from ), and therefore
As a consequence,
(24) |
Combining the estimates (23) and (24) and estimating the non-negative powers of by we find
(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 we have
Hence if one has
(26) |
The random variables being Gaussian, from the claim and (26) combined with the Kahane-Khintchine inequalities we deduce that for all and one has
(27) |
Now fix any and take . Then by (27) and the Kolmogorov-Chentsov criterion with -moments (see [11, Theorem 5]),
This inequality shows that for all we have
for all sufficiently large . It is clear that once we know this, this inequality extends to all values . This completes the proof of the theorem (with instead of , which obviously suffices). ∎
Corollary 21.
Suppose that is analytic on and that is a Brownian motion in for some . Let satisfy and . Then for almost all there exists a constant such that the solution of (SCP0) satisfies
Proof.
Set
Pick in such a way that and let be so large that . By Theorem 19, applied with instead of , and Chebyshev’s inequality,
with constant independent of . By the choice of we have , and therefore by the Borel-Cantelli lemma
For the belonging to this set we have
∎
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 :
(28) |
Following the approach of [10] we put and , where the exponent is to be chosen later on. In order to formulate the problem (28) as an abstract stochastic evolution equation of the form
(29) |
where is a Brownian motion with values in a suitable Banach space , we fix an arbitrary real number , to be chosen in a moment, and let denote the extrapolation space of order associated with the Dirichlet Laplacian in . It is shown in [10] (see also [5, Lemma 6.5]) that the identity operator on extends to a -radonifying embedding from into . As a result, the -cylindrical Brownian motion canonically associated with (see (5)) may be identified with a Brownian motion in . Furthermore the extrapolated Dirichlet Laplacian, henceforth denoted by , satisfies the assumptions of Theorem 19 in .
Let be the solution of (29) in . By definition, we shall regard as the solution of (28). Suppose now that we are given real numbers satisfy
This ensures that one can choose and in such a way that and . By Theorem 19 (with ), for all the splitting scheme associated with problem (29) satisfies
Putting we have , and this space embeds into since .
Choose so large that . We have
with equivalent norms. By the Sobolev embedding theorem,
with continuous inclusion. Here is the space of all Hölder continuous functions of exponent . We denote . Putting things together we obtain a continuous inclusion
We have proved the following theorem (cf. Example 4).
Theorem 23.
For all the stochastic heat equation (28) admits a solution in , and for all satisfying we have
By Corollary 21, we also obtain almost sure convergence with respect to the norm of with rate .
5 A counterexample for convergence
We shall now present an example of a -semigroup on a Banach space and an -valued Brownian motion 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 , with and , and consider the -valued Brownian motion , where is a standard real-valued Brownian motion and is a fixed element. With this notation a function is stochastically integrable with respect to if and only if is is stochastically integrable with respect to , in which case we have
Let and be fixed. For and define the intervals . As in particular , for all the intervals and are disjoint for . Let and denote the basic sequence of unit vectors in by . Inspired by [38, Example 3.2] we define by
Observe that for and is well-defined: because and are disjoint for one has, for any ,
For a given interval , , we write .
Claim 24.
The function is stochastically integrable on and
(30) | ||||
where is the basic sequence of unit vectors in , .
Proof.
We shall deduce this from [32, Theorem 2.3, ]. Define the -valued Gaussian random variable
This sum converges absolutely in . Indeed, let denote a standard Gaussian random variable. Then by Fubini’s theorem one has
By the Kahane-Khintchine inequalities, the sum defining converges absolutely in for all .
By similar reasoning (or an application of [32, Corollary 2.7]), for all the function is stochastically integrable on and
Let and let be the left-shift group on defined by
Claim 25.
For any the -valued function is stochastically integrable on and
for almost all almost surely.
Proof.
For the function is identically on , and for we have
As a consequence, -valued function defines an element of . Under the natural isometry we may identify this function with an element . To establish the claim, with an appeal to [32, Theorem 2.3] it suffices to check that for all and Borel sets we have
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 the stochastic integrals are well-defined. Because the process is a martingale having a continuous version by Doob’s maximal inequality, we also know that the convolution process
has a continuous version. However, as we shall see, the splitting scheme for fails to converge.
For define
Observe that for any
(31) |
Similarly to the above one checks that
for almost all almost surely.
The clue to this example is that for fixed and the function always ‘picks up’ the values of at the left parts of the dyadic intervals where is defined to be non-zero. Thus for these values of the function it is nowhere zero and its stochastic integral blows up as . We shall make this precise. Our aim is to prove that for certain values of (to be determined later on) one has
(32) |
By Minkowski’s inequality we have, for any and ,
Now fix . For any and any there exists a unique such that . Now observe that by definition of one has for that
Using this and representation (31) one obtains that for , , , and any ,
(33) |
To prove (32), we now estimate
where in the second inequality we use (33) and denotes a standard Gaussian random variable. Thus if , that is, if (recall that ), the left-hand side expression diverges as .
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 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. -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 -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 -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 -regularity for parabolic equations, Fourier multiplier theorems and -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 . 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. -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 -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 -regularity. Math. Ann., 319(4):735–758, 2001.