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

Weak convergence of the backward Euler method for stochastic Cahn–Hilliard equation with additive noise

Meng Cai mcai1993@126.com Siqing Gan sqgan@csu.edu.cn Yaozhong Hu yaozhong@ualberta.ca School of Mathematics and Statistics, HNP-LAMA, Central South University, 410083, Hunan, China Department of Mathematical and Statistical Sciences, University of Alberta, T6G 2G1, Edmonton, Canada
Abstract

We prove a weak rate of convergence of a fully discrete scheme for stochastic Cahn–Hilliard equation with additive noise, where the spectral Galerkin method is used in space and the backward Euler method is used in time. Compared with the Allen–Cahn type stochastic partial differential equation, the error analysis here is much more sophisticated due to the presence of the unbounded operator in front of the nonlinear term. To address such issues, a novel and direct approach has been exploited which does not rely on a Kolmogorov equation but on the integration by parts formula from Malliavin calculus. To the best of our knowledge, the rates of weak convergence are revealed in the stochastic Cahn–Hilliard equation setting for the first time.

keywords:
stochastic Cahn–Hilliard equation , weak convergence rate , backward Euler method
MSC:
60H35 , 60H15 , 65C30
volume: 00
\journalname\runauth

M. Cai, S.Gan and Y. Hu \jidanms \jnltitlelogo \CopyrightLine2011Published by Elsevier Ltd.

\dochead

1 Introduction

During the last decades, there have been overwhelming activities on the analysis of numerical stochastic partial differential equation (SPDE) under globally Lipschitz condition and a fast growing number of studies on Allen–Cahn type SPDE with non-globally Lipschitz coefficients. However, numerical analysis of stochastic Cahn–Hilliard equation, which is another prominent SPDE model with non-globally Lipschitz coefficients, is at its beginning and is far from being well understood. The Cahn–Hilliard equation is of fundamental importance in various applications to, such as, the complicated phase separation and coarsening phenomena in a melted alloy [6, 8], spinodal decomposition for binary mixture [7], the diffusive process of populations and oil film spreading over a solid surface [12]. Our motivating example arises from a simplified mesoscopic physical model for phase separation. The aim of this article is to investigate the weak convergence rate of a full discretization for stochastic Cahn–Hilliard equation driven by additive noise,

{dX(t)+A(AX(t)+F(X(t)))dt=dW(t),t(0,T],X(0)=X0.\left\{\begin{array}[]{lll}\mathrm{d}X(t)+A(AX(t)+F(X(t)))\,\mathrm{d}t=\mathrm{d}W(t),\quad t\in(0,T],\\ X(0)=X_{0}.\end{array}\right. (1)

Let 𝐃\mathbf{D} be a bounded connected open domain of d,d=1,2,3\mathbb{R}^{d},d=1,2,3 with smooth boundary and let H:=L2(𝐃,)H:=L^{2}(\mathbf{D},\mathbb{R}) be the Hilbert space with the usual scalar product ,\langle\cdot,\cdot\rangle and norm \|\cdot\|. The space H˙:={vH:𝐃vdx=0}\dot{H}:=\{v\in H:\int_{\mathbf{D}}v\mathrm{d}x=0\} is a subspace of HH. We make the following assumptions.

Assumption 1.1.

A:dom(A)H˙H˙-A:\mathrm{dom}(A)\subseteq\dot{H}\to\dot{H} is the Neumann Laplacian defined by Au=Δu,udom(A)={uH2(𝐃)H˙:un=0on𝐃}-Au=\Delta u,u\in\mathrm{dom}(A)=\{u\in H^{2}(\mathbf{D})\cap\dot{H}:\frac{\partial u}{\partial n}=0\,\,\mathrm{on}\,\,\partial\mathbf{D}\}.

Assumption 1.2.

F:L6(𝐃,)HF:L^{6}(\mathbf{D},\mathbb{R})\rightarrow H is the Nemytskii operator given by

F(v)(x)=f(v(x))=v3(x)v(x),x𝐃,vL6(𝐃,).F(v)(x)=f(v(x))=v^{3}(x)-v(x),\quad x\in\mathbf{D},v\in L^{6}(\mathbf{D},\mathbb{R}). (2)
Assumption 1.3.

The noise process {W(t)}t[0,T]\{W(t)\}_{t\in[0,T]} is an H˙\dot{H}-valued QQ-Wiener process with the covariance operator QQ satisfying

A12Q122<.\big{\|}A^{\frac{1}{2}}Q^{\frac{1}{2}}\big{\|}_{\mathcal{L}_{2}}<\infty. (3)
Assumption 1.4.

The initial value X0X_{0} is deterministic and satisfies

|X0|4<,|X_{0}|_{4}<\infty, (4)

where the norm ||4|\cdot|_{4} is defined in (14) below.

We point out that Assumption 1.3 is the same as that in [21, 24, 28]. The assumption on the initial datum can be relaxed, but at the expense of having the constant CC depending on t1t^{-1}, by exploiting the smoothing effect of the semigroup E(t),t[0,T]E(t),t\in[0,T] and standard non–smooth data error estimates.

Based on the above assumptions and following the semigroup framework in [19], we see that the model (1) admits a unique mild solution

X(t)=E(t)X00tE(ts)APF(X(s))ds+0tE(ts)dW(s),t[0,T],\displaystyle X(t)=E(t)X_{0}-\int_{0}^{t}E(t-s)APF(X(s))\,\mathrm{d}s+\int_{0}^{t}E(t-s)\,\mathrm{d}W(s),\quad t\in[0,T],

where E(t)E(t) denotes the analytic semigroup generated by A2-A^{2}. We refer the readers to [3, 10, 14, 15, 18, 20, 27] for the existence and uniqueness of the mild solution for such equation. Since the exact solutions are rarely known explicitly, numerical simulations are often used to investigate the behavior of the solutions. We choose the spatial semi-discretization by the spectral Galerkin method, i.e., projecting the equation to vector space HNH_{N}, spanned by the first NN eigenvectors of AA. The approximated equation of (1) is in the form

dXN(t)+A(AXN(t)+PNF(XN(t)))dt=PNdW(t),t(0,T];XN(0)=PNX0,\mathrm{d}X^{N}(t)+A(AX^{N}(t)+P_{N}F(X^{N}(t)))\mathrm{d}t=P_{N}\mathrm{d}W(t),\,t\in(0,T];\,\,X^{N}(0)=P_{N}X_{0}\,,

where PNP_{N} is the spectral Galerkin projection operator onto the space HNH_{N}. In the temporal direction, we apply the backward Euler method to the above equation. The fully discrete scheme is then given by

XtmM,NXtm1M,N+τA2XtmM,N+τPNAF(XtmM,N)=PNΔWm,m{1,2,,M}.X_{t_{m}}^{M,N}-X_{t_{m-1}}^{M,N}+\tau A^{2}X_{t_{m}}^{M,N}+\tau P_{N}AF(X_{t_{m}}^{M,N})=P_{N}\Delta W_{m},\quad m\in\{1,2,\cdots,M\}.

Here ΔWm:=W(tm)W(tm1)\Delta W_{m}:=W(t_{m})-W(t_{m-1}), τ=TM\tau=\tfrac{T}{M} is the time stepsize and tm=mτt_{m}=m\tau. The main result, concerning the weak convergence rates of the full discretization, reads

|𝔼[Φ(X(T))]𝔼[Φ(XTM,N)]|C(λN2+τ),ΦCb2(H˙,).\big{|}\mathbb{E}[\Phi(X(T))]-\mathbb{E}[\Phi(X_{T}^{M,N})]\big{|}\leq C\big{(}\lambda_{N}^{-2}+\tau\big{)},\,\forall\,\Phi\in C_{b}^{2}(\dot{H},\mathbb{R}). (5)

Here and throughout this article, CC denotes a generic positive constant that is independent of the discretization parameters M,NM,N and may change from line to line and Cb2(H˙,)C_{b}^{2}(\dot{H},\mathbb{R}) (or Cb2C_{b}^{2}) represents the space of not necessarily bounded mappings from H˙\dot{H} to \mathbb{R} that have continuous and bounded Fréchet derivatives up to order 2. We split the weak error into two terms, both the spatial error and the temporal error, which are analyzed in Section 3 and Section 4, respectively. The result given by the above inequality (5) is on the weak rate of convergence. It is strictly greater than the strong ones (see Corollary 4.1) as expected. It is seen that the weak rate (which is 1.0 in time) is not twice as the strong one, contrary to the common belief. Indeed, the order is limited to 11 since an implicit Euler scheme is used.

The idea for error analysis to obtain (5) goes as follows. At first, the weak error is separated into two parts, the spatial error and the temporal error,

𝔼[Φ(X(T))]𝔼[Φ(XTM,N)]=(𝔼[Φ(X(T))]𝔼[Φ(XN(T))])+(𝔼[Φ(XN(T))]𝔼[Φ(XTM,N)]).\mathbb{E}\big{[}\Phi(X(T))\big{]}-\mathbb{E}\big{[}\Phi(X_{T}^{M,N})\big{]}\\ =\big{(}\mathbb{E}\big{[}\Phi(X(T))\big{]}-\mathbb{E}\big{[}\Phi(X^{N}(T))\big{]}\big{)}+\big{(}\mathbb{E}\big{[}\Phi(X^{N}(T))\big{]}-\mathbb{E}\big{[}\Phi(X_{T}^{M,N})\big{]}\big{)}. (6)

To simplify the notation, we often write 𝒪t\mathcal{O}_{t} for 0tE(tr)dW(r)\int_{0}^{t}E(t-r)\mathrm{d}W(r) and 𝒪tN=PN𝒪t\mathcal{O}^{N}_{t}=P_{N}\mathcal{O}_{t}. By introducing two processes X¯(t):=X(t)𝒪t\bar{X}(t):=X(t)-\mathcal{O}_{t} and X¯N(t):=XN(t)𝒪tN\bar{X}^{N}(t):=X^{N}(t)-\mathcal{O}_{t}^{N}, we can further split the spatial error as

𝔼[Φ(X(T))]𝔼[Φ(XN(T))]=(𝔼[Φ(X¯(T)+𝒪T)]𝔼[Φ(X¯N(T)+𝒪T)])+(𝔼[Φ(X¯N(T)+𝒪T)]𝔼[Φ(X¯N(T)+𝒪TN)]).\begin{split}\mathbb{E}\big{[}\Phi(X(T))\big{]}-\mathbb{E}\big{[}\Phi(X^{N}(T))\big{]}&=\big{(}\mathbb{E}\big{[}\Phi(\bar{X}(T)+\mathcal{O}_{T})\big{]}-\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T})\big{]}\big{)}\\ &\quad+\big{(}\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T})\big{]}-\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T}^{N})\big{]}\big{)}.\end{split} (7)

To proceed, one relies on the Taylor expansion of the test function Φ\Phi. The key argument to estimate the first term on the right hand of (7) is to bound the error between X¯N(T)\bar{X}^{N}(T) and X¯(T)\bar{X}(T) by that in a strong sense,

|𝔼[Φ(X¯(T)+𝒪T)]𝔼[Φ(X¯N(T)+𝒪T)]|C|𝔼01Φ(X(T)+λ(X¯N(T)X¯(T)))(X¯N(T)X¯(T))dλ|CX¯(T)PNX¯(T)L2(Ω,H˙)+CPNX¯(T)X¯N(T)L2(Ω,H˙).\displaystyle\begin{split}&\big{|}\mathbb{E}\big{[}\Phi(\bar{X}(T)+\mathcal{O}_{T})\big{]}-\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T})\big{]}\big{|}\\ &\quad\leq C\Big{|}\mathbb{E}\int_{0}^{1}\Phi^{\prime}\big{(}X(T)+\lambda(\bar{X}^{N}(T)-\bar{X}(T))\big{)}\big{(}\bar{X}^{N}(T)-\bar{X}(T)\big{)}\mathrm{d}\lambda\Big{|}\\ &\quad\leq C\,\|\bar{X}(T)-P_{N}\bar{X}(T)\|_{L^{2}(\Omega,\dot{H})}+C\,\|P_{N}\bar{X}(T)-\bar{X}^{N}(T)\|_{L^{2}(\Omega,\dot{H})}.\end{split} (8)

The error term X¯(T)PNX¯(T)L2(Ω,H˙)\|\bar{X}(T)-P_{N}\bar{X}(T)\|_{L^{2}(\Omega,\dot{H})} can be easily controlled owing to the higher spatial regularity of the stochastic process X¯(T)\bar{X}(T), in the absence of the stochastic convolution. The remaining term eN(t):=PNX¯(t)X¯N(t)e^{N}(t):=P_{N}\bar{X}(t)-\bar{X}^{N}(t), satisfying the following random PDE,

ddteN(t)+A2eN(t)+PNA[F(X(t))F(XN(t))]=0,eN(0)=0,\tfrac{\mathrm{d}}{\mathrm{d}t}e^{N}(t)+A^{2}e^{N}(t)+P_{N}A\big{[}F(X(t))-F(X^{N}(t))\big{]}=0,\quad e^{N}(0)=0, (9)

must be carefully treated due to the presence of the unbounded operator AA before the nonlinear term FF. We use the monotonicity of the nonlinearity of FF and the regularities of X(T)X(T), XN(T)X^{N}(T) and 𝒪t\mathcal{O}_{t} to derive 0T|eN(t)|12dtLp(Ω,)CλN4\Big{\|}\int_{0}^{T}|e^{N}(t)|_{1}^{2}\mathrm{d}t\Big{\|}_{L^{p}(\Omega,\mathbb{R})}\leq C\,\lambda_{N}^{-4}. Then, combining it with the mild solution of (9) leads to the desired weak orders (c.f. (69)-(73) below). Subsequently, we turn our attention to the second term in (7). Applying the Taylor expansion gives

|𝔼[Φ(X¯N(T)+𝒪T)]𝔼[Φ(X¯N(T)+𝒪TN)]||𝔼[Φ(XN(T))(𝒪T𝒪TN)]|+|𝔼[01Φ′′(XN(T)+λ(𝒪T𝒪TN))(𝒪T𝒪TN,𝒪T𝒪TN)(1λ)dλ]|.\displaystyle\begin{split}&\big{|}\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T})\big{]}-\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T}^{N})\big{]}\big{|}\leq\Big{|}\mathbb{E}\big{[}\Phi^{{}^{\prime}}(X^{N}(T))(\mathcal{O}_{T}-\mathcal{O}_{T}^{N})\big{]}\Big{|}\\ &\quad+\Big{|}\mathbb{E}\Big{[}\int_{0}^{1}\Phi^{{}^{\prime\prime}}(X^{N}(T)+\lambda(\mathcal{O}_{T}-\mathcal{O}_{T}^{N}))(\mathcal{O}_{T}-\mathcal{O}_{T}^{N},\mathcal{O}_{T}-\mathcal{O}_{T}^{N})(1-\lambda)\mathrm{d}\lambda\Big{]}\Big{|}.\end{split} (10)

The Malliavin integration by parts formula is the key ingredient to deal with the first term (c.f. (77)) and the second term can be easily estimated due to the boundedness of Φ′′\Phi^{\prime\prime}. It is now easy to explain why the weak rate of convergence is expected to be higher than strong convergence rate. As a byproduct of the weak error analysis, one can easily obtain the rate of the strong error,

X(t)XN(t)L2(Ω,H˙)X¯(t)X¯N(t)L2(Ω,H˙)+𝒪t𝒪tNL2(Ω,H˙)CλN32,\|X(t)-X^{N}(t)\|_{L^{2}(\Omega,\dot{H})}\leq\|\bar{X}(t)-\bar{X}^{N}(t)\|_{L^{2}(\Omega,\dot{H})}+\|\mathcal{O}_{t}-\mathcal{O}^{N}_{t}\|_{L^{2}(\Omega,\dot{H})}\leq C\lambda_{N}^{-\frac{3}{2}}, (11)

which is consistent with the results in [16, 27] and is lower than the weak convergence rate in (5), due to the presence of the second error. The basic idea to estimate temporal error is the same as that of the spatial error by essentially exploiting the discrete analogue of the arguments. The main point is that error must be uniform on the spatially discrete parameter NN.

Having sketched the central ideas of the weak error analysis, we review some relevant results in the literature. For the linearized stochastic Cahn–Hilliard equations, we refer to [11, 23, 25] for some strong convergence results of the finite element method. The authors in [21, 24] studied the strong convergence of the fully discrete finite element approximation for Cahn–Hilliard–Cook equation under spatial regular noise, but with no rates obtained. Later, the authors in [28] derives strong convergence rates of the mixed finite element method by using a priori strong moment bounds of the numerical approximations. For unbounded noise diffusion, the existence and regularity of solution have been investigated in [3, 14] and the absolute continuity has been studied in [2, 15]. Recently, the strong convergence rates of the spatial spectral Galerkin method and the temporal accelerated implicit Euler method for the stochastic Cahn–Hilliard equation were obtained in [16]. For weak convergence analysis in the non–globally Lipschitz setting, we are only aware of the papers [5, 9, 13, 17] concerning the stochastic Allen–Cahn equation. To the best of our knowledge, the weak convergence rates of a fully discrete method for the stochastic Cahn–Hilliard equation are absent in the literature. It is worthwhile to point out that issues from the presence of the unbounded operator in front of the nonlinear term make the weak error analysis much more challenging. To be more specific, in addition to the aforementioned difficulty in the weak analysis, the estimate of the Malliavin derivative for the spatial approximation process is also completely different, much more efforts are needed (c.f. Proposition 3.2). More recently, while this work was under review, we were aware of the preprint [4] posted in arXiv, concerning with numerical approximations of similar SPDEs, where Bréhier, Cui and Wang provide weak error estimates for another class of numerical schemes, whose weak order is twice as the strong order, for less regular problems. It is worth mentioning that the approach in the two works are substantially different. Different methods and different regularity regimes are dealt with.

The outline of the article is as follows. In the next section, we present some preliminaries, including the well-posedness and regularity of the mild solution and give a brief introduction to Malliavin calculus. Section 3 is devoted to the weak analysis of the spectral Galerkin method in space and Section 4 is concerned with the weak convergence rates of the backward Euler method in time.

2 Preliminaries

In this section, the mathematical setting, well-posedness and regularity of the model and a brief introduction to Malliavin calculus are given.

2.1 Mathematical setting

Given two real separable Hilbert spaces (H,,,)(H,\langle\cdot,\cdot\rangle,\|\cdot\|) and (U,,U,U)(U,\langle\cdot,\cdot\rangle_{U},\|\cdot\|_{U}), (U,H)\mathcal{L}(U,H) stands for the space of all bounded linear operators from UU to HH with the operator norm (U,H)\|\cdot\|_{\mathcal{L}(U,H)} and 2(U,H)((U,H)\mathcal{L}_{2}(U,H)(\subset\mathcal{L}(U,H)) denotes the space of all Hilbert-Schmidt operators from UU to HH. For simplicity, we write (H)\mathcal{L}(H) and 2(H)\mathcal{L}_{2}(H) (or 2\mathcal{L}_{2} for short) instead of (H,H)\mathcal{L}(H,H) and 2(H,H)\mathcal{L}_{2}(H,H), respectively. It is known, see e.g., [19], that 2(U,H)\mathcal{L}_{2}(U,H) is a Hilbert space equipped with the inner product and norm,

T1,T22(U,H)=i+T1ϕi,T2ϕi,T2(U,H)=(i+Tϕi2)12,\displaystyle\left<T_{1},T_{2}\right>_{\mathcal{L}_{2}(U,H)}=\sum_{i\in\mathbb{N}^{+}}\left<T_{1}\phi_{i},T_{2}\phi_{i}\right>,\;\|T\|_{\mathcal{L}_{2}(U,H)}=\Big{(}\sum_{i\in\mathbb{N}^{+}}\|T\phi_{i}\|^{2}\Big{)}^{\frac{1}{2}}, (12)

where {ϕi}\{\phi_{i}\} is an arbitrary orthonormal basis of UU. Let H=L2(𝐃,)H=L^{2}(\mathbf{D},\mathbb{R}) and H˙={vH:v,1=0}\dot{H}=\{v\in H:\langle v,1\rangle=0\}. V:=C(𝐃,)V:=C(\mathbf{D},\mathbb{R}) denotes the Banach space of all continuous functions with supremum norm V\|\cdot\|_{V} and Lr(𝐃,):={f:𝐃,𝐃|f(x)|r𝑑x<}L^{r}(\mathbf{D},\mathbb{R}):=\{f:\mathbf{D}\to\mathbb{R},\int_{\mathbf{D}}|f(x)|^{r}dx<\infty\}. We define P:HH˙P:H\rightarrow\dot{H} the generalized orthogonal projection by Pv=v|𝐃|1𝐃vdxPv=v-|\mathbf{D}|^{-1}\int_{\mathbf{D}}v\mathrm{d}x, then (IP)v=|𝐃|1𝐃vdx(I-P)v=|\mathbf{D}|^{-1}\int_{\mathbf{D}}v\mathrm{d}x is the average of vv over 𝐃\mathbf{D}.

It is easy to check that AA is a positive definite, self-adjoint and unbounded linear operator on H˙\dot{H} with compact inverse. For any vHv\in H, we define Av=APvAv=APv, then there exists a family of eigenpairs {ej,λj}j\{e_{j},\lambda_{j}\}_{j\in\mathbb{N}} such that

Aej=λjejand0=λ0<λ1λ2λjwithλj,\displaystyle Ae_{j}=\lambda_{j}e_{j}\quad\text{and}\quad 0=\lambda_{0}<\lambda_{1}\leq\lambda_{2}\leq\cdots\leq\lambda_{j}\leq\cdots\quad\text{with}\quad\lambda_{j}\rightarrow\infty, (13)

where e0=|𝐃|12e_{0}=|\mathbf{D}|^{-\frac{1}{2}} and {ej,j=1,}\{e_{j},j=1,\cdots\} forms an orthonormal basis of H˙\dot{H}. Straightforward applications of the spectral theory yield the fractional powers of AA on H˙\dot{H}, e.g., Aαv=j=1λjαv,ejejA^{\alpha}v=\sum_{j=1}^{\infty}\lambda_{j}^{\alpha}\langle v,e_{j}\rangle e_{j}, α\alpha\in\mathbb{R}, vH˙v\in\dot{H}. The space H˙α=dom(Aα2),α\dot{H}^{\alpha}=\mathrm{dom}(A^{\frac{\alpha}{2}}),\alpha\in\mathbb{R} is a Hilbert space with the inner product ,α\langle\cdot,\cdot\rangle_{\alpha} and the associated norm ||α|\cdot|_{\alpha} given by

v,wα=j=1λjαv,ejw,ej,|v|α=Aα2v=(j=1λjα|v,ej|2)12.\langle v,w\rangle_{\alpha}=\sum_{j=1}^{\infty}\lambda_{j}^{\alpha}\langle v,e_{j}\rangle\langle w,e_{j}\rangle,\quad|v|_{\alpha}=\|A^{\frac{\alpha}{2}}v\|=\Big{(}\sum_{j=1}^{\infty}\lambda_{j}^{\alpha}|\langle v,e_{j}\rangle|^{2}\Big{)}^{\frac{1}{2}}. (14)

We also define uα=(|u|α2+|u,e0|2)12\|u\|_{\alpha}=\big{(}|u|_{\alpha}^{2}+|\langle u,e_{0}\rangle|^{2}\big{)}^{\frac{1}{2}} for uHu\in H and the corresponding space is Hα:={uH:uα<}.H^{\alpha}:=\{u\in H:\|u\|_{\alpha}<\infty\}. A basic fact shows that for α=1,2\alpha=1,2, the norm ||α|\cdot|_{\alpha} on H˙α\dot{H}^{\alpha} is equivalent to the standard Sobolev norm Hα(𝐃)\|\cdot\|_{H^{\alpha}(\mathbf{D})} (see [22, Theorems 2.9, 2.12] and [30, Theorem 16.9]). Since H2(𝐃)H^{2}(\mathbf{D}) is an algebra, there is a constant C>0C>0 such that, for any f,gH˙2f,g\in\dot{H}^{2},

fgH2(𝐃)CfH2(𝐃)gH2(𝐃)C|f|2|g|2.\|fg\|_{H^{2}(\mathbf{D})}\leq C\|f\|_{H^{2}(\mathbf{D})}\|g\|_{H^{2}(\mathbf{D})}\leq C|f|_{2}|g|_{2}. (15)

We recall that the operator A2-A^{2} generates an analytic semigroup E(t)=etA2E(t)=e^{-tA^{2}} on HH due to (13) and we have

E(t)v=etA2v=PetA2v+(IP)v,vH.\displaystyle\begin{split}E(t)v=e^{-tA^{2}}v=Pe^{-tA^{2}}v+(I-P)v,\quad v\in H.\end{split} (16)

With the aid of the eigenbasis of AA and Parseval’s identity, we have

AμE(t)(H˙)\displaystyle\|A^{\mu}E(t)\|_{\mathcal{L}(\dot{H})} Ctμ2,t>0,μ0,\displaystyle\leq Ct^{-\frac{\mu}{2}},\;t>0,\;\mu\geq 0, (17)
Aν(IE(t))(H˙)\displaystyle\|A^{-\nu}(I-E(t))\|_{\mathcal{L}(\dot{H})} Ctν2,t0,ν[0,2],\displaystyle\leq Ct^{\frac{\nu}{2}},\quad t\geq 0,\;\nu\in[0,2], (18)
t1t2AϱE(s)v2ds\displaystyle\int_{t_{1}}^{t_{2}}\|A^{\varrho}E(s)v\|^{2}\,\mathrm{d}s C|t2t1|1ϱv2,vH˙,ϱ[0,1],\displaystyle\leq C|t_{2}-t_{1}|^{1-\varrho}\|v\|^{2},\;\forall v\in\dot{H},\varrho\in[0,1], (19)
A2ρt1t2E(t2σ)vdσ\displaystyle\Big{\|}A^{2\rho}\int_{t_{1}}^{t_{2}}E(t_{2}-\sigma)v\,\mathrm{d}\sigma\Big{\|} C|t2t1|1ρv,vH˙,ρ[0,1].\displaystyle\leq C|t_{2}-t_{1}|^{1-\rho}\|v\|,\;\forall v\in\dot{H},\rho\in[0,1]. (20)

By Assumption 1.2, there exists a constant C>0C>0 such that

F(u)F(v),uv\displaystyle-\langle F(u)-F(v),u-v\rangle uv2,u,vL6(𝐃,),\displaystyle\leq\|u-v\|^{2},\quad u,v\in L^{6}(\mathbf{D},\mathbb{R}), (21)
F(u)F(v)\displaystyle\|F(u)-F(v)\| C(1+uV2+vV2)uv,u,vV.\displaystyle\leq C(1+\|u\|_{V}^{2}+\|v\|_{V}^{2})\|u-v\|,\quad u,v\in V. (22)

2.2 Well-posedness and regularity results of the model

First at all, similar to [16, (2.5) &(2.7)\&\,\,(2.7)], we give the following lemma concerning the spatio-temporal regularity result of stochastic convolution 𝒪t:=0tE(ts)dW(s)\mathcal{O}_{t}:=\int_{0}^{t}E(t-s)\mathrm{d}W(s).

Lemma 2.1.

Suppose Assumptions 1.1 and 1.3 hold. Then for all p1p\geq 1, the stochastic convolution 𝒪t\mathcal{O}_{t} satisfies

𝔼[supt[0,T]|𝒪t|Vp]+supt[0,T]𝔼[|𝒪t|3p]<,\mathbb{E}\Big{[}\sup_{t\in[0,T]}|\mathcal{O}_{t}|_{V}^{p}\Big{]}+\sup_{t\in[0,T]}\mathbb{E}\Big{[}|\mathcal{O}_{t}|_{3}^{p}\Big{]}<\infty, (23)

and for α[0,3]\alpha\in[0,3],

𝒪t𝒪sLp(Ω,H˙α)C|ts|min{12,3α4}.\|\mathcal{O}_{t}-\mathcal{O}_{s}\|_{L^{p}(\Omega,\dot{H}^{\alpha})}\leq C|t-s|^{\text{min}\{\frac{1}{2},\frac{3-\alpha}{4}\}}. (24)

The following theorem states the well-posedness and spatio-temporal regularity of the mild solution for stochastic Cahn-Hilliard equation (1), whose proofs can be found for example in [28, Theorem 2.1 & Theorem 2.2].

Theorem 2.1 (Well-posedness and regularity of the mild solution).

Under Assumptions 1.1-1.4, there is a unique mild solution of (1) satisfying

X(t)=E(t)X00tE(ts)AF(X(s))ds+0tE(ts)dW(s),t[0,T].\displaystyle X(t)=E(t)X_{0}-\int_{0}^{t}\!\!E(t-s)AF(X(s))\,\mathrm{d}s+\int_{0}^{t}\!\!E(t-s)\mathrm{d}W(s),\,t\in[0,T]. (25)

Furthermore, for p1p\geq 1,

supt[0,T]X(t)Lp(Ω,H˙3)<,\displaystyle\sup_{t\in[0,T]}\|X(t)\|_{L^{p}(\Omega,\dot{H}^{3})}<\infty, (26)

and for any α[0,3]\alpha\in[0,3],

X(t)X(s)Lp(Ω,H˙α)C(ts)min{12,3α4}, 0s<tT.\displaystyle\|X(t)-X(s)\|_{L^{p}(\Omega,\dot{H}^{\alpha})}\leq C(t-s)^{\text{min}\{\frac{1}{2},\frac{3-\alpha}{4}\}},\,0\leq s<t\leq T. (27)

Combining (26) and (15) yields the following result.

Corollary 2.1.

If Assumptions 1.1-1.4 are valid, then for all p1p\geq 1,

supt[0,T]F(X(t))Lp(Ω,H˙2)<.\sup_{t\in[0,T]}\|F(X(t))\|_{L^{p}(\Omega,\dot{H}^{2})}<\infty. (28)

2.3 Introduction to Malliavin calculus

A brief introduction to Malliavin calculus is given in this subsection. For more details, one can consult the classical monograph [26]. Define a Hilbert space U0=Q12(H˙)U_{0}=Q^{\frac{1}{2}}(\dot{H}) with inner product u,vU0=Q12u,Q12v\langle u,v\rangle_{U_{0}}=\langle Q^{-\frac{1}{2}}u,Q^{-\frac{1}{2}}v\rangle. Let 𝒢:L2([0,T],U0)L2(Ω)\mathcal{G}:L^{2}([0,T],U_{0})\rightarrow L^{2}(\Omega) be an isonormal Gaussian process. More precisely, for any deterministic mapping ϕL2([0,T],U0)\phi\in L^{2}([0,T],U_{0}), 𝒢(ϕ)\mathcal{G}(\phi) is centered Gaussian with the covariance structure

𝔼[𝒢(ϕ1)𝒢(ϕ2)]=ϕ1,ϕ2L2([0,T],U0),ϕ1,ϕ2L2([0,T],U0).\mathbb{E}\big{[}\mathcal{G}(\phi_{1})\mathcal{G}(\phi_{2})\big{]}=\langle\phi_{1},\phi_{2}\rangle_{L^{2}([0,T],U_{0})},\,\,\phi_{1},\phi_{2}\in L^{2}([0,T],U_{0}). (29)

For example (see e.g., [1]), we define the cylindrical QQ-Wiener process

W(t)u=𝒢(χ[0,t]u),uU0,t[0,T].W(t)u=\mathcal{G}(\chi_{[0,t]}\otimes u),\,u\in U_{0},\,t\in[0,T]. (30)

Given uU0u\in U_{0}, the process W(t)u,t[0,T]W(t)u,t\in[0,T], is a Brownian motion and we have

𝔼[W(t)uW(s)v]=min{s,t}u,vU0,u,vU0.\mathbb{E}[W(t)uW(s)v]=\text{min}\{s,t\}\langle u,v\rangle_{U_{0}},\,\,u,v\in U_{0}. (31)

Let Cp(M,)C_{p}^{\infty}(\mathbb{R}^{M},\mathbb{R}) be the space of all CC^{\infty}-mappings with polynomial growth. We define the family of smooth H˙\dot{H}-valued cylindrical random variables as

𝒮(H)={G=i=1Ngi(𝒢(ϕ1),,𝒢(ϕM))hi:ϕ1,,ϕML2([0,T],U0),giCp(M,),hiH˙,i{1,,N}}.\mathcal{S}(H)=\Big{\{}G=\sum_{i=1}^{N}g_{i}\big{(}\mathcal{G}(\phi_{1}),\ldots,\mathcal{G}(\phi_{M})\big{)}h_{i}:\phi_{1},\cdots,\phi_{M}\in L^{2}([0,T],U_{0}),\,g_{i}\in C_{p}^{\infty}(\mathbb{R}^{M},\mathbb{R}),\,h_{i}\in\dot{H},\,i\in\{1,\cdots,N\}\Big{\}}. (32)

The Malliavin derivative of G𝒮(H˙)G\in\mathcal{S}(\dot{H}) is an element of 2(U0,H˙)\mathcal{L}_{2}(U_{0},\dot{H}) and given by

𝒟tG:=i=1Nj=1Mjgi(𝒢(ϕ1),,𝒢(ϕM))hiϕj(t),\mathcal{D}_{t}G:=\sum_{i=1}^{N}\sum_{j=1}^{M}\partial_{j}g_{i}\big{(}\mathcal{G}(\phi_{1}),\ldots,\mathcal{G}(\phi_{M})\big{)}h_{i}\otimes\phi_{j}(t), (33)

where hiϕj(t)h_{i}\otimes\phi_{j}(t) denotes the tensor product, that is, for 1jM1\leq j\leq M and 1iN1\leq i\leq N,

(hiϕj(t))(u)=ϕj(t),uU0hiH˙,uU0,hiH˙,t[0,T].\big{(}h_{i}\otimes\phi_{j}(t)\big{)}(u)=\langle\phi_{j}(t),u\rangle_{U_{0}}h_{i}\in\dot{H},\quad\forall\,\,u\in U_{0},~{}h_{i}\in\dot{H},~{}t\in[0,T]. (34)

If GG is t\mathcal{F}_{t}-measurable, then 𝒟sG=0{\mathcal{D}}_{s}G=0 for s>ts>t. The derivative operator 𝒟{\mathcal{D}} is known to be closable and we define 𝔻1,2(H˙)\mathbb{D}^{1,2}(\dot{H}) as the closure of 𝒮(H˙)\mathcal{S}(\dot{H}) with respect to the norm

G𝔻1,2(H˙)=(𝔼[G2]+𝔼0T𝒟tG2(U0,H˙)2dt)12.\|G\|_{\mathbb{D}^{1,2}(\dot{H})}=\Bigl{(}\mathbb{E}\big{[}\|G\|^{2}\big{]}+\mathbb{E}\int_{0}^{T}\|{\mathcal{D}}_{t}G\|_{\mathcal{L}_{2}(U_{0},\dot{H})}^{2}\mathrm{d}t\Bigr{)}^{\frac{1}{2}}. (35)

We are now ready to give the Malliavin integration by parts formula. For any G𝔻1,2(H˙)G\in\mathbb{D}^{1,2}(\dot{H}) and an adapted process ΨL2([0,T]×Ω,2(U0,H˙))\Psi\in L^{2}([0,T]\times\Omega,\mathcal{L}_{2}(U_{0},\dot{H})),

𝔼[0TΨ(t)dW(t),G]=𝔼[0TΨ(t),𝒟tG2(U0,H˙)dt],\mathbb{E}\left[\left\langle\int_{0}^{T}\Psi(t)\mathrm{d}W(t),G\right\rangle\right]=\mathbb{E}\left[\int_{0}^{T}\left\langle\Psi(t),\mathcal{D}_{t}G\right\rangle_{\mathcal{L}_{2}(U_{0},\dot{H})}\mathrm{d}t\right], (36)

where the stochastic integral is Itô integral. To simplify the notation, we often write 20\mathcal{L}_{2}^{0} instead of 2(U0,H˙)\mathcal{L}_{2}(U_{0},\dot{H}). Next, we define 𝒟suG=𝒟sG,u{\mathcal{D}}_{s}^{u}G=\langle{\mathcal{D}}_{s}G,u\rangle the derivative in the direction uU0u\in U_{0}. Then the Malliavin derivative acting on the Itô integral 0tΨ(r)dW(r)\int_{0}^{t}\Psi(r)\mathrm{d}W(r) satisfies for all uU0u\in U_{0},

𝒟su0tΨ(r)dW(r)=0t𝒟suΨ(r)dW(r)+Ψ(s)u,0stT.\mathcal{D}_{s}^{u}\int_{0}^{t}\Psi(r)\mathrm{d}W(r)=\int_{0}^{t}\mathcal{D}_{s}^{u}\Psi(r)\mathrm{d}W(r)+\Psi(s)u,\quad 0\leq s\leq t\leq T. (37)

Given another separable Hilbert space \mathcal{H}, if σCb1(H˙,)\sigma\in C_{b}^{1}(\dot{H},\mathcal{H}) and G𝔻1,2(H˙)G\in\mathbb{D}^{1,2}(\dot{H}), then σ(G)𝔻1,2()\sigma(G)\in\mathbb{D}^{1,2}(\mathcal{H}) and the chain rule holds as 𝒟t(σ(G))=σ(G)𝒟tG{\mathcal{D}}_{t}(\sigma(G))=\sigma^{\prime}(G)\mathcal{D}_{t}G.

3 Weak convergence rate of the spectral Galerkin method

This section is devoted to the weak error analysis of the spatial spectral Galerkin semi-discretization. In the beginning, we define a finite dimension space HN=span{e1,,eN}H_{N}=\mathrm{span}\{e_{1},\cdots,e_{N}\} and the projection operator PN:H˙βHNP_{N}:\dot{H}^{\beta}\to H_{N} by PNx=i=1Nx,eieiP_{N}x=\sum_{i=1}^{N}\langle x,e_{i}\rangle e_{i} for all xH˙β,βx\in\dot{H}^{\beta},\beta\in\mathbb{R}. As a result, AA commutes with PNP_{N} and

(PNI)xCλNβ2|x|β,β0.\big{\|}\big{(}P_{N}-I\big{)}x\big{\|}\leq C\lambda_{N}^{-\frac{\beta}{2}}|x|_{\beta},\quad\forall~{}\beta\geq 0. (38)

Applying the spectral Galerkin approximation to (1) results in the finite-dimensional stochastic differential equation, given by

dXN(t)+A2XN(t)dt+APNF(XN(t))dt=PNdW(t),t(0,T];XN(0)=PNX0,\mathrm{d}X^{N}(t)+A^{2}X^{N}(t)\mathrm{d}t+AP_{N}F(X^{N}(t))\mathrm{d}t=P_{N}\mathrm{d}W(t),\,t\in(0,T];\,\,X^{N}(0)=P_{N}X_{0}, (39)

whose unique solution, in the mild form, is written as

XN(t)=E(t)PNX00tE(ts)APNF(XN(s))ds+0tE(ts)PNdW(s).X^{N}(t)=E(t)P_{N}X_{0}-\int_{0}^{t}E(t-s)AP_{N}F(X^{N}(s))\mathrm{d}s+\int_{0}^{t}E(t-s)P_{N}\mathrm{d}W(s). (40)

Similarly to Lemma 2.1, the spatio-temporal regularity of the discrete stochastic convolution 0tE(ts)PNdW(s)\int_{0}^{t}E(t-s)P_{N}\mathrm{d}W(s) (𝒪tN\mathcal{O}_{t}^{N} for short) (see e.g., [27]) enjoys

supNsupt[0,T]𝔼[|𝒪tN|3p]<,p1,\sup_{N\in\mathbb{N}}\sup_{t\in[0,T]}\mathbb{E}\Big{[}|\mathcal{O}_{t}^{N}|_{3}^{p}\Big{]}<\infty,\,\forall p\geq 1, (41)

and for α[0,3]\alpha\in[0,3],

supN𝒪tN𝒪sNLp(Ω,H˙α)C|ts|min{12,3α4},p1.\sup_{N\in\mathbb{N}}\big{\|}\mathcal{O}_{t}^{N}-\mathcal{O}_{s}^{N}\|_{L^{p}(\Omega,\dot{H}^{\alpha})}\leq C|t-s|^{\text{min}\{\frac{1}{2},\frac{3-\alpha}{4}\}},\,\forall p\geq 1. (42)

It has to be noted that essential difficulties exist for analyzing a finite element method for the considered SPDE. Indeed, the orthogonal projection PhP_{h} can not commute with operator AA, although PNP_{N} commutes with AA. Moreover, compared with finite difference method, the spectral Galerkin method admits a simpler analysis, whose approximated solution is smooth and allows better control of the Lipschitz constant. The proof of the following regularity results is given in [27, Lemma 3.4].

Proposition 3.1 (Spatio-temporal regularity of spatial semi-discretization).

If Assumptions 1.1-1.4 are satisfied, then the mild solution of the spatial approximation process (40) admits for all p1p\geq 1,

supN𝔼[supt[0,T]XN(t)L6(𝐃,)p]<.\sup_{N\in\mathbb{N}}\mathbb{E}\big{[}\sup_{t\in[0,T]}\|X^{N}(t)\|_{L^{6}(\mathbf{D},\mathbb{R})}^{p}\big{]}<\infty. (43)

Moreover, we have

supNsupt[0,T]XN(t)Lp(Ω,H˙3)<,\sup_{N\in\mathbb{N}}\sup_{t\in[0,T]}\|X^{N}(t)\|_{L^{p}(\Omega,\dot{H}^{3})}<\infty, (44)

and for any α[0,3]\alpha\in[0,3],

supNXN(t)XN(s)Lp(Ω,H˙α)C(ts)min{12,3α4}, 0s<tT.\sup_{N\in\mathbb{N}}\|X^{N}(t)-X^{N}(s)\|_{L^{p}(\Omega,\dot{H}^{\alpha})}\leq C(t-s)^{\text{min}\{\frac{1}{2},\frac{3-\alpha}{4}\}},\,0\leq s<t\leq T. (45)

Combining (44) and (15) gives the next result.

Corollary 3.1.

Under Assumptions 1.1-1.4,

supt[0,T]F(XN(t))Lp(Ω,H˙2)<,p1.\sup_{t\in[0,T]}\|F(X^{N}(t))\|_{L^{p}(\Omega,\dot{H}^{2})}<\infty,\,\forall p\geq 1. (46)

The next result shows that XN(t)X^{N}(t) is differentiable in Malliavin sense.

Proposition 3.2 (Boundedness of the Malliavin derivative).

Let Assumptions 1.1-1.4 hold. Then the Malliavin derivative of XN(t)X^{N}(t) satisfies

𝔼[𝒟sXN(t)2(U0,H˙)2]<,0stT.\displaystyle\mathbb{E}\big{[}\|\mathcal{D}_{s}X^{N}(t)\|_{\mathcal{L}_{2}(U_{0},\dot{H})}^{2}\big{]}<\infty,\quad 0\leq s\leq t\leq T. (47)
Proof.

The existence of the Malliavin derivative 𝒟syXN(t)\mathcal{D}_{s}^{y}X^{N}(t) can be obtained by the standard argument such as the Picard iteration. Here, we will focus on the bound (47). Taking the Malliavin derivative on the equation (40) in the direction yU0y\in U_{0} and using the chain rule yield that for 0stT0\leq s\leq t\leq T,

𝒟syXN(t)=E(ts)PNystE(tr)APNF(XN(r))𝒟syXN(r)dr.\displaystyle\mathcal{D}_{s}^{y}X^{N}(t)=E(t-s)P_{N}y-\int_{s}^{t}E(t-r)AP_{N}F^{\prime}(X^{N}(r))\mathcal{D}_{s}^{y}X^{N}(r)\mathrm{d}r. (48)

Following a standard strategy for the analysis of the Cahn–Hilliard equations, the proof of the upper bounds for 𝒟syXN(t)\mathcal{D}_{s}^{y}X^{N}(t) requires to exploit two energy estimates, in the ||1|\cdot|_{-1} and |||\cdot| norms. First, observe that for all tst\geq s, 𝒟syXN(t)\mathcal{D}_{s}^{y}X^{N}(t) is differentiable and satisfies

d𝒟syXN(t)dt+A2𝒟syXN(t)+APNF(XN(t))𝒟syXN(t)=0.\displaystyle\frac{\mathrm{d}\mathcal{D}_{s}^{y}X^{N}(t)}{\mathrm{d}t}+A^{2}\mathcal{D}_{s}^{y}X^{N}(t)+AP_{N}F^{\prime}(X^{N}(t))\mathcal{D}_{s}^{y}X^{N}(t)=0. (49)

Multiplying A1𝒟syXN(t)A^{-1}\mathcal{D}_{s}^{y}X^{N}(t) on both sides of the above equation yields

d𝒟syXN(t)dt,A1𝒟syXN(t)+A2𝒟syXN(t),A1𝒟syXN(t)+APNF(XN(t))𝒟syXN(t),A1𝒟syXN(t)=0.\displaystyle\begin{split}\Big{\langle}\frac{\mathrm{d}\mathcal{D}_{s}^{y}X^{N}(t)}{\mathrm{d}t},A^{-1}\mathcal{D}_{s}^{y}X^{N}(t)\Big{\rangle}+\langle A^{2}\mathcal{D}_{s}^{y}X^{N}(t),A^{-1}\mathcal{D}_{s}^{y}X^{N}(t)\rangle+\langle AP_{N}F^{\prime}(X^{N}(t))\mathcal{D}_{s}^{y}X^{N}(t),A^{-1}\mathcal{D}_{s}^{y}X^{N}(t)\rangle=0.\end{split} (50)

Next, integrating (50) over [s,t][s,t] one obtains

|𝒟syXN(t)|12=|y|122st|𝒟syXN(r)|12dr2stF(XN(r))𝒟syXN(r),𝒟syXN(r)dr=|y|122st|𝒟syXN(r)|12dr+2stA12𝒟syXN(r),A12𝒟syXN(r)dr6st(XN(r))2𝒟syXN(r),𝒟syXN(r)dr|y|12st|𝒟syXN(r)|12dr+st|𝒟syXN(r)|12dr,\displaystyle\begin{split}|\mathcal{D}_{s}^{y}X^{N}(t)|_{-1}^{2}&=|y|_{-1}^{2}\!-\!2\!\!\int_{s}^{t}\!\!\!|\mathcal{D}_{s}^{y}X^{N}(r)|_{1}^{2}\mathrm{d}r\!-\!2\!\!\int_{s}^{t}\!\!\!\left\langle F^{\prime}(X^{N}(r))\mathcal{D}_{s}^{y}X^{N}(r),\mathcal{D}_{s}^{y}X^{N}(r)\right\rangle\mathrm{d}r\\ &=|y|_{-1}^{2}\!-\!2\int_{s}^{t}\!\!\!|\mathcal{D}_{s}^{y}X^{N}(r)|_{1}^{2}\mathrm{d}r+2\int_{s}^{t}\!\!\!\left\langle A^{\frac{1}{2}}\mathcal{D}_{s}^{y}X^{N}(r),A^{-\frac{1}{2}}\mathcal{D}_{s}^{y}X^{N}(r)\right\rangle\mathrm{d}r\\ &\quad-6\int_{s}^{t}\left\langle(X^{N}(r))^{2}\mathcal{D}_{s}^{y}X^{N}(r),\mathcal{D}_{s}^{y}X^{N}(r)\right\rangle\mathrm{d}r\\ &\leq|y|_{-1}^{2}-\int_{s}^{t}|\mathcal{D}_{s}^{y}X^{N}(r)|_{1}^{2}\mathrm{d}r+\int_{s}^{t}|\mathcal{D}_{s}^{y}X^{N}(r)|_{-1}^{2}\mathrm{d}r,\end{split} (51)

where in the last step the elementary inequality 2aba2+b22ab\leq a^{2}+b^{2} was used. Hence, by Gronwall’s inequality we have

|𝒟syXN(t)|12C|y|12.|\mathcal{D}_{s}^{y}X^{N}(t)|_{-1}^{2}\leq C|y|_{-1}^{2}. (52)

Therefore, one has

st|𝒟syXN(r)|12drC|y|12.\int_{s}^{t}|\mathcal{D}_{s}^{y}X^{N}(r)|_{1}^{2}\mathrm{d}r\leq C|y|_{-1}^{2}. (53)

Next, we may multiply by 𝒟syXN(t)\mathcal{D}_{s}^{y}X^{N}(t) both sides of (49) to get

d𝒟syXN(t)dt,𝒟syXN(t)+A2𝒟syXN(t),𝒟syXN(t)+APNF(XN(t))𝒟syXN(t),𝒟syXN(t)=0.\displaystyle\begin{split}\left\langle\frac{\mathrm{d}\mathcal{D}_{s}^{y}X^{N}(t)}{\mathrm{d}t},\mathcal{D}_{s}^{y}X^{N}(t)\right\rangle+\left\langle A^{2}\mathcal{D}_{s}^{y}X^{N}(t),\mathcal{D}_{s}^{y}X^{N}(t)\right\rangle+\left\langle AP_{N}F^{\prime}(X^{N}(t))\mathcal{D}_{s}^{y}X^{N}(t),\mathcal{D}_{s}^{y}X^{N}(t)\right\rangle=0.\end{split} (54)

Similarly, the energy estimate in the |||\cdot| norm is treated as follows:

|𝒟syXN(t)|2=|y|22stA𝒟syXN(r)2dr2stF(XN(r))𝒟syXN(r),A𝒟syXN(r)dr|y|22stA𝒟syXN(r)2dr+2stA𝒟syXN(r)2dr+12stF(XN(r))𝒟syXN(r)2dr|y|2+C(supr[s,t]XN(r)L64+1)st|𝒟syXN(r)|12dr|y|2+C|y|12(supr[s,t]XN(r)L64+1)|y|2+C|y|2(supr[s,t]XN(r)L64+1),\displaystyle\begin{split}|\mathcal{D}_{s}^{y}X^{N}(t)|^{2}&=|y|^{2}-2\int_{s}^{t}\|A\mathcal{D}_{s}^{y}X^{N}(r)\|^{2}\mathrm{d}r-2\int_{s}^{t}\langle F^{\prime}(X^{N}(r))\mathcal{D}_{s}^{y}X^{N}(r),A\mathcal{D}_{s}^{y}X^{N}(r)\rangle\mathrm{d}r\\ &\leq|y|^{2}-2\int_{s}^{t}\|A\mathcal{D}_{s}^{y}X^{N}(r)\|^{2}\mathrm{d}r+2\int_{s}^{t}\|A\mathcal{D}_{s}^{y}X^{N}(r)\|^{2}\mathrm{d}r\\ &\quad+\tfrac{1}{2}\int_{s}^{t}\|F^{\prime}(X^{N}(r))\mathcal{D}_{s}^{y}X^{N}(r)\|^{2}\mathrm{d}r\\ &\leq|y|^{2}+C\big{(}\sup_{r\in[s,t]}\|X^{N}(r)\|_{L^{6}}^{4}+1\big{)}\int_{s}^{t}|\mathcal{D}_{s}^{y}X^{N}(r)|_{1}^{2}\mathrm{d}r\\ &\leq|y|^{2}+C\,|y|_{-1}^{2}\big{(}\sup_{r\in[s,t]}\|X^{N}(r)\|_{L^{6}}^{4}+1\big{)}\\ &\leq|y|^{2}+C\,|y|^{2}\big{(}\sup_{r\in[s,t]}\|X^{N}(r)\|_{L^{6}}^{4}+1\big{)},\end{split} (55)

where in the first inequality the elementary inequality 2ab2a2+12b22ab\leq 2a^{2}+\tfrac{1}{2}b^{2} was used. What’s more, Hölder’s inequality fgCfL3gL6\|fg\|\leq C\|f\|_{L^{3}}\|g\|_{L^{6}} and Sobolev embedding inequality H˙d3L6,d=1,2,3\dot{H}^{\frac{d}{3}}\subset L^{6},d=1,2,3 were used in the above second inequality. Choosing y=Q12ei,i={1,2,}y=Q^{\frac{1}{2}}e_{i},i=\{1,2,\cdots\} and taking expectation yield

𝔼[𝒟sXN(t)2(U0,H˙)2]CQ1222CA12Q1222<,\mathbb{E}\big{[}\|\mathcal{D}_{s}X^{N}(t)\|_{\mathcal{L}_{2}(U_{0},\dot{H})}^{2}\big{]}\leq C\,\|Q^{\frac{1}{2}}\|_{\mathcal{L}_{2}}^{2}\leq C\,\|A^{\frac{1}{2}}Q^{\frac{1}{2}}\|_{\mathcal{L}_{2}}^{2}<\infty, (56)

where (3) and (43) were used. ∎

Remark 3.1.

The trace-class noise (i.e., Q122<\|Q^{\frac{1}{2}}\|_{\mathcal{L}_{2}}<\infty) is sufficient to obtain (56) and thus Proposition 3.2.

Let us now turn to some useful results on the nonlinear term FF.

Lemma 3.1.

Let FF be the Nemytskii operator defined in Assumption 1.2, then for d=1,2,3d=1,2,3,

|F(x)y|1C(1+|x|22)|y|1,xH˙2,yH˙1,|F^{\prime}(x)y|_{1}\leq C\big{(}1+|x|_{2}^{2}\big{)}|y|_{1},\,x\in\dot{H}^{2},\,y\in\dot{H}^{1}, (57)

and

|F(ς)ψ|1C(1+|ς|22)|ψ|1,ςH˙2,ψH˙.\displaystyle\begin{split}|F^{\prime}(\varsigma)\psi|_{-1}\leq C\big{(}1+|\varsigma|_{2}^{2}\big{)}|\psi|_{-1},\quad\forall\varsigma\in\dot{H}^{2},\psi\in\dot{H}.\end{split} (58)
Proof.

The estimate (57) is an immediate consequence of [27, Lemma 3.2]. To see (58), with the aid of the self-adjointness of A12A^{-\frac{1}{2}} and F(ς)F^{\prime}(\varsigma), we have

A12F(ς)ψ=supξ1A12F(ς)ψ,ξ=supξ1ψ,F(ς)A12ξ=supξ1A12ψ,A12F(ς)A12ξ|ψ|1supξ1|F(ς)A12ξ|1C(1+|ς|22)|ψ|1.\begin{split}\|A^{-\frac{1}{2}}F^{\prime}(\varsigma)\psi\|&=\sup_{\|\xi\|\leq 1}\langle A^{-\frac{1}{2}}F^{\prime}(\varsigma)\psi,\xi\rangle=\sup_{\|\xi\|\leq 1}\langle\psi,F^{\prime}(\varsigma)A^{-\frac{1}{2}}\xi\rangle=\sup_{\|\xi\|\leq 1}\langle A^{-\frac{1}{2}}\psi,A^{\frac{1}{2}}F^{\prime}(\varsigma)A^{-\frac{1}{2}}\xi\rangle\\ &\leq|\psi|_{-1}\sup_{\|\xi\|\leq 1}|F^{\prime}(\varsigma)A^{-\frac{1}{2}}\xi|_{1}\leq C\big{(}1+|\varsigma|_{2}^{2}\big{)}|\psi|_{-1}.\end{split} (59)

The condition ψH˙\psi\in\dot{H} is used in the above first and second identities. This finishes the proof. ∎

Now, we are well prepared to carry out the weak error analysis of the spatial semi-discretization.

Theorem 3.1 (Weak convergence rate of the spatial approximation).

Let X(T)X(T) and XN(T)X^{N}(T), given by (25) and (40), be the solution of problems (1) and (39) respectively. Let Assumptions 1.1-1.4 hold. Then for ΦCb2\Phi\in C_{b}^{2}, there exists a constant C>0C>0 such that

|𝔼[Φ(X(T))]𝔼[Φ(XN(T))]|CλN2.\big{|}\mathbb{E}[\Phi(X(T))]-\mathbb{E}[\Phi(X^{N}(T))]\big{|}\leq C\,\lambda_{N}^{-2}. (60)
Proof.

By introducing two processes X¯(t)=X(t)𝒪t\bar{X}(t)=X(t)-\mathcal{O}_{t} and X¯N(t)=XN(t)𝒪tN\bar{X}^{N}(t)=X^{N}(t)-\mathcal{O}^{N}_{t}, we can separate the error 𝔼[Φ(X(T))]𝔼[Φ(XN(T))]\mathbb{E}\big{[}\Phi(X(T))\big{]}-\mathbb{E}\big{[}\Phi(X^{N}(T))\big{]} into two terms as follows

𝔼[Φ(X(T))]𝔼[Φ(XN(T))]=(𝔼[Φ(X¯(T)+𝒪T)]𝔼[Φ(X¯N(T)+𝒪T)])+(𝔼[Φ(X¯N(T)+𝒪T)]𝔼[Φ(X¯N(T)+𝒪TN)])=:I1+I2.\begin{split}\mathbb{E}\big{[}\Phi(X(T))\big{]}-\mathbb{E}\big{[}\Phi(X^{N}(T))\big{]}&=\Big{(}\mathbb{E}\big{[}\Phi(\bar{X}(T)+\mathcal{O}_{T})\big{]}-\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T})\big{]}\Big{)}\\ &\quad+\Big{(}\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T})\big{]}-\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T}^{N})\big{]}\Big{)}\\ &=:I_{1}+I_{2}.\end{split} (61)

To estimate I1I_{1}, it suffices to consider the strong convergence between X¯(T)\bar{X}(T) and X¯N(T)\bar{X}^{N}(T). To be specific, by the Taylor expansion and triangle inequality we have

|I1|=|𝔼[Φ(X¯(T)+𝒪T)]𝔼[Φ(X¯N(T)+𝒪T)]|C|𝔼[X¯(T)X¯N(T)]|CX¯(T)PNX¯(T)L2(Ω,H˙)+CPNX¯(T)X¯N(T)L2(Ω,H˙).\begin{split}|I_{1}|&=\Big{|}\mathbb{E}\big{[}\Phi(\bar{X}(T)+\mathcal{O}_{T})\big{]}-\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T})\big{]}\Big{|}\leq C\big{|}\mathbb{E}[\|\bar{X}(T)-\bar{X}^{N}(T)\|]\big{|}\\ &\leq C\|\bar{X}(T)-P_{N}\bar{X}(T)\|_{L^{2}(\Omega,\dot{H})}+C\|P_{N}\bar{X}(T)-\bar{X}^{N}(T)\|_{L^{2}(\Omega,\dot{H})}.\end{split} (62)

To bound the first error term X¯(T)PNX¯(T)L2(Ω,H˙)\|\bar{X}(T)-P_{N}\bar{X}(T)\|_{L^{2}(\Omega,\dot{H})}, we need an estimate on X¯(t),t[0,T]\bar{X}(t),t\in[0,T], that is,

X¯(t)Lp(Ω,H˙4)=A2X¯(t)Lp(Ω,H˙)A2E(t)X0Lp(Ω,H˙)+0tA2E(ts)APF(X(t))dsLp(Ω,H˙)+0tA2E(ts)AP(F(X(t))F(X(s)))dsLp(Ω,H˙)C(|X0|4+F(X(t))Lp(Ω,H˙2)+0t(ts)1P(F(X(t))F(X(s)))Lp(Ω,H˙2)ds)C(|X0|4+F(X(t))Lp(Ω,H˙2))+C(1+supr[0,t]X(r)L4p(Ω,H˙2)2)0t(ts)1X(t)X(s)L2p(Ω,H˙2)dsC(1+0t(ts)1(ts)14ds)<,\begin{split}\|\bar{X}(t)&\|_{L^{p}(\Omega,\dot{H}^{4})}=\|A^{2}\bar{X}(t)\|_{L^{p}(\Omega,\dot{H})}\\ &\leq\|A^{2}E(t)X_{0}\|_{L^{p}(\Omega,\dot{H})}+\left\|\int_{0}^{t}A^{2}E(t-s)APF(X(t))\,\mathrm{d}s\right\|_{L^{p}(\Omega,\dot{H})}\\ &\quad+\left\|\int_{0}^{t}A^{2}E(t-s)AP(F(X(t))-F(X(s)))\,\mathrm{d}s\right\|_{L^{p}(\Omega,\dot{H})}\\ &\leq C\Big{(}|X_{0}|_{4}+\|F(X(t))\|_{L^{p}(\Omega,\dot{H}^{2})}+\int_{0}^{t}(t-s)^{-1}\|P(F(X(t))-F(X(s)))\|_{L^{p}(\Omega,\dot{H}^{2})}\mathrm{d}s\Big{)}\\ &\leq C\Big{(}|X_{0}|_{4}+\|F(X(t))\|_{L^{p}(\Omega,\dot{H}^{2})}\Big{)}\\ &\quad+C\Big{(}1+\sup_{r\in[0,t]}\|X(r)\|_{L^{4p}(\Omega,\dot{H}^{2})}^{2}\Big{)}\cdot\int_{0}^{t}(t-s)^{-1}\|X(t)-X(s)\|_{L^{2p}(\Omega,\dot{H}^{2})}\mathrm{d}s\\ &\leq C\Big{(}1+\int_{0}^{t}(t-s)^{-1}\cdot(t-s)^{\frac{1}{4}}\mathrm{d}s\Big{)}\\ &<\infty,\end{split} (63)

where (17) and (20) were used in the above second inequality, (15) was used in the above third inequality, (26)-(28) were used in the above fourth inequality. As a result, by using (38), we get

X¯(T)PNX¯(T)Lp(Ω,H˙)=(IPN)A2A2X¯(T)Lp(Ω,H˙)CλN2.\|\bar{X}(T)-P_{N}\bar{X}(T)\|_{L^{p}(\Omega,\dot{H})}=\|(I-P_{N})A^{-2}A^{2}\bar{X}(T)\|_{L^{p}(\Omega,\dot{H})}\leq C\lambda_{N}^{-2}. (64)

In the next step, we consider the second term of (62) in the treatment of I1I_{1}. For convenience, we denote eN(t)=PNX¯(t)X¯N(t)e^{N}(t)=P_{N}\bar{X}(t)-\bar{X}^{N}(t), which satisfies

ddteN(t)+A2eN(t)+APN[F(X¯(t)+𝒪t)F(X¯N(t)+𝒪tN)]=0.\tfrac{\mathrm{d}}{\mathrm{d}t}e^{N}(t)+A^{2}e^{N}(t)+AP_{N}\big{[}F(\bar{X}(t)+\mathcal{O}_{t})-F(\bar{X}^{N}(t)+\mathcal{O}^{N}_{t})\big{]}=0. (65)

We multiply the above identity by A1eN(t)A^{-1}{e^{N}(t)} to get

12ddt|eN(t)|12+|eN(t)|12=eN(t),F(X¯(t)+𝒪t)F(PNX¯(t)+𝒪t)eN(t),F(PNX¯(t)+𝒪t)F(X¯N(t)+𝒪t)eN(t),F(X¯N(t)+𝒪t)F(X¯N(t)+𝒪tN)12eN(t)2+12F(X¯(t)+𝒪t)F(PNX¯(t)+𝒪t)2+eN(t)2+|eN(t)|1|F(X¯N(t)+𝒪t)F(X¯N(t)+𝒪tN)|132eN(t)2+12F(X¯(t)+𝒪t)F(PNX¯(t)+𝒪t)2+14|eN(t)|12+|F(X¯N(t)+𝒪t)F(X¯N(t)+𝒪tN|1234|eN(t)|12+98|eN(t)|12+CX¯(t)PNX¯(t)2(1+|X¯(t)|24+|𝒪t|24)+C|𝒪t𝒪tN|12(1+|X¯N(t)|24+|𝒪t|24),\begin{split}\tfrac{1}{2}\tfrac{\mathrm{d}}{\mathrm{d}t}|e^{N}(t)|_{-1}^{2}+|e^{N}(t)|_{1}^{2}&=-\langle e^{N}(t),F(\bar{X}(t)+\mathcal{O}_{t})-F(P_{N}\bar{X}(t)+\mathcal{O}_{t})\rangle\\ &\quad-\langle{e^{N}(t)},F(P_{N}\bar{X}(t)+\mathcal{O}_{t})-F(\bar{X}^{N}(t)+\mathcal{O}_{t})\rangle\\ &\quad-\langle{e^{N}(t)},F(\bar{X}^{N}({t})+\mathcal{O}_{t})-F(\bar{X}^{N}(t)+\mathcal{O}^{N}_{t})\rangle\\ &\leq\tfrac{1}{2}\|e^{N}(t)\|^{2}+\tfrac{1}{2}\|F(\bar{X}(t)+\mathcal{O}_{t})-F(P_{N}\bar{X}(t)+\mathcal{O}_{t})\|^{2}\\ &\quad+\|e^{N}(t)\|^{2}+|e^{N}(t)|_{1}\cdot|F(\bar{X}^{N}(t)+\mathcal{O}_{t})-F(\bar{X}^{N}(t)+\mathcal{O}^{N}_{t})|_{-1}\\ &\leq\tfrac{3}{2}\|e^{N}(t)\|^{2}+\tfrac{1}{2}\|F(\bar{X}(t)+\mathcal{O}_{t})-F(P_{N}\bar{X}(t)+\mathcal{O}_{t})\|^{2}\\ &\quad+\tfrac{1}{4}|e^{N}(t)|_{1}^{2}+|F(\bar{X}^{N}(t)+\mathcal{O}_{t})-F(\bar{X}^{N}(t)+\mathcal{O}^{N}_{t}|_{-1}^{2}\\ &\leq\tfrac{3}{4}|e^{N}(t)|_{1}^{2}+\tfrac{9}{8}|e^{N}(t)|_{-1}^{2}+C\|\bar{X}(t)-P_{N}\bar{X}(t)\|^{2}(1+|\bar{X}(t)|_{2}^{4}+|\mathcal{O}_{t}|_{2}^{4})\\ &\quad+C|\mathcal{O}_{t}-\mathcal{O}^{N}_{t}|_{-1}^{2}(1+|\bar{X}^{N}({t})|_{2}^{4}+|\mathcal{O}_{t}|_{2}^{4}),\end{split} (66)

where in the above first inequality we used Young’s inequality ab12a2+12b2ab\leq\frac{1}{2}a^{2}+\frac{1}{2}b^{2}, (21) and Cauchy-Schwartz inequality. Also, (22), Sobolev embedding inequality H˙2V\dot{H}^{2}\subset V, Young’s inequality 32ab12a2+98b2\frac{3}{2}ab\leq\frac{1}{2}a^{2}+\frac{9}{8}b^{2}, Taylor’s expansion and Lemma 3.1 were used in the above last inequality. By Gronwall’s inequality, we further deduce

|eN(T)|12+0T|eN(t)|12dtC0TX¯(t)PNX¯(t)2(1+|X¯(t)|24+|𝒪t|24)dt+C0T|𝒪t𝒪tN|12(1+|X¯N(t)|24+|𝒪t|24)dt.\begin{split}|e^{N}(T)|_{-1}^{2}+\int_{0}^{T}|e^{N}(t)|_{1}^{2}\mathrm{d}t&\leq C\int_{0}^{T}\|\bar{X}(t)-P_{N}\bar{X}(t)\|^{2}(1+|\bar{X}(t)|_{2}^{4}+|\mathcal{O}_{t}|_{2}^{4})\mathrm{d}t\\ &\quad+C\int_{0}^{T}|\mathcal{O}_{t}-\mathcal{O}^{N}_{t}|_{-1}^{2}(1+|\bar{X}^{N}(t)|_{2}^{4}+|\mathcal{O}_{t}|_{2}^{4})\mathrm{d}t.\end{split}

Applying (23) and (38) gives

𝒪t𝒪tNLp(Ω,H˙1)=(IPN)A2A32𝒪tLp(Ω,H˙)CλN2𝒪tLp(Ω,H˙3)CλN2.\|\mathcal{O}_{t}-\mathcal{O}^{N}_{t}\|_{L^{p}(\Omega,\dot{H}^{-1})}=\|(I-P_{N})A^{-2}A^{\frac{3}{2}}\mathcal{O}_{\color[rgb]{1,0,0}t}\|_{L^{p}(\Omega,\dot{H})}\leq C\lambda_{N}^{-2}~{}\|\mathcal{O}_{t}\|_{L^{p}(\Omega,\dot{H}^{3})}\leq C\lambda_{N}^{-2}. (67)

With the aid of the regularity of X(T)X(T) and XN(T)X^{N}(T), (64), Hölder’s inequality and (67), one can find that

0T|eN(t)|12dtLp(Ω,)C0TX¯(t)PNX¯(t)L4p(Ω,H˙)2dt+C0T𝒪t𝒪tNL4p(Ω,H˙1)2dtCλN4.\begin{split}\Big{\|}\int_{0}^{T}|e^{N}(t)|_{1}^{2}\mathrm{d}t\Big{\|}_{L^{p}(\Omega,\mathbb{R})}&\leq C\int_{0}^{T}\|\bar{X}(t)-P_{N}\bar{X}(t)\|_{L^{4p}(\Omega,\dot{H})}^{2}\mathrm{d}t+C\int_{0}^{T}\|\mathcal{O}_{t}-\mathcal{O}^{N}_{t}\|_{L^{4p}(\Omega,\dot{H}^{-1})}^{2}\mathrm{d}t\\ &\leq C\lambda_{N}^{-4}.\end{split} (68)

We are now ready to estimate

eN(T)Lp(Ω,H˙)=PN(E(T)X00TE(Ts)AF(X(s))ds)(E(T)PNX00TE(Ts)APNF(XN(s))ds)Lp(Ω,H˙)=0TE(Ts)APN(F(X(s))F(XN(s)))dsLp(Ω,H˙)0TE(Ts)APN(F(X¯(s)+𝒪s)F(PNX¯(s)+𝒪s))dsLp(Ω,H˙)+0TE(Ts)APN(F(PNX¯(s)+𝒪s)F(X¯N(s)+𝒪s))dsLp(Ω,H˙)+0TE(Ts)APN(F(X¯N(s)+𝒪s)F(X¯N(s)+𝒪sN))dsLp(Ω,H˙)=:e1N(T)+e2N(T)+e3N(T).\begin{split}\|e^{N}(T)\|_{L^{p}(\Omega,\dot{H})}&=\Big{\|}P_{N}\Big{(}E(T)X_{0}-\int_{0}^{T}E(T-s)AF(X(s))\mathrm{d}s\Big{)}\\ &\quad-\Big{(}E(T)P_{N}X_{0}-\int_{0}^{T}E(T-s)AP_{N}F(X^{N}(s))\mathrm{d}s\Big{)}\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &=\left\|\int_{0}^{T}E(T-s)AP_{N}(F(X(s))-F(X^{N}(s)))\mathrm{d}s\right\|_{L^{p}(\Omega,\dot{H})}\\ &\leq\left\|\int_{0}^{T}E(T-s)AP_{N}(F(\bar{X}(s)+\mathcal{O}_{s})-F(P_{N}\bar{X}(s)+\mathcal{O}_{s}))\mathrm{d}s\right\|_{L^{p}(\Omega,\dot{H})}\\ &\quad+\left\|\int_{0}^{T}E(T-s)AP_{N}(F(P_{N}\bar{X}(s)+\mathcal{O}_{s})-F(\bar{X}^{N}(s)+\mathcal{O}_{s}))\mathrm{d}s\right\|_{L^{p}(\Omega,\dot{H})}\\ &\quad+\left\|\int_{0}^{T}E(T-s)AP_{N}(F(\bar{X}^{N}(s)+\mathcal{O}_{s})-F(\bar{X}^{N}(s)+\mathcal{O}^{N}_{s}))\mathrm{d}s\right\|_{L^{p}(\Omega,\dot{H})}\\ &=:e^{N}_{1}(T)+e^{N}_{2}(T)+e^{N}_{3}(T).\end{split} (69)

Again, by (17), (22), (23),(38), (63) and Sobolev embedding inequality H˙2V\dot{H}^{2}\subset V, we have

e1N(T)=0TE(Ts)APN(F(X¯(s)+𝒪s)F(PNX¯(s)+𝒪s))dsLp(Ω,H˙)C0T(Ts)12X¯(s)PNX¯(s)L2p(Ω,H˙)ds×(1+sups[0,T]X¯(s)L4p(Ω,H˙2)2+sups[0,T]𝒪sL4p(Ω,H˙2)2)CλN2.\begin{split}e^{N}_{1}(T)&=\left\|\int_{0}^{T}E(T-s)AP_{N}(F(\bar{X}(s)+\mathcal{O}_{s})-F(P_{N}\bar{X}(s)+\mathcal{O}_{s}))\mathrm{d}s\right\|_{L^{p}(\Omega,\dot{H})}\\ &\leq C\int_{0}^{T}(T-s)^{-\frac{1}{2}}\|\bar{X}(s)-P_{N}\bar{X}(s)\|_{L^{2p}(\Omega,\dot{H})}\mathrm{d}s\times\big{(}1+\sup_{s\in[0,T]}\|\bar{X}(s)\|_{L^{4p}(\Omega,\dot{H}^{2})}^{2}+\sup_{s\in[0,T]}\|\mathcal{O}_{s}\|_{L^{4p}(\Omega,\dot{H}^{2})}^{2}\big{)}\\ &\leq C\lambda_{N}^{-2}.\end{split} (70)

From (17), (57) in Lemma 3.1, Hölder’s inequality, (68) and regularity of X(t)X(t) and XN(t)X^{N}(t), it follows that

e2N(T)C0T(Ts)14|F(PNX¯(s)+𝒪s)F(X¯N(s)+𝒪s)|1dsLp(Ω,)C0T(Ts)14|eN(s)|1(1+|X¯(s)|22+|X¯N(s)|22+|𝒪s|22)dsLp(Ω,)C0T|eN(s)|12dsLp(Ω,)12(0T(Ts)12ds)12CλN2.\begin{split}e^{N}_{2}(T)&\leq C\left\|\int_{0}^{T}(T-s)^{{-\frac{1}{4}}}\big{|}F(P_{N}\bar{X}(s)+\mathcal{O}_{s})-F(\bar{X}^{N}(s)+\mathcal{O}_{s})\big{|}_{1}\mathrm{d}s\right\|_{L^{p}(\Omega,\mathbb{R})}\\ &\leq C\left\|\int_{0}^{T}(T-s)^{-\frac{1}{4}}|e^{N}(s)|_{1}\big{(}1+|\bar{X}(s)|_{2}^{2}+|\bar{X}^{N}(s)|_{2}^{2}+|\mathcal{O}_{s}|_{2}^{2}\big{)}\mathrm{d}s\right\|_{L^{p}(\Omega,\mathbb{R})}\\ &\leq C\left\|\int_{0}^{T}|e^{N}(s)|_{1}^{2}\mathrm{d}s\right\|_{L^{p}(\Omega,\mathbb{R})}^{\frac{1}{2}}\left(\int_{0}^{T}(T-s)^{-\frac{1}{2}}\mathrm{d}s\right)^{\frac{1}{2}}\\ &\leq C\lambda_{N}^{-2}.\end{split} (71)

Similarly to the estimate of (70) with (58) and (67) instead, we obtain

e3N(T)=0TE(Ts)A32A12PN(F(X¯N(s)+𝒪s)F(X¯N(s)+𝒪sN))dsLp(Ω,H˙)C0T(Ts)34𝒪s𝒪sNL2p(Ω,H˙1)ds(1+sups[0,T]X¯N(s)L4p(Ω,H˙2)2+sups[0,T]𝒪sL4p(Ω,H˙2)2)CλN2.\begin{split}e^{N}_{3}(T)&=\left\|\int_{0}^{T}E(T-s)A^{\frac{3}{2}}A^{-\frac{1}{2}}P_{N}(F(\bar{X}^{N}(s)+\mathcal{O}_{s})-F(\bar{X}^{N}(s)+\mathcal{O}^{N}_{s}))\mathrm{d}s\right\|_{L^{p}(\Omega,\dot{H})}\\ &\leq C\int_{0}^{T}(T-s)^{-\frac{3}{4}}\|\mathcal{O}_{s}-\mathcal{O}^{N}_{s}\|_{L^{2p}(\Omega,\dot{H}^{-1})}\mathrm{d}s\big{(}1+\sup_{s\in[0,T]}\|\bar{X}^{N}(s)\|_{L^{4p}(\Omega,\dot{H}^{2})}^{2}+\sup_{s\in[0,T]}\|\mathcal{O}_{s}\|_{L^{4p}(\Omega,\dot{H}^{2})}^{2}\big{)}\\ &\leq C\lambda_{N}^{-2}.\end{split} (72)

Therefore, gathering estimates of e1N(T)e^{N}_{1}(T), e2N(T)e^{N}_{2}(T) and e3N(T)e^{N}_{3}(T) together yields

PNX¯(T)X¯N(T)L2(Ω,H˙)CλN2.\|P_{N}\bar{X}(T)-\bar{X}^{N}(T)\|_{L^{2}(\Omega,\dot{H})}\leq C\lambda_{N}^{-2}. (73)

Combining it with (64) yields

X¯(T)X¯N(T)L2(Ω,H˙)CλN2,\|\bar{X}(T)-\bar{X}^{N}(T)\|_{L^{2}(\Omega,\dot{H})}\leq C\lambda_{N}^{-2}, (74)

and thus |I1|CλN2|I_{1}|\leq C\lambda_{N}^{-2}. Next, we turn to the estimate of |I2||I_{2}|. Using Taylor’s expansion and the triangle inequality, we get

|I2|=|𝔼[Φ(XN(T))(𝒪T𝒪TN)+01Φ′′(XN(T)+λ(𝒪T𝒪TN))(𝒪T𝒪TN,𝒪T𝒪TN)(1λ)dλ]||𝔼[Φ(XN(T))(IPN)𝒪T]|+C𝔼[𝒪T𝒪TN2].\displaystyle\begin{split}|I_{2}|&=\Big{|}\mathbb{E}\Big{[}\Phi^{{}^{\prime}}(X^{N}(T))(\mathcal{O}_{T}-\mathcal{O}_{T}^{N})+\int_{0}^{1}\Phi^{{}^{\prime\prime}}(X^{N}(T)+\lambda(\mathcal{O}_{T}-\mathcal{O}_{T}^{N}))(\mathcal{O}_{T}-\mathcal{O}_{T}^{N},\mathcal{O}_{T}-\mathcal{O}_{T}^{N})(1-\lambda)\mathrm{d}\lambda\Big{]}\Big{|}\\ &\leq\Big{|}\mathbb{E}\big{[}\Phi^{\prime}(X^{N}(T))(I-P_{N})\mathcal{O}_{T}\big{]}\Big{|}+C\,\mathbb{E}\big{[}\|\mathcal{O}_{T}-\mathcal{O}^{N}_{T}\|^{2}\big{]}.\end{split} (75)

The second term can be easily bounded by utilizing (38) and the moment bound for |𝒪T|3|\mathcal{O}_{T}|_{3} in Lemma 2.1, that is

𝔼[𝒪T𝒪TN2]=𝔼[(IPN)𝒪T2]CλN3.\displaystyle\mathbb{E}\Big{[}\Big{\|}\mathcal{O}_{T}-\mathcal{O}^{N}_{T}\Big{\|}^{2}\Big{]}=\mathbb{E}\Big{[}\Big{\|}(I-P_{N})\mathcal{O}_{T}\Big{\|}^{2}\Big{]}\leq C\lambda_{N}^{-3}. (76)

For the first term, (47) in Proposition 3.2, the Malliavin integration by parts formula (36), the chain rule of the Malliavin derivative, (17), (38) and (3) enable us to obtain

|𝔼[Φ(XN(T))(IPN)𝒪T]|=|𝔼[0T(IPN)E(Ts)dW(s),Φ(XN(T))]|=|𝔼0T(IPN)E(Ts),𝒟sΦ(XN(T))20ds|C𝔼0T(IPN)E(Ts)20Φ′′(XN(T))𝒟sXN(T)20dsC0T(IPN)E(Ts)A12A12Q122dsCλN20T(Ts)34dsCλN2.\displaystyle\begin{split}\Big{|}\mathbb{E}\big{[}\Phi^{\prime}(X^{N}(T))(I-P_{N})\mathcal{O}_{T}\big{]}\Big{|}&=\Big{|}\mathbb{E}\Big{[}\Big{\langle}\int_{0}^{T}(I-P_{N})E(T-s)\mathrm{d}W(s),\Phi^{{}^{\prime}}(X^{N}(T))\Big{\rangle}\Big{]}\Big{|}\\ &=\Big{|}\mathbb{E}\int_{0}^{T}\left<(I-P_{N})E(T-s),\mathcal{D}_{s}\Phi^{{}^{\prime}}(X^{N}(T))\right>_{\mathcal{L}_{2}^{0}}\mathrm{d}s\Big{|}\\ &\leq C\,\mathbb{E}\int_{0}^{T}\big{\|}(I-P_{N})E(T-s)\big{\|}_{\mathcal{L}_{2}^{0}}\|\Phi^{{}^{\prime\prime}}(X^{N}(T))\|_{\mathcal{L}}\|\mathcal{D}_{s}X^{N}(T)\|_{\mathcal{L}_{2}^{0}}\,\mathrm{d}s\\ &\leq C\,\int_{0}^{T}\|(I-P_{N})E(T-s)A^{-\frac{1}{2}}\|_{\mathcal{L}}\|A^{\frac{1}{2}}Q^{\frac{1}{2}}\|_{\mathcal{L}_{2}}\mathrm{d}s\\ &\leq C\,\lambda_{N}^{-2}\int_{0}^{T}(T-s)^{-\frac{3}{4}}\mathrm{d}s\leq C\,\lambda_{N}^{-2}.\end{split} (77)

Hence, we obtain |I2|CλN2|I_{2}|\leq C\lambda_{N}^{-2}. Gathering it with |I1|CλN2|I_{1}|\leq C\lambda_{N}^{-2} then concludes the proof. ∎

4 Weak convergence rate of the backward Euler method

Based on the spatial spectral Galerkin approximation (39), this section concerns the weak error analysis of a backward Euler method in the temporal direction. We divide the interval [0,T][0,T] into MM equidistant subintervals with the time step-size τ=TM\tau=\tfrac{T}{M} and denote the nodes tm=mτt_{m}=m\tau for m{0,1,,M},M+m\in\{0,1,\cdots,M\},\,M\in\mathbb{N}^{+}. Then, the fully discrete scheme reads

XtmM,NXtm1M,N+τA2XtmM,N+τPNAF(XtmM,N)=PNΔWm,X0M,N=PNX0,X_{t_{m}}^{M,N}-X_{t_{m-1}}^{M,N}+\tau A^{2}X_{t_{m}}^{M,N}+\tau P_{N}AF(X_{t_{m}}^{M,N})=P_{N}\Delta W_{m},\quad X_{0}^{M,N}=P_{N}X_{0}, (78)

where ΔWm:=W(tm)W(tm1)\Delta W_{m}:=W(t_{m})-W(t_{m-1}) for short. By introducing a family of operators {Eτ,Nm}m=1M\{E_{\tau,N}^{m}\}_{m=1}^{M}: Eτ,Nmv=(I+τA2)mPNv=j=1N(1+τλj2)mv,ejejE_{\tau,N}^{m}v=(I+\tau A^{2})^{-m}P_{N}v=\sum_{j=1}^{N}(1+\tau\lambda_{j}^{2})^{-m}\langle v,e_{j}\rangle e_{j}, vH˙\forall\ v\in\dot{H}, we have

XtmM,N=Eτ,NmX0τj=1mEτ,Nmj+1AF(XtjM,N)+j=1mEτ,Nmj+1ΔWj.X_{t_{m}}^{M,N}=E_{\tau,N}^{m}X_{0}-\tau\sum_{j=1}^{m}E_{\tau,N}^{m-j+1}AF(X_{t_{j}}^{M,N})+\sum_{j=1}^{m}E_{\tau,N}^{m-j+1}\Delta W_{j}. (79)

Thanks to [29, Theorem C.2], the implicit scheme (78) is well-defined. More details can be found in [27]. Following the proof of [21, (2.10)], it is easy to check that the operator Eτ,NmE_{\tau,N}^{m} satisfies

AμEτ,NmvCtmμ2v,μ[0,2],vH˙,m{1,2,,M}\|A^{\mu}E_{\tau,N}^{m}v\|\leq Ct_{m}^{-\frac{\mu}{2}}\|v\|,\quad\mu\in[0,2],\,v\in\dot{H},\,m\in\{1,2,\cdots,M\} (80)

and there exists a constant CC such that for all vH˙v\in\dot{H},

(τj=1mAEτ,Njv2)12Cv.\Big{(}\tau\sum_{j=1}^{m}\|AE_{\tau,N}^{j}v\|^{2}\Big{)}^{\frac{1}{2}}\leq C\|v\|. (81)

The regularity of the fully discrete approximation is derived in the following result.

Proposition 4.1.

Let Assumptions 1.1-1.4 be satisfied, then we have for all p1p\geq 1,

supNsupm{0,1,,M}XtmM,NLp(Ω,H˙2)<.\sup_{N\in\mathbb{N}}\sup_{m\in\{0,1,\cdots,M\}}\|X^{M,N}_{t_{m}}\|_{L^{p}(\Omega,\dot{H}^{2})}<\infty. (82)
Proof.

Firstly, by the proof in [27, Theorem 4.1], we have for η(32,2)\eta\in(\frac{3}{2},2) and all p1p\geq 1,

supNsupm{0,1,,M}XtmM,NLp(Ω,H˙η)<.\sup_{N\in\mathbb{N}}\sup_{m\in\{0,1,\cdots,M\}}\|X^{M,N}_{t_{m}}\|_{L^{p}(\Omega,\dot{H}^{\eta})}<\infty. (83)

Next, from (80), the Burkholder-Davis-Gundy-type inequality, (81), (3), (4) and Sobolev embedding inequality H˙ηV\dot{H}^{\eta}\subset V, it follows that

supNsupm{0,1,,M}XtmM,NLp(Ω,H˙2)X0Lp(Ω,H˙2)+τsupNsupm{0,1,,M}j=1mtmj+134PF(XtjM,N)Lp(Ω,H˙1)+supNsupm{0,1,,M}(τj=1mA32Eτ,Nmj+1Q1222)12C(1+A12Q1222+τsupNsupm{0,1,,M}j=1mtmj+134|XtjM,N|1XtjM,NV2Lp(Ω,))C(1+supm{0,1,,M}τj=1mtmj+134supNsupj{0,1,,M}XtjM,NL3p(Ω,H˙η))<.\begin{split}\sup_{N\in\mathbb{N}}&\sup_{m\in\{0,1,\cdots,M\}}\|X^{M,N}_{t_{m}}\|_{L^{p}(\Omega,\dot{H}^{2})}\\ &\leq\|X_{0}\|_{L^{p}(\Omega,\dot{H}^{2})}+\tau\sup_{N\in\mathbb{N}}\sup_{m\in\{0,1,\cdots,M\}}\sum_{j=1}^{m}t_{m-j+1}^{-\frac{3}{4}}\|PF(X_{t_{j}}^{M,N})\|_{L^{p}(\Omega,\dot{H}^{1})}\\ &\quad+\sup_{N\in\mathbb{N}}\sup_{m\in\{0,1,\cdots,M\}}\Big{(}\tau\sum_{j=1}^{m}\big{\|}A^{\frac{3}{2}}E_{\tau,N}^{m-j+1}Q^{\frac{1}{2}}\big{\|}_{\mathcal{L}_{2}}^{2}\Big{)}^{\frac{1}{2}}\\ &\leq C\Big{(}1+\|A^{\frac{1}{2}}Q^{\frac{1}{2}}\|_{\mathcal{L}_{2}}^{2}+\tau\sup_{N\in\mathbb{N}}\sup_{m\in\{0,1,\cdots,M\}}\sum_{j=1}^{m}t_{m-j+1}^{-\frac{3}{4}}\Big{\|}|X_{t_{j}}^{M,N}|_{1}~{}\|X_{t_{j}}^{M,N}\|_{V}^{2}\Big{\|}_{L^{p}(\Omega,\mathbb{R})}\Big{)}\\ &\leq C\Big{(}1+\sup_{m\in\{0,1,\cdots,M\}}\tau\sum_{j=1}^{m}t_{m-j+1}^{-\frac{3}{4}}\sup_{N\in\mathbb{N}}\sup_{j\in\{0,1,\cdots,M\}}\|X_{t_{j}}^{M,N}\|_{L^{3p}(\Omega,\dot{H}^{\eta})}\Big{)}\\ &<\infty.\end{split} (84)

This completes the proof. ∎

Before presenting the main theorem, we introduce the notation s:=max{0,τ,,mτ,}[0,s]\lfloor s\rfloor:=\max\{0,\tau,\cdots,m\tau,\cdots\}\cap[0,s], s:=min{0,τ,,mτ,}[s,T]\lceil s\rceil:=\min\{0,\tau,\cdots,m\tau,\cdots\}\cap[s,T] and [s]:=sτ[s]:=\frac{\lfloor s\rfloor}{\tau}. The fully discrete approximation operator is then defined by

ΨτM,N(t):=E(t)PNEτ,Nk,t[tk1,tk),k{1,2,,M}.\Psi_{\tau}^{M,N}(t):=E(t)P_{N}-E_{\tau,N}^{k},\quad t\in[t_{k-1},t_{k}),\quad k\in\{1,2,\cdots,M\}. (85)

The following lemma of the fully discrete approximation operator plays a pivotal role in the weak convergence analysis.

Lemma 4.1.

Under Assumption 1.1, we have the following statements.

  1. (i)

    Let ρ[0,4]\rho\in[0,4], there exists a constant CC such that for t>0t>0,

    ΨτM,N(t)uCtρ4|u|ρ,uH˙ρ.\|\Psi_{\tau}^{M,N}(t)u\|\leq C\,t^{-\frac{\rho}{4}}\,|u|_{-\rho},\,u\in\dot{H}^{-\rho}. (86)
  2. (ii)

    Let β[0,4]\beta\in[0,4], there exists a constant CC such that for t>0t>0,

    ΨτM,N(t)uCτβ4|u|β,uH˙β.\|\Psi_{\tau}^{M,N}(t)u\|\leq C\,\tau^{\frac{\beta}{4}}\,|u|_{\beta},\,u\in\dot{H}^{\beta}. (87)
  3. (iii)

    Let α[0,4]\alpha\in[0,4], there exists a constant CC such that for t>0t>0,

    ΨτM,N(t)uCτ4α4t1|u|α,uH˙α.\|\Psi_{\tau}^{M,N}(t)u\|\leq C\,\tau^{\frac{4-\alpha}{4}}\,t^{-1}\,|u|_{-\alpha},\,u\in\dot{H}^{-\alpha}. (88)
  4. (iv)

    Let μ[0,4]\mu\in[0,4], there exists a constant CC such that for t>0t>0,

    ΨτM,N(t)uCτt4μ4|u|μ,uH˙μ.\|\Psi_{\tau}^{M,N}(t)u\|\leq C\,\tau\cdot t^{-\frac{4-\mu}{4}}\,|u|_{\mu},\,u\in\dot{H}^{\mu}. (89)
  5. (v)

    Let ν[0,4]\nu\in[0,4], there exists a constant CC such that for t>0t>0,

    (0tΨτM,N(s)u2ds)12Cτν4|u|ν2,uH˙ν2.\Big{(}\int_{0}^{t}\|\Psi_{\tau}^{M,N}(s)u\|^{2}\mathrm{d}s\Big{)}^{\frac{1}{2}}\leq C\,\tau^{\frac{\nu}{4}}|u|_{\nu-2},\,u\in\dot{H}^{\nu-2}. (90)
  6. (vi)

    Let δ[0,4]\delta\in[0,4], there exists a constant CC such that for t>0t>0,

    0tΨτM,N(s)udsCτ4δ4|u|δ,uH˙δ.\Big{\|}\int_{0}^{t}\Psi_{\tau}^{M,N}(s)u\mathrm{d}s\Big{\|}\leq C\,\tau^{\frac{4-\delta}{4}}|u|_{-\delta},\,u\in\dot{H}^{-\delta}. (91)
Proof.

Elementary fact in [27, Lemma 5.3] yields (i), (ii), (iii), (v) and (vi). We then use the standard interpolation argument to prove (iv). For μ=0\mu=0, it is a consequence of (iii) with α=0\alpha=0 and for μ=4\mu=4, it is a consequence of (ii) with β=4\beta=4. ∎

For clarity of exposition, we denote 𝒪TM,N:=j=1MEτ,NMj+1ΔWj=0TEτ,NM[s]dW(s).\mathcal{O}_{T}^{M,N}:=\sum_{j=1}^{M}E_{\tau,N}^{M-j+1}\Delta W_{j}=\int_{0}^{T}E_{\tau,N}^{M-[s]}\mathrm{d}W(s). The next lemma gives the estimate between 𝒪tmN\mathcal{O}_{t_{m}}^{N} and 𝒪tmM,N\mathcal{O}_{t_{m}}^{M,N}.

Lemma 4.2.

Under Assumptions 1.1 and 1.3, we have for p1p\geq 1,

supm{1,2,,M}𝒪tmM,N𝒪tmNLp(Ω,H˙β)Cτ3+β4,β[3,1].\sup_{m\in\{1,2,\cdots,M\}}\big{\|}\mathcal{O}_{t_{m}}^{M,N}-\mathcal{O}_{t_{m}}^{N}\big{\|}_{L^{p}(\Omega,\dot{H}^{-\beta})}\leq C\,\tau^{\frac{3+\beta}{4}},\,\beta\in[-3,1]. (92)
Proof.

The Burkholder-Davis-Gundy inequality and (v) in Lemma 4.1 with ν=3+β\nu=3+\beta yield

𝒪tmM,N𝒪tmNLp(Ω,H˙β)C(0tmΨτM,N(tms)Aβ2Q1222ds)12Cτ3+β4A12Q122Cτ3+β4.\begin{split}\big{\|}\mathcal{O}_{t_{m}}^{M,N}-\mathcal{O}_{t_{m}}^{N}\big{\|}_{L^{p}(\Omega,\dot{H}^{-\beta})}&\leq C~{}\Big{(}\int_{0}^{t_{m}}\|\Psi_{\tau}^{M,N}(t_{m}-s)A^{-\frac{\beta}{2}}Q^{\frac{1}{2}}\|_{\mathcal{L}_{2}}^{2}\mathrm{d}s\Big{)}^{\frac{1}{2}}\\ &\leq C\,\tau^{\frac{3+\beta}{4}}\|A^{\frac{1}{2}}Q^{\frac{1}{2}}\|_{\mathcal{L}_{2}}\leq C\,\tau^{\frac{3+\beta}{4}}.\end{split} (93)

This finishes the proof. ∎

The following theorem shows the weak convergence rate of the temporal semi-discretization.

Theorem 4.1 (Weak convergence rate of the temporal approximation).

Suppose Assumptions 1.1-1.4 are satisfied. Let XN(T)X^{N}(T) and XTM,NX_{T}^{M,N} be given by (40) and (79), respectively. Then, we have for ΦCb2\Phi\in C_{b}^{2},

|𝔼[Φ(XN(T))]𝔼[Φ(XTM,N)]|Cτ.\big{|}\mathbb{E}[\Phi(X^{N}(T))]-\mathbb{E}[\Phi(X_{T}^{M,N})]\big{|}\leq C\,\tau. (94)
Proof.

At first, we define X¯TM,N=XTM,N𝒪TM,N\bar{X}_{T}^{M,N}={X}_{T}^{M,N}-\mathcal{O}_{T}^{M,N} and separate the above error into

𝔼[Φ(XN(T))]𝔼[Φ(XTM,N)]=(𝔼[Φ(X¯N(T)+𝒪TM,N)]𝔼[Φ(X¯TM,N+𝒪TM,N)])+(𝔼[Φ(X¯N(T)+𝒪TN)]𝔼[Φ(X¯N(T)+𝒪TM,N)])=:K1+K2.\displaystyle\begin{split}\mathbb{E}\Big{[}\Phi(X^{N}(T))\Big{]}-\mathbb{E}\Big{[}\Phi(X_{T}^{M,N})\Big{]}&=\Big{(}\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T}^{M,N})\big{]}-\mathbb{E}\big{[}\Phi(\bar{X}_{T}^{M,N}+\mathcal{O}_{T}^{M,N})\big{]}\Big{)}\\ &\quad+\Big{(}\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T}^{N})\big{]}-\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T}^{M,N})\big{]}\Big{)}\\ &=:K_{1}+K_{2}.\end{split} (95)

To estimate K1K_{1}, it suffices to bound X¯N(T)X¯TM,NL2(Ω,H˙)\big{\|}\bar{X}^{N}(T)-\bar{X}_{T}^{M,N}\big{\|}_{L^{2}(\Omega,\dot{H})} . To this end, we introduce an auxiliary process YtmM,NY_{t_{m}}^{M,N} by

YtmM,N=Eτ,NmX0τj=1mEτ,Nmj+1AF(XN(tj))+j=1mEτ,Nmj+1ΔWjY_{t_{m}}^{M,N}=E_{\tau,N}^{m}X_{0}-\tau\sum_{j=1}^{m}E_{\tau,N}^{m-j+1}AF(X^{N}(t_{j}))+\sum_{j=1}^{m}E_{\tau,N}^{m-j+1}\Delta W_{j} (96)

and define Y¯tmM,N=YtmM,N𝒪tmM,N\bar{Y}_{t_{m}}^{M,N}=Y_{t_{m}}^{M,N}-\mathcal{O}_{t_{m}}^{M,N}. Note that the application of an appropriate auxiliary process was used in [16, 28] to deduce the strong convergence rates for the numerical approximations of similar SPDEs. Owning to (3), (4), (46), (80), (81) and discrete Burkholder-Davis-Gundy-type inequality, one can easily derive that for any m{0,1,,2,,M}m\in\{0,1,,2,\cdots,M\},

YtmM,NLp(Ω,H˙3)<.\|Y_{t_{m}}^{M,N}\|_{L^{p}(\Omega,\dot{H}^{3})}<\infty. (97)

Subsequently, by the triangle inequality, we have

X¯N(T)X¯TM,NL2(Ω,H˙)X¯N(T)Y¯TM,NL2(Ω,H˙)+Y¯TM,NX¯TM,NL2(Ω,H˙).\big{\|}\bar{X}^{N}(T)-\bar{X}_{T}^{M,N}\big{\|}_{L^{2}(\Omega,\dot{H})}\leq\big{\|}\bar{X}^{N}(T)-\bar{Y}_{T}^{M,N}\big{\|}_{L^{2}(\Omega,\dot{H})}+\big{\|}\bar{Y}_{T}^{M,N}-\bar{X}_{T}^{M,N}\big{\|}_{L^{2}(\Omega,\dot{H})}. (98)

The error term X¯N(T)Y¯TM,NLp(Ω,H˙)\big{\|}\bar{X}^{N}(T)-\bar{Y}_{T}^{M,N}\big{\|}_{L^{p}(\Omega,\dot{H})} can be further divided into three terms

X¯N(T)Y¯TM,NLp(Ω,H˙)=(E(T)PNEτ,NM)X0(0TE(Ts)PNAF(XN(s))dsτj=1MEτ,NMj+1AF(XN(tj)))Lp(Ω,H˙)(E(T)PNEτ,NM)X0Lp(Ω,H˙)+0T(E(Ts)PNEτ,NM[s])AF(XN(s))dsLp(Ω,H˙)+0TEτ,NM[s]A(F(XN(s))F(XN(s)))dsLp(Ω,H˙)=:K11+K12+K13.\begin{split}\big{\|}&\bar{X}^{N}(T)-\bar{Y}_{T}^{M,N}\big{\|}_{L^{p}(\Omega,\dot{H})}=\Big{\|}(E(T)P_{N}-E_{\tau,N}^{M})X_{0}\\ &\quad-\Big{(}\int_{0}^{T}E(T-s)P_{N}AF(X^{N}(s))\mathrm{d}s-\tau\sum_{j=1}^{M}E_{\tau,N}^{M-j+1}AF(X^{N}(t_{j}))\Big{)}\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &\leq\big{\|}(E(T)P_{N}-E_{\tau,N}^{M})X_{0}\big{\|}_{L^{p}(\Omega,\dot{H})}\\ &\quad+\Big{\|}\int_{0}^{T}\big{(}E(T-s)P_{N}-E_{\tau,N}^{M-[s]}\big{)}AF(X^{N}(s))\mathrm{d}s\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &\quad+\Big{\|}\int_{0}^{T}E_{\tau,N}^{M-[s]}A\big{(}F(X^{N}(s))-F(X^{N}(\lceil s\rceil))\big{)}\mathrm{d}s\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &=:K_{11}+K_{12}+K_{13}.\end{split} (99)

By (ii) of Lemma 4.1 with β=4\beta=4 and Assumption 1.4, we deduce

K11Cτ|X0|4Cτ.K_{11}\leq C\,\tau|X_{0}|_{4}\leq C\,\tau. (100)

Concerning the term K12K_{12}, by use of (vi) and (iv) in Lemma 4.1, (15), (44), (45), (46), we obtain

K120T(E(Ts)PNEτ,NM[s])APF(XN(T))dsLp(Ω,H˙)+0T(E(Ts)PNEτ,NM[s])AP(F(XN(s))F(XN(T)))Lp(Ω,H˙)dsCτF(XN(T))Lp(Ω,H˙2)+Cτ0T(Ts)1P(F(XN(s))F(XN(T)))Lp(Ω,H˙2)dsCτ+Cτ(1+supr[0,t]XN(r)L4p(Ω,H˙2)2)0T(Ts)1XN(s)XN(T)L2p(Ω,H˙2)dsCτ+Cτ0T(Ts)1(Ts)14dsCτ.\begin{split}K_{12}&\leq\Big{\|}\int_{0}^{T}\big{(}E(T-s)P_{N}-E_{\tau,N}^{M-[s]}\big{)}APF(X^{N}(T))\mathrm{d}s\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &\quad+\int_{0}^{T}\Big{\|}\big{(}E(T-s)P_{N}-E_{\tau,N}^{M-[s]}\big{)}AP(F(X^{N}(s))-F(X^{N}(T)))\Big{\|}_{L^{p}(\Omega,\dot{H})}\mathrm{d}s\\ &\leq C\,\tau\,\|F(X^{N}(T))\|_{L^{p}(\Omega,\dot{H}^{2})}+C\,\tau\,\int_{0}^{T}(T-s)^{-1}\|P(F(X^{N}(s))-F(X^{N}(T)))\|_{L^{p}(\Omega,\dot{H}^{2})}\mathrm{d}s\\ &\leq C\,\tau+C\,\tau\,\Big{(}1+\sup_{r\in[0,t]}\|X^{N}(r)\|_{L^{4p}(\Omega,\dot{H}^{2})}^{2}\Big{)}\cdot\int_{0}^{T}(T-s)^{-1}\|X^{N}(s)-X^{N}(T)\|_{L^{2p}(\Omega,\dot{H}^{2})}\mathrm{d}s\\ &\leq C\,\tau+C\,\tau\,\int_{0}^{T}(T-s)^{-1}(T-s)^{\frac{1}{4}}\mathrm{d}s\\ &\leq C\,\tau.\end{split} (101)

To handle K13K_{13}, we decompose it into four terms with the aid of the Taylor expansion and the mild form of XN(t)X^{N}(t),

K130TEτ,NM[s]A(F(XN(s))(E(ss)I)XN(s))dsLp(Ω,H˙)+0TEτ,NM[s]A(F(XN(s))ssE(sr)PNAF(XN(r))dr)dsLp(Ω,H˙)+0TEτ,NM[s]A(F(XN(s))ssE(sr)PNdW(r))dsLp(Ω,H˙)+0TEτ,NM[s]A(01F′′(XN(s)+λ(XN(s)XN(s)))(XN(s)XN(s),XN(s)XN(s))(1λ)dλ)dsLp(Ω,H˙)=:K131+K132+K133+K134.\begin{split}K_{13}&\leq\Big{\|}\int_{0}^{T}E_{\tau,N}^{M-[s]}A\Big{(}F^{\prime}(X^{N}(s))(E(\lceil s\rceil-s)-I)X^{N}(s)\Big{)}\mathrm{d}s\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &\quad+\Big{\|}\int_{0}^{T}E_{\tau,N}^{M-[s]}A\Big{(}F^{\prime}(X^{N}(s))\int_{s}^{\lceil s\rceil}\!\!\!E(\lceil s\rceil-r)P_{N}AF(X^{N}(r))\mathrm{d}r\Big{)}\mathrm{d}s\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &\quad+\Big{\|}\int_{0}^{T}E_{\tau,N}^{M-[s]}A\Big{(}F^{\prime}(X^{N}(s))\int_{s}^{\lceil s\rceil}E(\lceil s\rceil-r)P_{N}\mathrm{d}W(r)\Big{)}\mathrm{d}s\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &\quad+\Big{\|}\int_{0}^{T}E_{\tau,N}^{M-[s]}A\Big{(}\int_{0}^{1}F^{\prime\prime}\big{(}X^{N}(s)+\lambda(X^{N}(\lceil s\rceil)-X^{N}(s))\big{)}\\ &\qquad\qquad\qquad\big{(}X^{N}(\lceil s\rceil)-X^{N}(s),X^{N}(\lceil s\rceil)-X^{N}(s)\big{)}(1-\lambda)\mathrm{d}\lambda\Big{)}\mathrm{d}s\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &=:K_{131}+K_{132}+K_{133}+K_{134}.\end{split} (102)

The smoothness of Eτ,NmE_{\tau,N}^{m} in (80), (18), (58) and the regularity of XN(t)X^{N}(t) lead to

K131=0TEτ,NM[s]A32A12(F(XN(s))(E(ss)I)XN(s))dsLp(Ω,H˙)C0T(Ts)34(1+|XN(s)|22)|(E(ss)I)XN(s)|1Lp(Ω,)dsCτ.\begin{split}K_{131}&=\Big{\|}\int_{0}^{T}E_{\tau,N}^{M-[s]}A^{\frac{3}{2}}A^{-\frac{1}{2}}\Big{(}F^{\prime}(X^{N}(s))(E(\lceil s\rceil-s)-I)X^{N}(s)\Big{)}\mathrm{d}s\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &\leq C\!\int_{0}^{T}\!\!\!(T-\lfloor s\rfloor)^{-\frac{3}{4}}\left\|\big{(}1+|X^{N}(s)|_{2}^{2}\big{)}\big{|}(E(\lceil s\rceil-s)-I)X^{N}(s)\big{|}_{-1}\right\|_{L^{p}(\Omega,\mathbb{R})}\!\!\!\mathrm{d}s\\ &\leq C\,\tau.\end{split} (103)

Following similar approach as above and utilizing (46) yield

K132C0T(Ts)34(1+|XN(s)|22)ss|E(sr)PNAF(XN(r))|1drLp(Ω,)dsCτ0T(Ts)34dssupr[0,T]F(XN(r))L2p(Ω,H˙2)Cτ.\begin{split}K_{132}&\leq C\int_{0}^{T}(T-\lfloor s\rfloor)^{-\frac{3}{4}}\Big{\|}\big{(}1+|X^{N}(s)|_{2}^{2}\big{)}\int_{s}^{\lceil s\rceil}\big{|}E(\lceil s\rceil-r)P_{N}AF(X^{N}(r))\big{|}_{-1}\mathrm{d}r\Big{\|}_{L^{p}(\Omega,\mathbb{R})}\mathrm{d}s\\ &\leq C\,\tau\int_{0}^{T}(T-\lfloor s\rfloor)^{-\frac{3}{4}}\mathrm{d}s\sup_{r\in[0,T]}\|F(X^{N}(r))\|_{L^{2p}(\Omega,\dot{H}^{2})}\\ &\leq C\,\tau.\end{split} (104)

From stochastic Fubini theorem, the Burkholder-Davis-Gundy-type inequality and Hölder’s inequality, it follows that

K133=j=1Mtj1tjtj1tjχ[s,tj)(r)Eτ,NM[s]AF(XN(s))E(sr)PNdW(r)dsLp(Ω,H˙)=j=1Mtj1tjtj1tjχ[s,tj)(r)Eτ,NM[s]AF(XN(s))E(sr)PNdsdW(r)Lp(Ω,H˙)(j=1Mtj1tjtj1tjχ[s,tj)(r)Eτ,NM[s]AF(XN(s))E(sr)dsLp(Ω,20)2dr)12Cτ12(j=1Mtj1tjtj1tjEτ,NM[s]A12A12F(XN(s))E(sr)Lp(Ω,20)2dsdr)12Cτ12(j=1Mtj1tjtj1tj(Ts)12(1+XN(s)L2p(Ω,H˙2)4)A12Q1222dsdr)12Cτ,\begin{split}K_{133}&=\Big{\|}\sum_{j=1}^{M}\int_{t_{j-1}}^{t_{j}}\!\int_{t_{j-1}}^{t_{j}}\!\!\chi_{[s,t_{j})}(r)E_{\tau,N}^{M-[s]}AF^{\prime}(X^{N}(s))E(\lceil s\rceil-r)P_{N}\mathrm{d}W(r)\mathrm{d}s\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &=\Big{\|}\sum_{j=1}^{M}\int_{t_{j-1}}^{t_{j}}\!\int_{t_{j-1}}^{t_{j}}\!\!\chi_{[s,t_{j})}(r)E_{\tau,N}^{M-[s]}AF^{\prime}(X^{N}(s))E(\lceil s\rceil-r)P_{N}\mathrm{d}s\mathrm{d}W(r)\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &\leq\left(\sum_{j=1}^{M}\int_{t_{j-1}}^{t_{j}}\Big{\|}\int_{t_{j-1}}^{t_{j}}\chi_{[s,t_{j})}(r)E_{\tau,N}^{M-[s]}AF^{\prime}(X^{N}(s))E(\lceil s\rceil-r)\mathrm{d}s\Big{\|}_{L^{p}(\Omega,\mathcal{L}_{2}^{0})}^{2}\,\mathrm{d}r\right)^{\frac{1}{2}}\\ &\leq C\,\tau^{\frac{1}{2}}\left(\sum_{j=1}^{M}\int_{t_{j-1}}^{t_{j}}\!\int_{t_{j-1}}^{t_{j}}\!\!\Big{\|}E_{\tau,N}^{M-[s]}A^{\frac{1}{2}}A^{\frac{1}{2}}F^{\prime}(X^{N}(s))E(\lceil s\rceil-r)\Big{\|}_{L^{p}(\Omega,\mathcal{L}_{2}^{0})}^{2}\mathrm{d}s\,\mathrm{d}r\right)^{\frac{1}{2}}\\ &\leq C\,\tau^{\frac{1}{2}}\left(\sum_{j=1}^{M}\int_{t_{j-1}}^{t_{j}}\!\int_{t_{j-1}}^{t_{j}}\!\!(T-\lfloor s\rfloor)^{-\frac{1}{2}}\big{(}1+\|X^{N}(s)\|_{L^{2p}(\Omega,\dot{H}^{2})}^{4}\big{)}\big{\|}A^{\frac{1}{2}}Q^{\frac{1}{2}}\big{\|}_{\mathcal{L}_{2}}^{2}\mathrm{d}s\,\mathrm{d}r\right)^{\frac{1}{2}}\\ &\leq C\,\tau,\end{split} (105)

where χ[0,t]\chi_{[0,t]} denotes the indicate function on [0,t][0,t]. Additionally, (57), (80) and the stability of E(sr)E(\lceil s\rceil-r) were used in the third inequality and in the last inequality we used (3) and (44). Owing to Hölder’s inequality, the Sobolev embedding inequality H˙δV\dot{H}^{\delta}\subset V for 32<δ<2\frac{3}{2}<\delta<2 and Proposition 3.1, we obtain

K134=0TEτ,NM[s]A(01F′′(XN(s)+λ(XN(s)XN(s)))(XN(s)XN(s),XN(s)XN(s))(1λ)dλ)dsLp(Ω,H˙)C0T(Ts)2+δ4XN(s)XN(s)L4p(Ω,H˙)2(1+sups[0,T]XN(s)L2p(Ω,V))dsCτ.\begin{split}K_{134}&=\Big{\|}\int_{0}^{T}E_{\tau,N}^{M-[s]}A\Big{(}\int_{0}^{1}F^{\prime\prime}\big{(}X^{N}(s)+\lambda(X^{N}(\lceil s\rceil)-X^{N}(s))\big{)}\\ &\qquad\qquad\qquad\big{(}X^{N}(\lceil s\rceil)-X^{N}(s),X^{N}(\lceil s\rceil)-X^{N}(s)\big{)}(1-\lambda)\mathrm{d}\lambda\Big{)}\mathrm{d}s\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &\leq C\int_{0}^{T}(T-\lfloor s\rfloor)^{-\frac{2+\delta}{4}}\|X^{N}(\lceil s\rceil)-X^{N}(s)\|_{L^{4p}(\Omega,\dot{H})}^{2}\big{(}1+\sup_{s\in[0,T]}\|X^{N}(s)\|_{L^{2p}(\Omega,V)}\big{)}\mathrm{d}s\\ &\leq C\,\tau.\end{split} (106)

Gathering the above estimates and (99) together gives

X¯N(T)Y¯TM,NLp(Ω,H˙)Cτ.\big{\|}\bar{X}^{N}(T)-\bar{Y}_{T}^{M,N}\big{\|}_{L^{p}(\Omega,\dot{H})}\leq C\,\tau. (107)

Finally, we turn to remaining error term Y¯TM,NX¯TM,NL2(Ω,H˙)\big{\|}\bar{Y}_{T}^{M,N}-\bar{X}_{T}^{M,N}\big{\|}_{L^{2}(\Omega,\dot{H})} in (98). Denoting etM,N=Y¯tM,NX¯tM,Ne_{t}^{M,N}=\bar{Y}_{t}^{M,N}-\bar{X}_{t}^{M,N}, we have

etmM,Netm1M,N+τA2etmM,N=τPNAF(XtmM,N)τPNAF(XN(tm)).e_{t_{m}}^{M,N}-e_{t_{m-1}}^{M,N}+\tau A^{2}e_{t_{m}}^{M,N}=\tau P_{N}AF(X_{t_{m}}^{M,N})-\tau P_{N}AF(X^{N}(t_{m})). (108)

Multiplying both sides by A1etmM,NA^{-1}e_{t_{m}}^{M,N} shows

etmM,Netm1M,N,A1etmM,N+τA2etmM,N,A1etmM,N=τF(X¯tmM,N+𝒪tmM,N)F(Y¯tmM,N+𝒪tmM,N),etmM,N+τF(Y¯tmM,N+𝒪tmM,N)F(X¯N(tm)+𝒪tmM,N),etmM,N+τF(X¯N(tm)+𝒪tmM,N)F(X¯N(tm)+𝒪tmN),etmM,N.\begin{split}\langle e_{t_{m}}^{M,N}-&e_{t_{m-1}}^{M,N},A^{-1}e_{t_{m}}^{M,N}\rangle+\tau\langle A^{2}e_{t_{m}}^{M,N},A^{-1}e_{t_{m}}^{M,N}\rangle\\ &=\tau\langle F(\bar{X}_{t_{m}}^{M,N}+\mathcal{O}_{t_{m}}^{M,N})-F(\bar{Y}_{t_{m}}^{M,N}+\mathcal{O}_{t_{m}}^{M,N}),e_{t_{m}}^{M,N}\rangle\\ &\quad+\tau\langle F(\bar{Y}_{t_{m}}^{M,N}+\mathcal{O}_{t_{m}}^{M,N})-F(\bar{X}^{N}(t_{m})+\mathcal{O}_{t_{m}}^{M,N}),e_{t_{m}}^{M,N}\rangle\\ &\quad+\tau\langle F(\bar{X}^{N}(t_{m})+\mathcal{O}_{t_{m}}^{M,N})-F(\bar{X}^{N}(t_{m})+\mathcal{O}_{t_{m}}^{N}),e_{t_{m}}^{M,N}\rangle.\end{split} (109)

Following similar approach in (66) and using the inequality etmM,Netm1M,N,A1etmM,N12(|etmM,N|12|etm1M,N|12)\langle e_{t_{m}}^{M,N}-e_{t_{m-1}}^{M,N},A^{-1}e_{t_{m}}^{M,N}\rangle\geq\tfrac{1}{2}\big{(}|e_{t_{m}}^{M,N}|_{-1}^{2}-|e_{t_{m-1}}^{M,N}|_{-1}^{2}\big{)} and the monotonicity of FF in (21), we further obtain

12(|etmM,N|12|etm1M,N|12)+τ|etmM,N|1234τ|etmM,N|12+98τ|etmM,N|12+CτF(Y¯tmM,N+𝒪tmM,N)F(X¯N(tm)+𝒪tmM,N)2+Cτ|F(X¯N(tm)+𝒪tmM,N)F(X¯N(tm)+𝒪tmN)|12.\begin{split}\tfrac{1}{2}\big{(}|e_{t_{m}}^{M,N}|_{-1}^{2}-|e_{t_{m-1}}^{M,N}|_{-1}^{2}\big{)}+\tau|e_{t_{m}}^{M,N}|_{1}^{2}&\leq\tfrac{3}{4}\tau|e_{t_{m}}^{M,N}|_{1}^{2}+\tfrac{9}{8}\tau|e_{t_{m}}^{M,N}|_{-1}^{2}\\ &\quad+C\,\tau\big{\|}F(\bar{Y}_{t_{m}}^{M,N}+\mathcal{O}_{t_{m}}^{M,N})-F(\bar{X}^{N}(t_{m})+\mathcal{O}_{t_{m}}^{M,N})\big{\|}^{2}\\ &\quad+C\,\tau\big{|}F(\bar{X}^{N}(t_{m})+\mathcal{O}_{t_{m}}^{M,N})-F(\bar{X}^{N}(t_{m})+\mathcal{O}_{t_{m}}^{N})\big{|}_{-1}^{2}.\end{split} (110)

By iteration in mm and Gronwall’s inequality, we obtain

|eTM,N|12+τj=1M|etjM,N|12Cτj=1M(F(Y¯tjM,N+𝒪tjM,N)F(X¯N(tj)+𝒪tjM,N)2)+Cτj=1M(|F(X¯N(tj)+𝒪tjM,N)F(X¯N(tj)+𝒪tjN)|12).\begin{split}|e_{T}^{M,N}|_{-1}^{2}+\tau\sum_{j=1}^{M}|e_{t_{j}}^{M,N}|_{1}^{2}&\leq C\,\tau\sum_{j=1}^{M}\Big{(}\big{\|}F(\bar{Y}_{t_{j}}^{M,N}+\mathcal{O}_{t_{j}}^{M,N})-F(\bar{X}^{N}(t_{j})+\mathcal{O}_{t_{j}}^{M,N})\big{\|}^{2}\Big{)}\\ &\quad+C\,\tau\sum_{j=1}^{M}\Big{(}\big{|}F(\bar{X}^{N}(t_{j})+\mathcal{O}_{t_{j}}^{M,N})-F(\bar{X}^{N}(t_{j})+\mathcal{O}_{t_{j}}^{N})\big{|}_{-1}^{2}\Big{)}.\end{split} (111)

It is worth mentioning that (107) also holds for arbitrary tj,j{1,,M}t_{j},j\in\{1,\cdots,M\} by repeating the same argument from (99) to (107). Then, employing (22), (44), (58), (97) and Lemma 4.2 results in

τj=1M|etjM,N|12Lp(Ω,)Cτj=1MY¯tjM,NX¯N(tj)2(1+Y¯tjM,NV4+X¯N(tj)V4+𝒪tjM,NV4)Lp(Ω,)+Cτj=1M|𝒪tjM,N𝒪tjN|12(1+|X¯N(tj)|24+|𝒪tjM,N|24+|𝒪tjN|24)Lp(Ω,)Cτ2.\begin{split}\Big{\|}\tau\sum_{j=1}^{M}|e_{t_{j}}^{M,N}|_{1}^{2}\Big{\|}_{L^{p}(\Omega,\mathbb{R})}&\leq C~{}\tau\sum_{j=1}^{M}\!\Big{\|}\big{\|}\bar{Y}_{t_{j}}^{M,N}\!-\!\bar{X}^{N}(t_{j})\big{\|}^{2}\big{(}1\!+\!\|\bar{Y}_{t_{j}}^{M,N}\|_{V}^{4}+\|\bar{X}^{N}(t_{j})\|_{V}^{4}+\|\mathcal{O}_{t_{j}}^{M,N}\|_{V}^{4}\big{)}\Big{\|}_{L^{p}(\Omega,\mathbb{R})}\\ &\quad+C~{}\tau\sum_{j=1}^{M}\!\Big{\|}\big{|}\mathcal{O}_{t_{j}}^{M,N}-\mathcal{O}_{t_{j}}^{N}\big{|}_{-1}^{2}\big{(}1\!+\!|\bar{X}^{N}(t_{j})|_{2}^{4}+|\mathcal{O}_{t_{j}}^{M,N}|_{2}^{4}+|\mathcal{O}_{t_{j}}^{N}|_{2}^{4}\big{)}\Big{\|}_{L^{p}(\Omega,\mathbb{R})}\\ &\leq C\,\tau^{2}.\end{split} (112)

Furthermore, since

eTM,N=Y¯TM,NX¯TM,N=τj=1MEτ,NMj+1A(F(XtjM,N)F(XN(tj))),e_{T}^{M,N}=\bar{Y}_{T}^{M,N}-\bar{X}_{T}^{M,N}=\tau\sum_{j=1}^{M}E_{\tau,N}^{M-j+1}A\big{(}F(X_{t_{j}}^{M,N})-F(X^{N}(t_{j}))\big{)}, (113)

we split eTM,NLp(Ω,H˙)\|e_{T}^{M,N}\|_{L^{p}(\Omega,\dot{H})} into three parts

eTM,NLp(Ω,H˙)=τj=1MEτ,NMj+1A(F(XN(tj))F(XtjM,N))Lp(Ω,H˙)τj=1MEτ,NMj+1A(F(X¯N(tj)+𝒪tjN)F(Y¯tjM,N+𝒪tjN))Lp(Ω,H˙)+τj=1MEτ,NMj+1A(F(Y¯tjM,N+𝒪tjN)F(Y¯tjM,N+𝒪tjM,N))Lp(Ω,H˙)+τj=1MEτ,NMj+1A(F(Y¯tjM,N+𝒪tjM,N)F(X¯tjM,N+𝒪tjM,N))Lp(Ω,H˙)=:Err1+Err2+Err3.\begin{split}\|e_{T}^{M,N}\|_{L^{p}(\Omega,\dot{H})}&=\tau\Big{\|}\sum_{j=1}^{M}E_{\tau,N}^{M-j+1}A\big{(}F(X^{N}(t_{j}))-F(X_{t_{j}}^{M,N})\big{)}\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &\leq\tau\sum_{j=1}^{M}\Big{\|}E_{\tau,N}^{M-j+1}A\big{(}F(\bar{X}^{N}(t_{j})+\mathcal{O}_{t_{j}}^{N})-F(\bar{Y}_{t_{j}}^{M,N}+\mathcal{O}_{t_{j}}^{N})\big{)}\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &\quad+\tau\sum_{j=1}^{M}\Big{\|}E_{\tau,N}^{M-j+1}A\big{(}F(\bar{Y}_{t_{j}}^{M,N}+\mathcal{O}_{t_{j}}^{N})-F(\bar{Y}_{t_{j}}^{M,N}+\mathcal{O}_{t_{j}}^{M,N})\big{)}\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &\quad+\tau\Big{\|}\sum_{j=1}^{M}E_{\tau,N}^{M-j+1}A\big{(}F(\bar{Y}_{t_{j}}^{M,N}+\mathcal{O}_{t_{j}}^{M,N})-F(\bar{X}_{t_{j}}^{M,N}+\mathcal{O}_{t_{j}}^{M,N})\big{)}\Big{\|}_{L^{p}(\Omega,\dot{H})}\\ &=:Err_{1}+Err_{2}+Err_{3}.\end{split} (114)

Taking (80), (22), (107), Hölder’s inequality and moment bounds of YtmM,NY_{t_{m}}^{M,N} and XN(t)X^{N}(t) into account, we arrive at

Err1Cτj=1MtMj+112X¯N(tj)Y¯tjM,NL2p(Ω,H˙)(1+X¯N(tj)L4p(Ω,V)2+Y¯tjM,NL4p(Ω,V)2+𝒪tjNL4p(Ω,V)2)Cτj=1MtMj+112τCτ.\begin{split}Err_{1}&\leq C\,\tau\sum_{j=1}^{M}t_{M-j+1}^{-\frac{1}{2}}\|\bar{X}^{N}(t_{j})-\bar{Y}_{t_{j}}^{M,N}\|_{L^{2p}(\Omega,\dot{H})}\\ &\qquad\qquad\big{(}1+\|\bar{X}^{N}(t_{j})\|_{L^{4p}(\Omega,V)}^{2}+\|\bar{Y}_{t_{j}}^{M,N}\|_{L^{4p}(\Omega,V)}^{2}+\|\mathcal{O}_{t_{j}}^{N}\|_{L^{4p}(\Omega,V)}^{2}\big{)}\\ &\leq C\,\tau\,\sum_{j=1}^{M}t_{M-j+1}^{-\frac{1}{2}}\,\tau\leq C\,\tau.\end{split} (115)

Analogously to the above estimate but with (58) instead, we derive

Err2Cτj=1MtMj+134𝒪tjN𝒪tjM,NL2p(Ω,H˙1)(1+Y¯tjM,NL4p(Ω,H˙2)2+𝒪tjNL4p(Ω,H˙2)2+𝒪tjM,NL4p(Ω,H˙2)2)Cτ.\begin{split}Err_{2}&\leq C\tau\sum_{j=1}^{M}t_{M-j+1}^{-\frac{3}{4}}\|\mathcal{O}_{t_{j}}^{N}-\mathcal{O}_{t_{j}}^{M,N}\|_{L^{2p}(\Omega,\dot{H}^{-1})}\\ &\quad\big{(}1+\|\bar{Y}_{t_{j}}^{M,N}\|_{L^{4p}(\Omega,\dot{H}^{2})}^{2}+\|\mathcal{O}_{t_{j}}^{N}\|_{L^{4p}(\Omega,\dot{H}^{2})}^{2}+\|\mathcal{O}_{t_{j}}^{M,N}\|_{L^{4p}(\Omega,\dot{H}^{2})}^{2}\big{)}\\ &\leq C\,\tau.\end{split} (116)

At last, combining (57), Hölder’s inequality, (112) and regularity of YtmM,NY_{t_{m}}^{M,N} and XtmM,NX_{t_{m}}^{M,N} leads to

Err3Cτj=1MtMj+114|F(Y¯tjM,N+𝒪tjM,N)F(X¯tjM,N+𝒪tjM,N)|1Lp(Ω,)Cτj=1MtMj+114|etjM,N|1(1+|Y¯tjM,N|22+|X¯tjM,N|22+|𝒪tjM,N|22)Lp(Ω,)Cτj=1M|etjM,N|12Lp(Ω,)12×τj=1MtMj+112(1+|Y¯tjM,N|24+|X¯tjM,N|24+|𝒪tjM,N|24)Lp(Ω,)12Cτ.\begin{split}Err_{3}&\leq C\left\|\tau\sum_{j=1}^{M}t_{M-j+1}^{-\frac{1}{4}}\big{|}F(\bar{Y}_{t_{j}}^{M,N}+\mathcal{O}_{t_{j}}^{M,N})-F(\bar{X}_{t_{j}}^{M,N}+\mathcal{O}_{t_{j}}^{M,N})\big{|}_{1}\right\|_{L^{p}(\Omega,\mathbb{R})}\\ &\leq C\left\|\tau\sum_{j=1}^{M}t_{M-j+1}^{-\frac{1}{4}}|e_{t_{j}}^{M,N}|_{1}\big{(}1+|\bar{Y}_{t_{j}}^{M,N}|_{2}^{2}+|\bar{X}_{t_{j}}^{M,N}|_{2}^{2}+|\mathcal{O}_{t_{j}}^{M,N}|_{2}^{2}\big{)}\right\|_{L^{p}(\Omega,\mathbb{R})}\\ &\leq C\left\|\tau\sum_{j=1}^{M}|e_{t_{j}}^{M,N}|_{1}^{2}\right\|_{L^{p}(\Omega,\mathbb{R})}^{\frac{1}{2}}\times\left\|\tau\sum_{j=1}^{M}t_{M-j+1}^{-\frac{1}{2}}\big{(}1+|\bar{Y}_{t_{j}}^{M,N}|_{2}^{4}+|\bar{X}_{t_{j}}^{M,N}|_{2}^{4}+|\mathcal{O}_{t_{j}}^{M,N}|_{2}^{4}\big{)}\right\|_{L^{p}(\Omega,\mathbb{R})}^{\frac{1}{2}}\\ &\leq C\,\tau.\end{split} (117)

Combining the above estimates together yields

X¯N(T)X¯TM,NL2(Ω,H˙)Cτ,\big{\|}\bar{X}^{N}(T)-\bar{X}_{T}^{M,N}\big{\|}_{L^{2}(\Omega,\dot{H})}\leq C\,\tau, (118)

and thus |K1|Cτ|K_{1}|\leq C\,\tau. The estimate of K2K_{2} relies on a second-order Taylor expansion and the triangle inequality:

|K2|=|𝔼[Φ(X¯N(T)+𝒪TN)]𝔼[Φ(X¯N(T)+𝒪TM,N)]||𝔼[Φ(XN(T))(𝒪TM,N𝒪TN)]|+|𝔼[01Φ′′(XN(T)+λ(𝒪TM,N𝒪TN))(𝒪TM,N𝒪TN,𝒪TM,N𝒪TN)(1λ)dλ]||𝔼[Φ(XN(T))(𝒪TM,N𝒪TN)]|+C𝔼[𝒪TM,N𝒪TN2].\begin{split}|K_{2}|&=\Big{|}\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T}^{N})\big{]}-\mathbb{E}\big{[}\Phi(\bar{X}^{N}(T)+\mathcal{O}_{T}^{M,N})\big{]}\Big{|}\\ &\leq\Big{|}\mathbb{E}\Big{[}\Phi^{\prime}(X^{N}(T))\big{(}\mathcal{O}_{T}^{M,N}-\mathcal{O}_{T}^{N}\big{)}\Big{]}\Big{|}\\ &\quad+\Big{|}\mathbb{E}\Big{[}\int_{0}^{1}\Phi^{\prime\prime}\big{(}X^{N}(T)+\lambda(\mathcal{O}_{T}^{M,N}-\mathcal{O}_{T}^{N})\big{)}\big{(}\mathcal{O}_{T}^{M,N}-\mathcal{O}_{T}^{N},\mathcal{O}_{T}^{M,N}-\mathcal{O}_{T}^{N}\big{)}(1-\lambda)\mathrm{d}\lambda\Big{]}\Big{|}\\ &\leq\Big{|}\mathbb{E}\big{[}\Phi^{\prime}(X^{N}(T))\big{(}\mathcal{O}_{T}^{M,N}-\mathcal{O}_{T}^{N}\big{)}\big{]}\Big{|}+C\,\mathbb{E}\big{[}\|\mathcal{O}_{T}^{M,N}-\mathcal{O}_{T}^{N}\|^{2}\big{]}.\end{split} (119)

Thanks to Lemma 4.2 with β=0\beta=0, we have

𝔼[𝒪TM,N𝒪TN2](Cτ34)2Cτ32.\mathbb{E}\big{[}\|\mathcal{O}_{T}^{M,N}-\mathcal{O}_{T}^{N}\|^{2}\big{]}\leq(C\tau^{\frac{3}{4}})^{2}\leq C\,\tau^{\frac{3}{2}}. (120)

Then, we turn our attention to the first term,

|𝔼[Φ(XN(T))(𝒪TM,N𝒪TN)]|=|𝔼[0T(E(Ts)PNEτ,NM[s])dW(s),Φ(XN(T))]|=|𝔼0TE(Ts)PNEτ,NM[s],𝒟sΦ(XN(T))20ds|𝔼0TE(Ts)PNEτ,NM[s]20Φ′′(XN(T))𝒟sXN(T)20dsC0T(E(Ts)PNEτ,NM[s])A12A12Q122dsCτ0T(Ts)34dsCτ,\begin{split}\Big{|}\mathbb{E}\big{[}\Phi^{\prime}(X^{N}(T))\big{(}\mathcal{O}_{T}^{M,N}-\mathcal{O}_{T}^{N}\big{)}\big{]}\Big{|}&=\Big{|}\mathbb{E}\Big{[}\Big{\langle}\int_{0}^{T}\big{(}E(T-s)P_{N}-E_{\tau,N}^{M-[s]}\big{)}\mathrm{d}W(s),\Phi^{{}^{\prime}}(X^{N}(T))\Big{\rangle}\Big{]}\Big{|}\\ &=\Big{|}\mathbb{E}\int_{0}^{T}\big{\langle}E(T-s)P_{N}-E_{\tau,N}^{M-[s]},\mathcal{D}_{s}\Phi^{\prime}(X^{N}(T))\big{\rangle}_{\mathcal{L}_{2}^{0}}\mathrm{d}s\Big{|}\\ &\leq\mathbb{E}\int_{0}^{T}\big{\|}E(T-s)P_{N}-E_{\tau,N}^{M-[s]}\big{\|}_{\mathcal{L}_{2}^{0}}\|\Phi^{\prime\prime}(X^{N}(T))\mathcal{D}_{s}X^{N}(T)\|_{\mathcal{L}_{2}^{0}}\mathrm{d}s\\ &\leq C\int_{0}^{T}\big{\|}\big{(}E(T-s)P_{N}-E_{\tau,N}^{M-[s]}\big{)}A^{-\frac{1}{2}}\big{\|}_{\mathcal{L}}\|A^{\frac{1}{2}}Q^{\frac{1}{2}}\|_{\mathcal{L}_{2}}\mathrm{d}s\\ &\leq C\tau\int_{0}^{T}(T-s)^{-\frac{3}{4}}\mathrm{d}s\leq C\,\tau,\end{split} (121)

where (3), the Malliavin integration by parts formula (36), (47) in Proposition 3.2 and (iv) in Lemma 4.1 with μ=1\mu=1 were used. Therefore, we obtain |K1|Cτ|K_{1}|\leq C\,\tau and |K2|Cτ|K_{2}|\leq C\,\tau. The proof is thus complete. ∎

Corollary 4.1.

As a by-product of the weak error analysis, one can easily obtain the rates of the strong error, for NN\in\mathbb{N} and m{1,2,,M}m\in\{1,2,\cdots,M\},

X(tm)XtmM,NL2(Ω,H˙)X¯(tm)X¯tmM,NL2(Ω,H˙)+𝒪tm𝒪tmM,NL2(Ω,H˙)X¯(tm)X¯N(tm)L2(Ω,H˙)+X¯N(tm)X¯tmM,NL2(Ω,H˙)+𝒪tm𝒪tmNL2(Ω,H˙)+𝒪tmN𝒪tmM,NL2(Ω,H˙)C(λN2+τ+λN32+τ34)C(λN32+τ34),\begin{split}\|X(t_{m})-X_{t_{m}}^{M,N}\|_{L^{2}(\Omega,\dot{H})}&\leq\|\bar{X}(t_{m})-\bar{X}_{t_{m}}^{M,N}\|_{L^{2}(\Omega,\dot{H})}+\|\mathcal{O}_{t_{m}}-\mathcal{O}_{t_{m}}^{M,N}\|_{L^{2}(\Omega,\dot{H})}\\ &\leq\|\bar{X}(t_{m})-\bar{X}^{N}(t_{m})\|_{L^{2}(\Omega,\dot{H})}+\|\bar{X}^{N}(t_{m})-\bar{X}_{t_{m}}^{M,N}\|_{L^{2}(\Omega,\dot{H})}\\ &\quad+\|\mathcal{O}_{t_{m}}-\mathcal{O}_{t_{m}}^{N}\|_{L^{2}(\Omega,\dot{H})}+\|\mathcal{O}_{t_{m}}^{N}-\mathcal{O}_{t_{m}}^{M,N}\|_{L^{2}(\Omega,\dot{H})}\\ &\leq C(\lambda_{N}^{-2}+\tau+\lambda_{N}^{-\frac{3}{2}}+\tau^{\frac{3}{4}})\\ &\leq C(\lambda_{N}^{-\frac{3}{2}}+\tau^{\frac{3}{4}}),\end{split} (122)

where the third inequality follows from (74), (118), (76) and (120) with tmt_{m} instead of TT, successively. The strong error estimates here, the same as that in [16, 27, 28], coincide with the spatial regularity of X(t)X(t), and thus are optimal.

Remark 4.1.

It is worthwhile to mention that the obtained weak convergence rate in time (i.e., 𝒪(τ)\mathcal{O}(\tau)) is optimal for the Euler–type method applying to stochastic differential equation.

Acknowledgements

M. Cai and S. Gan are supported by NSF of China (No. 11971488). M. Cai is supported by the China Scholarship Council. Y. Hu is supported by an NSERC discovery grant. We are very grateful to the referees for the interesting and constructive comments and suggestions.

References

  • [1] A. Andersson and S. Larsson, Weak convergence for a spatial approximation of the nonlinear stochastic heat equation, Math. Comp. 85 (2016) 1335–1358.
  • [2] D. C. Antonopoulou, D. Farazakis and G. Karali, Malliavin calculus for the stochastic Cahn–Hilliard/Allen–Cahn equation with unbounded noise diffusion, J. Differential Equations 265 (2018) 3168–3211.
  • [3] D. C. Antonopoulou, G. Karali and A. Millet, Existence and regularity of solution for a stochastic Cahn–Hilliard/Allen–Cahn equation with unbounded noise diffusion, J. Differential Equations 260 (2016) 2383–2417.
  • [4] C-E. Bréhier, J. Cui and X. Wang, Weak error estimates of fully-discrete schemes for the stochastic Cahn-Hilliard equation, arXiv preprint arXiv:2207.09266, (2022).
  • [5] C-E. Bréhier and L. Goudenège, Weak convergence rates of splitting schemes for the stochastic Allen–Cahn equation, BIT 60 (2019) 543–582.
  • [6] J. W. Cahn, On spinodal decomposition, Acta Metall. 9 (1961) 795–801.
  • [7] J. W. Cahn and J. E. Hilliard, Free energy of a nonuniform system. I. Interfacial free energy, J. Chem. Phys. 28 (1958) 258–267.
  • [8] J. W. Cahn and J. E. Hilliard, Spinodal decomposition: A reprise, Acta Metall. 19 (1971) 151–161.
  • [9] M. Cai, S. Gan and X. Wang, Weak convergence rates for an explicit full-discretization of stochastic Allen–Cahn equation with additive noise, J. Sci. Comput. 86 (2021) Paper No. 34.
  • [10] C. Cardon-Weber, Cahn–Hilliard stochastic equation: existence of the solution and of its density, Bernoulli 7 (2001) 777–816.
  • [11] S. Chai, Y. Cao, Y. Zou and W. Zhao, Conforming finite element methods for the stochastic Cahn–Hilliard–Cook equation, Appl. Numer. Math. 124 (2018) 44–56.
  • [12] C.-K. Cho and H. J. Choe, The Asymptotic Behaviour of Solutions of a Porous Medium Equation with Bounded Measurable Coefficients, J. Math. Anal. Appl. 210 (1997) 241–256.
  • [13] J. Cui and J. Hong, Strong and weak convergence rates of a spatial approximation for stochastic partial differential equation with one-sided Lipschitz coefficient, SIAM J. Numer. Anal. 57 (2019) 1815–1841.
  • [14] J. Cui and J. Hong, Wellposedness and regularity estimate for stochastic Cahn–Hilliard equation with unbounded noise diffusion, arXiv preprint arXiv:1907.02249, (2019).
  • [15] J. Cui and J. Hong, Absolute continuity and numerical approximation of stochastic Cahn–Hilliard equation with unbounded noise diffusion, J. Differential Equations 269 (2020) 10143–10180.
  • [16] J. Cui, J. Hong and L. Sun, Strong convergence of full discretization for stochastic Cahn–Hilliard equation driven by additive noise, SIAM J. Numer. Anal. 59 (2021) 2866–2899.
  • [17] J. Cui, J. Hong and L. Sun, Weak convergence and invariant measure of a full discretization for parabolic SPDEs with non-globally Lipschitz coefficients, Stochastic Process. Appl. 134 (2021) 55–93.
  • [18] G. Da Prato and A. Debussche, Stochastic Cahn–Hilliard equation, Nonlinear Anal. 26 (1996) 241–263.
  • [19] G. Da Prato and J. Zabczyk, Stochastic Equations in Infinite Dimensions, 2nd edn. Cambridge University Press, 2014.
  • [20] N. Elezović and A. Mikelić, On the stochastic Cahn–Hilliard equation, Nonlinear Anal. 16 (1991) 1169–1200.
  • [21] D. Furihata, M. Kovács, S. Larsson and F. Lindgren, Strong convergence of a fully discrete finite element approximation of the stochastic Cahn–Hilliard equation, SIAM J. Numer. Anal. 56 (2018) 708–731.
  • [22] S. Kim, Fractional order sobolev spaces for the Neumann Laplacian and the vector Laplacian, J. Korean Math. Soc. 57 (2020) 721–745.
  • [23] G. T. Kossioris and G. E. Zouraris, Finite element approximations for a linear Cahn–Hilliard–Cook equation driven by the space derivative of a space-time white noise, Discrete Contin. Dyn. Syst. Ser. B 18 (2013) 1845–1872.
  • [24] M. Kovács, S. Larsson and A. Mesforush, Finite element approximation of the Cahn–Hilliard–Cook equation, SIAM J. Numer. Anal. 49 (2011) 2407–2429.
  • [25] S. Larsson and A. Mesforush, Finite-element approximation of the linearized Cahn–Hilliard–Cook equation, IMA J. Numer. Anal. 31 (2011) 1315–1333.
  • [26] D. Nualart, The Malliavin Calculus and Related Topics, Springer-Verlag GmbH, 2006.
  • [27] R. Qi, M. Cai and X. Wang, Strong convergence rates of a fully discrete scheme for the Cahn–Hilliard–Cook equation, arXiv preprint arXiv: 2204.01630, (2022).
  • [28] R. Qi and X. Wang, Error estimates of semidiscrete and fully discrete finite element methods for the Cahn–Hilliard–Cook equation, SIAM J. Numer. Anal. 58 (2020) 1613–1653.
  • [29] A. Stuart and A. Humphries, Dynamical Systems and Numerical Analysis, Cambridge University Press, Cambridge, 1996.
  • [30] A. Yagi, Abstract Parabolic Evolution Equations and their Applications, Springer Berlin Heidelberg, 2010.