1 Introduction

The aim of this paper is to study regularizing properties of the Ornstein–Uhlenbeck transition semigroup associated to an abstract linear wave equation of the following form

$$\begin{aligned} \left\{ \begin{array}{l} \frac{d^{2} y}{d \tau ^{2}} ( \tau ) = - \Lambda y (\tau ) + {\dot{W}}(\tau ), \\ y\left( 0\right) =x_{0}, \;\;\; \frac{d y}{d\tau }( 0) =x_{1},\;\;\; \tau \in (0,T], \end{array} \right. \end{aligned}$$
(1.1)

where \(\Lambda : {\mathscr {D}}(\Lambda ) \subset U \rightarrow U \) is a positive self-adjoint operator on a real separable Hilbert space U such that there exists \(\Lambda ^{-1}: U \rightarrow U\) of trace class (see, for instance, Example 5.8 and Section 5.5.2 in [14, 4] and the references therein) and \(\left\{ W(\tau )= W_{\tau },\tau \ge 0\right\} \) is a cylindrical Wiener process on U (see [21] and [11]). Many linear stochastic equations modelling the vibrations of elastic structures can be written in the form (1.1). This includes stochastic wave equations like

$$\begin{aligned} \left\{ \begin{array}{l} \frac{\partial ^{2}}{\partial \tau ^{2}}y\left( \tau ,\xi \right) =\frac{\partial ^{2}}{\partial \xi ^{2}}y\left( \tau ,\xi \right) +{\dot{W}}\left( \tau ,\xi \right) , \;\;\; \xi \in (0,1),\\ y\left( \tau ,0\right) =y\left( \tau ,1\right) =0,\\ y\left( 0,\xi \right) =x_{0}\left( \xi \right) , \;\;\;\; \frac{\partial y}{\partial \tau }\left( 0,\xi \right) =x_{1}\left( \xi \right) ,\;\;\; \tau \in (0,T],\;\;\; \xi \in [0,1], \end{array} \right. \end{aligned}$$
(1.2)

where \(x_0 \in H^1_0([0,1]) \), \(x_1 \in L^2([0,1]) = U\), and stochastic plate equations in dimension 2, cf. Sect. 3 in [24].

It is known that the Ornstein–Uhlenbeck Markov semigroup associated to (1.1) is not strong Feller (cf. Theorem 9.2.1 in [12]). Here we establish regularizing effects of the Ornstein–Uhlenbeck semigroup along special directions (cf. Sect. 3). These new results allow to study in Section 4.1 a vector-valued Kolmogorov equation also considered in [24] (see (1.8)); this equation is related to the study of well-posedness of singular semilinear stochastic wave equations.

The partial smoothing seems to be an important tool. It replaces the strong Feller property which we have improperly used for the Ornstein–Uhlenbeck semigroup in Sect. 4 of [24].

We also consider interpolation results involving spaces of Hölder continuous functions along special directions (cf. Sect. 3.1).

Related results have been considered in [5] in a different context to investigate infinite dimensional elliptic equations involving the Gross Laplacian; see in particular Lemma 3.2 which is proved in [5]. We partially extend Lemma 3.2 obtaining Lemma 3.4 which deals with interpolation of vector-valued functions. We apply the new regularity results of Section 4.1 to prove strong (or pathwise) uniqueness for semilinear stochastic wave equations involving nonlinearities of Hölder type (cf. (1.7) and see, in particular, the proof of Theorem 4.7).

Since without the noise \({\dot{W}}\left( \tau ,\xi \right) \) the corresponding nonlinear deterministic equation is in general not well-posed (cf. Section 3.3 in [24]) our result is a kind of regularization by additive noise for semilinear stochastic wave equations.

Coming into the details, to study Eq. (1.1) we consider two basic Hilbert spaces: K and H. The first one is

$$\begin{aligned} K={\mathscr {D}}(\Lambda ^{1/2})\times U = V \times U. \end{aligned}$$
(1.3)

This is the usual space for the deterministic wave equation obtained removing \({\dot{W}}(\tau )\); see for instance [22, 3] and Appendix A.5.4 in [14]. This space is also denoted by \(V \oplus U\) in the literature. However solutions to stochastic wave equations (1.1) do not evolve in K but in the larger space

$$\begin{aligned} H=U \times {\mathscr {D}}(\Lambda ^{-1/2}) = U \times V', \end{aligned}$$
(1.4)

(see Example 5.8 in [14]). Here \(V'\) is the dual space of V. On the other hand, as we mentioned before, the Ornstein–Uhlenbeck semigroup has regularizing effects only along the directions of K.

Notice that when \(\Lambda = - \dfrac{d^2}{\textrm{d}x^2}\) on the interval [0, 1] with Dirichlet boundary conditions, Eq. (1.1) reduces to (1.2) with \( U = L^{2}\left( \left[ 0,1\right] \right) , \;\; {\mathscr {D}}(\Lambda ) = H_{0}^{1}\left( \left[ 0,1\right] \right) \cap H_{}^{2}\left( \left[ 0,1\right] \right) , \;\; {\mathscr {D}}(\Lambda ^{1/2}) = H_{0}^{1}\left( \left[ 0,1\right] \right) = V \) and \({\mathscr {D}}(\Lambda ^{-1/2}) = H_{}^{-1}\left( \left[ 0,1\right] \right) \).

Equation (1.1) can be written as an evolution equation in H as

$$\begin{aligned} dX_{\tau }^{0,x} = A X_{\tau }^{0,x} d\tau + GdW_\tau ,\tau \in \left[ 0,T\right] , \displaystyle X_{0}^{0,x} =x \in H, \end{aligned}$$
(1.5)

where A is the generator of a unitary group in H and \(GdW_\tau =\) \(\Big ( \begin{array}{c} 0\\ dW_\tau \end{array} \Big ) \); see Sect. 2 about notations and preliminary results.

At the beginning of Sect. 3 we fix another real and separable Hilbert space J and investigate regularizing properties of the J-valued Ornstein–Uhlenbeck semigroup \((R_\tau )\) defined by \(R_{\tau } [\Phi ] (x) = {\mathbb {E}}[\Phi (X_{\tau }^{0,x})]\), \(\tau \ge 0\), \(\Phi \in B_b(H,J)\), where \(X_{\tau }^{0,x}\) is the Ornstein–Uhlenbeck process solving (1.5).

To prove the differentiability of \(R_{\tau } [\Phi ]\), \(\tau >0\), along the directions of K, we use sharp results on the behaviour of the minimal energy for the related deterministic linear control system (A.1), see “Appendix A” (see also the appendix in [25] and the reference therein for more details).

We also study second directional derivatives of \(R_{\tau } [\Phi ]\), \(\tau >0\); see in particular our estimate for

$$\begin{aligned} \sum _{m \ge 1} \sup _{\, a \in U, |a|_U =1}| \nabla _k \nabla _{Ga} {\mathbb {E}}[\, \langle \Phi (X_{\tau }^{0, \cdot \, }), f_m\rangle _J \, ](x)|^2, \end{aligned}$$
(1.6)

\(x \in H,\) \(k \in K\), \(\tau >0\), given in Lemma 3.9 which uses the interpolation result of Lemma 3.4 (here \((f_m)\) denotes any basis of J).

In Sect. 4 we use the regularity results of Sect. 3 to prove well-posedness of the following semilinear stochastic wave equation in H:

$$\begin{aligned} dX_{\tau }^{0,x} = A X_{\tau }^{0,x} d\tau +GB(\tau ,X_{\tau }^{0,x})d\tau +GdW_\tau ,\tau \in \left[ 0,T\right] , \; \displaystyle X_{0}^{0,x} =x \in H.\nonumber \\ \end{aligned}$$
(1.7)

Here \(B:[0,T]\times H\rightarrow U\) is a measurable and bounded function, Hölder continuous of exponent \(\alpha \in (2/3,1)\) with respect to the variable x (cf. Hypothesis 2 and see also Remark 4.8). The existence of a weak solution to equation (1.7) follows by the Girsanov theorem (cf. [14, 23, 26] and [25, Section 2.1]). We stress that we obtain pathwise uniqueness of equation (1.7) without the strong Feller property for the transition Ornstein Uhlenbeck semigroup associated to (1.5). This is in contrast with other papers dealing with strong uniqueness (see, for instance, [8,9,10].

To study (1.7) we first prove in Section 4.1 existence of solutions v to the following infinite dimensional PDE of Kolmogorov type for \(x\in H\), \(t\in [0,T]\):

$$\begin{aligned} v(t,x)&=\int _t^T R_{s-t}\left[ e^{-(s-t) {A}}G B(s,\cdot )\right] (x)\,\textrm{d}s \nonumber \\&+ \int _t^T R_{s-t}\left[ e^{-(s-t){A}} \nabla ^Gv(s,\cdot ) B(s,\cdot ) \right] (x)\,\textrm{d}s; \end{aligned}$$
(1.8)

here the unknown function v takes values in K and it is regular along the directions of K. Further \( \nabla ^G v(s,x)B(s,x)\) denotes the directional derivative \({\nabla _{GB(s,x)}v(s,x)} \) \( \in K \) (see (3.4)), \((s,x) \in [0,T] \times H\). To establish Theorem  4.3 about (1.8) we use the results of Sect. 3 when \(J=K\). In Remark 4.4 we discuss the use of Lemma 3.4 in the proof of Theorem 4.3.

In Section 4.3 we recall how to get the important identity

$$\begin{aligned} X_\tau ^{0,x}&= e^{\tau A}x+e^{\tau A}v(0,x)-v(\tau ,X_\tau ^{x}) \nonumber \\&+\int _0^\tau e^{(\tau -s)A}\nabla ^G v(s,X_s^x)\;dW_s +\int _0^\tau e^{(\tau -s)A}GdW_s, \end{aligned}$$
(1.9)

which holds for any weak mild solution \((X_\tau ^{0,x})\). Note that the irregular coefficient B is not present in (1.9). This identity involves v that solves the Kolmogorov equation (1.8). Identities like (1.9) are established in [8, 17, 9, 32] by the so-called Itô-Tanaka trick which is a variant of the Zvonkin method used in [31] (see also [15]). Here we refer to [24] where (1.9) is proved by using backward stochastic differential equations (BSDEs in the sequel), which, together with the group property of \(e^{tA}\), allow to remove the “bad term” B of (1.7) (cf. Section 4.2).

Note that in contrast with previous papers which use the Itô-Tanaka trick here we have a function v which is regular only along the directions of K (see Theorem 4.3). We can use the previous identity and prove pathwise uniqueness in Theorem 4.7 noting that (see (1.7))

$$\begin{aligned} \begin{array}{l} X^{0,x_1}_{\tau } - X^{0,x_2}_{\tau } \in K,\;\;\; \tau \in [0,T], \end{array} \end{aligned}$$

if \(x_1, x_2 \in H\) and \(x_1 - x_2 \in K\) (i.e., the difference of two solutions evolves in K when the difference of the initial data is in K, even if the single solution does not evolve on K; see (4.6) and the related comments).

By Theorem  4.7, using an extension of the Yamada-Watanabe theorem (see [26]) one can obtain that (1.1) has a unique strong mild solution, for any \(x \in H\).

2 Notations and Preliminary Results

Given two real separable Hilbert spaces H and J we denote by L(HJ) the space of bounded linear operators from H to J, endowed with the usual operator norm; \(L_2(H,J)\) is the subspace of all Hilbert-Schmidt operators endowed with the Hilbert-Schmidt norm \(\Vert \cdot \Vert _{L_2(H,J)}\). Let E be a Banach space. \(B_b(H,E)\) is the space of all Borel and bounded functions from H into E endowed with the supremum norm \(\Vert \cdot \Vert _{\infty }\), \(\Vert f\Vert _{\infty }\) \(= \sup _{x \in H} |f(x)|_E\), \(f \in B_b(H,E)\). \(C_b(H,E)\) is its subspace consisting of all uniformly continuous and bounded functions from H into E. The space \(C^1_b(H,E)\) is the space of all functions in \(C_b(H,E)\) which are Fréchet differentiable on H with bounded and uniformly continuous Fréchet derivative \(\nabla f: H \rightarrow L(H,E)\); it is a Banach space endowed with the norm \(\Vert \cdot \Vert _{C^1_b}\), \(\Vert f \Vert _{C^1_b}\) \(= \Vert f \Vert _{\infty }\) \(+ \Vert \nabla f \Vert _{\infty }\), \(f \in C^1_b(H,E)\). We define, for \(0<\alpha <1\), the space \(C_b^\alpha (H, E) \) of all functions f in \(C_b(H, E)\) such that

$$\begin{aligned}{}[f]_{\alpha } = \sup _{x',\; x \in H, x-x' \not = 0} {| f(x) - f(x')|_E}\, {|x - x'|_H^{-\alpha }} < \infty . \end{aligned}$$
(2.1)

It is a Banach space endowed with the norm \(\Vert \cdot \Vert _{\alpha } = \Vert \cdot \Vert _{\infty } + [\cdot ]_{\alpha }\).

By \(C([0,T]\times H, E)\) we denote the space of continuous functions from the product space \([0,T]\times H\) into E. Moreover, \(B_b([0,T]\times H, E)\) is the Banach space of bounded Borel measurable functions from \([0,T]\times H\) into E endowed with the sup norm.

Let U be a real separable Hilbert space with inner product \(\langle \cdot ,\cdot \rangle _U \) and norm \(|\cdot |_U\). To study (1.1) we assume that

Hypothesis 1

\(\Lambda : {\mathscr {D}}(\Lambda ) \subset U \rightarrow U \) is a given positive self-adjoint operator and there exists \(\Lambda ^{-1}\) which is a trace class operator from U into U.

We also consider the Hilbert space \( V = {\mathscr {D}}(\Lambda ^{1/2}) \) \(=\)Im\((\Lambda ^{-1/2})\) endowed with the inner product

$$\begin{aligned} \langle h, k \rangle _V = \langle \Lambda ^{1/2} h, \Lambda ^{1/2} k \rangle _U, \;\; h,k \in V \end{aligned}$$

and its dual space \(V' \) which is again a Hilbert space. Note that \(|\cdot |_{V'} \) is equivalent to \(| \Lambda ^{-1/2} \cdot |_{U}\). Moreover, \(V' \) can be identified with the completion of U with respect to the norm \(| \Lambda ^{-1/2} \cdot |_{U}\) (see Section 3.4 in [30]). \(V' \) is also denoted by \({\mathscr {D}}(\Lambda ^{-1/2})\). We have \( V \subset U \simeq U' \subset V' \) with continuous inclusions; \(\Lambda \) can be extended to an unbounded self-adjoint operator on \(V'\) with domain V, that we still denote by \(\Lambda \):

$$\begin{aligned} \Lambda : V \rightarrow V'. \end{aligned}$$
(2.2)

In a complete probability space \(\left( \Omega ,{\mathscr {F}},{\mathbb {P}}\right) \) with a filtration \(\left( {\mathscr {F}}_{\tau }\right) _{\tau \ge 0}\) satisfying the usual conditions, we consider the linear stochastic wave equation (1.1) where \(\left\{ W(\tau )= W_{\tau },\tau \ge 0\right\} \) is a cylindrical Wiener process in U with respect to the filtration \(\left( {\mathscr {F}}_{\tau }\right) _{\tau \ge 0}\). The process \(W_t\) is formally given by “\(W_t \) \( = \sum _{j \ge 1} \beta _j(t) e_j\)” where \(\beta _j(t)\) are independent real Wiener processes and \((e_j)\) denotes a basis in U (see [14] for more details). We introduce, see (1.4), the reference Hilbert space \(H= U \times V' \) for the solutions to (1.1); H is endowed with the inner product \(\langle x,y \rangle _H\) \(= \langle x_1, y_1 \rangle _U \) + \(\langle x_2, y_2 \rangle _{V'}\) \(= \langle x_1, y_1 \rangle _U \) + \(\langle \Lambda ^{-1/2} x_2, \Lambda ^{-1/2} y_2 \rangle _{U}\) and norm \(|x|_H = (\langle x,x \rangle _H)^{1/2}\), \(x,y \in H\). This space is also denoted by \(U \oplus V'\).

In the sequel we will also denote \(\langle \cdot , \cdot \rangle _H\) and \(|\cdot |_H\) by \(\langle \cdot , \cdot \rangle \) and \(|\cdot |\).

According to Example 5.8 in [14], the Eq. (1.1) is well-posed in H thanks to Hypothesis 1. On the other hand, (1.1) is not well-posed in the usual space \(K = V \times U = {{\mathscr {D}}(\Lambda ^{1/2})} \times {U}\) (see (1.3)) for the deterministic wave equation: (i.e., solutions to (1.1) do not evolve in K even if \(x_0 \in V\) and \(x_1 \in U\); see Example 5.8 in [14]). Recall the inner product \(\langle x,y \rangle _K\) \(= \langle x_1, y_1 \rangle _V \) + \(\langle x_2, y_2 \rangle _{U} \), \(x,y \in K.\) In H one considers the unbounded wave operator A which generates a unitary group \(e^{tA}\):

$$\begin{aligned} \begin{array}{l} {\mathscr {D}}\left( A\right) = V \times U, A\left( \begin{array}{c} y\\ z \end{array} \right) =\Big ( \begin{array}{cc} 0 &{} I\\ -\Lambda &{} 0 \end{array} \Big ) \Big ( \begin{array}{c} y\\ z \end{array} \Big ),\text { for every }\Big ( \begin{array}{c} y\\ z \end{array} \Big ) \in {\mathscr {D}}\left( A\right) , \end{array} \end{aligned}$$
(2.3)
$$\begin{aligned} \begin{array}{l} e^{tA}\Big ( \begin{array}{c} y\\ z \end{array} \Big ) =\Big ( \begin{array}{cc} \cos \sqrt{\Lambda }t &{} \frac{1}{\sqrt{\Lambda }}\sin \sqrt{\Lambda }t\\ -\sqrt{\Lambda }\sin \sqrt{\Lambda }t &{} \cos \sqrt{\Lambda }t \end{array} \Big ) \Big ( \begin{array} {c} y\\ z \end{array} \Big ),t\in {\mathbb {R}},\;\; \left( \begin{array} {c} y\\ z \end{array} \right) \in H \end{array} \end{aligned}$$

(see also Appendix A.5.4 in [14]). Let \(G: U \rightarrow K \subset H\),

$$\begin{aligned} \begin{array}{l} G u=\Big ( \begin{array}{c} 0\\ u \end{array} \Big ) =\Big ( \begin{array}{c} 0\\ I \end{array} \Big ) u,\;\;\; u \in U. \end{array} \end{aligned}$$
(2.4)

Notice that \(e^{tA}:K\rightarrow K\) and \(e^{tA}:H\rightarrow H\), and moreover since \((e^{tA})_{t}\) is a group of linear operators, then

$$\begin{aligned} e^{tA}(K)= K, \quad e^{tA}(H)= H,\;\;\; t \in {\mathbb {R}}. \end{aligned}$$
(2.5)

We still denote by A the generator \(A_K\) of \((e^{tA})\) in K which has domain \( {\mathscr {D}}(\Lambda )\times V\). Clearly, the operator defined in (2.3) is an extension of \(A_K\).

Equation (1.1) can be rewritten in an abstract form as

$$\begin{aligned} \left\{ \begin{array}{l} dX_\tau =AX_\tau d\tau + G dW_{\tau },\tau \in \left[ 0,T\right] . \\ X_0 =x \in H, \end{array} \right. \end{aligned}$$
(2.6)

A solution to (2.6) is a particular Ornstein–Uhlenbeck process. We study (1.1) in H since the operators

$$\begin{aligned} \begin{array}{l} Q_{\tau } = \displaystyle \int _{0}^{\tau }e^{sA}GG^{*}e^{sA^{*}}\textrm{d}s, \;\;\; \tau \ge 0, \end{array} \end{aligned}$$
(2.7)

are of trace class from H into H thanks to Hypothesis 1 (cf. Example 5.8 in [14]); here \(G^*\) denotes the adjoint operator of G in H. Thus the stochastic convolution (i.e., the solution to (2.6) when \(x=0\))

$$\begin{aligned} \begin{array}{l} S_{\tau } =\displaystyle \int _{0}^{\tau }e^{\left( \tau -s\right) A}GdW_s \;\; \text {is well defined in } H. \end{array} \end{aligned}$$
(2.8)

Its law at time \(\tau \) is the Gaussian measure \(\mathcal{N}(0, Q_{\tau })\) with mean 0 and covariance operator \(Q_{\tau }\) (cf. [14]). Moreover, since \( \sup _{t \in [0,T]} \Vert e^{tA} G \Vert _{L_2(U, H)} < \infty , \;\;\; T>0, \) we can apply Theorem 5.11 in [14] and deduce that the process \((S_{\tau })\) has a continuous version with values in H.

3 The J-Valued Transition Semigroup for the Stochastic Wave Equation

Let J be a real separable Hilbert space. As in Sect. 2 we consider the Hilbert spaces \(H= U \times V',\) and \( K = V \times U \subset H.\) Moreover \((e_j)\) is a basis in U such that \((e_j) \subset {\mathscr {D}} (\Lambda ) \subset U\) and

$$\begin{aligned} \begin{array}{l} \Lambda e_j = \lambda _j e_j, \;\; \lambda _j >0,\;\; j \ge 1; \;\;\; \sum _{j \ge 1} \lambda _j^{-1} < \infty . \end{array} \end{aligned}$$
(3.1)

We will prove some regularizing effects for the Ornstein–Uhlenbeck semigroup \((R_t)\) related to stochastic wave equation (1.5) and acting on J-valued functions \(\Phi \). Recall that

$$\begin{aligned} R_{\tau }\left[ \Phi \right] \left( x\right) = R_{\tau } \Phi \left( x\right) ={\mathbb {E}}\Phi \left( X_\tau ^{0,x} \right) ,\quad \Phi \in B_b(H,J), \;\; x \in H, \; \tau \ge 0, \nonumber \\ \end{aligned}$$
(3.2)

where X, defined by (2.6), is the Ornstein–Uhlenbeck process (cf. [8]). Since X is time homogeneous, we have \(R_{\tau - t}\left[ \Phi \right] \left( x\right) ={\mathbb {E}}\Phi \left( X_\tau ^{t,x} \right) , \;\; \Phi \in B_b(H,J), \,\tau \ge t \ge 0\), \(x \in H\). Similarly, we consider the usual Ornstein–Uhlenbeck semigroup \((P_t)\) acting on scalar functions \(\phi \in B_b(H)\):

$$\begin{aligned} P_{\tau }\left[ \phi \right] \left( x\right) = P_{\tau } \phi \left( x\right) ={\mathbb {E}}\phi \left( X_\tau ^{0,x} \right) ,\quad \phi \in B_b(H), \;\; \tau \ge 0. \end{aligned}$$
(3.3)

Using also the results in “Appendix”, for \(t>0,\) we show the differentiability of \(R_t \Phi \) along the directions of K. Moreover, we prove that, for any \(x \in H,\) \(k \in K\), \(t>0\), the series in (1.6) is finite, and we provide a bound independent of x and k (see Lemma 3.11 and compare with Chapter 6 of [13] and Sect. 3 of [8]).

In order to study differentiability properties of \(R_t[\Phi ]\) for \(t>0\) we fix some basic definitions. We say that \(F: H \rightarrow J\) is differentiable along the subspace \(K = V \times U \subset H\) if there exists at any \(x \in H\) the directional derivative along any direction \(k \in K\) (i.e., \(\lim _{s \rightarrow 0} \frac{ F (x+ s k )- F (x)}{s}\in J\)). We denote the directional derivative at x along the direction \(k \in K\) as \(\nabla _k F(x) \in J.\)

If in addition \(k \mapsto \nabla _k F(x)\) belongs to L(KJ) we indicate such linear operator with \(\nabla ^K F(x)\). We say that F is K-differentiable on H if it is differentiable along the subspace K and there exists \(\nabla ^K F(x) \in L(K,J)\) for any \(x \in H\) (if \(J={\mathbb {R}}\) then \(\nabla ^K F(x)\) can be identified with an element in K by the Riesz theorem).

Note that the concept of differentiability along subspaces arises naturally in the Malliavin calculus (see also the related concept of Gross differentiability; we refer to [28] and the references therein).

Let \(G: U \rightarrow H\), \(Ga = \Big ( \begin{array}{c} 0\\ a \end{array} \Big ) \in K \subset H\), \(a \in U\). If \(F: H \rightarrow J \) is differentiable along the subspace G(U) we set

$$\begin{aligned} \nabla _a^G F(x) = \nabla _{Ga} F(x) \in J,\;\;\;\; a \in U,\; x \in H. \end{aligned}$$
(3.4)

If in addition \(a \mapsto \nabla _a^G F(x)\) belongs to L(UJ) we denote such linear operator with \(\nabla ^G F(x)\). We say that F is G-differentiable on H if it is differentiable along the subspace G(U) and there exists \(\nabla ^G F(x) \in L(U,J)\) for any \(x \in H\). Note that if \(F: H \rightarrow J\) is K-differentiable on H then it is also G-differentiable on H and \(\nabla ^G F(x) = \nabla ^K F(x) G \in L(U,J)\).

3.1 Interpolation Results Involving K-Differentiable Functions

We first introduce spaces of functions that are K-differentiable. We say that \(f \in C_K^1( H, J)\) if \(f \in C_b(H, J)\), f is K-differentiable on H and \(\nabla ^K f: H \rightarrow L(K, J)\) is uniformly continuous and bounded. It is a Banach space endowed with the norm

$$\begin{aligned} \Vert f \Vert _{C_K^1( H, J)} = \Vert \nabla ^K f \Vert _{\infty } + \Vert f \Vert _{\infty },\;\; f \in C_K^1( H,J), \end{aligned}$$

setting \(\sup _{x \in H} | \nabla ^K f(x) |_J = \Vert \nabla ^K f \Vert _{\infty }\). When \(J = {\mathbb {R}}\) we set \( C_K^1( H,{\mathbb {R}}) = C_K^1( H )\). Recall that for \(f \in C_K^1( H )\) one has: \(\nabla ^K f: H \rightarrow K\) uniformly continuous and bounded.

Let us consider the following operator \(Q: H \rightarrow H\),

$$\begin{aligned} Q= \left( \begin{array}{cc} \Lambda ^{-1} &{} 0\\ 0 &{} \Lambda ^{-1} \end{array} \right) ,\;\;\;\; Q h = \Big ( \begin{array}{c} \Lambda ^{-1} h_1 \\ \Lambda ^{-1} h_2 \end{array} \Big ),\;\;\; h = (h_1, h_2 ) \in H = U \times V'.\nonumber \\ \end{aligned}$$
(3.5)

Let \((e_j)\) be the basis in U defined in (3.1). Then \((\sqrt{\lambda _j} e_j) \) is a basis of \(V'\) and \(\{(e_j,0)\}_{j \ge 1} \cup \{(0, \sqrt{\lambda _j} e_j)\}_{j \ge 1} \) is a basis of H.

It is not difficult to check that Q is a symmetric positive trace class operator and that

$$\begin{aligned} C^1_K(H) \text {coincides with } C^1_Q(H) \text { introduced in [5]} \text { with equivalence of norms.} \end{aligned}$$
(3.6)

To this purpose we note that \(Q^{1/2} H = K\). Then we consider conditions (i), (ii) and (iii) used in the definition of \(C^1_Q(H)\) in Section 2.1 of [5]. Let \(f \in C_b (H)\). Condition (i) says that there exist all the directional derivatives of f in the directions of \(K = Q^{1/2} H\). Let \(k = Q^{1/2}h\) with \(h \in H \). The directional derivative in x along the direction k is denoted by

$$\begin{aligned} \nabla _k f(x) = \nabla _{Q^{1/2}h} f(x). \end{aligned}$$

Condition (ii) says that for any \(x \in H\), there exists \(D_Q f(x) \in H\) such that

$$\begin{aligned} \nabla _{Q^{1/2}h} f(x) = \langle D_Q f(x), h \rangle _H. \end{aligned}$$

If \(k \in K\) then \(k = Q^{1/2} h\) for a unique \(h \in H.\) We have \(\langle D_Q f(x), h \rangle _H = \langle Q^{1/2} D_Q f(x), Q^{1/2} h \rangle _K = \langle Q^{1/2} D_Q f(x), k \rangle _K \). Thus condition (ii) is equivalent to say that \( k \mapsto \nabla _k f(x)\) is linear and continuous from K into \({\mathbb {R}}\). Moreover such linear functional can be identified with \(Q^{1/2} D_Q f(x)\). According to our previous notation we can write

$$\begin{aligned} \nabla ^{K} f(x) = Q^{1/2} D_Q f(x),\;\;\; x \in H. \end{aligned}$$

Condition (iii) requires that the mapping: \(x \mapsto D_Q f(x)\) is uniformly continuous and bounded from H into H. This is equivalent to say that the mapping: \(x \mapsto Q^{1/2} D_Q f(x) = \nabla ^{K} f(x)\) is uniformly continuous and bounded from H into K. This shows (3.6).

Similarly to [5], we define, for \(0<\alpha <1\), the space \(C_K^\alpha (H, J) \) of all functions f in \(C_b(H, J)\) such that

$$\begin{aligned}{}[f]_{\alpha , K} = \sup _{k',\; k \in K, k-k' \not = 0} \frac{| f(k) - f(k')|_J}{|k - k'|_K^{\alpha }} < \infty . \end{aligned}$$
(3.7)

It is a Banach space endowed with the norm \(\Vert \cdot \Vert _{\alpha , K} = \Vert \cdot \Vert _{\infty } + [\cdot ]_{\alpha , K}\), where \( \Vert f\Vert _{\infty } = \sup _{x \in H} |f(x)|_J\).

Note that \(C_b^\alpha (H, J) \subset C_K^\alpha (H, J) \) and in general the inclusion is strict (cf. Remark 3.3).

Remark 3.1

Condition (3.7) is equivalent to

$$\begin{aligned} \sup _{x \in H = U \times V',\; k \in K, k \not = 0} \frac{| f(x+ k) - f(x)|_J}{|k|_K^{\alpha }} < \infty . \end{aligned}$$

Indeed if \(x \in H\) and \(k \in K\) there exists a sequence \((k_n) \subset K\) such that \(k_n \rightarrow x\) in H. Then by (3.7) we find \( | f(k_n+ k) - f(k_n)|_J \le C |k|_K^{\alpha }. \) Passing to the limit as \(n \rightarrow \infty \) we obtain \({| f(x+ k) - f(x)|_J} \le C |k|_K^{\alpha }.\)

The space \(C_K^\alpha (H) = C_K^\alpha (H, {\mathbb {R}})\) coincides with the space \(C_Q^\alpha (H)\) introduced in Section 2.2 in [5] as the space of all functions \(f \in C_b(H) \) such that

$$\begin{aligned}{}[f]_{\alpha , Q} = \sup _{h',\; h \in H, h-h' \not = 0} \; {| f( Q^{1/2} h ) - f(Q^{1/2} h')|}\, \cdot {|h - h'|_H^{-\alpha }} < \infty \end{aligned}$$

with equivalence of norms. By [5] we now obtain the following useful result.

Lemma 3.2

We have, for \(\alpha \in (0,1)\), with equivalence of norms,

$$\begin{aligned} (C_b(H), C^1_K(H))_{\alpha , \infty } = C^{\alpha }_K(H). \end{aligned}$$
(3.8)

Proof

The result is proved in Proposition 2.1 in [5] in the form

$$\begin{aligned} (C_b(H), C^1_Q(H))_{\alpha , \infty } = C^{\alpha }_Q(H). \end{aligned}$$

We only recall that \(f \in X_{\alpha } = (C_b(H), C^1_K(H))_{\alpha , \infty } \) if \(\Vert f \Vert _{X_{\alpha }} = \sup _{t \in (0,1]} t^{- \alpha } L(t,f)\) \( < \infty \) where \( L(t,f) =\inf \{ \Vert a \Vert _{C_b(H)} + t \Vert b \Vert _{C^1_K(H)},\) \(f = a + b,\) \(a \in C_b(H),\, b \in C_K^1(H) \}\) (see, for instance, Section 2.3 in [14]). \(\square \)

Remark 3.3

Theorem 3.1 in [27] implies that \(C^{\alpha }_b(H)\) (the space of real \(\alpha \)-Hölder continuous and bounded functions defined on H) is strictly included in \(C^{\alpha }_Q(H)\), \(\alpha \in (0,1) \). Indeed \(C^{\alpha }_b(H)\) is contained in the interpolation space \(\mathcal{D}_\mathcal{A} (\alpha /2, \infty )\) (see the notation in [27]) which by Theorem 3.1 is strictly included in \(C^{\alpha }_Q(H)\). \(\square \)

When J is infinite dimensional it is an open problem to characterize both \(\big (C_b(H,J), C^1_K(H,J) \big )_{\alpha , \infty }\) and \(\big (C_b(H,J), C^1_b(H,J) \big )_{\alpha , \infty }\). However we can prove the following inclusion which will be important for the sequel (see in particular the proof of Lemma 3.9).

Lemma 3.4

For any real separable Hilbert space J we have

$$\begin{aligned} C^{\alpha }_b(H, J) \subset \big (C_b(H,J), C^1_K(H,J) \big )_{\alpha , \infty },\;\; \alpha \in (0,1), \;\; \text { with a continous inclusion.} \end{aligned}$$
(3.9)

Proof

Let \(f \in C^{\alpha }_b(H, J)\) and \(t \in (0,1]\). Taking into account Remark 2.3.1 in [13], we prove that there exists \(a_t \in C_b(H,J)\) and \(b_t \in C^1_K(H,J)\) such that \(f = a_t + b_t\) and

$$\begin{aligned} \Vert a_t\Vert _{ C_b(H,J)} + t \Vert b_t\Vert _{C^1_K(H,J)} \le c \Vert f\Vert _{C^{\alpha }_b(H, J)} \, t^{\alpha } \end{aligned}$$
(3.10)

with \(c>0\) independent of t and f. This gives (3.9).

Let us consider the trace class operator \(Q: H \rightarrow H\) given in (3.5). Recall that Q is injective and \(Q^{1/2}(H) = K\). As in Chapter 3 of [13] we consider the heat semigroup \((V_t)\) acting on functions in \(C_b(H,J)\):

$$\begin{aligned} V_r g(x) = \int _{H} g(x+ y) \mathscr {N}(0,rQ)(\textrm{d}y), \;\; x \in H,\;\; g \in C_b(H,J),\; r \ge 0. \end{aligned}$$

For \(t \in (0,T]\) we set

$$\begin{aligned} a_t = f - V_{t^2} f,\;\;\; b_t = V_{t^2} f \end{aligned}$$

and we prove that (3.10) holds. Let us first consider \(a_t\). It is easy to prove that \(a_t \in C_b(H,J) \). Moreover,

$$\begin{aligned} |a_t (x)|_J&\le \int _{H} |f(x+ y) - f(x) |_J \, \mathscr {N}(0,t^2 Q)(\textrm{d}y) \nonumber \\&\le \Vert f\Vert _{C^{\alpha }_b(H, J)} \int _{H} | t y |_{H}^{\alpha } \, \mathscr {N}(0, Q)(\textrm{d}y) \le c_{\alpha } \Vert f\Vert _{C^{\alpha }_b(H, J)} t^{\alpha }. \end{aligned}$$

To prove that \(b_t \in C^1_K(H,J)\) we consider \(k = Q^{1/2}h \in K\) with \(h = Q^{-1/2} k\in H \). Arguing as in Theorem 3.3.3 in [14], using the Cameron-Martin theorem, one can prove that, for any \(x \in H,\) there exists the directional derivative

$$\begin{aligned} \nabla _k b_t (x)&= \lim _{s \rightarrow 0} \frac{ V_{t^2} f (x+ s Q^{1/2}h )- V_{t^2} f (x)}{s} \nonumber \\&= \frac{1}{t}\int _{H} f(x+ y) \langle (t^{2} Q)^{-1/2} y , h \rangle \mathscr {N}(0,t^2 Q)(\textrm{d}y) \nonumber \\&= \frac{1}{t}\int _{H} f(x+ y) \langle (t^{2} Q)^{-1/2} y , Q^{-1/2} k \rangle \mathscr {N}(0,t^2 Q)(\textrm{d}y). \end{aligned}$$
(3.11)

It is not difficult to prove that \(k \mapsto \nabla _k b_t (x) \) is linear and continuous from K into J. We note that

$$\begin{aligned} |\nabla _k b_t (x)|_J^2&\le \frac{1}{t^2} \Vert f\Vert _{\infty }^2 \int _{H} | \langle (t^{2} Q)^{-1/2} y , Q^{-1/2} k \rangle |^2\, \mathscr {N}(0,t^2 Q)(\textrm{d}y) \nonumber \\&\le \frac{1}{t^2} \Vert f\Vert _{\infty }^2 \, | Q^{-1/2} k|^2 = \frac{|k|_K^2}{t^2} \Vert f\Vert _{\infty }^2, \end{aligned}$$
(3.12)

where in the last passage we have used that \(| Q^{-1/2} k|= |k|_K\). We take into account (3.11) and (3.12), and for \(k \in K\), \(x,x' \in H\), we consider the difference

$$\begin{aligned}&|\nabla _k b_t (x) - \nabla _k b_t (x')|_J^2 \\&\quad \le \frac{1}{t^2} \int _{H} |f(x+ y)-f(x'+y) \langle (t^{2} Q)^{-1/2} y , Q^{-1/2} k \rangle |^2 \mathscr {N}(0,t^2 Q)(\textrm{d}y)\\&\quad \le \frac{1}{t^2} \int _{H} | \langle (t^{2} Q)^{-1/2} y , Q^{-1/2} k \rangle |^2\, \mathscr {N}(0,t^2 Q)(\textrm{d}y) \\&\quad \cdot \int _{H} | f(x + y ) - f(x' +y) |^2_J\, \mathscr {N}(0,t^2 Q)(\textrm{d}y) \le \frac{|k|_K^2}{t^2} \Vert f\Vert _{C^{\alpha }_b(H, J)}^2\, |x-x'|^{2\alpha }. \end{aligned}$$

We check easily that \(\nabla ^K b_t: H \rightarrow L(K, J)\) is uniformly continuous and bounded. Finally, since

$$\begin{aligned} \frac{1}{t}\int _{H} f(x) \langle (t^{2} Q)^{-1/2} y, h \rangle \mathscr {N}(0,t^2 Q)(\textrm{d}y) =0, \end{aligned}$$

we obtain by the Cauchy–Schwarz inequality, for any \(k \in K,\) \(x \in H,\)

$$\begin{aligned} |\nabla _k b_t (x)|_J^2&= \Big | \frac{1}{t} \int _{H} \langle (t^{2} Q)^{-1/2} y , Q^{-1/2} k \rangle \, [f(x + y ) - f(x)]\mathscr {N}(0,t^2 Q)(\textrm{d}y) \Big |_J^2 \\&\le \Vert f\Vert _{C^{\alpha }_b(H, J)}^2 \frac{1}{t^2} \int _{H} | \langle (t^{2} Q)^{-1/2} y , Q^{-1/2} k \rangle |^2\, \mathscr {N}(0,t^2 Q)(\textrm{d}y) \, \cdot \\&\quad \, \cdot \, \int _{H} | t y |_{H}^{2\alpha } \, \mathscr {N}(0, Q)(\textrm{d}y) \, \le \, c_{\alpha } \Vert f\Vert _{C^{\alpha }_b(H, J)}^2 \, t^{2 \alpha -2}. \end{aligned}$$

Collecting the previous estimates we get (3.10). \(\square \)

3.2 Partial Regularizing Properties of the Transition Semigroup

We first collect some useful properties that we prove in Appendix by classical control theoretic arguments (for more details see Appendix in [25] which is based on similar results proved in [24]).

Remark 3.5

For any \(t>0\), we have \(e^{tA}( K) = Q_t^{1/2}(H)=K\) and \(Q_t^{-1/2} e^{tA} \) belongs to L(KH) (we recall that \(Q_t\) has been defined in (2.7)). Let \(T>0\). There exists \(c= c_T>0\) such that for any \(t \in (0,T]\) we have

$$\begin{aligned} |Q^{-1/2}_t e^{tA} k|_H\le & {} \frac{c}{t^{3/2}} |k|_K,\; \;\; k \in K = V \times U; \end{aligned}$$
(3.13)
$$\begin{aligned} |Q^{-1/2}_t e^{tA} Ga|_H\le & {} \frac{c}{t^{1/2}} |a|_U,\; \;\; a \in U. \end{aligned}$$
(3.14)

Let \(\Phi \in B_b(H,J )\) and \(x \in H\). Arguing as in the proof of Theorem 6.2.2 of [13] (similar arguments are used in Section 9.4 of [14] and Section 3 of [8]) one can prove the existence of the directional derivative of \(R_t[\Phi ]\) along the directions of K:

$$\begin{aligned}{} & {} \lim _{s \rightarrow 0} \frac{ R_t[\Phi ](x+ s k )-R_t[\Phi ](x)}{s} \nonumber \\{} & {} \quad = \nabla _k R_t[\Phi ](x) = \int _H \langle Q_t^{- \frac{1}{2}}e^{tA}k,Q_t^{- \frac{1}{2}}y\rangle \, \Phi (e^{tA}x+y)\, \mathscr {N}(0,Q_t)(\textrm{d}y),\; \nonumber \\{} & {} \quad { k\in K, } \,\; t>0. \end{aligned}$$
(3.15)

In the sequel we often write \(\mu _t = \mathscr {N}(0,Q_t)\) and \(|\cdot |_H = |\cdot |\). In the next result we will use (3.15) together with the estimates (3.13) and (3.14).

Lemma 3.6

Assume Hypothesis 1 and let \(R= (R_t)\) be the OU semigroup defined in (3.2). If \(\Phi \in {B}_b(H,J)\) and \(t>0\) then \(R_t \Phi \) is K-differentiable on H. The directional derivative \(\nabla _k R_t[\Phi ](x)\in J\) is given by (3.15), for \(x \in H\). In particular \(R_t[\Phi ]\) is G-differentiable on H; further

$$\begin{aligned} \nabla ^G_{a} R_t[\Phi ](x) =\int _H \langle Q_t^{- \frac{1}{2}}e^{tA}Ga,Q_t^{- \frac{1}{2}}y\rangle \, \Phi (e^{tA}x+y)\mathscr {N}(0,Q_t)(\textrm{d}y),\,\;\; a\in U.\nonumber \\ \end{aligned}$$
(3.16)

Moreover, for \(t \in (0,T]\), we have

$$\begin{aligned}&\sup _{x \in H}| \nabla _k R_t[\Phi ](x) \vert _J \le \frac{c}{t^{\frac{3}{2}}}\Vert \Phi \Vert _\infty \vert k\vert _K,\,k\in K; \end{aligned}$$
(3.17)
$$\begin{aligned}&\sup _{x \in H} \vert \nabla ^G_{a} R_t[\Phi ](x) \vert _J \le \frac{c}{t^{\frac{1}{2}}}\Vert \Phi \Vert _\infty \vert G a \vert _{K} = \frac{c}{t^{\frac{1}{2}}}\Vert \Phi \Vert _\infty \vert a \vert _{U},\, a \in U. \end{aligned}$$
(3.18)

If in addition \(\Phi \in {C}_b(H,J)\) then \( \nabla ^K R_t [\Phi ] \in C_b(H, L(K,J)) \), \(\nabla ^G R_t[\Phi ] \in C_b(H, L(U,J)) \) for \(t >0\).

Proof

Let us fix \(t \in (0,T] \) and \(x \in H\). The integral in (3.15) defines a linear operator in L(KJ). Let

$$\begin{aligned} I_{t,x}k:=\int _H \langle Q_t^{-\frac{1}{2}}e^{tA}k,Q_t^{-\frac{1}{2}}y\rangle \Phi (e^{tA}x+y)\mathscr {N}(0,Q_t)(\textrm{d}y),\;\; k \in K. \end{aligned}$$

We have the well-known estimate (cf. the proof of Theorem 6.2.2 in [13])

$$\begin{aligned} \vert I_{t,x}k\vert _J{} & {} \le \int _H \vert \langle Q_t^{-\frac{1}{2}}e^{tA}k,Q_t^{-\frac{1}{2}}y\rangle \Phi (e^{tA}x+y)\vert _J \mathscr {N}(0,Q_t)(\textrm{d}y) \nonumber \\{} & {} \le \Vert \Phi \Vert _\infty \Big (\int _H\vert \langle Q_t^{-\frac{1}{2}}e^{tA}k,Q_t^{-\frac{1}{2}}y\rangle \vert ^2 \mathscr {N}(0,Q_t)(\textrm{d}y) \Big )^{1/2}\nonumber \\{} & {} =\Vert \Phi \Vert _\infty \vert Q_t^{-\frac{1}{2}}e^{tA}k\vert _H \nonumber \\{} & {} \le \frac{c}{t^{\frac{3}{2}}}\Vert \Phi \Vert _\infty \vert k\vert _K. \end{aligned}$$
(3.19)

Similarly, we get (3.18) using (3.14) since

$$\begin{aligned} \int _H \vert \langle Q_t^{-\frac{1}{2}}e^{tA} G a,Q_t^{-\frac{1}{2}}y\rangle \vert ^2 \mathscr {N}(0,Q_t)(\textrm{d}y) \le |Q_t^{-\frac{1}{2}}e^{tA} G a|^2 \le \frac{c |a|^2_U }{t},\;\; a \in {U}.\nonumber \\ \end{aligned}$$
(3.20)

Computing the directional derivative as in (3.15) we obtain the differentiability of \(R_t[\Phi ]\) at x along the directions of K. We also obtain that \(R_t[\Phi ]\) is G-differentiable and K-differentiable on H.

If \(\Phi \in {C}_b(H,J)\) we compute, for any \(k \in K\), \(|k|_K=1\), \(z \in H\),

$$\begin{aligned}&|I_{t,x}k- I_{t, x+ z}k |^2_J \\&\quad \le \int _H\vert \langle Q_t^{-\frac{1}{2}}e^{tA}k,Q_t^{-\frac{1}{2}}y\rangle \vert ^2 \mathscr {N}(0,Q_t)(\textrm{d}y) \\&\qquad \cdot \, \int _H | \Phi (e^{tA}x+y) - \Phi (e^{tA}x + e^{tA}z +y) |^2_J \mathscr {N}(0,Q_t)(\textrm{d}y) \\&\quad \le \frac{c^2 |k|^2_K}{t^{3} } \, \int _H \sup _{x\in H}| \Phi (e^{tA}x+y) - \Phi (e^{tA}x + e^{tA}z +y) |^2_J \, \mathscr {N}(0,Q_t)(\textrm{d}y) \end{aligned}$$

and so we obtain easily \(\lim _{z \rightarrow 0} \; \sup _{y \in H} \, \sup _{|k|_K =1} |I_{t,y}k - I_{t, y+ z}k | _J=0\) by the dominated convergence theorem. We deduce \( \nabla ^K R_t[\Phi ] \in C_b(H, L(K,J))\) and \(\nabla ^G R_t[\Phi ] \in C_b(H, L(U,J)). \) \(\square \)

In a similar way we get

Lemma 3.7

Under the assumptions of Lemma 3.6 let \(t >0\). If \(\Phi \in C_b(H, J)\) and \(\xi \in U\) the J-valued mapping:

$$\begin{aligned} x \mapsto \nabla _{\xi }^G R_t[\Phi ](x) \end{aligned}$$

is K-differentiable on H. The second order directional derivative is

$$\begin{aligned} \nabla _k \nabla _{\xi }^G R_t[\Phi ](x)= & {} \int _H \big (\langle \Gamma _t k,Q_t^{-\frac{1}{2}}y\rangle \, \langle \Gamma _t G \xi ,Q_t^{-\frac{1}{2}}y\rangle \nonumber \\{} & {} - \langle \Gamma _t k, \Gamma _t G \xi \rangle \big ) \Phi (e^{tA}x+y)\mu _t(\textrm{d}y), \end{aligned}$$
(3.21)

for \(x\in H, { k \in K}\), \(\xi \in U\), and where \(\Gamma _t:=Q_t^{-\frac{1}{2}}e^{tA}\) is well defined, see Remark 3.5. Moreover, for each \(x \in H\), \(k \in K\), the map: \( \xi \rightarrow \nabla _k \nabla _{ \xi \,}^G R_t[\Phi ](x)\) belongs to L(UJ) and, for any \(t \in (0,T]\),

$$\begin{aligned}{} & {} \sup _{x \in H} \Vert \nabla _k \nabla _{\cdot }^G R_t[\Phi ](x) \Vert _{L(U,J)} \le \frac{c_T |k|_{K}}{t^{2}}\Vert \Phi \Vert _\infty , \, { k\in K. } \end{aligned}$$
(3.22)
$$\begin{aligned}{} & {} \lim _{x \rightarrow 0} \sup _{y \in H } \, \Vert \nabla ^K \nabla _{\xi }^G R_t[\Phi ](x +y) - \nabla \nabla _{\xi }^G R_t[\Phi ](y) \Vert _{L(K,J)} \nonumber \\{} & {} \quad = \lim _{x \rightarrow 0} \sup _{y \in H } \sup _{|k|_K = 1} \, | \nabla _k \nabla _{\xi }^G R_t[\Phi ](x +y) - \nabla _k \nabla _{\xi }^G R_t[\Phi ](y) |_{J} =0,\;\; \xi \in U.\nonumber \\ \end{aligned}$$
(3.23)

Proof

Let us fix \(T>0\) and \(t \in (0,T]\), \(x \in H\). Let \(\xi \in U\). First define \(\Gamma _{t,x,k, \xi }\) as the integral in the right hand side of (3.21). Proceeding as in the proof of Lemma 3.6 it is not difficult to show that

$$\begin{aligned} k \mapsto \Gamma _{t,x,k, \cdot } \;\; \text { is defined from } K \text { into } L(U,J). \end{aligned}$$
(3.24)

By (3.13) and using the Cauchy-Schwarz inequality, we get

$$\begin{aligned} |\Gamma _{t,x,k, \xi } |^2_J&\le \Big | \int _H [\langle \Gamma _t k ,Q_t^{-\frac{1}{2}}y\rangle \, \langle \Gamma _t G \xi ,Q_t^{-\frac{1}{2}}y\rangle - \langle \Gamma _t k, \Gamma _t G \xi \rangle \big ] \, \Phi (e^{tA}x+y) \, \mu _t(\textrm{d}y) \Big |_J^2 \nonumber \\&\le \int _H \big |\langle \Gamma _t k ,Q_t^{-\frac{1}{2}}y\rangle \, \langle \Gamma _t G \xi ,Q_t^{-\frac{1}{2}}y\rangle - \langle \Gamma _t k, \Gamma _t G \xi \rangle \big |^2 \mu _t(\textrm{d}y)\nonumber \\&\quad \cdot \int _H \, |\Phi (e^{tA}x+y)|^2_J \, \mu _t(\textrm{d}y) \nonumber \\&\le \frac{c |k|^2_K }{t^{4}} |\xi |^2_U \Vert \Phi \Vert _\infty ^2. \end{aligned}$$
(3.25)

Thus we have proved (3.22). Arguing as in Section 9.4 of [14] and as in Sect. 3 of [8] we find that

$$\begin{aligned} \lim _{s \rightarrow 0} \frac{ \nabla ^G_{\xi } R_t[\Phi ](x+ s k )- \nabla ^G_{\xi } R_t[\Phi ](x)}{s} = \Gamma _{t,x,k, \xi }, \;\; k \in K. \end{aligned}$$

Moreover, arguing as in the proof of Lemma 4.2 in [24], for any \(z \in H\), \(\xi \in U\),

$$\begin{aligned}{} & {} |\Gamma _{t,x, k, \xi } - \Gamma _{t, x+ z, k, \xi } |^2_J \\{} & {} \quad \le \frac{c |k|^2_K}{t^{4}} \vert \xi \vert _{U}^2 \int _H | \Phi (e^{tA}x+y) - \Phi (e^{tA}x + e^{tA}z +y) |^2_J \mu _t(\textrm{d}y) \end{aligned}$$

and so

$$\begin{aligned} \lim _{z \rightarrow 0} \sup _{x \in H} \sup _{|k|_K =1} |\Gamma _{t,x, k, \xi } - \Gamma _{t, x+ z, k, \xi } |^2_J =0. \end{aligned}$$
(3.26)

This shows in particular that the mapping \( x \mapsto \nabla _{\xi }^G R_t[\Phi ](x) \) verifies (3.23). \(\square \)

Now we improve the previous estimates in the case when \(\Phi \) is Hölder continuous along the directions of K using Lemma 3.2.

Lemma 3.8

Let \(T>0.\) Under the assumptions of Lemma 3.6 let \(\Phi \in C^{\alpha }_K (H, J)\), \(\alpha \in (0,1)\), see (3.7) and Remark 3.1. We have all the assertions of Lemmas 3.6 and 3.7 and the following new estimates, for \(t \in (0,T]\),

$$\begin{aligned}&\sup _{x \in H} \vert \nabla _k R_t[\Phi ](x)\vert _J \, \le \frac{c}{t^{\frac{3}{2} (1- \alpha )} }\Vert \Phi \Vert _{\alpha , K} \vert k\vert _K,\,k\in K; \nonumber \\&\quad \sup _{x \in H} \Vert \nabla _k \nabla _{\cdot }^G R_t[\Phi ](x) \Vert _{L(U,J)} \le \frac{c}{t^{\frac{4 - 3\alpha }{2}}}\Vert \Phi \Vert _{\alpha ,K} \vert k\vert _K,\,k\in K. \end{aligned}$$
(3.27)

Proof

Let us fix \(t \in (0,T]\), \(k \in K\) and \(\xi \in U\). Using the OU process X defined by (4.5) we can define the Ornstein–Uhlenbeck semigroup \((P_t)\) acting on scalar functions \(\phi \in B_b(H)\) (see (3.3)).

For \(h \in J\), we introduce the scalar function \( \Phi _h(x) = \langle \Phi (x), h \rangle _J \), \(x \in H\), which belongs to \(C_K^{\alpha }(H)\) with \(\Vert \Phi _h \Vert _{{\alpha ,K}} \) \(\le \Vert \Phi \Vert _{{\alpha ,K}} \, |h|_J \). We note arguing as in Sect. 3 of [8] that

$$\begin{aligned} \langle \nabla _k R_t[\Phi ](x), h \rangle _J = \nabla _k P_t[\Phi _h](x),\;\; x \in H. \end{aligned}$$

To prove the first estimate we consider the linear operators

$$\begin{aligned} \nabla _k P_t: C^1_K (H) \rightarrow {C}_b (H),\;\;\; \nabla _k P_t: {C}_b(H) \rightarrow {C}_b (H). \end{aligned}$$

These operators are well defined by Lemma 3.6. When \(\phi \in C^1_K(H)\) we find that (recall that \(e^{tA} k \in K\))

$$\begin{aligned} \nabla _k P_t[\phi ](x)= & {} \lim _{s \rightarrow 0}\int _H \frac{ \phi (e^{tA}x + s e^{tA}k +y) - \phi (e^{tA}x +y)}{s} \mu _t(\textrm{d}y) \\= & {} \int _H \nabla _{ e^{tA} k} \, \phi (e^{tA}x+y) \, \mu _t(\textrm{d}y) = \int _H \langle \nabla ^{K } \, \phi (e^{tA}x+y), e^{tA} k \rangle _K \, \mu _t(\textrm{d}y) \end{aligned}$$

(by the Riesz theorem we identify \(\nabla ^{K } \, \phi (e^{tA}x+y)\) with an element in K). We get the estimate

$$\begin{aligned} \sup _{x \in H} \vert \nabla _k P_t[\phi ](x) \vert \le C \Vert \nabla ^K \phi \Vert _\infty \vert k\vert _K,\;\;\; \phi \in C^1_K(H). \end{aligned}$$
(3.28)

On the other hand we have (cf. (3.17))

$$\begin{aligned} \sup _{x \in H}| \nabla _k P_t[f](x) \vert \le \frac{c}{t^{\frac{3}{2}}}\Vert f \Vert _\infty \vert k\vert _K,\;\;\; f \in C_b(H). \end{aligned}$$
(3.29)

Interpolating between (3.28) and (3.29) (see Theorem A.1.1 in [14]) we obtain that, for any \(\alpha \in (0,1),\)

$$\begin{aligned} \nabla _k P_t: (C_b(H), C^1_K(H))_{\alpha , \infty } \, \rightarrow \, {C}_b (H) \end{aligned}$$

is linear and bounded; further

$$\begin{aligned} \sup _{x \in H} \vert \nabla _k P_t[\psi ](x)\vert \le \frac{{\tilde{c}}}{t^{\frac{3}{2} (1- \alpha )} }\Vert \psi \Vert _{ (C_b(H), C^1_K(H))_{\alpha , \infty } } \vert k\vert _K, \;\;\; \psi \in (C_b(H), C^1_K(H))_{\alpha , \infty }. \end{aligned}$$

Now thanks to Lemma 3.2 we deduce

$$\begin{aligned} \sup _{x \in H} \vert \nabla _k P_t[\psi ](x)\vert \le \frac{c}{t^{\frac{3}{2} (1- \alpha )} }\Vert \psi \Vert _{\alpha , K} \, \vert k\vert _K, \;\;\; \psi \in C^{\alpha }_K(H). \end{aligned}$$

If we consider now \(\psi = \Phi _h\), we have, for each \(x \in H\), \(h \in J\),

$$\begin{aligned}&|\langle \nabla _k R_t[\Phi ](x), h \rangle _J | = \vert \nabla _k P_t[\Phi _h](x)\vert \\&\quad \le \frac{c}{t^{\frac{3}{2} (1- \alpha )} }\Vert \Phi _h\Vert _{\alpha , K} \, \vert k\vert _K \le \frac{c}{t^{\frac{3}{2} (1- \alpha )} }\Vert \Phi \Vert _{\alpha , K} \, \vert k\vert _K |h|_J. \end{aligned}$$

By taking the supremum over \(\{h \in J \,: \, |h|_J =1\}\) we get the first estimate in (3.27).

To prove the second estimate we fix \(t>0,\) \(k \in K\), with \(|k|_K=1\), \(\xi \in U\) with \(|\xi |_U=1\) and argue as before. We first introduce the following linear operators (cf. Lemma 3.7)

$$\begin{aligned} \nabla _k \nabla _{\xi }^G P_t: C^1_K(H) \rightarrow {C}_b (H),\;\;\; \nabla _k \nabla _{\xi }^G P_t: {C}_b(H) \rightarrow {C}_b (H). \end{aligned}$$
(3.30)

When \(\phi \in C^1_K(H)\) we know that

$$\begin{aligned} \nabla _k \nabla _{\xi }^G P_t[\phi ](x) =\int _H \langle \Gamma _t G \xi ,Q_t^{-\frac{1}{2}}y\rangle \nabla _{e^{tA}k} \phi (e^{tA}x+y) \mu _t(\textrm{d}y). \end{aligned}$$
(3.31)

Moreover, we have, with \(\Gamma _t = Q^{-1/2}_t e^{tA}\),

$$\begin{aligned} \sup _{x \in H} | \nabla _k \nabla ^G_{\xi } P_t[\phi ](x)|^2 \le \Vert \phi \Vert _{C^1_K}^2 \, \vert Q^{-1/2}_t e^{tA} G \xi \vert _{H}^2 \le \frac{c}{t} |\xi |_U^2 \Vert \phi \Vert _{C^1_K}^2= \frac{c}{t} \Vert \phi \Vert _{C^1_K}^2.\nonumber \\ \end{aligned}$$
(3.32)

On the other hand if \(\phi \in C_b(H)\) then

$$\begin{aligned}&\sup _{x \in H} | \nabla _k \nabla ^G_{\xi } P_t[\phi ](x)|^2 \le C \Vert \phi \Vert _{\infty }^2 \, | \Gamma _t k|_H^2 \, \vert Q^{-1/2}_t e^{tA} G \xi \vert _{H}^2 \nonumber \\&\quad \le c \Vert \phi \Vert _{\infty }^2 \; \frac{1}{t^{3}} |k|_K^2 \, \frac{1}{t} |\xi |_U^2 \le c \Vert \phi \Vert _{\infty }^2 \; \frac{1}{t^{4}}. \end{aligned}$$
(3.33)

Interpolating between (3.32) and (3.33) as we have done before we get

$$\begin{aligned} \sup _{x \in H} \vert \nabla _k (\nabla _{\xi }^G P_t[\psi ])(x) \vert _{} \le \frac{c}{t^{\frac{4 - 3\alpha }{2}}} \Vert \psi \Vert _{{\alpha , K }},\;\; \psi \in C^{\alpha }_K(H). \end{aligned}$$
(3.34)

Now for \(x \in H\), \(k \in K\), \(\Phi \in C^{\alpha }_K (H, J)\), we compute

$$\begin{aligned}&\Vert \nabla _k (\nabla _{}^G R_t[\Phi ])(x) \Vert _{L(U,J)}^2 = \sup _{a \in U, \, |a|_U =1} |\nabla _k (\nabla _{a}^G R_t[\Phi ])(x)|^2_J \\&\quad = \sup _{|a|_U =1} \, \, \, \sup _{h \in J, \; |h|_J =1}| \langle \nabla _k (\nabla _{a}^G R_t[\Phi ])(x), h \rangle _J |^2 = \sup _{|a|_U =1} \sup _{|h|_J=1}| \nabla _k (\nabla _{a}^G P_t[\Phi _h])(x) |^2 \\&\quad \le \frac{c\, \vert k\vert ^2_K}{t^{{4 - 3\alpha }}} \; \sup _{|h|_J=1} \Vert \Phi _h\Vert _{{\alpha , K }}^2 \; \le \; \frac{c\, \vert k\vert ^2_K }{t^{{4 - 3\alpha }}} \, \Vert \Phi \Vert _{{\alpha , K}}^2. \end{aligned}$$

The second estimate in (3.27) follows easily. \(\square \)

The next result is crucial for our approach to get pathwise uniqueness (see in particular Theorem 4.3 and the proof of Theorem 4.7). We can only prove the result when \(\Phi \in C^{\alpha }_b (H, J)\) using Lemma 3.4. We do not know if such result holds more generally when \(\Phi \in C^{\alpha }_K (H, J)\) (see the end of the next proof).

We fix a basis \((f_m)\) of J and set \(\Phi _m = \langle \Phi , f_m \rangle _J \). We will use the OU semigroup \((P_t)\) given in (3.3).

Lemma 3.9

Let \(T>0.\) Under the assumptions of Lemma 3.6 let \(\Phi \in C^{\alpha }_b (H, J)\), \(\alpha \in (0,1)\). We have, for \(t \in (0,T]\), \(k \in K\),

$$\begin{aligned} \Big ( \sum _{m \ge 1} \sup _{|a|_U =1}| \nabla _k \nabla _{a}^G P_t[\Phi _m](x)|^2 \Big )^{1/2}\le \frac{C}{t^{\frac{4 - 3\alpha }{2}}} |k|_K \, \Vert \Phi \Vert _{ C^{\alpha }_b(H, J)}. \end{aligned}$$
(3.35)

where \(C >0\) is independent of xkt, \(\Phi \) and the basis \((f_m)\) of J.

Proof

We recall that for \(k \in K\), \(a \in U\)

$$\begin{aligned}&\nabla _k \nabla _{a}^G P_t[\Phi _m](x) = \int _H \Big (\langle \Gamma _t k,Q_t^{-\frac{1}{2}}y\rangle \, \langle \Gamma _t G a ,Q_t^{-\frac{1}{2}}y\rangle -\\&\quad - \langle \Gamma _t k, \Gamma _t G a \rangle \Big ) \, \Phi _m (e^{tA}x +y) \, \mu _t(\textrm{d}y). \end{aligned}$$

We fix \(k \in K\). We have by the Cauchy-Schwarz inequality (cf. (3.25)) for any \(x \in H,\) \(m \ge 1,\)

$$\begin{aligned}&\sum _{m \ge 1} \sup _{|a|_U =1}| \nabla _k \nabla _{a}^G P_t[\Phi _m](x)|^2 \nonumber \\&\quad \, \le \sum _{m \ge 1} \sup _{|a|_U =1} \int _H \big | \big (\langle \Gamma _t k,Q_t^{-\frac{1}{2}}y\rangle \, \langle \Gamma _t G a ,Q_t^{-\frac{1}{2}}y\rangle - \langle \Gamma _t k, \Gamma _t G a \rangle \big ) \big |^2 \, \mu _t(\textrm{d}y)\, \cdot \nonumber \\&\qquad \; \cdot \, \int _H |\Phi _m (e^{tA}x +y)|^2 \, \mu _t(\textrm{d}y) \nonumber \\ {}&\quad \, \le \frac{c}{t^4} |k|_K^2 \sum _{m \ge 1} \int _H |\Phi _m (e^{tA}x +y)|^2 \, \mu _t(\textrm{d}y) = \frac{c}{t^4} |k|_K^2 \int _H |\Phi (e^{tA}x +y)|^2_J \, \mu _t(\textrm{d}y). \end{aligned}$$
(3.36)

Now we fix also \(x \in H\). For any \(l \in J\) we define \(\Phi _l: H \rightarrow {\mathbb {R}}\),

$$\begin{aligned} \Phi _l(y) = \langle \Phi (y), l\rangle _J,\;\;\; l \in J. \end{aligned}$$

We have \(\Phi _m = \Phi _ {f_m} \).

We can consider the linear operator \(T_{x,k}: C_b^{} (H, J) \rightarrow L_2 (J, U)\),

$$\begin{aligned} T_{x,k}(\Phi ) (l) = \nabla _k \nabla _{\cdot }^G P_t[ \Phi _l ](x) \in U, \;\;\; \Phi \in C_b^{} (H, J),\;\; l \in J \end{aligned}$$

(we identify U with \(L(U, {\mathbb {R}})\)). We have

$$\begin{aligned}&| T_{x,k}(\Phi ) (l)|^2_U \\&\quad \,= \sup _{|a|_U =1} \Big | \int _H \big (\langle \Gamma _t k,Q_t^{-\frac{1}{2}}y\rangle \, \langle \Gamma _t G a ,Q_t^{-\frac{1}{2}}y\rangle \\&\quad - \langle \Gamma _t k, \Gamma _t G a \rangle \big ) \, \langle \Phi (e^{tA}x +y) , l \rangle _J \,\, \mu _t(\textrm{d}y) \Big |^2. \end{aligned}$$

and

$$\begin{aligned} \Vert T_{x,k}(\Phi ) \Vert _{ L_2 (J, U)}^2 = \sum _{m \ge 1} | T_{x,k}(\Phi ) (f_m)|^2_U = \sum _{m \ge 1} \sup _{|a|_U=1} | \nabla _k \nabla _{a}^G P_t[ \Phi _m ](x)|^2. \end{aligned}$$

By the bound in (3.36) we deduce that \(\Vert T_{x,k}(\Phi ) \Vert _{ L_2 (J, U)}^2 \le \frac{c}{t^4} |k|_K^2 \, \Vert \Phi \Vert _{\infty }^2 \), where c is independent of \(t>0\), \(x \in H\), \(\Phi \) and \(k \in K.\) The linear operator: \(T_{x,k}: C_b^{} (H, J) \rightarrow L_2 (J, U)\) is well defined and continuous; we have

$$\begin{aligned} \Vert T_{x,k} \Vert _{L ( C_b^{} (H, J),L_2 (J, U) )} \le \frac{c}{t^2} |k|_K. \end{aligned}$$
(3.37)

Now if \(\Phi \in C^1_{K}(H, J)\) we find for \(k \in K\), \(a \in U\) (cf. (3.31) and (3.20))

$$\begin{aligned}&\sup _{a \in U, |a|_U =1} | \nabla _k \nabla _{a}^G P_t[\Phi _m ](x) |^2 \\&\quad \, = \sup _{a \in U, |a|_U =1} | |\int _H \langle \Gamma _t G a,Q_t^{-\frac{1}{2}}y\rangle \, \langle \nabla ^{K}\Phi _m(e^{tA}x+y), {e^{tA}k} \rangle _K \; \mu _t(\textrm{d}y)|^2\\&\quad \, \le \sup _{|a|_U =1} \int _H \big | \langle \Gamma _t G a ,Q_t^{-\frac{1}{2}}y\rangle \big | ^2 \, \mu _t(\textrm{d}y)\, \cdot \, \int _H | \langle \nabla ^{K}\Phi _m(e^{tA}x+y), {e^{tA}k} \rangle _K |^2 \, \mu _t(\textrm{d}y) \\&\quad \, \le \frac{c}{t} \, \cdot \, \int _H | \langle \nabla ^{K}\Phi _m(e^{tA}x+y), {e^{tA}k} \rangle _K |^2 \, \mu _t(\textrm{d}y) \end{aligned}$$

and so

$$\begin{aligned}&\sum _{m \ge 1} \sup _{ |a|_U =1 } | \nabla _k \nabla _{a}^G P_t[\Phi _m](x)|^2 \le \frac{c}{t} \sum _{m \ge 1} \int _H | \langle \nabla ^{K}\Phi _m(e^{tA}x+y), {e^{tA}k} \rangle _K |^2 \, \mu _t(\textrm{d}y) \\&\quad = \frac{c}{t} \int _H | \nabla ^{K} \Phi (e^{tA}x +y) [e^{tA}k] |^2_J \, \mu _t(\textrm{d}y). \end{aligned}$$

Since \(| \nabla ^{K} \Phi (e^{tA}x +y) [e^{tA}k] |_J \le \Vert \nabla ^{K} \Phi (e^{tA}x +y)\Vert _{L(K, J)} \, |k |_K \), \(t \ge 0,\) it follows that

$$\begin{aligned} \sum _{m \ge 1} \sup _{|a|_U =1}| \nabla _k \nabla _{a}^G P_t[\Phi _m](x)|^2 \le \frac{c}{t} |k|_K^2 \, \Vert \Phi \Vert _{C^1_K(H, J)}^2. \end{aligned}$$
(3.38)

Hence

$$\begin{aligned} \Vert T_{x,k} \Vert _{L ( C_K^{1} (H, J),L_2 (J, U) )} \le \frac{c}{t^{1/2}} |k|_K. \end{aligned}$$
(3.39)

Interpolating between (3.37) and (3.39) as in the proof of Lemma 3.8, for any \(\Phi \in \big (C_b(H,J), C^1_K(H,J) \big )_{\alpha , \infty }\), we get

$$\begin{aligned} \Big ( \sum _{m \ge 1} \sup _{|a|_U =1}| \nabla _k \nabla _{a}^G P_t[\Phi _m](x)|^2 \Big )^{1/2} \, \le \frac{c}{t^{\frac{4 - 3\alpha }{2}}} |k|_K \, \Vert \Phi \Vert _{ \big (C_b(H,J), C^1_K(H,J) \big )_{\alpha , \infty }}. \end{aligned}$$
(3.40)

Now we consider any \(\Phi \in C^{\alpha }_b(H, J) \subset \big (C_b(H,J), C^1_K(H,J) \big )_{\alpha , \infty } \) (see Lemma 3.4). We finally obtain, for any \(t \in (0,T],\)

$$\begin{aligned} \Big (\sum _{m \ge 1} \sup _{|a|_U =1}| \nabla _k \nabla _{a}^G P_t[\Phi _m](x)|^2 \Big )^{1/2} \, \le \frac{C}{t^{\frac{4 - 3\alpha }{2}}} |k|_K \, \Vert \Phi \Vert _{ C^{\alpha }_b(H, J)}. \end{aligned}$$

\(\square \)

Remark 3.10

We provide here an equivalent formulation of Lemma 3.9 when \(J=K\) (this will be used in the proof of Theorem 4.3). Let \(\Phi \in C^{\alpha }_b (H, K)\), \(\alpha \in (0,1)\). Let us fix \(t \in (0,T]\), \(x \in H\) and \(k \in K\).

Recall the notation \(\Phi _l \) \( = \langle \Phi , l \rangle _K \), \(l \in K.\) We know that the linear operator \(\nabla _k \nabla _{}^G R_t[ \Phi ](x)\) is well defined from K into U by the formula

$$\begin{aligned} l \mapsto \nabla _k \nabla _{}^G P_t[ \Phi _l](x) \end{aligned}$$

(cf. Lemma 3.7; note that the operator \(a \mapsto \nabla _k \nabla _{a}^G P_t[ \Phi _l](x)\) belongs to \(L (U, {\mathbb {R}})\) and so it can be identified with an element of U). Assertion of Lemma 3.9 is equivalent to say that

$$\begin{aligned}&\nabla _k \nabla _{}^G R_t[ \Phi ](x) \in L_2(K,U) \;\; \text {and} \nonumber \\&\Vert \nabla _k \nabla _{}^G R_t[ \Phi ](x) \Vert _{L_2(K,U)} = \Big (\sum _{m \ge 1} \sup _{|a|_U =1}| \nabla _k \nabla _{a}^G P_t[\langle \Phi , f_m \rangle _K ](x)|^2 \Big )^{1/2} \nonumber \\&\quad \le \frac{C}{t^{\frac{4 - 3\alpha }{2}}} |k|_K \, \Vert \Phi \Vert _{ C^{\alpha }_b(H, K),}\;\; t \in (0,T]; \end{aligned}$$
(3.41)

here \((f_m)\) is a basis in K. Recall that \(C >0\) is independent of xkt and \(\Phi \). \(\square \)

We will also use the following additional regularity result.

Lemma 3.11

Let \(T>0.\) Under the assumptions of Lemma 3.6 let \(\Phi \in C^{\alpha }_b (H, J)\), \(\alpha \in (0,1)\). We have the following estimate, for \(t \in (0,T]\), \(x, y \in H\)

$$\begin{aligned} \Vert \nabla _{}^G R_t[\Phi ](x) - \nabla _{}^G R_t[\Phi ](x') \Vert _{L(U,J)} \le \frac{c}{t^{\frac{1}{2}}} |x-x'|^{\alpha } \Vert \Phi \Vert _{\alpha }. \end{aligned}$$
(3.42)

Proof

The assertion follows easily by the formula

$$\begin{aligned}&\nabla ^G_{a} R_t[\Phi ](x) - \nabla ^G_{a} R_t[\Phi ](x') \\&\quad \;\,=\int _H \langle Q_t^{-\frac{1}{2}}e^{tA}Ga,Q_t^{-\frac{1}{2}}y\rangle \, [\Phi (e^{tA}x+y) - [\Phi (e^{tA}x'+y)] \mathscr {N}(0,Q_t)(dy), \end{aligned}$$

\( a\in U \), using that \( |\Phi (e^{tA}x+y) - \Phi (e^{tA}x'+y)|_J \le \Vert \Phi \Vert _{\alpha } |x-x'|^{\alpha },\) \(t \ge 0, \, y \in H. \) \(\square \)

4 Well-Posedness of Singular Wave Equations

In this section we apply our regularity results to prove strong well-posedness of a semilinear abstract stochastic wave equation. Let us consider the following SPDE

$$\begin{aligned} \left\{ \begin{array}{l} \frac{\partial ^{2} y}{\partial \tau ^{2}}\left( \tau \right) =-\Lambda y\left( \tau \right) +b\left( \tau ,y(\tau )\right) +{\dot{W}}\left( \tau \right) , \;\;\; \\ y\left( 0\right) =x_{0} \in U, \;\;\; \;\; \frac{\partial y}{\partial \tau }\left( 0\right) =x_{1} \in V',\;\;\; \tau \in (0,T]. \end{array} \right. \end{aligned}$$
(4.1)

This equation is a semilinear extension of (1.1), with a drift \(b:[0,T] \times U \rightarrow U\) which can be a bounded measurable function, Hölder continuous of exponent \(\alpha \in (2/3,1)\) with respect to the y-variable, uniformly in \(t \in [0,T]\).

An example of Eq. (4.1) is obtained by adding the term \(b\left( \tau ,\xi , y(\tau , \xi ) \right) \) to to the stochastic wave equation (1.2), i.e.,

$$\begin{aligned} \left\{ \begin{array}{l} \frac{\partial ^{2}}{\partial \tau ^{2}}y\left( \tau ,\xi \right) =\frac{\partial ^{2}}{\partial \xi ^{2}}y\left( \tau ,\xi \right) +b\left( \tau ,\xi , y(\tau , \xi ) \right) +{\dot{W}}\left( \tau ,\xi \right) , \;\;\; \xi \in (0,1),\\ y\left( \tau ,0\right) =y\left( \tau ,1\right) =0,\\ y\left( 0,\xi \right) =x_{0}\left( \xi \right) , \;\;\; \frac{\partial y}{\partial \tau }\left( 0,\xi \right) =x_{1}\left( \xi \right) ,\;\;\; \tau \in (0,T],\;\;\; \xi \in [0,1]. \end{array} \right. \end{aligned}$$
(4.2)

Here \(b: \left[ 0,T\right] \times \left[ 0,1\right] \times {\mathbb {R}} \rightarrow {\mathbb {R}}\) is measurable and, for \(\tau \in \left[ 0,T\right] ,\) a.e. \(\xi \in \left[ 0,1\right] ,\) the map \(b\left( \tau ,\xi ,\cdot \right) :\mathbb {{\mathbb {R}}} \rightarrow {\mathbb {R}}\) is continuous. Moreover, there exists \(c_{1} \in L^{\infty }([0,1])\), \(\alpha \in (2/3,1)\), such that, for \(\tau \in \left[ 0,T\right] \) and a.e. \(\xi \in \left[ 0,1\right] ,\)

$$\begin{aligned} \left| b\left( \tau ,\xi ,x\right) -b\left( \tau ,\xi ,y\right) \right| \le c_{1}\left( \xi \right) \left| x-y\right| ^\alpha , \end{aligned}$$

\( x,\,y \in {\mathbb {R}}\). Further, we require that \(\left| b\left( \tau ,\xi ,x\right) \right| \le c_{2}\left( \xi \right) ,\) for \(\tau \in [0,T]\), \(x \in {\mathbb {R}}\) and a.e. \(\xi \in [0,1]\), with \(c_{2} \in L^2 (\left[ 0,1\right] ) \). This example is discussed in Section 3.1 of [24]. In [24] it is also shown that 2-dimensional semilinear stochastic plate equations are of the form (4.1).

Equation (4.1) can be reformulated as an abstract evolution equation in H:

$$\begin{aligned} dX_{\tau }^{0,x} = A X_{\tau }^{0,x} d\tau +GB(\tau ,X_{\tau }^{0,x})d\tau +GdW_\tau ,\; \tau \in \left[ 0,T\right] , \; \displaystyle X_{0}^{0,x} =x \in H, \end{aligned}$$
(4.3)

where A is the generator of the wave group in H and \(GdW_\tau =\) \(\Big ( \begin{array}{c} 0\\ dW_\tau \end{array} \Big ) \). Concerning the semilinear stochastic Eq. (4.3), we assume that

Hypothesis 2

\(B: [0,T] \times H \rightarrow U\) is (Borel) measurable and bounded and moreover there exists \(C>0\) such that

$$\begin{aligned} |B(t, x+ h) - B(t,x) |_{U} \le C |h|^{\alpha }_H,\;\; x,h \in H, \; t \in [0,T], \end{aligned}$$
(4.4)

for some \(\alpha \in (2/3,1)\). We also write that \(B \in B_b([0,T]; C^{\alpha }_b(H,U))\).

Remark 4.1

We point out that the critical exponent \(\alpha =2/3\) agrees with the one which appears in the study of pathwise uniqueness for degenerate finite dimentional SDEs like (4.3) (with \(H = {\mathbb {R}}^{2d}\)); see [6] more details.

Let \(x \in H\). Recall that a (weak) mild solution to (4.3) is a tuple

$$\begin{aligned} \left( \Omega , {{\mathscr {F}}}, ({ {\mathscr {F}}}_{t}), {\mathbb {P}}, W, X\right) \end{aligned}$$

where \(\left( \Omega , {{\mathscr {F}}}, ({{\mathscr {F}}}_{t}), {\mathbb {P}}\right) \) is a stochastic basis on which it is defined a cylindrical U-valued \({{\mathscr {F}}}_{t}\)-Wiener process W and a continuous \({{\mathscr {F}}}_{t}\)-adapted H-valued process \(X = (X_t) = (X_t)_{t \in [ 0,T] }\) such that, \({\mathbb {P}}\)-a.s.,

$$\begin{aligned} X_{t}=e^{tA}x+\int _{0}^{t}e^{\left( t-s\right) A}GB\left( s, X_{s}\right) ds+\int _{0}^{t}e^{\left( t-s\right) A}GdW_{s},\;\;\; t \in [0,T]. \end{aligned}$$
(4.5)

According to Chapter 1 in [26] we say that strong existence holds for equation (4.3) if, for every stochastic basis \((\Omega ,\mathcal{F}, (\mathcal{F}_t), {\mathbb {P}})\) on which there is defined an U-valued cylindrical \(\mathcal{F}_t\)-Wiener process W, for any initial condition \(x \in H,\) there exists an H-valued continuous \((\mathcal{F}_t)\)-adapted process \(X= (X_t)= (X_t)_{t \in [0,T]}\) such that \((\Omega ,\mathcal{F}, (\mathcal{F}_t), {\mathbb {P}},W, X) \) is a weak mild solution. We also write \(X_t^{0,x}\) or \(X_t^x\) instead of \(X_t\). Similarly, we denote by \((X_{\tau }^{t,x})_{\tau \ge t}\) the solution to (4.3) starting from \(x \in H\) at time \(t \in [0,T]\).

We say that for Eq. (4.3) strong or pathwise uniqueness holds (starting from any initial \(x \in H\)) if given two weak mild solutions \(X= (X_{t})\) and \(Y = (Y_{t})\) (defined on the same stochastic basis, solutions with respect to the same cylindrical Wiener process) starting at the same x, we have, \({\mathbb {P}}\)-a.s., \(X_t = Y_t\), \(t \in [0,T]\).

Note that if \(a \in U\), \( Ga = \left( \begin{array}{c} 0\\ a \end{array} \right) \in K \;\; \text {and } \;\; e^{tA} \left( \begin{array}{c} 0\\ a \end{array} \right) =\Big ( \begin{array}{c} \frac{1}{\sqrt{{\Lambda }}}\sin (\sqrt{{\Lambda }}\, t) \, a \\ \cos (\sqrt{{\Lambda }}t)a \end{array} \Big ) \in K, \; t\in {\mathbb {R}}. \) Hence, since in Hypotheses 2 we assume that the drift B takes its values in U, then \(\displaystyle \int _{0}^{t}e^{\left( t-s\right) A}GB\left( s, X_{s}\right) ds\) evolves in K: it is K-valued and the map \(t\mapsto \displaystyle \int _{0}^{t}e^{\left( t-s\right) A}GB\left( s, X_{s}\right) ds\) is continuous due to the boundedness of B (indeed, let \(T>0\); for any \(\omega \), \({\mathbb {P}}\)-a.s., \(s \mapsto GB\left( s, X_{s}(\omega )\right) \) is Borel and bounded from [0, T] with values in K and so we can apply Lemma 3.1.5 in [7]).

Therefore even if in general a solution \((X_t^x)\) does not evolve in K (cf. (2.8)) we know that for any initial condition \(x_1,\,x_2\in H\) such that \( x_1 - x_2 \in K \) then any couple of weak mild solutions \(X_t^{x_1}\) and \( X_t^{x_2}\) (starting at \(x_1\) and \(x_2\), respectively, and driven by the same noise) verifies the property

$$\begin{aligned} ( X_t^{x_1} - X_t^{x_2})_{ \in [0,T]} \;\; \text {evolves in} \; \; K,\;\;\; \text {if }x_1-x_2\in K. \end{aligned}$$
(4.6)

Indeed

$$\begin{aligned} X_t^{x_1} - X_t^{x_2}=e^{ t A}(x_1-x_2)+\displaystyle \int _{0}^{t}e^{\left( t-s\right) A}G\Big (B\left( s,X^{x_1}_{s}\right) -B\left( s,X^{x_2}_{s}\right) \Big )\,ds,\;\;\; t \in [0,T]; \end{aligned}$$

the stochastic integral has disappeared, and since \(x_1 - x_2 \in K\), also \(e^{ t A}(x_1-x_2)\in K\); the other term we have already discussed that belongs to K. Note that, \({\mathbb {P}}\)-a.s., the paths of \( (X_t^{x_1} - X_t^{x_2})\) are continuous functions from [0, T] with values in K. Property (4.6) will be important in the proof of our uniqueness result (see Theorem 4.7). Indeed recall that the Ornstein–Uhlenbeck semigroup regularizes only in the directions of K (see Sect. 3).

Note that thanks to the boundedness of B we can apply the Girsanov Theorem, see [24, Remark 2.1] changing H in K, or see [25, Remark 2.1] for more details.

4.1 The Related Infinite Dimensional PDE

In this section we will apply the regularity results proved in Sect. 3 with \(J = K = V \times U\). Let \(T>0.\) We consider the following integral equation of Kolmogorov type which will be important in the sequel:

$$\begin{aligned} u(t,x)&=\int _t^T R_{s-t}\left[ e^{-(s-t) {A}}G B(s,\cdot )\right] (x)\,ds \nonumber \\&+ \int _t^T R_{s-t}\left[ e^{-(s-t){A}} \nabla ^Gu(s,\cdot ) B(s,\cdot ) \right] (x)\,ds, \end{aligned}$$
(4.7)

where u(tx) takes values in K and \( \nabla ^G u(s,x)B(s,x) = \nabla _{GB(s,x)}u(s,x) \in K \), \((s,x) \in [0,T] \times H\).

Using results of Sect. 3 we will solve the equation in the Banach space \(E_0 \) consisting of all \(u \in B_b([0,T]\times H, K)\) such that \(u(t, \cdot ) \) is K-differentiable on H, with

\(\nabla ^{K} u \in B_b([0,T]\times H, L(K,K))\). We also require that there exists \(C =C_T >0\) such that for any \(x, y \in H\)

$$\begin{aligned} \sup _{s \in [0,T]}|\nabla ^G u(s, x) - \nabla ^G u(s, y)|_{L(U,K)} \, \le \, C |x-y|_H^{\alpha }, \end{aligned}$$
(4.8)

where \(\alpha \in (2/3, 1)\) is given in Hypothesis 2. Finally, to define \(E_0\) we require that, for each \(\xi \in U\), \(t \in [0,T]\), the mapping:

$$\begin{aligned} x \mapsto \nabla _{\xi }^G u (t,x) \;\; \text { is } K \text { -differentiable on } H \end{aligned}$$
(4.9)
$$\begin{aligned} \text { with } \;\; \sup _{(t, x) \in [0,T] \times H } \; \sup _{|\xi |_U =1} \Vert \nabla ^K \nabla ^G_{\xi } u(t,x) \Vert _{L\left( K, K \right) } < \infty . \end{aligned}$$

Let \(\beta \ge 0\) to be fixed later. It is not difficult to prove that \(E_0\) is a Banach space endowed with the norm

$$\begin{aligned} \Vert u\Vert _{E_0, \beta } = \sup _{(t, x) \in [0,T] \times H} e^{\beta t} | u (t,x) |_K + \sup _{(t, x) \in [0,T] \times H} e^{\beta t} \Vert \nabla ^K u (t,x) \Vert _{L(K,K)} \end{aligned}$$
$$\begin{aligned}{} & {} + \sup _{t \in [0,T]} e^{\beta t} \Vert \nabla ^G u (t, \cdot ) \Vert _{C^{\alpha }_b(H, L(U,K))} \nonumber \\{} & {} \quad + \sup _{(t, x ) \in [0,T] \times H} \, \sup _{|\xi |_U =1} e^{\beta t} \Vert \nabla ^K \nabla ^G_{\xi } u(t,x) \Vert _{L\left( K, K \right) }. \end{aligned}$$
(4.10)

Lemma 4.2

Let Hypotheses 1 and  2 hold true. There exists a unique solution \(u \in E_0\) to (4.7). Moreover, there exists a function \(h(r) = h(r,\alpha ) > 0\), \(r\ge 0\), such that \(h(r) \rightarrow 0\) as \(r \rightarrow 0^+\) and if \(S \in [0,T]\) verifies \(h(T- S) \cdot ( \sup _{t \in [0,T]}\Vert B(t,\cdot ) \Vert _{\alpha }) \) \( \le 1/4\), then

$$\begin{aligned} \sup _{t\in [S,T],\,x\in H} \Vert \nabla ^K u (t,x) \Vert _{L(K,K)} \le 1/3. \end{aligned}$$
(4.11)

Proof

We introduce the following operator \(\mathcal{T}\) defined on \(E_0\):

$$\begin{aligned} \mathcal{T} u (t,x)= & {} \int _t^T R_{s-t}\left[ e^{-(s-t){A}}G B(s,\cdot )\right] (x)\,ds\\{} & {} + \int _t^T R_{s-t}\left[ e^{-(s-t){A}} \nabla ^G u(s,\cdot ) B(s,\cdot )\right] (x)\,ds, \end{aligned}$$

\(u \in E_0,\) \((t,x) \in [0,T] \times H\). In order to apply Lemmas 3.8 and 3.11 with \(J=K\) we first check the Hölder regularity of the terms \(G B(s, \cdot )\) and \( \nabla ^G u(s,\cdot ) B(s,\cdot ).\)

Since, for any \(x \in H\), \(h \in H,\) \(s \in [0,T]\),

$$\begin{aligned} |G B(s, x) - G B(s, x+h ) |_K = | B(s, x) - B(s, x+ h) |_U \le C |h|_H^{\alpha } \end{aligned}$$

we get that \(G B \in B_b([0,T]; C^{\alpha }_b(H,K))\), see the definition in Hypothesis (2).

We need to prove that also

$$\begin{aligned} \nabla ^G u(s,\cdot )B(s,\cdot ) \;\; \text { belongs to } B_b([0,T]; C^{\alpha }_b(H,K)). \end{aligned}$$
(4.12)

We write for \(x, y \in H\) with \(|x-y|_H \le 1\)

$$\begin{aligned}{}[\nabla ^G u(s, x)B(s,x) - \nabla ^G u(s, y)B(s,x)] + \nabla ^G u(s, y)[B(s,x) - B(s,y)]. \end{aligned}$$

We bound the second term with

$$\begin{aligned}&| \nabla ^G u(s, y)[B(s,x) - B(s,y)] |_K \\&\quad \; \le \sup _{(t, x) \in [0,T] \times H} e^{\beta t} \Vert \nabla ^G u (t,x) \Vert _{L(U,K)} \, | B(s,x) - B(s,y)|_U \le C |x-y|_H^{\alpha }. \end{aligned}$$

Moreover we have

$$\begin{aligned}&|[\nabla ^G u(s, x) - \nabla ^G u(s, y)] B(s,x)|_K \le |B(s,x)|_U |\nabla ^G u(s, x) - \nabla ^G u(s, y)|_{L(K,U)} \\&\quad \le \Vert B \Vert _{\infty } \,\sup _{t \in [0,T]} e^{\beta t} \Vert \nabla ^G u (t, \cdot ) \Vert _{C^{\alpha }_b(H, L(U,K))} \, |x-y|_H^{\alpha }. \end{aligned}$$

By Lemmas 3.8 and 3.11, using that \(\alpha >2/3\), it is not difficult to prove that \(\mathcal{T}: E_0 \rightarrow E_0\). Let us check that for a suitable value of \(\beta \) the map \(\mathcal{T}\) is a strict contraction (see (4.10)). We have to consider \(\Vert \mathcal{T}u_1 - \mathcal{T}u_2\Vert _{E_0, \beta }\), \(u_1, u_2 \in E_0\); we only treat the term

$$\begin{aligned} \sup _{t,x} \sup _{|\xi |_U =1} e^{\beta t} \Vert \nabla ^K \nabla _{\xi }^G [\mathcal{T} u_1 - \mathcal{T} u_2] (t,x) \Vert _{L(K,K)}. \end{aligned}$$

Indeed the other terms of \(\Vert \mathcal{T}u_1 - \mathcal{T}u_2\Vert _{E_0, \beta }\) can be estimated in a similar way. We have for any \(k \in K\) with \(|k|_K =1\)

$$\begin{aligned}&e^{\beta t} | \nabla _k \nabla _{\xi }^G [\mathcal{T} u_1(t,x) - \mathcal{T} u_2(t,x)] \, |_{K} \\&\quad \le \int _t^T e^{-\beta (s-t)} \Big | \nabla _k \nabla ^G_{\xi } R_{s-t}\left[ e^{-(s-t){A}}\, e^{\beta s} \{ \nabla ^G u_1(s,\cdot ) - \nabla ^G u_2(s,\cdot ) \} B(s,\cdot ) \right] (x) \,\Big |_{K}ds \\&\quad \le \int _t^T \frac{c e^{- \beta (s-t)}}{(s-t)^{\frac{4 - 3\alpha }{2}}} ds \sup _{t \in [0,T]}\Vert B(t,\cdot ) \Vert _{\alpha } \Vert u_1 - u_2 \Vert _{E_0, \beta }\\&\quad \le C_{\beta ,T} \sup _{t \in [0,T]}\Vert B(t,\cdot ) \Vert _{\alpha } \Vert u_1 - u_2 \Vert _{E_0, \beta }, \end{aligned}$$

where \(C_{\beta ,T} >0\) tends to 0 as \(\beta \rightarrow + \infty .\) Therefore taking the supremum over \(k \in K\) with \(|k|_K =1\) we infer

$$\begin{aligned} e^{\beta t} \Vert \nabla ^K \nabla _{\xi }^G [\mathcal{T} u_1(t,x) - \mathcal{T} u_2(t,x)] \Vert _{L(K,K)} \le C_{\beta ,T} \sup _{t \in [0,T]}\Vert B(t,\cdot ) \Vert _{\alpha } \Vert u_1 - u_2 \Vert _{E_0, \beta }. \end{aligned}$$

Choosing \(\beta \) large enough, we can apply the fixed point theorem and obtain that there exists a unique solution \(u \in E_0\).

In order to prove (4.11), we first introduce \(\Vert u\Vert _{E_0, 0, S,T}\) which is defined as \(\Vert u\Vert _{E_0, 0}\) in (4.10) (with \(\beta =0)\) but taking all the supremums over \([S, T] \times H\) instead of \([0,T] \times H\). We proceed as before:

$$\begin{aligned}&\Vert u\Vert _{E_0,0, S, T} \le \sup _{t \in [S,T]}\int _t^T \frac{c }{(s-t)^{\frac{4 - 3\alpha }{2}}} ds \, \sup _{t \in [0,T]}\Vert B(t,\cdot ) \Vert _{\alpha , K } \, (\Vert u_0 \Vert _{E_0, 0, S,T} + 1) \\&\quad \le h(T-S) \, \sup _{t \in [0,T]}\Vert B(t,\cdot ) \Vert _{\alpha } \, (\Vert u_0 \Vert _{E_0, 0, S,T} + 1), \end{aligned}$$

where \( h( r) = \int _0^r \frac{c }{s^{\frac{4 - 3\alpha }{2}}} ds \); now (4.11) follows since we have \(\frac{3}{4}\Vert u\Vert _{E_0,0, S,T} \le 1/4.\) \(\square \)

To prove the next result we will apply Lemma 3.9 (see also Remark 3.10). To this purpose we fix a basis \((f_m)\) of K and set \(u_m = \langle u, f_m \rangle _K \), where u is the solution given in Lemma 4.2.

Theorem 4.3

Let Hypotheses 1 and  2 hold true. Then the unique solution \(u \in E_0\) to (4.7) (see Lemma 4.2) verifies in addition

$$\begin{aligned} \Big ( \sum _{m \ge 1} \sup _{|a|_U =1}| \nabla _k \nabla _{a}^G u_m (t,x)|^2 \Big )^{1/2}\le C |k|_K \, \sup _{t \in [0,T]}\Vert B(t, \cdot )\Vert _{ C^{\alpha }_b(H, U)}\;\; \;\;\; k \in K;\nonumber \\ \end{aligned}$$
(4.13)

here \(C >0\) is independent of \(x \in H, k \in K\), \(t \in [0,T]\), u and the basis \((f_m)\) in K.

Proof

First following the proof of Lemma 4.2 it is not difficult to prove that there exists \(C_T>0\) independent of u such that

$$\begin{aligned} \sup _{t \in [0,T]} \Vert \nabla ^G u (t, \cdot ) \Vert _{C^{\alpha }_b(H, L(U,K))} \le C_T \sup _{t \in [0,T]}\Vert B(t, \cdot )\Vert _{ C^{\alpha }_b(H, U)}. \end{aligned}$$
(4.14)

Then we write \(u = v + w\), where

$$\begin{aligned} \,v(t,x)&= \int _t^T R_{s-t}\left[ e^{-(s-t){A}}G B(s,\cdot )\right] (x)\,ds,\;\; \\ w(t,x)&= \int _t^T R_{s-t}\left[ e^{-(s-t){A}} \nabla ^G u(s,\cdot ) B(s,\cdot )\right] (x)\,ds. \end{aligned}$$

We need to prove that (4.13) holds when u is replaced by v and w. We concentrate on w (the proof for v is similar). We define, for any \(s \in [0,T]\), \(x\in H,\) \(\Phi (s, x) = e^{-(s-t){A}} \nabla ^G u(s,x) B(s,x)\).

Note that \(\Phi \in B_b([0,T]; C^{\alpha }_b(H,K)) \) (see the computations to verify (4.12)). By (4.14) we obtain

$$\begin{aligned}{} & {} \sup _{s \in [0,T]} \Vert \Phi (s, \cdot )\Vert _{ C^{\alpha }_b(H, K)} \le C_T \sup _{t \in [0,T]}\Vert B(t, \cdot )\Vert _{ C^{\alpha }_b(H, U)} \;\; \text {with} \\{} & {} \quad w(t,x)= \int _t^T R_{s-t}[ \Phi (s,\cdot )](x)\,ds,\;\;\; t \in [0,T],\; x \in H. \end{aligned}$$

Let us fix \(k \in K\) and \(t \in [0,T)\); set \(w_m = \langle w, f_m \rangle _K \). By Lemma 3.9 (see also Remark 3.10) we know that, for any \(s \in (t,T]\), the linear operator \(\nabla _k \nabla _{}^G R_{s-t}[ \Phi (s, \cdot )](x)\) is well defined from K into U by the formula \( l \mapsto \nabla _k \nabla _{}^G P_{s-t}[ \langle \Phi (s, \cdot ), l \rangle _K ](x) \) and it is a Hilbert-Schmidt operator. Moreover, for \(s>t,\)

$$\begin{aligned}{} & {} \Vert \nabla _k \nabla _{}^G R_{s-t}[ \Phi (s, \cdot )](x) \Vert _{L_2(K,U)} = \Big (\sum _{m \ge 1} \sup _{|a|_U =1}| \nabla _k \nabla _{a}^G P_{s-t}[\langle \Phi (s, \cdot ), f_m \rangle _K ](x)|^2 \Big )^{1/2} \nonumber \\{} & {} \quad \le \frac{C}{(s -t)^{\frac{4 - 3\alpha }{2}}} |k|_K \, \sup _{r \in [0,T]}\Vert \Phi (r, \cdot )\Vert _{ C^{\alpha }_b(H, K)} \le \frac{c |k|_K }{(s -t)^{\frac{4 - 3\alpha }{2}}} \sup _{t \in [0,T]}\Vert B(t, \cdot )\Vert _{ C^{\alpha }_b(H, U)}.\nonumber \\ \end{aligned}$$
(4.15)

Note that also the linear operator \(\nabla _k \nabla _{}^G w(t, x)\) is well defined from K into U (\(l \mapsto \nabla _k \nabla _{}^G [\langle w(s, \cdot ), l \rangle _K ](x) \)). Further, we have, for any \(x \in H,\)

$$\begin{aligned} \nabla _k \nabla _{}^G w(t, x) = \int _t^T \nabla _k \nabla _{}^G R_{s-t}[ \Phi (s, \cdot )](x) ds. \end{aligned}$$

Using (4.15) we deduce that \(\nabla _k \nabla _{}^G w(t, x) \in L_2(K,U)\) and

$$\begin{aligned}&\Vert \nabla _k \nabla _{}^G w(t, x)\Vert _{L_2(K,U)} \\&\quad \le \int _t^T \Vert \nabla _k \nabla _{}^G R_{s-t}[ \Phi (s, \cdot )](x) \Vert _{L_2(K,U)} ds \le C_{\alpha , T} |k|_K \, \sup _{t \in [0,T]}\Vert B(t, \cdot )\Vert _{ C^{\alpha }_b(H, U)}, \end{aligned}$$

where \( C_{\alpha , T} >0 \) is independent of \(x \in H\), \(t \in [0,T]\), B and \(k \in K\). \(\square \)

Remark 4.4

If we try to prove the previous theorem using Lemma 3.2 instead of Lemma 3.4 we could eventually obtain a bound for \(\sum _{m \ge 1} \sup _{|a|_U =1}| \nabla _k \nabla _{a}^G u_m (t,x)|^2 \) by requiring that

$$\begin{aligned} \sum _{j \ge 1} \sup _{t \in [0,T]} \Vert \langle B(t, \cdot ), e_j \rangle _U \Vert _{C^{\alpha }_K(H)}^2 < \infty . \end{aligned}$$

However this condition is restrictive when it is applied to an example like (4.2) (see also Section 3 in [24]).

4.2 The Related Infinite Dimensional Forward-Backward System

We first collect some results on a related forward-backward system (FBSDE), whose solution will be useful “in removing the bad term B” in (4.3) (cf. Proposition 4.6). Comparing with [8] this procedure turns out to be an alternative to the use of the Itô’s formula (for recent advances on infinite dimensional Itô formula we refer to [16]). Here, following [24], we use it in the present case when A is the generator of a group of operators. In [1] this approach has been extended to the case when A is the generator of a semigroup of operators.

In a complete probability space \((\Omega , \mathscr {F}, {\mathbb {P}})\), let us consider the following FBSDE

$$\begin{aligned} \left\{ \begin{array}{l} d\Xi _\tau ^{t,x} =A\Xi _\tau ^{t,x} d\tau +Gd W_\tau ,\tau \in \left[ t,T\right] , \\ \displaystyle \Xi _{t}^{t,x} =x,\\ \displaystyle -d Y_{\tau }^{t,x}=-{A}Y_\tau ^{t,x}+GB(\tau ,\Xi ^{t,x}_\tau )\;d\tau + Z^{t,x}_{\tau } \, B(\tau ,\Xi ^{t,x}_\tau ) d\tau - Z^{t,x}_{\tau }\;dW_\tau , \,\tau \in [0,T], \\ \displaystyle Y_{T}^{t,x}=0, \end{array} \right. \end{aligned}$$
(4.16)

\(t \in [0,T]\), \(x \in H\), and the forward equation is the abstract formulation of the wave equation (1.1) given in (2.6) under Hypothesis 1; here \(B: [0,T] \times H \rightarrow U\) satisfies Hypothesis 2, G is defined by (2.4) and W is a cylindrical Wiener process in U.

We extend \(\Xi ^{t,x}\) to the whole [0, T] by setting \(\Xi _\tau ^{t,x}=x\) for \(0\le \tau \le t\), so \((Y^{t,x},Z^{t,x})\) are well defined on [0, T]. The precise meaning of the BSDE in (4.16) is given by its mild formulation: for \( \tau \in [0,T]\)

$$\begin{aligned} Y_{\tau }^{t,x}&=\int _\tau ^Te^{-(s-\tau ){A}}G B(s,\Xi ^{t,x}_s)\,ds \nonumber \\&+\int _\tau ^Te^{-(s-\tau ){A}} Z_s^{t,x}\, B(s,\Xi ^{t,x}_s)\,ds- \int _\tau ^Te^{-(s-\tau ){A}} Z^{t,x}_{s}\;d W_s, \end{aligned}$$
(4.17)

\({\mathbb {P}}\)-a.s. (cf. [20, 21] and the references therein).

Notice that the forward equation evolves in H as we have already discussed in Sect. 2, while the backward equation takes values in K. Indeed the term \(GB(\cdot , \cdot )\) takes values in K, and we can use an explicit representation of the solution in terms of \(GB(\cdot )\), see [21], Sects. 2 and 3.

We denote by \(L^2_\mathscr {P}(\Omega , C([0,T],{K}))\) the space of all predictable (with respect to the completed natural filtration generated by W) K-valued processes Y with continuous paths and such that

$$\begin{aligned} {\mathbb {E}}[ \sup _{\tau \in [0,T]}\vert Y^{}_\tau \vert ^2] = \Vert Y\Vert _{L^2_\mathscr {P}(\Omega , C([0,T],{K}))}^2 < \infty . \end{aligned}$$

We refer to [18], Section 2.2, for the definition of the spaces of Gâteaux differentiable functions \(\mathscr {G}^{1}( {H},{U})\) and \(\mathscr {G}^{0,1}([0,T]\times {H},{K}) \subset C([0,T]\times {H,K})\) (such spaces are also considered in [24] and [25]).

Following [21] and Remark 4.5 and estimate (4.19) in [19] (see also Section 5 of [24] where these results are summed up) it is immediate to get existence and pathwise uniqueness of a solution \((Y^{t,x},Z^{t,x})\in L^2_\mathscr {P}(\Omega , C([0,T],{K})) \times L^2_\mathscr {P}(\Omega \times [0,T],L_2(U,{K}))\). such that

$$\begin{aligned} {\mathbb {E}}\big [ \sup _{\tau \in [0,T]}\vert Y^{t,x}_\tau \vert _{{K}}^2 \big ]+{\mathbb {E}}\int _0^T\Vert Z_\tau ^{t,x}\Vert ^2_{L_2(U,{K})} d \tau \le C_T \Vert B\Vert _{\infty }. \end{aligned}$$
(4.18)

Moreover if we further assume that the map \(x\mapsto B(\tau ,x),\; H\rightarrow U,\) belongs to \(\mathscr {G}^{1}( {H},{U})\) then, the map: \((t,x)\mapsto Y_t^{t,x}\) belongs to \(\mathscr {G}^{0,1}([0,T]\times H,{K})\). In addition, \((t,x)\mapsto Y_t^{t,x} \), \([0,T]\times H \rightarrow {K}\), is deterministic.

So we can define the function

$$\begin{aligned} v(t,x)= Y_t^{t,x}, \quad (t,x)\in [0,T]\times H. \end{aligned}$$
(4.19)

and by the Markov property we have:

$$\begin{aligned} v(\tau ,\Xi _\tau ^{t,x})=Y_\tau ^{t,x},\, \;\;\; \tau \in [t,T], \; {\mathbb {P}}-a.s.. \end{aligned}$$
(4.20)

Theorem 4.5

Assume Hypothesis 1 and let B satisfy Hypothesis 2. Let v be the function defined in (4.19). Then \(v \in B_b([0,T] \times H, H)\) and, for any \(t \in [0,T],\) \(v (t, \cdot ): H \rightarrow {K}\) is G-differentiable on H (see the definition after (3.4)), for any \((t,x) \in [0,T] \times H\), the map: \(\xi \mapsto \) \(\nabla _{G \xi }v(t,x) = \nabla _{ \xi }^Gv(t,x)\) \( \in L(U, {K})\) and, for any \(\xi \in U\), \(\nabla ^G_{\xi } v \in B_b([0,T] \times H, {K})\) with \(\sup _{(t,x) \in [0,T] \times H} \Vert \nabla ^Gv(t,x) \Vert _{L(U, {K})} < \infty \). For any \( \tau \in [0,T] \), a.e., we have

$$\begin{aligned} \nabla ^G v(\tau ,\Xi _\tau ^{t,x})= Z_\tau ^{t,x}, \;\; \,{\mathbb {P}}\text {-a.s.}. \end{aligned}$$
(4.21)

Finally, by the representation formula given in (4.17) and using the OU semigroup \((R_t)\), the function v defined in (4.19) satisfies the integral equation

$$\begin{aligned} v(t,x)&=\int _t^T R_{s-t}\left[ e^{-(s-t){A}}G B(s,\cdot )\right] (x)\,ds \nonumber \\&+ \int _t^T R_{s-t}\left[ e^{-(s-t){A}} \nabla ^Gv(s,\cdot ) B(s,\cdot ) \right] (x)\,ds. \end{aligned}$$
(4.22)

and coincides with the function u, unique solution of (4.7) given in Lemma  4.2 and Theorem 4.3.

Proof

For the proof we refer to [24, Lemma 5.3 and Theorem 5.4] for the identification of Z, and to [24, Lemma 5.5] for the identification of v with u. \(\square \)

4.3 Applications to Strong Uniqueness

We study Eq. (4.3). We show how to remove the “bad” term B and how to apply our regularity results to get pathwise uniqueness. Let \(x \in H\). Recall that we consider a (weak) mild solution \((X_\tau ^{t,x})\) \(=(X_\tau ^{t,x})_{\tau \in [0,T]}\):

$$\begin{aligned} X_\tau ^{t,x} =e^{(\tau -t)A}x+\int _t^\tau e^{(\tau -s)A}GB(s,X_s)ds+\int _t^\tau e^{(\tau -s)A}GdW_s,\tau \in [ t,T], \nonumber \\ \end{aligned}$$
(4.23)

\( X_\tau ^{t,x} =x, \; \;\tau \le t\); cf. (4.5). This is a continuous H-valued process defined and adapted on a stochastic basis \(\left( \Omega , {{\mathscr {F}}}, ({ {\mathscr {F}}}_{t}), {\mathbb {P}}\right) \), on which it is defined a cylindrical U-valued \({{\mathscr {F}}}_{t}\)-Wiener process W. The next result shows how to remove the “bad" term B using the function v defined in (4.19), and without applying the Itô formula, but using the BSDE (4.17) in its mild formulation (we refer to [24, Section 6]).

Proposition 4.6

Let Hypotheses 1 and 2 hold true. Then a (weak) mild solution \(X^x = (X_{\tau }^x)\) of (4.23) starting at \(t=0\) satisfies, for any \(\tau \in [0,T]\), \({\mathbb {P}}\)-a.s.,

$$\begin{aligned} X_\tau ^{x}= & {} {e^{\tau A}x+e^{\tau A}v(0,x)-v(\tau ,X_\tau ^{x})+\int _0^\tau e^{(\tau -s)A}{\widetilde{Z}}_s^x\;dW_s +\int _0^\tau e^{(\tau -s)A}GdW_s}\nonumber \\= & {} e^{\tau A}x+e^{\tau A}v(0,x)-v(\tau ,X_\tau ^{x})+\int _0^\tau e^{(\tau -s)A}\nabla ^G v(s,X_s^x)\;dW_s\nonumber \\{} & {} \quad +\int _0^\tau e^{(\tau -s)A}GdW_s. \end{aligned}$$
(4.24)

Proof

The proof is given in [24, Proposition 6.1] and it is based on BSDEs, see also [24, Remark 6.2] where the novelty of the approach is discussed and [1, Proposition 4.1]. \(\square \)

We now prove strong (or pathwise) uniqueness of solutions.

Theorem 4.7

Let Hypotheses  1 and 2 hold true. Then for equation (4.3) pathwise uniqueness holds (starting from any initial condition \(x \in H\)).

Moreover, there exists \(c_T>0\) such that if \(X_\tau ^{x_1}\) and \( X_\tau ^{x_2}\) are two (weak) mild solutions starting from \(x_1\) and \(x_2 \in H\) at \(t=0\) (defined on the same stochastic basis) such that \(x_1-x_2\in K\) then \({\mathbb {P}}\)-a.s. \(X^{x_1}_{\tau }-X^{x_2}_{\tau } \in K \) for any \(\tau \in [0,T]\) and

$$\begin{aligned} { \sup _{\tau \in [0,T]} {\mathbb {E}}\vert X_\tau ^{x_1}-X_\tau ^{x_2}\vert ^2_K \le c_T \vert x_1-x_2\vert ^2_{K}. } \end{aligned}$$
(4.25)

Proof

We prove (4.25) which implies the pathwise uniqueness starting from any \(x \in H\). Indeed if \(x_1 = x_2 =x\) then (4.25) implies that \({\mathbb {P}}\)-a.s., \(X_\tau ^{x_1}=X_\tau ^{x_2}\), \(\tau \in [0,T]\).

Let us fix \(x_1, x_2 \in H\) with \(x_1-x_2 \in K\) and consider two (weak) mild solutions \(X^1\) and \(X^2\) defined on the same stochastic basis, with respect to the same cylindrical Wiener process and starting respectively from \(x_1\) and \(x_2\) at time \(t=0.\) Notice that

$$\begin{aligned} X^1_{\tau } - X^2_{\tau } = e^{\tau A}(x_1 - x_2) +\int _0^\tau e^{(\tau -s)A}\big ( GB(s,X^1_s)-GB(X^2_s)\big )\,ds, \end{aligned}$$

and both \(e^{\tau A}(x_1 - x_2) \) and the integral take their values in K. Indeed \(x_1 - x_2 \in K\), \(GB(s, \cdot ) \in K\) and \(e^{rA}: K \rightarrow K\) (cf. (4.6)).

Let \(T_0 \in (0,T]\) be such that \(h(T_0) \cdot ( \sup _{t \in [0,T]}\Vert B(t,\cdot ) \Vert _{\alpha }) \) \( \le 1/4\) (see (4.11)).

We consider the FBSDE (4.16) with \(T = T_0\) and we denote its solution again \(({\widetilde{Y}}^{x},{\widetilde{Z}}^{x})\). We find the function \(v^{(0)}: [0,T_0] \times H \rightarrow K\) according to (4.19) with \(T= T_0\). By (4.24) we know that

$$\begin{aligned} X^1_{\tau } - X^2_{\tau }&= e^{\tau A}(x_1 - x_2) + e^{\tau A}[v^{(0)}(0,x_1) - v^{(0)}(0, x_2)] \nonumber \\&- [v^{(0)}(\tau ,X_\tau ^{1}) - v^{(0)}(\tau ,X_\tau ^{2})] +{\int _0^\tau e^{(\tau -s)A}[{\widetilde{Z}}^{x_1}_s- {\widetilde{Z}}^{x_2}_s]\;dW_s,}\;\; \tau \in [0,T_0], \end{aligned}$$
(4.26)

where \(\int _0^\tau e^{(\tau -s)A}[{\widetilde{Z}}^{x_1}_s- {\widetilde{Z}}^{x_2}_s]\;dW_s = \int _0^\tau e^{(\tau -s)A}[ \nabla ^G v^{(0)}(s,X_s^1) - \nabla ^G v^{(0)}(s,X_s^2) ]\;dW_s. \)

By the regularity properties of \(v^{(0)}\), see Theorem  4.3, Lemma 4.2 and Theorem 4.5, we get

$$\begin{aligned}{} & {} \vert e^{\tau A}(x_1 - x_2)\vert _K + \vert e^{\tau A}[v^{(0)}(0,x_1) - v^{(0)}(0, x_2)]\vert _K + \vert v^{(0)}(\tau ,X_\tau ^{1}) - v^{(0)}(\tau ,X_\tau ^{2})\vert _K \\{} & {} \quad \le C\vert x_1-x_2\vert _K +\frac{1}{3}\vert X_\tau ^1-X_\tau ^2\vert _K,\;\;\; \tau \in [0,T_0]. \end{aligned}$$

Concerning the stochastic integral, we have

(see [12] page 57 or [14], Section 4.3)

$$\begin{aligned}{} & {} {\mathbb {E}}\Big | \int _0^\tau e^{(\tau -s)A}[ \nabla ^G v^{(0)}(s,X_s^1) - \nabla ^G v^{(0)}(s,X_s^2) ]\;dW_s \Big |^2_K \nonumber \\{} & {} \quad \le {\mathbb {E}}\int _0^{\tau } \Vert \nabla ^G v^{(0)}(s,X_s^1) - \nabla ^G v^{(0)}(s,X_s^2) \Vert _{L_2(U, K)}^2 ds. \end{aligned}$$
(4.27)

It the sequel we will prove that \({\mathbb {E}}\int _0^{\tau } \Vert \nabla ^G v^{(0)}(s,X_s^1) - \nabla ^G v^{(0)}(s,X_s^2) \Vert _{L_2(U, K)}^2 ds\) is finite and we will provide a bound for it.

Let us consider a basis \((e_k) \) in U; by the regularity properties of \(v^{(0)}\) we get

$$\begin{aligned}{} & {} {\mathbb {E}}\int _0^{\tau } \Vert \nabla ^G v^{(0)}(s,X_s^1) - \nabla ^G v^{(0)}(s,X_s^2) \Vert _{L_2(U,K)}^2 ds \\{} & {} \quad = \sum _{j \ge 1} {\mathbb {E}}\int _0^{\tau } | \nabla ^G_{e_j} v^{(0)}(s,X_s^1) - \nabla ^G_{e_j} v^{(0)}(s,X_s^2) |_{K}^2 ds. \end{aligned}$$

Now, using a basis \((f_m)\) in K, we write, for any \(s \in [0,\tau ]\), \({\mathbb {P}}\)-a.s.,

$$\begin{aligned}&\sum _{j \ge 1} | \nabla ^G_{e_j} v^{(0)}(s,X_s^1) - \nabla ^G_{e_j} v^{(0)}(s,X_s^2) |_{K}^2 \\&\quad \; = \sum _{m \ge 1} \sum _{j \ge 1} \langle \nabla ^G_{e_j} v^{(0)}(s,X_s^1) - \nabla ^G_{e_j} v^{(0)}(s,X_s^2), f_m \rangle _K ^2 \\&\quad \;= \sum _{m \ge 1} \sum _{j \ge 1} [ \nabla ^G_{e_j} v^{(0)}_m (s,X_s^1) - \nabla ^G_{e_j} v^{(0)}_m (s,X_s^2) ] ^2 \\&\quad = \sum _{m \ge 1} | \nabla ^G v^{(0)}_m (s,X_s^1) - \nabla ^G v^{(0)}_m (s,X_s^2) |^2_U, \end{aligned}$$

using \( v^{(0)}_m = \langle v^{(0)}, f_m \rangle _K \) and noting that \(\nabla ^G v^{(0)}_m (s,X_s^1) \in L(U, {\mathbb {R}})\) can be identified by the Riesz theorem with a unique element in U. We have obtained

$$\begin{aligned}{} & {} \sum _{j \ge 1} \int _0^{\tau } | \nabla ^G_{e_j} v^{(0)}(s,X_s^1) - \nabla ^G_{e_j} v^{(0)}(s,X_s^2) |_{K}^2 ds\\{} & {} \quad = \int _0^{\tau } \sum _{m \ge 1} | \nabla ^G v^{(0)}_m (s,X_s^1) - \nabla ^G v^{(0)}_m (s,X_s^2) |^2_U ds\\{} & {} \quad = \int _0^{\tau } \sum _{m \ge 1} \sup _{|a|_U =1} | \nabla ^G_a v^{(0)}_m (s,X_s^1) - \nabla ^G_a v^{(0)}_m (s,X_s^2) |^2 ds. \end{aligned}$$

Hence we have, \({\mathbb {P}}\)-a.s.,

$$\begin{aligned}&\sum _{j \ge 1} \int _0^{\tau } | \nabla ^G_{e_j} v^{(0)}(s,X_s^1) - \nabla ^G_{e_j} v^{(0)}(s,X_s^2) |_{K}^2 ds \\&\quad \, = \sum _{m \ge 1} \int _0^{\tau } \sup _{|a|_U =1} \Big | \int _0^1 \nabla ^K \nabla ^G_a v^{(0)}_m (s,X_s^1 + r (X_s^2 - X_s^1)) \, [X_s^2 - X_s^1]dr \Big |^2 ds \\&\quad \, = \int _0^{\tau } \sum _{m \ge 1} \sup _{|a|_U =1} \Big | \int _0^1 \nabla _{[X_s^2 - X_s^1]}\nabla ^G_a v^{(0)}_m (s,X_s^1 + r (X_s^2 - X_s^1)) \, dr \Big |^2 ds. \end{aligned}$$

Moreover, we find, \({\mathbb {P}}\)-a.s., \(r \in [0,1]\), \(s \in [0, \tau ]\),

$$\begin{aligned}&\sum _{m \ge 1} \sup _{|a|_U =1} \Big | \int _0^1 \nabla _{[X_s^2 - X_s^1]}\nabla ^G_a v^{(0)}_m (s,X_s^1 + r (X_s^2 - X_s^1)) \, dr \Big |^2\nonumber \\&\quad \le \int _0^1 \sum _{m \ge 1} \sup _{|a|_U =1} | \nabla _{[X_s^2 - X_s^1]}\nabla ^G_a v^{(0)}_m (s,X_s^1 + r (X_s^2 - X_s^1)) |^2 \, dr \nonumber \\&\quad \le C_T |X_s^2 - X_s^1|_K^2 \, \sup _{t \in [0,T]}\Vert B(t, \cdot )\Vert _{ C^{\alpha }_b(H, U)}^2. \end{aligned}$$
(4.28)

In the last inequality we have used Theorem 4.3 with \(u= v^{(0)}\) and \(k = X_s^2 - X_s^1\).

Coming back to (4.27) we obtain

$$\begin{aligned}&{\mathbb {E}}\Big | \int _0^\tau e^{(\tau -s)A}[ \nabla ^G v^{(0)}(s,X_s^1) - \nabla ^G v^{(0)}(s,X_s^2) ]\;dW_s \Big |^2_K \\&\quad \le \sum _{j \ge 1} {\mathbb {E}}\int _0^{\tau } | \nabla ^G_{e_j} v^{(0)}(s,X_s^1) - \nabla ^G_{e_j} v^{(0)}(s,X_s^2) |_{K}^2 ds \\&\quad \le {\mathbb {E}}\int _0^{\tau } ds \int _0^1 \sum _{m \ge 1} \sup _{|a|_U =1} | \nabla _{[X_s^2 - X_s^1]}\nabla ^G_a v^{(0)}_m (s,X_s^1 + r (X_s^2 - X_s^1)) |^2 \, dr \\&\quad \le C_T \sup _{t \in [0,T]}\Vert B(t, \cdot )\Vert _{ C^{\alpha }_b(H, U)}^2 \, {\mathbb {E}}\int _0^{\tau } |X_s^2 - X_s^1|_K^2 ds. \end{aligned}$$

Starting from (4.26) and using the previous estimates, we can apply the Gronwall lemma and obtain

$$\begin{aligned} \sup _{\tau \in [0,T_0]} {\mathbb {E}}\vert X_\tau ^{x_1}-X_\tau ^{x_2}\vert ^2_K \le c_T \vert x_1-x_2\vert ^2_K. \end{aligned}$$
(4.29)

If \(T_0 <T\) we consider the FBSDE (4.16) with terminal time \( (2T_0) \wedge T \). We find \(v^{(1)}: [0,(2T_0) \wedge T] \times H \rightarrow K\) according to (4.22) with T replaced by \((2T_0) \wedge T\). By (4.24) we obtain in particular

$$\begin{aligned} X_\tau ^{1}- X_\tau ^{2}&= e^{\tau A}(x_1 - x_2) + e^{\tau A}[v^{(1)}(0,x_1) - v^{(1)}(0, x_2)] \\&- [v^{(1)}(\tau ,X_\tau ^{1}) - v^{(1)}(\tau ,X_\tau ^{2})] +\int _0^\tau e^{(\tau -s)A}[ \nabla ^G v^{(1)}(s,X_s^1)\\&- \nabla ^G v^{(1)}(s,X_s^2) ]\;dW_s, \end{aligned}$$

\(\tau \in [T_0,(2T_0) \wedge T].\) Arguing as before, we get, for \(\tau \in [T_0,(2T_0) \wedge T]\),

$$\begin{aligned}&\vert e^{\tau A}x_1 - x_2)\vert _K +\vert e^{\tau A}[v^{(1)}(0,x_1) - v^{(1)}(0, x_2)]\vert _K + \vert v^{(1)}(\tau ,X_\tau ^{1}) - v^{(1)}(\tau ,X_\tau ^{2})\vert _K \\&\quad \le C_T\vert x_1-x_2\vert _K +\frac{1}{3}\vert X_\tau ^1-X_\tau ^2\vert _K \end{aligned}$$

and

$$\begin{aligned}&{\mathbb {E}}\Big | \int _0^\tau e^{(\tau -s)A}[ \nabla ^G v^{(1)}(s,X_s^1) - \nabla ^G v^{(1)}(s,X_s^2) ]\;dW_s \Big |^2_K \\&\quad \le C_T \sup _{t \in [0,T]}\Vert B(t, \cdot )\Vert _{ C^{\alpha }_b(H, U)}^2 \int _0^{\tau } {\mathbb {E}}|X_s^1 - X_s^2 |^2_K ds \\&\quad = C_T \sup _{t \in [0,T]}\Vert B(t, \cdot )\Vert _{ C^{\alpha }_b(H, U)}^2 \Big ( \int _{0}^{T_0 } {\mathbb {E}}|X_s^1 - X_s^2 |^2_K ds + \int _{T_0 }^{\tau } {\mathbb {E}}|X_s^1 - X_s^2 |^2_K ds \Big ) \\&\quad \le C_T \sup _{t \in [0,T]}\Vert B(t, \cdot )\Vert _{ C^{\alpha }_b(H, U)}^2 \Big ( c_T T_0 \, |x_1 - x_2|^2_K + \int _{T_0 }^{\tau } {\mathbb {E}}|X_s^1 - X_s^2 |^2_K ds \big ). \end{aligned}$$

We have obtained, for \(\tau \in [T_0,(2T_0) \wedge T]\),

$$\begin{aligned} {\mathbb {E}}|X^1_{\tau } - X^2_{\tau }|^2_K \le C_T' |x_1 - x_2|^2_K + C_T' \int _{T_0 }^{\tau } {\mathbb {E}}|X_s^1 - X_s^2 |^2_K ds. \end{aligned}$$

By the Gronwall lemma we find

\( \sup _{\tau \in [T_0, (2T_0) \wedge T ]} {\mathbb {E}}\vert X_\tau ^{x_1}-X_\tau ^{x_2}\vert ^2_K \le c_T \vert x_1-x_2\vert ^2_K.\)

Proceeding in this way, in finite steps, we get (4.25). \(\square \)

Remark 4.8

Using a localization argument as in [10], the boundeness of B in (1.7) could be dispensed. In particular, one can prove strong well-posedness of (1.1), for any \(x \in H\), under Hypothesis  1 and assuming that \(B: [0,T] \times H \rightarrow U\) is continuous on \([0,T] \times H \) and growths at most linearly, uniformly in \(t \in [0,T]\); moreover, one requires that for any ball \(S \subset H\) the function \(B (t, \cdot ): S \rightarrow U\) is \(\alpha \)-Hölder continuous, for some \(\alpha > 2/3\), uniformly in \(t \in [0,T]\) (cf. (4.4)).