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

An extrapolation result in the variational setting: improved regularity, compactness, and applications to quasilinear systems

Sebastian Bechtel Delft Institute of Applied Mathematics, Delft University of Technology, P.O. Box 5031, 2600 GA Delft, The Netherlands S.Bechtel@tudelft.nl  and  Mark Veraar M.C.Veraar@tudelft.nl
(Date: September 13, 2025)
Abstract.

In this paper we consider the variational setting for SPDE on a Gelfand triple (V,H,V)(V,H,V^{*}). Under the standard conditions on a linear coercive pair (A,B)(A,B), and a symmetry condition on AA we manage to extrapolate the classical L2\mathrm{L}^{2}-estimates in time to Lp\mathrm{L}^{p}-estimates for some p>2p>2 without any further conditions on (A,B)(A,B). As a consequence we obtain several other a priori regularity results of the paths of the solution.

Under the assumption that VV embeds compactly into HH, we derive a universal compactness result quantifying over all (A,B)(A,B). As an application of the compactness result we prove global existence of weak solutions to a system of second order quasi-linear equations.

Key words and phrases:
Stochastic evolution equations, variational methods, compactness, tightness, stochastic partial differential equations, quasi- and semi-linear, coercivity, extrapolation, Sneiberg’s lemma
2020 Mathematics Subject Classification:
Primary: 60H15. Secondary: 35A01, 35B65, 35K59, 35K90, 47H05, 47J35.
The first author is supported by the Alexander von Humboldt foundation by a Feodor Lynen grant. The second author is supported by the VICI subsidy VI.C.212.027 of the Netherlands Organisation for Scientific Research (NWO)

1. Introduction

In this paper we consider stochastic evolution equations in the variational setting. This is the stochastic version of Lions classical setting [23] and goes back to the work of [6, 21, 28]. For details we refer the reader to the monographs [24, 31].

This paper consists of two main parts. In the first part we prove an extrapolation result, which improves the usual a priori regularity estimates for linear equations with operators that satisfy the usual coercivity conditions. As a consequence we derive a new type of compactness result for linear equations, which is universal in the sense that for given constants in the estimates it varies over all admissible operators and data. In the second part we apply the compactness result to obtain global existence for a class of stochastic parabolic systems which is not well-understood yet.

The triple (V,H,V)(V,H,V^{*}) are Hilbert spaces such that VHVV\hookrightarrow H\hookrightarrow V^{*} densely, where we identify the Hilbert space HH with its dual, and where VV^{*} is the dual with respect to the inner product of HH. The stochastic equations we consider can be written as:

(LP) {du(t)+A(t)u(t)dt=f(t)dt+(B(t)u(t)+g(t))dW(t),u(0)=u0.\left\{\begin{aligned} \,\mathrm{d}u(t)+A(t)u(t)\,\,\mathrm{d}t&=f(t)\,\,\mathrm{d}t+(B(t)u(t)+g(t))\,\,\mathrm{d}W(t),\\ u(0)&=u_{0}.\end{aligned}\right.

Here A:(0,T)×Ω(V,V)A:(0,T)\times\Omega\to\mathcal{L}(V,V^{*}) and B:(0,T)×Ω(V,2(U,H))B:(0,T)\times\Omega\to\mathcal{L}(V,\mathcal{L}_{2}(U,H)), and WW is a UU-cylindrical Brownian motion, where UU is a real separable Hilbert space. The time T>0T>0 is finite. For the inhomogeneities (f,g,u0)(f,g,u_{0}) we assume f:(0,T)×ΩVf:(0,T)\times\Omega\to V^{*}, g:(0,T)×Ω2(U,H)g:(0,T)\times\Omega\to\mathcal{L}_{2}(U,H) and u0:ΩHu_{0}:\Omega\to H, and the standard measurability conditions are supposed to be satisfied.

Consider the following condition on the pair (A,B)(A,B).

Assumption 1.1.

There exist constants Λ,λ>0\Lambda,\lambda>0 and M0M\geq 0 such that pointwise in [0,T]×Ω[0,T]\times\Omega for all vVv\in V,

ReAv,v12Bv2(U,H)2\displaystyle\operatorname{Re}\,\langle Av,v\rangle-\frac{1}{2}\|Bv\|_{\mathcal{L}_{2}(U,H)}^{2} λvV2MvH2\displaystyle\geq\lambda\|v\|_{V}^{2}-M\|v\|_{H}^{2} (coercivity),
AvVΛvV\displaystyle\|Av\|_{V^{*}}\leq\Lambda\|v\|_{V}\quad &Bv2(U,H)ΛvV\displaystyle\&\quad\|Bv\|_{\mathcal{L}_{2}(U,H)}\leq\Lambda\|v\|_{V} (boundedness).\displaystyle\text{(boundedness)}.

Under Assumption 1.1 it is standard that (LP) has a unique strong solution uL2(0,T;V)C([0,T];H)u\in\mathrm{L}^{2}(0,T;V)\cap\mathrm{C}([0,T];H) a.s. and the following a priori estimate holds for a constant CC only depending on T1T\vee 1, λ\lambda, Λ\Lambda, and MM:

(1) uL2(Ω×(0,T);V)+uL2(Ω;C([0,T];H))C[u0L2(Ω;H)+fL2(Ω×(0,T);V)+gL2(Ω×(0,T);2(U,H))].\displaystyle\begin{aligned} \|u\|_{\mathrm{L}^{2}(\Omega\times(0,T);V)}+\|u\|_{\mathrm{L}^{2}(\Omega;\mathrm{C}([0,T];H))}\leq C\big{[}\|u_{0}\|_{\mathrm{L}^{2}(\Omega;H)}&+\|f\|_{\mathrm{L}^{2}(\Omega\times(0,T);V^{*})}\\ &+\|g\|_{\mathrm{L}^{2}(\Omega\times(0,T);\mathcal{L}_{2}(U,H))}\big{]}.\end{aligned}

In the deterministic setting it follows from [14] that one can find some p0>2p_{0}>2 depending on T1T\vee 1, λ\lambda, Λ\Lambda, and MM such that for every p[2,p0]p\in[2,p_{0}] there exists a constant CpC_{p} such that

(2) uLp(0,T;V)+uLp(0,T;V)Cp[u0(H,V)12/p\displaystyle\|u\|_{\mathrm{L}^{p}(0,T;V)}+\|u^{\prime}\|_{\mathrm{L}^{p}(0,T;V^{*})}\leq C_{p}\big{[}\|u_{0}\|_{(H,V)_{1-\nicefrac{{2}}{{p}}}} +fLp(0,T;V)].\displaystyle+\|f\|_{\mathrm{L}^{p}(0,T;V^{*})}\big{]}.

Actually this even holds true outside the setting of Hilbert space and variational problems. On the other hand, (2) does not hold for all p(2,)p\in(2,\infty). In [14] the estimate (2) is used to obtain a compactness result, which in turn is used to prove global existence of solutions to a quasi-linear parabolic PDE. On the other hand, if AA does not depend on time, then (2) holds for all p(1,)p\in(1,\infty) (see [19, Theorem 17.2.26 (4)]). A stochastic version of an extrapolation result holds if B=0B=0 (or small), and was proved in [25]. In the general case B0B\neq 0 the result is false since the Lp(Ω)\mathrm{L}^{p}(\Omega)-integrability can fail to hold (see [11]) and a more restrictive coercivity condition is needed (see [16]). However, it would be interesting to obtain some result with mixed integrability L2(Ω;Lp(0,T;V))\mathrm{L}^{2}(\Omega;\mathrm{L}^{p}(0,T;V)) in case of general (A,B)(A,B) satisfying Assumption 1.1 and any p[2,)p\in[2,\infty).

To prove (2) one can use Sneiberg’s lemma for complex interpolation. In the latter argument one needs an operator T(Xi,Yi)T\in\mathcal{L}(X_{i},Y_{i}), where (X0,X1)(X_{0},X_{1}) and (Y0,Y1)(Y_{0},Y_{1}) are Banach couples. By complex interpolation also T(Xθ,Yθ)T\in\mathcal{L}(X_{\theta},Y_{\theta}) for all θ(0,1)\theta\in(0,1). Sneiberg’s result states that the interval of θ(0,1)\theta\in(0,1) for which TT is invertible is open. In the deterministic setting one can take Xi=W1,pi0(0,T;V)Lpi(0,T;V)X_{i}={}_{0}\mathrm{W}^{1,p_{i}}(0,T;V^{*})\cap\mathrm{L}^{p_{i}}(0,T;V) and Yi=Lpi(0,T;V)Y_{i}=\mathrm{L}^{p_{i}}(0,T;V^{*}) with 1<p0<2<p1<1<p_{0}<2<p_{1}<\infty and Tu=u+AuTu=u^{\prime}+Au, where W1,pi0(0,T;V){}_{0}\mathrm{W}^{1,p_{i}}(0,T;V^{*}) is defined as the subspace of W1,pi(0,T;V)\mathrm{W}^{1,p_{i}}(0,T;V^{*}) of function that vanish at zero. The boundedness of TT is trivial. Now if one takes θ(0,1)\theta\in(0,1) such that 1θp0+θp1=12\frac{1-\theta}{p_{0}}+\frac{\theta}{p_{1}}=\frac{1}{2}, then T:XθYθT:X_{\theta}\to Y_{\theta} is invertible by (2) for p=2p=2 and with u0=0u_{0}=0. Thus, Sneiberg’s lemma provides us with an open interval around this value of θ\theta, which yields the required result. As far as we could see it seems impossible to define an operator TT in the stochastic framework so that the above steps can be completed.

New estimates for the linear problem

In this paper we circumvent the above problem at the price of a symmetry condition on AA, and prove the following result, where we emphasise that we do not need any regularity conditions on (A,B)(A,B).

Theorem 1.2 (Extrapolation of integrability and regularity).

Suppose that Assumption 1.1 holds and that on [0,T]×Ω[0,T]\times\Omega and for u,vVu,v\in V the symmetry condition Au,v=Av,u\langle Au,v\rangle=\langle Av,u\rangle holds. Then there exists p0>2p_{0}>2 only depending on Λ,λ\Lambda,\lambda such that for all p[2,p0]p\in[2,p_{0}], fLp(Ω×(0,T);V)f\in\mathrm{L}^{p}_{\mathcal{F}}(\Omega\times(0,T);V^{*}), gLp(Ω×(0,T);2(U,H))g\in\mathrm{L}^{p}_{\mathcal{F}}(\Omega\times(0,T);\mathcal{L}_{2}(U,H)) and u0L0p(Ω;(H,V)12/p,p)u_{0}\in\mathrm{L}^{p}_{\mathcal{F}_{0}}(\Omega;(H,V)_{1-\nicefrac{{2}}{{p}},p}), problem (LP) admits a unique solution uLp(Ω×(0,T);V)u\in\mathrm{L}^{p}(\Omega\times(0,T);V). Moreover, for δ(1/p,1/2)\delta\in(1/p,1/2) and θ[0,1/2)\theta\in[0,1/2) we have

uLp(Ω;Cδ1/p([0,T];[H,V]12δ)),anduLp(Ω;C([0,T];(H,V)12/p,p)).u\in\mathrm{L}^{p}(\Omega;\mathrm{C}^{\delta-\nicefrac{{1}}{{p}}}([0,T];[H,V]_{1-2\delta})),\ \ \text{and}\ \ u\in\mathrm{L}^{p}(\Omega;\mathrm{C}([0,T];(H,V)_{1-\nicefrac{{2}}{{p}},p})).

In the above we used the notation [,]θ[\cdot,\cdot]_{\theta} and (,)θ,p(\cdot,\cdot)_{\theta,p} for complex and real interpolation, respectively. The main novelties are Lp\mathrm{L}^{p}-integrability in Ω×(0,T)\Omega\times(0,T) and the improved smoothness of the paths, which is much better than the classical path space C([0,T];H)\mathrm{C}([0,T];H) used in (1).

Theorem 1.2 is a special case of Theorem 2.2, where additional linear estimates are stated as well. As mentioned, we cannot use Sneiberg’s lemma to prove Theorem 1.2 . Instead we will use abstract Stein interpolation. The proof was inspired by the recent work [8] on Lp\mathrm{L}^{p}-bounds for heat semigroups.

The symmetry condition can be written equivalently as A=AA^{*}=A. We do not know if the symmetry condition is needed in Theorem 1.2. It is certainly not necessary. For many operators AA satisfying Assumption 1.1, one can always apply Theorem 1.2 to its symmetric part A+A2\frac{A+A^{*}}{2} and use a perturbation argument A=A+A2+RA=\frac{A+A^{*}}{2}+R with R=AA2R=\frac{A-A^{*}}{2} (see [3, Theorem 3.2]). The only thing which needs to be checked is the following relative boundedness condition: for every ε>0\varepsilon>0 there exists a CεC_{\varepsilon} such that

RvVεvV+CεvH.\|Rv\|_{V^{*}}\leq\varepsilon\|v\|_{V}+C_{\varepsilon}\|v\|_{H}.

For instance for 2m2m-th elliptic differential operators with smooth coefficients in space the highest order differentiation cancels out and RR turns out to be of order 2m12m-1, and thus it satisfies the above relative boundedness condition.

Theorem 1.2 implies that the L2\mathrm{L}^{2}-theory in [2] can be extended to Lp\mathrm{L}^{p}-theory for some range p(2,p0]p\in(2,p_{0}]. Although this leads to minimal changes in the proof, it has many consequences for applications since for several problems it turns out that any p>2p>2 is sufficient. For instance this occurs for SPDEs in dimension d=2d=2. Due to the fact that taking for instance H=H1H=\mathrm{H}^{1}, V=H2V=\mathrm{H}^{2} and V=L2V^{*}=\mathrm{L}^{2}, the space HH does not embed into L\mathrm{L}^{\infty} as this instance of the Sobolev embedding fails. Using Lp\mathrm{L}^{p}-theory instead one only needs that [H,V]12/p=H21/p[H,V]_{1-\nicefrac{{2}}{{p}}}=\mathrm{H}^{2-\nicefrac{{1}}{{p}}} embeds into L\mathrm{L}^{\infty}, which is true in dimension d=2d=2. We will investigate these applications in a subsequent paper.

Universal compactness

Thanks to Theorem 1.2 we obtain a compactness/tightness result for the solution mapping (f,g,u0)u(f,g,u_{0})\mapsto u for (LP) for the path space C([0,T];H)\mathrm{C}([0,T];H). We expect that this has many consequences. In earlier works tightness of laws is obtained with HH replaced by VV^{*} or even larger spaces for different but related settings (cf. [12] and [30]). Tightness plays a key role in the stochastic compactness method (see [9]) for obtaining weak solutions. Using a smaller path space can give more information on the weak solution and its approximation.

Theorem 1.3 (Universal compactness for variational problems).

Suppose that the embedding VHV\hookrightarrow H is compact. Let λ,Λ,T>0\lambda,\Lambda,T>0, M0M\geq 0 and K>0K>0 be fixed. Then there exists a p0>2p_{0}>2 only depending on λ,Λ,T1\lambda,\Lambda,T\vee 1 such that for all p(2,p0]p\in(2,p_{0}], the laws {(u):u}\{\mathscr{L}(u):u\} on C([0,T];H)C([0,T];H) are tight, where uu runs over all strong solutions to (LP) and all (A,B)(A,B) satisfying Assumption 1.1 with A=AA=A^{*}, and all fLp(Ω×(0,T);V)f\in\mathrm{L}^{p}_{\mathcal{F}}(\Omega\times(0,T);V^{*}), gLp(Ω×(0,T);2(U,H))g\in\mathrm{L}^{p}_{\mathcal{F}}(\Omega\times(0,T);\mathcal{L}_{2}(U,H)) and u0L0p(Ω;(H,V)12/p,p)u_{0}\in\mathrm{L}^{p}_{\mathcal{F}_{0}}(\Omega;(H,V)_{1-\nicefrac{{2}}{{p}},p}) satisfying fK\|f\|\leq K, gK\|g\|\leq K and u0K\|u_{0}\|\leq K in the respective norms.

As before, the condition A=AA=A^{*} can be weakened by a perturbation argument.

To illustrate the power of Theorem 1.3 we will show that it can be effectively used in the stochastic compactness method to prove global existence of a system of quasi-linear SPDEs (see Theorem 6.6 below). A technical issue in applying the stochastic compactness method is that the a.s. limit, obtained by tightness and the Skorohod embedding theorem, needs to be identified as a solution. The following advantages turn up as a consequence of the uniform estimate we obtain in the proof of Theorem 1.2:

  1. (i)

    Tightness in C([0,T];H)\mathrm{C}([0,T];H) (if VHV\hookrightarrow H is compact);

  2. (ii)

    uniform estimates in Lp(Ω;C([0,T];H))\mathrm{L}^{p}(\Omega;\mathrm{C}([0,T];H));

  3. (iii)

    uniform estimates in Lp(Ω×(0,T);V)\mathrm{L}^{p}(\Omega\times(0,T);V).

Here we let uu run over all solutions with data (A,B,f,g,u0)(A,B,f,g,u_{0}) as in Theorem 1.3. As a consequence of (1) we obtain subsequences which converge in C([0,T];H)\mathrm{C}([0,T];H). As a consequence of (2) and (3) we obtain weak compactness in Lp(Ω×(0,T);V)\mathrm{L}^{p}(\Omega\times(0,T);V) and uniform integrability of {uC([0,T];H)2:u}\{\|u\|_{\mathrm{C}([0,T];H)}^{2}:u\} and {uL2(0,T;V)2:u}\{\|u\|_{\mathrm{L}^{2}(0,T;V)}^{2}:u\}, where uu runs over all strong solutions for the given data (A,B,f,g,u0)(A,B,f,g,u_{0}) The uniform integrability can be effectively combined with Vitali’s convergence theorem.

Application to a quasi-linear system of SPDEs

On an open and bounded set DdD\subseteq\mathbb{R}^{d} consider the quasilinear system

(QLP) {duα=[i(aijαβ(u)juβ)+iΦiα(u)+ϕα(u)]dt+n1[bn,jαβ(uβ)juβ+gnα(u)]dwn, on D,uα=0, on D,uα(0)=u0α, on D.\displaystyle\left\{\quad\begin{aligned} \,\mathrm{d}u^{\alpha}&=\big{[}\partial_{i}(a^{\alpha\beta}_{ij}(u)\partial_{j}u^{\beta})+\partial_{i}\Phi^{\alpha}_{i}(u)+\phi^{\alpha}(u)\big{]}\,\mathrm{d}t\\ &\qquad\qquad+\sum_{n\geq 1}\big{[}b^{\alpha\beta}_{n,j}(u^{\beta})\partial_{j}u^{\beta}+g^{\alpha}_{n}(u)\big{]}\,\mathrm{d}w_{n},\ \ &\text{ on }D,\\ u^{\alpha}&=0,\ \ &\text{ on }\partial D,\\ u^{\alpha}(0)&=u^{\alpha}_{0},\ \ &\text{ on }D.\end{aligned}\right.

We state a special case of Theorem 6.6.

Theorem 1.4 (Weak existence for quasi-linear systems).

Let h>1h>1. Suppose Assumption 6.1. Then there is p0>2p_{0}>2, depending on T1T\vee 1, λ\lambda, Λ\Lambda, qq and CC from the assumption such that, if p(2,p0]p\in(2,p_{0}] and q(ph,)q\in(ph,\infty), then, given u0L0p(Ω;B2,p,012/p(D))L0q(Ω×D)u_{0}\in\mathrm{L}^{p}_{\mathcal{F}_{0}}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))\cap\mathrm{L}^{q}_{\mathcal{F}_{0}}(\Omega\times D), there exists a weak solution (u~,W~,Ω~,~,~,(~t)t0)(\widetilde{u},\widetilde{W},\widetilde{\Omega},\widetilde{\mathcal{F}},\widetilde{\mathbb{P}},(\widetilde{\mathcal{F}}_{t})_{t\geq 0}) of (QLP).

Besides that we are able to treat highly coupled systems, even in the case of scalar quasilinear equations the result of Theorem 1.4 contains new features:

  1. (i)

    less regularity on the initial data is required;

  2. (ii)

    equations on domains with boundary condition can be considered;

  3. (iii)

    gradient/transport noise terms are included.

Often when applying the stochastic compactness method only the torus is considered for simplicity. Actually, equations on domains with for instance Dirichlet boundary condition can lead to complications:

  • Bootstrapping regularity can be problematic due the presence of boundary values;

  • It is often unclear in what extrapolation space to formulate compactness/tightness.

Open problems

Given the well-posedness theory and the deterministic extrapolation results it is natural to state the following open problem.

Problem 1.5.

Do Theorems 1.2 and 1.3 hold without the symmetry condition A=AA^{*}=A ?

The well-posedness theory of Lions is usually stated in the setting of nonlinear operators (A,B)(A,B) satisfying a continuity, monotonicity and coercivity condition. We covered only the linear setting in Theorems 1.2 and 1.3. This leads to the following natural question.

Problem 1.6.

Do Theorems 1.2 and 1.3 hold in the setting of nonlinear monotone operators (A,B)(A,B) ?

This problem seems to be open even in the deterministic setting.

Organization of the paper

In Section 2 we present the precise assumptions and state a more general version of Theorem 1.2, and we derive Theorem 1.3 from it. In Section 3 we present a result on analytic dependency of solutions on the equation. This will be needed later on in order to apply Stein interpolation. In Section 4 we discuss a case in which we have Lp(Ω×(0,T);V)\mathrm{L}^{p}(\Omega\times(0,T);V)-estimates for all p[2,)p\in[2,\infty), which is used as one of the endpoints in the Stein interpolation; the other endpoint is L2(Ω×(0,T);V)\mathrm{L}^{2}(\Omega\times(0,T);V). Finally, in Section 5 we prove the main result Theorem 2.2, a more general version of Theorem 1.2.

The main weak existence result on quasi-linear systems is stated in Section 6. First, a proof in the case of Lipschitz nonlinearities will be given in Section 7 via Theorem 1.3 and the stochastic compactness method. This intermediate result has its own value as it imposes even less structural assumptions. The general case is covered in Section 8, again via an a priori estimate in Lq\mathrm{L}^{q} and the stochastic compactness method.

Notation

  • The notation aPba\lesssim_{P}b means that there is a constant CC only depending on the parameter PP such that aCba\leq Cb.

  • (V,H,V)(V,H,V^{*}) is the notation for the Gelfand triple of Hilbert spaces.

  • (,)θ,p(\cdot,\cdot)_{\theta,p} stands for real interpolation.

  • [,]θ[\cdot,\cdot]_{\theta} stands for complex interpolation.

  • WW denotes a cylindrical Brownian motion on the real separable Hilbert space UU.

  • 2\mathcal{L}_{2} denotes the Hilbert–Schmidt operators.

For further unexplained notation the reader is referred to [2, 17, 24].

Acknowledgements

The authors thank Max Sauerbrey for useful discussions on the stochastic compactness method and the referee for helpful comments and careful reading.

2. Main extrapolation result for linear equations

We consider a linear stochastic partial differential equation of the form

(P) du(t)+A(t)u(t)dt=f(t)dt+(B(t)u(t)+g(t))dW(t),u(0)=u0.\displaystyle\,\mathrm{d}u(t)+A(t)u(t)\,\mathrm{d}t=f(t)\,\mathrm{d}t+(B(t)u(t)+g(t))\,\mathrm{d}W(t),\quad u(0)=u_{0}.

Let us make precise the setting and assumptions for this part.

Assumption 2.1.

Fix 0<T<0<T<\infty. Let UU be a real separable Hilbert space, and HH and VV be complex separable Hilbert spaces, where VHV\subseteq H continuously and densely. Fix a probability space Ω\Omega with filtration =(t)0tT\mathcal{F}=(\mathcal{F}_{t})_{0\leq t\leq T} and let WW be a real-valued UU-cylindrical Brownian motion adapted to \mathcal{F}. Consider A:Ω×(0,T)(V,V)A\colon\Omega\times(0,T)\to\mathcal{L}(V,V^{*}) and B:Ω×(0,T)(V,2(U,H))B\colon\Omega\times(0,T)\to\mathcal{L}(V,\mathcal{L}_{2}(U,H)), where 2(U,H)\mathcal{L}_{2}(U,H) denotes the space of Hilbert Schmidt operators from UU to HH, both progressively measurable. Suppose that there exist Λ(A),Λ(B),λ>0\Lambda(A),\Lambda(B),\lambda>0 and M0M\geq 0 such that pointwise on [0,T]×Ω[0,T]\times\Omega and vVv\in V one has

(3) ReAv,v12Bv2(U,H)2λvV2MvH2,\displaystyle\operatorname{Re}\,\langle Av,v\rangle-\frac{1}{2}\|Bv\|_{\mathcal{L}_{2}(U,H)}^{2}\geq\lambda\|v\|_{V}^{2}-M\|v\|_{H}^{2},

and

(4) AvVΛ(A)vV&Bv2(U,H)Λ(B)vV.\displaystyle\|Av\|_{V^{*}}\leq\Lambda(A)\|v\|_{V}\quad\&\quad\|Bv\|_{\mathcal{L}_{2}(U,H)}\leq\Lambda(B)\|v\|_{V}.

Put Λ=Λ(A)+Λ(B)\Lambda=\Lambda(A)+\Lambda(B).

The main result of this part is the following.

Theorem 2.2.

Suppose that Assumption 2.1 holds and Au,v=Av,u\langle Au,v\rangle=\langle Av,u\rangle pointwise in [0,T]×Ω[0,T]\times\Omega. Then there exists a p0>2p_{0}>2 such that, for all p[2,p0]p\in[2,p_{0}], fLp(Ω×(0,T);V)f\in\mathrm{L}^{p}_{\mathcal{F}}(\Omega\times(0,T);V^{*}), gLp(Ω×(0,T);2(U,H))g\in\mathrm{L}^{p}_{\mathcal{F}}(\Omega\times(0,T);\mathcal{L}_{2}(U,H)) and u0L0p(Ω;(H,V)12/p,p)u_{0}\in\mathrm{L}^{p}_{\mathcal{F}_{0}}(\Omega;(H,V)_{1-\nicefrac{{2}}{{p}},p}), problem (P) admits a unique solution uLp(Ω×(0,T);V)u\in\mathrm{L}^{p}(\Omega\times(0,T);V), and for any θ[0,1/2)\theta\in[0,\nicefrac{{1}}{{2}}) there is a constant KθK_{\theta} depending only on Λ(A)\Lambda(A), Λ(B)\Lambda(B), λ\lambda, MM, T1T\vee 1, θ\theta such that

(5) uLp(Ω;Hθ,p(0,T;[H,V]12θ))KθCu0,f,g,\displaystyle\|u\|_{\mathrm{L}^{p}(\Omega;\mathrm{H}^{\theta,p}(0,T;[H,V]_{1-2\theta}))}\leq K_{\theta}C_{u_{0},f,g},

where

(6) Cu0,f,g=u0Lp(Ω;(H,V)12/p,p)+fLp(Ω×(0,T);V)+gLp(Ω×(0,T);2(U,H)).\displaystyle C_{u_{0},f,g}=\|u_{0}\|_{\mathrm{L}^{p}(\Omega;(H,V)_{1-\nicefrac{{2}}{{p}},p})}+\|f\|_{\mathrm{L}^{p}(\Omega\times(0,T);V^{*})}+\|g\|_{\mathrm{L}^{p}(\Omega\times(0,T);\mathcal{L}_{2}(U,H))}.

Also, there is a constant K0K_{0} depending only on Λ(A)\Lambda(A), Λ(B)\Lambda(B), λ\lambda, MM, T1T\vee 1, and pp such that

(7) uLp(Ω;C([0,T];(H,V)12/p,p))K0Cu0,f,g.\displaystyle\|u\|_{\mathrm{L}^{p}(\Omega;\mathrm{C}([0,T];(H,V)_{1-\nicefrac{{2}}{{p}},p}))}\leq K_{0}C_{u_{0},f,g}.

Moreover, for θ(1/p,1/2)\theta\in(\nicefrac{{1}}{{p}},\nicefrac{{1}}{{2}}), there is a constant KθK_{\theta} depending only on Λ(A)\Lambda(A), Λ(B)\Lambda(B), λ\lambda, T1T\vee 1, and θ\theta such that

(8) uLp(Ω;Cθ1/p([0,T];[H,V]12θ))\displaystyle\|u\|_{\mathrm{L}^{p}(\Omega;\mathrm{C}^{\theta-\nicefrac{{1}}{{p}}}([0,T];[H,V]_{1-2\theta}))} KθCu0,f,g.\displaystyle\leq K_{\theta}C_{u_{0},f,g}.

The symbol Hθ,p\mathrm{H}^{\theta,p} denotes the (vector-valued) Bessel potential space (see [17, 19]), and the subscripts \mathcal{F} and 0\mathcal{F}_{0} refer to the subspaces of Lp\mathrm{L}^{p} of \mathcal{F}-progressively measurable and 0\mathcal{F}_{0}-adapted functions.

Remark 2.3 (Real problems).

In virtue of complexification, the result does also apply to problems over real Hilbert spaces. We are going to use this fact later on.

As a consequence of Theorem 2.2 we can already prove the compactness/tightness result of Theorem 1.3.

Proof of Theorem 1.3.

Let SS denote the set of solutions described in Theorem 1.3. Fix θ(1/p,1/2)\theta\in(1/p,1/2). By Theorem 2.2, SS is bounded in Lp(Ω;Cθ1/p([0,T];[H,V]12θ))\mathrm{L}^{p}(\Omega;\mathrm{C}^{\theta-\nicefrac{{1}}{{p}}}([0,T];[H,V]_{1-2\theta})) by KθCu0,f,gK_{\theta}C_{u_{0},f,g}. Let ε>0\varepsilon>0. Choose R>0R>0 such that KθpCu0,f,gpRpεK_{\theta}^{p}C_{u_{0},f,g}^{p}R^{-p}\leq\varepsilon. Then for all uSu\in S,

(uCθ1/p([0,T];[H,V]12θ)R)Rp𝔼uCθ1/p([0,T];[H,V]12θ)pε.\mathbb{P}(\|u\|_{\mathrm{C}^{\theta-\nicefrac{{1}}{{p}}}([0,T];[H,V]_{1-2\theta})}\geq R)\leq R^{-p}\mathbb{E}\|u\|_{\mathrm{C}^{\theta-\nicefrac{{1}}{{p}}}([0,T];[H,V]_{1-2\theta})}^{p}\leq\varepsilon.

By the vector-valued Arzéla–Ascoli theorem [22, Theorem III.3.1], the embedding

Cθ1/p([0,T];[H,V]12θ)C([0,T];H)\mathrm{C}^{\theta-\nicefrac{{1}}{{p}}}([0,T];[H,V]_{1-2\theta})\hookrightarrow\mathrm{C}([0,T];H)

is compact, and therefore {(u):uS}\{\mathcal{L}(u):u\in S\} is tight. ∎

3. L2\mathrm{L}^{2}-estimates and analytic dependency

The central finding of this section is that solutions to a complex family of variational problems have analytic dependence provided this was the case for the complex family, see Proposition 3.3 and Corollary 3.4. Before, we will briefly discuss well-posedness of (P) in a complex setting.

To ease notation, we introduce data and solution spaces for the variational setting.

Definition 3.1.

Define the data spaces E0L2(Ω×(0,T);V)\mathrm{E}_{0}\coloneqq\mathrm{L}^{2}_{\mathcal{F}}(\Omega\times(0,T);V^{*}) and 𝔼12L2(Ω×(0,T);2(U,H)){\mathbb{E}_{\frac{1}{2}}}\coloneqq\mathrm{L}^{2}_{\mathcal{F}}(\Omega\times(0,T);\mathcal{L}_{2}(U,H)), and the solution space E1L2(Ω×(0,T);V)L2(Ω;C([0,T];H))\mathrm{E}_{1}\coloneqq\mathrm{L}^{2}_{\mathcal{F}}(\Omega\times(0,T);V)\cap\mathrm{L}^{2}_{\mathcal{F}}(\Omega;\mathrm{C}([0,T];H)), where the subscript \mathcal{F} indicates the subspace of progressively measurable functions.

3.1. A (complex) variational setting

The following variational well-posedness result is well-known in its real formulation [24]. The complex version follows by “forgetting” the complex structure. For instance, we can interpret a complex vector space HH as a real vector space if we equip it with the inner product (|)H=Re(|)H(\cdot|\cdot)_{H_{\mathbb{R}}}=\operatorname{Re}(\cdot|\cdot)_{H} and so on.

Proposition 3.2.

Let fE0f\in\mathrm{E}_{0}, g𝔼12g\in{\mathbb{E}_{\frac{1}{2}}}. Suppose Assumption 2.1 holds. Then (P) with u0L02(Ω;H)u_{0}\in\mathrm{L}^{2}_{\mathcal{F}_{0}}(\Omega;H) has a unique solution uE1u\in\mathrm{E}_{1}, and there exists a constant CLC_{L} depending on Λ\Lambda, λ\lambda, MM, and T1T\vee 1, such that

(9) uE1CL(u0L2(Ω;H)+fE0+g𝔼12).\displaystyle\|u\|_{\mathrm{E}_{1}}\leq C_{L}\bigl{(}\|u_{0}\|_{\mathrm{L}^{2}(\Omega;H)}+\|f\|_{\mathrm{E}_{0}}+\|g\|_{{\mathbb{E}_{\frac{1}{2}}}}\bigr{)}.

3.2. Analytic dependence of solutions

In the following result we show analytic dependence of solutions.

Proposition 3.3.

Let OO\subseteq\mathbb{C} be open and for zOz\in O let AzA_{z} and BzB_{z} be operator functions that satisfy Assumption 2.1 uniformly in zz and that depend analytically on zz in the uniform operator topology. Furthermore, let fE0f\in\mathrm{E}_{0}, g𝔼12g\in{\mathbb{E}_{\frac{1}{2}}} and u0Hu_{0}\in H. Then, for fixed zz, the unique solution uzu_{z} of the problem

(10) du(t)+Az(t)u(t)dt=f(t)dt+(Bz(t)u(t)+g(t))dW(t),u(0)=u0,\displaystyle\,\mathrm{d}u(t)+A_{z}(t)u(t)\,\mathrm{d}t=f(t)\,\mathrm{d}t+(B_{z}(t)u(t)+g(t))\,\mathrm{d}W(t),\quad u(0)=u_{0},

gives rise to an analytic function OzuzE1O\ni z\mapsto u_{z}\in\mathrm{E}_{1}.

Proof.

Since we consider differences below, we have zero initial data in the equations in this proof. For fixed zz, Proposition 3.2 yields a unique solution uzu_{z} of (10).

Step 1: OzuzE1O\ni z\mapsto u_{z}\in\mathrm{E}_{1} is continuous.

Let zOz\in O and hh small enough so that z+hOz+h\in O. The function vuz+huzv\coloneqq u_{z+h}-u_{z} is the unique solution of

(11) dv(t)+Az+hv(t)dt=(Az+hAz)uzdt+(Bz+hv(t)+(Bz+hBz)uz)dW(t).\displaystyle\,\mathrm{d}v(t)+A_{z+h}v(t)\,\mathrm{d}t=-(A_{z+h}-A_{z})u_{z}\,\mathrm{d}t+(B_{z+h}v(t)+(B_{z+h}-B_{z})u_{z})\,\mathrm{d}W(t).

Therefore, by Proposition 3.2,

(12) vE1\displaystyle\|v\|_{\mathrm{E}_{1}} CL((Az+hAz)uzE0+(Bz+hBz)uz𝔼12).\displaystyle\leq C_{L}\bigl{(}\|(A_{z+h}-A_{z})u_{z}\|_{\mathrm{E}_{0}}+\|(B_{z+h}-B_{z})u_{z}\|_{{\mathbb{E}_{\frac{1}{2}}}}\bigr{)}.

Hence, continuity of AzA_{z} and BzB_{z} yield the claim.

Step 2: OzuzE1O\ni z\mapsto u_{z}\in\mathrm{E}_{1} is analytic.

Write DhD^{h} for difference quotients of uu, AA, and BB with respect to zz, for example Dhu=u(z+h)u(z)hD^{h}u=\frac{u(z+h)-u(z)}{h}. One has that DhuD^{h}u is the unique solution of

(13) dv(t)+Azvdt=(DhAz)uz+hdt+(Bzv+(DhBz)uz+h)dW(t).\displaystyle\,\mathrm{d}v(t)+A_{z}v\,\mathrm{d}t=-(D^{h}A_{z})u_{z+h}\,\mathrm{d}t+(B_{z}v+(D^{h}B_{z})u_{z+h})\,\mathrm{d}W(t).

We show that DhuD^{h}u is a Cauchy sequence in E1\mathrm{E}_{1}. Then, by the very definition of complex differentiability, zuzz\mapsto u_{z} is analytic. To this end, consider h1h2h_{1}\neq h_{2}. The difference wDh1uDh2uw\coloneqq D^{h_{1}}u-D^{h_{2}}u solves

(14) dw(t)+Azwdt=((Dh1ADh2A)uz+h1+(Dh2A)(uz+h1uz+h2))dt\displaystyle\,\mathrm{d}w(t)+A_{z}w\,\mathrm{d}t=-\Bigl{(}\bigl{(}D^{h_{1}}A-D^{h_{2}}A\bigr{)}u_{z+h_{1}}+(D^{h_{2}}A)(u_{z+h_{1}}-u_{z+h_{2}})\Bigr{)}\,\mathrm{d}t
(15) +(Bzw+(Dh1BDh2B)uz+h1+(Dh2B)(uz+h1uz+h2))dW(t).\displaystyle+\Bigl{(}B_{z}w+\bigl{(}D^{h_{1}}B-D^{h_{2}}B\bigr{)}u_{z+h_{1}}+(D^{h_{2}}B)(u_{z+h_{1}}-u_{z+h_{2}})\Bigr{)}\,\mathrm{d}W(t).

A calculation using Proposition 3.2 shows

(16) Dh1uDh2uE1\displaystyle\|D^{h_{1}}u-D^{h_{2}}u\|_{\mathrm{E}_{1}}
(17) \displaystyle\leq{} CL2(Dh1ADh2A(V,V)+Dh1BDh2B(V,2(U,H)))(fE0+g𝔼12)\displaystyle C_{L}^{2}\Bigl{(}\|D^{h_{1}}A-D^{h_{2}}A\|_{\mathcal{L}(V,V^{*})}+\|D^{h_{1}}B-D^{h_{2}}B\|_{\mathcal{L}(V,\mathcal{L}_{2}(U,H))}\Bigr{)}(\|f\|_{\mathrm{E}_{0}}+\|g\|_{{\mathbb{E}_{\frac{1}{2}}}})
(18) +CL(Dh2A(V,V)+Dh2B(V,2(U,H)))uz+h1uz+h2E1.\displaystyle\quad+C_{L}\Bigl{(}\|D^{h_{2}}A\|_{\mathcal{L}(V,V^{*})}+\|D^{h_{2}}B\|_{\mathcal{L}(V,\mathcal{L}_{2}(U,H))}\Bigr{)}\|u_{z+h_{1}}-u_{z+h_{2}}\|_{\mathrm{E}_{1}}.

First, the difference quotients of AA and BB are Cauchy sequences for AA and BB are holomorphic. Second, the difference quotients of AA and BB are, again by analyticity, uniformly bounded. Third, uz+h1uz+h2E1\|u_{z+h_{1}}-u_{z+h_{2}}\|_{\mathrm{E}_{1}} tends to zero by Step 1. We conclude that DhuD^{h}u is a Cauchy sequence as claimed. ∎

Corollary 3.4.

Let OO\subseteq\mathbb{C} be open and for zOz\in O let AzA_{z} and BzB_{z} be operator functions that satisfy Assumption 2.1 uniformly in zz and that depend analytically on zz in the uniform operator topology. Furthermore, let f:OE0f:O\to\mathrm{E}_{0}, g:O𝔼12g:O\to{\mathbb{E}_{\frac{1}{2}}} and u0:OHu_{0}:O\to H be analytic. Then, for fixed zz, the unique solution uzu_{z} of the problem

(19) du(t)+Az(t)u(t)dt=fz(t)dt+(Bz(t)u(t)+gz(t))dW(t),u(0)=u0,\displaystyle\,\mathrm{d}u(t)+A_{z}(t)u(t)\,\mathrm{d}t=f_{z}(t)\,\mathrm{d}t+(B_{z}(t)u(t)+g_{z}(t))\,\mathrm{d}W(t),\quad u(0)=u_{0},

gives rise to an analytic function OzuzE1O\ni z\mapsto u_{z}\in\mathrm{E}_{1}.

Proof.

From Proposition 3.3 it follows that Sz:O(E0×𝔼12×H,E1)S_{z}:O\to\mathcal{L}(\mathrm{E}_{0}\times{\mathbb{E}_{\frac{1}{2}}}\times H,\mathrm{E}_{1}) given by Sz=uzS_{z}=u_{z}, where uzu_{z} is the solution to (10), is analytic in the strong operator topology. Therefore, SS is analytic in the uniform operator topology (see [5, Proposition A.3]). Now the analyticity follows as the solution to (19) is given by Sz(fz,gz,u0,z)S_{z}(f_{z},g_{z},u_{0,z}) and is the composition of analytic functions. ∎

4. Lp\mathrm{L}^{p}-estimates for small perturbations of the autonomous case

In this section, we establish Lp\mathrm{L}^{p}-theory by means of a perturbation result with the autonomous case. The quantities in the assumption of Theorem 4.6 look cumbersome at first glance, but we will see in Section 5 that they are, in fact, very closely tied to the concept of coercivity.

Throughout this section, fix some p>2p>2. We extend Definition 3.1.

Definition 4.1.

Put E0pLp(Ω×(0,T);V)\mathrm{E}^{p}_{0}\coloneqq\mathrm{L}^{p}_{\mathcal{F}}(\Omega\times(0,T);V^{*}) and 𝔼12pLp(Ω×(0,T);2(U,H)){\mathbb{E}^{p}_{\frac{1}{2}}}\coloneqq\mathrm{L}^{p}_{\mathcal{F}}(\Omega\times(0,T);\mathcal{L}_{2}(U,H)) for the data spaces and E1pLp(Ω×(0,T);V)Lp(Ω;C([0,T];(H,V)12/p,p))\mathrm{E}^{p}_{1}\coloneqq\mathrm{L}^{p}_{\mathcal{F}}(\Omega\times(0,T);V)\cap\mathrm{L}^{p}_{\mathcal{F}}(\Omega;\mathrm{C}([0,T];(H,V)_{1-\nicefrac{{2}}{{p}},p})) for the solution space.

Let us introduce a reference operator for the perturbation argument.

Definition 4.2.

Consider the operator A0:VVA_{0}\colon V\to V^{*} given by A0(u),v(u|v)V\langle A_{0}(u),v\rangle\coloneqq(u|v)_{V}.

Remark 4.3.

The operator A0A_{0} is invertible, positive and self-adjoint.

The following deterministic maximal LpL^{p}-regularity result holds.

Lemma 4.4.

There is a constant CpC_{p} such that, for all μ>0\mu>0, fLp(0,T;V)f\in\mathrm{L}^{p}(0,T;V^{*}), and uu a strong solution to u+(μA0)u=fu^{\prime}+(\mu A_{0})u=f with u(0)=0u(0)=0, one has the a-priori estimate

(20) μuLp(0,T;V)CpfLp(0,T;V).\displaystyle\mu\|u\|_{\mathrm{L}^{p}(0,T;V)}\leq C_{p}\|f\|_{\mathrm{L}^{p}(0,T;V^{*})}.
Proof.

The fact that A0A_{0} has maximal LpL^{p}-regularity on +\mathbb{R}_{+} follows from de Simon’s result (see [19, Corollary 17.3.8]). From the proof of [19, Theorem 17.2.26] it follows that μA0\mu A_{0} has maximal LpL^{p}-regularity on +\mathbb{R}_{+} with the same constant CpC_{p}. Therefore, by [19, Lemma 17.2.16]), μA0\mu A_{0} has maximal LpL^{p}-regularity on (0,T)(0,T) with the same constant CpC_{p}. This gives the desired bound with constant independent of μ\mu

Next we will present a similar result in the stochastic setting, but without the optimal scaling in μ\mu as this will not be needed later.

Lemma 4.5.

For any μ>0\mu>0, the operator μA0\mu A_{0} admits stochastic maximal Lp\mathrm{L}^{p}-regularity, that is to say, there is c(p,μ)>0c(p,\mu)>0 such that for all fE0pf\in\mathrm{E}^{p}_{0} and g𝔼12pg\in{\mathbb{E}^{p}_{\frac{1}{2}}} there is a unique strong solution uE1u\in\mathrm{E}_{1} to

(21) du(t)+μA0udt=f(t)dt+g(t)dW(t),u(0)=0,\displaystyle\,\mathrm{d}u(t)+\mu A_{0}u\,\mathrm{d}t=f(t)\,\mathrm{d}t+g(t)\,\mathrm{d}W(t),\qquad u(0)=0,

satisfying the estimate

(22) uE1pc(p,μ)(fE0p+g𝔼12p).\displaystyle\|u\|_{\mathrm{E}^{p}_{1}}\leq c(p,\mu)\bigl{(}\|f\|_{\mathrm{E}^{p}_{0}}+\|g\|_{{\mathbb{E}^{p}_{\frac{1}{2}}}}\bigr{)}.
Proof.

Let μ>0\mu>0. By linearity and Lemma 4.4 it suffices to consider f=0f=0. As μA0\mu A_{0} is positive and self-adjoint on VV^{*}, it is also positive and self-adjoint on HH (see [27, Proposition 1.24]). Therefore, it has a bounded H\mathrm{H}^{\infty}-calculus with constant 11 (see [18, Proposition 10.2.23]). Moreover, HH is, as a separable Hilbert space, isomorphic to L2()\mathrm{L}^{2}(\mathbb{R}). Therefore, from the definition of A0A_{0} and from [26] we obtain

μ1/2uE1p=(μA0)1/2uLp(Ω×+;H)c(p)gLp(Ω×+;2(U,H))=c(p)g𝔼12p.\mu^{1/2}\|u\|_{\mathrm{E}^{p}_{1}}=\|(\mu A_{0})^{1/2}u\|_{\mathrm{L}^{p}(\Omega\times\mathbb{R}_{+};H)}\leq c(p)\|g\|_{\mathrm{L}^{p}(\Omega\times\mathbb{R}_{+};\mathcal{L}_{2}(U,H))}=c(p)\|g\|_{{\mathbb{E}^{p}_{\frac{1}{2}}}}.

Theorem 4.6.

Let fE0pf\in\mathrm{E}^{p}_{0} and g𝔼12pg\in{\mathbb{E}^{p}_{\frac{1}{2}}}. Employ Assumptions 2.1 with M=0M=0 and recall the constant Λ(B)\Lambda(B). Suppose that there are ε,δ>0\varepsilon,\delta>0 such that

  1. (i)

    for some μ>0\mu>0 one has Cpμ1AμA0(V,V)1εC_{p}\mu^{-1}\|A-\mu A_{0}\|_{\mathcal{L}(V,V^{*})}\leq 1-\varepsilon, where CpC_{p} is the constant from Lemma 4.4, and

  2. (ii)

    cΛ(B)1δc\Lambda(B)\leq 1-\delta, where c=c(Λ,μ,T,ε,p)c=c(\Lambda,\mu,T,\varepsilon,p).

Then there exists a unique strong solution uE1pu\in\mathrm{E}^{p}_{1} to (P) with u0=0u_{0}=0 satisfying the estimate

(23) uE1pcδ(fE0p+g𝔼12p).\displaystyle\|u\|_{\mathrm{E}^{p}_{1}}\leq\frac{c}{\delta}\bigl{(}\|f\|_{\mathrm{E}^{p}_{0}}+\|g\|_{{\mathbb{E}^{p}_{\frac{1}{2}}}}\bigr{)}.
Proof.

We tacitly impose the initial condition u(0)=0u(0)=0 whenever we speak about equations in this proof. First, we investigate the problem with B=0B=0, that is to say, we want to show that there exists a unique strong solution uE1pu\in\mathrm{E}^{p}_{1} to the stochastic problem

(24) du(t)+A(t)u(t)dt=f(t)dt+g(t)dW(t)\displaystyle\,\mathrm{d}u(t)+A(t)u(t)\,\mathrm{d}t=f(t)\,\mathrm{d}t+g(t)\,\mathrm{d}W(t)

satisfying the estimate

(25) uE1pc(fE0p+g𝔼12p)\displaystyle\|u\|_{\mathrm{E}^{p}_{1}}\leq c(\|f\|_{\mathrm{E}^{p}_{0}}+\|g\|_{{\mathbb{E}^{p}_{\frac{1}{2}}}})

with a constant cc depending only on ε\varepsilon, Λ\Lambda, μ\mu, T1T\vee 1, and pp. This constant is the constant postulated in hypothesis (ii) of the theorem.

Step 1: Reduction to an a-priori estimate.

For θ[0,1]\theta\in[0,1] consider Aθ(1θ)μA0+θAA_{\theta}\coloneqq(1-\theta)\mu A_{0}+\theta A. Since AθμA0=θ(AμA0)A_{\theta}-\mu A_{0}=\theta(A-\mu A_{0}), one has

Cpμ1AθμA0(V,V)=Cpμ1θAμA0(V,V)1ε.C_{p}\mu^{-1}\|A_{\theta}-\mu A_{0}\|_{\mathcal{L}(V,V^{*})}=C_{p}\mu^{-1}\theta\|A-\mu A_{0}\|_{\mathcal{L}(V,V^{*})}\leq 1-\varepsilon.

Therefore, in the light of a stochastic version of the method of continuity (see [29, Prop. 3.10]), it suffices to show the a priori estimate (25) for any strong solution uu of (24).

Step 2: Show the a-priori estimate.

Let uu a strong solution of (24). By Lemma 4.5 there is a strong solution vE1pv\in\mathrm{E}^{p}_{1} of

dv(t)+(μA0)v(t)dt=f(t)dt+g(t)dW(t)\,\mathrm{d}v(t)+(\mu A_{0})v(t)\,\mathrm{d}t=f(t)\,\mathrm{d}t+g(t)\,\mathrm{d}W(t)

satisfying the estimate

(26) vE1pc(p,μ)(fE0p+g𝔼12p).\displaystyle\|v\|_{\mathrm{E}^{p}_{1}}\leq c(p,\mu)(\|f\|_{\mathrm{E}^{p}_{0}}+\|g\|_{{\mathbb{E}^{p}_{\frac{1}{2}}}}).

Define wuvE1pw\coloneqq u-v\in\mathrm{E}^{p}_{1}. Fix ωΩ\omega\in\Omega. Use the shorthand notation wω(t)=w(ω,t)w_{\omega}(t)=w(\omega,t), and so on. By construction, wωw_{\omega} is almost surely a strong solution to the deterministic problem

twω+(μA0)wω=(AωμA0)uω.\partial_{t}w_{\omega}+(\mu A_{0})w_{\omega}=-(A_{\omega}-\mu A_{0})u_{\omega}.

Lemma 4.4 lets us compute

(27) μwωLp(0,T;V)\displaystyle\mu\|w_{\omega}\|_{\mathrm{L}^{p}(0,T;V)} Cp(AωμA0)uωLp(0,T;V)\displaystyle\leq C_{p}\|(A_{\omega}-\mu A_{0})u_{\omega}\|_{\mathrm{L}^{p}(0,T;V^{*})}
(28) CpAωμA0(V,V)(vωLp(0,T;V)+wωLp(0,T;V)).\displaystyle\leq C_{p}\|A_{\omega}-\mu A_{0}\|_{\mathcal{L}(V,V^{*})}\Bigl{(}\|v_{\omega}\|_{\mathrm{L}^{p}(0,T;V)}+\|w_{\omega}\|_{\mathrm{L}^{p}(0,T;V)}\Bigr{)}.

Divide this by μ\mu and use the assumption to give

(29) wωLp(0,T;V)\displaystyle\|w_{\omega}\|_{\mathrm{L}^{p}(0,T;V)} Cpμ1(AωμA0)(V,V)(vωLp(0,T;V)+wωLp(0,T;V))\displaystyle\leq C_{p}\mu^{-1}\|(A_{\omega}-\mu A_{0})\|_{\mathcal{L}(V,V^{*})}\Bigl{(}\|v_{\omega}\|_{\mathrm{L}^{p}(0,T;V)}+\|w_{\omega}\|_{\mathrm{L}^{p}(0,T;V)}\Bigr{)}
(30) (1ε)vωLp(0,T;V)+(1ε)wωLp(0,T;V).\displaystyle\leq(1-\varepsilon)\|v_{\omega}\|_{\mathrm{L}^{p}(0,T;V)}+(1-\varepsilon)\|w_{\omega}\|_{\mathrm{L}^{p}(0,T;V)}.

Absorb the second term on the right-hand side into the left-hand side to deduce

(31) wωLp(0,T;V)1εεvωLp(0,T;V).\displaystyle\|w_{\omega}\|_{\mathrm{L}^{p}(0,T;V)}\leq\frac{1-\varepsilon}{\varepsilon}\|v_{\omega}\|_{\mathrm{L}^{p}(0,T;V)}.

According to [29, Step 1 in Thm. 3.9], uu is progressively measurable. Hence, averaging the last bound over Ω\Omega yields

(32) wLp(Ω×(0,T);V)1εεvLp(Ω×(0,T);V).\displaystyle\|w\|_{\mathrm{L}^{p}(\Omega\times(0,T);V)}\leq\frac{1-\varepsilon}{\varepsilon}\|v\|_{\mathrm{L}^{p}(\Omega\times(0,T);V)}.

In conjunction with u=v+wu=v+w and (26) we obtain in summary

(33) uLp(Ω×(0,T);V)\displaystyle\|u\|_{\mathrm{L}^{p}(\Omega\times(0,T);V)} vLp(Ω×(0,T);V)+wLp(Ω×(0,T);V)c(p,μ)ε(fE0p+g𝔼12p).\displaystyle\leq\|v\|_{\mathrm{L}^{p}(\Omega\times(0,T);V)}+\|w\|_{\mathrm{L}^{p}(\Omega\times(0,T);V)}\leq\frac{c(p,\mu)}{\varepsilon}\bigl{(}\|f\|_{\mathrm{E}^{p}_{0}}+\|g\|_{{\mathbb{E}^{p}_{\frac{1}{2}}}}\bigr{)}.

Step 3: Including BB.

As is standard (see [29, Prop. 3.10], for instance), we solve the problem (P) with u0=0u_{0}=0,

(34) du(t)+A(t)u(t)dt=f(t)dt+(B(t)u(t)+g(t))dW(t),\displaystyle\,\mathrm{d}u(t)+A(t)u(t)\,\mathrm{d}t=f(t)\,\mathrm{d}t+(B(t)u(t)+g(t))\,\mathrm{d}W(t),

using a fixed-point argument. For ΦE1p\Phi\in\mathrm{E}^{p}_{1}, replace BuBu by BΦB\Phi in (34). This problem possesses a unique solution in E1p\mathrm{E}^{p}_{1} in virtue of the first part of the proof. Write R(Φ)R(\Phi) for it. To show unique solvability of (34), it suffices to show that RR is a strict contraction.

Let Φ1,Φ2E1p\Phi_{1},\Phi_{2}\in\mathrm{E}^{p}_{1}. By linearity, R(Φ1)R(Φ2)R(\Phi_{1})-R(\Phi_{2}) is the unique strong solution to the problem

du(t)+A(t)u(t)dt=B(t)(Φ1(t)Φ2(t))dW(t).\,\mathrm{d}u(t)+A(t)u(t)\,\mathrm{d}t=B(t)(\Phi_{1}(t)-\Phi_{2}(t))\,\mathrm{d}W(t).

Hence, (25) with f=0f=0 and g=B(Φ1Φ2)g=B(\Phi_{1}-\Phi_{2}) gives

(35) R(Φ1)R(Φ2)E1p\displaystyle\|R(\Phi_{1})-R(\Phi_{2})\|_{\mathrm{E}^{p}_{1}} CB(Φ1Φ2)𝔼12pCΛ(B)Φ1Φ2E1p.\displaystyle\leq C\|B(\Phi_{1}-\Phi_{2})\|_{{\mathbb{E}^{p}_{\frac{1}{2}}}}\leq C\Lambda(B)\|\Phi_{1}-\Phi_{2}\|_{\mathrm{E}^{p}_{1}}.

Thus, RR is a strict contraction by the assumption CΛ(B)1δC\Lambda(B)\leq 1-\delta.

For the estimate, apply the a-priori estimate (25) once more to find

(36) uE1p\displaystyle\|u\|_{\mathrm{E}^{p}_{1}} C(fE0p+g𝔼12p+Bu𝔼12p)C(fE0p+g𝔼12p)+(1δ)uE1p.\displaystyle\leq C(\|f\|_{\mathrm{E}^{p}_{0}}+\|g\|_{{\mathbb{E}^{p}_{\frac{1}{2}}}}+\|Bu\|_{{\mathbb{E}^{p}_{\frac{1}{2}}}})\leq C(\|f\|_{\mathrm{E}^{p}_{0}}+\|g\|_{{\mathbb{E}^{p}_{\frac{1}{2}}}})+(1-\delta)\|u\|_{\mathrm{E}^{p}_{1}}.

We can absorb the second term of the right-hand side into the left-hand side to conclude. ∎

5. Extrapolation of integrability in time

Our main result follows from an interpolation argument using abstract Stein interpolation [33]. For details on interpolation the reader is referred to [7, 17, 32]. The last two sections constitute the endpoint cases of this interpolation. Before we can conclude in Section 5.2, we have to construct an analytic family of auxiliary problems first.

Assumption 5.1.

For fixed tt, the form V×V(u,v)A(ω,t)u,vV\times V\ni(u,v)\mapsto\langle A(\omega,t)u,v\rangle is hermitian almost surely. To ease notation, write a(u,v)aω,t(u,v)A(ω,t)u,va(u,v)\coloneqq a_{\omega,t}(u,v)\coloneqq\langle A(\omega,t)u,v\rangle.

This symmetry assumption is used in a crucial way in Lemmas 5.4 and 5.7.

Convention 5.2.

We ignore the dependence on (ω,t)(\omega,t) in our notation. For instance, we let AA denote the operator VVV\to V^{*} given by Au,vA(ω,t)u,v\langle Au,v\rangle\coloneqq\langle A(\omega,t)u,v\rangle. Also, we put a(u,v)Au,va(u,v)\coloneqq\langle Au,v\rangle and a[u]a(u,u)a[u]\coloneqq a(u,u).

The following lemma translates upper bounds on the diagonal of V×VV\times V to the whole form. That the optimal constant is 11 is a consequence of Assumption 5.1, but is actually not needed later. A simpler polarization argument without Assumption 5.1 would lead to an extra factor 22.

Lemma 5.3.

Suppose that Assumption 5.1 holds. If for some c>0c>0 one has |a[u]|cuV2|a[u]|\leq c\|u\|_{V}^{2}, then also |a(u,v)|cuVvV|a(u,v)|\leq c\|u\|_{V}\|v\|_{V} for all u,vVu,v\in V.

Proof.

Let u,vVu,v\in V. We can of course assume u0vu\neq 0\neq v. There is a complex number zz with |z|=1|z|=1 such that za(u,v)za(u,v) is real and positive. By the polarization identity one has

(37) za(u,v)=a(zu,v)=14(a[zu+v]a[zuv]+ia[zu+iv]ia[zuiv]).\displaystyle za(u,v)=a(zu,v)=\frac{1}{4}\bigl{(}a[zu+v]-a[zu-v]+\mathrm{i}a[zu+\mathrm{i}v]-\mathrm{i}a[zu-\mathrm{i}v]\bigr{)}.

Since aa is hermitian, a[u]a[u] is real for any uVu\in V. Hence, taking the real part of (37) gives

|a(u,v)|=za(u,v)=14(a[zu+v]a[zuv]).|a(u,v)|=za(u,v)=\frac{1}{4}\bigl{(}a[zu+v]-a[zu-v]\bigr{)}.

Expanding the quadratic forms in conjunction with the bound on the diagonal leads to

(38) |a(u,v)|=14(2a[zu]+2a[v])c2(uV2+vV2).\displaystyle|a(u,v)|=\frac{1}{4}\bigl{(}2a[zu]+2a[v]\bigr{)}\leq\frac{c}{2}\bigl{(}\|u\|_{V}^{2}+\|v\|_{V}^{2}\bigr{)}.

It remains to use |a(u,v)|=|a(λu,λ1v)||a(u,v)|=|a(\lambda u,\lambda^{-1}v)|, apply (38), and minimize over λ>0\lambda>0. ∎

5.1. Complex family of auxiliary problems

Lemma 5.4.

Suppose that Assumption 2.1 holds with M=0M=0 and that Assumption 5.1 holds. Put μ=Λ(A)\mu=\Lambda(A) and ρ=λΛ(A)\rho=\frac{\lambda}{\Lambda(A)}. Then one has

|μ1(a[v]12Bv2(U,H)2)vV2|(1ρ)vV2.\Bigl{|}\mu^{-1}\Bigl{(}a[v]-\frac{1}{2}\|Bv\|_{\mathcal{L}_{2}(U,H)}^{2}\Bigr{)}-\|v\|_{V}^{2}\Bigr{|}\leq(1-\rho)\|v\|_{V}^{2}.
Proof.

To ease notation, put c[v]a[v]12Bv2(U,H)2c[v]\coloneqq a[v]-\frac{1}{2}\|Bv\|_{\mathcal{L}_{2}(U,H)}^{2}. By Assumptions 2.1 (with M=0M=0) and 5.1 one has λvV2c[v]Λ(A)vV2\lambda\|v\|_{V}^{2}\leq c[v]\leq\Lambda(A)\|v\|_{V}^{2}. Let s0s\geq 0. Assumption 5.1 implies that sc[v]vV2sc[v]-\|v\|_{V}^{2} is real, hence

(39) |sc[v]vV2|=max(sc[v]vV2,vV2sc[v]).\displaystyle|sc[v]-\|v\|_{V}^{2}|=\max(sc[v]-\|v\|_{V}^{2},\|v\|_{V}^{2}-sc[v]).

Consider the first term. Since

sc[v]vV2(sΛ(A)1)vV2,sc[v]-\|v\|_{V}^{2}\leq(s\Lambda(A)-1)\|v\|_{V}^{2},

the maximum in (39) coincides with the second term for any choice s1Λ(A)s\leq\frac{1}{\Lambda(A)}. On the other hand,

vV2sc[v](1sλ)vV2,\|v\|_{V}^{2}-sc[v]\leq(1-s\lambda)\|v\|_{V}^{2},

so the right-hand side of the last estimate is minimal for the maximal admissible choice s=(Λ(A))1s=(\Lambda(A))^{-1}. Put μ=s1=Λ(A)\mu=s^{-1}=\Lambda(A) and ρ=sλ=λΛ(A)\rho=s\lambda=\frac{\lambda}{\Lambda(A)}. In summary, we obtain

(40) |μ1c[v]vV2|(1ρ)vV2.\displaystyle|\mu^{-1}c[v]-\|v\|_{V}^{2}|\leq(1-\rho)\|v\|_{V}^{2}.

The same calculation but applied to a[v]a[v] instead of a[v]12Bv2(U,H)2a[v]-\frac{1}{2}\|Bv\|_{\mathcal{L}_{2}(U,H)}^{2} in conjunction with Lemma 5.3 gives the following.

Corollary 5.5.

Suppose that Assumption 2.1 holds with M=0M=0 and that Assumption 5.1 holds. For the same μ\mu and ρ\rho as in Lemma 5.4 one has μ1AA0(V,V)1ρ\|\mu^{-1}A-A_{0}\|_{\mathcal{L}(V,V^{*})}\leq 1-\rho.

We use μ\mu and ρ\rho from the lemma to define complex perturbations of AA and BB.

Definition 5.6.

With μ\mu and ρ\rho from Lemma 5.4, fix numbers 0<r<1<R0<r<1<R satisfying

  1. (i)

    R(1ρ)<1R(1-\rho)<1, and

  2. (ii)

    rmin(Cp(1ρ),cΛ(B))<1r\min(C_{p}(1-\rho),c\Lambda(B))<1, where CpC_{p} is the constant from Lemma 4.4 and cc is the constant from Theorem 4.6.

Moreover, with the holomorphic function F(z)=rezlog(R/r)F(z)=r\mathrm{e}^{z\log(R/r)} defined on the open unit strip S={z:0<Re(z)<1}S=\{z\in\mathbb{C}:0<\operatorname{Re}(z)<1\}, define

(41) Azμ(F(z)(μ1AA0)+A0)&BzF(z)12B.\displaystyle A_{z}\coloneqq\mu\bigl{(}F(z)(\mu^{-1}A-A_{0})+A_{0}\bigr{)}\quad\&\quad B_{z}\coloneqq F(z)^{\frac{1}{2}}B.
Lemma 5.7.

Suppose that Assumption 2.1 holds with M=0M=0 and that Assumption 5.1 holds. Then the coercivity condition (3) in Assumption 2.1 is satisfied for AzA_{z} and BzB_{z}, uniformly in zS¯z\in\overline{S}. To be more precise, the implied constant depends only on λ\lambda, Λ(A)\Lambda(A), and RR.

Proof.

Write 22(U,H)\|\cdot\|_{\mathcal{L}_{2}}\coloneqq\|\cdot\|_{\mathcal{L}_{2}(U,H)}. Recall μ=Λ(A)\mu=\Lambda(A) from Lemma 5.4. Consequently, μ1A(V,V)1\|\mu^{-1}A\|_{\mathcal{L}(V,V^{*})}\leq 1. Since aa is hermitian by Assumption 5.1, it follows that μ1a[v]vV2\mu^{-1}a[v]-\|v\|_{V}^{2} is real and non-positive. Recall

μ1Az=F(z)(μ1AA0)+A0.\mu^{-1}A_{z}=F(z)(\mu^{-1}A-A_{0})+A_{0}.

Calculate for vV=1\|v\|_{V}=1 that

(42) Reμ1(az[v]12Bz(v)22)\displaystyle\operatorname{Re}\mu^{-1}\bigl{(}a_{z}[v]-\frac{1}{2}\|B_{z}(v)\|_{\mathcal{L}_{2}}^{2}\bigr{)} =1+(ReF(z))(μ1a[v]vV2)|F(z)|2μB(v)22\displaystyle={}1+\bigl{(}\operatorname{Re}F(z)\bigr{)}(\mu^{-1}a[v]-\|v\|_{V}^{2})-\frac{|F(z)|}{2\mu}\|B(v)\|_{\mathcal{L}_{2}}^{2}
(43) 1+|F(z)|(μ1a[v]vV2)|F(z)|2μB(v)22\displaystyle\geq{}1+|F(z)|(\mu^{-1}a[v]-\|v\|_{V}^{2})-\frac{|F(z)|}{2\mu}\|B(v)\|_{\mathcal{L}_{2}}^{2}
(44) =1+|F(z)|(μ1(a[v]12B(v)22)vV2)\displaystyle={}1+|F(z)|\Bigl{(}\mu^{-1}\bigl{(}a[v]-\frac{1}{2}\|B(v)\|_{\mathcal{L}_{2}}^{2}\bigr{)}-\|v\|_{V}^{2}\Bigr{)}
(45) 1|F(z)||μ1(a[v]12B(v)22)vV2|,\displaystyle\geq{}1-|F(z)|\Bigl{|}\mu^{-1}\bigl{(}a[v]-\frac{1}{2}\|B(v)\|_{\mathcal{L}_{2}}^{2}\bigr{)}-\|v\|_{V}^{2}\Bigr{|},

where we exploited in the first line that μ1a[v]vV2\mu^{-1}a[v]-\|v\|_{V}^{2} is real and in the second line that it is non-positive. Therefore, Lemma 5.4 reveals

(46) Reμ1(az[v]12Bz(v)22)1|F(z)|(1ρ),\displaystyle\operatorname{Re}\mu^{-1}\bigl{(}a_{z}[v]-\frac{1}{2}\|B_{z}(v)\|_{\mathcal{L}_{2}}^{2}\bigr{)}\geq 1-|F(z)|(1-\rho),

so that the right-hand side is strictly positive in virtue of Definition 5.6 (i) and since |F(z)|relog(R/r)=R|F(z)|\leq r\mathrm{e}^{\log(R/r)}=R. Eventually, multiplying by μ\mu completes the proof. ∎

Now consider on the unit strip SS the complex family of problems

(PzP_{z}) du(t)+Az(t)u(t)dt=f(t)dt+(Bz(t)u(t)+g(t))dW(t)\displaystyle\,\mathrm{d}u(t)+A_{z}(t)u(t)\,\mathrm{d}t=f(t)\,\mathrm{d}t+(B_{z}(t)u(t)+g(t))\,\mathrm{d}W(t)

with initial condition u(0)=0u(0)=0.

By construction, the mappings zAzz\mapsto A_{z} and zBzz\mapsto B_{z} are analytic, and Lemma 5.7 assures that Assumption 2.1 is satisfied uniformly in zz. We conclude with Proposition 3.3 that zuzE1z\mapsto u_{z}\in\mathrm{E}_{1}, where uzu_{z} is the unique solution of (PzP_{z}), is analytic.

5.2. Conclusion by Stein interpolation

The following proposition is the basis for our main result.

Proposition 5.8.

Under Assumptions 2.1 with M=0M=0 and 5.1 there exists a p0>2p_{0}>2 such that, for all p[2,p0]p\in[2,p_{0}], fE0pf\in\mathrm{E}_{0}^{p} and g𝔼12pg\in{\mathbb{E}^{p}_{\frac{1}{2}}}, problem (P) with u0=0u_{0}=0 admits a unique solution uE1pu\in\mathrm{E}_{1}^{p}, and there is a constant CC depending only on Λ(A)\Lambda(A), Λ(B)\Lambda(B), λ\lambda, T1T\vee 1 such that

(47) uE1pC(fE0p+g𝔼12p).\displaystyle\|u\|_{\mathrm{E}_{1}^{p}}\leq C\bigl{(}\|f\|_{\mathrm{E}_{0}^{p}}+\|g\|_{{\mathbb{E}^{p}_{\frac{1}{2}}}}\bigr{)}.
Proof.

Step 1: Applying abstract Stein interpolation to (PzP_{z}).

For zS¯z\in\overline{S} define an operator Tz:E0×𝔼12E1T_{z}\colon\mathrm{E}_{0}\times{\mathbb{E}_{\frac{1}{2}}}\to\mathrm{E}_{1} which maps a data pair (f,g)(f,g) to the unique solution uu of (PzP_{z}). By linearity of the equation, TzT_{z} is linear and the family of operator (Tz)zS¯(T_{z})_{z\in\overline{S}} is uniformly bounded according to Proposition 3.2.

Fix some q>2q>2. We apply the results of the last section with pp replaced by qq. Let (f,g)E0q×𝔼12qE0×𝔼12(f,g)\in\mathrm{E}^{q}_{0}\times{\mathbb{E}^{q}_{\frac{1}{2}}}\subseteq\mathrm{E}_{0}\times{\mathbb{E}_{\frac{1}{2}}}. We claim that Tit(f,g)E1qT_{\mathrm{i}t}(f,g)\in\mathrm{E}^{q}_{1} uniformly. Indeed, this follows from Theorem 4.6 if we check its smallness assumptions. Hypothesis (ii) follows immediately from Definition 5.6. Moreover, since A0μ1Az=F(z)(A0μ1A)A_{0}-\mu^{-1}A_{z}=F(z)(A_{0}-\mu^{-1}A), it follows from Corollary 5.5 that

(48) CqA0μ1Ait(V,V)=Cq|F(it)|A0μ1A(V,V)rCq(1ρ).\displaystyle C_{q}\|A_{0}-\mu^{-1}A_{\mathrm{i}t}\|_{\mathcal{L}(V,V^{*})}=C_{q}|F(\mathrm{i}t)|\|A_{0}-\mu^{-1}A\|_{\mathcal{L}(V,V^{*})}\leq rC_{q}(1-\rho).

Therefore, by choice of rr in Definition 5.6, CqA0μ1Ait(V,V)<1C_{q}\|A_{0}-\mu^{-1}A_{\mathrm{i}t}\|_{\mathcal{L}(V,V^{*})}<1 uniformly in z=itz=\mathrm{i}t.

Finally, we have argued in Section 3.2 that for fixed (f,g)(f,g) the map SzTz(f,g)E1S\ni z\mapsto T_{z}(f,g)\in\mathrm{E}_{1} is holomorphic. In summary, this allows us to invoke abstract Stein interpolation [33] to deduce

(49) uθE1qθCθ(fE0qθ+g𝔼12qθ),θ(0,1),fE0q,g𝔼12q,\displaystyle\|u_{\theta}\|_{\mathrm{E}_{1}^{q_{\theta}}}\leq C_{\theta}\bigl{(}\|f\|_{\mathrm{E}_{0}^{q_{\theta}}}+\|g\|_{\mathbb{E}^{q_{\theta}}_{\frac{1}{2}}}\bigr{)},\quad\theta\in(0,1),f\in\mathrm{E}^{q}_{0},g\in{\mathbb{E}^{q}_{\frac{1}{2}}},

where 1qθ=1θ2+θq\frac{1}{q_{\theta}}=\frac{1-\theta}{2}+\frac{\theta}{q} and Cθ=CL1θ(c/δ)θC_{\theta}=C_{L}^{1-\theta}(c/\delta)^{\theta}, uθu_{\theta} is the unique solution to (PzP_{z}) with z=θz=\theta, and where we use the constants from Proposition 3.2 and Theorem 4.6.

Step 2: Specializing to (P).

Choose θ=log(r)log(R/r)(0,1)\theta=-\frac{\log(r)}{\log(R/r)}\in(0,1). Then F(θ)=1F(\theta)=1 and we find Aθ=AA_{\theta}=A and Bθ=BB_{\theta}=B. Write p0qθp_{0}\coloneqq q_{\theta}, then Step 1 shows

(50) uE1p0C(fE0p0+g𝔼12p0),fE0q,g𝔼12q,\displaystyle\|u\|_{\mathrm{E}_{1}^{p_{0}}}\leq C\bigl{(}\|f\|_{\mathrm{E}_{0}^{p_{0}}}+\|g\|_{\mathbb{E}^{p_{0}}_{\frac{1}{2}}}\bigr{)},\quad f\in\mathrm{E}^{q}_{0},g\in{\mathbb{E}^{q}_{\frac{1}{2}}},

where 2<p0<q2<p_{0}<q and uu is the unique solution to (P) with data (f,g)(f,g). By an approximation argument, we extend well-posedness to (f,g)E0p0×𝔼12p0(f,g)\in\mathrm{E}_{0}^{p_{0}}\times\mathbb{E}^{p_{0}}_{\frac{1}{2}}.

Now to obtain the statement for p(2,p0]p\in(2,p_{0}], one can either use complex interpolation with the case p=2p=2, or lower the value of rr in the above proof. ∎

Now we can prove the main regularity result for linear equations stated in Theorem 2.2.

Proof of Theorem 2.2.

By considering eλtu(t)\mathrm{e}^{\lambda t}u(t) for suitable λ\lambda\in\mathbb{R} we can reduce to the case M=0M=0 in Assumption 2.1. Then Proposition 5.8 yields a unique strong solution to (P) with u0=0u_{0}=0 that satisfies the Lp\mathrm{L}^{p}-estimate corresponding to θ=0\theta=0 in the theorem. Still with u0=0u_{0}=0, we can first deduce higher regularity in time from [1, Prop. 3.8], using the reference operators B~=0\widetilde{B}=0 and A~=A0\widetilde{A}=A_{0}. In a second step, we include initial values u0L0p(Ω;(H,V)12/p,p)u_{0}\in\mathrm{L}^{p}_{\mathcal{F}_{0}}(\Omega;(H,V)_{1-\nicefrac{{2}}{{p}},p}) in virtue of [1, Prop. 3.10]. At both stages, [1, Prop. 2.10] can be used to get the maximal estimate missing in the case θ=0\theta=0. ∎

6. Quasilinear problem and well-posedness result

On an open and bounded set DdD\subseteq\mathbb{R}^{d} consider the quasilinear system

(QL) {duα=[i(aijαβ(u)juβ)+iΦiα(u)+ϕα(u)]dt+n1[bn,jαβ(uβ)juβ+gnα(u)]dwn, on D,uα=0, on D,uα(0)=u0α, on D,\displaystyle\left\{\quad\begin{aligned} \,\mathrm{d}u^{\alpha}&=\big{[}\partial_{i}(a^{\alpha\beta}_{ij}(u)\partial_{j}u^{\beta})+\partial_{i}\Phi^{\alpha}_{i}(u)+\phi^{\alpha}(u)\big{]}\,\mathrm{d}t\\ &\qquad\qquad+\sum_{n\geq 1}\big{[}b^{\alpha\beta}_{n,j}(u^{\beta})\partial_{j}u^{\beta}+g^{\alpha}_{n}(u)\big{]}\,\mathrm{d}w_{n},\ \ &\text{ on }D,\\ u^{\alpha}&=0,\ \ &\text{ on }\partial D,\\ u^{\alpha}(0)&=u^{\alpha}_{0},\ \ &\text{ on }D,\end{aligned}\right.

where W=(wn(t):t0)n1W=(w_{n}(t)\,:\,t\geq 0)_{n\geq 1} are independent standard Brownian motions on a probability space Ω\Omega, i,j=1,,di,j=1,\dots,d, and α,β=1,,N\alpha,\beta=1,\dots,N for some system size N1N\geq 1. For simplicity, we employed Einstein’s summation convention above and in the sequel. We will ignore the system size NN in the notation for function spaces if this cannot cause any confusion. Let \mathcal{F} be the filtration on Ω\Omega induced by WW.

Let h>1h>1. For our main result with growing non-linearities, Theorem 6.6, we introduce the following assumption. As an intermediate step, we also treat “the case h=1h=1 in Theorem 7.2, with reduced structural assumptions, and thus of independent interest.

Assumption 6.1 (𝒉h).
  1. (i)

    The system is diagonal, that is aijαβa_{ij}^{\alpha\beta} and bjαβb_{j}^{\alpha\beta} are non-zero only when α=β\alpha=\beta. Write aijαaijααa_{ij}^{\alpha}\coloneqq a_{ij}^{\alpha\alpha} and bjαbjααb_{j}^{\alpha}\coloneqq b_{j}^{\alpha\alpha}. For fixed α\alpha, the matrix aαa^{\alpha} does not need to be diagonal.

  2. (ii)

    The coefficients aijα:(0,T)×D×Na^{\alpha}_{ij}\colon(0,T)\times D\times\mathbb{R}^{N}\to\mathbb{R} and bjα(bn,jα)n1:(0,T)×D×2b^{\alpha}_{j}\coloneqq(b^{\alpha}_{n,j})_{n\geq 1}\colon(0,T)\times D\times\mathbb{R}\to\ell^{2} are continuous in the last component, measurable, and uniformly bounded by a constant Λ\Lambda. Moreover, supt,x,y|jbn,jαβ(t,x,y)|<\sup_{t,x,y}|\partial_{j}b_{n,j}^{\alpha\beta}(t,x,y)|<\infty.

  3. (iii)

    One has aijα=ajiαa^{\alpha}_{ij}=a^{\alpha}_{ji} and there exists λ>0\lambda>0 such that, for all t(0,T)t\in(0,T), xDx\in D, and yNy\in\mathbb{R}^{N},

    (aijα(t,x,y)12bn,iα(t,x,yα)bn,jα(t,x,yα))ξiαξjαλ|ξ|2,ξdN.\big{(}a^{\alpha}_{ij}(t,x,y)-\frac{1}{2}b^{\alpha}_{n,i}(t,x,y^{\alpha})b^{\alpha}_{n,j}(t,x,y^{\alpha})\big{)}\xi_{i}^{\alpha}\xi_{j}^{\alpha}\geq\lambda|\xi|^{2},\quad\xi\in\mathbb{R}^{dN}.
  4. (iv)

    The non-linearities Φiα:(0,T)×D×N\Phi^{\alpha}_{i}\colon(0,T)\times D\times\mathbb{R}^{N}\to\mathbb{R} and ϕα:(0,T)×D×N\phi^{\alpha}\colon(0,T)\times D\times\mathbb{R}^{N}\to\mathbb{R} are measurable, locally Lipschitz in the last component, and satisfy the following growth condition: there exists a constant CC such that,

    (G) |Φ(t,x,y)|+|ϕ(t,x,y)|\displaystyle|\Phi(t,x,y)|+|\phi(t,x,y)| C(1+|y|h),t(0,T),xD,yN.\displaystyle\leq C(1+|y|^{h}),\quad t\in(0,T),x\in D,y\in\mathbb{R}^{N}.
  5. (v)

    The non-linearity ϕ\phi is dissipative in the following sense: there exists a constant CC such that one has for all t(0,T)t\in(0,T), xDx\in D, yNy\in\mathbb{R}^{N} and α=1,,N\alpha=1,\dots,N that

    (51) ϕα(t,x,y)yαC(|y|2+1).\displaystyle\phi^{\alpha}(t,x,y)y^{\alpha}\leq C(|y|^{2}+1).

    The nonlinearity Φ\Phi is dissipative in the following sense:

    Φα(t,x,y)=Φ¯α(yα)+Φ^α(t,x,y),\Phi^{\alpha}(t,x,y)=\overline{\Phi}^{\alpha}(y^{\alpha})+\widehat{\Phi}^{\alpha}(t,x,y),

    where Φ¯i:N\overline{\Phi}_{i}:\mathbb{R}\to\mathbb{R}^{N} is measurable, and Φ^i:(0,T)×D×NN\widehat{\Phi}_{i}:(0,T)\times D\times\mathbb{R}^{N}\to\mathbb{R}^{N} is measurable and Lipschitz continuous in the last component (uniformly with respect to tt and xx).

  6. (vi)

    The stochastic non-linearity gα(gnα)n1:(0,T)×D×N2g^{\alpha}\coloneqq(g^{\alpha}_{n})_{n\geq 1}\colon(0,T)\times D\times\mathbb{R}^{N}\to\ell^{2} is measurable, Lipschitz continuous and of linear growth in the last component (uniformly with respect to the first two components).

Remark 6.2.

Some discussion of the hypotheses is in order.

  1. (i)

    When we work with non-linearities of linear growth, see Section 7, then the diagonal structure assumed in (i) is not needed. This is why (QL) is formulated for the non-diagonal case.

  2. (ii)

    The non-linearities for aa, Φ\Phi and ϕ\phi lead to a coupling between the components even when working with a diagonal structure.

  3. (iii)

    To identify the limit in the stochastic compactness method, it seems crucial that bjαβb_{j}^{\alpha\beta} only depends on the β\betath component. In fact, it is a surprising strength of our approach that this structural assumption is not needed for aa.

Remark 6.3 (Diagonal structure).

Instead of imposing a diagonal structure, we could also strengthen the coercivity condition as follows: for q[0,2(h1)+ε]q\in[0,2(h-1)+\varepsilon] assume that there exists λq>0\lambda_{q}>0 such that, for all t(0,T)t\in(0,T), xDx\in D, and yNy\in\mathbb{R}^{N},

|yα|q(aijα(t,x,y)12bn,iα(t,x,yα)bn,jα(t,x,yα))ξiαξjαλq|yα|q|ξα|2 for all ξdN.|y^{\alpha}|^{q}\Big{(}a^{\alpha}_{ij}(t,x,y)-\frac{1}{2}b^{\alpha}_{n,i}(t,x,y^{\alpha})b^{\alpha}_{n,j}(t,x,y^{\alpha})\Big{)}\xi_{i}^{\alpha}\xi_{j}^{\alpha}\geq\lambda_{q}|y^{\alpha}|^{q}|\xi^{\alpha}|^{2}\quad\text{ for all }\xi\in\mathbb{R}^{dN}.

For instance, such a condition can be ensured for perturbations of diagonal systems in which |aijαβ|c(1+|y|q)1|a^{\alpha\beta}_{ij}|\leq c(1+|y|^{q})^{-1} whenever αβ\alpha\neq\beta, where cc is sufficiently small. For clarity of exposition, we only work with the diagonal case. This case is also the most important one in reaction–diffusion equations. However, we do not know whether the structural condition we have made is essential. It is needed at a technical point in the proof in the identification of the limit in the stochastic compactness method.

To make precise the homogeneous Dirichlet boundary condition in (QL), we introduce the following function spaces.

Definition 6.4.

Write H01(D)\mathrm{H}^{1}_{0}(D) for the closure in H1(D)\mathrm{H}^{1}(D) of C0(D)\mathrm{C}^{\infty}_{0}(D), the smooth and compactly supported functions in DD. Moreover, for s(0,1)s\in(0,1) put

(52) H0s(D)[L2(D),H01(D)]s&B2,p,0s(D)(L2(D),H01(D))s,p.\displaystyle\mathrm{H}^{s}_{0}(D)\coloneqq[\mathrm{L}^{2}(D),\mathrm{H}^{1}_{0}(D)]_{s}\quad\&\quad\mathrm{B}^{s}_{2,p,0}(D)\coloneqq(\mathrm{L}^{2}(D),\mathrm{H}^{1}_{0}(D))_{s,p}.
Remark 6.5.

The spaces H0s(D)\mathrm{H}^{s}_{0}(D) and B2,p,0s(D)\mathrm{B}^{s}_{2,p,0}(D) can often be identified with closed subspaces of the usual Bessel potential and Besov spaces on DD.

The main result on the quasi-linear system (QL) reads as follows.

Theorem 6.6 (Weak existence).

Let h>1h>1. Suppose Assumption 6.1 holds. Then there is p0>2p_{0}>2, depending only on C,Λ,λC,\Lambda,\lambda from the assumption as well as dimensions, such that, if p(2,p0]p\in(2,p_{0}] and q(ph,)q\in(ph,\infty), then, given u0L0p(Ω;B2,p,012/p(D))L0q(Ω×D)u_{0}\in\mathrm{L}^{p}_{\mathcal{F}_{0}}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))\cap\mathrm{L}^{q}_{\mathcal{F}_{0}}(\Omega\times D), there exists a weak solution (u~,W~,Ω~,~,~,(~t)t0)(\widetilde{u},\widetilde{W},\widetilde{\Omega},\widetilde{\mathcal{F}},\widetilde{\mathbb{P}},(\widetilde{\mathcal{F}}_{t})_{t\geq 0}) of (QL) in the sense of Definition 6.7. One has the estimates

(53) u~Lp(Ω~;C([0,T];B2,p,012/p(D)))\displaystyle\|\widetilde{u}\|_{\mathrm{L}^{p}(\widetilde{\Omega};\mathrm{C}([0,T];\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D)))} θ1+u0Lq(Ω;Lq(D))h+u0Lp(Ω;B2,p,012/p(D)),\displaystyle\lesssim_{\theta}1+\|u_{0}\|_{\mathrm{L}^{q}(\Omega;\mathrm{L}^{q}(D))}^{h}+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))},
(54) u~Lq(Ω~×(0,T)×D)\displaystyle\|\widetilde{u}\|_{\mathrm{L}^{q}(\widetilde{\Omega}\times(0,T)\times D)} 1+u0Lq(Ω×D)h.\displaystyle\lesssim 1+\|u_{0}\|_{\mathrm{L}^{q}(\Omega\times D)}^{h}.
Moreover, for all θ[0,1/2)\theta\in[0,1/2) one has the bound
(55) u~Lp(Ω~;Hθ,p(0,T;H012θ(D)))\displaystyle\|\widetilde{u}\|_{\mathrm{L}^{p}(\widetilde{\Omega};\mathrm{H}^{\theta,p}(0,T;\mathrm{H}^{1-2\theta}_{0}(D)))} θ1+u0Lq(Ω;Lq(D))h+u0Lp(Ω;B2,p,012/p(D)).\displaystyle\lesssim_{\theta}1+\|u_{0}\|_{\mathrm{L}^{q}(\Omega;\mathrm{L}^{q}(D))}^{h}+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))}.

Our notion of a weak solution employed in Theorem 6.6 is the following.

Definition 6.7.

Call the tuple (u~,W~,Ω~,~,~,(~t)t0)(\widetilde{u},\widetilde{W},\widetilde{\Omega},\widetilde{\mathcal{F}},\widetilde{\mathbb{P}},(\widetilde{\mathcal{F}}_{t})_{t\geq 0}) a weak solution of (QL) if W~=(w~n)n1\widetilde{W}=(\widetilde{w}_{n})_{n\geq 1} is an 2\ell^{2}-cylindrical Brownian motion on Ω~\widetilde{\Omega}, u~(0)\widetilde{u}(0) has the same distribution as u0u_{0}, u~L2(0,T;H01(D))C([0,T];L2(D))\widetilde{u}\in\mathrm{L}^{2}(0,T;\mathrm{H}^{1}_{0}(D))\cap\mathrm{C}([0,T];\mathrm{L}^{2}(D)) almost surely and is (~t)(\widetilde{\mathcal{F}}_{t})-progressively measurable, Φiα(u~),ϕα(u~)L1(0,T;L1(D))\Phi^{\alpha}_{i}(\widetilde{u}),\phi^{\alpha}(\widetilde{u})\in\mathrm{L}^{1}(0,T;\mathrm{L}^{1}(D)), g(u~)L2(0,T;L2(D;2))g(\widetilde{u})\in\mathrm{L}^{2}(0,T;\mathrm{L}^{2}(D;\ell^{2})), and for all ξC0(D)\xi\in\mathrm{C}^{\infty}_{0}(D) and t[0,T]t\in[0,T], one has almost surely

(56) u~α(t),ξ=u~α(0),ξ\displaystyle\langle\widetilde{u}^{\alpha}(t),\xi\rangle=\langle\widetilde{u}^{\alpha}(0),\xi\rangle +0taijαβ(u~)ju~,iξΦiα(u~),iξ+ϕα(u~),ξdt\displaystyle+\int_{0}^{t}-\langle a^{\alpha\beta}_{ij}(\widetilde{u})\partial_{j}\widetilde{u},\partial_{i}\xi\rangle-\langle\Phi^{\alpha}_{i}(\widetilde{u}),\partial_{i}\xi\rangle+\langle\phi^{\alpha}(\widetilde{u}),\xi\rangle\,\mathrm{d}t
(57) +n10tbn,jαβ(u~)ju~,ξ+gnα(u~),ξdwn.\displaystyle+\sum_{n\geq 1}\int_{0}^{t}\langle b^{\alpha\beta}_{n,j}(\widetilde{u})\partial_{j}\widetilde{u},\xi\rangle+\langle g^{\alpha}_{n}(\widetilde{u}),\xi\rangle\,\mathrm{d}w_{n}.

The proof of Theorem 6.6 will occupy the rest of this article.

Remark 6.8.

In the scalar case N=1N=1 with b=0b=0 and ϕ=0\phi=0 a result such as Theorem 6.6 is proved in [12] in the case of periodic boundary conditions. Moreover, even the degenerate case and initial data in L1\mathrm{L}^{1} are considered via a different solution concept called kinetic solutions. In this framework uniqueness is proved as well. It would be interesting to see if some of these results can be obtained in the generality of Theorem 6.6.

In the scalar case N=1N=1 one can also make a comparison to [30, Example 4.1] in case the parameter α\alpha used there satisfies α=2\alpha=2. The main differences in this special case are that, in Theorem 6.6, we allow systems. Moreover, our condition on the gradient noise is more flexible, and for N=1N=1 it coincides with the classical stochastic parabolicity condition (cf. [2, Assumption 5.9(2)]). To compare the condition on the gradient noise more precisely, note that the condition in [30, Theorem 3.2] reduces to χLB<2LA\chi L_{B}<2L_{A}, where χh+1\chi\geq h+1, where h1h\geq 1 is the growth of our nonlinearity. Moreover, LAL_{A} and LBL_{B} are uniform constants over the coefficients. Our condition in Assumption 6.1(iii) is pointwise in xx and yy instead of uniform. Moreover, the growth of ϕ\phi and Φ\Phi does not influence this coercivity condition in any way.

7. Quasilinear problem with Lipschitz non-linearities

First, we study (QL) under the following set of hypotheses. In Section 8 we are going to reduce the general case to the current one in virtue of an approximation procedure.

Assumption 7.1.
  1. (i)

    The coefficient functions aijαβ:(0,T)×D×Na^{\alpha\beta}_{ij}\colon(0,T)\times D\times\mathbb{R}^{N}\to\mathbb{R} and bjαβ(bn,jαβ)n1:(0,T)×D×2b^{\alpha\beta}_{j}\coloneqq(b^{\alpha\beta}_{n,j})_{n\geq 1}\colon(0,T)\times D\times\mathbb{R}\to\ell^{2} are continuous in the last component, measurable, and uniformly bounded by a constant Λ\Lambda. Moreover, supt,x,y|jbn,jαβ(t,x,y)|<\sup_{t,x,y}|\partial_{j}b_{n,j}^{\alpha\beta}(t,x,y)|<\infty.

  2. (ii)

    One has aijαβ=ajiβαa^{\alpha\beta}_{ij}=a^{\beta\alpha}_{ji} and there exists λ>0\lambda>0 such that, for all t(0,T)t\in(0,T), xDx\in D, and yNy\in\mathbb{R}^{N},

    (aijαβ(t,x,y)12bn,iγα(t,x,yα)bn,jγβ(t,x,yβ))ξiαξjβλ|ξ|2 for all ξdN.\big{(}a^{\alpha\beta}_{ij}(t,x,y)-\frac{1}{2}b^{\gamma\alpha}_{n,i}(t,x,y^{\alpha})b^{\gamma\beta}_{n,j}(t,x,y^{\beta})\big{)}\xi_{i}^{\alpha}\xi_{j}^{\beta}\geq\lambda|\xi|^{2}\quad\text{ for all }\xi\in\mathbb{R}^{dN}.
  3. (iii)

    The non-linearities Φiα:(0,T)×D×N\Phi^{\alpha}_{i}\colon(0,T)\times D\times\mathbb{R}^{N}\to\mathbb{R} and ϕα:(0,T)×D×N\phi^{\alpha}\colon(0,T)\times D\times\mathbb{R}^{N}\to\mathbb{R}, and the stochastic non-linearity gα(gnα)n1:(0,T)×D×N2g^{\alpha}\coloneqq(g^{\alpha}_{n})_{n\geq 1}\colon(0,T)\times D\times\mathbb{R}^{N}\to\ell^{2} are measurable, Lipschitz continuous and of linear growth in the last component (uniformly with respect to the first two components).

Observe that Assumption 7.1 (iii) is the combination of Assumption 6.1 (iv) with h=1h=1 and Assumption 6.1 (vi). Assumption 6.1 (v) becomes obsolete.

We are going to show the following well-posedness result.

Theorem 7.2.

Suppose Assumption 7.1 holds. There is p0>2p_{0}>2, depending only on C,Λ,λC,\Lambda,\lambda from the assumption as well as dimensions, such that if p(2,p0]p\in(2,p_{0}], then, given u0L0p(Ω;B2,p,012/p(D))u_{0}\in\mathrm{L}^{p}_{\mathcal{F}_{0}}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D)), there exists a weak solution (u~,W~,Ω~,~,~,(~t)t0)(\widetilde{u},\widetilde{W},\widetilde{\Omega},\widetilde{\mathcal{F}},\widetilde{\mathbb{P}},(\widetilde{\mathcal{F}}_{t})_{t\geq 0}) of (QL). Moreover, for all θ[0,1/2)\theta\in[0,1/2) one has

(58) u~Lp(Ω~;C([0,T];B2,p,012/p(D)))+u~Lp(Ω~;Hθ,p(0,T;H012θ(D)))θ1+u0Lp(Ω;B2,p,012/p(D)).\displaystyle\|\widetilde{u}\|_{\mathrm{L}^{p}(\widetilde{\Omega};\mathrm{C}([0,T];\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D)))}+\|\widetilde{u}\|_{\mathrm{L}^{p}(\widetilde{\Omega};\mathrm{H}^{\theta,p}(0,T;\mathrm{H}^{1-2\theta}_{0}(D)))}\lesssim_{\theta}1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))}.

7.1. Smoothing of the coefficients

Let ζ:[0,)\zeta\colon\mathbb{R}\to[0,\infty) be a smooth and compactly supported function with integral equal to 11 and ρ:N[0,)\rho\colon\mathbb{R}^{N}\to[0,\infty) the NN-fold product of ζ\zeta. For m1m\geq 1, write ζm\zeta_{m} and ρm\rho_{m} for the induced mollifier sequences. Now, define the quasi-linear operators

(59) Am(t):L2(D)N\displaystyle{}_{m}A(t)\colon\mathrm{L}^{2}(D)^{N} (H01(D)N,H1(D)N),\displaystyle\to\mathcal{L}(\mathrm{H}^{1}_{0}(D)^{N},\mathrm{H}^{-1}(D)^{N}),
(60) Bm(t):L2(D)\displaystyle{}_{m}B(t)\colon\mathrm{L}^{2}(D) (H01(D)N,2(2,L2(D)N))\displaystyle\to\mathcal{L}(\mathrm{H}^{1}_{0}(D)^{N},\mathcal{L}_{2}(\ell^{2},\mathrm{L}^{2}(D)^{N}))

by

(61) Am(t,v)u\displaystyle{}_{m}A(t,v)u =(i(aijαβm(t,x,Rmv)juβ))α=1N,(uH01(D)N,vL2(D)N),\displaystyle=\bigl{(}\partial_{i}({}_{m}a_{ij}^{\alpha\beta}(t,x,R_{m}v)\partial_{j}u^{\beta})\bigr{)}_{\alpha=1}^{N},\qquad(u\in\mathrm{H}^{1}_{0}(D)^{N},v\in\mathrm{L}^{2}(D)^{N}),
(62) Bm(t,w)u,en\displaystyle\langle{}_{m}B(t,w)u,e_{n}\rangle =(bn,jαβm(t,x,Rmwβ)juβ)α=1N,(uH01(D)N,wL2(D)),\displaystyle=\bigl{(}{}_{m}b_{n,j}^{\alpha\beta}(t,x,R_{m}w^{\beta})\partial_{j}u^{\beta}\bigr{)}_{\alpha=1}^{N},\qquad(u\in\mathrm{H}^{1}_{0}(D)^{N},w\in\mathrm{L}^{2}(D)),

where Rmf=ρmEfR_{m}f=\rho_{m}\ast Ef, aijαβm(t,x,)=ρmaijαβ(t,x,){}_{m}a_{ij}^{\alpha\beta}(t,x,\cdot)=\rho_{m}\ast a_{ij}^{\alpha\beta}(t,x,\cdot) and bn,jαβm(t,x,)=ζmbn,jαβ(t,x,){}_{m}b_{n,j}^{\alpha\beta}(t,x,\cdot)=\zeta_{m}\ast b_{n,j}^{\alpha\beta}(t,x,\cdot), where the convolution is taken with respect to the third variable. Here, EE denotes the zero extension of a given function, and we apply the convolution in the definition of RmR_{m} component-wise if ff is vector-valued.

Consider now the family of equations

(m-QL) {du=[Am(u)u+iΦi(u)+ϕ(u)]dt+n1[Bm(u)u,en+gn(u)]dwn, on D,u=0, on D,u(0)=u0, on D.\displaystyle\left\{\quad\begin{aligned} \,\mathrm{d}u&=\big{[}{}_{m}A(u)u+\partial_{i}\Phi_{i}(u)+\phi(u)\big{]}\,\mathrm{d}t\\ &\qquad\qquad+\sum_{n\geq 1}\big{[}\langle{}_{m}B(u)u,e_{n}\rangle+g_{n}(u)\big{]}\,\mathrm{d}w_{n},\ \ &\text{ on }D,\\ u&=0,\ \ &\text{ on }\partial D,\\ u(0)&=u_{0},\ \ &\text{ on }D.\end{aligned}\right.

Following [20, p. 211] we show that the smoothed coefficients satisfy the same coercivity condition as the original ones.

Lemma 7.3.

For fixed mm, the smoothed coefficients aijαβm{}_{m}a_{ij}^{\alpha\beta} and bn,jαβm{}_{m}b_{n,j}^{\alpha\beta} satisfy Assumptions 7.1 (i) and  (ii) with the same constants.

Proof.

Fix mm. By definition of the convolution, symmetry of aijαβm{}_{m}a^{\alpha\beta}_{ij} is immediate. The same holds for uniform boundedness, since the convolution kernel has normalized L1\mathrm{L}^{1}-norm.

For better readability, suppress dependence on tt and xx of the coefficients. Introduce the vector notation bn,jα(y)(bn,jαβ(yβ))β=1N\vec{b}_{n,j}^{\alpha}(y)\coloneqq\bigl{(}b_{n,j}^{\alpha\beta}(y^{\beta})\bigr{)}_{\beta=1}^{N}, likewise for bn,jαm(y){}_{m}\vec{b}_{n,j}^{\alpha}(y). In this way, βbn,jαβ(yβ)ξjβ=bn,jα(y)ξj\sum_{\beta}b^{\alpha\beta}_{n,j}(y^{\beta})\xi_{j}^{\beta}=\vec{b}_{n,j}^{\alpha}(y)\cdot\xi_{j} and (bn,jαβm)β=1N=ρmbn,jα({}_{m}b_{n,j}^{\alpha\beta})_{\beta=1}^{N}=\rho_{m}\ast\vec{b}_{n,j}^{\alpha}, where again we only convolve in the last coordinate.

Observe first that for yNy\in\mathbb{R}^{N},

(63) ni,jα,β,γbn,iγαm(yα)bn,jγβm(yβ)ξiαξjβ=nγ|jbjγm(y)ξj|22,\displaystyle\sum_{n}\sum_{i,j}\sum_{\alpha,\beta,\gamma}{}_{m}b^{\gamma\alpha}_{n,i}(y^{\alpha}){}_{m}b^{\gamma\beta}_{n,j}(y^{\beta})\xi_{i}^{\alpha}\xi_{j}^{\beta}=\sum_{n}\sum_{\gamma}\Bigl{|}\sum_{j}{}_{m}\vec{b}^{\gamma}_{j}(y)\cdot\xi_{j}\Bigr{|}_{2}^{2},

where ||2|\cdot|_{2} is the Euclidean norm on N\mathbb{R}^{N}. Write 𝒛:Nzz\bm{z}\colon\mathbb{R}^{N}\ni z\mapsto z for the identity map. Using that the convolution kernel is positive with normalized integral, Jensen’s inequality gives

(64) nγ|jbjγm(y)ξj|22\displaystyle\sum_{n}\sum_{\gamma}\Bigl{|}\sum_{j}{}_{m}\vec{b}^{\gamma}_{j}(y)\cdot\xi_{j}\Bigr{|}_{2}^{2} nγ|jbjγ(𝒛)ξj|22ρm(y)\displaystyle\leq\sum_{n}\sum_{\gamma}\Bigl{|}\sum_{j}\vec{b}^{\gamma}_{j}(\bm{z})\cdot\xi_{j}\Bigr{|}_{2}^{2}\ast\rho_{m}(y)
(65) =ni,jα,β,γ(bn,iγα(𝒛α)bn,jγβ(𝒛β)ξiαξjβ)ρm(y).\displaystyle=\sum_{n}\sum_{i,j}\sum_{\alpha,\beta,\gamma}\Bigl{(}b^{\gamma\alpha}_{n,i}(\bm{z}^{\alpha})b^{\gamma\beta}_{n,j}(\bm{z}^{\beta})\xi_{i}^{\alpha}\xi_{j}^{\beta}\Bigr{)}\ast\rho_{m}(y).

Hence, it follows from Assumptions 7.1 (ii) and positivity of the convolution kernel that

(66) (aijαβm(y)12bn,iγαm(yα)bn,jγβm(yβ))ξiαξjβ\displaystyle\quad\Big{(}{}_{m}a^{\alpha\beta}_{ij}(y)-\frac{1}{2}{}_{m}b^{\gamma\alpha}_{n,i}(y^{\alpha}){}_{m}b^{\gamma\beta}_{n,j}(y^{\beta})\Big{)}\xi_{i}^{\alpha}\xi_{j}^{\beta}
(67) N(aijαβ(z)12bn,iγα(zα)bn,jγβ(zβ))ξiαξjβρm(yz)dzλ|ξ|2.\displaystyle\geq\int_{\mathbb{R}^{N}}\Big{(}a^{\alpha\beta}_{ij}(z)-\frac{1}{2}b^{\gamma\alpha}_{n,i}(z^{\alpha})b^{\gamma\beta}_{n,j}(z^{\beta})\Big{)}\xi_{i}^{\alpha}\xi_{j}^{\beta}\rho_{m}(y-z)\,\mathrm{d}z\geq\lambda|\xi|^{2}.\qed

Using [2] we show that for each mm, problem (m-QL) admits a unique solution.

Lemma 7.4.

Suppose Assumption 7.1 holds. Then, for each mm, there is a unique strong solution umL2(Ω×(0,T);H01(D))L2(Ω;C([0,T];L2(D)))u_{m}\in\mathrm{L}^{2}(\Omega\times(0,T);\mathrm{H}^{1}_{0}(D))\cap\mathrm{L}^{2}(\Omega;\mathrm{C}([0,T];\mathrm{L}^{2}(D))) to (m-QL).

Proof.

Fix mm. Write 2=2(2,L2(D))\|\cdot\|_{\mathcal{L}_{2}}=\|\cdot\|_{\mathcal{L}_{2}(\ell^{2},\mathrm{L}^{2}(D))}. The assertion follows from [2, Thm. 3.5] applied with A0(v)uA0(t,v)uAm(t,v)uA_{0}(v)u\coloneqq A_{0}(t,v)u\coloneqq{}_{m}A(t,v)u and B0(v)uB0(t,v)uBm(t,v)uB_{0}(v)u\coloneqq B_{0}(t,v)u\coloneqq{}_{m}B(t,v)u, the non-linearities F(v)F(t,v)iΦi(t,v)+ϕ(t,v)F(v)\coloneqq F(t,v)\coloneqq\partial_{i}\Phi_{i}(t,v)+\phi(t,v) and G(v)enG(t,v)engn(t,v)G(v)e_{n}\coloneqq G(t,v)e_{n}\coloneqq g_{n}(t,v), and the cylindrical Brownian motion WenwnWe_{n}\coloneqq w_{n}.

Indeed, note first that, in the light of Lemma 7.3, all assertions from Assumption 7.1 remain valid for the problem (m-QL). Besides boundedness of A0A_{0} and B0B_{0}, and linear growth of FF and GG in the respective norms, which readily follow from uniform boundedness of the coefficients in the first two cases, and linear growth of the non-linearities in the last two cases, we have to check the regularity conditions

(68) A0(u)wA0(v)wH1(D)+B0(u)wB0(v)w2uvL2(D)wH01(D),\displaystyle\|A_{0}(u)w-A_{0}(v)w\|_{\mathrm{H}^{-1}(D)}+\|B_{0}(u)w-B_{0}(v)w\|_{\mathcal{L}_{2}}\lesssim\|u-v\|_{\mathrm{L}^{2}(D)}\|w\|_{\mathrm{H}^{1}_{0}(D)},

for all u,vL2(D)u,v\in\mathrm{L}^{2}(D) and wH01(D)w\in\mathrm{H}^{1}_{0}(D), and

(69) F(u)F(v)H1(D)+G(u)G(v)2uvL2(D),\displaystyle\|F(u)-F(v)\|_{\mathrm{H}^{-1}(D)}+\|G(u)-G(v)\|_{\mathcal{L}_{2}}\lesssim\|u-v\|_{\mathrm{L}^{2}(D)},

for all u,vL2(D)u,v\in\mathrm{L}^{2}(D), as well as the following coercivity condition: there exists θ,η>0\theta,\eta>0 and M0M\geq 0 such that a.s.

(70) A0(v)u,u(1/2+η)B0(v)u22θuH01(D)2MuL2(D)2,\displaystyle\langle A_{0}(v)u,u\rangle-(\nicefrac{{1}}{{2}}+\eta)\|B_{0}(v)u\|_{\mathcal{L}_{2}}^{2}\geq\theta\|u\|_{\mathrm{H}^{1}_{0}(D)}^{2}-M\|u\|_{\mathrm{L}^{2}(D)}^{2},

for all uH01(D)u\in\mathrm{H}^{1}_{0}(D) and vL2(D)v\in\mathrm{L}^{2}(D).

The second regularity condition, (69), follows immediately from Lischitz continuity of Φ\Phi, ϕ\phi, and gg. For brevity, we present the regularity condition (68) only for A0A_{0}. The calculation for B0B_{0} is similar. Fix an integer k>d/2k>\nicefrac{{d}}{{2}}. Using the Sobolev embedding and Young’s convolution inequality, calculate

(71) A0(u)wA0(v)wH1(D)\displaystyle\|A_{0}(u)w-A_{0}(v)w\|_{\mathrm{H}^{-1}(D)} aijαβm(Rmu)aijαβm(Rmv)jwβL2(D)\displaystyle\leq\|{}_{m}a^{\alpha\beta}_{ij}(R_{m}u)-{}_{m}a^{\alpha\beta}_{ij}(R_{m}v)\|_{\infty}\|\partial_{j}w^{\beta}\|_{\mathrm{L}^{2}(D)}
(72) aijαβmLipRmuRmvjwβL2(D)\displaystyle\lesssim\|{}_{m}a^{\alpha\beta}_{ij}\|_{\mathrm{Lip}}\|R_{m}u-R_{m}v\|_{\infty}\|\partial_{j}w^{\beta}\|_{\mathrm{L}^{2}(D)}
(73) maijαβRmuRmvWk,2(D)jwβL2(D)\displaystyle\lesssim_{m}\|a^{\alpha\beta}_{ij}\|_{\infty}\|R_{m}u-R_{m}v\|_{\mathrm{W}^{k,2}(D)}\|\partial_{j}w^{\beta}\|_{\mathrm{L}^{2}(D)}
(74) ρmWk,1(D)uvL2(D)wH01(D)\displaystyle\lesssim\|\rho_{m}\|_{\mathrm{W}^{k,1}(D)}\|u-v\|_{\mathrm{L}^{2}(D)}\|w\|_{\mathrm{H}^{1}_{0}(D)}
(75) muvL2(D)wH01(D).\displaystyle\lesssim_{m}\|u-v\|_{\mathrm{L}^{2}(D)}\|w\|_{\mathrm{H}^{1}_{0}(D)}.

To complete the proof, note first that Assumption 7.1 (ii) implies (70) with η=0\eta=0 and θλM\theta\coloneqq\lambda\eqqcolon M. Therefore, since B0(v):H01(D)NL2(D)NB_{0}(v)\colon\mathrm{H}^{1}_{0}(D)^{N}\to\mathrm{L}^{2}(D)^{N} is bounded (uniformly in vv), we get (70) with η>0\eta>0 if we replace θ\theta by θ/2\nicefrac{{\theta}}{{2}}. ∎

7.2. Higher integrability of semilinear equations

It will turn out useful in the sequel to consider a semi-linear version of (QL). More precisely, consider

(SL) {duα=[i(𝒂ijαβjuβ)+iΦiα(u)+ϕα(u)]dt+n1[𝒃n,jαβjuβ+gnα(u)]dwn, on D,uα=0, on D,uα(0)=u0α, on D.\displaystyle\left\{\quad\begin{aligned} \,\mathrm{d}u^{\alpha}&=\big{[}\partial_{i}(\bm{a}^{\alpha\beta}_{ij}\partial_{j}u^{\beta})+\partial_{i}\Phi^{\alpha}_{i}(u)+\phi^{\alpha}(u)\big{]}\,\mathrm{d}t\\ &\qquad\qquad+\sum_{n\geq 1}\big{[}\bm{b}^{\alpha\beta}_{n,j}\partial_{j}u^{\beta}+g^{\alpha}_{n}(u)\big{]}\,\mathrm{d}w_{n},\ \ &\text{ on }D,\\ u^{\alpha}&=0,\ \ &\text{ on }\partial D,\\ u^{\alpha}(0)&=u^{\alpha}_{0},\ \ &\text{ on }D.\end{aligned}\right.

We impose the following assumption, which is similar to Assumption 7.1.

Assumption 7.5.
  1. (i)

    The coefficients 𝒂ijαβ:Ω×(0,T)×D\bm{a}^{\alpha\beta}_{ij}\colon\Omega\times(0,T)\times D\to\mathbb{R} as well as 𝒃jαβ(𝒃n,jαβ)n1:Ω×(0,T)×D2\bm{b}^{\alpha\beta}_{j}\coloneqq(\bm{b}^{\alpha\beta}_{n,j})_{n\geq 1}\colon\Omega\times(0,T)\times D\to\ell^{2} are progressively measurable and uniformly bounded by a constant Λ\Lambda.

  2. (ii)

    One has 𝒂ijαβ=𝒂jiβα\bm{a}^{\alpha\beta}_{ij}=\bm{a}^{\beta\alpha}_{ji} and there exists λ>0\lambda>0 such that, for all t(0,T),xDt\in(0,T),x\in D, one has almost surely

    (𝒂ijαβ(t,x)12𝒃n,iγα(t,x)𝒃n,jγβ(t,x))ξiαξjβλ|ξ|2 for all ξdN.\Big{(}\bm{a}^{\alpha\beta}_{ij}(t,x)-\frac{1}{2}\bm{b}^{\gamma\alpha}_{n,i}(t,x)\bm{b}^{\gamma\beta}_{n,j}(t,x)\Big{)}\xi_{i}^{\alpha}\xi_{j}^{\beta}\geq\lambda|\xi|^{2}\quad\text{ for all }\xi\in\mathbb{R}^{dN}.
  3. (iii)

    The deterministic non-linearities Φiα:(0,T)×D×N\Phi^{\alpha}_{i}\colon(0,T)\times D\times\mathbb{R}^{N}\to\mathbb{R} and ϕα:(0,T)×D×N\phi^{\alpha}\colon(0,T)\times D\times\mathbb{R}^{N}\to\mathbb{R} and the stochastic non-linearity gα(gnα)n1:(0,T)×D×N2g^{\alpha}\coloneqq(g^{\alpha}_{n})_{n\geq 1}\colon(0,T)\times D\times\mathbb{R}^{N}\to\ell^{2} are measurable and Lipschitz continuous and of linear growth in the last variable (uniformly with respect to the first two components).

A prototypical example for Assumption 7.5 is 𝒂ijαβ(t,x)aijαβ(t,x,u(t,x))\bm{a}^{\alpha\beta}_{ij}(t,x)\coloneqq a^{\alpha\beta}_{ij}(t,x,u(t,x)) and 𝒃n,jαβ(t,x)bn,jαβ(t,x,uβ(t,x))\bm{b}^{\alpha\beta}_{n,j}(t,x)\coloneqq b^{\alpha\beta}_{n,j}(t,x,u^{\beta}(t,x)), where aijαβa^{\alpha\beta}_{ij} and bn,jαβb^{\alpha\beta}_{n,j} are coefficients for (QL) subject to Assumption 7.1 and the progressively measurable function u:Ω×(0,T)×DNu\colon\Omega\times(0,T)\times D\to\mathbb{R}^{N} is frozen.

It is well-known that (SL) admits a unique variational solution in L2(Ω×(0,T);H01(D))L2(Ω;C([0,T];L2(D)))\mathrm{L}^{2}(\Omega\times(0,T);\mathrm{H}^{1}_{0}(D))\cap\mathrm{L}^{2}(\Omega;\mathrm{C}([0,T];\mathrm{L}^{2}(D))). As a consequence of Theorem 2.2, the main result of the first part of this article, we can improve integrability in time of this variational solution to p>2p>2.

Proposition 7.6 (Higher integrability for (SL)).

Suppose Assumption 7.5 holds. Then there is some p0>2p_{0}>2 such that for any p(2,p0]p\in(2,p_{0}] and any u0L0p(Ω;B2,p,012/p(D))u_{0}\in\mathrm{L}^{p}_{\mathcal{F}_{0}}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D)) there is a unique solution uLp(Ω×(0,T);H01(D))Lp(Ω;C([0,T];B2,p,012/p(D)))u\in\mathrm{L}^{p}(\Omega\times(0,T);\mathrm{H}^{1}_{0}(D))\cap\mathrm{L}^{p}(\Omega;\mathrm{C}([0,T];\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))) to (SL) satisfying the estimate

(76) uLp(Ω;C([0,T];B2,p,012/p(D)))1+u0Lp(Ω;B2,p,012/p(D)),\displaystyle\|u\|_{\mathrm{L}^{p}(\Omega;\mathrm{C}([0,T];\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D)))}\lesssim 1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))},

and for all θ[0,1/2)\theta\in[0,\nicefrac{{1}}{{2}}) the estimate

(77) uLp(Ω;Hθ,p(0,T;H012θ(D)))θ1+u0Lp(Ω;B2,p,012/p(D)).\displaystyle\|u\|_{\mathrm{L}^{p}(\Omega;\mathrm{H}^{\theta,p}(0,T;\mathrm{H}^{1-2\theta}_{0}(D)))}\lesssim_{\theta}1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))}.

If θ(1/p,1/2)\theta\in(\nicefrac{{1}}{{p}},\nicefrac{{1}}{{2}}), then one has moreover the estimate

(78) uLp(Ω;Cθ1/p([0,T];H012θ(D)))θ1+u0Lp(Ω;B2,p,012/p(D)).\displaystyle\|u\|_{\mathrm{L}^{p}(\Omega;\mathrm{C}^{\theta-\nicefrac{{1}}{{p}}}([0,T];\mathrm{H}^{1-2\theta}_{0}(D)))}\lesssim_{\theta}1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))}.
Proof.

Step 1: Existence of variational solution and bootstrapping integrability.

Consider the progressively measurable maps

(79) A0(t,u)\displaystyle A_{0}(t,u) (i(𝒂ijαβ(t,x)juβ))α=1N,\displaystyle\coloneqq(\partial_{i}(\bm{a}^{\alpha\beta}_{ij}(t,x)\partial_{j}u^{\beta}))_{\alpha=1}^{N},
(80) B0(t,u)en\displaystyle B_{0}(t,u)e_{n} (𝒃n,jαβ(t,x)juβ)α=1N,\displaystyle\coloneqq(\bm{b}^{\alpha\beta}_{n,j}(t,x)\partial_{j}u^{\beta})_{\alpha=1}^{N},
(81) F(u)\displaystyle F(u) (iΦiα(u)+ϕα(u))α=1N,\displaystyle\coloneqq(\partial_{i}\Phi^{\alpha}_{i}(u)+\phi^{\alpha}(u))_{\alpha=1}^{N},
(82) G(u)en\displaystyle G(u)e_{n} (gnα(u))α=1N.\displaystyle\coloneqq(g^{\alpha}_{n}(u))_{\alpha=1}^{N}.

The well-posedness result [16, Thm. 2.4] applied with A(t,v)A0(t,v)+F(v)A(t,v)\coloneqq A_{0}(t,v)+F(v) and B(t,v)enB0(t,v)en+G(u)enB(t,v)e_{n}\coloneqq B_{0}(t,v)e_{n}+G(u)e_{n} yields a unique strong solution

uLp(Ω;L2(0,T;H01(D)))Lp(Ω;C([0,T];L2(D)))u\in\mathrm{L}^{p}(\Omega;\mathrm{L}^{2}(0,T;\mathrm{H}^{1}_{0}(D)))\cap\mathrm{L}^{p}(\Omega;\mathrm{C}([0,T];\mathrm{L}^{2}(D)))

together with the estimate

(83) uLp(Ω;L2(0,T;H01(D)))+uLp(Ω;C([0,T];L2(D)))1+u0Lp(Ω;L2(D))\displaystyle\|u\|_{\mathrm{L}^{p}(\Omega;\mathrm{L}^{2}(0,T;\mathrm{H}^{1}_{0}(D)))}+\|u\|_{\mathrm{L}^{p}(\Omega;\mathrm{C}([0,T];\mathrm{L}^{2}(D)))}\lesssim 1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{L}^{2}(D))}

for some p>2p>2 depending only on the quantities from Assumption 7.5, provided we can verify the hypotheses listed in [16, Ass. 2.1]. Using boundedness of the coefficients and linear growth of the non-linearities, (H4) and (H5) with f1f\equiv 1 follow readily. As in the proof of Lemma 7.4 one deduces (H3) with p>2p>2 sufficiently close to 22 from Assumption 7.5 (ii) and boundedness of B0B_{0}. Condition (H2) is void for A0(t,u)A_{0}(t,u) and B0(t,u)B_{0}(t,u) in virtue of (H3) and linearity, and is fulfilled for F(u)F(u) and G(u)G(u) by their linear growth. Likewise, (H1) is immediate from linearity and Lipschitz continuity.

Step 2: bootstrapping time-regularity.

A particular consequence of (83) is F(u)Lp(Ω×(0,T);H1(D))F(u)\in\mathrm{L}^{p}(\Omega\times(0,T);\mathrm{H}^{-1}(D)) and G(u)Lp(Ω×(0,T);2(2,L2(D)))G(u)\in\mathrm{L}^{p}(\Omega\times(0,T);\mathcal{L}_{2}(\ell^{2},\mathrm{L}^{2}(D))), with the estimate

(84) F(u)Lp(Ω×(0,T);H1(D))+G(u)Lp(Ω×(0,T);2(2,L2(D)))1+u0Lp(Ω;L2(D)).\displaystyle\|F(u)\|_{\mathrm{L}^{p}(\Omega\times(0,T);\mathrm{H}^{-1}(D))}+\|G(u)\|_{\mathrm{L}^{p}(\Omega\times(0,T);\mathcal{L}_{2}(\ell^{2},\mathrm{L}^{2}(D)))}\lesssim 1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{L}^{2}(D))}.

Consequently, if we freeze uu in the semi-linearities of (SL), that is to say, if we consider (SL) as a linear problem with right-hand sides f=F(u)f=F(u) and g=G(u)g=G(u), then Theorem 2.2 becomes applicable (of course we can restrict pp from Step 1 further so that the smallness condition in the theorem is verified) and yields a unique solution vLp(Ω×(0,T);H01(D))Lp(Ω;C([0,T];B2,p,012/p(D)))v\in\mathrm{L}^{p}(\Omega\times(0,T);\mathrm{H}^{1}_{0}(D))\cap\mathrm{L}^{p}(\Omega;\mathrm{C}([0,T];\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))). For the first estimate in the theorem calculate

(85) vLp(Ω;C([0,T];B2,p,012/p(D)))\displaystyle\|v\|_{\mathrm{L}^{p}(\Omega;\mathrm{C}([0,T];\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D)))}
(86) \displaystyle\lesssim{} u0Lp(Ω;B2,p,012/p(D))+F(u)Lp(Ω×(0,T),H1(D))+G(u)Lp(Ω×(0,T),2(2,L2(D)))\displaystyle\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))}+\|F(u)\|_{\mathrm{L}^{p}(\Omega\times(0,T),\mathrm{H}^{-1}(D))}+\|G(u)\|_{\mathrm{L}^{p}(\Omega\times(0,T),\mathcal{L}_{2}(\ell^{2},\mathrm{L}^{2}(D)))}
(87) \displaystyle\lesssim{} 1+u0Lp(Ω;B2,p,012/p(D)),\displaystyle 1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))},

where we used (84) and the embedding B2,p,012/p(D)L2(D)\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D)\subseteq\mathrm{L}^{2}(D) in the last step. The other estimates in the theorem follow also from Theorem 2.2 using the same argument. By uniqueness with p=2p=2, u=vu=v, which completes the proof. ∎

7.3. Uniform bounds with p>2p>2

Recall the family of solutions

(um)mL2(Ω;C([0,T];L2(D)))L2(Ω×(0,T);H01(D))(u_{m})_{m}\subseteq\mathrm{L}^{2}(\Omega;\mathrm{C}([0,T];\mathrm{L}^{2}(D)))\cap\mathrm{L}^{2}(\Omega\times(0,T);\mathrm{H}^{1}_{0}(D))

to the approximate problems (m-QL) from Lemma 7.4. Using the result from Section 7.2, we derive uniform Lp\mathrm{L}^{p}-bounds for some p>2p>2.

Proposition 7.7.

Suppose Assumption 7.1 holds. Then there is some p0>2p_{0}>2 such that for all mm, for all p(2,p0]p\in(2,p_{0}], and any u0L0p(Ω;B2,p,012/p(D))u_{0}\in\mathrm{L}^{p}_{\mathcal{F}_{0}}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D)) there is a unique solution

umLp(Ω×(0,T);H01(D))Lp(Ω;C([0,T];B2,p,012/p(D)))u_{m}\in\mathrm{L}^{p}(\Omega\times(0,T);\mathrm{H}^{1}_{0}(D))\cap\mathrm{L}^{p}(\Omega;\mathrm{C}([0,T];\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D)))

to (m-QL) satisfying the estimate

(88) umLp(Ω;C([0,T];B2,p,012/p(D)))1+u0Lp(Ω;B2,p,012/p(D)),\displaystyle\|u_{m}\|_{\mathrm{L}^{p}(\Omega;\mathrm{C}([0,T];\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D)))}\lesssim 1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))},

and for all θ[0,1/2)\theta\in[0,\nicefrac{{1}}{{2}}) the estimate

(89) umLp(Ω;Hθ,p(0,T;H012θ(D)))θ1+u0Lp(Ω;B2,p,012/p(D)).\displaystyle\|u_{m}\|_{\mathrm{L}^{p}(\Omega;\mathrm{H}^{\theta,p}(0,T;\mathrm{H}^{1-2\theta}_{0}(D)))}\lesssim_{\theta}1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))}.

If θ(1/p,1/2)\theta\in(\nicefrac{{1}}{{p}},\nicefrac{{1}}{{2}}), then one has moreover the estimate

(90) umLp(Ω;Cθ1/p([0,T];H012θ(D)))θ1+u0Lp(Ω;B2,p,012/p(D)).\displaystyle\|u_{m}\|_{\mathrm{L}^{p}(\Omega;\mathrm{C}^{\theta-\nicefrac{{1}}{{p}}}([0,T];\mathrm{H}^{1-2\theta}_{0}(D)))}\lesssim_{\theta}1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))}.
Proof.

Fix some mm. As already mentioned in Section 7.2, we can consider the coefficients 𝒂ijαβ(t,x)aijαβm(t,x,um(t,x))\bm{a}^{\alpha\beta}_{ij}(t,x)\coloneqq{}_{m}a^{\alpha\beta}_{ij}(t,x,u_{m}(t,x)) and 𝒃n,jαβ(t,x)bn,jαβm(t,x,umβ(t,x))\bm{b}^{\alpha\beta}_{n,j}(t,x)\coloneqq{}_{m}b^{\alpha\beta}_{n,j}(t,x,u_{m}^{\beta}(t,x)), which then satisfy Assumption 7.5. Now by Proposition 7.6, (SL) with these coefficients admits a unique solution vv that satisfies the bounds claimed in the proposition. But since

(91) v\displaystyle v\in{} Lp(Ω×(0,T);H01(D))Lp(Ω;C([0,T];B2,p,012/p(D)))\displaystyle\mathrm{L}^{p}(\Omega\times(0,T);\mathrm{H}^{1}_{0}(D))\cap\mathrm{L}^{p}(\Omega;\mathrm{C}([0,T];\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D)))
(92) \displaystyle\subseteq{} L2(Ω×(0,T);H01(D))L2(Ω;C([0,T];L2(D))),\displaystyle\mathrm{L}^{2}(\Omega\times(0,T);\mathrm{H}^{1}_{0}(D))\cap\mathrm{L}^{2}(\Omega;\mathrm{C}([0,T];\mathrm{L}^{2}(D))),

it follows v=umv=u_{m} by uniqueness for p=2p=2 stated in Lemma 7.4, which completes the proof. ∎

7.4. Stochastic compactness argument

In Theorem 1.3 we have already seen that a general tightness result can be deduced from a priori estimates such as the ones of Proposition 7.7. In the application to the quasi-linear system we need a slight variation of the tightness result in which a certain weak compactness is also taken into account.

In this section we suppose Assumption 7.1 holds, and we fix p0p_{0}, p(2,p0]p\in(2,p_{0}], and (um)m1(u_{m})_{m\geq 1} as in Proposition 7.6. Put

(93) 𝒳uC([0,T];L2(D))Lwp(0,T;H01(D)),𝒳WC([0,T];𝒰0),𝒳𝒳u×𝒳W.\displaystyle\mathcal{X}_{u}\coloneqq\mathrm{C}([0,T];\mathrm{L}^{2}(D))\cap\mathrm{L}^{p}_{\mathrm{w}}(0,T;\mathrm{H}^{1}_{0}(D)),\quad\mathcal{X}_{W}\coloneqq\mathrm{C}([0,T];\mathcal{U}_{0}),\quad\mathcal{X}\coloneqq\mathcal{X}_{u}\times\mathcal{X}_{W}.

With the subscript w\mathrm{w} we indicate that the space Lp(0,T;H01(D))\mathrm{L}^{p}(0,T;\mathrm{H}^{1}_{0}(D)) is equipped with the weak topology. As a consequence of separability, this has no consequence for questions of measurability. The space 𝒰0𝒰\mathcal{U}_{0}\supseteq\mathcal{U} is chosen in such a way that the cylindrical Brownian motion converges almost surely. The random vectors (um,W)(u_{m},W) take values in 𝒳\mathcal{X}. Write m\mathcal{L}_{m} for their (joint) laws.

We claim that the family (m)m(\mathcal{L}_{m})_{m} is tight. It is sufficient to show tightness of the laws (um)\mathcal{L}(u_{m}) on 𝒳u\mathcal{X}_{u}. Let us emphasise that no “diagonal structure” of the coefficients is needed.

Lemma 7.8 (Tightness).

The family of laws ((um))m(\mathcal{L}(u_{m}))_{m} on 𝒳u\mathcal{X}_{u} is tight.

Proof.

Let θ(1/p,1/2)\theta\in(\nicefrac{{1}}{{p}},\nicefrac{{1}}{{2}}). For brevity, put

𝒳uθ,pCθ1/p([0,T];H12θ(D))Lp(0,T;H01(D))\mathcal{X}_{u}^{\theta,p}\coloneqq\mathrm{C}^{\theta-\nicefrac{{1}}{{p}}}([0,T];\mathrm{H}^{1-2\theta}(D))\cap\mathrm{L}^{p}(0,T;\mathrm{H}^{1}_{0}(D))

and write BR\mathrm{B}_{R} for the open ball of radius RR in 𝒳uθ,p\mathcal{X}_{u}^{\theta,p}. First, by Chebyshev’s inequality and Proposition 7.7,

(94) (um𝒳uθ,pR)Rp𝔼um𝒳uθ,ppRp(1+u0Lp(Ω;B2,p,012/p(D))p)0 as R.\displaystyle\begin{split}\mathbb{P}(\|u_{m}\|_{\mathcal{X}_{u}^{\theta,p}}\geq R)&\leq R^{-p}\mathbb{E}\|u_{m}\|_{\mathcal{X}_{u}^{\theta,p}}^{p}\\ &\lesssim R^{-p}\Bigl{(}1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))}^{p}\Bigr{)}\to 0\quad\text{ as }R\to\infty.\end{split}

Second, we claim that the closure of BR\mathrm{B}_{R} is compact in 𝒳u\mathcal{X}_{u}. Indeed, pre-compactness in C([0,T];L2(D))\mathrm{C}([0,T];\mathrm{L}^{2}(D)) follows from the vector-valued Arzéla–Ascoli theorem [22, Theorem III.3.1], taking into account the compact embedding H01(D)L2(D)\mathrm{H}^{1}_{0}(D)\subseteq\mathrm{L}^{2}(D), and compactness in the space Lwp(0,T;H01(D))\mathrm{L}^{p}_{\mathrm{w}}(0,T;\mathrm{H}^{1}_{0}(D)) is clear, for BR\mathrm{B}_{R} is bounded in the norm topology. ∎

Therefore, Prokhorov’s theorem allows to pass to a weakly convergent subsequence (for convenience, we use the same symbol for this subsequence). Second, the Jakubowski–Skorohod theorem [9, Thm. 2.7.1] gives the following almost sure convergence.

Lemma 7.9 (Skorohod).

There exists a probability space (Ω~,~,~)(\tilde{\Omega},\tilde{\mathcal{F}},\tilde{\mathbb{P}}), a sequence of 𝒳\mathcal{X}-valued random variables (u~m,W~m)(\widetilde{u}_{m},\widetilde{W}_{m}), and a limiting 𝒳\mathcal{X}-valued random variable (u~,W~)(\widetilde{u},\widetilde{W}), such that

  1. (i)

    for each mm, the law of (u~m,W~m)(\widetilde{u}_{m},\widetilde{W}_{m}) under ~\tilde{\mathbb{P}} coincides with the law m\mathcal{L}_{m}, and

  2. (ii)

    one has ~\tilde{\mathbb{P}}-almost surely that u~mu~\widetilde{u}_{m}\to\widetilde{u} in C([0,T];L2(D))\mathrm{C}([0,T];\mathrm{L}^{2}(D)), u~mu~\nabla\widetilde{u}_{m}\to\nabla\widetilde{u} weakly in Lp(0,T;L2(D))\mathrm{L}^{p}(0,T;\mathrm{L}^{2}(D)), and W~mW~\widetilde{W}_{m}\to\widetilde{W} in C([0,T];𝒰0)\mathrm{C}([0,T];\mathcal{U}_{0}) as mm\to\infty.

Corollary 7.10.

The inclusion

(u~m)mLp(Ω~;C([0,T];L2(D)))Lp(Ω~×(0,T);H01(D))(\widetilde{u}_{m})_{m}\subseteq\mathrm{L}^{p}(\widetilde{\Omega};\mathrm{C}([0,T];\mathrm{L}^{2}(D)))\cap\mathrm{L}^{p}(\widetilde{\Omega}\times(0,T);\mathrm{H}^{1}_{0}(D))

holds with the uniform bound

(95) supm(u~mLp(Ω~;C([0,T];L2(D)))+u~mLp(Ω~×(0,T);H01(D)))1+u0Lp(Ω;B2,p,012/p(D)).\displaystyle\sup_{m}\left(\|\widetilde{u}_{m}\|_{\mathrm{L}^{p}(\widetilde{\Omega};\mathrm{C}([0,T];\mathrm{L}^{2}(D)))}+\|\widetilde{u}_{m}\|_{\mathrm{L}^{p}(\widetilde{\Omega}\times(0,T);\mathrm{H}^{1}_{0}(D))}\right)\lesssim 1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))}.

According to [12, Lem. 4.8], W~\widetilde{W} is a cylindrical Brownian motion. Fix α=1,,N\alpha=1,\dots,N and ξCc(D)\xi\in\mathrm{C}^{\infty}_{c}(D). Define

(96) M~α(t)\displaystyle\widetilde{M}^{\alpha}(t) =u~α(t),ξu~α(0),ξ+0taijαβ(u~)ju~β,iξdr\displaystyle=\langle\widetilde{u}^{\alpha}(t),\xi\rangle-\langle\widetilde{u}^{\alpha}(0),\xi\rangle+\int_{0}^{t}\langle a^{\alpha\beta}_{ij}(\widetilde{u})\partial_{j}\widetilde{u}^{\beta},\partial_{i}\xi\rangle\,\mathrm{d}r
(97) +0tΦiα(u~),iξdr0tϕα(u~),ξdr.\displaystyle\qquad+\int_{0}^{t}\langle\Phi^{\alpha}_{i}(\widetilde{u}),\partial_{i}\xi\rangle\,\mathrm{d}r-\int_{0}^{t}\langle\phi^{\alpha}(\widetilde{u}),\xi\rangle\,\mathrm{d}r.

As presented in the proof of [12, Prop. 4.7], well-posedness of (QL) under Assumption 7.1, that is to say, validity of Theorem 7.2, follows directly from the following lemma.

Lemma 7.11.

The processes

(98) M~,M~2n10[bn,jαβ(u~β)ju~β,ξ+gnα(u~),ξ]2dr,andM~wn0bn,jαβ(u~β)ju~β,ξ+gnα(u~),ξdr\displaystyle\begin{aligned} \widetilde{M},&\ \widetilde{M}^{2}-\sum_{n\geq 1}\int_{0}^{\cdot}\big{[}\langle b^{\alpha\beta}_{n,j}(\widetilde{u}^{\beta})\partial_{j}\widetilde{u}^{\beta},\xi\rangle+\langle g^{\alpha}_{n}(\widetilde{u}),\xi\rangle\big{]}^{2}\,\mathrm{d}r,\\ &\ \text{and}\ \widetilde{M}w_{n}-\int_{0}^{\cdot}\langle b^{\alpha\beta}_{n,j}(\widetilde{u}^{\beta})\partial_{j}\widetilde{u}^{\beta},\xi\rangle+\langle g^{\alpha}_{n}(\widetilde{u}),\xi\rangle\,\mathrm{d}r\end{aligned}

are (~t)(\widetilde{\mathcal{F}}_{t})-martingales, where ~tσ({u~(s),w~n(s):st,n1})\widetilde{\mathcal{F}}_{t}\coloneqq\sigma(\{\widetilde{u}(s),\widetilde{w}_{n}(s)\colon s\leq t,n\geq 1\}).

Proof.

In a first step, we showcase the general strategy for the process M~\widetilde{M}. Afterwards, we explain the necessary changes for the two remaining processes.

Step 1: M~\widetilde{M} is an (~t)(\widetilde{\mathcal{F}}_{t})-martingale.

Introduce the processes

(99) Mmα(t)\displaystyle M^{\alpha}_{m}(t) =umα(t),ξu0α,ξ+0taijαβm(Rmum)jumβ,iξdr\displaystyle=\langle u^{\alpha}_{m}(t),\xi\rangle-\langle u^{\alpha}_{0},\xi\rangle+\int_{0}^{t}\langle{}_{m}a^{\alpha\beta}_{ij}(R_{m}u_{m})\partial_{j}u^{\beta}_{m},\partial_{i}\xi\rangle\,\mathrm{d}r
(100) +0tΦiα(um),iξdr0tϕα(um),ξdr,\displaystyle\qquad+\int_{0}^{t}\langle\Phi^{\alpha}_{i}(u_{m}),\partial_{i}\xi\rangle\,\mathrm{d}r-\int_{0}^{t}\langle\phi^{\alpha}(u_{m}),\xi\rangle\,\mathrm{d}r,
(101) M~mα(t)\displaystyle\widetilde{M}^{\alpha}_{m}(t) =u~mα(t),ξu~mα(0),ξ+0taijαβm(Rmu~m)ju~mβ,iξdr\displaystyle=\langle\widetilde{u}^{\alpha}_{m}(t),\xi\rangle-\langle\widetilde{u}^{\alpha}_{m}(0),\xi\rangle+\int_{0}^{t}\langle{}_{m}a^{\alpha\beta}_{ij}(R_{m}\widetilde{u}_{m})\partial_{j}\widetilde{u}^{\beta}_{m},\partial_{i}\xi\rangle\,\mathrm{d}r
(102) +0tΦiα(u~m),iξdr0tϕα(u~m),ξdr.\displaystyle\qquad+\int_{0}^{t}\langle\Phi^{\alpha}_{i}(\widetilde{u}_{m}),\partial_{i}\xi\rangle\,\mathrm{d}r-\int_{0}^{t}\langle\phi^{\alpha}(\widetilde{u}_{m}),\xi\rangle\,\mathrm{d}r.

For each mm, MmαM^{\alpha}_{m} is an (t)(\mathcal{F}_{t})-martingale because umu_{m} is a solution to (m-QL).

Fix α\alpha and sts\leq t throughout the proof. Let ρs\rho_{s} denote the canonical restriction C([0,T];𝒳)C([0,s];𝒳)\mathrm{C}([0,T];\mathcal{X})\to\mathrm{C}([0,s];\mathcal{X}) and fix any continuous function γ:C([0,s];𝒳)[0,1]\gamma\colon\mathrm{C}([0,s];\mathcal{X})\to[0,1]. One has that

γmγ(ρs(um,W))\gamma_{m}\coloneqq\gamma(\rho_{s}(u_{m},W))

is s\mathcal{F}_{s}-measurable. Therefore, conditioning and the martingale property yield

(103) 𝔼(γm[Mmα(t)Mmα(s)])=0.\displaystyle\mathbb{E}\Bigl{(}\gamma_{m}[M^{\alpha}_{m}(t)-M^{\alpha}_{m}(s)]\Bigr{)}=0.

We claim that γm\gamma_{m}, Mmα(t)M^{\alpha}_{m}(t), and Mmα(s)M^{\alpha}_{m}(s) depend on uu in a measurable way. Indeed, these quantities even depend continuously on uu if we equip Lp(0,T;H01(D))\mathrm{L}^{p}(0,T;\mathrm{H}^{1}_{0}(D)) with the strong topology in the definition of 𝒳\mathcal{X}. As already mentioned, passing to the weak topology preserves measurability, which gives the claim. Therefore, with Lemma 7.9 (i) deduce from (103) that

(104) 𝔼~(γ~m[M~mα(t)M~mα(s)])=0,\displaystyle\widetilde{\mathbb{E}}\Bigl{(}\widetilde{\gamma}_{m}[\widetilde{M}^{\alpha}_{m}(t)-\widetilde{M}^{\alpha}_{m}(s)]\Bigr{)}=0,

where γ~m\widetilde{\gamma}_{m} and M~mα\widetilde{M}^{\alpha}_{m} are defined by the same expressions as γm\gamma_{m} and MmαM^{\alpha}_{m}, but with umu_{m} replaced by u~m\widetilde{u}_{m} and WW replaced by W~\widetilde{W}.

Next, we take the limit mm\to\infty in (104). To do so, we appeal to Vitali’s convergence theorem. Owing to Corollary 7.10, we have for instance

(105) 𝔼~|0taijαβm(Rmu~m)ju~mβ,iξdr|p\displaystyle\widetilde{\mathbb{E}}\left|\int_{0}^{t}\langle{}_{m}a^{\alpha\beta}_{ij}(R_{m}\widetilde{u}_{m})\partial_{j}\widetilde{u}^{\beta}_{m},\partial_{i}\xi\rangle\,\mathrm{d}r\right|^{p} Tp1𝔼~0Tu~mL2(D)p\displaystyle\lesssim T^{p-1}\widetilde{\mathbb{E}}\int_{0}^{T}\|\nabla\widetilde{u}_{m}\|_{\mathrm{L}^{2}(D)}^{p}
(106) 1+u0Lp(Ω;B2,p,012/p(D))p,\displaystyle\lesssim 1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))}^{p},

which justifies the application of Vitali’s convergence theorem. It remains to determine the almost-sure limit of γ~m[M~mα(t)M~α(t)]\widetilde{\gamma}_{m}[\widetilde{M}^{\alpha}_{m}(t)-\widetilde{M}^{\alpha}(t)]. On the one hand, by continuity of γρs\gamma\circ\rho_{s} on 𝒳\mathcal{X}, deduce γ~mγ~\widetilde{\gamma}_{m}\to\widetilde{\gamma} almost surely from Lemma 7.9 (ii). On the other hand, we claim that M~mα(t)M~α(t)\widetilde{M}^{\alpha}_{m}(t)\to\widetilde{M}^{\alpha}(t) almost surely. Convergence of the first two terms of M~mα(t)\widetilde{M}^{\alpha}_{m}(t) is immediate. For the fourth and fifth term, use that Φ\Phi and ϕ\phi are Lipschitz continuous along with almost sure convergence in C([0,T];L2(D))L1(0,T;L2(D))\mathrm{C}([0,T];\mathrm{L}^{2}(D))\subseteq\mathrm{L}^{1}(0,T;\mathrm{L}^{2}(D)). The most challenging term is the third. Rewrite

0taijαβm(Rmu~m)ju~mβ,iξdr0taijαβ(u~)ju~β,iξdr\displaystyle\int_{0}^{t}\langle{}_{m}a^{\alpha\beta}_{ij}(R_{m}\widetilde{u}_{m})\partial_{j}\widetilde{u}^{\beta}_{m},\partial_{i}\xi\rangle\,\mathrm{d}r-\int_{0}^{t}\langle a^{\alpha\beta}_{ij}(\widetilde{u})\partial_{j}\widetilde{u}^{\beta},\partial_{i}\xi\rangle\,\mathrm{d}r
=\displaystyle={} 0tju~mβ,(aijαβm(Rmu~m)aijαβ(u~))iξdr+0tj(u~mβu~β),aijαβ(u~)iξdr\displaystyle\int_{0}^{t}\langle\partial_{j}\widetilde{u}^{\beta}_{m},({}_{m}a^{\alpha\beta}_{ij}(R_{m}\widetilde{u}_{m})-a^{\alpha\beta}_{ij}(\widetilde{u}))\partial_{i}\xi\rangle\,\mathrm{d}r+\int_{0}^{t}\langle\partial_{j}(\widetilde{u}^{\beta}_{m}-\widetilde{u}^{\beta}),a^{\alpha\beta}_{ij}(\widetilde{u})\partial_{i}\xi\rangle\,\mathrm{d}r
=\displaystyle={} I+II.\displaystyle\textrm{I}+\textrm{II}.

We use the convergence properties from Lemma 7.9 (ii) as follows to conclude: using the weak convergence of u~m\nabla\widetilde{u}_{m} in L2(0,t;L2(D))\mathrm{L}^{2}(0,t;\mathrm{L}^{2}(D)) we see that II0\textrm{II}\to 0 as mm\to\infty almost surely. Moreover, u~m\nabla\widetilde{u}_{m} is uniformly bounded in Lp(0,T;L2(D))L2(0,t;L2(D))\mathrm{L}^{p}(0,T;\mathrm{L}^{2}(D))\subseteq\mathrm{L}^{2}(0,t;\mathrm{L}^{2}(D)) almost surely. Consequently, I goes to zero if (aijαβm(Rmu~m)aijαβ(u~))iξ0({}_{m}a^{\alpha\beta}_{ij}(R_{m}\widetilde{u}_{m})-a^{\alpha\beta}_{ij}(\widetilde{u}))\partial_{i}\xi\to 0 in L2(0,t;L2(D))\mathrm{L}^{2}(0,t;\mathrm{L}^{2}(D)). Split further

(107) aijαβm(Rmu~m)aijαβ(u~)=(aijαβm(Rmu~m)aijαβ(Rmu~m))+(aijαβ(Rmu~m)aijαβ(u~)).\displaystyle{}_{m}a^{\alpha\beta}_{ij}(R_{m}\widetilde{u}_{m})-a^{\alpha\beta}_{ij}(\widetilde{u})={}({}_{m}a^{\alpha\beta}_{ij}(R_{m}\widetilde{u}_{m})-a^{\alpha\beta}_{ij}(R_{m}\widetilde{u}_{m}))+(a^{\alpha\beta}_{ij}(R_{m}\widetilde{u}_{m})-a^{\alpha\beta}_{ij}(\widetilde{u})).

Since the coefficients are uniformly bounded and ξ\xi is smooth and compactly supported, we have Vitali’s convergence theorem in (r,x)(r,x) at our disposal. Up to passing to another subsequence, we can suppose Rmu~mu~R_{m}\widetilde{u}_{m}\to\widetilde{u} pointwise almost everywhere in (r,x)(r,x). In particular, for almost every (r,x)(r,x), the closure of the set {Rmu~m(r,x):m}\{R_{m}\widetilde{u}_{m}(r,x)\colon m\in\mathbb{N}\} is compact. Then, the claim follows by continuity of aijαβa^{\alpha\beta}_{ij} in conjunction with the convergence aijαβmaijαβ{}_{m}a^{\alpha\beta}_{ij}\to a^{\alpha\beta}_{ij} uniformly on compact subsets.

In conclusion, we deduce

(108) 𝔼~(γ~[M~α(t)M~α(s)])=0.\displaystyle\widetilde{\mathbb{E}}\Bigl{(}\widetilde{\gamma}[\widetilde{M}^{\alpha}(t)-\widetilde{M}^{\alpha}(s)]\Bigr{)}=0.

But (108) remains valid if we replace γ~\widetilde{\gamma} by any bounded and ~s\widetilde{\mathcal{F}}_{s}-measurable function in virtue of the monotone class theorem, since ~s\widetilde{\mathcal{F}}_{s} is by definition the σ\sigma-field induced by u~\widetilde{u} and W~\widetilde{W}. It follows that M~α\widetilde{M}^{\alpha} is an (~t)(\widetilde{\mathcal{F}}_{t})-martingale as claimed.

Step 2: the remaining two processes.

The general strategy is as in Step 1 and we only discuss the necessary modifications. The martingale property for the processes related to umu_{m} is again clear. We follow the proof of Step 1 until we reach the analogues of (104). The first step in the identification of the limit, the moment condition (105), can be reused with the following observations: first of all, since the moment condition for one term holds with p>2p>2, any product appearing in (M~mα)2(\widetilde{M}_{m}^{\alpha})^{2} satisfies a moment condition with p/2>1\nicefrac{{p}}{{2}}>1 in virtue of Hölder’s inequality. The same is true for M~mwk\widetilde{M}_{m}w_{k}, of course. Likewise, one has

(109) 𝔼~|0tbn,jαβm(Rmu~mβ)ju~mβ,ξ2dr|p/2\displaystyle\widetilde{\mathbb{E}}\left|\int_{0}^{t}\langle{}_{m}b^{\alpha\beta}_{n,j}(R_{m}\widetilde{u}_{m}^{\beta})\partial_{j}\widetilde{u}_{m}^{\beta},\xi\rangle^{2}\,\mathrm{d}r\right|^{\nicefrac{{p}}{{2}}} Tp/21𝔼~0Tu~mL2(D)p\displaystyle\lesssim T^{\nicefrac{{p}}{{2}}-1}\widetilde{\mathbb{E}}\int_{0}^{T}\|\nabla\widetilde{u}_{m}\|_{\mathrm{L}^{2}(D)}^{p}
(110) 1+u0Lp(Ω;B2,p,012/p(D))p.\displaystyle\lesssim 1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))}^{p}.

It remains to verify almost sure convergence, for instance

(111) 0tbn,jαβm(Rmu~mβ)ju~mβ,ξ2dr0tbn,jαβ(u~β)ju~β,ξ2dr,\displaystyle\int_{0}^{t}\langle{}_{m}b^{\alpha\beta}_{n,j}(R_{m}\widetilde{u}_{m}^{\beta})\partial_{j}\widetilde{u}_{m}^{\beta},\xi\rangle^{2}\,\mathrm{d}r\to\int_{0}^{t}\langle b^{\alpha\beta}_{n,j}(\widetilde{u}^{\beta})\partial_{j}\widetilde{u}^{\beta},\xi\rangle^{2}\,\mathrm{d}r,

for fixed nn. First, we claim that we can replace bn,jαβm(Rmu~mβ){}_{m}b^{\alpha\beta}_{n,j}(R_{m}\widetilde{u}_{m}^{\beta}) by bn,jαβ(u~β)b^{\alpha\beta}_{n,j}(\widetilde{u}^{\beta}) on the left-hand side of (111). Indeed, the calculation is similar to the treatment of term I in Step 1, but one has to use uniform boundedness of u~m\nabla\widetilde{u}_{m} in Lp(0,T;L2(D))\mathrm{L}^{p}(0,T;\mathrm{L}^{2}(D)) with p>2p>2. Next, define the auxiliary function

(112) b¯n,jαβ(r,x,y)0ybn,jαβ(r,x,z)dz,(y).\displaystyle\bar{b}_{n,j}^{\alpha\beta}(r,x,y)\coloneqq\int_{0}^{y}b_{n,j}^{\alpha\beta}(r,x,z)\,\mathrm{d}z,\qquad(y\in\mathbb{R}).

By the chain rule, jb¯n,jαβ(r,x,u~(r,x))=bn,jαβ(r,x,u~(r,x))ju~(r,x)+Θn,jαβ(r,x,u~β(r,x))\partial_{j}\bar{b}_{n,j}^{\alpha\beta}(r,x,\widetilde{u}(r,x))=b_{n,j}^{\alpha\beta}(r,x,\widetilde{u}(r,x))\partial_{j}\widetilde{u}(r,x)+\Theta^{\alpha\beta}_{n,j}(r,x,\widetilde{u}^{\beta}(r,x)), where Θn,jαβ(r,x,y)=0yjbn,jαβ(r,x,z)dz\Theta^{\alpha\beta}_{n,j}(r,x,y)=\int_{0}^{y}\partial_{j}b_{n,j}^{\alpha\beta}(r,x,z)\,\mathrm{d}z. Therefore, to prove (111) it is enough to show

(113) 0tb¯n,jαβ(u~β)b¯n,jαβ(u~mβ),jξ2dr0and0tΘn,jαβ(,u~β)Θn,jαβ(,u~mβ),ξ2dr0.\displaystyle\int_{0}^{t}\langle\bar{b}^{\alpha\beta}_{n,j}(\widetilde{u}^{\beta})-\bar{b}^{\alpha\beta}_{n,j}(\widetilde{u}_{m}^{\beta}),\partial_{j}\xi\rangle^{2}\,\mathrm{d}r\to 0\ \ \text{and}\ \ \int_{0}^{t}\langle\Theta^{\alpha\beta}_{n,j}(\cdot,\widetilde{u}^{\beta})-\Theta^{\alpha\beta}_{n,j}(\cdot,\widetilde{u}_{m}^{\beta}),\xi\rangle^{2}\,\mathrm{d}r\to 0.

For the first term, using Vitali’s convergence theorem once more, we only need to show convergence for a fixed time r(0,t)r\in(0,t). Now since u~mβu~β\widetilde{u}_{m}^{\beta}\to\widetilde{u}^{\beta} in L2(D)\mathrm{L}^{2}(D), we can conclude the proof by passing to another subsequence and using continuity of b¯n,jαβ\bar{b}^{\alpha\beta}_{n,j}. For the second term, the uniform boundedness of jbn,jαβ\partial_{j}b_{n,j}^{\alpha\beta} implies that |Θn,jαβ(,u~β)Θn,jαβ(,u~mβ)||u~βu~mβ||\Theta^{\alpha\beta}_{n,j}(\cdot,\widetilde{u}^{\beta})-\Theta^{\alpha\beta}_{n,j}(\cdot,\widetilde{u}_{m}^{\beta})|\lesssim|\widetilde{u}^{\beta}-\widetilde{u}^{\beta}_{m}|. Therefore, the desired convergence follows from u~mu~\widetilde{u}_{m}\to\widetilde{u} in L2(0,t;L2(D))\mathrm{L}^{2}(0,t;\mathrm{L}^{2}(D)).

Let us stress that the last argument heavily relies on the structural assumption on bb, and is the reason why we cannot allow full non-linear dependence on all components of uu as is the case for aa. ∎

Remark 7.12.

We stress that the above proof critically uses the pp-integrability of u~m\nabla\widetilde{u}_{m} in Ω~×(0,T)\widetilde{\Omega}\times(0,T). Indeed, to get L1(Ω)\mathrm{L}^{1}(\Omega)-convergence of (109) we used Vitali’s convergence theorem based on the uniform boundedness of u~m\nabla\widetilde{u}_{m} in Lp(Ω×(0,T);L2(D))\mathrm{L}^{p}(\Omega\times(0,T);\mathrm{L}^{2}(D)) with p>2p>2. That such a bound is available for p>2p>2 is a novelty of our approach.

8. Quasilinear problem with growing non-linearities

We reduce the general case with growing non-linearities to the Lipschitz case from the last section. This happens using a truncation argument. We use the following classical result from convex analysis (see [10, Proposition 5.3]) to include systems in a neat way.

Lemma 8.1 (Phelps).

Let CC be a non-empty, closed, convex subset of a Hilbert space KK. Then the projection onto CC is Lipschitz continuous with Lipschitz constant 11.

With the preceding lemma, we can consider truncated non-linearities ΦP\Phi\circ P and ϕP\phi\circ P, where PP is a projection to a bounded and convex set. As in Section 7.2, we would like to consider f=Φi(P(u))+ϕ(P(u))f=\Phi_{i}(P(u))+\phi(P(u)) as a fixed right-hand side. However, that ff is an admissible right-hand side is more difficult with growing non-linearities. We will take care of this in Proposition 8.4. The next two lemmas are a preparation for the last-mentioned proposition.

Lemma 8.2 ([13, Lem. 8]).

For q>2q>2 and m1m\geq 1, the function ψm:\psi_{m}\colon\mathbb{R}\to\mathbb{R} given by

(114) ψm(ξ)={|ξ|q,|ξ|m,mq2[q(q1)2ξ2q(q2)m|ξ|+(q1)(q2)2m2],|ξ|>m\displaystyle\psi_{m}(\xi)=\begin{cases}|\xi|^{q},&|\xi|\leq m,\\ m^{q-2}\Bigl{[}\frac{q(q-1)}{2}\xi^{2}-q(q-2)m|\xi|+\frac{(q-1)(q-2)}{2}m^{2}\Bigr{]},&|\xi|>m\end{cases}

is twice continuously differentiable, has bounded second derivative, and satisfies the following properties:

(115) |ψm(ξ)|\displaystyle|\psi_{m}^{\prime}(\xi)| |ξ|ψm′′(ξ),\displaystyle\leq|\xi|\psi_{m}^{\prime\prime}(\xi),
(116) ψm(ξ)ξ\displaystyle\frac{\psi_{m}^{\prime}(\xi)}{\xi} 0,(ξ0),\displaystyle\geq 0,\mathrlap{\quad(\xi\neq 0),}
(117) ψm′′(ξ)\displaystyle\psi_{m}^{\prime\prime}(\xi) q(q1)(1+ψm(ξ)),\displaystyle\leq q(q-1)(1+\psi_{m}(\xi)),
(118) ξ2ψm′′(ξ)\displaystyle\xi^{2}\psi_{m}^{\prime\prime}(\xi) q(q1)ψm(ξ),\displaystyle\leq q(q-1)\psi_{m}(\xi),
(119) ψm′′(ξ1)\displaystyle\psi_{m}^{\prime\prime}(\xi_{1}) ψm′′(ξ2),(|ξ1||ξ2|).\displaystyle\leq\psi_{m}^{\prime\prime}(\xi_{2}),\mathrlap{\quad(|\xi_{1}|\leq|\xi_{2}|).}

Moreover, ψm||q\psi_{m}\to|\cdot|^{q} and ψm′′||q2\psi_{m}^{\prime\prime}\to|\cdot|^{q-2} pointwise.

The following technical lemma relates the dissipativity of ϕ\phi stated in Assumption 6.1 (v) with the auxiliary function ψm\psi_{m}.

Lemma 8.3.

Let vNv\in\mathbb{R}^{N}. For all q(2,)q\in(2,\infty) and m1m\geq 1 one has for ψm\psi_{m} defined as in Lemma 8.2 the estimate

(120) αψm(vα)ϕα(v)q1+βψm(vβ).\displaystyle\sum_{\alpha}\psi_{m}^{\prime}(v^{\alpha})\phi^{\alpha}(v)\lesssim_{q}1+\sum_{\beta}\psi_{m}(v^{\beta}).
Proof.

Let vNv\in\mathbb{R}^{N} and fix α\alpha for the moment. We can assume vα0v^{\alpha}\neq 0. Calculate using (116), Assumption 6.1 (v), and (115) that

(121) ψm(vα)ϕα(v)=ψm(vα)vαvαϕα(v)ψm(vα)vα(1+|v|2)ψm′′(vα)(1+|v|2).\displaystyle\psi_{m}^{\prime}(v^{\alpha})\phi^{\alpha}(v)=\frac{\psi_{m}^{\prime}(v^{\alpha})}{v^{\alpha}}v^{\alpha}\phi^{\alpha}(v)\lesssim\frac{\psi_{m}^{\prime}(v^{\alpha})}{v^{\alpha}}(1+|v|^{2})\lesssim\psi_{m}^{\prime\prime}(v^{\alpha})(1+|v|^{2}).

Using (117), the term ψm′′(vα)\psi_{m}^{\prime\prime}(v^{\alpha}) can be controlled by 1+ψm(vα)1+\psi_{m}(v^{\alpha}). Likewise, ψm′′(vα)|vα|2\psi_{m}^{\prime\prime}(v^{\alpha})|v^{\alpha}|^{2} is controlled by ψm(vα)\psi_{m}(v^{\alpha}) owing to (118). It remains to consider ψm′′(vα)|vβ|2\psi_{m}^{\prime\prime}(v^{\alpha})|v^{\beta}|^{2}. We distinguish the cases |vα||vβ||v^{\alpha}|\leq|v^{\beta}| and |vα||vβ||v^{\alpha}|\geq|v^{\beta}|. In the first case, we use that ψm′′\psi_{m}^{\prime\prime} is increasing, (119), to reduce to the known case ψm′′(vβ)|vβ|2\psi_{m}^{\prime\prime}(v^{\beta})|v^{\beta}|^{2}. Likewise, we reduce in the second case using that ||2|\cdot|^{2} is increasing. Finally, summing in α\alpha gives the claim. ∎

Proposition 8.4 (Bootstrapping integrability).

Suppose Assumption 6.1 holds, and for the moment suppose that there is a constant LL such that

|ϕ(t,x,y)|+|Φ(t,x,y)|L(1+|y|),t[0,T],xD,yN.|\phi(t,x,y)|+|\Phi(t,x,y)|\leq L(1+|y|),\ \ \ t\in[0,T],x\in D,y\in\mathbb{R}^{N}.

Let u0L0q(Ω×D)u_{0}\in\mathrm{L}^{q}_{\mathcal{F}_{0}}(\Omega\times D) with q[2,)q\in[2,\infty) fixed and let vL2(Ω×(0,T);H01(D))L2(Ω;C([0,T];L2(D)))v\in\mathrm{L}^{2}(\Omega\times(0,T);\mathrm{H}^{1}_{0}(D))\cap\mathrm{L}^{2}(\Omega;\mathrm{C}([0,T];\mathrm{L}^{2}(D))) be a solution to (QL) with initial datum u0u_{0}. Then one has

(122) vLq(Ω×(0,T)×D)q+𝔼0TD|v|q2|v|2dxds1+u0Lq(Ω;Lq(D))q,\displaystyle\|v\|_{\mathrm{L}^{q}(\Omega\times(0,T)\times D)}^{q}+\mathbb{E}\int_{0}^{T}\int_{D}|v|^{q-2}|\nabla v|^{2}\,\mathrm{d}x\,\mathrm{d}s\lesssim 1+\|u_{0}\|_{\mathrm{L}^{q}(\Omega;\mathrm{L}^{q}(D))}^{q},

where the implicit constant does not depend on LL.

The additional growth condition on ϕ\phi and Φ\Phi are needed to make sure that the functions in the proof below are integrable. Later on, we will apply the lemma to truncated versions of ϕ\phi and Φ\Phi, and therefore the growth condition does not lead to additional assumptions.

Proof.

Fix α\alpha and let m1m\geq 1, t[0,T]t\in[0,T]. Define the linear functional vDψm(v)dxv\mapsto\int_{D}\psi_{m}(v)\,\mathrm{d}x on L2(D)\mathrm{L}^{2}(D). Since ψm′′\psi_{m}^{\prime\prime} is bounded and, taking (115) into account, ψm\psi_{m}^{\prime} is of linear growth, the Itô formula from [28, Thm. 4.2] is applicable and yields

(123) Dψm(vα(t))dx\displaystyle\int_{D}\psi_{m}(v^{\alpha}(t))\,\mathrm{d}x =Dψm(vα(0))dx\displaystyle=\int_{D}\psi_{m}(v^{\alpha}(0))\,\mathrm{d}x
(124) 0tDψm′′(vα)aijα(v)ivαjvαdxds\displaystyle\quad-\int_{0}^{t}\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})a_{ij}^{\alpha}(v)\partial_{i}v^{\alpha}\partial_{j}v^{\alpha}\,\mathrm{d}x\,\mathrm{d}s
(125) 0tDψm′′(vα)Φiα(v)ivαdxds\displaystyle\quad-\int_{0}^{t}\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})\Phi_{i}^{\alpha}(v)\partial_{i}v^{\alpha}\,\mathrm{d}x\,\mathrm{d}s
(126) +0tDψm(vα)ϕα(v)dxds\displaystyle\quad+\int_{0}^{t}\int_{D}\psi_{m}^{\prime}(v^{\alpha})\phi^{\alpha}(v)\,\mathrm{d}x\,\mathrm{d}s
(127) +n0tDψm(vα)[bn,jα(vα)jvα+gnα(v)]dxdwn\displaystyle\quad+\sum_{n}\int_{0}^{t}\int_{D}\psi_{m}^{\prime}(v^{\alpha})\bigl{[}b_{n,j}^{\alpha}(v^{\alpha})\partial_{j}v^{\alpha}+g_{n}^{\alpha}(v)\bigr{]}\,\mathrm{d}x\,\mathrm{d}w_{n}
(128) +n120tDψm′′(vα)|jbn,jα(vα)jvα+gnα(v)|2dxds\displaystyle\quad+\sum_{n}\frac{1}{2}\int_{0}^{t}\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})\Bigl{|}\sum_{j}b_{n,j}^{\alpha}(v^{\alpha})\partial_{j}v^{\alpha}+g_{n}^{\alpha}(v)\Bigr{|}^{2}\,\mathrm{d}x\,\mathrm{d}s
(129) =IIIIII+IV+V+VI.\displaystyle\quad=\textrm{I}-\textrm{II}-\textrm{III}+\textrm{IV}+\textrm{V}+\textrm{VI}.

Recall that Φα(s,x,y)=Φ^α(s,x,y)+Φ¯α(yα)\Phi^{\alpha}(s,x,y)=\widehat{\Phi}^{\alpha}(s,x,y)+\overline{\Phi}^{\alpha}(y^{\alpha}). Therefore, the part of III involving Φ¯α\overline{\Phi}^{\alpha} vanishes as can be seen in the same way as in [4, Lemma 3.5]. Indeed,

Dψm′′(vα)Φ¯iα(vα)ivαdx=Ddivx0vαψm′′(r)Φ¯iα(r)drdx=0,\displaystyle\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})\overline{\Phi}_{i}^{\alpha}(v^{\alpha})\partial_{i}v^{\alpha}\,\mathrm{d}x=\int_{D}\operatorname{div}_{x}\int_{0}^{v^{\alpha}}\psi_{m}^{\prime\prime}(r)\overline{\Phi}_{i}^{\alpha}(r)\,\mathrm{d}r\,\mathrm{d}x=0,

where we applied the divergence theorem and the fact that vαv^{\alpha} vanishes at the boundary. Thus III can be bounded as follows

III\displaystyle-\textrm{III} |0tDψm′′(vα)Φ^iα(v)ivαdxds|\displaystyle\leq\Big{|}\int_{0}^{t}\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})\widehat{\Phi}_{i}^{\alpha}(v)\partial_{i}v^{\alpha}\,\mathrm{d}x\,\mathrm{d}s\Big{|}
Cδ0tDψm′′(vα)|Φ^α(v)|2dxds+δ0tDψm′′(vα)|vα|2dxds\displaystyle\leq C_{\delta}\int_{0}^{t}\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})|\widehat{\Phi}^{\alpha}(v)|^{2}\,\mathrm{d}x\,\mathrm{d}s+\delta\int_{0}^{t}\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})|\nabla v^{\alpha}|^{2}\,\mathrm{d}x\,\mathrm{d}s
=:III(Φ^α(v))+III(vα).\displaystyle=:\textrm{III}(\widehat{\Phi}^{\alpha}(v))+\textrm{III}(\nabla v^{\alpha}).

Use the inequality 1/2|b+g|2(1/2+ε)b2+Cεg2\nicefrac{{1}}{{2}}|b+g|^{2}\leq(\nicefrac{{1}}{{2}}+\varepsilon)b^{2}+C_{\varepsilon}g^{2} for all ε>0\varepsilon>0 in VI to get

(130) VI (1/2+ε)n0tDψm′′(vα)|jbn,jα(vα)jvα|2dxds\displaystyle\leq(\nicefrac{{1}}{{2}}+\varepsilon)\sum_{n}\int_{0}^{t}\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})\Bigl{|}\sum_{j}b_{n,j}^{\alpha}(v^{\alpha})\partial_{j}v^{\alpha}\Bigr{|}^{2}\,\mathrm{d}x\,\mathrm{d}s
(131) +Cεn0tDψm′′(vα)|gnα(v)|2dxds\displaystyle\qquad+C_{\varepsilon}\sum_{n}\int_{0}^{t}\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})|g_{n}^{\alpha}(v)|^{2}\,\mathrm{d}x\,\mathrm{d}s
(132) =VI(b)+VI(g).\displaystyle=\textrm{VI}(b)+\textrm{VI}(g).

By the “diagonal structure” of aa and bb stated in Assumption 6.1 (i), the coercivity condition in Assumption 6.1 (iii) can be applied componentwise. Therefore, by choosing ε\varepsilon and δ\delta small enough (similar to the proof of Lemma 7.4), we get the lower bound

(133) IIIII(vα)VI(b)λ20tDψm′′(vα)|vα|2dxds.\displaystyle\textrm{II}-\textrm{III}(\nabla v^{\alpha})-\textrm{VI}(b)\geq\frac{\lambda}{2}\int_{0}^{t}\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})|\nabla v^{\alpha}|^{2}\,\mathrm{d}x\,\mathrm{d}s.

This is the only argument that uses Assumption 6.1 (i), see Remark 6.3 for further discussion. Hence, we can absorb II+III(vα)+VI(b)-\textrm{II}+\textrm{III}(\nabla v^{\alpha})+\textrm{VI}(b) into the left-hand side, to give

(134) Dψm(vα)dx+λ20tDψm′′(vα)|vα|2dxdsDψm(vα(0))dx+0tDψm(vα)ϕα(v)dxds+n0tDψm(vα)[bn,jα(vα)jvα+gnα(v)]dxdwn+Cεn0tDψm′′(vα)|gnα(v)|2dxds+Cδ0tDψm′′(vα)|Φ^α(v)|2dxds.\displaystyle\begin{split}&\int_{D}\psi_{m}(v^{\alpha})\,\mathrm{d}x+\frac{\lambda}{2}\int_{0}^{t}\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})|\nabla v^{\alpha}|^{2}\,\mathrm{d}x\,\mathrm{d}s\\ \leq{}&\int_{D}\psi_{m}(v^{\alpha}(0))\,\mathrm{d}x\\ &+\int_{0}^{t}\int_{D}\psi_{m}^{\prime}(v^{\alpha})\phi^{\alpha}(v)\,\mathrm{d}x\,\mathrm{d}s\\ &+\sum_{n}\int_{0}^{t}\int_{D}\psi_{m}^{\prime}(v^{\alpha})\bigl{[}b_{n,j}^{\alpha}(v^{\alpha})\partial_{j}v^{\alpha}+g_{n}^{\alpha}(v)\bigr{]}\,\mathrm{d}x\,\mathrm{d}w_{n}\\ &+C_{\varepsilon}\sum_{n}\int_{0}^{t}\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})|g_{n}^{\alpha}(v)|^{2}\,\mathrm{d}x\,\mathrm{d}s+C_{\delta}\int_{0}^{t}\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})|\widehat{\Phi}^{\alpha}(v)|^{2}\,\mathrm{d}x\,\mathrm{d}s.\end{split}

The first term on the right-hand side of (134) is controlled by u0Lq(D)q\|u_{0}\|_{\mathrm{L}^{q}(D)}^{q} by definition of ψm\psi_{m}. We come to the second term. Applying Lemma 8.3 with v=v(s,x)v=v(s,x), estimate

(135) 0tDψm(vα)ϕα(v)dxds1+0tDβψm(vβ)dxds.\displaystyle\int_{0}^{t}\int_{D}\psi_{m}^{\prime}(v^{\alpha})\phi^{\alpha}(v)\,\mathrm{d}x\,\mathrm{d}s\lesssim 1+\int_{0}^{t}\int_{D}\sum_{\beta}\psi_{m}(v^{\beta})\,\mathrm{d}x\,\mathrm{d}s.

Since gg is of linear growth, we get nψm′′(vα)|gnα(v)|2ψm′′(vα)(1+|v|2)\sum_{n}\psi_{m}^{\prime\prime}(v^{\alpha})|g_{n}^{\alpha}(v)|^{2}\lesssim\psi_{m}^{\prime\prime}(v^{\alpha})(1+|v|^{2}), so the proof of Lemma 8.3 reveals also

(136) 0tDψm′′(vα)|gnα(v)|2dxds1+0tDβψm(vβ)dxds.\displaystyle\int_{0}^{t}\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})|g_{n}^{\alpha}(v)|^{2}\,\mathrm{d}x\,\mathrm{d}s\lesssim 1+\int_{0}^{t}\int_{D}\sum_{\beta}\psi_{m}(v^{\beta})\,\mathrm{d}x\,\mathrm{d}s.

Since Φ^α\widehat{\Phi}^{\alpha} is Lipschitz, a similar estimate holds for ψm′′(vα)|Φ^α(v)|2\psi_{m}^{\prime\prime}(v^{\alpha})|\widehat{\Phi}^{\alpha}(v)|^{2}. So far, we have in summary

(137) Dψm(vα)dx+λ20tDψm′′(vα)|vα|2dxds1+u0Lq(D)q+0tDβψm(vβ)dxds+n0tDψm(vα)[bn,jα(vα)jvα+gnα(v)]dwnds.\displaystyle\begin{split}&\int_{D}\psi_{m}(v^{\alpha})\,\mathrm{d}x+\frac{\lambda}{2}\int_{0}^{t}\int_{D}\psi_{m}^{\prime\prime}(v^{\alpha})|\nabla v^{\alpha}|^{2}\,\mathrm{d}x\,\mathrm{d}s\\ \lesssim{}&1+\|u_{0}\|_{\mathrm{L}^{q}(D)}^{q}+\int_{0}^{t}\int_{D}\sum_{\beta}\psi_{m}(v^{\beta})\,\mathrm{d}x\,\mathrm{d}s\\ &\quad+\sum_{n}\int_{0}^{t}\int_{D}\psi_{m}^{\prime}(v^{\alpha})\bigl{[}b_{n,j}^{\alpha}(v^{\alpha})\partial_{j}v^{\alpha}+g_{n}^{\alpha}(v)\bigr{]}\,\mathrm{d}w_{n}\,\mathrm{d}s.\end{split}

We take the expectation, so that the stochastic integral vanishes, and sum in α\alpha afterwards, to find

(138) 𝔼Dαψm(vα)dx+λ2𝔼0tDαψm′′(vα)|vα|2dxds1+𝔼u0Lq(D)q+𝔼0tDαψm(vα)dxds.\displaystyle\begin{split}&\mathbb{E}\int_{D}\sum_{\alpha}\psi_{m}(v^{\alpha})\,\mathrm{d}x+\frac{\lambda}{2}\mathbb{E}\int_{0}^{t}\int_{D}\sum_{\alpha}\psi_{m}^{\prime\prime}(v^{\alpha})|\nabla v^{\alpha}|^{2}\,\mathrm{d}x\,\mathrm{d}s\\ \lesssim{}&1+\mathbb{E}\|u_{0}\|_{\mathrm{L}^{q}(D)}^{q}+\mathbb{E}\int_{0}^{t}\int_{D}\sum_{\alpha}\psi_{m}(v^{\alpha})\,\mathrm{d}x\,\mathrm{d}s.\end{split}

Note that for all t[0,T]t\in[0,T] one has 𝔼0tDαψm(vα(s))dxds\mathbb{E}\int_{0}^{t}\int_{D}\sum_{\alpha}\psi_{m}(v^{\alpha}(s))\,\mathrm{d}x\,\mathrm{d}s is finite, for ψm(v(s))|v(s)|2+C\psi_{m}(v(s))\leq|v(s)|^{2}+C. Therefore, Gronwall’s inequality is applicable with t𝔼Dαψm(vα(t))dxt\mapsto\mathbb{E}\int_{D}\sum_{\alpha}\psi_{m}(v^{\alpha}(t))\,\mathrm{d}x and yields

(139) 𝔼Dαψm(vα)dx1+𝔼u0Lq(D)q\displaystyle\mathbb{E}\int_{D}\sum_{\alpha}\psi_{m}(v^{\alpha})\,\mathrm{d}x\lesssim 1+\mathbb{E}\|u_{0}\|_{\mathrm{L}^{q}(D)}^{q}

for all t(0,T)t\in(0,T). Integrate this bound in tt, use Fubini’s theorem, and take (in virtue of the pointwise convergence of ψm\psi_{m} from Lemma 8.2) the limit mm\to\infty to conclude the first claim of the proposition. Finally, plug (139) back into (138) to deduce

(140) λ2𝔼0tDαψm′′(vα)|v|2dxds\displaystyle\frac{\lambda}{2}\mathbb{E}\int_{0}^{t}\int_{D}\sum_{\alpha}\psi_{m}^{\prime\prime}(v^{\alpha})|\nabla v|^{2}\,\mathrm{d}x\,\mathrm{d}s\lesssim{} 1+𝔼u0Lq(D)q.\displaystyle 1+\mathbb{E}\|u_{0}\|_{\mathrm{L}^{q}(D)}^{q}.

This time using the pointwise convergence of ψm′′\psi_{m}^{\prime\prime} stated in Lemma 8.2, take once more the limit mm\to\infty. Afterwards, take the limit tTt\to T to conclude the second claim in the proposition. ∎

Remark 8.5.

Even more is true in the last proposition: if we take first the supremum over tt in (137) and then expectations, a calculation based on the Burkholder–Davis–Gundy inequality even yields the estimate

(141) vLq(Ω;C([0,T];Lq(D)))1+u0Lq(Ω;Lq(D)).\displaystyle\|v\|_{\mathrm{L}^{q}(\Omega;\mathrm{C}([0,T];\mathrm{L}^{q}(D)))}\lesssim 1+\|u_{0}\|_{\mathrm{L}^{q}(\Omega;\mathrm{L}^{q}(D))}.

We will not need this extra information and decided therefore to stick to the shorter proof that leads to Proposition 8.4.

Proof of Theorem 6.6.

Write PRP_{R} for the projection onto the closed ball of radius RR around 0 in N\mathbb{R}^{N}. Then, for R>0R>0, define the non-linearities ϕαR(t,x,y)ϕα(t,x,PR(y)){}_{R}\phi^{\alpha}(t,x,y)\coloneqq\phi^{\alpha}(t,x,P_{R}(y)) and ΦiαR(t,x,y)Φiα(t,x,PR(y)){}_{R}\Phi^{\alpha}_{i}(t,x,y)\coloneqq\Phi^{\alpha}_{i}(t,x,P_{R}(y)). They are Lipschitz continuous and bounded (in particular, of linear growth) with constant depending on RR, and satisfy Assumptions 6.1 (iv) and (v) uniformly in RR by virtue of Lemma 8.1. Consider the equation

(R-QL) {duα=[i(aijαβ(u)juβ)+i(ΦiαR(u))+ϕαR(u)]dt+n1[bn,jαβ(uβ)juβ+gnα(u)]dwn, on D,uα=0, on D,uα(0)=u0α, on D.\displaystyle\left\{\quad\begin{aligned} \,\mathrm{d}u^{\alpha}&=\big{[}\partial_{i}(a^{\alpha\beta}_{ij}(u)\partial_{j}u^{\beta})+\partial_{i}({}_{R}\Phi^{\alpha}_{i}(u))+{}_{R}\phi^{\alpha}(u)\big{]}\,\mathrm{d}t\\ &\qquad\qquad+\sum_{n\geq 1}\big{[}b^{\alpha\beta}_{n,j}(u^{\beta})\partial_{j}u^{\beta}+g^{\alpha}_{n}(u)\big{]}\,\mathrm{d}w_{n},\ \ &\text{ on }D,\\ u^{\alpha}&=0,\ \ &\text{ on }\partial D,\\ u^{\alpha}(0)&=u^{\alpha}_{0},\ \ &\text{ on }D.\end{aligned}\right.

Problem (R-QL) fulfills Assumption 7.1. Therefore, by the well-posedness result of Section 7, there are p0>2p_{0}>2 and solutions

uRLp(Ω×(0,T);H01(D))Lp(Ω;C([0,T];B2,p,012/p(D)))u_{R}\in\mathrm{L}^{p}(\Omega\times(0,T);\mathrm{H}^{1}_{0}(D))\cap\mathrm{L}^{p}(\Omega;\mathrm{C}([0,T];\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D)))

to (R-QL).

Apply Proposition 8.4 to uRu_{R} to find uRLq(Ω×(0,T)×D)1+u0Lq(Ω×D)\|u_{R}\|_{\mathrm{L}^{q}(\Omega\times(0,T)\times D)}\lesssim 1+\|u_{0}\|_{\mathrm{L}^{q}(\Omega\times D)}. With the growth condition for the non-linearities we find, keeping phqph\leq q in mind,

(142) ϕαR(uR)Lp(Ω×(0,T);L2(D))+ΦiαR(uR)Lp(Ω×(0,T);L2(D))\displaystyle\|{}_{R}\phi^{\alpha}(u_{R})\|_{\mathrm{L}^{p}(\Omega\times(0,T);L^{2}(D))}+\|{}_{R}\Phi^{\alpha}_{i}(u_{R})\|_{\mathrm{L}^{p}(\Omega\times(0,T);L^{2}(D))}
(143) \displaystyle\lesssim{} 1+uRLq(Ω×(0,T)×D)h\displaystyle 1+\|u_{R}\|^{h}_{L^{q}(\Omega\times(0,T)\times D)}
(144) \displaystyle\lesssim{} 1+u0Lq(Ω×D)h.\displaystyle 1+\|u_{0}\|^{h}_{L^{q}(\Omega\times D)}.

Therefore, we can freeze uRu_{R}, ΦR{}_{R}\Phi and ϕR{}_{R}\phi in (R-QL) to obtain, as in Section 7.3, a linear problem to which Theorem 2.2 applies. This results in the uniform bounds

(145) uRLp(Ω;C([0,T];B2,p,012/p(D)))1+u0Lp(Ω;B2,p,012/p(D))+u0Lq(Ω×D)h,\displaystyle\|u_{R}\|_{\mathrm{L}^{p}(\Omega;\mathrm{C}([0,T];\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D)))}\lesssim 1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))}+\|u_{0}\|^{h}_{L^{q}(\Omega\times D)},

and for all θ[0,1/2)\theta\in[0,\nicefrac{{1}}{{2}}) the estimate

(146) uRLp(Ω;Hθ,p(0,T;H012θ(D)))θ1+u0Lp(Ω;B2,p,012/p(D))+u0Lq(Ω×D)h.\displaystyle\|u_{R}\|_{\mathrm{L}^{p}(\Omega;\mathrm{H}^{\theta,p}(0,T;\mathrm{H}^{1-2\theta}_{0}(D)))}\lesssim_{\theta}1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))}+\|u_{0}\|^{h}_{L^{q}(\Omega\times D)}.

If θ(1/p,1/2)\theta\in(\nicefrac{{1}}{{p}},\nicefrac{{1}}{{2}}), then one has moreover the estimate

(147) uRLp(Ω;Cθ1/p([0,T];H012θ(D)))θ1+u0Lp(Ω;B2,p,012/p(D))+u0Lq(Ω×D)h.\displaystyle\|u_{R}\|_{\mathrm{L}^{p}(\Omega;\mathrm{C}^{\theta-\nicefrac{{1}}{{p}}}([0,T];\mathrm{H}^{1-2\theta}_{0}(D)))}\lesssim_{\theta}1+\|u_{0}\|_{\mathrm{L}^{p}(\Omega;\mathrm{B}^{1-\nicefrac{{2}}{{p}}}_{2,p,0}(D))}+\|u_{0}\|^{h}_{L^{q}(\Omega\times D)}.

Then, the conclusion using the stochastic compactness argument works analogous, with one additional argument: first, by (142), the terms corresponding to Φ\Phi and ϕ\phi in (99) satisfy the same moment condition as before. Second, for the almost sure convergence, decompose

(148) ϕ(u~)ϕR(u~R)=(ϕ(u~)ϕ(u~R))+(ϕ(u~R)ϕR(u~R)),\displaystyle\phi(\widetilde{u})-{}_{R}\phi(\widetilde{u}_{R})=(\phi(\widetilde{u})-\phi(\widetilde{u}_{R}))+(\phi(\widetilde{u}_{R})-{}_{R}\phi(\widetilde{u}_{R})),

likewise for ΦiR{}_{R}\Phi_{i}. As seen before, by Vitali’s convergence theorem, it suffices to show (r,x)(r,x) almost everywhere convergence of ϕR(u~R),ξ\langle{}_{R}\phi(\widetilde{u}_{R}),\xi\rangle and ΦiR(u~R),iξ\langle{}_{R}\Phi_{i}(\widetilde{u}_{R}),\partial_{i}\xi\rangle, and upon passing to a subsequence, we can assume that u~R\widetilde{u}_{R} converges to u~\widetilde{u} for almost every (s,x)(s,x). Therefore, the second term on the right-hand side of (148) goes to zero since u~R\widetilde{u}_{R} is a bounded sequence for fixed (s,x)(s,x), so that eventually ϕ\phi and ϕR{}_{R}\phi coincide. For the first term on the right-hand side of (148), we upgrade almost everywhere convergence of the subsequence to Lph((0,T)×D)\mathrm{L}^{ph}((0,T)\times D) convergence using Vitali’s theorem (recall that q>ph)q>ph). Then the claim follows by continuity of the Nemytski operator [15, Thm. 3.2.24], that is to say, continuity of the operator vϕ(t,x,v)v\mapsto\phi(t,x,v) from Lph((0,T)×D)Lp((0,T)×D)\mathrm{L}^{ph}((0,T)\times D)\to\mathrm{L}^{p}((0,T)\times D). Here, we use that ϕ\phi is a Carathéodory function satisfying the uniform growth condition stated in Assumption 6.1 (iv). ∎

References

  • [1] A. Agresti and M. Veraar. Nonlinear parabolic stochastic evolution equations in critical spaces part I. Stochastic maximal regularity and local existence. Nonlinearity 35 (2022), no. 8, 4100–4210. doi:10.1088/1361-6544/abd613.
  • [2] A. Agresti and M. Veraar. The critical variational setting for stochastic evolution equations. Probab. Theory Related Fields 188 (2024), no. 3-4, 957–1015. doi:10.1007/s00440-023-01249-x.
  • [3] A. Agresti and M. Veraar. Stochastic maximal Lp(Lq)L^{p}(L^{q})-regularity for second order systems with periodic boundary conditions. Ann. Inst. Henri Poincaré Probab. Stat. 60 (2024), no. 1, 413–430. doi:10.1214/22-AIHP1333.
  • [4] A. Agresti and M. Veraar. Reaction–Diffusion equations with transport noise and critical superlinear diffusion: Global well-posedness of weakly dissipative systems.. SIAM J. Math. Anal. 56 (2024), no 4, pp 4870-4927. doi:10.1137/23M1562482.
  • [5] W. Arendt, C. Batty, M. Hieber, and F. Neubrander. Vector-valued Laplace Transforms and Cauchy Problems. Monographs in Mathematics, vol. 96, Birkhäuser, Basel-Boston-Berlin, 2001. doi:10.1007/978-3-0348-0087-7.
  • [6] A. Bensoussan and R. Temam. Équations aux dérivées partielles stochastiques non linéaires. Israel Journal of Mathematics 11 (1972), 95–129. doi:10.1007/BF02761449.
  • [7] J. Bergh and J. Löfström. Interpolation Spaces. An Introduction. Grundlehren der Mathematischen Wissenschaften, vol. 223. Springer, Berlin, 1976.
  • [8] T. Böhnlein and M. Egert. Explicit improvements for LpL^{p}-estimates related to elliptic systems. Bull. Lond. Math. Soc. 56 (2024), no. 3, 914–930. doi:10.1112/blms.12973.
  • [9] D. Breit, E. Feireisl, and M. Hofmanová. Stochastically forced compressible fluid flows. De Gruyter, Berlin, 2018. doi:10.1515/9783110492552.
  • [10] H. Brezis. Functional analysis, Sobolev spaces and partial differential equations. Universitext. Springer, New York, 2011. doi:10.1007/978-0-387-70914-7.
  • [11] Z. Brzeźniak and M. Veraar. Is the stochastic parabolicity condition dependent on pp and qq?. Electron. J. Probab. 17 (2012), 1–24. doi:10.1214/EJP.v17-2186.
  • [12] A. Debussche, M. Hofmanová, and J. Vovelle. Degenerate parabolic stochastic partial differential equations: quasilinear case. Ann. Probab. 44 (2016), no. 3, 1916–1955. doi:10.1214/15-AOP1013.
  • [13] L. Denis, A. Matoussi, and L. Stoica. LpL^{p} estimates for the uniform norm of solutions of quasilinear SPDE’s. Probab. Theory Relat. Fields 133 (2005), 437–463. doi:10.1007/s00440-005-0436-5.
  • [14] K. Disser, A. ter Elst, and J. Rehberg. On maximal parabolic regularity for non-autonomous parabolic operators. Journal of Differential Equations 262 (2017), 2039–2072. doi:10.1016/j.jde.2016.10.033.
  • [15] J. Drábek and J. Milota. Methods of nonlinear analysis. Birkhäuser/Springer Basel AG, Basel, 2013. doi:10.1007/978-3-0348-0387-8.
  • [16] M. Gnann, J. Hoogendijk, and M. Veraar. Higher order moments for SPDE with monotone nonlinearities. Stochastics 96 (2024), no. 7, 1948–1983. doi:10.1080/17442508.2024.2384554.
  • [17] T. Hytönen, J. van Neerven, M. Veraar, and L. Weis. Analysis in Banach spaces. Vol. I. Martingales and Littlewood-Paley theory. Ergebnisse der Mathematik und ihrer Grenzgebiete, vol. 63, Springer, Cham, 2016. doi:10.1007/978-3-319-48520-1.
  • [18] T. Hytönen, J. van Neerven, M. Veraar, and L. Weis. Analysis in Banach spaces. Vol. II. Probabilistic Methods and Operator Theory. Ergebnisse der Mathematik und ihrer Grenzgebiete, vol. 67, Springer, Cham, 2016. doi:10.1007/978-3-319-69808-3.
  • [19] T. Hytönen, J. van Neerven, M. Veraar, and L. Weis. Analysis in Banach spaces. Vol. III. Harmonic Analysis and Spectral Theory. Ergebnisse der Mathematik und ihrer Grenzgebiete, vol. 76 Springer, Cham, 2023. doi:10.1007/978-3-031-46598-7.
  • [20] N. Krylov. An analytic approach to SPDEs. In: Stochastic partial differential equations: six perspectives, Math. Surveys Monogr., vol. 64. Birkhäuser, Basel, 1999.
  • [21] N. Krylov and B. Rozovskii. Stochastic evolution equations. In: Current problems in mathematics, Vol. 14 (Russian). Akad. Nauk SSSR, Moscow, 1979.
  • [22] S. Lang. Real and functional analysis. Springer-Verlag, New York, 1993. doi:10.1007/978-1-4612-0897-6.
  • [23] J.-L. Lions. Quelques méthodes de résolution des problèmes aux limites non linéaires. Dunod, 1969.
  • [24] W. Liu and M. Röckner. Stochastic partial differential equations: an introduction. Springer, Cham, 2015. doi:10.1007/978-3-319-22354-4.
  • [25] E. Lorist and M. Veraar. Singular stochastic integral operators. Anal. PDE 14 (2021), no. 5, 1443–1507. doi:10.2140/apde.2021.14.1443.
  • [26] J. van Neerven, M. Veraar, and L. Weis. Stochastic maximal LpL^{p}-regularity. Ann. Probab. 40 (2012), no. 2, 788–812. doi:10.1214/10-AOP626.
  • [27] E.-M. Ouhabaz. Analysis of Heat Equations on Domains. London Mathematical Society Monographs Series, vol. 31, Princeton University Press, Princeton NJ, 2005. doi:10.1515/9781400826483.
  • [28] E. Pardoux. Équations aux dérivées partielles stochastiques nonlinéares monotones: étude de solutions fortes de type Itô. PhD thesis, Université Paris–Orsay, 1975.
  • [29] P. Portal and M. Veraar. Stochastic maximal regularity for rough time-dependent problems. Stoch PDE: Anal Comp 7 (2019), no. 4, 541–597. doi:10.1007/s40072-019-00134-w.
  • [30] M. Röckner, S. Shang, and T. Zhang. Well-posedness of stochastic partial differential equations with fully local monotone coefficients. Math. Ann.390 (2024), no. 3, 3419–3469. doi:10.1007/s00208-024-02836-6.
  • [31] B. Rozovskii. Stochastic evolution systems. Linear Theory and Applications to Non-Linear Filtering. Mathematics and its Applications (Soviet Series), Vol. 35. Kluwer Academic Publishers Group, Dordrecht, 1990. doi:10.1007/978-3-319-94893-5.
  • [32] H. Triebel. Interpolation Theory, Function Spaces, Differential Operators. North-Holland Mathematical Library, vol. 18, North-Holland Publishing, Amsterdam, 1978.
  • [33] J. Voigt. Abstract Stein interpolation. Math. Nachr. 157 (1992), 197–199. doi:10.1002/mana.19921570115.