1 Introduction

1.1 Motivation and background

Gaussian processes play an important role for modeling in spatial statistics. Typical applications arise in the environmental sciences, where geographically indexed data is collected, including climatology [3, 66], oceanography [8], meteorology [38], and forestry [7, 43, 54]. More generally, hierarchical models based on Gaussian processes have been used in various disciplines, where spatially dependent (or spatiotemporal) data is recorded, such as demography [28, 62], epidemiology [48], finance [33], and neuroimaging [56].

Since a Gaussian process \((X(j))_{j\in {\mathcal {I}}}\) is fully characterized by its mean and its covariance function, second-order-based approaches focus on the construction of appropriate covariance classes. In the case that the index set \({\mathcal {I}}\) is given by a spatial domain in the Euclidean space \({\mathcal {I}}= {\mathcal {D}}\subseteq {\mathbb {R}}^d\), the Matérn covariance class [54] is an important and widely used model. The Matérn covariance function is given by

$$\begin{aligned} \varrho (x,y) = 2^{1-\nu }\sigma ^2[{\varGamma }(\nu )]^{-1} (\kappa \Vert x-y\Vert _{{\mathbb {R}}^d})^\nu K_\nu (\kappa \Vert x-y\Vert _{{\mathbb {R}}^d}), \quad x,y\in {\mathcal {D}}, \end{aligned}$$
(1.1)

where \(K_\nu \) denotes the modified Bessel function of the second kind. It is indexed by the three interpretable parameters \(\nu ,\kappa ,\sigma ^2\in (0,\infty )\), which determine smoothness, correlation length and variance of the process. It is this feature that renders the Matérn class particularly suitable for making inference about spatial data [69].

When considering spatiotemporal phenomena, the following difficulties occur:

  1. 1.

    It is desirable to control the properties of the stochastic process named above (in particular, smoothness and correlation lengths) separately in space and time. For this reason, considering (1.1) in \(d+1\) dimensions is not expedient and it is a difficult task to construct appropriate spatiotemporal covariance models, see e.g. [22, 34, 36, 64, 65, 70].

  2. 2.

    Second-order-based approaches require the factorization of, in general, dense covariance matrices, causing computational costs which are cubic in the number of observations. The two common assumptions imposed on spatiotemporal covariance models to reduce the computational costs—separability (factorization into merely spatial and temporal covariance functions) and stationarity (invariance under translations)—have proven unrealistic in many situations, see [23, 55, 70]. In particular, Stein [70] criticized the behavior of separable covariance functions with respect to their differentiability.

Owing to these problems, the class of dynamical models has gained popularity. The name originates from focusing on the dynamics of the stochastic process which are described either by means of conditional probability distributions or by representing the process as a solution of a stochastic partial differential equation (SPDE). The latter approach was originally proposed in the merely spatial case, motivated by the following observation made by Whittle [73]: A stationary process \((X(x))_{x\in {\mathcal {D}}}\) indexed by the entire Euclidean space \({\mathcal {D}}={\mathbb {R}}^d\) which solves the SPDE

$$\begin{aligned} \bigl ( \kappa ^2-{\varDelta }\bigr )^\beta X(x) = {\mathcal {W}}(x), \quad x\in {\mathcal {D}}, \end{aligned}$$
(1.2)

has a covariance function of Matérn type (1.1) with \(\nu = 2\beta - {d}/{2}\). Here, \({\varDelta }\) denotes the Laplacian and \({\mathcal {W}}\) is Gaussian white noise. This relation gave rise to the SPDE approach proposed by Lindgren, Rue, and Lindström [50], where the SPDE (1.2) is considered on a bounded domain \({\mathcal {D}}\subsetneq {\mathbb {R}}^d\) and augmented with Dirichlet or Neumann boundary conditions. Besides enabling the applicability of efficient numerical methods available for (S)PDEs, such as finite element methods [11, 13, 14, 21, 40, 50] or wavelets [16, 39], this approach has the advantage of allowing for

  1. (a)

    nonstationary or anisotropic generalizations by replacing the operator \(\kappa ^2-{\varDelta }\) in (1.2) with more general strongly elliptic second-order differential operators,

    $$\begin{aligned} (L v)(x) = \kappa ^2(x) v(x) - \nabla \cdot ( a(x) \, \nabla v(x) ), \quad x \in {\mathcal {D}}, \end{aligned}$$
    (1.3)

    where \(\kappa :{\mathcal {D}}\rightarrow {\mathbb {R}}\) and \(a:{\mathcal {D}}\rightarrow {\mathbb {R}}_{\textrm{sym}}^{d\times d}\) are functions [6, 11, 13, 14, 21, 35, 40, 50];

  2. (b)

    more general domains, such as surfaces [15, 40] or manifolds [39].

In the SPDE (1.2) the fractional exponent \(\beta \) defines the (spatial) differentiability of its solution, see e.g. [21]. A realistic description of spatiotemporal phenomena necessitates controllable differentiability in space and time. This motivates to consider the space–time fractional SPDE model

$$\begin{aligned} \left\{ \begin{aligned} \left( \partial _t + L^\beta \right) ^\gamma X(t,x)&= \dot{\mathcal {W}}(t,x),&\;&t \in [0,T], \quad x \in \mathcal {D}, \\ X(0,x)&= X_0(x),&\;&x\in \mathcal {D}, \end{aligned} \right. \end{aligned}$$
(1.4)

where L in (1.3) is augmented with boundary conditions on \(\partial {\mathcal {D}}\), \((X_0(x))_{x\in {\mathcal {D}}}\) is the initial random field, \(\dot{{\mathcal {W}}}\) denotes space–time Gaussian white noise, and \(T \in (0,\infty )\) is the time horizon. Whenever \(\beta =\gamma =1\), the SPDE (1.4) simplifies to the stochastic heat equation and this spatiotemporal model had already been mentioned in [50] and it was used for statistical inference in [18, 67]. The novelty and sophistication of the SPDE model (1.4) lies in the fractional power \(\gamma \in (0,\infty )\) of the parabolic operator. Notably, it is the interplay of the parameters \(\beta \) and \(\gamma \) that will facilitate controlling spatial and temporal smoothness of the solution process. For \({\mathcal {D}}={\mathbb {R}}^d\), this has recently been investigated via Fourier techniques in [49], see also [4, 19, 44].

Besides the aforementioned benefits of the SPDE approach and in contrast to the SPDE \(\bigl ( \partial _t^\gamma + L^\beta \bigr ) X = {\dot{{\mathcal {W}}}}\), considered for instance in [17, 30], the SPDE model (1.4) furthermore exhibits a long-time behavior resembling the spatial model (1.2).

1.2 Contributions

We introduce a novel interpretation of (1.4) with \(X_0=0\) as a fractional parabolic stochastic evolution equation, and correspondingly define mild and weak solutions for it. To this end, we first give a meaning to fractional powers of an operator of the form \(\partial _t + A\), where \(-A\) generates a \(C_0\)-semigroup. Generalizing the approach taken for \(\gamma =1\) in [25, Chapter 5], we prove that mild and weak solutions are equivalent under natural assumptions, and we investigate their existence, uniqueness, regularity, and covariance. Our main findings are that the problem (1.4) is well-posed, and the properties of its solution X with respect to smoothness and covariance structure generalize those of the spatial Whittle–Matérn SPDE model (1.2) and relate to the parameters \(\beta ,\gamma \in (0,\infty )\) in the desired way. Restricting the analysis to a zero initial field is justified by our primary interest in regularity related to the dynamics of (1.4) and the long-time behavior of solutions.

In comparison with [9, 10, 51, 57, 71]—the only previous works on an equation of the form \((\partial _t + L)^\gamma u = f\) known to the authors—the main contributions of this work, besides considering a stochastic right-hand side, are the fractional power \(\beta \) in (1.4) and the method of proving regularity using semigroups. As opposed to the extension approach in [9, 10, 51, 57, 71], this setting does not require a Euclidean structure.

1.3 Outline

Preliminary notation and theory will be introduced in Sect. 2. In Sect. 3 we give a meaning to the parabolic operator \(\partial _t + A\) and its fractional powers in order to introduce well-defined mild and weak solutions of (1.4) with \({X_0=0}\). Subsequently, we analyze these in terms of spatiotemporal regularity. Section 4 is concerned with the covariance structure of solutions. Finally, in Sect. 5 we apply our results to the space–time Whittle–Matérn SPDE (1.4) considered on a bounded Euclidean domain or on a surface. This article is supplemented by two appendices: “Appendix A” contains several technical auxiliary results used in the proofs of Sect. 3. “Appendix B” collects necessary definitions and results from functional calculus.

2 Preliminaries

2.1 Notation

The sets \({\mathbb {N}}:= \{1,2,3,\dots \}\) and \({\mathbb {N}}_0:= {\mathbb {N}}\cup \{0\}\) denote the positive and non-negative integers, respectively. We write \(s\wedge t\) (or \(s \vee t\)) for the minimum (or maximum) of two real numbers \(s,t\in {\mathbb {R}}\). The real and imaginary parts of a complex number \(z \in {\mathbb {C}}\) are denoted by \({\text {Re}}z\) and \({\text {Im}}z\), respectively; its argument, denoted \(\arg z\), takes its values in \((-\pi ,\pi ]\). We write \({\textbf{1}}_D\) for the indicator function of a set D. The restriction of a function \(f :D \rightarrow E\) to a subset \(D_0 \subseteq D\) is denoted by \(f|_{D_0}:D_0 \rightarrow E\); the image of \(D_0\) under a linear mapping T is written as \(T D_0\). Given two parameter sets \({\mathscr {P}},{\mathscr {Q}}\) and two mappings \({\mathscr {F}},{\mathscr {G}}:{\mathscr {P}}\times {\mathscr {Q}}\rightarrow {\mathbb {R}}\), we use the expression \({\mathscr {F}} (p,q) \lesssim _{q} {\mathscr {G}} (p,q)\) to indicate that for each \(q \in {\mathscr {Q}}\) there exists a constant \({C_q \in (0,\infty )}\) such that \({\mathscr {F}} (p,q) \le C_q \, {\mathscr {G}} (p,q)\) for all \({p\in {\mathscr {P}}}\). We write \({\mathscr {F}} (p,q) \eqsim _q {\mathscr {G}} (p,q)\) if both relations, \({\mathscr {F}} (p,q) \lesssim _q {\mathscr {G}}(p,q)\) and \({\mathscr {G}} (p,q) \lesssim _q {\mathscr {F}}(p,q)\), hold simultaneously.

2.2 Banach spaces and operators

If not specified otherwise, E or F denote separable Banach spaces. We instead write H or U if we work with separable Hilbert spaces and wish to emphasize this. The scalar field \({\mathbb {K}}\) is either given by the real numbers \({\mathbb {R}}\) or the complex numbers \({\mathbb {C}}\). A norm on E will be denoted by \(\Vert \,\cdot \,\Vert _{E}\) and an inner product on H by \((\,\cdot , \,\cdot \,)_{H}\). We write I for the identity operator. The notation \({E \hookrightarrow F}\) indicates that E is continuously embedded in F, i.e., there exists a bounded injective map from E to F. The dual space of E is denoted by \(E^*\). We write \(\overline{E_0}^E\) for the closure of a subset \(E_0\subseteq E\) with respect to the norm on E; the superscript may be omitted when there is no risk of confusion. The Borel \(\sigma \)-algebra of E is denoted by \({\mathcal {B}}(E)\).

We write \({T \in {\mathscr {L}}(E;F)}\) if the linear operator \(T:E \rightarrow F\) is bounded. Whenever \(E=F\), we abbreviate \({\mathscr {L}}(E):= {\mathscr {L}}(E;E)\), and this convention holds also for all other spaces of operators to be introduced. The space \({\mathscr {L}}(E;F)\) is rendered a Banach space when equipped with the usual operator norm; the space of Hilbert–Schmidt operators \({\mathscr {L}}_2(U;H)\subseteq {\mathscr {L}}(U;H)\) is a Hilbert space with respect to the inner product \((T,S)_{{\mathscr {L}}_2(U;H)}:= \sum _{j \in {\mathbb {N}}}(Te_j, Se_j)_H\), where \((e_j)_{j\in {\mathbb {N}}}\) is any orthonormal basis for U. We write \({T^* \in {\mathscr {L}}(F^*; E^*)}\) for the adjoint operator of \(T\in {\mathscr {L}}(E;F)\). In the case that \(T \in {\mathscr {L}}(U;H)\), we identify \(U^* = U\) and \(H^* = H\) via the Riesz maps, so that \(T^*\in {\mathscr {L}}(H;U)\). An operator \(T\in {\mathscr {L}}(H)\) is said to be self-adjoint if \(T^* = T\), non-negative if \((Tx,x)_{H}\ge 0\) holds for all \(x\in H\), and strictly positive if there exists a constant \(\theta \in (0,\infty )\) such that \((Tx,x)_{H} \ge \theta \Vert x\Vert _H^2\) holds for all \(x\in H\).

A linear operator A on E with domain \({\textsf{D}}(A)\) is denoted by \({A:{\textsf{D}}(A) \subseteq E \rightarrow E}\) and its range by \({\textsf{R}}(A)\). We call A closed if its graph \({\textsf{G}}(A):= {\{(x,Ax): x \in {\textsf{D}}(A)\}}\) is closed with respect to the graph norm \(\Vert (x,Ax)\Vert _{{\textsf{G}}(A)}:= \Vert x\Vert _{E} + \Vert Ax\Vert _{E}\), and densely defined if \({\textsf{D}}(A)\) is dense in E. The definition \(\Vert x\Vert _{{\textsf{D}}(A)}:= \Vert (x,Ax)\Vert _{{\textsf{G}}(A)}\) yields a norm on \({\textsf{D}}(A)\). If \({\textsf{G}}(A) \subseteq {\textsf{G}}({{\widetilde{A}}})\) for another linear operator \({{\widetilde{A}}}\) on E, then \({{\widetilde{A}}}\) is called an extension of A and we write \(A \subseteq {{\widetilde{A}}}\). If \(\overline{ {\textsf{G}}(A) }\) is the graph of a linear operator, then we call this operator the closure of A, denoted \(\overline{ A }\).

2.3 Function spaces

Let a measure space \((S, {\mathcal {S}}, \mu )\) be given. We abbreviate the phrases “almost everywhere” and “almost all” by “a.e.” and “a.a.”, respectively.

We say that a function \(f :S \rightarrow E\) is strongly measurable if it is the \(\mu \)-a.e. limit of measurable simple functions. For \(p \in [1,\infty ]\), the Bochner space of (equivalence classes of) strongly measurable, p-integrable functions is denoted by \(L^p(S;E)\). It is equipped with the norm

$$\begin{aligned} \Vert f\Vert _{L^p(S;E)}:= {\left\{ \begin{array}{ll} \left( \int _S \Vert f(t)\Vert _{E}^p \mathop {}\!\textrm{d}\mu (t) \right) ^{{1}/{p}} &{}\text {if}\, p \in [1,\infty ), \\ {\text {ess\,sup}}_{t\in S} \Vert f(t)\Vert _{E} &{}\text {if}\, p = \infty , \end{array}\right. } \end{aligned}$$

where \({\text {ess\,sup}}\) denotes the essential supremum. The norm on \(L^2(S;H)\) is induced by the inner product \((f,g)_{L^2(S;H)}:= \int _S (f(t),g(t))_H \mathop {}\!\textrm{d}\mu (t)\).

Now let S be an interval \(S:= J \subseteq {\mathbb {R}}\), equipped with the Borel \(\sigma \)-algebra and the Lebesgue measure. The space of continuous functions from J to E will be denoted by C(JE) or \(C^{0,0}(J;E)\) and be endowed with the supremum norm if J is compact. For \(\alpha \in (0,1]\) and compact J, we consider the space \(C^{0,\alpha }(J;E)\) of \(\alpha \)-Hölder continuous functions with norm

$$\begin{aligned} \Vert f\Vert _{C^{0,\alpha }(J;E)}:= & {} |f|_{C^{0,\alpha }(J;E)} + \Vert f\Vert _{C(J;E)}, \; \text {where} \;\; \\ |f|_{C^{0,\alpha }(J;E)}:= & {} \sup _{ t, s\in J, \, t\ne s } \frac{\Vert f(t)-f(s)\Vert _{E}}{|t-s|^\alpha } \end{aligned}$$

is the \(\alpha \)-Hölder seminorm. For \(n \in {\mathbb {N}}_0\) and \(0 \le \alpha \le 1\), the space \(C^{n,\alpha }(J;E)\) consists of functions whose nth derivative exists and belongs to \(C^{0,\alpha }(J;E)\). If J is compact we use the norm \(\Vert f\Vert _{C^{n,\alpha }(J;E)}:= \Vert f^{(n)}\Vert _{C^{0,\alpha }(J;E)} + \sum _{k=0}^{n-1} \Vert f^{(k)}\Vert _{C(J;E)}\), where \(f^{(k)}\) denotes the kth derivative of f. Moreover, we define \(C^{\infty }(J;E):= \bigcap _{n\in {\mathbb {N}}} C^{n,0}(J;E)\). We say that \(f \in C^{n,\alpha }(J;E)\) is compactly supported if the support of f, defined by

$$\begin{aligned} {\text {supp}} f:= \overline{\{t\in J: f(t)\ne 0\}}^J, \end{aligned}$$

is compact. The space consisting of such functions is denoted by \(C_c^{n,\alpha }(J;E)\). If f vanishes at a point \(t \in J\), then we use the notation \({f \in C^{n,\alpha }_{0,\{t\}}(J;E)}\). The spaces \(C_c^\infty (J;E)\) and \(C_{0,\{t\}}^\infty (J;E)\) are defined analogously.

For an open interval J, we say that \(u \in L^2(J;E)\) belongs to \(H^1(J;E)\) if there exists a function \(v \in L^2(J;E)\) such that \(\int _J v(t) \phi (t)\mathop {}\!\textrm{d}t = -\int _J u(t) \phi '(t)\mathop {}\!\textrm{d}t\) holds for all \(\phi \in C_c^\infty (J;{\mathbb {R}})\). The function \({\partial _t u:= v}\) is called the weak derivative of u and the norm on \(H^1(J;E)\) is \(\Vert u\Vert _{H^1(J;E)}:= \bigl ( \Vert u\Vert _{L^2(J;E)}^2 + \Vert \partial _t u\Vert _{L^2(J;E)}^2 \bigr )^{{1}/{2}}\). The completion of \(C_c^\infty ((0,\infty );E)\) with respect to the norm \(\Vert \,\cdot \,\Vert _{H^1(0,\infty ;E)}\) defines the space \(H^1_{0,\{0\}}(0,\infty ;E)\). Elements of \(H^1_{0,\{0\}}(J;E)\) are restrictions of functions in \(H^1_{0,\{0\}}(0,\infty ;E)\) to \({J\subseteq (0,\infty )}\).

Whenever the function space contains functions mapping to \(E={\mathbb {R}}\), we omit the codomain, e.g., we write \(L^p(S):= L^p(S;{\mathbb {R}})\) for the Lebesgue spaces.

2.4 Vector-valued stochastic processes

Throughout this article, \(({\varOmega },{\mathcal {F}},{\mathbb {P}})\) denotes a complete probability space equipped with a normal filtration \(({\mathcal {F}}_t)_{t\ge 0}\), i.e., \({\mathcal {F}}_0\) contains all elements \(B\in {\mathcal {F}}\) with \({\mathbb {P}}(B)=0\) and \({\mathcal {F}}_t = \bigcap _{s>t} {\mathcal {F}}_s\) for all \(t\ge 0\). Statements which hold \({\mathbb {P}}\)-almost surely are marked with “\({\mathbb {P}}\)-a.s.”.

We call every strongly measurable function \(Z:{\varOmega } \rightarrow E\) a (vector-valued) random variable, and the expectation of \(Z \in L^1({\varOmega };E)\) is defined as the Bochner integral \({\mathbb {E}}[Z]:= \int _{{\varOmega }} Z(\omega ) \mathop {}\!\textrm{d}{\mathbb {P}}(\omega )\). An E-valued stochastic process \(X=(X(t))_{t\in [0,T]}\) indexed by the interval [0, T], \(T\in (0,\infty )\), is called integrable if \({(X(t))_{t\in [0,T]} \subseteq L^p({\varOmega };E)}\) holds for \(p=1\), and square-integrable if this inclusion is true for \(p=2\). It is said to be predictable if it is strongly measurable as a mapping from \([0,T]\times {\varOmega }\) to E, where the former set is equipped with the \(\sigma \)-algebra generated by the family

$$\begin{aligned} \{ (s,t] \times F_s: 0\le s < t\le T,\, F_s\in {\mathcal {F}}_s \} \cup \{ \{0\}\times F_0: F_0\in {\mathcal {F}}_0 \}. \end{aligned}$$

Given another E-valued process \({{\widetilde{X}}}:= ({{\widetilde{X}}}(t))_{t\in [0,T]}\), we call \({{\widetilde{X}}}\) a modification of X, provided that \({\mathbb {P}}(X(t)={{\widetilde{X}}}(t))=1\) holds for all \(t\in [0,T]\). Moreover, X and \({{\widetilde{X}}}\) are said to be indistinguishable if \({\mathbb {P}}(\forall t\in [0,T]: X(t) = {{\widetilde{X}}}(t))=1\).

For a self-adjoint strictly positive operator \(Q \in {\mathscr {L}}(H)\), \((W^Q(t))_{t\ge 0}\) denotes a cylindrical Q-Wiener process with respect to \(({\mathcal {F}}_t)_{t\ge 0}\) which takes its values in H, cf. [52, Proposition 2.5.2]; if \(Q=I\), we omit the superscript and call \((W(t))_{t\ge 0}\) a cylindrical Wiener process.

3 Analysis of the fractional stochastic evolution equation

The aim of this section is to define and analyze solutions to the following stochastic evolution equation of the general fractional order \(\gamma \in (0,\infty )\):

$$\begin{aligned} (\partial _t + A)^\gamma X (t) = {\dot{W}}^{Q}(t), \quad t\in [0,T], \qquad X(0)=0. \end{aligned}$$
(3.1)

We interpret this as an abstraction of (1.4) with \(X_0 = 0\). As noted in the introduction, we restrict the discussion to a zero initial field, since we are primarily interested in properties resulting from the dynamics of the SPDE (1.4), respectively (3.1), and the long-time behavior for \(0\ll T < \infty \) of its solution. We also note that imposing non-zero boundary data for fractional problems is, in general, highly non-trivial, see e.g. the recent works [1, 5] on the fractional Laplacian.

In Sect. 3.1 we investigate the parabolic operator \({\mathcal {B}}\), which is defined as the closure of the sum operator \(\partial _t + A\) on an appropriate domain. In particular, we consider the \(C_0\)-semigroup generated by \(-{\mathcal {B}}\), which is used to define fractional powers \({\mathcal {B}}^\gamma \) for \(\gamma \in {\mathbb {R}}\). Interpreting the expression \((\partial _t + A)^\gamma \) appearing in (3.1) as \({\mathcal {B}}^\gamma \), we use this result to define mild solutions in Sect. 3.2. In this part, we furthermore introduce a weak solution concept for (3.1), and prove equivalence of the two solution concepts as well as existence and uniqueness of mild and weak solutions. Spatiotemporal regularity of solutions is the subject of Sect. 3.3.

3.1 The parabolic operator and its fractional powers

In this subsection we define the parabolic operator \({\mathcal {B}}\) and fractional powers \({\mathcal {B}}^\gamma \). We start by formulating several assumptions on the linear operator A, to which we shall refer throughout the remainder of this work. For an overview of the theory of \(C_0\)-semigroups, we refer the reader to [31] or [61]. The complexification of a normed space or operator is indicated by the subscript \({\mathbb {C}}\); see Sect. B.2.1 in “Appendix B” for details.

Assumption 3.1

Let H be a separable Hilbert space over the real scalar field \({\mathbb {R}}\). We assume that the linear operator \({A:{\textsf{D}}(A)\subseteq H \rightarrow H}\) satisfies

  1. (i)

    \(-A\) generates a \(C_0\)-semigroup \((S(t))_{t\ge 0}\).

Sometimes we additionally require one or more of the following conditions:

  1. (ii)

    \((S(t))_{t\ge 0}\) is (uniformly) bounded analytic, i.e., the mapping \({t \mapsto S_{{\mathbb {C}}}(t)}\), where \({S_{{\mathbb {C}}}(t):= [S(t)]_{{\mathbb {C}}}}\), extends to a bounded holomorphic function on an open sector \({\varSigma }_\omega \subseteq {\mathbb {C}}\) for some angle \(\omega \in (0,\pi )\) (see Definition B.1 in “Appendix B”);

  2. (iii)

    \(A_{{\mathbb {C}}}\) admits a bounded \(H^\infty \)-calculus with \(\omega _{H^\infty }(A_{\mathbb {C}})<\tfrac{\pi }{2}\), see Definition B.3;

  3. (iv)

    A has a bounded inverse.

Under Assumption 3.1(i), Lemma B.6 allows us to use several results from [31, 37, 61] for \(C_0\)-semigroups and their generators on complex spaces also for \((S(t))_{t\ge 0}\) and \(-A\). For instance, by [31, Theorem II.1.4] and [61, Chapter 1, Theorem 2.2] the operator A is closed and densely defined, and the \(C_0\)-semigroup \((S(t))_{t\ge 0}\) satisfies

$$\begin{aligned} \exists M \in [1,\infty ), \; w \in {\mathbb {R}}: \quad \Vert S(t)\Vert _{{\mathscr {L}}(H)} = \Vert S_{\mathbb {C}}(t)\Vert _{{\mathscr {L}}(H_{\mathbb {C}})} \le M e^{-wt} \quad \forall t \ge 0.\nonumber \\ \end{aligned}$$
(3.2)

If the conditions (i), (ii) and (iv) are satisfied, then (3.2) holds for some \(w \in (0,\infty )\), see e.g. [61, p. 70]. In this case, \((S(t))_{t\ge 0}\) is said to be exponentially stable. Moreover, we note that Assumption 3.1(ii) is equivalent to the operator \(A_{\mathbb {C}}\) being sectorial with \(\omega (A_{\mathbb {C}}) < \tfrac{\pi }{2}\) by Theorem B.2, and that consequently condition (iii) implies (ii) since \(\omega (A_{\mathbb {C}}) \le \omega _{H^\infty }(A_{\mathbb {C}})\) by Remark B.5. Whenever the conditions (i) and (ii) are satisfied, we have the following useful estimate (see [37, Proposition 3.4.3]):

$$\begin{aligned} \forall c \in [0,\infty ): \quad \Vert A^c S(t) \Vert _{{\mathscr {L}}(H)} = \Vert A_{\mathbb {C}}^c S_{\mathbb {C}}(t) \Vert _{{\mathscr {L}}(H_{\mathbb {C}})} \lesssim _c t^{-c} \quad \forall t \in (0,\infty ). \end{aligned}$$
(3.3)

As a first step towards defining the parabolic operator \({\mathcal {B}}\), we define the Bochner space counterpart \({\mathcal {A}}:{\textsf{D}}({\mathcal {A}}) \subseteq L^2(0,T;H) \rightarrow L^2(0,T;H)\) of A by

$$\begin{aligned} \begin{aligned}&[{\mathcal {A}}v](\vartheta ):= Av(\vartheta ), \quad v \in {\textsf{D}}({\mathcal {A}}), \ \text {a.a.\ }\vartheta \in (0,T),\\&{\textsf{D}}({\mathcal {A}}) = L^2(0,T;{\textsf{D}}(A)):= \bigl \{ v \in L^2(0,T;H): \Vert Av(\,\cdot \,)\Vert _{L^2(0,T;H)} < \infty \bigr \}. \end{aligned} \end{aligned}$$
(3.4)

The \(C_0\)-semigroup \((S(t))_{t\ge 0}\) on H, generated by \(-A\), can be associated to a family of operators \(({\mathcal {S}}(t))_{t\ge 0}\) on \(L^2(0,T;H)\) in a similar way:

$$\begin{aligned} {[}{\mathcal {S}}(t)v](\vartheta ):= S(t)v(\vartheta ), \quad \;\; t \ge 0, \ v \in L^2(0,T;H), \ \text {a.a.\ }\vartheta \in (0,T). \end{aligned}$$
(3.5)

It turns out that \(({\mathcal {S}}(t))_{t\ge 0}\subseteq {\mathscr {L}}(L^2(0,T;H))\) is again a \(C_0\)-semigroup, with infinitesimal generator \(-{\mathcal {A}}\), see Proposition A.3 in “Appendix A”.

In addition, we consider the family of zero-padded right-translation operators \(({\mathcal {T}}(t))_{t\ge 0}\) on \(L^2(0,T; H)\), defined by

$$\begin{aligned} {[}{\mathcal {T}}(t)v](\vartheta ):= {{\widetilde{v}}}(\vartheta - t), \quad \;\; t\ge 0, \ v \in L^2(0,T; H), \ \text {a.a.\ }\vartheta \in (0,T), \end{aligned}$$
(3.6)

where \({{\widetilde{v}}} \in L^2(-\infty ,T;H)\) denotes the extension of v by zero to \((-\infty ,T)\). As shown in Proposition A.5 in “Appendix A”, also \(({\mathcal {T}}(t))_{t\ge 0} \subseteq {\mathscr {L}}(L^2(0,T;H))\) is a \(C_0\)-semigroup and its infinitesimal generator is given by \(-\partial _t\), where

$$\begin{aligned} \partial _t :{\textsf{D}}(\partial _t) \subseteq L^2(0,T;H) \rightarrow L^2(0,T;H), \qquad {\textsf{D}}(\partial _t) = H^1_{0,\{0\}}(0,T; H), \end{aligned}$$
(3.7)

denotes the Bochner–Sobolev vector-valued weak derivative. We point out that the domain \({\textsf{D}}(\partial _t)= H^1_{0,\{0\}}(0,T; H)\) encodes the zero initial condition of the SPDE (3.1). Furthermore, note that it readily follows from the definitions in (3.5) and (3.6) that, for all \(t\ge 0\), every \( v \in L^2(0,T;H)\), and a.a. \(\vartheta \in (0,T)\),

$$\begin{aligned} {[}{\mathcal {S}}(t) {\mathcal {T}}(t)v](\vartheta ) = [{\mathcal {T}}(t) {\mathcal {S}}(t)v](\vartheta ) = S(t) {{\widetilde{v}}}(\vartheta -t), \end{aligned}$$

i.e., the semigroups \(({\mathcal {S}}(t))_{t\ge 0}\) and \(({\mathcal {T}}(t))_{t\ge 0}\) commute.

We now define the sum operator \(\partial _t + {\mathcal {A}}:{\textsf{D}}(\partial _t+{\mathcal {A}})\subseteq L^2(0,T;H) \rightarrow L^2(0,T;H)\) on its natural domain, that is

$$\begin{aligned} \begin{aligned} ( \partial _t+ \mathcal {A})v&:= \partial _t v + \mathcal {A} v, \\ v\in \mathsf D(\partial _t + \mathcal {A})&= H^1_{0,\{0\}}(0,T; H) \cap L^2(0,T; \mathsf D(A)), \end{aligned} \end{aligned}$$
(3.8)

with \({\mathcal {A}}\) and \(\partial _t\) as given in (3.4) and (3.7), respectively. The next proposition shows that the closure of \(-(\partial _t + {\mathcal {A}})\) again generates a \(C_0\)-semigroup, namely the product semigroup of \(({\mathcal {S}}(t))_{t\ge 0}\) and \(({\mathcal {T}}(t))_{t\ge 0}\).

Proposition 3.2

Let Assumption 3.1(i) be satisfied. The closure \({\mathcal {B}}:= \overline{ \partial _t + {\mathcal {A}} }\) of the sum operator \(\partial _t + {\mathcal {A}}\) defined in (3.8) exists and \(-{\mathcal {B}}\) generates the \(C_0\)-semigroup \(({\mathcal {S}}(t) {\mathcal {T}}(t))_{t\ge 0}\) on \(L^2(0,T;H)\), which satisfies

$$\begin{aligned} \Vert {\mathcal {S}}(t) {\mathcal {T}}(t) \Vert _{{\mathscr {L}}(L^2(0,T;H))} = \Vert {\mathcal {T}}(t) {\mathcal {S}}(t) \Vert _{{\mathscr {L}}(L^2(0,T;H))} = {\left\{ \begin{array}{ll} \Vert S(t)\Vert _{{\mathscr {L}}(H)} &{} \text {if } 0 \le t < T,\\ 0 &{} \text {if } t \ge T, \end{array}\right. } \end{aligned}$$

where \(({\mathcal {S}}(t))_{t\ge 0}\) and \(({\mathcal {T}}(t))_{t\ge 0}\) are defined as in (3.5) and (3.6), respectively.

Proof

By the commutativity of the semigroups \(({\mathcal {S}}(t))_{t\ge 0}\) and \(({\mathcal {T}}(t))_{t\ge 0}\), we may conclude that \(( {\mathcal {T}}(t) {\mathcal {S}}(t) )_{t\ge 0}\) is a \(C_0\)-semigroup whose generator is an extension of \(-(\partial _t + {\mathcal {A}})\), and the domain of the generator contains \(H^1_{0,\{0\}}(0,T;H) \cap L^2(0,T;{\textsf{D}}(A))\) as a subspace that is dense with respect to the graph norm, see [31, Example II.2.7]. Subsequently, Lemma A.2 shows that the generator is the closure of \(-(\partial _t + {\mathcal {A}})\).

Fix \(t\in [0,T)\). The inequality \(\Vert {\mathcal {T}}(t) {\mathcal {S}}(t) \Vert _{{\mathscr {L}}(L^2(0,T;H))} \le \Vert S(t)\Vert _{{\mathscr {L}}(H)}\) follows by the contractivity of \({\mathcal {T}}(t)\) and the operator norm isometry from Lemma A.1(a). Now we turn to the reverse inequality. By definition of the operator norm on \({\mathscr {L}}(H)\), there exists a normalized sequence \((x_n)_{n\in {\mathbb {N}}}\) in H such that \(\Vert S(t)x_n\Vert _{H} \ge \Vert S(t)\Vert _{{\mathscr {L}}(H)} - \frac{1}{n}\) holds for all \(n\in {\mathbb {N}}\). Correspondingly, define the sequence \((v_n)_{n\in {\mathbb {N}}}\) in \(L^2(0,T;H)\) by \(v_n(\vartheta ):= (T-t)^{-{1}/{2}} {\textbf{1}}_{(0,T-t)}(\vartheta ) x_n\) for every \(\vartheta \in (0,T)\) and all \(n\in {\mathbb {N}}\). Note that \(\Vert v_n\Vert _{L^2(0,T;H)} = 1\) for every \(n \in {\mathbb {N}}\), and

$$\begin{aligned} \Vert {\mathcal {T}}(t) {\mathcal {S}}(t) v_n\Vert _{L^2(0,T;H)} = \Vert (T-t)^{-{1}/{2}}{\textbf{1}}_{(t,T)}\Vert _{L^2(0,T)} \Vert S(t)x_n\Vert _{H} \ge \Vert S(t)\Vert _{{\mathscr {L}}(H)} - \tfrac{1}{n}. \end{aligned}$$

As this holds for all \(n \in {\mathbb {N}}\), we conclude that \(\Vert {\mathcal {T}}(t) {\mathcal {S}}(t)\Vert _{{\mathscr {L}}(L^2(0,T;H))} \ge \Vert S(t)\Vert _{{\mathscr {L}}(H)}\). The final assertion for \(t\ge T\) follows from the fact that \({\mathcal {T}}(t) = 0\) for \(t \ge T\). \(\square \)

Remark 3.3

The closure \({\mathcal {B}}=\overline{ \partial _t + {\mathcal {A}} }\) appearing in Proposition 3.2 raises the question of when the sum operator itself is closed. The answer is intimately related to the subject of maximal \(L^p\)-regularity; we refer the reader to [29] or [47] for detailed accounts of this theory. In the Hilbert space setting, the sum turns out to be closed under Assumptions 3.1(i),(ii). Indeed, \([\partial _t]_{\mathbb {C}}\) has a bounded \(H^\infty \)-calculus with \({\omega _{H^\infty }([\partial _t]_{\mathbb {C}}) \le \tfrac{\pi }{2}}\) since \(({\mathcal {T}}(t))_{t\ge 0}\) and \(({\mathcal {T}}_{\mathbb {C}}(t))_{t\ge 0}\) are contractive, see Definition B.3 in “Appendix B” and [42, Theorem 10.2.24]. By Assumption 3.1(ii) and Theorem B.2, we have \(\omega (A_{\mathbb {C}}) < \tfrac{\pi }{2}\), and the same follows for \({\mathcal {A}}_{\mathbb {C}}\) by applying Lemma A.1(a) to its resolvent operators. Thus, we may conclude with [47, Theorem 12.13] that \([\partial _t + {\mathcal {A}}]_{\mathbb {C}}\) is closed, so that the same holds for \(\partial _t + {\mathcal {A}}\).

We are now in the position to define fractional powers of the parabolic operator. For \(\gamma \in (0,\infty )\) we work with the following representation (see “Appendix B.2.2”):

$$\begin{aligned} {\mathcal {B}}^{-\gamma }:= \frac{1}{{\varGamma }(\gamma )} \int _0^\infty s^{\gamma -1} {\mathcal {S}}(s) {\mathcal {T}}(s) \mathop {}\!\textrm{d}s = \frac{1}{{\varGamma }(\gamma )} \int _0^T s^{\gamma -1} {\mathcal {S}}(s) {\mathcal {T}}(s) \mathop {}\!\textrm{d}s. \end{aligned}$$
(3.9)

Note that, for any \(\gamma \in (0,\infty )\), this definition yields a well-defined bounded linear operator on \(L^2(0,T;H)\), since the product semigroup \(({\mathcal {S}}(t){\mathcal {T}}(t))_{t\ge 0}\) was seen to be exponentially stable (in fact, eventually zero) in Proposition 3.2.

The next result shows that the pointwise evaluation of \({\mathcal {B}}^{-\gamma }f\) at \(t\in [0,T]\) is meaningful, provided that \(\gamma >\frac{1}{2}\).

Proposition 3.4

Suppose Assumption 3.1(i) and let \(p\in (1,\infty ), \gamma \in ({1}/{p},\infty )\). Then

$$\begin{aligned} f \mapsto {\mathscr {B}}_{\gamma ,p} f, \quad \;\; [{\mathscr {B}}_{\gamma ,p} f](t):= \frac{1}{{\varGamma }(\gamma )} \int _0^t (t-s)^{\gamma -1} S(t-s)f(s) \mathop {}\!\textrm{d}s \quad \forall t \in [0,T], \end{aligned}$$
(3.10)

defines a bounded linear operator, mapping \(f\in L^p(0,T;H)\) into \(C_{0,\{0\}}([0,T];H)\).

In particular, if \(\gamma \in ({1}/{2}, \infty )\), we have for the negative fractional parabolic operator \({{\mathcal {B}}^{-\gamma }}\) defined by (3.9) when acting on \(f\in L^2(0,T;H)\) the pointwise formula

$$\begin{aligned}{}[{\mathcal {B}}^{-\gamma } f](t) = [{\mathscr {B}}_{\gamma ,2} f](t) = \frac{1}{{\varGamma }(\gamma )} \int _0^t (t-s)^{\gamma -1} S(t-s)f(s) \mathop {}\!\textrm{d}s \quad \forall t \in [0,T]. \end{aligned}$$
(3.11)

Proof

By [25, Proposition 5.9], for \(p\in (1,\infty )\) and \(\gamma \in ({1}/{p},\infty )\), the operator \({\mathscr {B}}_{\gamma ,p}\) defined by (3.10) maps continuously from \(L^p(0,T;H)\) to \(C_{0,\{0\}}([0,T];H)\).

Next, note that for all \(f \in L^2(0,T;H)\) and a.a. \(t \in [0,T]\), we obtain by (3.9)

$$\begin{aligned} {[}{\mathcal {B}}^{-\gamma }f ](t)&= \frac{1}{{\varGamma }(\gamma )} \int _0^\infty s^{\gamma -1} [ {\mathcal {S}}(s) {\mathcal {T}}(s) f ](t)\mathop {}\!\textrm{d}s = \frac{1}{{\varGamma }(\gamma )} \int _0^t s^{\gamma -1} S(s) f(t-s) \mathop {}\!\textrm{d}s\\&= \frac{1}{{\varGamma }(\gamma )} \int _0^t (t-s)^{\gamma -1} S(t-s) f(s) \mathop {}\!\textrm{d}s = [{\mathscr {B}}_{\gamma ,2} f](t). \end{aligned}$$

Thus, by the first part of this proposition, for every \(\gamma \in ({1}/{2},\infty )\), we obtain that \({\textsf{R}}({\mathcal {B}}^{-\gamma })\subseteq C_{0,\{0\}}([0,T];H)\) and the above identities hold pointwise in \(t\in [0,T]\). \(\square \)

Remark 3.5

Propositions 3.2 and 3.4 require only Assumption 3.1(i), i.e., that \(-A\) generates the \(C_0\)-semigroup \((S(t))_{t\ge 0}\). Exponential stability or uniform boundedness of \((S(t))_{t\ge 0}\) are not needed, since we consider linear operators on \(L^2(0,T;H)\) (instead of \(L^2(0,\infty ;H)\)), allowing us to use uniform boundedness of \((S(t))_{t\ge 0}\) on the compact interval [0, T] to derive exponential stability of \(({\mathcal {S}}(t){\mathcal {T}}(t))_{t\ge 0}\).

In what follows, we may also consider the adjoint operator \({\mathcal {B}}^{-\gamma *}:= ({\mathcal {B}}^{-\gamma })^*\). More specifically, we will use it in the next section to define a weak solution to the fractional parabolic SPDE (3.1). The following lemma provides useful results for the adjoint \({\mathcal {B}}^{-\gamma *}\) which are analogous to those for \({\mathcal {B}}^{-\gamma }\) in Proposition 3.4. For ease of presentation, the proof has been moved to Sect. A.3 of “Appendix A”.

Lemma 3.6

Suppose Assumption 3.1(i) and let \(\gamma \in ({1}/{2},\infty )\). The adjoint negative fractional parabolic operator \({{\mathcal {B}}^{-\gamma *}}\) maps \(g\in L^2(0,T;H)\) into \(C_{0,\{T\}}([0,T];H)\), and

$$\begin{aligned} {[} {\mathcal {B}}^{-\gamma *}g ](s) = \frac{1}{{\varGamma }(\gamma )} \int _s^T (t-s)^{\gamma -1} [S(t-s)] ^* g(t) \mathop {}\!\textrm{d}t \quad \forall s \in [0,T]. \end{aligned}$$
(3.12)

Finally, we note that \({\mathcal {B}}^{-\gamma *} = ({\mathcal {B}}^*)^{-\gamma }\). To see that the fractional power on the right-hand side is indeed well-defined, we use [61, Chapter 1, Corollary 10.6] and conclude that \(-{\mathcal {B}}^*\) is the generator of the \(C_0\)-semigroup \(([{\mathcal {S}}(t){\mathcal {T}}(t)]^*)_{t\ge 0}\), which clearly inherits the exponential stability from \(( {\mathcal {S}}(t) {\mathcal {T}}(t))_{t\ge 0}\) since their norms are equal. The identity is then obtained as follows,

$$\begin{aligned} {\mathcal {B}}^{-\gamma *} = \biggl (\frac{1}{{\varGamma }(\gamma )} \int _0^\infty s^{\gamma -1} {\mathcal {S}}(s) {\mathcal {T}}(s) \mathop {}\!\textrm{d}s\biggr )^* = \frac{1}{{\varGamma }(\gamma )} \int _0^\infty s^{\gamma -1} [ {\mathcal {S}}(s) {\mathcal {T}}(s)]^*\mathop {}\!\textrm{d}s = ({\mathcal {B}}^*)^{-\gamma }, \end{aligned}$$

where the first and last identities are due to (3.9) and the second is a consequence of the general ability to interchange Bochner integrals and duality pairings.

3.2 Solution concepts, existence and uniqueness

We now turn towards defining solutions to (3.1) for fractional powers \(\gamma \in (0,\infty )\). Recall from Sect. 2 that \(({\varOmega }, {\mathcal {F}}, {\mathbb {P}})\) is a complete probability space equipped with a normal filtration \(({\mathcal {F}}_t)_{t\ge 0}\), and that \((W^Q(t))_{t\ge 0}\) is a cylindrical Q-Wiener process on H with respect to \(({\mathcal {F}}_t)_{t\ge 0}\), where \(Q \in {\mathscr {L}}(H)\) is self-adjoint and strictly positive.

Having defined and investigated the parabolic operator \({\mathcal {B}}\), its domain and its fractional powers, we are now in particular able to invert the fractional operator \({\mathcal {B}}^\gamma \). Equation (3.11) suggests the following definition of a fractional stochastic convolution as a mild solution to (3.1).

Definition 3.7

Let Assumption 3.1(i) hold and define, for \(\gamma \in (0,\infty )\), the stochastic convolution

$$\begin{aligned} {\widetilde{Z}}_\gamma (t):= \frac{1}{{\varGamma }(\gamma )} \int _0^t (t-s)^{\gamma -1} S(t-s) \mathop {}\!\textrm{d}W^Q(s), \quad t\in [0,T]. \end{aligned}$$
(3.13)

A predictable H-valued stochastic process \(Z_\gamma := (Z_\gamma (t))_{t\in [0,T]}\) is called a mild solution to (3.1) if, for all \(t\in [0,T]\), it satisfies \(Z_\gamma (t) = {\widetilde{Z}}_\gamma (t)\), \({\mathbb {P}}\)-a.s.

We first address existence and mean-square continuity of mild solutions. Furthermore, we adapt the Da Prato–Kwapień–Zabczyk factorization method (see [24, 25, Section 5.3]) to establish the existence of a pathwise continuous modification.

Theorem 3.8

Let Assumption 3.1(i) be satisfied and let \(\gamma \in (0,\infty )\) be such that

$$\begin{aligned} \exists \, \delta \in [0,\gamma ): \quad \int _0^T \bigl \Vert t^{\gamma -1-\delta } S(t) Q^{ \frac{1}{2} } \bigr \Vert _{{\mathscr {L}}_2(H)}^2 \mathop {}\!\textrm{d}t < \infty . \end{aligned}$$
(3.14)

The stochastic convolution \({\widetilde{Z}}_\gamma (t)\) in (3.13) belongs to \(L^2({\varOmega };H)\) for all \(t\in [0,T]\) if and only if (3.14) holds with \(\delta =0\). In this case, the mapping \(t\mapsto {\widetilde{Z}}_\gamma (t)\) is an element of \(C([0,T];L^p({\varOmega };H))\) for all \(p\in [1,\infty )\); in particular, there exists a mild solution in the sense of Definition 3.7, and it is mean-square continuous.

Whenever (3.14) holds for some \(\delta \in (0,\gamma )\), then for every \(p \in [1,\infty )\) there exists a modification of \({{\widetilde{Z}}}_\gamma \) with continuous sample paths belonging to \(L^p({\varOmega };C([0,T];H))\). In particular, the mild solution has a modification with continuous sample paths.

Proof

We first consider the case \(\delta =0\) in (3.14). By the Itô isometry (see e.g. [52, Proposition 2.3.5 and p. 32]), we obtain the identity

$$\begin{aligned} \sup _{t\in [0,T]} \Vert {\widetilde{Z}}_\gamma (t) \Vert _{L^2({\varOmega };H)}^2 = \frac{1}{|{\varGamma }(\gamma )|^2} \int _0^T \bigl \Vert t^{\gamma -1} S(t) Q^{ \frac{1}{2} } \bigr \Vert _{{\mathscr {L}}_2(H)}^2 \mathop {}\!\textrm{d}t. \end{aligned}$$

Therefore, \({\widetilde{Z}}_\gamma (t)\in L^2({\varOmega };H)\) holds for all \(t\in [0,T]\) if and only if (3.14) is satisfied with \(\delta =0\). The fact that \(t\mapsto {\widetilde{Z}}_\gamma (t)\) belongs to \(C([0,T];L^p({\varOmega };H))\) for every \({p \in [1,\infty )}\) will be shown in greater generality in Proposition 3.18, see Sect. 3.3.3.

Moreover, note that \({\widetilde{Z}}_\gamma :[0,T]\times {\varOmega } \rightarrow H\) is measurable and \(({\mathcal {F}}_t)_{t\in [0,T]}\)-adapted, and that mean-square continuity implies continuity in probability, so that we may apply [63, Proposition 3.21] to conclude that there exists a predictable modification \(Z_\gamma \) of \({\widetilde{Z}}_\gamma \). Then, \(Z_\gamma \) is a mild solution to (3.1) in the sense of Definition 3.7.

Now suppose that (3.14) holds for some \(\delta \in (0,\gamma )\) and let \(p\in ( {1}/{\delta } \vee 1, \infty )\). By the above considerations, \({{\widetilde{Z}}}_{\gamma -\delta }\) and \({{\widetilde{Z}}}_{\gamma }\) exist as elements of \(C([0,T];L^p({\varOmega };H))\). In particular, \({{\widetilde{Z}}}_{\gamma -\delta }\) belongs to \(L^p(0,T; L^p({\varOmega }; H))\), hence to \(L^p({\varOmega }; L^p(0,T;H))\) by Fubini’s theorem. For this reason, there exists a set \({\varOmega }_0\in {\mathcal {F}}\) with \({\mathbb {P}}({\varOmega }_0)=0\) such that \({{\widetilde{Z}}}_{\gamma -\delta } (\,\cdot \,, \omega ) \in L^p(0,T;H)\) for all \(\omega \in {\varOmega }_0^c = {\varOmega }{\setminus }{\varOmega }_0\). We recall the linear operator \({\mathscr {B}}_{\delta ,p} :L^p(0,T;H) \rightarrow C_{0,\{0\}}([0,T];H)\) from (3.10) and claim that the process \({\widehat{Z}}_\gamma \) defined for \(t\in [0,T]\) and \(\omega \in {\varOmega }\) by

$$\begin{aligned} {\widehat{Z}}_\gamma (t, \omega ):= {\left\{ \begin{array}{ll} \bigl [{\mathscr {B}}_{\delta ,p} {{\widetilde{Z}}}_{\gamma -\delta }\bigr ](t, \omega ) &{}\text {if }(t,\omega ) \in [0,T]\times {\varOmega }_0^c,\\ 0 &{}\text {if }(t,\omega ) \in [0,T]\times {\varOmega }_0, \end{array}\right. } \end{aligned}$$

is the desired continuous modification of \({{\widetilde{Z}}}_\gamma \). To this end, firstly note that for all \(\omega \in {\varOmega }\) the mapping \(t \mapsto {\widehat{Z}}_\gamma (t, \omega )\) indeed is continuous and \(\widehat{Z}_\gamma \) belongs to \(L^p(\Omega ; C([0,T];H))\); this follows from Proposition 3.4 since \(\delta \in ( {1}/{p}, \infty )\). In order to show that \({\widehat{Z}}_\gamma \) is a modification of \({\widetilde{Z}}_\gamma \), we fix \(t \in [0,T]\) and employ formulas (3.10) and (3.13) along with the semigroup property to obtain

$$\begin{aligned}&{\widehat{Z}}_{\gamma }(t) = \bigl [{\mathscr {B}}_{\delta ,p} {{\widetilde{Z}}}_{\gamma -\delta }\bigr ](t) = \frac{1}{{\varGamma }(\delta )} \int _0^t (t-s)^{\delta -1} S(t-s) {{\widetilde{Z}}}_{\gamma -\delta }(s) \mathop {}\!\textrm{d}s \nonumber \\&\quad = \frac{1}{{\varGamma }(\delta ){\varGamma }(\gamma -\delta )} \int _0^t (t-s)^{\delta -1} S(t-s) \biggl [ \int _0^s (s-r)^{\gamma -\delta -1}S(s-r)\mathop {}\!\textrm{d}W^Q(r) \biggr ] \mathop {}\!\textrm{d}s \nonumber \\&\quad = \frac{1}{{\varGamma }(\delta ){\varGamma }(\gamma -\delta )} \int _0^t \int _0^s (t-s)^{\delta -1} (s-r)^{\gamma -\delta -1} S(t-r) \mathop {}\!\textrm{d}W^Q(r) \mathop {}\!\textrm{d}s , \quad {\mathbb {P}}\text {-a.s.} \end{aligned}$$
(3.15)

We set \({{\widetilde{M}}}_T:= \sup _{t\in [0,T]} \Vert S(t)\Vert _{{\mathscr {L}}(H)}\), \(K_T:= \int _0^T \bigl \Vert t^{\gamma -1-\delta } S(t) Q^{ \frac{1}{2} } \bigr \Vert _{{\mathscr {L}}_2(H)}^2 \mathop {}\!\textrm{d}t\) and find

$$\begin{aligned}&\int _0^t \biggl [ \int _0^s \bigl \Vert (t-s)^{\delta -1} (s-r)^{\gamma -\delta -1} S(t-r)Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}_2(H)}^2 \mathop {}\!\textrm{d}r \biggr ]^{{1}/{2}} \mathop {}\!\textrm{d}s\\&\quad \le {{\widetilde{M}}}_T \int _0^t (t-s)^{\delta -1} \biggl [ \int _0^s \bigl \Vert (s-r)^{\gamma -\delta -1} S(s-r)Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}_2(H)}^2 \mathop {}\!\textrm{d}r \biggr ]^{{1}/{2}} \mathop {}\!\textrm{d}s\\&\quad = {{\widetilde{M}}}_T \int _0^t (t-s)^{\delta -1} \biggl [ \int _0^s \bigl \Vert r^{\gamma -1-\delta } S(r)Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}_2(H)}^2 \mathop {}\!\textrm{d}r \biggr ]^{{1}/{2}} \mathop {}\!\textrm{d}s \le \frac{{{\widetilde{M}}}_T T^\delta \sqrt{ K_T } }{ \delta } < \infty . \end{aligned}$$

This estimate shows that

$$\begin{aligned}{} & {} s\mapsto {\textbf{1}}_{(0,t)}(s) {\textbf{1}}_{(0,s)}(\,\cdot \,) (t-s)^{\delta -1} (s-\,\cdot \,)^{\gamma -\delta -1} S(t-\,\cdot \,)Q^{\frac{1}{2}} \\{} & {} \quad \in L^1(0,T;L^2(0,T;{\mathscr {L}}_2(H))), \end{aligned}$$

and the stochastic Fubini theorem [63, Theorem 8.14] may be used in (3.15), yielding

$$\begin{aligned} {{\widehat{Z}}}_{\gamma }(t) = \frac{1}{{\varGamma }(\delta ){\varGamma }(\gamma -\delta )} \int _0^t \biggl [\int _r^t (t-s)^{\delta -1} (s-r)^{\gamma -\delta -1} \mathop {}\!\textrm{d}s\biggr ] S(t-r) \mathop {}\!\textrm{d}W^Q(r), \quad {\mathbb {P}}\text {-a.s.} \end{aligned}$$

Using the change of variables \(u(s):= \frac{s-r}{t-r}\) and [59, Formula 5.12.1], we derive that

$$\begin{aligned} (t-r)^{1-\gamma }\int _r^t (t-s)^{\delta -1} (s-r)^{\gamma -\delta -1} \mathop {}\!\textrm{d}s = \int _0^1 (1-u)^{\delta -1} u^{\gamma -\delta -1} \mathop {}\!\textrm{d}u = \frac{{\varGamma }(\gamma -\delta ){\varGamma }(\delta )}{{\varGamma }(\gamma )}, \end{aligned}$$

which shows that \({\widehat{Z}}_{\gamma }(t) = {\widetilde{Z}}_{\gamma }(t)\) holds \({\mathbb {P}}\)-a.s. Since \(t \in [0,T]\) was arbitrary this implies that \({\widehat{Z}}_{\gamma }\) is a modification of \({\widetilde{Z}}_{\gamma }\) and completes the proof for \(p \in ( {1}/{\delta } \vee 1,\infty )\). The case \(p\in [1, {1}/{\delta } \vee 1 ]\) follows from the nestedness of the \(L^p({\varOmega };C([0,T];H))\) spaces. \(\square \)

In order to provide a more rigorous justification for the Definition 3.7 of a mild solution to (3.1), we proceed as follows: We seek a further suitable solution concept of a weak solution, which follows “naturally” from (3.1) using \(L^2(0,T;H)\) inner products, and show that weak and mild solutions are equivalent.

For this, we first define the weak stochastic Itô integral for \(f:(0,T)\rightarrow {\mathscr {L}}(H)\) and \(g:(0,T) \rightarrow H\) by

$$\begin{aligned} \int _0^t \bigl ( f(s) \mathop {}\!\textrm{d}W^Q(s), g(s) \bigr )_{H}:= \int _0^t {\widetilde{f}}_g(s) \mathop {}\!\textrm{d}W^Q(s), \quad t \in [0,T], \end{aligned}$$

where \(\int _0^T \bigl \Vert Q^{\frac{1}{2}} [f(s)]^* g(s) \bigr \Vert _{H}^2 \mathop {}\!\textrm{d}s < \infty \) and \({\widetilde{f}}_g :(0,T) \rightarrow {\mathscr {L}}(H;{\mathbb {R}})\) is defined by

$$\begin{aligned} {\widetilde{f}}_g(s) x:= ( f(s) x, g(s) )_{H} \quad \forall x \in H, \quad \forall s\in (0,T), \end{aligned}$$

cf. [52, Lemma 2.4.2].

Definition 3.9

Let Assumption 3.1(i) hold and let \(\gamma \in (0,\infty )\). A predictable H-valued stochastic process \(Y_\gamma :=(Y_\gamma (t))_{t\in [0,T]}\) is called a weak solution to (3.1) if it is mean-square continuous and, in addition,

$$\begin{aligned} \forall \psi \in {\textsf{D}}({\mathcal {B}}^{\gamma *}): \quad ( Y_\gamma , {\mathcal {B}}^{\gamma *}\psi )_{L^2(0,T;H)} = \int _0^T \bigl ( \mathop {}\!\textrm{d}W^Q(t), \psi (t) \bigr )_H, \quad {\mathbb {P}}\text {-a.s.} \end{aligned}$$
(3.16)

Remark 3.10

For \(\gamma =1\), a natural weak solution concept is the formulation given in [63, Definition 9.11]: A predictable H-valued process \((Y_1(t))_{t\in [0,T]}\) is a weak solution to (3.1) if \(\sup _{t\in [0,T]} \Vert Y_1(t) \Vert _{L^2({\varOmega };H)} < \infty \) and, for all \(t \in [0,T]\) and \(y \in {\textsf{D}}(A^*)\),

$$\begin{aligned} (Y_1(t), y)_H = - \int _0^t (Y_1(s), A^* y)_H \mathop {}\!\textrm{d}s + \bigl ( W^Q(t), y \bigr )_H, \quad {\mathbb {P}}\text {-a.s.} \end{aligned}$$

Provided that Assumption 3.1(i) and (3.14) are satisfied, by [63, Theorem 9.15] an H-valued stochastic process is a weak solution in this sense if and only if it is a mild solution in the sense of Definition 3.7 with \(\gamma =1\).

In the next proposition we generalize this result to an arbitrary fractional power \(\gamma \) and show that, under the same conditions, the mild solution in the sense of Definition 3.7 is equivalent to the weak solution in the sense of Definition 3.9.

Proposition 3.11

Suppose that Assumption 3.1(i) holds and let \(\gamma \in (0,\infty )\) be such that (3.14) is satisfied. Then, a stochastic process is a mild solution in the sense of Definition 3.7 if and only if it is a weak solution in the sense of Definition 3.9. Moreover, mild and weak solutions are unique up to modification. If one requires continuity of the sample paths, mild and weak solutions are unique up to indistinguishability.

Proof

First, we show that a mild solution \(Z_\gamma \) is a weak solution. Note that mean-square continuity follows from Theorem 3.8. Fix an arbitrary \(\psi \in {\textsf{D}}({\mathcal {B}}^{\gamma *})\). Then,

$$\begin{aligned} ( Z_\gamma&, {\mathcal {B}}^{\gamma *}\psi )_{L^2(0,T;H)} = \frac{1}{{\varGamma }(\gamma )} \int _0^T \biggl (\int _0^t (t-s)^{\gamma -1} S(t-s) \mathop {}\!\textrm{d}W^Q(s), [{\mathcal {B}}^{\gamma *}\psi ](t) \biggr )_H \mathop {}\!\textrm{d}t \nonumber \\&= \frac{1}{{\varGamma }(\gamma )} \int _0^T \int _0^T \bigl ( {\textbf{1}}_{(0,t)}(s) (t-s)^{\gamma -1}S(t-s) \mathop {}\!\textrm{d}W^Q(s), [{\mathcal {B}}^{\gamma *}\psi ](t) \bigr )_H \mathop {}\!\textrm{d}t \end{aligned}$$
(3.17)

holds \({\mathbb {P}}\)-a.s. Here, we used that \(\bigl ( \int _0^T f(s) \mathop {}\!\textrm{d}W^Q(s), x \bigr )_H = \int _0^T (f(s) \mathop {}\!\textrm{d}W^Q(s), x)_H\) for all \(f :(0,T) \rightarrow {\mathscr {L}}(H)\) and \(x \in H\), which readily is derived from the definition of the weak stochastic integral and the continuity of inner products. We now would like to apply the stochastic Fubini theorem, see e.g. [63, Theorem 8.14], in order to interchange the inner weak stochastic integral and the outer deterministic integral. Again by the definition of the weak stochastic integral we have, for a.a. \(t\in (0,T)\),

$$\begin{aligned} \int _0^T \bigl ( {\textbf{1}}_{(0,t)}(s) (t-s)^{\gamma -1} S(t-s) \mathop {}\!\textrm{d}W^Q(s), [{\mathcal {B}}^{\gamma *}\psi ](t) \bigr )_H = \int _0^T {\varPsi }(s,t) \mathop {}\!\textrm{d}W^Q(s), \quad {\mathbb {P}}\text {-a.s.}, \end{aligned}$$

where the integrand \({\varPsi }(s,t):H \rightarrow {\mathbb {R}}\) is deterministic and, for \(s,t \in (0,T)\), defined by

$$\begin{aligned} {\varPsi }(s,t) x:= \bigl ( {\textbf{1}}_{(0,t)}(s) (t-s)^{\gamma -1} S(t-s) x, [{\mathcal {B}}^{\gamma *}\psi ](t) \bigr )_H \quad \forall x\in H. \end{aligned}$$
(3.18)

Thus, the usage of the stochastic Fubini theorem is justified if \(t\mapsto {\varPsi }(\,\cdot , t)Q^{\frac{1}{2}}\) is in \(L^1(0,T; L^2(0,T;{\mathscr {L}}_2(H;{\mathbb {R}})))\). Given an orthonormal basis \((g_j)_{j\in {\mathbb {N}}}\) for H, we obtain

$$\begin{aligned} \bigl \Vert {\varPsi }(s,t) Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}_2(H;{\mathbb {R}})}^2&= \sum _{j=1}^\infty \bigl | \bigl ( {\textbf{1}}_{(0,t)}(s) (t-s)^{\gamma -1}S(t-s) Q^\frac{1}{2} g_j, [{\mathcal {B}}^{\gamma *}\psi ](t) \bigr )_H \bigr |^2\\&\le \bigl \Vert {\textbf{1}}_{(0,t)}(s) (t-s)^{\gamma -1} S(t-s) Q^\frac{1}{2} \bigr \Vert _{{\mathscr {L}}_2(H)}^2 \bigl \Vert [{\mathcal {B}}^{\gamma *}\psi ](t) \bigr \Vert _{H}^2 \end{aligned}$$

by the Cauchy–Schwarz inequality on H. From this, it follows that

$$\begin{aligned}&\bigl \Vert t\mapsto {\varPsi }(\,\cdot \,,t) Q^{\frac{1}{2}} \bigr \Vert _{ L^1(0,T;L^2(0,T;{\mathscr {L}}_2(H;{\mathbb {R}})))} = \int _0^T \biggl ( \int _0^T \bigl \Vert {\varPsi }(s,t) Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}_2(H;{\mathbb {R}})}^2 \mathop {}\!\textrm{d}s \biggr )^{{1}/{2}} \mathop {}\!\textrm{d}t\\&\quad \le \int _0^T \biggl ( \int _0^t \bigl \Vert (t-s)^{\gamma -1}S(t-s) Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}_2(H)}^2 \bigl \Vert [{\mathcal {B}}^{\gamma *}\psi ](t) \bigr \Vert _{H}^2 \mathop {}\!\textrm{d}s \biggr )^{{1}/{2}} \mathop {}\!\textrm{d}t\\&\quad = \int _0^T \biggl (\int _0^t \bigl \Vert s^{\gamma -1}S(s)Q^\frac{1}{2} \bigr \Vert _{{\mathscr {L}}_2(H)}^2 \mathop {}\!\textrm{d}s\biggr )^{{1}/{2}} \bigl \Vert [{\mathcal {B}}^{\gamma *}\psi ](t) \bigr \Vert _{H} \mathop {}\!\textrm{d}t\\&\quad \le T^{{1}/{2}} \Vert {\mathcal {B}}^{\gamma *}\psi \Vert _{L^2(0,T;H)} \biggl ( \int _0^T \bigl \Vert s^{\gamma -1} S(s) Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}_2(H)}^2 \mathop {}\!\textrm{d}s \biggr )^{ {1}/{2} } < \infty , \end{aligned}$$

where we used the Cauchy–Schwarz inequality on \(L^2(0,T)\) in the last step. Owing to (3.14), the integral in the final expression is finite. Applying the stochastic Fubini theorem to (3.17), taking adjoints in (3.18) and using the continuity of \((\,\cdot ,\,\cdot \,)_H\) gives

$$\begin{aligned}&(Z_\gamma , {\mathcal {B}}^{\gamma *}\psi )_{L^2(0,T;H)} = \frac{1}{{\varGamma }(\gamma )} \int _0^T\int _0^T {\varPsi }(s,t) \mathop {}\!\textrm{d}t\mathop {}\!\textrm{d}W^Q(s) \\&\quad = \int _0^T \biggl ( \mathop {}\!\textrm{d}W^Q(s), \frac{1}{{\varGamma }(\gamma )} \int _s^T (t-s)^{\gamma -1}[S(t-s)]^*[{\mathcal {B}}^{\gamma *}\psi ](t) \mathop {}\!\textrm{d}t\biggr )_H \\&\quad = \int _0^T \bigl ( \mathop {}\!\textrm{d}W^Q(s), [{\mathcal {B}}^{-\gamma *}{\mathcal {B}}^{\gamma *}\psi ](s) \bigr )_H = \int _0^T \bigl ( \mathop {}\!\textrm{d}W^Q(s), \psi (s) \bigr )_H, \qquad {\mathbb {P}}\text {-a.s.}, \end{aligned}$$

where we used (3.12) in the third line. Therefore, \(Z_\gamma \) is a weak solution.

Conversely, suppose that \(Y_\gamma \) is a weak solution, let an arbitrary \(\phi \in L^2(0,T;H)\) be given and set \(\psi := {\mathcal {B}}^{-\gamma *} \phi \in {\textsf{D}}({\mathcal {B}}^{\gamma *})\). Substituting this into (3.16) gives

$$\begin{aligned} ( Y_\gamma , \phi )_{L^2(0,T;H)} = \int _0^T \bigl ( \mathop {}\!\textrm{d}W^Q(t), [{\mathcal {B}}^{-\gamma *} \phi ](t) \bigr )_H, \quad {\mathbb {P}}\text {-a.s.} \end{aligned}$$

Let \(({\widetilde{Z}}_\gamma (t))_{t\in [0,T]}\) be the stochastic convolution in (3.13). Since the condition for the stochastic Fubini theorem still holds after replacing \({\mathcal {B}}^{\gamma *}\psi \) by \(\phi \) in (3.18), the proof of the previous implication can be read backwards to see that

$$\begin{aligned} \forall \phi \in L^2(0,T;H): \quad {\mathbb {P}}\bigl ( ( Y_\gamma , \phi )_{L^2(0,T;H)} = ( {\widetilde{Z}}_\gamma , \phi )_{L^2(0,T;H)} \bigr ) =1. \end{aligned}$$

By separability of H, also \({\mathbb {P}}( Y_\gamma = {\widetilde{Z}}_\gamma \) in \(L^2(0,T;H) ) = 1\) holds so that by Fubini’s theorem \(Y_\gamma = {\widetilde{Z}}_\gamma \) in \(L^2(0,T;L^2({\varOmega };H))\) follows. Since both \(Y_\gamma \) and \({\widetilde{Z}}_\gamma \) are mean-square continuous, this shows that, for all \(t\in [0,T]\), \(Y_\gamma (t) = {\widetilde{Z}}_\gamma (t)\) in \(L^2({\varOmega };H)\). Therefore, for all \(t\in [0,T]\), we have that \(Y_\gamma (t) = {\widetilde{Z}}_\gamma (t)\), \({\mathbb {P}}\)-a.s., i.e., \(Y_\gamma \) is a mild solution.

It thus suffices to prove uniqueness only for mild solutions. By Definition 3.7, mild solutions are modifications of the stochastic convolution \({{\widetilde{Z}}}_\gamma \) in (3.13), hence of each other. If two mild solutions are moreover known to have continuous sample paths, then they are indistinguishable by [63, Proposition 3.17]. \(\square \)

3.3 Spatiotemporal regularity of solutions

We now investigate spatiotemporal regularity of the mild solution \(Z_\gamma \) in Definition 3.7. We start by stating our main results, Theorem 3.12 and Corollary 3.13, in Sect. 3.3.1. In Sect. 3.3.2 we derive a simplified condition for spatiotemporal regularity, which is easier to check in applications and sufficient whenever A satisfies Assumptions 3.1(i),(iii),(iv), see Proposition 3.14. In addition, we explicitly discuss the setting of a Gelfand triple \(V\hookrightarrow H \cong H^* \hookrightarrow V^*\) in the case that the operator A is induced by a (not necessarily symmetric) bilinear form \({\mathfrak {a}} :V\times V\rightarrow {\mathbb {R}}\) which is continuous and satisfies a Gårding inequality. Section 3.3.3 is devoted to the proof of Theorem 3.12.

3.3.1 Main results

In Theorem 3.12 below, the temporal regularity of the (weak or mild) solution is measured by the differentiability \(n\in {\mathbb {N}}_0\) as well as the Hölder exponent \(\tau \in [0,1)\). Spatial regularity is expressed by means of vector spaces which are defined in terms of fractional powers of A (see Sect. B.2.2 in “Appendix B”) as follows:

$$\begin{aligned} \dot{H}_A^\sigma := {\textsf{D}}\bigl ( A^{{\sigma }/{2}} \bigr ), \qquad (x, y)_{\dot{H}_A^\sigma }:= \bigl ( A^{{\sigma }/{2}} x, A^{{\sigma }/{2}} y \bigr )_H, \quad \sigma \in [0,\infty ). \end{aligned}$$

For \(\sigma \in (0,\infty )\), \(\dot{H}_A^\sigma \) is a Hilbert space provided that Assumptions 3.1(i),(ii),(iv) are satisfied. In this case, we have the embeddings \({\dot{H}}^{\sigma '}_A \hookrightarrow {\dot{H}}^{\sigma }_A\hookrightarrow H\) for all \(\sigma '\ge \sigma \ge 0\). Note, in particular, that we do not need to assume that A is self-adjoint.

Theorem 3.12

Suppose that Assumptions 3.1(i),(ii) are satisfied and let \(n \in {\mathbb {N}}_0\), \(\sigma \in [0,\infty )\) and \(\gamma \in \bigl ( \frac{\sigma -r}{2} +n,\infty \bigr )\), where \(r\in [0,\sigma ]\) is such that \(Q^{\frac{1}{2}} \in {\mathscr {L}}(H;\dot{H}_A^r)\). In the case that \(\sigma \in (0,\infty )\), suppose furthermore that Assumption 3.1(iv) is fulfilled. Under the condition

$$\begin{aligned} \int _0^T \bigl \Vert t^{\gamma -1-n} S(t) Q^{\frac{1}{2}} \bigr \Vert _{ {\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A)}^2 \mathop {}\!\textrm{d}t < \infty , \end{aligned}$$
(3.19)

the mild solution \(Z_\gamma \) (or, equivalently, the weak solution \(Y_\gamma \)) in the sense of Definition 3.7 (or 3.9) belongs to \(C^{n,0}([0,T];L^p({\varOmega };\dot{H}^\sigma _A))\) for every \(p \in [1,\infty )\).

If additionally \(\gamma \ge n + \tau + \frac{1}{2}\) and \(A^{n+\tau +\frac{1}{2} -\gamma } Q^{\frac{1}{2}} \in {\mathscr {L}}_2(H; \dot{H}_A^\sigma )\) are satisfied for some \(\tau \in (0,1)\), then we have \(Z_\gamma \in C^{n,\tau }([0,T];L^p({\varOmega };\dot{H}^\sigma _A))\) for every \(p \in [1,\infty )\).

An application of the Kolmogorov–Chentsov continuity theorem, see e.g. [20, Theorem 3.9], allows us to (partially) transport the temporal regularity result of Theorem 3.12 to the pathwise setting, as seen in the next corollary.

Corollary 3.13

Suppose that Assumptions 3.1(i),(ii) are satisfied. Let \(\sigma \in [0,\infty )\), \(r\in [0,\sigma ]\), \(\gamma \in \bigl (\frac{\sigma -r}{2},\infty \bigr )\) and \(\tau \in (0,1)\) be such that \(Q^{\frac{1}{2}} \in {\mathscr {L}}(H;\dot{H}_A^r)\) and \(\gamma \ge \tau + \frac{1}{2}\). If \({\sigma \in (0,\infty )}\), suppose also that Assumption 3.1(iv) holds. If the condition

$$\begin{aligned} \bigl \Vert A^{\tau +\frac{1}{2}-\gamma } Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}_2(H;\dot{H}_A^\sigma )} + \int _0^T \bigl \Vert t^{\gamma -1} S(t) Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A)}^2 \mathop {}\!\textrm{d}t < \infty \end{aligned}$$

is satisfied, then for all \(p \in [1,\infty )\) and every \(\tau '\in [0,\tau )\) there exists a modification \({\widehat{Z}}_\gamma \) of the mild solution \(Z_\gamma \) (or, equivalently, of the weak solution \(Y_\gamma \)) in the sense of Definition 3.7 (or 3.9) such that \({\widehat{Z}}_\gamma \) has \(\tau '\)-Hölder continuous sample paths and belongs to \(L^p\bigl ( {\varOmega }; C^{0,\tau '}([0,T]; \dot{H}^\sigma _A) \bigr )\).

Proof

We first invoke Theorem 3.12 with \(n = 0\) and \(\tau \in (0,1)\) to establish that \(Z_\gamma \) belongs to \(C^{0,\tau }([0,T]; L^q({\varOmega };\dot{H}_A^\sigma ))\) for every \(q \in [1,\infty )\). The result then follows by choosing \(q \ge 1\) sufficiently large, applying the Kolmogorov–Chentsov continuity theorem (see e.g. [20, Theorem 3.9]), and using nestedness of the \(L^p\) spaces. \(\square \)

3.3.2 A simplified condition and its application to the Gårding inequality case

Whenever also Assumption 3.1(iii) holds, it is possible to replace the condition (3.19) by one which is simpler to check in practice. In this case, the operator A satisfies square function estimates (see Sect. B.2.3 in “Appendix B”), one of which is used to prove the next result.

Proposition 3.14

Let Assumptions 3.1(i),(iii),(iv) be satisfied. Suppose that the constants \(\sigma , \delta \in [0, \infty )\) and \(\gamma \in \bigl (\frac{1}{2}+\delta ,\infty \bigr ) \cap \bigl [\frac{1}{2}+\delta +\frac{\sigma -r}{2},\infty \bigr )\) are given, where \(r \in [0,\sigma ]\) is taken such that \(Q^{\frac{1}{2}} \in {\mathscr {L}}(H; \dot{H}_A^r)\). Then,

$$\begin{aligned} \int _0^\infty \bigl \Vert t^{\gamma -1-\delta } S(t) Q^{\frac{1}{2}} \bigr \Vert _{ {\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A)}^2 \mathop {}\!\textrm{d}t \eqsim _{(\gamma ,\delta )} \bigl \Vert A^{ \delta + \frac{1}{2} - \gamma } Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}_2(H;{\dot{H}}^{\sigma }_A)}^2. \end{aligned}$$

Proof

Applying Lemma B.7, see “Appendix B”, with \(a:= \gamma - \delta - \frac{1}{2} \in (0,\infty )\) and \(x:= A^{ \frac{\sigma }{2}+\delta +\frac{1}{2}-\gamma }Q^\frac{1}{2} y \in H\) for \(y\in H\) shows that

$$\begin{aligned} \int _0^\infty \bigl \Vert t^{\gamma -1-\delta } A^{ \gamma -\delta -\frac{1}{2} } S(t) A^{ \delta +\frac{1}{2}-\gamma } Q^\frac{1}{2} y \bigr \Vert _{{\dot{H}}^{\sigma }_A}^2 \mathop {}\!\textrm{d}t \eqsim _{(\gamma ,\delta )} \bigl \Vert A^{ \delta +\frac{1}{2}-\gamma } Q^\frac{1}{2} y \bigr \Vert _{{\dot{H}}^{\sigma }_A}^2 \quad \forall y \in H. \end{aligned}$$

Summing both sides over an orthonormal basis for H and using the Fubini–Tonelli theorem to interchange integration and summation on the left-hand side yields the desired conclusion. \(\square \)

Remark 3.15

Proposition 3.14 shows that under the additional assumption that \(A_{\mathbb {C}}\) admits a bounded \(H^\infty \)-calculus with \(\omega _{H^\infty }(A_{\mathbb {C}})<\tfrac{\pi }{2}\), which e.g. is satisfied whenever A is self-adjoint and strictly positive, it suffices to check that \(\gamma > n + \frac{ (\sigma -r) \vee 1 }{2} \) and \(\gamma \ge n+\frac{1+ (\sigma -r)\vee (2\tau ) }{2}\) and that the Hilbert–Schmidt norm \(\Vert A^{ n+ \tau + \frac{1}{2} - \gamma } Q^{\frac{1}{2}} \Vert _{{\mathscr {L}}_2(H;{\dot{H}}^{\sigma }_A)}\) is bounded to conclude the regularity results of Theorem 3.12. This condition coincides with the one imposed in [46, Section 4, Theorem 6] to derive regularity in the non-fractional case \(\gamma =1\) for \(p=2\), \(\sigma =0\), \(n=0\) and \(\tau \in [0, {1}/{2}]\).

Corollary 3.16

Let \(\delta \in [0,\infty )\) and \(\gamma \in \bigl (\frac{1}{2}+\delta , \infty \bigr )\). Suppose that A satisfies Assumption 3.1(i) and that there exists a constant \(\eta \in [0,\infty )\) such that \({{\widehat{A}}}:= A + \eta I\) satisfies Assumptions 3.1(i),(iii),(iv) and \({{\widehat{A}}}^{\delta + \frac{1}{2}-\gamma } Q^{\frac{1}{2}} \in {\mathscr {L}}_2(H)\). Then, the mild solution \(Z_\gamma \) in the sense of Definition 3.7 exists and belongs to \(C([0,T];L^p({\varOmega };H))\) for every \(p\in [1,\infty )\). If \(\delta >0\), then for every \(p \in [1,\infty )\) there exists a modification of \(Z_\gamma \) in \(L^p({\varOmega };C([0,T];H))\) which has continuous sample paths.

Proof

Note that \(S(t) = e^{\eta t} {{\widehat{S}}}(t)\) holds for every \(t\ge 0\), where \(( {{\widehat{S}}}(t) )_{t\ge 0}\) denotes the \(C_0\)-semigroup generated by \(-{{\widehat{A}}}\). Hence, by Proposition 3.14 we find that

$$\begin{aligned} \int _0^T \bigl \Vert t^{\gamma -1-\delta } S(t) Q^{\frac{1}{2}} \bigr \Vert _{ {\mathscr {L}}_2(H)}^2 \mathop {}\!\textrm{d}t&\le e^{2 \eta T} \int _0^T \bigl \Vert t^{\gamma -1-\delta } {{\widehat{S}}}(t) Q^{\frac{1}{2}} \bigr \Vert _{ {\mathscr {L}}_2(H)}^2 \mathop {}\!\textrm{d}t\\&\lesssim _{(\gamma ,\delta )} e^{2 \eta T} \bigl \Vert {{\widehat{A}}}^{ \delta +\frac{1}{2}-\gamma } Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}_2(H)}^2 < \infty . \end{aligned}$$

The claim now follows from Theorem 3.8. \(\square \)

We illustrate the utility of Corollary 3.16 in the following example. It is concerned with the case that the operator A is induced by a bounded bilinear form \({\mathfrak {a}} :V\times V\rightarrow {\mathbb {R}}\), where \(V\hookrightarrow H\) is dense in H, and \({\mathfrak {a}}\) is not necessarily coercive on V; see also [37, Section 7.3.2]. We note that this setting applies to a variety of important applications, including symmetric and non-symmetric differential operators of even orders.

Example 3.17

Let \((V, (\,\cdot , \,\cdot \,)_V)\) be a Hilbert space which is densely and continuously embedded in H. Suppose that \(A :{\textsf{D}}(A)\subseteq H\rightarrow H\) is induced by a bounded bilinear form \({\mathfrak {a}} :V\times V\rightarrow {\mathbb {R}}\) which satisfies a Gårding inequality, i.e., there exist constants \(\alpha _0, \alpha _1 \in (0,\infty )\) and \(\eta \in [0,\infty )\) such that

$$\begin{aligned} \qquad \quad |{\mathfrak {a}}(u,v)|&\le \alpha _1 \Vert u\Vert _{V}\Vert v\Vert _{V}{} & {} \forall u,v\in V, \qquad \quad \end{aligned}$$
(3.20)
$$\begin{aligned} \qquad \quad {\mathfrak {a}}(u,u)&\ge \alpha _0 \Vert u\Vert _{V}^2 - \eta \Vert u\Vert _{H}^2{} & {} \forall u \in V. \qquad \quad \end{aligned}$$
(3.21)

The Gårding inequality (3.21) can be interpreted as coercivity of the bilinear form \(\hat{{\mathfrak {a}}}(u,v):= {\mathfrak {a}}(u,v) + \eta (u,v)_H\) on V, associated with \({{\widehat{A}}} = A + \eta I\), while (3.20) implies that \(\hat{{\mathfrak {a}}}\) is bounded. The complexified sesquilinear form \(\hat{{\mathfrak {a}}}_{{\mathbb {C}}} :V_{\mathbb {C}}\times V_{\mathbb {C}}\rightarrow {\mathbb {C}}\), which is defined analogously to (B.2) and induces the operator \({{\widehat{A}}}_{\mathbb {C}}\), inherits the boundedness and coercivity from \(\hat{{\mathfrak {a}}}\). Thus, there exist \({\widehat{\alpha }}_0, {\widehat{\alpha }}_1 \in (0,\infty )\) such that

$$\begin{aligned} |\hat{{\mathfrak {a}}}_{\mathbb {C}}(u,v)|&\le {\widehat{\alpha }}_1 \Vert u\Vert _{V_{\mathbb {C}}}\Vert v\Vert _{V_{\mathbb {C}}}{} & {} \forall u,v\in V_{\mathbb {C}}, \\ {\text {Re}}{\hat{{\mathfrak {a}}}_{\mathbb {C}}(u,u)}&\ge {\widehat{\alpha }}_0 \Vert u\Vert _{V_{\mathbb {C}}}^2{} & {} \forall u \in V_{\mathbb {C}}. \end{aligned}$$

Therefore, \({\widehat{\alpha }}_0 \Vert u\Vert _{V_{\mathbb {C}}}^2 \le {\text {Re}}{\hat{{\mathfrak {a}}}_{\mathbb {C}}(u,u)} \le |\hat{{\mathfrak {a}}}_{\mathbb {C}}(u,u)| \le {\widehat{\alpha }}_1 \Vert u\Vert _{V_{\mathbb {C}}}^2 \le \tfrac{{\widehat{\alpha }}_1}{{\widehat{\alpha }}_0} {\text {Re}}{\hat{{\mathfrak {a}}}_{\mathbb {C}}(u,u)}\) follows for every \(u \in V_{\mathbb {C}}\). If \(V_{\mathbb {C}}\ne \{0\}\), these estimates imply that \({\widehat{\alpha }}_0 \le {\widehat{\alpha }}_1\) and

$$\begin{aligned}{} & {} |{\text {Im}}{\hat{{\mathfrak {a}}}_{\mathbb {C}}}(u,u)|\\{} & {} \quad = \sqrt{ |\hat{{\mathfrak {a}}}_{\mathbb {C}}(u,u)|^2 - | {\text {Re}}{\hat{{\mathfrak {a}}}_{\mathbb {C}}}(u,u)|^2 } \le \Bigl ( \tfrac{{\widehat{\alpha }}_1^2}{{\widehat{\alpha }}_0^2} - 1 \Bigr )^{{1}/{2}} {\text {Re}}{\hat{{\mathfrak {a}}}_{\mathbb {C}}}(u,u) \quad \forall u \in V_{\mathbb {C}}. \end{aligned}$$

This shows that \(-{{\widehat{A}}}_{\mathbb {C}}\) generates a bounded analytic \(C_0\)-semigroup \(({\widehat{S}}_{\mathbb {C}}(t))_{t\ge 0}\) of contractions on \(H_{\mathbb {C}}\), cf. [60, Theorem 1.54], where we used that \((-\infty ,0) \subseteq \rho ({{\widehat{A}}}_{\mathbb {C}})\) by [60, Proposition 1.22]. Applying [42, Theorems 10.2.24 and 10.4.21] and using that \(\omega ({{\widehat{A}}}_{\mathbb {C}}) \in \bigl [0,\tfrac{\pi }{2} \bigr )\) because \(({\widehat{S}}_{\mathbb {C}}(t))_{t\ge 0}\) is bounded analytic (see Theorem B.2), we find that \({{\widehat{A}}}_{\mathbb {C}}\) admits a bounded \(H^\infty \)-calculus of angle \(\omega _{H^\infty }({{\widehat{A}}}_{\mathbb {C}}) = \omega ({{\widehat{A}}}_{\mathbb {C}}) \in \bigl [0,\tfrac{\pi }{2} \bigr )\). Thus, we are in the setting of Corollary 3.16. In particular, the existence of a mean-square continuous mild solution to (3.1) for \(\gamma > \frac{1}{2}\) follows if \(\Vert {{\widehat{A}}}^{\frac{1}{2}-\gamma } Q^{\frac{1}{2}} \Vert _{{\mathscr {L}}_2(H)}<\infty \).

3.3.3 The proof of Theorem 3.12

We split the proof of Theorem 3.12 into several intermediate results. Before stating and proving these, we introduce the following function, which generalizes the integrand in (3.13) used to define mild solutions. Given \(a\in {\mathbb {R}}\), \(b \in [0,\infty )\) and \(\sigma \in [0,\infty )\), define \({\varPhi }_{a,b}:(0,\infty ) \rightarrow {\mathscr {L}}(H; \dot{H}_A^\sigma )\) by

$$\begin{aligned} {\varPhi }_{a,b} (t):= t^{a} A^b S(t)Q^\frac{1}{2}, \qquad t\in (0,\infty ). \end{aligned}$$
(3.22)

Note that a mild solution \(Z_\gamma \) in the sense of Definition 3.7 satisfies the relation

$$\begin{aligned} \forall t\in [0,T]: \quad Z_\gamma (t) = \frac{1}{{\varGamma }(\gamma )} \int _0^t {\varPhi }_{\gamma -1,0} (t-s) \mathop {}\!\textrm{d}{\widehat{W}}(s), \quad {\mathbb {P}}\text {-a.s.}, \end{aligned}$$

where \({\widehat{W}}(t):=Q^{-\frac{1}{2}} W^Q(t)\), \(t\ge 0\), is a cylindrical Wiener process.

The first result quantifies spatial regularity of the continuous-in-time stochastic convolution with \({\varPhi }_{a,b}\) in \(L^p({\varOmega }; {\dot{H}}^{\sigma }_A)\)-sense. Recall from Sect. 2 that \((W(t))_{t\ge 0}\) denotes an (arbitrary) H-valued cylindrical Wiener process with respect to \(({\mathcal {F}}_t)_{t\ge 0}\).

Proposition 3.18

Let Assumption 3.1(i) hold. Suppose that the constants \(a \in {\mathbb {R}}\), \(b,\sigma \in [0,\infty )\) and \(T \in (0,\infty )\) are given. If \(\sigma \ne 0\), then suppose moreover that Assumptions 3.1(ii),(iv) are satisfied. If the function \({\varPhi }_{a,b}\) defined in (3.22) belongs to \(L^2(0,T; {\mathscr {L}}_2(H;{\dot{H}}^\sigma _A))\), i.e.,

$$\begin{aligned} \int _0^T \Vert {\varPhi }_{a,b}(t)\Vert _{{\mathscr {L}}_2(H;\dot{H}_A^\sigma )}^2 \mathop {}\!\textrm{d}t < \infty , \end{aligned}$$

then \(t\mapsto \int _0^t {\varPhi }_{a,b}(t-s)\mathop {}\!\textrm{d}W(s)\) belongs to \(C([0,T]; L^p({\varOmega };\dot{H}_A^\sigma ))\) for all \(p \in [1,\infty )\).

Proof

We first note that the assumption \({\varPhi }_{a,b} \in L^2(0,T;{\mathscr {L}}_2(H;{\dot{H}}^\sigma _A))\), combined with the Burkholder–Davis–Gundy inequality (see [52, Theorem 6.1.2]) and the continuous embedding

$$\begin{aligned} L^2({\varOmega };{\dot{H}}^{\sigma }_A) \hookrightarrow L^p({\varOmega };{\dot{H}}^{\sigma }_A), \qquad p\in [1,2), \; \sigma \in [0,\infty ), \end{aligned}$$
(3.23)

imply that \(\int _0^t {\varPhi }_{a,b}(t-s)\mathop {}\!\textrm{d}W(s)\) indeed is a well-defined element of \(L^p({\varOmega };\dot{H}_A^\sigma )\) for all \(t\in [0,T]\) and every \(p\in [1,\infty )\).

It remains to check the \(L^p({\varOmega }; {\dot{H}}^{\sigma }_A)\)-continuity of \(t \mapsto \int _0^t {\varPhi }_{a,b}(t-s)\mathop {}\!\textrm{d}W(s)\). For fixed \(t \in [0,T)\) and \(h \in (0, T-t]\), we split the stochastic integrals as follows:

$$\begin{aligned}&\int _0^{t+h} {\varPhi }_{a,b}(t+h-s)\mathop {}\!\textrm{d}W(s) - \int _0^t {\varPhi }_{a,b}(t-s)\mathop {}\!\textrm{d}W(s)\\&\quad = \int _t^{t+h} {\varPhi }_{a,b}(t+h-s) \mathop {}\!\textrm{d}W(s) + \int _0^{t} [{\varPhi }_{a,b}(t+h-s) - {\varPhi }_{a,b}(t-s)] \mathop {}\!\textrm{d}W(s). \end{aligned}$$

For \(p \in [2,\infty )\), the Burkholder–Davis–Gundy inequality yields

$$\begin{aligned}&\biggl \Vert \int _t^{t+h} {\varPhi }_{a,b}(t+h-s) \mathop {}\!\textrm{d}W(s) + \int _0^{t} [{\varPhi }_{a,b}(t+h-s) - {\varPhi }_{a,b}(t-s)] \mathop {}\!\textrm{d}W(s) \biggr \Vert _{L^p({\varOmega };{\dot{H}}^{\sigma }_A)}\\&\quad \lesssim _p \biggl [\int _t^{t+h} \Vert {\varPhi }_{a,b}(t+h-s)\Vert _{ {\mathscr {L}}_2(H;\dot{H}_A^\sigma )}^2 \mathop {}\!\textrm{d}s \biggr ]^{{1}/{2}}\\&\qquad + \biggl [ \int _0^t \Vert {\varPhi }_{a,b}(t+h-s) - {\varPhi }_{a,b}(t-s) \Vert _{ {\mathscr {L}}_2(H;\dot{H}_A^\sigma )}^2 \mathop {}\!\textrm{d}s \biggr ]^{{1}/{2}}\\&\quad = \biggl [ \int _0^h \Vert {\varPhi }_{a,b}(u)\Vert _{{\mathscr {L}}_2(H;\dot{H}_A^\sigma )}^2 \mathop {}\!\textrm{d}u \biggr ]^{{1}/{2}} + \biggl [ \int _0^t \Vert {\varPhi }_{a,b}(r+h) - {\varPhi }_{a,b}(r)\Vert _{ {\mathscr {L}}_2(H;\dot{H}_A^\sigma )}^2 \mathop {}\!\textrm{d}r \biggr ]^{{1}/{2}}, \end{aligned}$$

where \(u:= t+h-s\) and \(r:= t-s\). Since \({\varPhi }_{a,b}\in L^2(0,T;{\mathscr {L}}_2(H;\dot{H}_A^\sigma ))\) the first integral tends to zero as \(h \downarrow 0\) by dominated convergence. The second term tends to zero by Lemma A.4, see “Appendix A”.

For \(t\in (0,T]\) and \(h\in [-t, 0)\), the difference of stochastic integrals can be rewritten using \(\int _0^t = \int _0^{t+h} + \int _{t+h}^t\). Thus, we obtain, for every \(p\in [2,\infty )\), the bound

$$\begin{aligned}&\biggl \Vert \int _0^{t+h} {\varPhi }_{a,b}(t+h-s) \mathop {}\!\textrm{d}W(s) - \int _0^t {\varPhi }_{a,b}(t-s)\mathop {}\!\textrm{d}W(s) \biggr \Vert _{L^p({\varOmega };{\dot{H}}^{\sigma }_A)}\\&\quad \lesssim _p \biggl [ \int _0^{-h} \Vert {\varPhi }_{a,b}(r)\Vert _{{\mathscr {L}}_2(H;\dot{H}_A^\sigma )}^2 \mathop {}\!\textrm{d}r \biggr ]^{{1}/{2}} + \biggl [ \int _{-h}^t \Vert {\varPhi }_{a,b}(r+h) - {\varPhi }_{a,b}(r)\Vert _{{\mathscr {L}}_2(H;\dot{H}_A^\sigma )}^2 \mathop {}\!\textrm{d}r \biggr ]^{{1}/{2}}, \end{aligned}$$

where we again used the change of variables \(r:= t-s\). Both terms on the last line tend to zero, again by dominated convergence and Lemma A.4, respectively.

Finally, we note that the result for \(p=2\) implies that for \(p\in [1,2)\) by (3.23). \(\square \)

Furthermore, we obtain the following result regarding the temporal Hölder continuity of the stochastic convolution with the function \({\varPhi }_{a,b}\) in (3.22).

Proposition 3.19

Suppose that Assumptions 3.1(i),(ii) are fulfilled, let \(T \in (0,\infty )\), \({a \in \bigl (-\frac{1}{2},\infty \bigr )}\), \({b, \sigma \in [0,\infty )}\) and \(\tau \in \bigl ( 0,a + \frac{1}{2} \bigr ] \cap (0,1)\). If \(\sigma \ne 0\), then suppose also that Assumption 3.1(iv) holds. If \(A^{-a-\frac{1}{2} +b+\tau } Q^{\frac{1}{2}} \in {\mathscr {L}}_2(H; \dot{H}_A^\sigma )\) and \({\varPhi }_{a,b}\) is defined by (3.22), then \(t \mapsto \int _0^t {\varPhi }_{a,b}(t-s)\mathop {}\!\textrm{d}W(s)\) belongs to \(C^{0,\tau }([0,T];L^p({\varOmega };\dot{H}_A^\sigma ))\) for all \(p \in [1,\infty )\).

Proof

For \(t \in [0,T)\) and \(h \in (0,T-t]\), we obtain

$$\begin{aligned}&\biggl \Vert \int _0^{t+h} {\varPhi }_{a,b}(t+h -s) \mathop {}\!\textrm{d}W(s) - \int _0^t {\varPhi }_{a,b}(t-s)\mathop {}\!\textrm{d}W(s) \biggr \Vert _{L^p({\varOmega }; {\dot{H}}^{\sigma }_A)}\\&\quad \le \biggl \Vert \int _0^t \bigl [ {\varPhi }_{a,b}(t+h-s) - {\varPhi }_{a,b}(t-s) \bigr ] \mathop {}\!\textrm{d}W(s) \biggr \Vert _{L^p({\varOmega };{\dot{H}}^{\sigma }_A)} \\&\qquad + \biggl \Vert \int _t^{t+h} {\varPhi }_{a,b}(t+h-s) \mathop {}\!\textrm{d}W(s) \biggr \Vert _{L^p({\varOmega }; {\dot{H}}^{\sigma }_A)} \\&\quad \lesssim _{(p,a,\tau )} h^\tau \bigl \Vert A^{-a-\frac{1}{2} +b+\tau } Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}_2(H;{\dot{H}}^{\sigma }_A)} \end{aligned}$$

by Lemmas A.6 and A.7, see “Appendix A”. The analogous result for the case that \(t\in (0,T]\) and \(h\in [-t,0)\) follows upon splitting \(\int _0^t = \int _0^{t+h} + \int _{t+h}^t\) and applying the lemmas with \({\bar{t}}:= t+h \in [0,T)\) and \({\bar{h}}:= -h \in (0,T-{\bar{t}}\,]\). \(\square \)

We now investigate temporal mean-square differentiability. To this end, we need the following estimate which is implied by (3.3): For all \(a\in {\mathbb {R}}\), \(b\in [0,\infty )\), we have

$$\begin{aligned} \forall c \in [0,\infty ): \quad \Vert {\varPhi }_{a,b}(t)x\Vert _{H} \lesssim _{c} t^{a-c} \bigl \Vert A^{b-c} Q^{\frac{1}{2}} x \bigr \Vert _{H} \quad \forall x\in {\textsf{D}}\bigl ( A^{b-c} Q^{\frac{1}{2}} \bigr ). \end{aligned}$$
(3.24)

The next lemma records some information about the derivatives of \({\varPhi }_{a,b}\) in (3.22).

Lemma 3.20

Let Assumptions 3.1(i),(ii) be satisfied, and let \(a\in {\mathbb {R}}\), \(b,\sigma \in [0,\infty )\). If \(\sigma \in (0,\infty )\), suppose furthermore that Assumption 3.1(iv) holds. Then, the function \({\varPhi }_{a,b}\) defined by (3.22) belongs to \(C^\infty ((0,\infty ); {\mathscr {L}}(H; \dot{H}_A^\sigma ))\) with kth derivative

$$\begin{aligned} \frac{\mathrm d^k}{\mathrm d t^k} \, {\varPhi }_{a,b} (t) = \sum _{j=0}^k C_{a,j,k} t^{a -(k-j)} A^{b+j} S(t)Q^\frac{1}{2} = \sum _{j=0}^k C_{a,j,k} {\varPhi }_{a-(k-j),b+j}(t), \end{aligned}$$
(3.25)

where \(C_{a,j,k}:= (-1)^j \left( {\begin{array}{c}k\\ j\end{array}}\right) \prod _{i=1}^{k-j} (a-(k-j)+i)\) for \(a\in {\mathbb {R}}\), \(j,k\in {\mathbb {N}}_0\), \(j\le k\).

Moreover, if \(r\in [0, 2b+\sigma ]\) is such that \(Q^{\frac{1}{2}} \in {\mathscr {L}}(H;\dot{H}_A^r)\) and \(n\in {\mathbb {N}}_0\) satisfies \(n < a - b - \tfrac{\sigma -r}{2}\), then \({\varPhi }_{a,b}\) has a continuous extension in \(C^n( [0,\infty ); {\mathscr {L}}(H;{\dot{H}}^{\sigma }_A))\) with all n derivatives vanishing at zero.

Proof

Since \((S(t))_{t\ge 0}\) is assumed to be analytic, \(S(\,\cdot \,)\) is infinitely differentiable from \((0,\infty )\) to \({\mathscr {L}}(H)\), with jth derivative \((-A)^j S(\,\cdot \,)\) and, for \(t \in (0,\infty )\), \(\varepsilon := \frac{t}{2}\),

$$\begin{aligned} \bigl [ A^{ b + \frac{\sigma }{2} } S(\,\cdot \,) \bigr ]^{(j)}(t)&= \bigl [ S( \,\cdot \, - \varepsilon ) A^{ b+ \frac{\sigma }{2} } S( \varepsilon ) \bigr ]^{(j)}(t)\\&= (-A)^j S(t-\varepsilon ) A^{ b + \frac{\sigma }{2} } S(\varepsilon ) = (-1)^j A^{ j + b + \frac{\sigma }{2} } S(t). \end{aligned}$$

Here, the limits for the derivatives are taken in the \({\mathscr {L}}(H)\) norm. This is equivalent to \([ A^b S(\,\cdot \,) ]^{(j)}(t) = (-1)^j A^{j+b} S(t)\) with respect to the \({\mathscr {L}}(H;\dot{H}_A^\sigma )\) norm. The expression for the kth derivative of \({\varPhi }_{a,b}\) thus follows from the Leibniz rule.

Now let \(r\in [0, 2b+\sigma ]\), \(n\in {\mathbb {N}}_0\) be such that \(n < a-b-\frac{\sigma -r}{2}\) and \(Q^{\frac{1}{2}} \in {\mathscr {L}}(H;\dot{H}_A^r)\). To prove the second claim, we derive that for all \(k \in \{0,1,\dots ,n\}\) and \(t \in (0,\infty )\)

$$\begin{aligned} \biggl \Vert \frac{\mathrm d^k}{\mathrm d t^k} \, {\varPhi }_{a,b} (t) \biggr \Vert _{{\mathscr {L}}(H;{\dot{H}}^{\sigma }_A)}&= \biggl \Vert \sum _{j=0}^k C_{a,j,k} t^{a -(k-j)} A^{b+j+\frac{\sigma -r}{2}} S(t) A^{\frac{r}{2}} Q^\frac{1}{2} \biggr \Vert _{{\mathscr {L}}(H)}\\&\lesssim _{(a,b,k,r,\sigma )} t^{a-k-b-\frac{\sigma -r}{2}} \bigl \Vert Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}(H; \dot{H}_A^r)} \end{aligned}$$

by applying (3.24) to each summand with \(c:=b+j+\frac{\sigma -r}{2}\ge 0\). Furthermore, since \(a-k-b-\frac{\sigma -r}{2}\ge a-n-b-\frac{\sigma -r}{2}>0\), the above quantity tends to zero as \(t \downarrow 0\). Hence, extending \(t\mapsto \frac{\mathrm d^k}{\mathrm d t^k} \, {\varPhi }_{a,b}(t)\) by zero at \(t=0\) gives a function in \(C([0,\infty ); {\mathscr {L}}(H; {\dot{H}}^{\sigma }_A))\) for all \(k \in \{0,1,\dots ,n\}\). Inductively it follows then that the kth derivative of the zero extension is the zero extension of the original kth derivative. \(\square \)

Proposition 3.21

Let \(\sigma \in [0,\infty )\), and additionally require Assumptions 3.1(i),(ii),(iv) whenever \({\sigma \in (0,\infty )}\). Suppose that \({\varPsi } \in H^1_{0,\{0\}}(0, T; {\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A))\) and let \({\varPsi }'\) denote its weak derivative. Then, for every \(p \in [1,\infty )\), the stochastic convolution

$$\begin{aligned} t \mapsto \int _0^t {\varPsi }(t-s)\mathop {}\!\textrm{d}W(s) \end{aligned}$$

is differentiable from [0, T] to \(L^p({\varOmega }; {\dot{H}}^{\sigma }_A)\), with derivative

$$\begin{aligned} \frac{\mathop {}\!\textrm{d}}{\mathop {}\!\textrm{d}t}\int _0^{t} {\varPsi }(t-s)\mathop {}\!\textrm{d}W(s) = \int _0^t {\varPsi }'(t-s) \mathop {}\!\textrm{d}W(s) \quad \forall \, t \in [0,T]. \end{aligned}$$
(3.26)

Proof

For \(t \in [0,T)\) and \(h \in (0,T-t]\), we can write

$$\begin{aligned}&\frac{1}{h} \biggl [ \int _0^{t+h} {\varPsi }(t+h-s)\mathop {}\!\textrm{d}W(s) - \int _0^{t} {\varPsi }(t-s)\mathop {}\!\textrm{d}W(s) \biggr ] - \int _0^t {\varPsi }'(t-s) \mathop {}\!\textrm{d}W(s) \\&\quad = \int _0^t \biggl [ \frac{{\varPsi }(t+h-s) - {\varPsi }(t-s)}{h} - {\varPsi }'(t-s) \biggr ] \mathop {}\!\textrm{d}W(s) \\&\qquad + \frac{1}{h}\int _t^{t+h} {\varPsi }(t+h-s)\mathop {}\!\textrm{d}W(s) \\&\quad =: I^{h^+}_{1} + I^{h^+}_{2}. \end{aligned}$$

For \(t\in (0,T]\) and \(h \in [-t,0)\), we instead have

$$\begin{aligned}&\frac{1}{h}\biggl [ \int _0^{t+h} {\varPsi }(t+h-s)\mathop {}\!\textrm{d}W(s) - \int _0^{t} {\varPsi }(t-s)\mathop {}\!\textrm{d}W(s)\biggr ] - \int _0^t {\varPsi }'(t-s) \mathop {}\!\textrm{d}W(s) \\&\quad = \int _0^{t+h} \biggl [ \frac{{\varPsi }(t+h-s)-{\varPsi }(t-s)}{h} - {\varPsi }'(t-s)\biggr ] \mathop {}\!\textrm{d}W(s) \\&\qquad - \frac{1}{h}\int _{t+h}^t {\varPsi }(t-s) \mathop {}\!\textrm{d}W(s) - \int _{t+h}^t {\varPsi }'(t-s)\mathop {}\!\textrm{d}W(s) =: I_1^{h^-} + I_2^{h^-} + I_3^{h^-}. \end{aligned}$$

We first deal with the terms \(I_2^{h^\pm }\). Note that \({\varPsi } \in H^1_{0,\{0\}}(0, T; {\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A))\) implies \({\varPsi }(u) = \int _0^u {\varPsi }'(r)\mathop {}\!\textrm{d}r\) for all \(u \in (0, |h|)\), see [32, §5.9.2, Theorem 2]. In conjunction with the Burkholder–Davis–Gundy inequality (combined with the embedding (3.23) if \(p\in [1,2)\)) and the Cauchy–Schwarz inequality, this leads to

$$\begin{aligned}&\bigl \Vert I_2^{h^\pm } \bigr \Vert _{L^p({\varOmega };{\dot{H}}^{\sigma }_A)} \lesssim _{p} \frac{1}{|h|} \biggl [ \int _0^{|h|} \Vert {\varPsi }(u)\Vert _{{\mathscr {L}}_2(H;{\dot{H}}^{\sigma }_A)}^2 \mathop {}\!\textrm{d}u \biggr ]^{{1}/{2}} \\&\quad \le \frac{1}{|h|} \biggl [ \int _0^{|h|} \biggl (\int _0^u \Vert {\varPsi }'(r)\Vert _{{\mathscr {L}}_2(H;{\dot{H}}^{\sigma }_A)} \mathop {}\!\textrm{d}r\biggr )^2 \mathop {}\!\textrm{d}u\biggr ]^{{1}/{2}} \le \Vert {\varPsi }'\Vert _{L^2(0,|h|;{\mathscr {L}}_2(H;{\dot{H}}^{\sigma }_A))}. \end{aligned}$$

Moreover, we find that

$$\begin{aligned} \bigl \Vert I_3^{h^-} \bigr \Vert _{L^p({\varOmega };{\dot{H}}^{\sigma }_A)}&\lesssim _p \biggl [\int _{t+h}^t \Vert {\varPsi }'(t-s)\Vert _{{\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A)}^2 \mathop {}\!\textrm{d}s\biggr ]^{{1}/{2}}\\&= \biggl [\int _{0}^{|h|} \Vert {\varPsi }'(u)\Vert _{{\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A)}^2 \mathop {}\!\textrm{d}u \biggr ]^{{1}/{2}} = \Vert {\varPsi }'\Vert _{L^2(0,|h|;{\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A))}. \end{aligned}$$

Since \({\varPsi }' \in L^2(0,T; {\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A))\), we have that \(\Vert {\varPsi }'\Vert _{L^2(0,|h|;{\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A))} \rightarrow 0\) as \(h \rightarrow 0\) by dominated convergence. Thus, it remains to deal with the \(I_1^{h^\pm }\) terms. For the case of positive h, we find using the definition of the difference quotient \(D_h\) (see Eq. (A.6) in Sect. A.4 of “Appendix A”) that

$$\begin{aligned} \bigl \Vert I_1^{h^+} \bigr \Vert _{L^p({\varOmega };{\dot{H}}^{\sigma }_A)}&\lesssim _p \biggl [ \int _0^t \biggl \Vert \frac{{\varPsi }(t+h-s) - {\varPsi }(t-s)}{h} - {\varPsi }'(t-s) \biggr \Vert _{{\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A)}^2 \mathop {}\!\textrm{d}s\biggr ]^{{1}/{2}}\\&= \biggl [\int _0^t \biggl \Vert \frac{{\varPsi }(u+h) - {\varPsi }(u)}{h} - {\varPsi }'(u) \biggr \Vert _{{\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A)}^2 \mathop {}\!\textrm{d}u \biggr ]^{{1}/{2}}\\&= \Vert D_h {\varPsi } - {\varPsi }'\Vert _{L^2(0,t; {\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A))}. \end{aligned}$$

For the case of negative h, we arrive at

$$\begin{aligned} \bigl \Vert I_1^{h^-}\bigr \Vert _{L^p({\varOmega };{\dot{H}}^{\sigma }_A)}&\lesssim _p \biggl [\int _0^{t+h} \biggl \Vert \frac{{\varPsi }(t+h-s) - {\varPsi }(t-s)}{h} - {\varPsi }'(t-s) \biggr \Vert _{{\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A)}^2\mathop {}\!\textrm{d}s \biggr ]^{{1}/{2}}\\&= \biggl [\int _{-h}^t \biggl \Vert \frac{{\varPsi }(u+h) - {\varPsi }(u)}{h} - {\varPsi }'(u) \biggr \Vert _{{\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A)}^2\mathop {}\!\textrm{d}u \biggr ]^{{1}/{2}} \\&= \Vert D_h {\varPsi } - {\varPsi }'\Vert _{L^2(-h,t; {\mathscr {L}}_2(H; {\dot{H}}^{\sigma }_A))}. \end{aligned}$$

The convergence \(\lim _{h\rightarrow 0}\Vert I_1^{h^\pm }\Vert _{L^p({\varOmega };{\dot{H}}^{\sigma }_A)}=0\) follows then from Proposition A.8. \(\square \)

We are now ready to prove Theorem 3.12.

Proof of Theorem 3.12

We first claim that the mild solution, viewed as a mapping \(Z_\gamma :[0,T]\rightarrow L^p({\varOmega };{\dot{H}}^{\sigma }_A)\), is n times differentiable and, for every \(k \in \{0,1,\dots ,n\}\) and all \(t\in [0,T]\), its kth derivative satisfies

$$\begin{aligned} Z_\gamma ^{(k)}(t) = \frac{1}{{\varGamma }(\gamma )} \int _0^t {\varPhi }_{\gamma -1,0}^{(k)}(t-s) \mathop {}\!\textrm{d}{\widehat{W}}(s), \quad \mathbb {P}\text {-a.s.}, \end{aligned}$$
(3.27)

where \({\varPhi }_{\gamma -1,0}^{(k)}\) is the kth derivative of \({\varPhi }_{\gamma -1,0}\) given by (3.25), and \({\widehat{W}}\) is the cylindrical Wiener process \({\widehat{W}}(t):= Q^{-\frac{1}{2}} W^Q(t)\), \(t\ge 0\). We prove this by induction with respect to k. For \(k=0\), the identity (3.27) follows from Definition 3.7 and (3.22). Now let \({k \in \{0,1,\dots ,n-1\}}\) and suppose that \(Z_\gamma \) is k times differentiable and (3.27) holds. Then, the induction hypothesis and Lemma 3.20 show that, for all \(t\in [0,T]\),

$$\begin{aligned} \frac{\mathrm d^{k+1}}{\mathop {}\!\textrm{d}t^{k+1}} \, Z_\gamma (t)&= \frac{\mathrm d}{\mathop {}\!\textrm{d}t} \, Z_\gamma ^{(k)}(t) = \frac{\mathop {}\!\textrm{d}}{\mathop {}\!\textrm{d}t} \biggl [ \frac{1}{{\varGamma }(\gamma )} \int _0^t {\varPhi }^{(k)}_{\gamma -1,0}(t-s) \mathop {}\!\textrm{d}{\widehat{W}}(s) \biggr ]\\&= \frac{1}{{\varGamma }(\gamma )} \frac{\mathop {}\!\textrm{d}}{\mathop {}\!\textrm{d}t} \int _0^t \sum _{j=0}^k C_{\gamma -1,j,k} {\varPhi }_{\gamma -1-(k-j),j}(t-s) \mathop {}\!\textrm{d}{\widehat{W}}(s), \quad {\mathbb {P}}\text {-a.s.} \end{aligned}$$

Fixing an arbitrary \(j \in \{0,1,\dots ,k\}\), it suffices to verify that \({\varPsi }:= {\varPhi }_{\gamma -1-(k-j),j}\) satisfies the conditions of Proposition 3.21, so that (3.26) holds for the cylindrical Wiener process \({\widehat{W}}\). Indeed, having proved this for an arbitrary j, by linearity

$$\begin{aligned} \frac{\mathrm d^{k+1}}{\mathop {}\!\textrm{d}t^{k+1}} \, Z_\gamma (t)&= \frac{1}{{\varGamma }(\gamma )} \int _0^t \sum _{j=0}^k C_{\gamma -1,j,k} {\varPhi }_{\gamma -1-(k-j),j}'(t-s) \mathop {}\!\textrm{d}{\widehat{W}}(s)\\&= \frac{1}{{\varGamma }(\gamma )} \int _0^t{\varPhi }_{\gamma -1,0}^{(k+1)}(t-s) \mathop {}\!\textrm{d}{\widehat{W}}(s), \quad {\mathbb {P}}\text {-a.s.}, \end{aligned}$$

follows, where the latter identity is an equality of the operator-valued integrands.

Using (3.3) with \(c:=b\), the identity \(A^{\frac{\sigma }{2}} {\varPhi }_{a,b}(t) = 2^a (t/2)^a A^b S(t/2) A^{\frac{\sigma }{2}} S(t/2) Q^{\frac{1}{2}}\) and a change of variables \(u:=t/2\), we observe that

$$\begin{aligned} \begin{aligned}&\Vert {\varPhi }_{a,b} \Vert _{L^2(0,T;{\mathscr {L}}_2(H;\dot{H}_A^\sigma ))} \lesssim _{(a,b)} \biggl [ \int _0^T (t/2)^{2(a-b)} \bigl \Vert A^{\frac{\sigma }{2}} S(t/2) Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}_2(H)}^2 \frac{\mathop {}\!\textrm{d}t}{2} \biggr ]^{{1}/{2}}\\&\quad = \biggl [ \int _0^T \Vert {\varPhi }_{a-b,0}(t/2) \Vert _{{\mathscr {L}}_2(H;\dot{H}_A^\sigma )}^2 \frac{\mathop {}\!\textrm{d}t}{2} \biggr ]^{{1}/{2}} \le \Vert {\varPhi }_{a-b,0}\Vert _{L^2(0,T;{\mathscr {L}}_2(H;\dot{H}_A^\sigma ))} \end{aligned} \end{aligned}$$
(3.28)

holds for all \(a\in {\mathbb {R}}\) and \(b\in [0,\infty )\). For \({\varPsi }={\varPhi }_{\gamma -1-(k-j),j}\) we use (3.28) to obtain

$$\begin{aligned} \Vert {\varPsi }\Vert _{L^2(0,T;{\mathscr {L}}_2(H;\dot{H}_A^\sigma ))} \lesssim _{(\gamma ,k,j)} \Vert {\varPhi }_{\gamma -1-k,0}\Vert _{L^2(0,T;{\mathscr {L}}_2(H;\dot{H}_A^\sigma ))}. \end{aligned}$$

The norm on the right-hand side is finite by (3.19), since \(k \le n - 1 < n\). Next, noting that \(\gamma -1-k-\frac{\sigma -r}{2} \ge \gamma - n - \frac{\sigma -r}{2} > 0\), the second assertion of Lemma 3.20 implies that \(t \mapsto {\varPsi }(t)\) has a continuous extension in \(C_{0,\{0\}}([0,T]; {\mathscr {L}}(H;\dot{H}_A^\sigma ))\). Furthermore, also by Lemma 3.20, \({\varPsi }\) is differentiable from (0, T) to \({\mathscr {L}}(H;{\dot{H}}^{\sigma }_A)\), with derivative

$$\begin{aligned} {\varPsi }' = (\gamma -1-(k-j)){\varPhi }_{\gamma -1-(k-j)-1, j} - {\varPhi }_{\gamma -1-(k-j), j+1}. \end{aligned}$$

Applying the triangle inequality and (3.28) then shows that

$$\begin{aligned} \Vert {\varPsi }'\Vert _{L^2(0,T;{\mathscr {L}}_2(H;\dot{H}_A^\sigma ))} \lesssim _{(\gamma ,k,j)} \Vert {\varPhi }_{\gamma -1-(k+1),0}\Vert _{L^2(0,T;{\mathscr {L}}_2(H;\dot{H}_A^\sigma ))}, \end{aligned}$$

where the norm on the right-hand side is finite by (3.19), as \({k + 1 \le n }\). Since also \({\mathscr {L}}_2(H;{\dot{H}}^{\sigma }_A) \hookrightarrow {\mathscr {L}}(H;{\dot{H}}^{\sigma }_A)\), Lemma A.9 implies that \(\Psi \) is an element of \(H^1_{0,\{0\}}(0,T; \mathscr {L}_2(H; \dot{H}_A^\sigma ))\). Thus, we may indeed use Proposition 3.21, and the differentiability follows.

It remains to show that the nth derivative \(Z_\gamma ^{(n)}\) is (Hölder) continuous, i.e., that \(Z_\gamma ^{(n)} \in C^{0,\tau }([0,T]; L^p({\varOmega }; \dot{H}_A^\sigma ))\). To this end, we use (3.27) and (3.25), and write

$$\begin{aligned} \forall t\in [0,T]: \quad Z_\gamma ^{(n)}(t) = \frac{1}{{\varGamma }(\gamma )} \sum _{j=0}^n C_{\gamma -1,j,n} \int _0^t {\varPhi }_{\gamma -1-(n-j),j}(t-s) \mathop {}\!\textrm{d}{\widehat{W}}(s), \quad {\mathbb {P}}\text {-a.s.} \end{aligned}$$

The case \(\tau =0\) (i.e., continuity) follows after applying, for all \(j \in \{0,1,\ldots ,n\}\), Proposition 3.18 with \(a=\gamma -1-(n-j)\) and \(b = j\). Note that \({\varPhi }_{\gamma -1-(n-j),j}\) indeed is an element of \(L^2(0,T;{\mathscr {L}}_2(H;\dot{H}_A^\sigma ))\) for all \(j \in \{0,\ldots ,n\}\) by (3.19) and (3.28). For \(\tau \in \bigl (0,\gamma -n-\frac{1}{2}\bigr ] \cap (0,1)\), the Hölder continuity of \(Z_\gamma ^{(n)}\) follows from Proposition 3.19 which we may apply, for all \(j \in \{0,1,\ldots ,n\}\), with \(a=\gamma -1-(n-j)\) and \(b = j\), since \( A^{n+\tau +\frac{1}{2}-\gamma } Q^{\frac{1}{2}} \in {\mathscr {L}}_2(H;\dot{H}_A^\sigma )\) is assumed. \(\square \)

4 Covariance structure

In this section, we study the covariance structure of solutions to (3.1). More specifically, we consider the mild solution process \((Z_\gamma (t))_{t\in [0,T]}\) from Definition 3.7. The covariance structure of \(Z_\gamma \) will be expressed in terms of the family of covariance operators \((Q_{Z_\gamma }(s,t))_{s,t\in [0,T]} \subseteq {\mathscr {L}}(H)\) which satisfies, for all \(s,t\in [0,T]\), that

$$\begin{aligned} ( Q_{Z_\gamma }(s,t) x, y)_H = {\mathbb {E}}[ (Z_\gamma (s)-{\mathbb {E}}[Z_\gamma (s)], x)_H( Z_\gamma (t)-{\mathbb {E}}[Z_\gamma (t)], y)_H ] \quad \forall x,y\in H. \end{aligned}$$

Note that this family is well-defined whenever \(Z_\gamma \) is square-integrable, e.g., under the assumptions made in Theorem 3.8. Note also that \({\mathbb {E}}[Z_\gamma (t)] = 0\) for all \(t\in [0,T]\).

We present three results on the covariance operators of the mild solution \(Z_\gamma \). The most general result is Proposition 4.1, which provides an explicit integral representation of \(Q_{Z_\gamma }(s,t)\). Corollary 4.2 is concerned with the asymptotic behavior of the covariance operator \(Q_{Z_\gamma }(t,t)\) as \(t\rightarrow \infty \). Subsequently, in Corollary 4.3 we consider a situation in which the covariance is separable in time and space, and prove that the temporal part is asymptotically of Matérn type.

Proposition 4.1

Let Assumption 3.1(i) be satisfied and \(\gamma \in (0,\infty )\) be such that (3.14) holds. The covariance operators \((Q_{Z_\gamma }(s,t))_{s,t\in [0,T]}\) of \(Z_\gamma \) admit the representation

$$\begin{aligned} Q_{Z_\gamma }(s,t) = \frac{1}{{\varGamma }(\gamma )^2} \int _0^{s\wedge t} [(s-r)(t-r)]^{\gamma -1} S(t-r) Q [S(s-r)]^* \mathop {}\!\textrm{d}r. \end{aligned}$$
(4.1)

Proof

Square-integrability of \(Z_\gamma \) is a consequence of Theorem 3.8 and (3.14). In order to prove the integral representation (4.1), for \(s\in [0,T]\), \(r\in (0,s)\) and \(x\in H\), we define \(f(s,r;x) \in {\mathscr {L}}(H;{\mathbb {R}})\) by

$$\begin{aligned} f(s,r;x) z:= [{\varGamma }(\gamma )]^{-1} (z, (s-r)^{\gamma -1} [S(s-r)]^* x)_H, \quad z\in H. \end{aligned}$$

We proceed similarly as in [45, Lemma 3.10] and obtain (4.1) from the Itô isometry combined with the polarization identity:

$$\begin{aligned}&{\mathbb {E}}[ (Z_\gamma (s), x)_H ( Z_\gamma (t), y)_H ] = {\mathbb {E}}\left[ \int _0^s f(s,r;x) \mathop {}\!\textrm{d}W^Q(r) \int _0^t f(t,\tau ;y) \mathop {}\!\textrm{d}W^Q(\tau ) \right] \\&\quad = \int _0^{s\wedge t} \bigl ( f(s,r;x) Q^{\frac{1}{2}}, f(t,r;y) Q^{\frac{1}{2}} \bigr )_{{\mathscr {L}}_2(H;{\mathbb {R}})} \mathop {}\!\textrm{d}r\\&\quad = \frac{1}{{\varGamma }(\gamma )^2} \int _0^{s\wedge t} [(s-r)(t-r)]^{\gamma -1} ( S(t-r) Q [S(s-r)]^* x, y )_H \mathop {}\!\textrm{d}r . \end{aligned}$$

Then, (4.1) follows from exchanging the order of integration and taking the inner product, which is justified since \((0,s\!\wedge t)\!\ni r \!\mapsto [(s-r)(t-r)]^{\gamma -1}S(t-r) Q [S(s-r)]^* x\) is integrable by (3.14). \(\square \)

By imposing more assumptions on the operator A, one can obtain explicit representations of the asymptotic covariance structure of \(Z_\gamma \) as \(t\rightarrow \infty \), as the next two corollaries show. Note that, if (3.14) holds for \(\delta =0\) and \(T=\infty \), in Definition 3.7 the stochastic convolution \({{\widetilde{Z}}}_\gamma \) and the mild solution \(Z_\gamma \) are well-defined on the infinite time interval \([0,\infty )\). It is thus meaningful to consider the asymptotic behavior.

Corollary 4.2

Let Assumptions 3.1(i),(ii),(iv) be satisfied and let \({\gamma \in ({1}/{2},\infty )}\). Suppose that (3.14) holds for \(\delta =0\) and \(T=\infty \). If for every \(t \in [0,\infty )\) the operator S(t) is self-adjoint and commutes with the covariance operator Q of \(W^Q\), we have

$$\begin{aligned} \lim _{t\rightarrow \infty } Q_{Z_\gamma }(t,t) = {\varGamma }(\gamma -{1}/{2}) \bigl [ 2\sqrt{\pi } {\varGamma }(\gamma ) \bigr ]^{-1} A^{1-2\gamma } Q \quad \text {in } {\mathscr {L}}(H). \end{aligned}$$

Proof

Starting from the identity (4.1) for a fixed \(t = s \in [0,\infty )\), we recall self-adjointness of the operators \((S(t))_{t\ge 0}\) and the commutativity with Q to obtain that

$$\begin{aligned} Q_{Z_\gamma }(t,t)&= \frac{1}{{\varGamma }(\gamma )^2} \int _0^t (t-r)^{2(\gamma -1)} S(t-r) Q S(t-r) \mathop {}\!\textrm{d}r\\&= \frac{1}{{\varGamma }(\gamma )^2} \int _0^t (t-r)^{2\gamma -2} Q S(2t-2r) \mathop {}\!\textrm{d}r = \frac{2^{1-2\gamma }}{{\varGamma }(\gamma )^2} \int _0^{2t} u^{2\gamma -2} Q S(u) \mathop {}\!\textrm{d}u, \end{aligned}$$

where we also used the semigroup property and the change of variables \({u:= 2(t-r)}\). Now we interchange the bounded linear operator Q with the integral, and pass to the limit \(t \rightarrow \infty \) in \({\mathscr {L}}(H)\), which by (B.3) with \(\alpha := 2\gamma - 1 \in (0,\infty )\) gives

$$\begin{aligned} \lim _{t\rightarrow \infty } Q_{Z_\gamma }(t,t)&= 2^{1-2\gamma }\Gamma (2\gamma -1) [\Gamma (\gamma )]^{-2} A^{1-2\gamma } Q \\&= \Gamma (\gamma -{1}/{2}) \bigl [ 2\sqrt{\pi } \Gamma (\gamma ) \bigr ]^{-1} A^{1-2\gamma } Q. \end{aligned}$$

The last equality is a consequence of the Legendre duplication formula for the gamma function (see e.g. [59, Formula 5.5.5]) applied to \({\varGamma }(2\gamma -1) = {\varGamma }(2[\gamma -{1}/{2}])\). \(\square \)

Corollary 4.3

Suppose the setting of Corollary 4.2 and let \(A:= \kappa I\) for \({\kappa \in (0,\infty )}\). Then the covariance function of \(Z_\gamma \) is separable and its temporal part is asymptotically of Matérn type, i.e., there is a function \(\varrho _{Z_\gamma }:[0,\infty )\times [0,\infty ) \rightarrow {\mathbb {R}}\) such that

$$\begin{aligned}&\forall s,t \in [0,\infty ):{} & {} Q_{Z_\gamma }(s,t) = \varrho _{Z_\gamma }(s,t)\, Q , \nonumber \\&\quad \forall h \in {\mathbb {R}}{\setminus }\{0\}:{} & {} \lim _{t\rightarrow \infty }\varrho _{Z_\gamma }(t,t+h) = \frac{2^{\frac{1}{2}-\gamma }\kappa ^{1-2\gamma }}{\sqrt{\pi }{\varGamma }(\gamma )} (\kappa |h|)^{\gamma -\frac{1}{2}} K_{\gamma -\frac{1}{2}}(\kappa |h|) . \end{aligned}$$
(4.2)

Remark 4.4

On the right-hand side of (4.2), one recognizes the Matérn covariance function (1.1) with smoothness parameter \(\nu = \gamma -{1}/{2}\), correlation length parameter \(\kappa \) and variance \(\sigma ^2 = \kappa ^{1-2\gamma } {\varGamma }(\gamma -{1}/{2}) \bigl [ 2 \sqrt{\pi } {\varGamma }(\gamma ) \bigr ]^{-1}\).

Proof of Corollary 4.3

For \(s,t \ge 0\), the integral representation (4.1) yields

$$\begin{aligned} Q_{Z_\gamma }(s,t) = \frac{1}{{\varGamma }(\gamma )^2} \int _0^{s\wedge t} [(s-r)(t-r)]^{\gamma -1} e^{-\kappa (s+t-2r)} \mathop {}\!\textrm{d}r \, Q = \varrho _{Z_\gamma }(s,t) \, Q, \end{aligned}$$

where we moved the bounded operator \(Q\in {\mathscr {L}}(H)\) out of the integral. Next, we fix \(h \in (0,\infty )\), let \(t \in [0,\infty )\) and perform the change of variables \({u:=h + 2(t-r) }\),

$$\begin{aligned} \varrho _{Z_\gamma }(t,t+h) = \varrho _{Z_\gamma }(t+h,t) = \frac{2^{1-2\gamma }}{{\varGamma }(\gamma )^2} \int _h^{2t+h} [(u+h)(u-h)]^{\gamma -1} e^{-\kappa u} \mathop {}\!\textrm{d}u. \end{aligned}$$

Thus, by passing to the limit \(t \rightarrow \infty \), we obtain

$$\begin{aligned}&\lim _{t\rightarrow \infty } \varrho _{Z_\gamma }(t,t+h) = \frac{2^{1-2\gamma }}{{\varGamma }(\gamma )^2} \int _h^{\infty } \bigl ( u^2 - h^2 \bigr )^{\gamma -1} e^{-\kappa u} \mathop {}\!\textrm{d}u\\&\quad = \frac{2^{1-2\gamma }}{{\varGamma }(\gamma )^2} {\mathcal {L}}\bigl [ u \mapsto \bigl ( u^2-h^2 \bigr )^{\gamma -1} {\textbf{1}}_{(h,\infty )}(u) \bigr ] (\kappa ) = \frac{2^{1-2\gamma }}{{\varGamma }(\gamma )^2} \frac{(2h)^{\gamma -\frac{1}{2}} {\varGamma }(\gamma )}{\sqrt{\pi }\kappa ^{\gamma -\frac{1}{2}}} K_{\gamma -\frac{1}{2}}(\kappa h), \end{aligned}$$

where \({\mathcal {L}}[f](\kappa )\) denotes the Laplace transform of the function \(f :[0,\infty )\rightarrow {\mathbb {R}}\) evaluated at \(\kappa \), and the last identity follows from [58, Chapter I, Formula 3.13]. \(\square \)

5 Spatiotemporal Whittle–Matérn fields

In this section, we demonstrate how the results of the previous Sects. 3 and 4 can be related to the widely used statistical models involving generalized Whittle–Matérn operators (1.3) on \({H = L^2({\mathcal {X}})}\), where \({\mathcal {X}}={\mathcal {D}}\subsetneq {\mathbb {R}}^d\) is a bounded domain in the Euclidean space (see Sect. 5.1) or a surface \({\mathcal {X}}={\mathcal {M}}\) (see Sect. 5.2).

5.1 Bounded Euclidean domains

Throughout this subsection, let \({\emptyset \ne {\mathcal {D}}\subsetneq {\mathbb {R}}^d}\) be a bounded, connected and open domain. In order to rigorously define the symmetric, strongly elliptic second-order differential operator L, formally given by (1.3), as a linear operator on \(L^2({\mathcal {D}})\), we make the following assumptions on its coefficients \(\kappa :{\mathcal {D}}\rightarrow {\mathbb {R}}\) and \(a:{\mathcal {D}}\rightarrow {\mathbb {R}}^{d\times d}_{\textrm{sym}}\), as well as on the spatial domain \({\mathcal {D}}\subsetneq {\mathbb {R}}^d\).

Assumption 5.1

(Euclidean domain—minimal conditions)

  1. (i)

    \({\mathcal {D}}\) has a Lipschitz continuous boundary \(\partial {\mathcal {D}}\);

  2. (ii)

    \(a \in L^\infty \bigl ( {\mathcal {D}}; {\mathbb {R}}^{d\times d}_{\textrm{sym}} \bigr )\) is strongly elliptic, i.e.,

    $$\begin{aligned} \exists \, \theta > 0: \qquad \mathop {{\mathrm{ess\,inf}}}\limits _{x\in {\mathcal {D}}} \, \xi ^\top a(x) \xi \ge \theta \Vert \xi \Vert _{{\mathbb {R}}^d}^2 \qquad \forall \xi \in {\mathbb {R}}^d; \end{aligned}$$
  3. (iii)

    \(\kappa \in L^\infty ({\mathcal {D}})\).

Under these assumptions, we introduce the bilinear form

$$\begin{aligned} {\mathfrak {a}}_L :H_0^1({\mathcal {D}})\times H_0^1({\mathcal {D}}) \rightarrow {\mathbb {R}}, \qquad {\mathfrak {a}}_L(u,v):= (a\nabla u, \nabla v)_{L^2({\mathcal {D}})} + (\kappa ^2 u,v)_{L^2({\mathcal {D}})}, \end{aligned}$$

which is symmetric, continuous and coercive. We say that \(u \in H^1_0({\mathcal {D}})\) belongs to the domain \({\textsf{D}}(L)\) of the differential operator L if and only if \(|{\mathfrak {a}}_L(u,v) | \lesssim _u \Vert v\Vert _{L^2({\mathcal {D}})}\) holds for all \(v \in H^1_0({\mathcal {D}})\). In this case, we define Lu as the unique element of \(L^2({\mathcal {D}})\) which satisfies the relation \({\mathfrak {a}}_L(u,v) = (Lu,v)_{L^2({\mathcal {D}})}\) for all \(v \in H^1_0({\mathcal {D}})\).

By the Lax–Milgram theorem the inverse \(L^{-1} \in {\mathscr {L}}(L^2({\mathcal {D}}); H_0^1({\mathcal {D}})) \) exists and can be extended to \(L^{-1} \in {\mathscr {L}}(H_0^1({\mathcal {D}})^*; H_0^1({\mathcal {D}}))\). Moreover, it is a consequence of the Rellich–Kondrachov theorem (see [2, Theorem 6.3]) that \(L^{-1}\) is compact on \(L^2({\mathcal {D}})\). For this reason, the spectral theorem for self-adjoint compact operators is applicable and shows that there exist an orthonormal basis \((e_j)_{j\in {\mathbb {N}}}\) for \(L^2({\mathcal {D}})\) and a non-decreasing sequence \((\lambda _j)_{j\in {\mathbb {N}}}\) of positive real numbers accumulating only at infinity such that \(L e_j = \lambda _j e_j\) holds for all \(j\in {\mathbb {N}}\). Furthermore, the eigenvalues of L satisfy the following asymptotic behavior, known as Weyl’s law [27, Theorem 6.3.1]:

$$\begin{aligned} \lambda _j \eqsim j^{{2}/{d}} \quad \forall j \in {\mathbb {N}}. \end{aligned}$$
(5.1)

In this setting, for two differential operators L and \({\widetilde{L}}\) on \(L^2({\mathcal {D}})\) with coefficients \(a,\kappa \) and \({\widetilde{a}},{\widetilde{\kappa }}\), respectively, we obtain the following corollary from the regularity results in Sect. 3 for spatiotemporal Whittle–Matérn fields, where \(A:=L^\beta \) and \(Q:={\widetilde{L}}^{-\alpha }\).

Corollary 5.2

Let \(\alpha ,\beta ,\sigma \in [0,\infty )\), set \(r:=\frac{\alpha }{\beta } \wedge \sigma \) if \(\beta > 0\) and \(r:=\sigma \) if \(\beta =0\), and suppose that \(n \in {\mathbb {N}}_0\), \(\tau \in [0,1)\) and \(\gamma \in \bigl ( n + \frac{(\sigma -r)\vee 1}{2}, \infty \bigr )\) are such that

$$\begin{aligned} \gamma \ge n + \tfrac{1+(\sigma -r)\vee (2\tau )}{2} \quad \text {and} \quad \beta \gamma >\tfrac{d}{4} - \tfrac{\alpha }{2} + \beta \bigl ( n+\tau +\tfrac{1 + \sigma }{2} \bigr ). \end{aligned}$$
(5.2)

Let \(L:{\textsf{D}}(L) \subseteq H^1_0({\mathcal {D}}) \rightarrow L^2({\mathcal {D}})\) and \({{\widetilde{L}}:{\textsf{D}}({\widetilde{L}}) \subseteq H^1_0({\mathcal {D}}) \rightarrow L^2({\mathcal {D}})}\) be symmetric, strongly elliptic second-order differential operators as defined above, cf. (1.3). Suppose that Assumption 5.1(i) holds for \({\mathcal {D}}\subsetneq {\mathbb {R}}^d\), and that the coefficients \(a,\kappa \) of L and \({{\widetilde{a}}}, {{\widetilde{\kappa }}}\) of \({{\widetilde{L}}}\) satisfy Assumptions 5.1(ii),(iii). Assume further that L and \({{\widetilde{L}}}\) diagonalize with respect to the same orthonormal basis \((e_j)_{j\in {\mathbb {N}}}\) for \(L^2({\mathcal {D}})\), i.e., there exist non-decreasing sequences \((\lambda _j)_{j\in {\mathbb {N}}}\), \(({\widetilde{\lambda }}_j)_{j\in {\mathbb {N}}}\) of positive real numbers such that \({Le_j = \lambda _j e_j}\) and \({{\widetilde{L}}} e_j = {\widetilde{\lambda }}_j e_j\) for all \(j \in {\mathbb {N}}\).

Then, setting \(A:= L^\beta \) and \(Q:= {{\widetilde{L}}}^{-\alpha }\), the mild solution \(Z_\gamma \) to (3.1) in the sense of Definition 3.7, see also (1.4), belongs to \(C^{n,\tau }([0, T]; L^p({\varOmega };{\dot{H}}^{\sigma }_A))\) for all \({p \in [1,\infty )}\). If the above conditions hold with \(n = 0\) and \(\tau \in (0,1)\), then for every \(p\in [1,\infty )\) and all \(\tau '\in [0,\tau )\) the mild solution \(Z_\gamma \) has a modification \({{\widehat{Z}}}_\gamma \in L^p\bigl ({\varOmega }; C^{0,\tau '}([0,T];\dot{H}^\sigma _A)\bigr )\).

Proof

By the spectral mapping theorem for fractional powers of operators, see e.g. [53, Section 5.3], we obtain that \(Ae_j = L^\beta e_j = \lambda _j^\beta e_j\) and \({Qe_j = {\widetilde{L}}^{-\alpha } e_j = {{\widetilde{\lambda }}}_j^{-\alpha } e_j}\). In particular, it follows that A inherits the self-adjointness and strict positive-definiteness from L. This readily implies that \(0\in \rho (A)\). By [42, Proposition 10.2.23] we see that \(A_{\mathbb {C}}\) admits a bounded \(H^\infty \)-calculus of angle \(\omega _{H^\infty }(A_{\mathbb {C}})=0\), showing that Assumptions 3.1(i)–(iv) are satisfied for A.

Furthermore, we note that, for every \(\sigma ,s\in [0,\infty )\), we have that \(\dot{H}_A^\sigma = \dot{H}_L^{\sigma \beta }\) and the spaces \(\dot{H}_L^s\) and \(\dot{H}_{{{\widetilde{L}}}}^s\) are isomorphic. The latter fact follows from the asymptotic behavior (5.1) of the eigenvalues \((\lambda _j)_{j\in {\mathbb {N}}}\) and \(({\widetilde{\lambda }}_j)_{j\in {\mathbb {N}}}\), since L and \({\widetilde{L}}\) have the same eigenfunctions. Thus, we obtain that \(Q^{\frac{1}{2}} = {\widetilde{L}}^{-\frac{\alpha }{2}} \in {\mathscr {L}}(H;\dot{H}_L^{\alpha }) \subseteq {\mathscr {L}}(H; \dot{H}_A^r)\).

Since \(\gamma \in \bigl (\frac{1}{2}+n,\infty \bigr )\cap \bigl [\frac{1}{2}+n+\frac{\sigma -r}{2},\infty \bigr )\) is assumed, by Proposition 3.14 (see also Remark 3.15) the condition (3.19) of Theorem 3.12 is equivalent to requiring that \(A^{n+\frac{1}{2}-\gamma }Q^{\frac{1}{2}} \in {\mathscr {L}}_2(H;\dot{H}_A^\sigma )\).

Since also \(\gamma \in \bigl ( \frac{\sigma -r}{2} + n, \infty \bigr ) \cap \bigl [n + \tau +\frac{1}{2}, \infty \bigr )\), we therefore conclude with Theorem 3.12 that it suffices to check that the quantity

$$\begin{aligned} \begin{aligned}&\bigl \Vert A^{\frac{\sigma }{2} + n + \tau + \frac{1}{2} - \gamma } Q^{\frac{1}{2}} \bigr \Vert _{{\mathscr {L}}_2(H)}^2 = \bigl \Vert L^{\beta \left( \frac{\sigma }{2}+n+\tau + \frac{1}{2}-\gamma \right) } {{\widetilde{L}}}^{-\frac{\alpha }{2}} \bigr \Vert _{{\mathscr {L}}_2(H)}^2\\&\quad = \sum _{j=1}^\infty \bigl \Vert L^{\beta \left( \frac{\sigma }{2}+n+\tau + \frac{1}{2}-\gamma \right) } {{\widetilde{L}}}^{-\frac{\alpha }{2}} e_j \bigr \Vert _H^2 = \sum _{j=1}^\infty \lambda _j^{2\beta \left( \frac{\sigma }{2}+n+\tau + \frac{1}{2}-\gamma \right) } {\widetilde{\lambda }}_j^{-\alpha } \end{aligned} \end{aligned}$$
(5.3)

is finite. Indeed, applying Weyl’s law (5.1) to both L and \({{\widetilde{L}}}\), it follows that

$$\begin{aligned} \sum _{j=1}^\infty \lambda _j^{2\beta \left( \frac{\sigma }{2}+n+\tau + \frac{1}{2}-\gamma \right) } {\widetilde{\lambda }}_j^{-\alpha } \eqsim _{\text{( }\alpha ,\beta ,\gamma ,\sigma ,n,\tau )} \sum _{j=1}^\infty j^{\frac{4}{d} \left[ \beta \left( n+\tau + \frac{1+\sigma }{2} \right) -\beta \gamma -\frac{\alpha }{2}\right] }, \end{aligned}$$

so that (5.3) is finite if and only if (5.2) holds, as we assume. Then, for all \({p \in [1,\infty )}\), Theorem 3.8, Theorem 3.12 and Proposition 3.14 yield the existence of a mild solution \(Z_\gamma \in C^{n,\tau }([0,T]; L^p({\varOmega };{\dot{H}}^{\sigma }_A))\), which is unique up to modification. The last assertion for \(n = 0\) and \(\tau \in (0,1)\) follows from Corollary 3.13. \(\square \)

The spatial regularity obtained in Corollary 5.2 is measured by means of the spaces \({{\dot{H}}^{\sigma }_A= \dot{H}_L^{\beta \sigma }}\). It would be more practical to express this in terms of fractional-order Sobolev spaces \(H^s({\mathcal {D}})\), \(s\ge 0\). This raises the question of how \(\dot{H}_L^s\) and \(H^s({\mathcal {D}})\) relate. The answer to this question depends on the smoothness of the coefficients \(a,\kappa \) and of the boundary \(\partial {\mathcal {D}}\). We therefore introduce two additional sets of assumptions: Assumption 5.3 is only slightly more restrictive than the minimal conditions of Assumption 5.1, whereas Assumption 5.4 requires a high degree of smoothness.

Assumption 5.3

(Euclidean domain—\(H^2({\mathcal {D}})\)-regular setting)

  1. (i)

    \({\mathcal {D}}\) is convex.

  2. (ii)

    \(a :{\overline{{\mathcal {D}}}} \rightarrow {\mathbb {R}}_\text {sym}^{d\times d} \) is Lipschitz continuous, i.e.,

    $$\begin{aligned} |a_{ij}(x) - a_{ij}(y)| \lesssim \Vert x-y\Vert _{{\mathbb {R}}^d} \quad \forall x,y\in {\overline{{\mathcal {D}}}}, \quad \forall i,j \in \{1,\dots ,d\}. \end{aligned}$$

Assumption 5.4

(Euclidean domain—smooth setting)

  1. (i)

    The boundary \(\partial {\mathcal {D}}\) is of class \(C^\infty \);

  2. (ii)

    \(a_{ij} \in C^\infty ({\overline{{\mathcal {D}}}})\) holds for all \(i,j\in \{1,\ldots ,d\}\), i.e., for all entries of a;

  3. (iii)

    \(\kappa \in C^\infty ({\overline{{\mathcal {D}}}})\).

The results of the next lemma are taken from [21, Lemma 2] and [12, Lemma 4.3].

Lemma 5.5

Let \(L:{\textsf{D}}(L) \subseteq H^1_0({\mathcal {D}}) \rightarrow L^2({\mathcal {D}})\) denote a symmetric second-order differential operator defined as above, cf. (1.3). Then, the following assertions hold:

  1. (a)

    If Assumption 5.1 is satisfied, then \(\dot{H}_L^s \hookrightarrow H^s({\mathcal {D}})\) for all \({s \in [0,1]}\). Moreover, the norms \(\Vert \,\cdot \, \Vert _{\dot{H}_L^s}\) and \(\Vert \,\cdot \, \Vert _{H^s({\mathcal {D}})}\) are equivalent on \(\dot{H}_L^s\) for \(s \in [0,1]{\setminus }\{{1}/{2}\}\);

  2. (b)

    If Assumptions 5.1 and 5.3 are fulfilled, then

    $$\begin{aligned} \bigl ( \dot{H}_L^s, \Vert \,\cdot \, \Vert _{\dot{H}_L^s} \bigr ) \cong \bigl ( H^s({\mathcal {D}}) \cap H_0^1({\mathcal {D}}), \Vert \,\cdot \,\Vert _{H^s({\mathcal {D}})} \bigr ) \quad \forall s \in [1,2]; \end{aligned}$$
  3. (c)

    If Assumptions 5.1 and 5.4 are satisfied, then we have \(\dot{H}_L^s \hookrightarrow H^s({\mathcal {D}})\) for all \(s \in [0,\infty )\), and the norms \(\Vert \,\cdot \, \Vert _{\dot{H}_L^s}\), \(\Vert \,\cdot \,\Vert _{H^s({\mathcal {D}})}\) are equivalent on \(\dot{H}_L^s\) for every \(s \in [0,\infty ){\setminus } {\mathfrak {E}}\), where \({\mathfrak {E}}:= \{ 2k + {1}/{2}: k \in {\mathbb {N}}_0 \}\) is called the exclusion set.

Combining Lemma 5.5 with the results of Corollary 5.2 shows that the mild solution \(Z_\gamma \) is an element of \(C^{n,\tau }([0,T]; L^p({\varOmega }; H^{\beta \sigma }({\mathcal {D}})))\), provided that \(\sigma \beta \in [0,s']\), where \(s'\in [1,\infty )\) is prescribed by the smoothness of the coefficients \(a,\kappa \) and the boundary \(\partial {\mathcal {D}}\) via Lemma 5.5(a), (b) or (c). Note that we do not have to take the exclusion set \({\mathfrak {E}}\) into account, as we only need the embedding \(\dot{H}_L^s \hookrightarrow H^s({\mathcal {D}})\).

Lastly, we consider the covariance structure of the mild solution, as treated in the abstract setting in Sect. 4. The most illustrative results are the asymptotic formulas presented in Corollaries 4.2 and 4.3, which we translate to the current setting in Corollary 5.6. We see that (Whittle–)Matérn operators are recovered as marginal spatial or temporal covariance operators.

Corollary 5.6

Consider the setting of Corollary 5.2 with \(L={\widetilde{L}}\), i.e., \({Q:= L^{-\alpha }}\). Let \(\alpha ,\beta \in [0,\infty )\) and \(\gamma \in ({1}/{2},\infty )\) be such that \(\beta \gamma > \frac{1}{2}\bigl ( \frac{d}{2} - \alpha +\beta \bigr )\), and let \(Z_\gamma \) be the mild solution in the sense of Definition 3.7. Then the asymptotic marginal spatial covariance of \(Z_\gamma \) satisfies

$$\begin{aligned} \lim _{t\rightarrow \infty } Q_{Z_\gamma }(t,t) = {\varGamma }(\gamma -{1}/{2}) \bigl [ 2\sqrt{\pi } {\varGamma }(\gamma ) \bigr ]^{-1} L^{\beta (1-2\gamma )-\alpha } \quad \text {in}\;\; {\mathscr {L}}(L^2({\mathcal {D}})). \end{aligned}$$

For \(\beta = 0\), the covariance of \(Z_\gamma \) is separable in the sense that there exists a function \(\varrho _{Z_\gamma }:[0,\infty )\times [0,\infty ) \rightarrow {\mathbb {R}}\) such that

$$\begin{aligned} Q_{Z_\gamma }(s,t) = {\varrho _{Z_\gamma }}(s,t) \, L^{-\alpha } \quad \forall s,t\in [0,\infty ), \end{aligned}$$

and for all \(h \in {\mathbb {R}}{\setminus } \{0\}\) we have

$$\begin{aligned} \lim _{t\rightarrow \infty }Q_{Z_\gamma }(t,t+h) = 2^{\frac{1}{2}-\gamma } \bigl [ \sqrt{\pi }{\varGamma }(\gamma ) \bigr ]^{-1} |h|^{\gamma -\frac{1}{2}} K_{\gamma -\frac{1}{2}}(|h|)\, L^{-\alpha } \quad \text {in}\;\; {\mathscr {L}}(L^2({\mathcal {D}})). \end{aligned}$$

Proof

Existence and uniqueness of the mild solution \(Z_\gamma \) follow from Corollary 5.2 with \(L = {{\widetilde{L}}}\) and \(n = \tau = \sigma = 0\). Recall from its proof that A satisfies all of Assumptions 3.1(i)–(iv). Moreover, \(A = L^\beta \) is self-adjoint and \(Q = L^{-\alpha }\in {\mathscr {L}}(L^2({\mathcal {D}}))\) commutes with A, so that it also commutes with S(t) for all \(t \in [0,\infty )\), cf. [37, Theorem 1.3.2(a)]. All assertions follow thus from Corollaries 4.2 and 4.3. \(\square \)

Remark 5.7

The asymptotic results obtained in Corollary 5.6 are in accordance with the marginal spatial and temporal covariance functions derived in [49, Section 3, Proposition 1 and Corollary 1] for the case of the differential operator \(L= \gamma _s^2 - {\varDelta }\) acting on functions defined on all of \({\mathbb {R}}^2\), where \(\gamma _s\in (0,\infty )\). Note that, in order to exploit Fourier techniques, in [49] the “time” variable t is an element of the whole real axis, \(t\in {\mathbb {R}}\), instead of only its non-negative part.

Remark 5.8

Corollaries 5.2 and 5.6 explain and justify the roles of the parameters \(\alpha \), \(\beta \) and \(\gamma \). They control three important properties of spatiotemporal Whittle–Matérn fields. Besides the temporal and spatial smoothness, measured respectively by the quantities \(n+\tau \) and \(\sigma \), we identify a third degree of freedom: The degree of separability, expressed by the ratio \(\frac{\alpha }{\beta }\in [0,\infty ]\). Indeed, if \(\frac{\alpha }{\beta } = \infty \), i.e. \(\beta = 0\), we observe that the covariance of the field is separable and that its temporal and spatial behavior are exclusively governed by the parameters \(\gamma \) and \(\alpha \), respectively. In contrast, if \(\frac{\alpha }{\beta } = 0\), i.e. \(\alpha = 0\), the SPDE is driven by spatiotemporal Gaussian white noise and the “coloring” of its solution is fully determined by the fractional parabolic differential operator \(\bigl (\partial _t + L^\beta \bigr )^\gamma \).

5.2 Surfaces

In this subsection, we provide a brief demonstration of how the above results can be extended to spatiotemporal Whittle–Matérn fields on more general spatial domains. More precisely, we consider a smooth, closed, connected, orientable and compact 2-surface \({\mathcal {M}}\) immersed in \({\mathbb {R}}^3\) and endowed with the positive surface measure \(\nu _{\mathcal {M}}\) on \({\mathcal {B}}({\mathcal {M}})\), induced by the first fundamental form. An important example of such a surface is given by the 2-sphere, \({\mathcal {M}}={\mathbb {S}}^2\).

On \(H:= L^2({\mathcal {M}})\), we consider the following analog of the symmetric, strongly elliptic second-order differential operator from Sect. 5.1, formally given by

$$\begin{aligned} Lu:= -\nabla _{\mathcal {M}}\cdot (a\nabla _{\mathcal {M}}u) + \kappa ^2 u, \qquad u \in {\textsf{D}}(L)\subseteq L^2({\mathcal {M}}), \end{aligned}$$

where \(\nabla _{{\mathcal {M}}}\,\cdot \,\) and \(\nabla _{{\mathcal {M}}}\) denote the surface divergence and the surface gradient, respectively. We record the precise conditions on the surface \({\mathcal {M}}\) and on the coefficients \(a,\kappa \) in Assumption 5.9 below; with regard to smoothness, they are analogous to the setting of Assumption 5.4 in the case of a bounded Euclidean domain.

Assumption 5.9

(Surface—smooth setting)

  1. (i)

    a is a symmetric tensor field, i.e., \(a(x):T_x {\mathcal {M}}\rightarrow T_x {\mathcal {M}}\) is linear and symmetric for all \(x \in {\mathcal {M}}\), where \(T_x {\mathcal {M}}\) denotes the tangent space of x. Moreover, a is smooth and strongly elliptic in the following sense:

    $$\begin{aligned} \exists \, \theta > 0: \quad \forall x \in {\mathcal {M}}, \; \forall \xi \in T_x{\mathcal {M}}: \quad \xi ^\top a(x) \xi \ge \theta \Vert \xi \Vert _{{\mathbb {R}}^3}^2. \end{aligned}$$
  2. (ii)

    The coefficient \(\kappa :{\mathcal {M}}\rightarrow {\mathbb {R}}\) is smooth and bounded away from zero, i.e., there exists \(\kappa _0 \in (0,\infty )\) such that \(|\kappa (x)|\ge \kappa _0\) for all \(x \in {\mathcal {M}}\).

The conditions in Assumption 5.9 are sufficient to ensure that \(L :\dot{H}_L^1 \rightarrow (\dot{H}_L^{1})^*\) is boundedly invertible, and has a compact inverse on \(L^2({\mathcal {M}})\). This allows us to find an orthonormal basis \((e_j)_{j\in {\mathbb {N}}}\) for \(L^2({\mathcal {M}})\) and a non-decreasing sequence of positive real eigenvalues \((\lambda _j)_{j\in {\mathbb {N}}}\) of L accumulating only at infinity, as in Sect. 5.1. Moreover, fractional powers \(L^\beta \) are well-defined for all \(\beta \in {\mathbb {R}}\), the sequence of eigenvalues still satisfies Weyl’s law (5.1) (with \(d = 2\)), and a spectral mapping theorem holds, cf. [72, Theorems XII.1.3 and XII.2.1]. These facts are sufficient to repeat the proofs of Corollaries 5.2 and 5.6 yielding the analogous results, with \(d = 2\) and other obvious modifications to the conditions. In particular, the analog of Corollary 5.2 on the surface \({\mathcal {M}}\) implies regularity of the solution process in the space \(C^{n,\tau }([0,T]; L^p({\varOmega }; H^{\beta \sigma }({\mathcal {M}})))\).

An important difference from the (smooth) Euclidean setting of Assumption 5.4 is that under Assumption 5.9, the Sobolev space \(H^s({\mathcal {M}})\) and \(\dot{H}_L^s\) are isomorphic for every \(s \in [0,\infty )\), see [72, Example XII.2.1]. In other words, the absence of a boundary \(\partial {\mathcal {M}}\) implies that one does not need to exclude the exception set \({\mathfrak {E}}\) from the admissible exponents s in the analog of Lemma 5.5(c).