1 Introduction

Let \(T>0\), \(D:=(0,1)\), and \((\Omega ,{\mathcal {F}},P)\) be a complete probability space. Then, we consider the model initial- and Dirichlet boundary- value problem for a linearized Cahn–Hilliard–Cook equation formulated in Kossioris and Zouraris (2013), which is as follows: find a stochastic function \(u:[0,T]\times {\overline{D}}\rightarrow {\mathbb {R}}\), such that

$$\begin{aligned} \begin{array}{l} u_t+u_{xxxx}+\mu \,u_{xx} =\partial _x{\dot{W}}(t,x) \quad \forall \,(t,x)\in (0,T]\times D,\\ \qquad \qquad \quad \,\,\, u(t,\cdot )\big |_{\scriptscriptstyle \partial D}=u_{xx}(t,\cdot )\big |_{\scriptscriptstyle \partial D}=0 \quad \forall \,t\in (0,T], \\ \qquad \qquad \qquad \,\,\, u(0,x)=0\quad \forall \,x\in D, \end{array} \end{aligned}$$
(1.1)

a.s. in \(\Omega \), where \({\dot{W}}\) denotes a space-time white noise on \([0,T]\times D\) (see, e.g., Walsh 1986; Kallianpur and Xiong 1995) and \(\mu \) is a real constant. We recall that the mild solution to the problem above (cf. Debussche and Zambotti 2007) is given by

$$\begin{aligned} u(t,x)=\int _0^t\int _{\scriptscriptstyle D}{\mathsf{\Psi }}_{t-s}(x,y)\,\mathrm{d}W(s,y), \end{aligned}$$
(1.2)

where

$$\begin{aligned} {\mathsf{\Psi }}_t(x,y):= & {} \,-\sum _{k=1}^{\infty }\lambda _k\,e^{-\lambda _k^2\,(\lambda _k^2-\mu ) t} \,\varepsilon _k(x)\,\varphi _k(y)\nonumber \\= & {} \,-\partial _y\mathsf{G}_t(x,y) \quad \forall \,(t,x,y)\in (0,T]\times {\overline{D}}\times {\overline{D}}, \end{aligned}$$
(1.3)

\(\lambda _k:=k\,\pi \) for \(k\in {\mathbb {N}}\), \(\varepsilon _k(z):=\sqrt{2}\,\sin (\lambda _k\,z)\) and \(\varphi _k(z):=\sqrt{2}\,\cos (\lambda _k\,z)\) for \(z\in {\overline{D}}\) and \(k\in {\mathbb {N}}\), and \(\mathsf{G}_t(x,y)\) is the space-time Green kernel of the solution to the deterministic parabolic problem: find \(w:[0,T]\times {\overline{D}}\rightarrow {\mathbb {R}}\), such that

$$\begin{aligned} \begin{array}{l} w_t + w_{xxxx}+\mu \,w_{xx}= 0 \quad \forall \,(t,x)\in (0,T]\times D,\\ \qquad \qquad \quad \,\,\,\,\, w(t,\cdot )\big |_{\scriptscriptstyle \partial D}= w_{xx}(t,\cdot )\big |_{\scriptscriptstyle \partial D}=0 \quad \forall \,t\in (0,T],\\ \qquad \qquad \qquad \,\,\,\,\, w(0,x)=w_0(x)\quad \forall \,x\in D. \end{array} \end{aligned}$$
(1.4)

In the paper at hand, our goal is to propose and analyze a numerical method for the approximation of u that has less stability requirements and lower complexity than the method proposed in Kossioris and Zouraris (2013).

1.1 A canvas problem

A canvas problem is an initial- and boundary- value problem the solution to which: i) depends on a finite number of random variables and ii) is a regular approximation of the mild solution u to (1.1). Then, we can derive computable approximations of u by constructing numerical approximations of the canvas problem solution via the application of a discretization technique for stochastic partial differential equations with random coefficients. The formulation of the canvas problem depends on the way which we replace the infinite stochastic dimensionality of the problem (1.1) by a finite one.

In our case, the canvas problem is formulated as follows (cf. Allen et al. 1998; Kossioris and Zouraris 2010, 2013): Let \(\mathsf{M},\mathsf{N}\in {\mathbb {N}}\), \(\Delta {t}:=\frac{T}{\mathsf{N}}\), and \(t_n:=n\,\Delta {t}\) for \(n=0,\dots ,\mathsf{N}\), \(\mathsf{T}_n:=(t_{n-1},t_n)\) for \(n=1,\dots ,\mathsf{N}\), and \(\mathsf{u}:[0,T]\times {\overline{D}}\rightarrow {\mathbb {R}}\), such that

$$\begin{aligned} \begin{array}{l} \mathsf{u}_t +\mathsf{u}_{xxxx} +\mu \,\mathsf{u}_{xx}=\partial _x{\mathcal {W}} \quad {\mathrm{in}} \ \ (0,T]\times D,\\ \qquad \qquad \quad \,\,\, \mathsf{u}(t,\cdot )\big |_{\scriptscriptstyle \partial D}= \mathsf{u}_{xx}(t,\cdot )\big |_{\scriptscriptstyle \partial D}=0 \quad \forall \,t\in (0,T],\\ \qquad \qquad \qquad \,\,\, \mathsf{u}(0,x)=0\quad \forall \,x\in D, \end{array} \end{aligned}$$
(1.5)

where

$$\begin{aligned}&{\mathcal {W}}(\cdot ,x)|_{\scriptscriptstyle \mathsf{T}_n} := \tfrac{1}{\Delta {t}}\, \sum _{i=1}^{\scriptscriptstyle \mathsf{M}} R^n_i\,\varphi _i(x) \quad \forall \,x\in D,\quad n=1,\dots ,\mathsf{N}, \end{aligned}$$
(1.6)
$$\begin{aligned}&R^n_i:=\int _{\scriptscriptstyle \mathsf{T}_n}\int _{\scriptscriptstyle D}\varphi _i(x)\;\mathrm{d}W(t,x) =B^i(t_{n+1})-B^i(t_n), \quad i=1,\dots ,\mathsf{M},\quad n=1,\dots ,\mathsf{N},\nonumber \\ \end{aligned}$$
(1.7)

and \(B^i(t):=\int _0^t\int _{\scriptscriptstyle D}\varphi _i(x)\;\mathrm{d}W(s,x)\) for \(t\ge 0\) and \(i\in {\mathbb {N}}\). According to Walsh (1986), \((B^i)_{i=1}^{\infty }\) is a family of independent Brownian motions, and thus, the random variables \(\left( \left( R^n_i\right) _{n=1}^{\scriptscriptstyle \mathsf{N}}\right) _{i=1}^{\scriptscriptstyle \mathsf M}\) are independent and satisfy

$$\begin{aligned} R^n_i\sim {\mathcal {N}}(0,\Delta {t}),\quad i=1,\dots ,\mathsf{M},\quad n=1,\dots ,\mathsf{N}. \end{aligned}$$
(1.8)

Thus, the solution \(\mathsf{u}\) to (1.5) depends on \(\mathsf{N}\mathsf{M}\) random variables and the well-known theory for parabolic problems (see, e.g, Lions and Magenes 1972) yields its regularity along with the following representation formula:

$$\begin{aligned} \mathsf{u}(t,x)&=\,\int _0^t\int _{\scriptscriptstyle D}\mathsf{G}_{t-s}(x,y) \,\partial _y{\mathcal {W}}(s,y)\,\mathrm{d}s\mathrm{d}y \nonumber \\&=\,\int _0^t\int _{\scriptscriptstyle D}{\mathsf{\Psi }}_{t-s}(x,y) \,{\mathcal {W}}(s,y) \,\mathrm{d}s\mathrm{d}y \quad \forall \,(t,x)\in [0,T]\times {\overline{D}}. \end{aligned}$$
(1.9)

Remark 1.1

In Kossioris and Zouraris (2013), the definition of \({\mathcal {W}}\) is based on a uniform partition of [0, T] in N subintervals and on a uniform partition of D in J subintervals. At every time-slab, \({\mathcal {W}}\) has a constant value with respect to the time variable, but, with respect to the space variable, is defined as the \(L^2(D)\)-projection of a random, piecewise constant function onto the space of linear splines, the computation of which leads to the numerical solution of a \((J+1)\times (J+1)\) tridiagonal linear system of algebraic equations. Finally, \({\mathcal {W}}\) depends on \(N (J+1)\) random variables and its construction has \(O(N\,(J+1))\) complexity, which must to be added to the complexity of the numerical method used for the approximation of \(\mathsf{u}\). On the contrary, the stochastic load \({\mathcal {W}}\) of the canvas problem (1.5) which we propose here is given explicitly by the formula (1.6), and thus, no extra computational cost is required for its formation.

1.2 An IMEX finite element method

Let \(M\in {\mathbb {N}}\), \(\Delta \tau :=\frac{T}{M}\), and \(\tau _m:=m\,\Delta \tau \) for \(m=0,\dots ,M\), and \(\Delta _m:=(\tau _{m-1},\tau _m)\) for \(m=1,\dots ,M\). In addition, for \(r=2\) or 3, let \(\mathsf{M}_h^r\subset H^2(D)\cap H_0^1(D)\) be a finite element space consisting of functions which are piecewise polynomials of degree at most r over a partition of D in intervals with maximum mesh length h.

The fully discrete method which we propose for the numerical approximation of \(\mathsf{u}\) uses an implicit/explicit (IMEX) time-discretization treatment of the space differential operator along with a finite element variational formulation for space discretization. Its algorithm is as follows: first, sets

$$\begin{aligned} \mathsf{U}_h^0:=0, \end{aligned}$$
(1.10)

and then, for \(m=1,\dots ,M\), finds \(\mathsf{U}_h^m\in \mathsf{M}_h^r\), such that

$$\begin{aligned} \left( \,\mathsf{U}_h^m-\mathsf{U}_h^{m-1},\chi \,\right) _{\scriptscriptstyle 0,D} {+}\Delta \tau \left[ (\partial _x^2\mathsf{U}_h^{m},\partial _x^2\chi \,)_{\scriptscriptstyle 0,D} {+}\mu (\partial _x^2\mathsf{U}_h^{m-1},\chi \,)_{\scriptscriptstyle 0,D}\,\right] {=}\int _{\scriptscriptstyle \Delta _m}\left( \partial _x{\mathcal {W}},\chi \right) _{\scriptscriptstyle 0,D}\,\mathrm{d}\tau , \end{aligned}$$
(1.11)

for all \(\chi \in \mathsf{M}_h^r\), where \((\cdot ,\cdot )_{\scriptscriptstyle 0,D}\) is the usual \(L^2(D)\)-inner product.

Remark 1.2

It is easily seen that the numerical method above is unconditionally stable, while the Backward Euler finite element method is stable under the time-step restriction: \(\Delta \tau \,\mu ^2\le 4\) (see Kossioris and Zouraris 2013).

1.3 An overview of the paper

In Sect. 2, we introduce notation and we recall several results that are often used in the rest of the paper. In Sect. 3, we focus on the estimation of the error which we made by approximating the solution u to (1.1) by the solution \(\mathsf{u}\) to (1.5), arriving at the bound

$$\begin{aligned} \max _{\scriptscriptstyle [0,T]}\left( \,{\mathbb {E}}\left[ \, \Vert u-\mathsf{u}\Vert ^2_{\scriptscriptstyle L^2(D)}\,\right] \,\right) ^{\frac{1}{2}} \le \,C\,\left( \,\mathsf{M}^{-\frac{1}{2}}+{\Delta t}^{\frac{1}{8}}\,\right) \end{aligned}$$

(see Theorem 3.1). Section 4 is dedicated to the definition and the convergence analysis of modified IMEX time-discrete and fully discrete approximations of the solution w to the deterministic problem (1.4). The results obtained are used later in Sect. 5, where we analyze the numerical method for the approximation of \(\mathsf{u}\), given in Sect. 1.2. Its convergence is established by proving the following strong error estimate:

$$\begin{aligned} \max _{0\le {m}\le {\scriptscriptstyle M}}\left( \, {\mathbb {E}}\left[ \,\Vert \mathsf{U}_h^m-\mathsf{u}(\tau _m,\cdot )\Vert ^2_{\scriptscriptstyle 0,D}\,\right] \right) ^{\frac{1}{2}} \le \,C\, \,\left( \,\epsilon ^{-\frac{1}{2}}_1\,\Delta \tau ^{\frac{1}{8}-\epsilon _1} +\epsilon _2^{-\frac{1}{2}}\,\,\,h^{\frac{r}{6}-\epsilon _2}\,\right) \end{aligned}$$

for all \(\epsilon _1\in \left( 0,\frac{1}{8}\right] \) and \(\epsilon _2\in \left( 0,\frac{r}{6}\right] \) (see Theorem 5.3). We obtain the latter error bound, by applying a discrete Duhamel principle technique to estimate separately the time-discretization error and the space-discretization error, which are defined using as an intermediate the corresponding IMEX time-discrete approximations of \(\mathsf{u}\), specified by (5.1) and (5.2) (cf., e.g., Kossioris and Zouraris 2010, 2013; Yan 2005).

Since we have no assumptions on the sign, or, the size of \(\mu \), the elliptic operator in (1.5) is, in general, not invertible. This is the reason that the Backward Euler/finite element method is stable and convergent after adopting a restriction on the time-step size (see Kossioris and Zouraris 2013, Remark 1.2). On the contrary, the IMEX/finite element method which we propose here is unconditionally stable and convergent, because the principal part of the elliptic operator is treated implicitly and its lower order part explicitly. Another characteristic in our method is the choice to build up the canvas problem using spectral functions, which allow us to avoid the numerical solution of an extra linear system of algebraic equation at every time-step that is required in the approach of Kossioris and Zouraris (2013) (see Remark 1.1).

The error analysis of the IMEX finite element method is more technical than that in Kossioris and Zouraris (2013) for the Backward Euler finite element method. The main difference is due to the fact that the representation of the time-discrete and fully discrete approximations of \(\mathsf{u}\) is related to a modified version of the IMEX time-stepping method for the approximation of the solution to the deterministic problem (1.4), the error analysis of which is necessary in obtaining the desired error estimate and is of independent interest (see Sect. 4).

2 Preliminaries

We denote by \(L^2(D)\) the space of the Lebesgue measurable functions which are square integrable on D with respect to the Lebesgue measure \(\mathrm{d}x\). The space \(L^2(D)\) is provided with the standard norm \(\Vert g\Vert _{\scriptscriptstyle 0,D}:= \left( \int _{\scriptscriptstyle D}|g(x)|^2\,\mathrm{d}x\right) ^{\frac{1}{2}}\) for \(g\in L^2(D)\), which is derived by the usual inner product \((g_1,g_2)_{\scriptscriptstyle 0,D}:=\int _{\scriptscriptstyle D}g_1(x)\,g_2(x)\,\mathrm{d}x\) for \(g_1\), \(g_2\in L^2(D)\). In addition, we employ the symbol \({\mathbb {N}}_0\) for the set of all nonnegative integers.

For \(s\in {\mathbb {N}}_0\), we denote by \(H^s(D)\) the Sobolev space of functions having generalized derivatives up to order s in \(L^2(D)\), and by \(\Vert \cdot \Vert _{\scriptscriptstyle s,D}\) its usual norm, i.e., \(\Vert g\Vert _{\scriptscriptstyle s,D}:=\left( \sum _{\ell =0}^s \Vert \partial _x^{\ell }g\Vert _{\scriptscriptstyle 0,D}^2\right) ^{\scriptscriptstyle 1/2}\) for \(g\in H^s(D)\). In addition, by \(H_0^1(D)\), we denote the subspace of \(H^1(D)\) consisting of functions which vanish at the endpoints of D in the sense of trace.

The sequence of pairs \(\left\{ \left( \lambda _i^2,\varepsilon _i\right) \right\} _{i=1}^{\infty }\) is a solution to the eigenvalue/eigenfunction problem: find nonzero \(\varphi \in H^2(D)\cap H_0^1(D)\) and \(\lambda \in {\mathbb {R}}\), such that \(-\varphi ''=\lambda \,\varphi \) in D. Since \((\varepsilon _i)_{i=1}^{\infty }\) is a complete \((\cdot ,\cdot )_{\scriptscriptstyle 0,D}\)-orthonormal system in \(L^2(D)\), for \(s\in {\mathbb {R}}\), we define by

$$\begin{aligned} {\mathcal {V}}^s(D):=\Bigg \{v\in L^2(D):\quad \sum _{i=1}^{\infty } \lambda _{i}^{2s} \,(v,\varepsilon _i)^2_{\scriptscriptstyle 0,D}<\infty \,\Bigg \} \end{aligned}$$

a subspace of \(L^2(D)\) provided with the natural norm \(\Vert v\Vert _{\scriptscriptstyle {\mathcal {V}}^s}:=\big (\,\sum _{i=1}^{\infty } \lambda _{i}^{2s}\,(v,\varepsilon _i)^2_{\scriptscriptstyle 0,D}\,\big )^{\scriptscriptstyle 1/2}\) for \(v\in {\mathcal {V}}^s(D)\). For \(s\ge 0\), the space \(({\mathcal {V}}^s(D),\Vert \cdot \Vert _{\scriptscriptstyle {\mathcal {V}}^s})\) is a complete subspace of \(L^2(D)\) and we define \(({\dot{\mathbf{H}}}^{{s}}(D),\Vert \cdot \Vert _{\scriptscriptstyle {\dot{\mathbf{H}}}^s}) :=({\mathcal {V}}^s(D),\Vert \cdot \Vert _{\scriptscriptstyle {\mathcal {V}}^s})\). For \(s<0\), the space \(({\dot{\mathbf{H}}}^s(D),\Vert \cdot \Vert _{\scriptscriptstyle {\dot{\mathbf{H}}}^s})\) is defined as the completion of \(({\mathcal {V}}^s(D),\Vert \cdot \Vert _{\scriptscriptstyle {\mathcal {V}}^s})\), or, equivalently, as the dual of \(({\dot{\mathbf{H}}}^{-s}(D),\Vert \cdot \Vert _{\scriptscriptstyle {\dot{\mathbf{H}}}^{-s}})\).

Let \(m\in {\mathbb {N}}_0\). It is well known (see Thomée 1997) that

$$\begin{aligned} {\dot{\mathbf{H}}}^m(D)=\left\{ \,v\in H^m(D): \quad \partial ^{2\ell }v\left| _{\scriptscriptstyle \partial D}\right. =0 \quad \mathrm{if} \ \ 0\le {2\ell }<m\,\right\} \end{aligned}$$

and that there exist constants \(C_{m,{\scriptscriptstyle A}}\) and \(C_{m,{\scriptscriptstyle B}}\), such that

$$\begin{aligned} C_{m,{\scriptscriptstyle A}}\,\Vert v\Vert _{\scriptscriptstyle m,D} \le \Vert v\Vert _{\scriptscriptstyle {\dot{\mathbf{H}}}^m} \le \,C_{m,{\scriptscriptstyle B}}\,\Vert v\Vert _{\scriptscriptstyle m,D}\quad \forall \,v\in {\dot{\mathbf{H}}}^m(D). \end{aligned}$$
(2.1)

In addition, we define on \(L^2(D)\) the negative norm \(\Vert \cdot \Vert _{\scriptscriptstyle -m, D}\) by

$$\begin{aligned} \Vert v\Vert _{\scriptscriptstyle -m, D}:=\sup \Big \{ \tfrac{(v,\varphi )_{\scriptscriptstyle 0,D}}{\Vert \varphi \Vert _{\scriptscriptstyle m,D}}:\quad \varphi \in {\dot{\mathbf{H}}}^m(D) \ \ \mathrm{and} \ \ \varphi \not =0\Big \} \quad \forall \,v\in L^2(D), \end{aligned}$$

for which, using (2.1), follows that there exists a constant \(C_{-m}>0\), such that:

$$\begin{aligned} \Vert v\Vert _{\scriptscriptstyle -m,D}\le \,C_{-m}\,\Vert v\Vert _{{\dot{\mathbf{H}}}^{-m}} \quad \forall \,v\in L^2(D). \end{aligned}$$
(2.2)

Let \({\mathbb {L}}_2=(L^2(D),(\cdot ,\cdot )_{\scriptscriptstyle 0,D})\) and \({\mathcal {L}}({\mathbb {L}}_2)\) be the space of linear, bounded operators from \({\mathbb {L}}_2\) to \({\mathbb {L}}_2\). An operator \(\Gamma \in {\mathcal {L}}({\mathbb {L}}_2)\) is Hilbert–Schmidt, when \(\Vert \Gamma \Vert _{\scriptscriptstyle \mathrm HS}:=\left( \sum _{i=1}^{\infty } \Vert \Gamma \varepsilon _i\Vert ^2_{\scriptscriptstyle 0,D}\right) ^{\frac{1}{2}}<+\infty \), where \(\Vert \Gamma \Vert _{\scriptscriptstyle \mathrm HS}\) is the so-called Hilbert–Schmidt norm of \(\Gamma \). We note that the quantity \(\Vert \Gamma \Vert _{\scriptscriptstyle \mathrm HS}\) does not change when we replace \((\varepsilon _i)_{i=1}^{\infty }\) by another complete orthonormal system of \({\mathbb {L}}_2\). It is well known (see, e.g., Dunford and Schwartz 1988; Lord et al. 2014) that an operator \(\Gamma \in {\mathcal {L}}({\mathbb {L}}_2)\) is Hilbert–Schmidt iff there exists a measurable function \(\gamma :D\times D\rightarrow {\mathbb {R}}\), such that \(\Gamma [v](\cdot )=\int _{\scriptscriptstyle D}\gamma (\cdot ,y)\,v(y)\,dy\) for \(v\in L^2(D)\), and then, it holds that

$$\begin{aligned} \Vert \Gamma \Vert _{\scriptscriptstyle \mathrm HS} =\left( \, \iint _{\scriptscriptstyle D\times D}\gamma ^2(x,y)\,\mathrm{d}x\mathrm{d}y\,\right) ^{\frac{1}{2}}. \end{aligned}$$
(2.3)

Let \({\mathcal {L}}_{\scriptscriptstyle \mathrm HS}({\mathbb {L}}_2)\) be the set of Hilbert–Schmidt operators of \({\mathcal {L}}({\mathbb {L}}^2)\) and \(\Phi :[0,T]\rightarrow {\mathcal {L}}_{\scriptscriptstyle \mathrm HS}({\mathbb {L}}_2)\). In addition, for a random variable X, let \({\mathbb {E}}[X]\) be its expected value, i.e., \({\mathbb {E}}[X]:=\int _{\scriptscriptstyle \Omega }X\,\mathrm{d}P\). Then, the Itô isometry property for stochastic integrals reads

$$\begin{aligned} {\mathbb {E}}\left[ \Big \Vert \int _0^{\scriptscriptstyle T}\Phi \,\mathrm{d}W\Big \Vert _{\scriptscriptstyle 0,D}^2\right] =\int _0^{\scriptscriptstyle T}\Vert \Phi (t)\Vert _{\scriptscriptstyle \mathrm HS}^2\,\mathrm{d}t. \end{aligned}$$
(2.4)

For later use, we recall that if \(({\mathcal {H}},(\cdot ,\cdot )_{\scriptscriptstyle {\mathcal {H}}})\) is a real inner product space with induced norm \(|\cdot |_{\scriptscriptstyle {\mathcal {H}}}\), then

$$\begin{aligned} 2\,(g-v,g)_{\scriptscriptstyle {\mathcal {H}}} =|g|^2_{\scriptscriptstyle {\mathcal {H}}}-|v|^2_{\scriptscriptstyle {\mathcal {H}}} +|g-v|^2_{\scriptscriptstyle {\mathcal {H}}}\quad \forall \,g,v\in {\mathcal {H}}. \end{aligned}$$
(2.5)

Finally, for any nonempty set A, we denote by \({\mathcal {X}}_{\scriptscriptstyle A}\) the indicator function of A.

2.1 A projection operator

Let \({\mathcal {O}}:=(0,T)\times D\), \({\mathfrak S}_{\scriptscriptstyle \mathsf{M}}:={\mathrm{span}}(\varphi _i)_{i=1}^{\scriptscriptstyle \mathsf{M}}\), \({\mathfrak S}_{\scriptscriptstyle \mathsf{N}}:=\mathrm{span}({\mathcal {X}}_{\scriptscriptstyle T_n})_{n=1}^{\scriptscriptstyle \mathsf{N}}\) and \(\mathsf{\Pi }:L^2({\mathcal {O}})\rightarrow {\mathfrak S}_{\scriptscriptstyle \mathsf{N}}\otimes {\mathfrak S}_{\scriptscriptstyle \mathsf{M}}\), the usual \(L^2({\mathcal {O}})\)-projection operator which is given by the formula:

$$\begin{aligned} \mathsf{\Pi }g:=\tfrac{1}{\Delta {t}}\,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\,\left( \,\sum _{n=1}^{\scriptscriptstyle \mathsf{N}} {\mathcal {X}}_{\scriptscriptstyle {T_n}}\,\int _{\scriptscriptstyle T_n} (g,\varphi _i)_{\scriptscriptstyle 0,D}\,\mathrm{d}t\right) \,\varphi _i \quad \forall \,g\in L^2({\mathcal {O}}). \end{aligned}$$
(2.6)

Then, the following representation of the stochastic integral of \(\mathsf{\Pi }\) holds [cf. Lemma 2.1 in Kossioris and Zouraris (2010)].

Lemma 2.1

For \(g\in L^2({\mathcal {O}})\), it holds that

$$\begin{aligned} \int _0^{\scriptscriptstyle T}\int _{\scriptscriptstyle D}\mathsf{\Pi }g(t,x)\,\mathrm{d}W(t,x) =\iint _{\scriptscriptstyle {\mathcal {O}}}{\mathcal {W}}(s,y)\,g(s,y)\,\mathrm{d}s\mathrm{d}y. \end{aligned}$$
(2.7)

Proof

Using (2.6) and (1.7), we have

$$\begin{aligned} \begin{aligned} \int _0^{\scriptscriptstyle T}\int _{\scriptscriptstyle D}\mathsf{\Pi }g(t,x)\,\mathrm{d}W(t,x) =&\tfrac{1}{\Delta {t}}\, \sum _{n=1}^{\scriptscriptstyle \mathsf{N}} \,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\left( \, \int _{\scriptscriptstyle T_n}\int _{\scriptscriptstyle D}g(s,y)\,\varphi _i(y)\;\mathrm{d}s\mathrm{d}y\,\right) \, R_i^{n}\\ =&\tfrac{1}{\Delta {t}}\, \sum _{n=1}^{\scriptscriptstyle \mathsf{N}} \,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\left( \, \iint _{\scriptscriptstyle {\mathcal {O}}}{\mathcal {X}}_{\scriptscriptstyle T_n}(s)\,R_i^{n}\,g(s,y)\,\varphi _i(y)\;\mathrm{d}s\mathrm{d}y \,\right) \\ =&\,\iint _{\scriptscriptstyle {\mathcal {O}}}g(s,y)\,\left( \,\tfrac{1}{\Delta {t}}\,\sum _{n=1}^{\scriptscriptstyle \mathsf{N}} \,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}{\mathcal {X}}_{\scriptscriptstyle T_n}(s)\,R_i^{n}\,\varphi _i(y)\,\right) \;\mathrm{d}s\mathrm{d}y\\ \end{aligned} \end{aligned}$$

which along (1.6) yields (2.7). \(\square \)

2.2 Linear elliptic and parabolic operators

Let \(T_{\scriptscriptstyle E}:L^2(D)\rightarrow {\dot{\mathbf{H}}}^2(D)\) be the solution operator of the Dirichlet two-point boundary-value problem: for given \(f\in L^2(D)\) find \(v_{\scriptscriptstyle E}\in {\dot{\mathbf{H}}}^2(D)\), such that \(v_{\scriptscriptstyle E}''=f\) in D, i.e., \(T_{\scriptscriptstyle E}f:=v_{\scriptscriptstyle E}\). It is well known that

$$\begin{aligned} (T_{\scriptscriptstyle E}f,g)_{\scriptscriptstyle 0,D}=(f,T_{\scriptscriptstyle E}g)_{\scriptscriptstyle 0,D} \quad \forall \,f,g\in L^2(D), \end{aligned}$$
(2.8)

and, for \(m\in {\mathbb {N}}_0\), there exists a constant \(C_{\scriptscriptstyle E}^m>0\), such that

$$\begin{aligned} \Vert T_{\scriptscriptstyle E}f\Vert _{\scriptscriptstyle m,D}\le \,C_{\scriptscriptstyle E}^m \,\Vert f\Vert _{\scriptscriptstyle m-2, D} \quad \forall \,f\in H^{\max \{0,m-2\}}(D). \end{aligned}$$
(2.9)

Let, also, \(T_{\scriptscriptstyle B}:L^2(D)\rightarrow {\dot{\mathbf{H}}}^4(D)\) be the solution operator of the following Dirichlet biharmonic two-point boundary-value problem: for given \(f\in L^2(D)\) find \(v_{\scriptscriptstyle B}\in {\dot{\mathbf{H}}}^4(D)\), such that

$$\begin{aligned} v_{\scriptscriptstyle B}''''=f\quad \mathrm{in} \ \ D, \end{aligned}$$
(2.10)

i.e., \(T_{\scriptscriptstyle B}f:=v_{\scriptscriptstyle B}\). It is well known that, for \(m\in {\mathbb {N}}_0\), there exists a constant \(C^{m}_{\scriptscriptstyle B}>0\), such that

$$\begin{aligned} \Vert T_{\scriptscriptstyle B}f\Vert _{\scriptscriptstyle m,D}\le \,C^m_{\scriptscriptstyle B}\,\Vert f\Vert _{\scriptscriptstyle m-4, D} \quad \forall \,f\in H^{\max \{0,m-4\}}(D). \end{aligned}$$
(2.11)

Due to the type of boundary conditions of (2.10), we have

$$\begin{aligned} T_{\scriptscriptstyle B}f= T_{\scriptscriptstyle E}^2f\quad \forall \,f\in L^2(D), \end{aligned}$$
(2.12)

which, after using (2.8), yields

$$\begin{aligned} (T_{\scriptscriptstyle B}v_1,v_2)_{\scriptscriptstyle 0,D} =(T_{\scriptscriptstyle E}v_1,T_{\scriptscriptstyle E}v_2)_{\scriptscriptstyle 0,D}=(v_1,T_{\scriptscriptstyle B}v_2)_{\scriptscriptstyle 0,D} \quad \forall \,v_1,v_2\in L^2(D). \end{aligned}$$
(2.13)

Let \(({\mathcal {S}}(t)w_0)_{\scriptscriptstyle t\in [0,T]}\) be the standard semigroup notation for the solution w to (1.4). Then [see Appendix A in Kossioris and Zouraris (2013)], for \(\ell \in {\mathbb {N}}_0\), \(\beta \ge 0\) and \(p\ge 0\), there exists a constant \({\mathcal {C}}_{\beta ,\ell ,\mu ,\mu ^2 T}>0\), such that

$$\begin{aligned} \int _{t_a}^{t_b}(\tau -t_a)^{\beta }\, \big \Vert \partial _t^{\ell }{\mathcal {S}}(\tau )w_0 \big \Vert _{\scriptscriptstyle {\dot{\mathbf{H}}}^p}^2\,\mathrm{d}\tau \le \,{\mathcal {C}}_{\beta ,\ell ,\mu ,\mu ^2 T}\, \Vert w_0\Vert ^2_{\scriptscriptstyle {\dot{\mathbf{H}}}^{p+4\ell -2\beta -2}} \end{aligned}$$
(2.14)

for all \(w_0\in {\dot{\mathbf{H}}}^{p+4\ell -2\beta -2}(D)\) and \(t_a\), \(t_b\in [0,T]\) with \(t_b>t_a\).

2.3 Discrete operators

Let \(r=2\) or 3, and \(\mathsf{M}_h^r\subset H_0^1(D)\cap H^2(D)\) be a finite element space consisting of functions which are piecewise polynomials of degree at most r over a partition of D in intervals with maximum length h. It is well known (cf., e.g., Bramble and Hilbert 1970) that

$$\begin{aligned} \inf _{\chi \in \mathsf{M}_h^r} \Vert v-\chi \Vert _{\scriptscriptstyle 2,D} \le \,C_{r}\,h^{s-2}\,\Vert v\Vert _{\scriptscriptstyle s,D} \quad \,\forall \,v\in H^{s}(D)\cap H_0^1(D),\quad \,s=3,\dots ,r+1, \end{aligned}$$
(2.15)

where \(C_{r}\) is a positive constant that depends on r and D, and is independent of h and v. Then, we define the discrete biharmonic operator \(B_h:\mathsf{M}_h^r\rightarrow \mathsf{M}_h^r\) by \((B_h\varphi ,\chi )_{\scriptscriptstyle 0,D}=(\partial _x^2\varphi ,\partial _x^2\chi )_{\scriptscriptstyle 0,D}\) for \(\varphi ,\chi \in \mathsf{M}_h^r\), the \(L^2(D)\)-projection operator \(P_h:L^2(D)\rightarrow \mathsf{M}_h^r\) by \((P_hf,\chi )_{\scriptscriptstyle 0,D}=(f,\chi )_{\scriptscriptstyle 0,D}\) for \(\chi \in \mathsf{M}_h^r \) and \(f\in L^2(D)\), and the standard Galerkin finite element approximation \(v_{{\scriptscriptstyle B},h}\in \mathsf{M}_h^r\) of the solution \(v_{\scriptscriptstyle B}\) to (2.10) by requiring

$$\begin{aligned} B_h(v_{{\scriptscriptstyle B},h})=P_hf. \end{aligned}$$
(2.16)

Let \(T_{{\scriptscriptstyle B},h}:L^2(D)\rightarrow {\mathsf{M}^r_h}\) be the solution operator of the finite element method (2.16), i.e., \(T_{{\scriptscriptstyle B},h}f:=v_{{\scriptscriptstyle B},h}=B_h^{-1}P_hf\) for all \(f\in L^2(D)\). Then, we can easily conclude that

$$\begin{aligned} \left( T_{{\scriptscriptstyle B},h}f,g\right) _{\scriptscriptstyle 0,D}=\left( \partial _x^2\left( T_{{\scriptscriptstyle B},h}f\right) , \partial _x^2\left( T_{{\scriptscriptstyle B},h}g\right) \right) _{\scriptscriptstyle 0,D} =\left( f,T_{{\scriptscriptstyle B},h}g\right) _{\scriptscriptstyle 0,D} \quad \forall \,f,g\in L^2(D) \end{aligned}$$
(2.17)

and

$$\begin{aligned} \Vert \partial _x^2(T_{\scriptscriptstyle B,h}f)\Vert _{\scriptscriptstyle 0,D} \le \,C\,\Vert f\Vert _{\scriptscriptstyle -2,D}\quad \forall \,f\in L^2(D). \end{aligned}$$
(2.18)

Finally, the approximation property (2.15) of the finite element space \(\mathsf{M}_h^r\) yields (see, e.g., Proposition 2.2 in Kossioris and Zouraris 2010) the following error estimate:

$$\begin{aligned} \Vert T_{\scriptscriptstyle B}f-T_{{\scriptscriptstyle B},h}f\Vert _{\scriptscriptstyle 0,D} \le \,C\, h^{r}\,\Vert f\Vert _{\scriptscriptstyle -1,D}\quad \forall \,f\in L^2(D),\quad r=2,3. \end{aligned}$$
(2.19)

3 An approximation estimate for the canvas problem solution

Here, we establish the convergence of \(\mathsf{u}\) towards u with respect to the \(L^{\infty }_t(L^2_{\scriptscriptstyle P}(L^2_x))\) norm, when \(\Delta {t}\rightarrow 0\) and \(\mathsf{M}\rightarrow \infty \) (cf. Kossioris and Zouraris 2010, 2013).

Theorem 3.1

Let u be the solution to (1.1), \(\mathsf{u}\) be the solution to (1.5), and \(\kappa \in {\mathbb {N}}\), such that \(\kappa ^2\,\pi ^2>\mu \). Then, there exists a constant \({\widehat{c}}_{\scriptscriptstyle \mathrm{CER}}>0\), independent of \(\Delta {t}\) and \(\mathsf{M}\), such that

$$\begin{aligned} \max _{\scriptscriptstyle [0,T]}{\mathsf{\Theta }} \le \,{\widehat{c}}_{\scriptscriptstyle \mathrm{CER}}\,\left( \,\Delta {t}^\frac{1}{8} +\mathsf{M}^{-\frac{1}{2}}\,\right) \quad \forall \,\mathsf{M}\ge \kappa , \end{aligned}$$
(3.1)

where \({\mathsf{\Theta }}(t):=\left( {\mathbb {E}}\left[ \Vert u(t,\cdot )-\mathsf{u}(t,\cdot )\Vert _{\scriptscriptstyle 0,D}^2 \right] \right) ^{\frac{1}{2}}\) for \(t\in [0,T]\).

Proof

In the sequel, we will use the symbol C to denote a generic constant that is independent of \(\Delta {t}\) and \(\mathsf{M}\) and may change value from one line to the other.

Using (1.2), (1.9), and Lemma 2.1, we conclude that

$$\begin{aligned} u(t,x)-\mathsf{u}(t,x)=\int _0^{\scriptscriptstyle T}\int _{\scriptscriptstyle D} \big [{\mathcal {X}}_{(0,t)}(s)\,{\mathsf{\Psi }}_{t-s}(x,y) -{\widetilde{\mathsf{\Psi }}}(t,x;s,y)\big ]\,\mathrm{d}W(s,y), \end{aligned}$$
(3.2)

for \((t,x)\in [0,T]\times {\overline{D}}\), where \({\widetilde{\mathsf{\Psi }}}: (0,T)\times {D}\rightarrow L^2({\mathcal {D}})\) is given by

$$\begin{aligned} {\widetilde{\mathsf{\Psi }}}(t,x;s,y):=\tfrac{1}{\Delta {t}}\sum _{i=1}^{\scriptscriptstyle \mathsf{M}} \left[ \,\int _{\scriptscriptstyle T_n}{\mathcal {X}}_{(0,t)}(s') \left( \int _{\scriptscriptstyle D}{\mathsf{\Psi }}_{t-s'}(x,y')\,\varphi _i(y')\,dy'\right) \mathrm{d}s'\right] \varphi _i(y) \end{aligned}$$

for \((s,y)\in T_n\times {D}\), \(n=1,\dots ,\mathsf{N}\), and for \((t,x)\in (0,T]\times D\). Now, we use (1.3) and the \(L^2(D)\)-orthogonality of \((\varphi _k)_{k=1}^{\infty }\) to obtain

$$\begin{aligned} {\widetilde{\mathsf{\Psi }}}(t,x;s,y) =\tfrac{1}{\Delta {t}}\, \int _{\scriptscriptstyle T_n}{\mathcal {X}}_{(0,t)}(s')\left( \, \sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\lambda _i e^{-\lambda _i^2(\lambda _i^2-\mu )(t-s')} \varepsilon _i(x)\,\varphi _i(y)\right) \,\mathrm{d}s' \end{aligned}$$
(3.3)

for \((s,y)\in T_n\times {D}\), \(n=1,\dots ,\mathsf{N}\), and for \((t,x)\in (0,T]\times D\). In addition, we use (3.2), (2.4), and (2.3), to get

$$\begin{aligned} {\mathsf{\Theta }}(t)&=\left( \,\int _0^{\scriptscriptstyle T}\int _{\scriptscriptstyle D}\int _{\scriptscriptstyle D} \big [{\mathcal {X}}_{(0,t)}(s)\,{\mathsf{\Psi }}_{t-s}(x,y) -{\widetilde{\mathsf{\Psi }}}(t,x;s,y)\big ]^2\;\mathrm{d}x\mathrm{d}y\mathrm{d}s\,\right) ^{\frac{1}{2}}\nonumber \\&\le \sqrt{{\mathsf{\Theta }}_{\scriptscriptstyle A}(t)}+\sqrt{{\mathsf{\Theta }}_{\scriptscriptstyle B}(t)}\quad \forall \,t\in (0,T], \end{aligned}$$
(3.4)

where

$$\begin{aligned} {\mathsf{\Theta }}_{\scriptscriptstyle A}(t)\!:=\! \sum _{n=1}^{\scriptscriptstyle \mathsf{N}}\int _{\scriptscriptstyle D}\int _{\scriptscriptstyle D}\int _{\scriptscriptstyle T_n} \left[ {\mathcal {X}}_{(0,t)}(s)\,{\mathsf{\Psi }}_{t-s}(x,y) -\tfrac{1}{\Delta {t}}\,\int _{\scriptscriptstyle T_n} {\mathcal {X}}_{(0,t)}(s')\,{\mathsf{\Psi }}_{t-s'}(x,y)\,\mathrm{d}s'\,\right] ^2\;\mathrm{d}x\mathrm{d}y\mathrm{d}s \end{aligned}$$

and

$$\begin{aligned} {\mathsf{\Theta }}_{\scriptscriptstyle B}(t):=\sum _{n=1}^{\scriptscriptstyle \mathsf{N}} \int _{\scriptscriptstyle D}\int _{\scriptscriptstyle D}\int _{\scriptscriptstyle T_n}\left[ \, \tfrac{1}{\Delta {t}}\int _{\scriptscriptstyle T_n}{\mathcal {X}}_{(0,t)}(s')\,{\mathsf{\Psi }}_{t-s'}(x,y)\,\mathrm{d}s' -{\widetilde{\mathsf{\Psi }}}(t,x;s,y)\,\right] ^2\mathrm{d}x\mathrm{d}y\mathrm{d}s. \end{aligned}$$

Proceeding as in the proof of Theorem 3.1 in Kossioris and Zouraris (2013), we arrive at

$$\begin{aligned} \sqrt{{\mathsf{\Theta }}_{\scriptscriptstyle A}(t)}\le \,C\,\Delta {t}^{\frac{1}{8}}\quad \forall \,t\in (0,T]. \end{aligned}$$
(3.5)

Combining (3) and (3.3) and using the \(L^2(D)\)-orthogonality of \((\varepsilon _k)_{k=1}^{\infty }\) and \((\varphi _k)_{k=1}^{\infty }\), we have

$$\begin{aligned} {\mathsf{\Theta }}_{\scriptscriptstyle B}(t)= & {} \,\tfrac{1}{\Delta {t}}\, \sum _{n=1}^{\scriptscriptstyle \mathsf{N}} \int _{\scriptscriptstyle D}\int _{\scriptscriptstyle D}\left[ \int _{\scriptscriptstyle T_n}{\mathcal {X}}_{(0,t)}(s') \left( \,{\mathsf{\Psi }}_{t-s'}(x,y)-\sum _{i=1}^{\scriptscriptstyle \mathsf{M}} \lambda _ie^{-\lambda _i^2(\lambda _i^2-\mu )(t-s')} \varepsilon _i(x)\,\varphi _i(y)\right) \mathrm{d}s'\right] ^2\mathrm{d}x\mathrm{d}y\\= & {} \tfrac{1}{\Delta {t}}\, \sum _{n=1}^{\scriptscriptstyle \mathsf{N}} \int _{\scriptscriptstyle D}\int _{\scriptscriptstyle D}\left[ \int _{\scriptscriptstyle T_n}{\mathcal {X}}_{(0,t)}(s') \left( \,\sum _{i={\scriptscriptstyle \mathsf{M}}+1}^{\infty } \lambda _ie^{-\lambda _i^2(\lambda _i^2-\mu )(t-s')} \varepsilon _i(x)\,\varphi _i(y)\right) \mathrm{d}s' \right] ^2\mathrm{d}x\mathrm{d}y\\= & {} \tfrac{1}{\Delta {t}}\, \sum _{n=1}^{\scriptscriptstyle \mathsf{N}} \int _{\scriptscriptstyle D}\int _{\scriptscriptstyle D}\left[ \,\sum _{i={\scriptscriptstyle \mathsf{M}}+1}^{\infty } \left( \int _{\scriptscriptstyle T_n}{\mathcal {X}}_{(0,t)}(s') \,\lambda _ie^{-\lambda _i^2(\lambda _i^2-\mu )(t-s')}\;\mathrm{d}s' \right) \varepsilon _i(x)\,\varphi _i(y) \right] ^2\mathrm{d}x\mathrm{d}y\\= & {} \,\tfrac{1}{\Delta {t}}\, \sum _{n=1}^{\scriptscriptstyle \mathsf{N}} \,\sum _{i={\scriptscriptstyle \mathsf{M}}+1}^{\infty } \left( \int _{\scriptscriptstyle T_n}{\mathcal {X}}_{(0,t)}(s')\,\lambda _i\,e^{-\lambda _i^2(\lambda _i^2-\mu )(t-s')}\;\mathrm{d}s' \right) ^2 \quad \forall \,t\in (0,T]. \end{aligned}$$

For \(\mathsf{M}\ge \kappa \), using the Cauchy–Schwarz inequality, we obtain

$$\begin{aligned} \sqrt{{\mathsf{\Theta }}_{\scriptscriptstyle B}(t)}&\le \left[ \,\sum _{i={\scriptscriptstyle \mathsf{M}}+1}^{\infty } \lambda _i^2\,\left( \,\int _0^t\,e^{-2\,\lambda _i^2(\lambda _i^2-\mu )(t-s')}\;\mathrm{d}s'\,\right) \right] ^{\frac{1}{2}}\nonumber \\&\le \tfrac{1}{\sqrt{2}}\,\left( \,\sum _{i={\scriptscriptstyle \mathsf{M}}+1}^{\infty } \tfrac{1}{\lambda _i^2-\mu }\,\right) ^{\frac{1}{2}}\nonumber \\&\le \tfrac{\kappa +1}{\sqrt{2+4\kappa }} \,\left( \,\sum _{i={\scriptscriptstyle \mathsf{M}}+1}^{\infty } \tfrac{1}{\lambda _i^2}\,\right) ^{\frac{1}{2}}\nonumber \\&\le \tfrac{\kappa +1}{\pi \,\sqrt{2+4\kappa }} \,\left( \,\int _{\scriptscriptstyle \mathsf{M}}^{\infty }\tfrac{1}{x^2}\;\mathrm{d}x\,\right) ^{\frac{1}{2}}\nonumber \\&\le \tfrac{\kappa +1}{\pi \,\sqrt{2+4\kappa }}\,\mathsf{M}^{-\frac{1}{2}}\quad \forall \,t\in (0,T]. \end{aligned}$$
(3.6)

The error bound (3.1) follows by observing that \(\Theta (0)=0\) and by combining the bounds (3.4), (3.5) and (3.6). \(\square \)

4 Deterministic time-discrete and fully discrete approximations

In this section, we define and analyze auxiliary time-discrete and fully discrete approximations of the solution to the deterministic problem (1.4). The results of the convergence analysis will be used in Sect. 5 for the derivation of an error estimate for the numerical approximations of \(\mathsf{u}\) introduced in Sect. 1.2.

4.1 Time-discrete approximations

We define an auxiliary modified-IMEX time-discrete method to approximate the solution w to (1.4), which has the following structure: First, sets

$$\begin{aligned} W^0:=w_0 \end{aligned}$$
(4.1)

and determines \(W^1\in {\dot{\mathbf{H}}}^4(D)\) by

$$\begin{aligned} W^1-W^{0} +\Delta {\tau }\,\partial _x^4W^1=0. \end{aligned}$$
(4.2)

Then, for \(m=2,\dots ,M\), finds \(W^m\in {\dot{\mathbf{H}}}^4(D)\), such that

$$\begin{aligned} W^m-W^{m-1}+\Delta \tau \,\left( \,\partial _x^4W^m +\mu \,\partial _x^2W^{m-1}\,\right) =0. \end{aligned}$$
(4.3)

In the proposition below, we derive a low regularity priori error estimate in a discrete in time \(L^2_t(L^2_x)\)-norm.

Proposition 4.1

Let \((W^m)_{m=0}^{\scriptscriptstyle M}\) be the time-discrete approximations defined in (4.1)–(4.3), and w be the solution to the problem (1.4). Then, there exists a constant \(C>0\), independent of \(\Delta \tau \), such that

$$\begin{aligned} \left( \,\Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M} \Vert W^m-w^m\Vert _{\scriptscriptstyle 0,D}^2 \,\right) ^{\frac{1}{2}} \le \,C\,\Delta \tau ^{\theta } \,\Vert w_0\Vert _{\scriptscriptstyle {\dot{\mathbf{H}}}^{4\theta -2}} \quad \forall \,\theta \in [0,1], \quad \forall \,w_0\in {\dot{\mathbf{H}}}^2(D), \end{aligned}$$
(4.4)

where \(w^{\ell }(\cdot ):=w(\tau _{\ell },\cdot )\) for \(\ell =0,\dots ,M\).

Proof

In the sequel, we will use the symbol C to denote a generic constant that is independent of \(\Delta \tau \) and may changes value from one line to the other.

Let \(\mathsf{E}^m:=w^m-W^m\) for \(m=0,\dots ,M\), and

$$\begin{aligned} \sigma _{m}(\cdot ):=\int _{\scriptscriptstyle \Delta _m} \left( \,w(\tau _m,\cdot )-w(\tau ,\cdot )\,\right) \,\mathrm{d}\tau \, +\mu \,\int _{\scriptscriptstyle \Delta _m} T_{\scriptscriptstyle E}\left( \,w(\tau _{m-1},\cdot )-w(\tau ,\cdot )\,\right) \,\mathrm{d}\tau , \end{aligned}$$

for \(m=1,\dots ,M\). Thus, combining (1.4), (4.2) and (4.3), we conclude that

$$\begin{aligned}&T_{\scriptscriptstyle B}(\mathsf{E}^1-\mathsf{E}^0)+\Delta \tau \,\mathsf{E}^1=\sigma _1-\Delta \tau \,\mu \,T_{\scriptscriptstyle E}w_0, \end{aligned}$$
(4.5)
$$\begin{aligned}&T_{\scriptscriptstyle B}(\mathsf{E}^m-\mathsf{E}^{m-1})+\Delta \tau \,\left( \,\mathsf{E}^m+\mu \,T_{\scriptscriptstyle E}\mathsf{E}^{m-1}\,\right) =\sigma _m,\quad m=2,\dots ,M. \end{aligned}$$
(4.6)

First, take the \(L^2(D)\)-inner product of both sides of (4.5) with \(\mathsf{E}^1\) and of (4.6) with \(\mathsf{E}^m\), and then use (2.13) to obtain

$$\begin{aligned}&(T_{\scriptscriptstyle E}\mathsf{E}^1-T_{\scriptscriptstyle E}\mathsf{E}^0,T_{\scriptscriptstyle E}\mathsf{E}^1)_{\scriptscriptstyle 0,D} +\Delta \tau \,\Vert \mathsf{E}^1\Vert _{\scriptscriptstyle 0,D}^2=(\sigma _1,\mathsf{E}^1)_{\scriptscriptstyle 0,D} -\Delta \tau \,\mu \,(T_{\scriptscriptstyle E}w_0,\mathsf{E}^1)_{\scriptscriptstyle 0,D},\\&\quad (T_{\scriptscriptstyle E}\mathsf{E}^m-T_{\scriptscriptstyle E}\mathsf{E}^{m-1},T_{\scriptscriptstyle E}\mathsf{E}^m)_{\scriptscriptstyle 0,D} +\Delta \tau \,\Vert \mathsf{E}^m\Vert _{\scriptscriptstyle 0,D}^2= -\mu \,\Delta \tau \,(T_{\scriptscriptstyle E}\mathsf{E}^{m-1},\mathsf{E}^m)_{\scriptscriptstyle 0,D} +(\sigma _m,\mathsf{E}^m)_{\scriptscriptstyle 0,D} \end{aligned}$$

for \(m=2,\dots ,M\). Then, using that \(\mathsf{E}^0=0\) and applying (2.5) along with the arithmetic mean inequality, we get

$$\begin{aligned}&\Vert T_{\scriptscriptstyle E}\mathsf{E}^1\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau \,\Vert \mathsf{E}^1\Vert _{\scriptscriptstyle 0,D}^2 \le \Delta \tau ^{-1}\,\Vert \sigma _1\Vert _{\scriptscriptstyle 0,D}^2 -2\,\Delta \tau \,\mu \,(T_{\scriptscriptstyle E}w_0,\mathsf{E}^1)_{\scriptscriptstyle 0,D}, \end{aligned}$$
(4.7)
$$\begin{aligned}&\Vert T_{\scriptscriptstyle E}\mathsf{E}^m\Vert _{\scriptscriptstyle 0,D}^2 +\tfrac{1}{2}\,\Delta \tau \,\Vert \mathsf{E}^m\Vert _{\scriptscriptstyle 0,D}^2\le (1+2\,\mu ^2\,\Delta \tau )\,\Vert T_{\scriptscriptstyle E}\mathsf{E}^{m-1}\Vert _{\scriptscriptstyle 0,D}^2 \nonumber \\&\qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad +\Delta \tau ^{-1}\,\Vert \sigma _m\Vert _{\scriptscriptstyle 0,D}^2,\quad m=2,\dots ,M. \end{aligned}$$
(4.8)

Observing that (4.8) yields

$$\begin{aligned} \Vert T_{\scriptscriptstyle E}\mathsf{E}^m\Vert _{\scriptscriptstyle 0,D}^2\le \,(1+2\,\mu ^2\,\Delta \tau ) \,\Vert T_{\scriptscriptstyle E}\mathsf{E}^{m-1}\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau ^{-1}\,\Vert \sigma _m\Vert _{\scriptscriptstyle 0,D}^2,\quad m=2,\dots ,M, \end{aligned}$$

we use a standard discrete Gronwall argument to arrive at

$$\begin{aligned} \max _{1\le {m}\le {\scriptscriptstyle M}}\Vert T_{\scriptscriptstyle E}\mathsf{E}^m\Vert _{\scriptscriptstyle 0,D}^2 \le \,C\,\left( \,\Vert T_{\scriptscriptstyle E}\mathsf{E}^1\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau ^{-1}\,\sum _{m=2}^{\scriptscriptstyle M}\,\Vert \sigma _m\Vert _{\scriptscriptstyle 0,D}^2\,\right) . \end{aligned}$$
(4.9)

Summing both sides of (4.8) with respect to m, from 2 up to M, we obtain

$$\begin{aligned} \Vert T_{\scriptscriptstyle E}\mathsf{E}^{\scriptscriptstyle M}\Vert _{\scriptscriptstyle 0,D}^2 +\tfrac{\Delta \tau }{2}\,\sum _{m=2}^{\scriptscriptstyle M}\Vert \mathsf{E}^m\Vert _{\scriptscriptstyle 0,D}^2\le \Vert T_{\scriptscriptstyle E}\mathsf{E}^1\Vert _{\scriptscriptstyle 0,D}^2 +2\,\mu ^2\,\Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M-1}\Vert T_{\scriptscriptstyle E}\mathsf{E}^{m}\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau ^{-1}\sum _{m=2}^{\scriptscriptstyle M}\Vert \sigma _m\Vert _{\scriptscriptstyle 0,D}^2, \end{aligned}$$

which, along with (4.9), yields

$$\begin{aligned} \Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M}\Vert \mathsf{E}^m\Vert _{\scriptscriptstyle 0,D}^2\le \,C\,\left( \, \Vert T_{\scriptscriptstyle E}\mathsf{E}^1\Vert _{\scriptscriptstyle 0,D}^2+\Delta \tau \,\Vert \mathsf{E}^1\Vert ^2_{\scriptscriptstyle 0,D} +\Delta \tau ^{-1}\,\sum _{m=2}^{\scriptscriptstyle M}\Vert \sigma _m\Vert _{\scriptscriptstyle 0,D}^2\,\right) . \end{aligned}$$
(4.10)

Using (4.7), (2.8), the Cauchy–Schwarz inequality and the arithmetic mean inequality, we have

$$\begin{aligned} \begin{aligned} \Vert T_{\scriptscriptstyle E}\mathsf{E}^1\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau \,\Vert \mathsf{E}^1\Vert _{\scriptscriptstyle 0,D}^2\le&\,\Delta \tau ^{-1}\,\Vert \sigma _1\Vert _{\scriptscriptstyle 0,D}^2 -2\,\Delta \tau \,\mu \,(w_0,T_{\scriptscriptstyle E}\mathsf{E}^1)_{\scriptscriptstyle 0,D}\\ \le&\,\Delta \tau ^{-1}\,\Vert \sigma _1\Vert _{\scriptscriptstyle 0,D}^2 +2\,\Delta \tau \,|\mu |\,\Vert w_0\Vert _{\scriptscriptstyle 0,D}\,\Vert T_{\scriptscriptstyle E}\mathsf{E}^1\Vert _{\scriptscriptstyle 0,D}\\ \le&\,\Delta \tau ^{-1}\,\Vert \sigma _1\Vert _{\scriptscriptstyle 0,D}^2 +\tfrac{1}{2}\,\Vert T_{\scriptscriptstyle E}\mathsf{E}^1\Vert _{\scriptscriptstyle 0,D}^2 +2\,\Delta \tau ^2\,\mu ^2\,\Vert w_0\Vert ^2_{\scriptscriptstyle 0,D}\\ \end{aligned} \end{aligned}$$

which, finally, yields

$$\begin{aligned} \Vert T_{\scriptscriptstyle E}\mathsf{E}^1\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau \,\Vert \mathsf{E}^1\Vert _{\scriptscriptstyle 0,D}^2\le \,C\,\left( \,\Delta \tau ^2\,\Vert w_0\Vert ^2_{\scriptscriptstyle 0,D} +\Delta \tau ^{-1}\,\Vert \sigma _1\Vert _{\scriptscriptstyle 0,D}^2\,\right) . \end{aligned}$$
(4.11)

Next, we use the Cauchy–Schwarz inequality and (2.9) to get

$$\begin{aligned} \Vert \sigma _m\Vert _{\scriptscriptstyle 0,D}^2&\le 2\,\Delta \tau ^3\, \int _{\scriptscriptstyle \Delta _m}\Vert \partial _{\tau }w(s,\cdot )\Vert _{\scriptscriptstyle 0,D}^2\;\mathrm{d}s +2\,\mu ^2\,\Delta \tau ^3\, \int _{\scriptscriptstyle \Delta _m}\Vert T_{\scriptscriptstyle E}(\partial _{\tau }w(s,\cdot ))\Vert _{\scriptscriptstyle 0,D}^2\;\mathrm{d}s\nonumber \\&\le C\,(\Delta \tau )^{3}\,\int _{\scriptscriptstyle \Delta _m} \Vert \partial _{\tau }w(s,\cdot )\Vert _{\scriptscriptstyle 0,D}^2\,\mathrm{d}s, \quad m=1,\dots ,M. \end{aligned}$$
(4.12)

Finally, we use (4.10), (4.11), (4.12), and (2.14) (with \(\beta =0\), \(\ell =1\), and \(p=0\)) to obtain

$$\begin{aligned} \Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M}\Vert \mathsf{E}^m\Vert _{\scriptscriptstyle 0,D}^2&\le \,C\,\left( \, \Delta \tau ^2\,\Vert w_0\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau ^{-1}\,\sum _{m=1}^{\scriptscriptstyle M}\Vert \sigma _m\Vert _{\scriptscriptstyle 0,D}^2\,\right) \\&\le \,C\,\left( \, \Delta \tau ^2\,\Vert w_0\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau ^2\,\int _0^{\scriptscriptstyle T}\Vert \partial _{\tau }w(s,\cdot )\Vert _{\scriptscriptstyle 0,D}^2\,\mathrm{d}s\,\right) \\&\le \,C\,\Delta \tau ^2\,\Vert w_0\Vert ^2_{\scriptscriptstyle {\dot{\mathbf{H}}}^2},\\ \end{aligned}$$

which establishes (4.4) for \(\theta =1\).

From (4.2), (4.3), and (2.12), it follows that:

$$\begin{aligned}&T_{\scriptscriptstyle B}(W^1-W^0)+\Delta \tau \,W^1=0,\\&\quad T_{\scriptscriptstyle B}(W^m-W^{m-1})+\Delta \tau \,\left( \,W^m+\mu \,T_{\scriptscriptstyle E}W^{m-1}\,\right) =0, \quad m=2,\dots ,M. \end{aligned}$$

Taking the \(L^2(D)\)-inner product of both sides of the first equation above with \(W^1\) and of the second one with \(W^m\), and then applying (2.13), (2.5) and the arithmetic mean inequality, we obtain

$$\begin{aligned}&\Vert T_{\scriptscriptstyle E}W^1\Vert _{\scriptscriptstyle 0,D}^2 -\Vert T_{\scriptscriptstyle E}W^0\Vert _{\scriptscriptstyle 0,D}^2+2\,\Delta \tau \,\Vert W^1\Vert _{\scriptscriptstyle 0,D}^2\le \,0, \end{aligned}$$
(4.13)
$$\begin{aligned}&\Vert T_{\scriptscriptstyle E}W^m\Vert _{\scriptscriptstyle 0,D}^2 -\Vert T_{\scriptscriptstyle E}W^{m-1}\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau \,\Vert W^m\Vert _{\scriptscriptstyle 0,D}^2\le \,\mu ^2\,\Delta \tau \,\Vert T_{\scriptscriptstyle E}W^{m-1}\Vert _{\scriptscriptstyle 0,D}^2,\quad m=2,\dots ,M.\nonumber \\ \end{aligned}$$
(4.14)

The inequalities (4.13) and (4.14), easily, yield that

$$\begin{aligned} \Vert T_{\scriptscriptstyle E}W^m\Vert _{\scriptscriptstyle 0,D}^2\le \,(1+\mu ^2\,\Delta \tau )\,\Vert T_{\scriptscriptstyle E}W^{m-1}\Vert ^2_{\scriptscriptstyle 0,D},\quad m=1,\dots ,M, \end{aligned}$$

from which, after the use of a standard discrete Gronwall argument, we arrive at

$$\begin{aligned} \max _{0\le {m}\le {\scriptscriptstyle M}}\Vert T_{\scriptscriptstyle E}W^m\Vert _{\scriptscriptstyle 0,D}^2 \le \,C\,\Vert T_{\scriptscriptstyle E}W^0\Vert _{\scriptscriptstyle 0,D}^2. \end{aligned}$$
(4.15)

We sum both sides of (4.14) with respect to m, from 2 up to M, and then use (4.15), to have

$$\begin{aligned} \Delta \tau \,\sum _{m=2}^{\scriptscriptstyle M}\Vert W^m\Vert _{\scriptscriptstyle 0,D}^2&\le \Vert T_{\scriptscriptstyle E}W^1\Vert _{\scriptscriptstyle 0,D}^2 +\mu ^2\,\Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M-1}\Vert T_{\scriptscriptstyle E}W^m\Vert _{\scriptscriptstyle 0,D}^2\nonumber \\&\le C\,\left( \,\Vert T_{\scriptscriptstyle E}W^1\Vert _{\scriptscriptstyle 0,D}^2 +\Vert T_{\scriptscriptstyle E}W^0\Vert _{\scriptscriptstyle 0,D}^2\,\right) . \end{aligned}$$
(4.16)

Thus, using (4.16), (4.13), (4.1), (2.9), and (2.2), we obtain

$$\begin{aligned} \Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M}\Vert W^m\Vert _{\scriptscriptstyle 0,D}^2&\le C\,\left( \,\Vert T_{\scriptscriptstyle E}W^1\Vert _{\scriptscriptstyle 0,D}^2+ \Delta \tau \,\Vert W^1\Vert _{\scriptscriptstyle 0,D}^2+\Vert T_{\scriptscriptstyle E}w_0\Vert _{\scriptscriptstyle 0,D}^2\,\right) \nonumber \\&\le C\,\Vert T_{\scriptscriptstyle E}w_0\Vert _{\scriptscriptstyle 0, D}^2\nonumber \\&\le C\,\Vert w_0\Vert _{\scriptscriptstyle -2, D}^2\nonumber \\&\le C\,\Vert w_0\Vert _{\scriptscriptstyle {\dot{\mathbf{H}}}^{-2}}^2. \end{aligned}$$
(4.17)

In addition, we have

$$\begin{aligned} \begin{aligned} \Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M}\Vert w^m\Vert _{\scriptscriptstyle 0,D}^2 =&\,\sum _{m=1}^{\scriptscriptstyle M}\,\int _{\scriptscriptstyle D}\, \left( \,\int _{\scriptscriptstyle \Delta _m}\partial _{\tau }\left[ \,(\tau -\tau _{m-1}) \,w^2(\tau ,x)\,\right] \,\mathrm{d}\tau \,\right) \,\mathrm{d}x\\ =&\,\sum _{m=1}^{\scriptscriptstyle M}\,\int _{\scriptscriptstyle D}\, \left( \int _{\scriptscriptstyle \Delta _m}\left[ \,w^2(\tau ,x) +2\,(\tau -\tau _{m-1})\,w_{\tau }(\tau ,x)\,w(\tau ,x)\,\right] \,\mathrm{d}\tau \right) \,\mathrm{d}x\\ \le&\,\sum _{m=1}^{\scriptscriptstyle M}\int _{\scriptscriptstyle \Delta _m}\, \left( \,2\,\Vert w(\tau ,\cdot )\Vert _{\scriptscriptstyle 0,D}^2 +(\tau -\tau _{m-1})^2\, \Vert w_{\tau }(\tau ,\cdot )\Vert _{\scriptscriptstyle 0,D}^2\,\right) \;\mathrm{d}\tau \\ \le&\,2\,\int _0^{\scriptscriptstyle T}\Vert w(\tau ,\cdot )\Vert _{\scriptscriptstyle 0,D}^2\;\mathrm{d}\tau +\int _0^{\scriptscriptstyle T}\tau ^2\,\Vert w_{\tau }(\tau ,\cdot )\Vert _{\scriptscriptstyle 0,D}^2\,\mathrm{d}\tau , \end{aligned} \end{aligned}$$

which, along with (2.14) (with \((\beta ,\ell ,p)=(0,0,0)\) and \((\beta ,\ell ,p)=(2,1,0)\)), yields

$$\begin{aligned} \Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M}\Vert w^m\Vert _{\scriptscriptstyle 0,D}^2 \le \,C\,\Vert w_0\Vert ^2_{\scriptscriptstyle {\dot{\mathbf{H}}}^{-2}}. \end{aligned}$$
(4.18)

Thus, (4.17) and (4.18) establish (4.4) for \(\theta =0\).

Finally, the estimate (4.4) follows by interpolation, since it is valid for \(\theta =1\) and \(\theta =0\). \(\square \)

We close this section by deriving, for later use, the following a priori bound.

Lemma 4.1

Let \((W^m)_{m=0}^{\scriptscriptstyle M}\) be the time-discrete approximations defined by (4.1)–(4.3). Then, there exist a constant \(C>0\), independent of \(\Delta \tau \), such that

$$\begin{aligned} \left( \,\Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M} \Vert \partial _x^3W^m\Vert _{\scriptscriptstyle 0,D}^2\,\right) ^{\frac{1}{2}} \le \,C\,\Vert w_0\Vert _{\scriptscriptstyle {\dot{\mathbf{H}}}^{1}} \quad \forall \,w_0\in {\dot{\mathbf{H}}}^1(D). \end{aligned}$$
(4.19)

Proof

In the sequel, we will use the symbol C to denote a generic constant that is independent of \(\Delta \tau \) and may changes value from one line to the other.

Taking the \((\cdot ,\cdot )_{\scriptscriptstyle 0,D}\)-inner product of (4.3) with \(\partial _x^2W^m\) and of (4.2) with \(\partial _x^2W^1\), and then integrating by parts, we obtain

$$\begin{aligned}&\big (\partial _xW^1-\partial _xW^0, \partial _xW^1\big )_{\scriptscriptstyle 0,D} +\Delta \tau \,\Vert \partial _x^3W^1\Vert _{\scriptscriptstyle 0,D}^2=0, \end{aligned}$$
(4.20)
$$\begin{aligned}&\big (\partial _xW^m-\partial _xW^{m-1}, \partial _xW^m\big )_{\scriptscriptstyle 0,D} +\Delta \tau \,\left[ \,\Vert \partial _x^3W^m\Vert _{\scriptscriptstyle 0,D}^2 +\mu \,(\partial _x^3W^{m},\partial _xW^{m-1})_{\scriptscriptstyle 0,D}\,\right] =0\nonumber \\ \end{aligned}$$
(4.21)

for \(m=2,\dots ,M\). Using (2.5) and the arithmetic mean inequality, from (4.20) and (4.21), it follows that:

$$\begin{aligned}&\Vert \partial _xW^1\Vert ^2_{\scriptscriptstyle 0,D}-\Vert \partial _xW^0\Vert _{\scriptscriptstyle 0,D}^2+2\,\Delta \tau \,\Vert \partial _x^3W^1\Vert _{\scriptscriptstyle 0,D}^2\le 0, \end{aligned}$$
(4.22)
$$\begin{aligned}&\Vert \partial _xW^m\Vert ^2_{\scriptscriptstyle 0,D}-\Vert \partial _xW^{m-1}\Vert _{\scriptscriptstyle 0,D}^2+\Delta \tau \,\Vert \partial _x^3W^m\Vert _{\scriptscriptstyle 0,D}^2\le \,\Delta \tau \,\mu ^2\,\Vert \partial _xW^{m-1}\Vert _{\scriptscriptstyle 0,D}^2,\quad m=2,\dots ,M.\nonumber \\ \end{aligned}$$
(4.23)

Now, (4.23) and (4.22), easily, yield that

$$\begin{aligned} \Vert \partial _xW^m\Vert _{\scriptscriptstyle 0,D}^2\le \,(1+\mu ^2\,\Delta \tau ) \,\Vert \partial _xW^{m-1}\Vert _{\scriptscriptstyle 0,D}^2,\quad m=2,\dots ,M, \end{aligned}$$

which, after a standard induction argument, leads to

$$\begin{aligned} \max _{1\le {m}\le {\scriptscriptstyle M}}\Vert \partial _xW^m\Vert _{\scriptscriptstyle 0,D}^2\le \,C\,\Vert \partial _xW^1\Vert _{\scriptscriptstyle 1,D}^2. \end{aligned}$$
(4.24)

After summing both sides of (4.23) with respect to m, from 2 up to M, we obtain

$$\begin{aligned} \Delta \tau \,\sum _{m=2}^{\scriptscriptstyle M}\Vert \partial _x^3W^m\Vert _{\scriptscriptstyle 0,D}^2\le \Vert \partial _xW^1\Vert _{\scriptscriptstyle 0,D}^2+\mu ^2\,\Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M-1} \Vert \partial _xW^{m}\Vert _{\scriptscriptstyle 0,D}^2 \end{aligned}$$

which, after using (4.24), yields

$$\begin{aligned} \Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M}\Vert \partial _x^3W^m\Vert _{\scriptscriptstyle 0,D}^2\le \,C\,\left( \, \Vert \partial _xW^1\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau \,\Vert \partial _x^3W^1\Vert _{\scriptscriptstyle 0,D}^2\,\right) . \end{aligned}$$
(4.25)

Finally, we combine (4.25), (4.22), and (2.1) to get

$$\begin{aligned} \begin{aligned} \Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M} \Vert \partial _x^3W^m\Vert _{\scriptscriptstyle 0,D}^2\le&\,C\,\Vert \partial _xW^0\Vert _{\scriptscriptstyle 0,D}^2\\ \le&\,C\,\Vert w_0\Vert _{\scriptscriptstyle 1,D}^2\\ \le&\,C\,\Vert w_0\Vert _{\scriptscriptstyle {\dot{\mathbf{H}}}^1}^2,\\ \end{aligned} \end{aligned}$$

which, easily, yields (4.19). \(\square \)

4.2 Fully discrete approximations

The modified-IMEX time-stepping method along with a finite element space discretization yields a fully discrete method for the approximation of the solution to the deterministic problem (1.4). The method begins by setting

$$\begin{aligned} W_h^0:=P_hw_0 \end{aligned}$$
(4.26)

and specifying \(W_h^1\in \mathsf{M}_h^r\), such that

$$\begin{aligned} W_h^1-W_h^0+\Delta \tau \,B_hW_h^1=0. \end{aligned}$$
(4.27)

Then, for \(m=2,\dots ,M\), it finds \(W_h^m\in \mathsf{M}_h^r\), such that

$$\begin{aligned} W_h^m-W_h^{m-1}+\Delta \tau \,\left[ \,B_hW_h^m+\mu \,P_h\left( \, \partial _x^2W_h^{m-1}\,\right) \,\right] =0. \end{aligned}$$
(4.28)

Adopting the viewpoint that the fully discrete approximations defined above are approximations of the time-discrete ones defined in the previous section, we estimate below the corresponding approximation error in a discrete in time \(L^2_t(L^2_x)\)-norm.

Proposition 4.2

Let \(r=2\) or 3, \((W^m)_{m=0}^{\scriptscriptstyle M}\) be the time-discrete approximations defined by (4.1)–(4.3), and \((W_h^m)_{m=0}^{\scriptscriptstyle M}\subset \mathsf{M}_h^r\) be the fully discrete approximations specified in (4.26)–(4.28). Then, there exists a constant \(C>0\), independent of \(\Delta \tau \) and h, such that

$$\begin{aligned} \left( \,\Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M}\Vert W^m-W_h^m\Vert ^2_{\scriptscriptstyle 0,D}\,\right) ^{\frac{1}{2}} \le \,C\,h^{r \theta }\,\Vert w_0\Vert _{\scriptscriptstyle {\dot{\mathbf{H}}}^{3\theta -2}} \quad \forall \,w_0\in {\dot{\mathbf{H}}}^1(D),\quad \forall \,\theta \in [0,1]. \end{aligned}$$
(4.29)

Proof

In the sequel, we will use the symbol C to denote a generic constant which is independent of \(\Delta \tau \) and h, and may changes value from one line to the other.

Let \(\mathsf{Z}^m:=W^m-W_h^m\) for \(m=0,\dots ,M\). Then, from (4.2), (4.3), (4.27), and (4.28), we obtain the following error equations:

$$\begin{aligned}&T_{\scriptscriptstyle B,h}(\mathsf{Z}^1-\mathsf{Z}^0)+\Delta \tau \,\mathsf{Z}^1=\Delta \tau \,\xi ^1, \end{aligned}$$
(4.30)
$$\begin{aligned}&T_{\scriptscriptstyle B,h}(\mathsf{Z}^m-\mathsf{Z}^{m-1})+\Delta \tau \,\left[ \,\mathsf{Z}^m +\mu \,T_{\scriptscriptstyle B,h}(\partial _x^2\mathsf{Z}^{m-1})\,\right] =\Delta \tau \,\xi ^m, \quad m=2,\dots ,M,\qquad \qquad \end{aligned}$$
(4.31)

where

$$\begin{aligned} \xi ^m:=(T_{\scriptscriptstyle B}-T_{{\scriptscriptstyle B},h})\partial _x^4W^m,\quad m=1,\dots ,M. \end{aligned}$$
(4.32)

Taking the \(L^2(D)\)-inner product of both sides of (4.31) with \(\mathsf{Z}^m\), we obtain

$$\begin{aligned} \begin{aligned} (T_{\scriptscriptstyle B,h}(\mathsf{Z}^m-\mathsf{Z}^{m-1}),\mathsf{Z}^m)_{\scriptscriptstyle 0,D} +\Delta \tau \,\Vert \mathsf{Z}^m\Vert ^2_{\scriptscriptstyle 0,D}=&\,-\mu \,\Delta \tau \,\left( T_{\scriptscriptstyle B,h}(\partial _x^2\mathsf{Z}^{m-1}),\mathsf{Z}^m\right) _{\scriptscriptstyle 0,D}\\&+\Delta \tau \,(\xi ^m,\mathsf{Z}^m)_{\scriptscriptstyle 0,D}, \quad m=2,\dots ,M,\\ \end{aligned} \end{aligned}$$

which, along with (2.17) and (2.5), yields

$$\begin{aligned}&\Vert \partial _x^2(T_{\scriptscriptstyle B,h}\mathsf{Z}^m)\Vert _{\scriptscriptstyle 0,D}^2 -\Vert \partial _x^2(T_{\scriptscriptstyle B,h}\mathsf{Z}^{m-1})\Vert _{\scriptscriptstyle 0,D}^2+\Vert \partial _x^2\left( T_{\scriptscriptstyle B,h}\left( \mathsf{Z}^m-\mathsf{Z}^{m-1}\right) \right) \Vert _{\scriptscriptstyle 0,D}^2\nonumber \\&\quad +2\,\Delta \tau \,\Vert \mathsf{Z}^m\Vert ^2_{\scriptscriptstyle 0,D} ={\mathcal {A}}_1^m+{\mathcal {A}}_2^m, \end{aligned}$$
(4.33)

for \(m=2,\dots ,M\), where

$$\begin{aligned} \begin{aligned} {\mathcal {A}}^m_1:=&\,2\,\Delta \tau \,(\xi ^m,\mathsf{Z}^m)_{\scriptscriptstyle 0,D},\\ {\mathcal {A}}_2^m:=&\,-2\,\mu \,\Delta \tau \,\left( T_{\scriptscriptstyle B,h}\left( \partial _x^2\mathsf{Z}^{m-1}\right) ,\mathsf{Z}^m\right) _{\scriptscriptstyle 0,D}.\\ \end{aligned} \end{aligned}$$

Using (2.17), integration by parts, the Cauchy–Schwarz inequality, the arithmetic mean inequality, we have

$$\begin{aligned} {\mathcal {A}}_1^m\le \Delta \tau \,\left( \,\Vert \mathsf{Z}^m\Vert _{\scriptscriptstyle 0,D}^2+\Vert \xi ^m\Vert ^2_{\scriptscriptstyle 0,D}\,\right) \end{aligned}$$
(4.34)

and

$$\begin{aligned} {\mathcal {A}}_2^m&= -2\,\mu \,\Delta \tau \,(\partial _x^2\mathsf{Z}^{m-1},T_{\scriptscriptstyle B,h}\mathsf{Z}^m)_{\scriptscriptstyle 0,D}\nonumber \\&=-2\,\mu \,\Delta \tau \,(\mathsf{Z}^{m-1},\partial _x^2(T_{\scriptscriptstyle B,h}\mathsf{Z}^m))_{\scriptscriptstyle 0,D}\nonumber \\&=-2\,\mu \,\Delta \tau \,\left( \mathsf{Z}^{m-1},\partial _x^2 \left( T_{\scriptscriptstyle B,h}\left( \mathsf{Z}^m-\mathsf{Z}^{m-1}\right) \right) \right) _{\scriptscriptstyle 0,D}\nonumber \\&\quad -2\,\mu \,\Delta \tau \,(\mathsf{Z}^{m-1}, \partial _x^2(T_{\scriptscriptstyle B,h}\mathsf{Z}^{m-1}))_{\scriptscriptstyle 0,D}\nonumber \\&\le 2\,|\mu |\,\Delta \tau \,\Vert \mathsf{Z}^{m-1}\Vert _{\scriptscriptstyle 0,D}\, \left\| \partial _x^2\left( T_{\scriptscriptstyle B,h}\left( \mathsf{Z}^m-\mathsf{Z}^{m-1}\right) \right) \right\| _{\scriptscriptstyle 0,D}\nonumber \\&\quad +2\,|\mu |\,\Delta \tau \,\Vert \mathsf{Z}^{m-1}\Vert _{\scriptscriptstyle 0,D}\, \left\| \partial _x^2\left( T_{\scriptscriptstyle B,h}\mathsf{Z}^{m-1}\right) \right\| _{\scriptscriptstyle 0,D}\nonumber \\&\le \Delta \tau ^2\,\mu ^2\,\Vert \mathsf{Z}^{m-1}\Vert ^2_{\scriptscriptstyle 0,D} +\Vert \partial _x^2\left( T_{\scriptscriptstyle B,h}\left( \mathsf{Z}^m-\mathsf{Z}^{m-1}\right) \right) \Vert _{\scriptscriptstyle 0,D}^2\nonumber \\&\quad +\tfrac{\Delta \tau }{2}\,\Vert \mathsf{Z}^{m-1}\Vert ^2_{\scriptscriptstyle 0,D} +2\,\Delta \tau \,\mu ^2\,\Vert \partial _x^2(T_{\scriptscriptstyle B,h}\mathsf{Z}^{m-1})\Vert _{\scriptscriptstyle 0,D}^2,\quad m=2,\dots ,M. \end{aligned}$$
(4.35)

Now, we combine (4.33), (4.34) and (4.35) to get

$$\begin{aligned} \Vert \partial _x^2(T_{\scriptscriptstyle B,h}\mathsf{Z}^m)\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau \,\Vert \mathsf{Z}^m\Vert ^2_{\scriptscriptstyle 0,D}&\le \Vert \partial _x^2(T_{\scriptscriptstyle B,h}\mathsf{Z}^{m-1})\Vert _{\scriptscriptstyle 0,D}^2 +\tfrac{\Delta \tau }{2}\,\Vert \mathsf{Z}^{m-1}\Vert _{\scriptscriptstyle 0,D}^2+\Delta \tau \,\Vert \xi ^m\Vert _{\scriptscriptstyle 0,D}^2\nonumber \\&\qquad +2\,\Delta \tau \,\mu ^2\,\left( \,\Vert \partial _x^2(T_{\scriptscriptstyle B,h}\mathsf{Z}^{m-1})\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau \,\Vert \mathsf{Z}^{m-1}\Vert _{\scriptscriptstyle 0,D}^2\,\right) \end{aligned}$$
(4.36)

for \(m=2,\dots ,M\). Let \(\Upsilon ^{\ell }:=\Vert \partial _x^2(T_{\scriptscriptstyle B,h}\mathsf{Z}^{\ell })\Vert _{\scriptscriptstyle 0,D}^2+\Delta \tau \,\Vert \mathsf{Z}^{\ell }\Vert _{\scriptscriptstyle 0,D}^2\) for \(\ell =1,\dots ,M\). Then, (4.36) yields

$$\begin{aligned} \Upsilon ^m\le (1+2\,\mu ^2\,\Delta \tau )\,\Upsilon ^{m-1} +\Delta \tau \,\Vert \xi ^m\Vert _{\scriptscriptstyle 0,D}^2,\quad m=2,\dots ,M, \end{aligned}$$

from which, after applying a standard discrete Gronwall argument, we conclude that

$$\begin{aligned} \max _{1\le {m}\le {\scriptscriptstyle M}} \Upsilon ^m\le \,C\left( \,\Upsilon ^1 +\Delta \tau \,\sum _{m=2}^{\scriptscriptstyle M}\Vert \xi ^m\Vert _{\scriptscriptstyle 0,D}^2\,\right) . \end{aligned}$$
(4.37)

Since \(T_{\scriptscriptstyle B,h}\mathsf{Z}^0=0\), after taking the \(L^2(D)\)-inner product of both sides of (4.30) with \(\mathsf{Z}^1\), and then, using (2.17) and the arithmetic mean inequality, we obtain

$$\begin{aligned} \Vert \partial _x^2(T_{\scriptscriptstyle B,h}\mathsf{Z}^1)\Vert _{\scriptscriptstyle 0,D}^2+\tfrac{\Delta \tau }{2}\,\Vert \mathsf{Z}^1\Vert _{\scriptscriptstyle 0,D}^2\le \,\tfrac{\Delta \tau }{2}\,\Vert \xi ^1\Vert ^2_{\scriptscriptstyle 0,D}, \end{aligned}$$
(4.38)

which, along with (4.37), yields

$$\begin{aligned} \max _{1\le {m}\le {\scriptscriptstyle M}}\Upsilon ^m\le \,C\,\Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M} \Vert \xi ^m\Vert _{\scriptscriptstyle 0,D}^2. \end{aligned}$$
(4.39)

Now, summing both sides of (4.36) with respect to m, from 2 up to M, we obtain

$$\begin{aligned} \begin{aligned} \Delta \tau \,\sum _{m=2}^{\scriptscriptstyle M}\Vert \mathsf{Z}^m\Vert _{\scriptscriptstyle 0,D}^2 \le&\,\Vert \partial _x^2(T_{\scriptscriptstyle B,h}\mathsf{Z}^1)\Vert _{\scriptscriptstyle 0,D}^2 +\tfrac{\Delta \tau }{2}\,\sum _{m=1}^{\scriptscriptstyle M-1}\Vert \mathsf{Z}^m\Vert _{\scriptscriptstyle 0,D}^2\\&+\Delta \tau \,\sum _{m=2}^{\scriptscriptstyle M}\Vert \xi ^m\Vert _{\scriptscriptstyle 0,D}^2 +2\,\mu ^2\,\Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M-1}\Upsilon ^{m},\\ \end{aligned} \end{aligned}$$

which, along with (4.39), yields

$$\begin{aligned} \tfrac{\Delta \tau }{2}\,\sum _{m=1}^{\scriptscriptstyle M}\Vert \mathsf{Z}^m\Vert _{\scriptscriptstyle 0,D}^2&\le \Vert \partial _x^2(T_{\scriptscriptstyle B,h}\mathsf{Z}^1)\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau \,\Vert \mathsf{Z}^1\Vert _{\scriptscriptstyle 0,D}^2\nonumber \\&\quad +\Delta \tau \,\sum _{m=2}^{\scriptscriptstyle M}\Vert \xi ^m\Vert _{\scriptscriptstyle 0,D}^2 +2\,\mu ^2\,\Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M-1}\Upsilon ^m\nonumber \\&\le C\,\left( \,\max _{1\le {m}\le {\scriptscriptstyle M}-1}\Upsilon ^m +\Delta \tau \,\sum _{m=2}^{\scriptscriptstyle M}\Vert \xi ^m\Vert _{\scriptscriptstyle 0,D}^2\,\right) \nonumber \\&\le C\,\Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M}\Vert \xi ^m\Vert _{\scriptscriptstyle 0,D}^2. \end{aligned}$$
(4.40)

Combining (4.40), (4.32), (2.19), and (4.19), we obtain

$$\begin{aligned} \Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M}\Vert \mathsf{Z}^m\Vert _{\scriptscriptstyle 0,D}^2&\le C\,h^{2 r}\,\Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M}\Vert \partial _x^3W^m\Vert _{\scriptscriptstyle 0,D}^2\nonumber \\&\le C\,h^{2 r}\,\Vert w_0\Vert ^2_{\scriptscriptstyle {\dot{\mathbf{H}}}^1}. \end{aligned}$$
(4.41)

Thus, (4.41) yields (4.29) for \(\theta =1\).

From (4.27) and (4.28), we conclude that

$$\begin{aligned}&T_{\scriptscriptstyle B,h}(W_h^1-W_h^0)+\Delta \tau \,W_h^1=0,\\&\quad T_{\scriptscriptstyle B,h}(W_h^m-W_h^{m-1})+\Delta \tau \,W_h^m =-\mu \,\Delta \tau \,T_{\scriptscriptstyle B,h}(\partial _x^2W_h^{m-1}), \quad m=2,\dots ,M. \end{aligned}$$

Taking the \(L^2(D)\)-inner product of both sides of the first equation above with \(W_h^1\) and of the second one with \(W_h^m\), and then, applying (2.17) and (2.5), we obtain

$$\begin{aligned}&\Vert \partial _x^2(T_{\scriptscriptstyle B,h}W_h^1)\Vert _{\scriptscriptstyle 0,D}^2 -\Vert \partial _x^2(T_{\scriptscriptstyle B,h}W_h^0)\Vert _{\scriptscriptstyle 0,D}^2 +2\,\Delta \tau \,\Vert W_h^1\Vert _{\scriptscriptstyle 0,D}^2\le \,0, \end{aligned}$$
(4.42)
$$\begin{aligned}&\Vert \partial _x^2(T_{\scriptscriptstyle B,h}W_h^m)\Vert _{\scriptscriptstyle 0,D}^2 \,+\Vert \partial _x^2(T_{\scriptscriptstyle B,h}(W_h^m-W_h^{m-1}))\Vert _{\scriptscriptstyle 0,D}^2\nonumber \\&\quad +2\Delta \tau \,\Vert W_h^m\Vert _{\scriptscriptstyle 0,D}^2=\Vert \partial _x^2(T_{\scriptscriptstyle B,h}W_h^{m-1})\Vert _{\scriptscriptstyle 0,D}^2 +{\mathcal {A}}_3^m,\quad m=2,\dots ,M, \end{aligned}$$
(4.43)

where

$$\begin{aligned} {\mathcal {A}}_3^m:=-2\,\mu \,\Delta \tau \,\left( T_{\scriptscriptstyle B,h} \left( \partial _x^2W_h^{m-1}\right) ,W_h^{m}\right) _{\scriptscriptstyle 0,D}. \end{aligned}$$

Using (2.17), integration by parts, the Cauchy–Schwarz inequality, and the arithmetic mean inequality, we have

$$\begin{aligned} {\mathcal {A}}_3^m&= -2\,\mu \,\Delta \tau \,\left( W_h^{m-1},\partial _x^2\left( T_{\scriptscriptstyle B,h}W_h^{m}\right) \right) _{\scriptscriptstyle 0,D}\nonumber \\&=-2\,\mu \,\Delta \tau \,\left( W_h^{m-1},\partial _x^2 \left( T_{\scriptscriptstyle B,h}\left( W_h^m-W_h^{m-1}\right) \right) \right) _{\scriptscriptstyle 0,D}\nonumber \\&\quad -2\,\mu \,\Delta \tau \,(W_h^{m-1}, \partial _x^2(T_{\scriptscriptstyle B,h}W_h^{m-1}))_{\scriptscriptstyle 0,D}\nonumber \\&\le \Delta \tau ^2\,\mu ^2\,\Vert W_h^{m-1}\Vert ^2_{\scriptscriptstyle 0,D} +\Vert \partial _x^2\left( T_{\scriptscriptstyle B,h}\left( W_h^m-W_h^{m-1}\right) \right) \Vert _{\scriptscriptstyle 0,D}^2\nonumber \\&\quad +\tfrac{\Delta \tau }{2}\,\Vert W_h^{m-1}\Vert ^2_{\scriptscriptstyle 0,D} +2\,\Delta \tau \,\mu ^2\,\Vert \partial _x^2(T_{\scriptscriptstyle B,h}W_h^{m-1})\Vert _{\scriptscriptstyle 0,D}^2, \quad m=2,\dots ,M. \end{aligned}$$
(4.44)

Combining (4.43) and (4.44), we arrive at

$$\begin{aligned}&\Vert \partial _x^2(T_{\scriptscriptstyle B,h}W_h^m)\Vert _{\scriptscriptstyle 0,D}^2\,+2\,\Delta \tau \,\Vert W^m_h\Vert _{\scriptscriptstyle 0,D}^2 \le \Vert \partial _x^2(T_{\scriptscriptstyle B,h}W_h^{m-1})\Vert _{\scriptscriptstyle 0,D}^2 +\tfrac{\Delta \tau }{2}\,\Vert W^{m-1}_h\Vert _{\scriptscriptstyle 0,D}^2\nonumber \\&\quad +2\,\Delta \tau \,\mu ^2\,\left( \,\Vert \partial _x^2(T_{\scriptscriptstyle B,h}W_h^{m-1})\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau \,\Vert W_h^{m-1}\Vert _{\scriptscriptstyle 0,D}^2\,\right) , \quad m=2,\dots ,M. \end{aligned}$$
(4.45)

Let\(\Upsilon _h^{\ell }:=\Vert \partial _x^2(T_{\scriptscriptstyle B,h}W_h^{\ell })\Vert _{\scriptscriptstyle 0,D}^2 +\Delta \tau \,\Vert W_h^{\ell }\Vert _{\scriptscriptstyle 0,D}^2\) for \(\ell =1,\dots ,M\). Then, we use (4.42), (4.26), (2.18), (2.2), and (4.45) to obtain

$$\begin{aligned} \Upsilon _h^1&\le \Vert \partial _x^2(T_{\scriptscriptstyle B,h}P_hw_0)\Vert _{\scriptscriptstyle 0,D}^2\nonumber \\&\le \Vert \partial _x^2(T_{\scriptscriptstyle B,h}w_0)\Vert _{\scriptscriptstyle 0,D}^2\nonumber \\&\le \Vert w_0\Vert _{\scriptscriptstyle -2,D}^2\nonumber \\&\le \Vert w_0\Vert _{\scriptscriptstyle {\dot{\mathbf{H}}}^{-2}}^2 \end{aligned}$$
(4.46)

and

$$\begin{aligned} \Upsilon _h^{m}\le (1+2\,\mu ^2\Delta \tau )\,\Upsilon _h^{m-1}, \quad m=2,\dots ,M. \end{aligned}$$
(4.47)

From (4.47), after the application of a standard discrete Gronwall argument and the use of (4.46), we conclude that

$$\begin{aligned} \max _{1\le {m}\le {\scriptscriptstyle M}} \Upsilon _h^m&\le C\,\Upsilon _h^1\nonumber \\&\le C\,\Vert w_0\Vert _{\scriptscriptstyle {\dot{\mathbf{H}}}^{-2}}^2. \end{aligned}$$
(4.48)

Summing both sides of (4.45) with respect to m, from 2 up to M, we have

$$\begin{aligned} \begin{aligned} \Delta \tau \,\sum _{m=2}^{\scriptscriptstyle M}\Vert W_h^m\Vert _{\scriptscriptstyle 0,D}^2 \le&\,\Vert \partial _x^2(T_{\scriptscriptstyle B,h}W_h^1)\Vert _{\scriptscriptstyle 0,D}^2 +\tfrac{\Delta \tau }{2}\,\sum _{m=1}^{\scriptscriptstyle M-1}\Vert W_h^m\Vert _{\scriptscriptstyle 0,D}^2 +2\,\mu ^2\,\Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M-1}\Upsilon _h^m, \end{aligned} \end{aligned}$$

which, along with (4.48), yields

$$\begin{aligned} \tfrac{\Delta \tau }{2}\,\sum _{m=1}^{\scriptscriptstyle M}\Vert W_h^m\Vert _{\scriptscriptstyle 0,D}^2&\le \Upsilon _h^1 +2\,\mu ^2\,\Delta \tau \,\sum _{m=1}^{\scriptscriptstyle M-1}\Upsilon _h^m\nonumber \\&\le C\,\Vert w_0\Vert _{\scriptscriptstyle {\dot{\mathbf{H}}}^{-2}}^2. \end{aligned}$$
(4.49)

Thus, (4.49) and (4.17) yield (4.29) for \(\theta =0\).

Thus, the error estimate (4.29) follows by interpolation, since it holds for \(\theta =1\) and \(\theta =0\). \(\square \)

5 Convergence analysis of the IMEX finite element method

To estimate the approximation error of the IMEX finite element method given in Sect. 1.2, we use, as a tool, the corresponding IMEX time-discrete approximations of \(\mathsf{u}\), which are defined first by setting

$$\begin{aligned} \mathsf{U}^0:=0 \end{aligned}$$
(5.1)

and then, for \(m=1,\dots ,M\), by seeking \(\mathsf{U}^m\in {\dot{\mathbf{H}}}^4(D)\), such that

$$\begin{aligned} \mathsf{U}^m-\mathsf{U}^{m-1}+\Delta \tau \,\left( \,\partial _x^4\mathsf{U}^m +\mu \,\partial _x^2\mathsf{U}^{m-1}\right) =\int _{\scriptscriptstyle \Delta _m}\partial _x{\mathcal {W}}\,\mathrm{d}\tau \quad \mathrm{a.s.}. \end{aligned}$$
(5.2)

Thus, we split the total error of the IMEX finite element method as follows:

$$\begin{aligned} \max _{0\le {m}\le {\scriptscriptstyle M}}\left( {\mathbb {E}}\left[ \, \Vert \mathsf{u}^m-\mathsf{U}_h^m\Vert _{\scriptscriptstyle 0,D}^2\,\right] \right) ^{\frac{1}{2}} \le \max _{0\le {m}\le {\scriptscriptstyle M}}{\mathcal {E}}_{{\scriptscriptstyle \mathrm TDR}}^m +\max _{0\le {m}\le {\scriptscriptstyle M}}{\mathcal {E}}^m_{{\scriptscriptstyle \mathrm SDR}}, \end{aligned}$$
(5.3)

where \(\mathsf{u}^m:=\mathsf{u}(\tau _m,\cdot )\), \({\mathcal {E}}^m_{{\scriptscriptstyle \mathrm TDR}}:= \left( {\mathbb {E}}\left[ \Vert \mathsf{u}^m-\mathsf{U}^m\Vert _{\scriptscriptstyle 0,D}^2 \right] \right) ^{\scriptscriptstyle 1/2}\) is the time-discretization error at \(\tau _m\), and \({\mathcal {E}}_{{\scriptscriptstyle \mathrm SDR}}^m:= \left( {\mathbb {E}}\left[ \Vert \mathsf{U}^m-\mathsf{U}^m_h\Vert _{\scriptscriptstyle 0,D}^2 \right] \right) ^{\scriptscriptstyle 1/2}\) is the space-discretization error at \(\tau _m\).

5.1 Estimating the time-discretization error

The convergence estimate of Proposition 4.1 is the main tool in providing a discrete in time \(L^{\infty }_t(L^2_{\scriptscriptstyle P}(L^2_x))\) error estimate of the time-discretization error (cf. Yan 2005; Kossioris and Zouraris 2010, 2013).

Proposition 5.1

Let \(\mathsf{u}\) be the solution to (1.5) and \((\mathsf{U}^m)_{m=0}^{\scriptscriptstyle M}\) be the time-discrete approximations of \(\mathsf{u}\) defined by (5.1)–(5.2). Then, there exists a constant \({\widehat{c}}_{{\scriptscriptstyle \mathrm TDR}}\), independent of \(\Delta {t}\), \(\mathsf{M}\) and \(\Delta \tau \), such that

$$\begin{aligned} \max _{0\le m \le {\scriptscriptstyle M}}{\mathcal {E}}_{\scriptscriptstyle \mathrm{TDR}}^{m} \le \,{\widehat{c}}_{{\scriptscriptstyle \mathrm TDR}}\,\epsilon ^{-\frac{1}{2}} \,\Delta \tau ^{\frac{1}{8}-\epsilon } \quad \forall \,\epsilon \in \left( 0,\tfrac{1}{8}\right] . \end{aligned}$$
(5.4)

Proof

In the sequel, we will use the symbol C to denote a generic constant that is independent of \(\Delta {t}\), \(\mathsf{M}\), and \(\Delta \tau \), and may change value from one line to the other.

First, we introduce some notation by letting \(\mathsf{I}:L^2(D)\rightarrow L^2(D)\) be the identity operator, \(\mathsf{Y}:H^2(D)\rightarrow L^2(D)\) be the differential operator \(\mathsf{Y}:=\mathsf{I}-\Delta \tau \,\mu \,\partial _x^2\), and \(\mathsf{\Lambda }:L^2(D)\rightarrow {\dot{\mathbf{H}}}^4(D)\) be the inverse elliptic operator \(\mathsf{\Lambda }:=(\mathsf{I}+\Delta \tau \,\partial _x^4)^{-1}\). Then, for \(m=1,\dots ,M\), we define the operator \(\mathsf{Q}^m:L^2(D)\rightarrow {\dot{\mathbf{H}}}^4(D)\) by \(\mathsf{Q}^m:=(\mathsf{\Lambda }\circ \mathsf{Y})^{m-1}\circ \mathsf{\Lambda }\). In addition, for given \(w_0\in {\dot{\mathbf{H}}}^2(D)\), let \(({\mathcal {S}}_{\scriptscriptstyle {\Delta \tau }}^m(w_0))_{m=0}^{\scriptscriptstyle M}\) be time-discrete approximations of the solution to the deterministic problem (1.4), defined by (4.1)–(4.3). Then, using a simple induction argument, we conclude that

$$\begin{aligned} {\mathcal {S}}_{\scriptscriptstyle {\Delta \tau }}^m(w_0)=\mathsf{Q}^{m}(w_0),\quad m=1,\dots ,M. \end{aligned}$$
(5.5)

Let \(m\in \{1,\dots ,M\}\). Applying a simple induction argument on (5.2), we conclude that

$$\begin{aligned} \mathsf{U}^m=\sum _{\ell =1}^{\scriptscriptstyle m} \int _{\scriptscriptstyle \Delta _\ell } \mathsf{Q}^{m-\ell +1}\left( \partial _x{\mathcal {W}}(\tau ,\cdot )\right) \,\mathrm{d}\tau , \end{aligned}$$

which, along with (1.6) and (5.5), yields

$$\begin{aligned} \mathsf{U}^m&=-\tfrac{1}{\Delta {t}}\,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\sum _{n=1}^{\scriptscriptstyle \mathsf{N}} R_i^n\,\lambda _i\,\left( \,\sum _{\ell =1}^{\scriptscriptstyle m}\int _{\scriptscriptstyle \Delta _\ell } {\mathcal {X}}_{\scriptscriptstyle T_n}(\tau ) \,{\mathcal {S}}_{\scriptscriptstyle \Delta \tau }^{m-\ell +1}(\varepsilon _i)\,\mathrm{d}\tau \,\right) \nonumber \\&=-\tfrac{1}{\Delta {t}}\,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\sum _{n=1}^{\scriptscriptstyle \mathsf{N}} R_i^n\,\lambda _i\,\left[ \,\int _0^{\scriptscriptstyle T} {\mathcal {X}}_{\scriptscriptstyle T_n}(\tau )\,\left( \,\sum _{\ell =1}^{\scriptscriptstyle m} {\mathcal {X}}_{\scriptscriptstyle \Delta _{\ell }}(\tau ) \,{\mathcal {S}}_{\scriptscriptstyle \Delta \tau }^{m-\ell +1}(\varepsilon _i)\,\right) \,\mathrm{d}\tau \,\right] \nonumber \\&=-\tfrac{1}{\Delta {t}}\,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\sum _{n=1}^{\scriptscriptstyle \mathsf{N}} R_i^n\,\lambda _i\,\left[ \,\int _{\scriptscriptstyle T_n} \left( \,\sum _{\ell =1}^{\scriptscriptstyle m} {\mathcal {X}}_{\scriptscriptstyle \Delta _{\ell }}(\tau ) \,{\mathcal {S}}_{\scriptscriptstyle \Delta \tau }^{m-\ell +1}(\varepsilon _i)\,\right) \,\mathrm{d}\tau \,\right] . \end{aligned}$$
(5.6)

In addition, using (1.9) and (1.6), and proceeding in similar manner, we arrive at

$$\begin{aligned} \mathsf{u}^m&=\int _0^{\tau _m}{\mathcal {S}}(\tau _m-\tau ) \,\left( \partial _x{\mathcal {W}}(\tau ,\cdot )\right) \,\mathrm{d}\tau \nonumber \\&=-\tfrac{1}{\Delta {t}}\,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\,\sum _{n=1}^{\scriptscriptstyle \mathsf{N}} R_i^n\,\lambda _i \,\left[ \,\int _{\scriptscriptstyle T_n}\left( \,\sum _{\ell =1}^m {\mathcal {X}}_{\scriptscriptstyle {\Delta _{\ell }}}(\tau ) \,{\mathcal {S}}(\tau _m-\tau ) \left( \varepsilon _i\right) \,\right) \,\mathrm{d}\tau \,\right] . \end{aligned}$$
(5.7)

Thus, using (5.6) and (5.7) along with Remark 1.8, we obtain

$$\begin{aligned} \begin{aligned} \left( \,{\mathcal {E}}^m_{{\scriptscriptstyle \mathrm TDR}}\,\right) ^2=&\,\tfrac{1}{\Delta {t}} \sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\,\sum _{n=1}^{\scriptscriptstyle \mathsf{N}} \lambda _i^2\,\\&\int _{\scriptscriptstyle D}\left( \,\int _{\scriptscriptstyle T_n} \left( \,\sum _{\ell =1}^m {\mathcal {X}}_{\scriptscriptstyle {\Delta _{\ell }}}(\tau ) \,\,\left[ {\mathcal {S}}_{\scriptscriptstyle \Delta \tau }^{m-\ell +1}(\varepsilon _i) -{\mathcal {S}}(\tau _m-\tau )(\varepsilon _i)\right] \,\right) \,\mathrm{d}\tau \,\right) ^2\,\mathrm{d}x\\ \le&\,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\,\sum _{n=1}^{\scriptscriptstyle \mathsf{N}} \lambda _i^2\,\int _{\scriptscriptstyle D}\int _{\scriptscriptstyle T_n} \left( \,\sum _{\ell =1}^m{\mathcal {X}}_{\scriptscriptstyle {\Delta _{\ell }}}(\tau ) \,\left[ {\mathcal {S}}_{\scriptscriptstyle \Delta \tau }^{m-\ell +1}(\varepsilon _i) -{\mathcal {S}}(\tau _m-\tau )(\varepsilon _i)\right] \,\right) ^2\,\mathrm{d}\tau \,\mathrm{d}x\\ \le&\,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\, \lambda _i^2\,\int _0^{\scriptscriptstyle T}\int _{\scriptscriptstyle D} \left( \,\sum _{\ell =1}^m{\mathcal {X}}_{\scriptscriptstyle {\Delta _{\ell }}}(\tau ) \,\left[ {\mathcal {S}}_{\scriptscriptstyle \Delta \tau }^{m-\ell +1}(\varepsilon _i) -{\mathcal {S}}(\tau _m-\tau )(\varepsilon _i)\right] \,\right) ^2\,\mathrm{d}x\,\mathrm{d}\tau \\ \le&\,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\, \lambda _i^2\,\left( \,\sum _{\ell =1}^m\int _{\scriptscriptstyle \Delta _{\ell }} \Vert {\mathcal {S}}_{\scriptscriptstyle \Delta \tau }^{m-\ell +1}(\varepsilon _i) -{\mathcal {S}}(\tau _m-\tau )(\varepsilon _i)\Vert _{\scriptscriptstyle 0,D}^2\,\mathrm{d}\tau \,\right) ,\\ \end{aligned} \end{aligned}$$

which, easily, yields

$$\begin{aligned} {\mathcal {E}}^m_{{\scriptscriptstyle \mathrm TDR}}\le \,\sqrt{{\mathcal {B}}_1^m}+\sqrt{{\mathcal {B}}_2^m}, \end{aligned}$$
(5.8)

with

$$\begin{aligned} \begin{aligned} {\mathcal {B}}_1^m:=&\,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\, \lambda _i^2\,\left( \,\sum _{\ell =1}^m\Delta \tau \, \left\| {\mathcal {S}}_{\scriptscriptstyle \Delta \tau }^{m-\ell +1}(\varepsilon _i) -{\mathcal {S}}(\tau _{m-\ell +1})(\varepsilon _i)\right\| _{\scriptscriptstyle 0,D}^2\,\right) ,\\ {\mathcal {B}}_2^m:=&\,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\, \lambda _i^2\,\left( \,\sum _{\ell =1}^m\int _{\scriptscriptstyle \Delta _{\ell }} \left\| {\mathcal {S}}(\tau _{m-\ell +1})(\varepsilon _i) -{\mathcal {S}}(\tau _m-\tau )(\varepsilon _i)\right\| _{\scriptscriptstyle 0,D}^2\,\mathrm{d}\tau \,\right) .\\ \end{aligned} \end{aligned}$$

Proceeding as in the proof of Theorem 4.1 in Kossioris and Zouraris (2013), we get

$$\begin{aligned} \sqrt{{\mathcal {B}}_2^m}\le \,C\,\Delta \tau ^{\frac{1}{8}}. \end{aligned}$$
(5.9)

In addition, using the error estimate (4.4), it follows that:

$$\begin{aligned} \begin{aligned} \sqrt{{\mathcal {B}}_1^m}\le&\,C\,\Delta \tau ^{\theta } \,\left( \,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\lambda _i^2\,\Vert \varepsilon _i\Vert ^2_{\scriptscriptstyle {\dot{\mathbf{H}}}^{4\theta -2}}\,\right) ^{\frac{1}{2}}\\ \le&\,C\,\Delta \tau ^{\theta }\,\left( \,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}} \tfrac{1}{\lambda _i^{2-8\theta }}\,\right) ^{\frac{1}{2}}\quad \forall \,\theta \in [0,1].\\ \end{aligned} \end{aligned}$$

Setting \(\theta =\tfrac{1}{8}-\epsilon \) with \(\epsilon \in \left( 0,\tfrac{1}{8}\right] \), we have

$$\begin{aligned} \sqrt{{\mathcal {B}}_1^m}&\le C\,\Delta \tau ^{\frac{1}{8}-\epsilon } \,\left( \,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}} \tfrac{1}{i^{1+8\epsilon }}\, \right) ^{\frac{1}{2}} \nonumber \\&\le C\,\Delta \tau ^{\frac{1}{8}-\epsilon }\,\left( \,1+\int _1^{\scriptscriptstyle \mathsf{M}}x^{-1-8\epsilon }\,\mathrm{d}x\,\right) ^{\frac{1}{2}} \nonumber \\&\le C\,\Delta \tau ^{\frac{1}{8}-\epsilon }\,\epsilon ^{-\frac{1}{2}} \,\left( 1-\tfrac{1}{\mathsf{M}^{8\epsilon }}\right) ^{\frac{1}{2}}. \end{aligned}$$
(5.10)

Thus, the estimate (5.4) follows, easily, as a simple consequence of (5.8), (5.9), and (5.10). \(\square \)

5.2 Estimating the space-discretization error

The outcome of Proposition 4.2 will be used below in the derivation of a discrete in time \(L^{\infty }_t(L^2_{\scriptscriptstyle P}(L^2_x))\) error estimate of the space-discretization error (cf. Yan 2005; Kossioris and Zouraris 2010, 2013).

Proposition 5.2

Let \(r=2\) or 3, \((\mathsf{U}_h^m)_{m=0}^{\scriptscriptstyle M}\) be the fully discrete approximations defined by (1.10)–(1.11) and \((\mathsf{U}^m)_{m=0}^{\scriptscriptstyle M}\) be the time-discrete approximations defined by (5.1)–(5.2). Then, there exists a constant \({\widehat{c}}_{{\scriptscriptstyle \mathrm SDR}}>0\), independent of \(\mathsf{M}\), \(\Delta {t}\), \(\Delta \tau \) and h, such that

$$\begin{aligned} \max _{0\le {m}\le {\scriptscriptstyle M}}{\mathcal {E}}^m_{{\scriptscriptstyle \mathrm SDR}} \le \,{\widehat{c}}_{{\scriptscriptstyle \mathrm SDR}} \,\epsilon ^{-\frac{1}{2}} \,\,\,h^{\frac{r}{6}-\epsilon } \quad \forall \,\epsilon \in \left( 0,\tfrac{r}{6}\right] . \end{aligned}$$
(5.11)

Proof

In the sequel, we will use the symbol C to denote a generic constant that is independent of \(\Delta {t}\), \(\mathsf{M}\), \(\Delta \tau \), and h, and may change value from one line to the other.

Let us denote by \(\mathsf{I}:L^2(D)\rightarrow L^2(D)\) the identity operator, by \(\mathsf{Y}_h:\mathsf{M}_h^r\rightarrow \mathsf{M}_h^r\) the discrete differential operator \(\mathsf{Y}_h:=\mathsf{I}-\mu \,\Delta \tau \,(P_h\circ \partial _x^2)\), \(\mathsf{\Lambda }_h:L^2(D)\rightarrow \mathsf{M}^r_h\) be the inverse discrete elliptic operator \(\mathsf{\Lambda }_h:=(I+\Delta \tau \,B_h)^{-1}\circ P_h\). Then, for \(m=1,\dots ,M\), we define the auxiliary operator \(\mathsf{Q}_h^m:L^2(D)\rightarrow \mathsf{M}_h^r\) by \(\mathsf{Q}^m_h:=(\mathsf{\Lambda }_h\circ \mathsf{Y}_h)^{m-1}\circ \mathsf{\Lambda }_h\). In addition, for given \(w_0\in {\dot{\mathbf{H}}}^2(D)\), let \(({\mathcal {S}}_{h}^m(w_0))_{m=0}^{\scriptscriptstyle M}\) be fully discrete approximations of the solution to the deterministic problem (1.4), defined by (4.26)–(4.28). Then, using a simple induction argument, we conclude that

$$\begin{aligned} {\mathcal {S}}_{h}^m(w_0)=\mathsf{Q}_h^{m}(w_0),\quad m=1,\dots ,M. \end{aligned}$$
(5.12)

Let \(m\in \{1,\dots ,M\}\). Using a simple induction argument on (1.11), (1.6) and (5.12), we conclude that

$$\begin{aligned} \mathsf{U}_h^m&=\sum _{\ell =1}^{\scriptscriptstyle m} \int _{\scriptscriptstyle \Delta _\ell } \mathsf{Q}_h^{m-\ell +1}\left( \partial _x{\mathcal {W}}(\tau ,\cdot )\right) \,\mathrm{d}\tau \nonumber \\&=-\tfrac{1}{\Delta {t}}\,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\sum _{n=1}^{\scriptscriptstyle \mathsf{N}} R_i^n\,\lambda _i\,\left[ \,\int _{\scriptscriptstyle T_n} \left( \,\sum _{\ell =1}^{\scriptscriptstyle m} {\mathcal {X}}_{\scriptscriptstyle \Delta _{\ell }}(\tau ) \,{\mathcal {S}}_{h}^{m-\ell +1}(\varepsilon _i)\,\right) \,\mathrm{d}\tau \,\right] . \end{aligned}$$
(5.13)

After, using (5.13), (5.6), and Remark 1.8, and proceeding as in the proof of Proposition 5.1, we arrive at

$$\begin{aligned} {\mathcal {E}}^m_{{\scriptscriptstyle \mathrm SDR}}\le \left[ \,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\, \lambda _i^2\,\left( \,\sum _{\ell =1}^m\Delta \tau \, \Vert {\mathcal {S}}_{\scriptscriptstyle \Delta \tau }^{m-\ell +1}(\varepsilon _i) -{\mathcal {S}}^{m-\ell +1}_{h}(\varepsilon _i)\Vert _{\scriptscriptstyle 0,D}^2\,\mathrm{d}\tau \,\right) \,\right] ^{\frac{1}{2}}, \end{aligned}$$

which, along (4.29), yields

$$\begin{aligned} {\mathcal {E}}^m_{{\scriptscriptstyle \mathrm SDR}}&\le C\,h^{r\theta } \,\left( \,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\lambda _i^2\,\Vert \varepsilon _i\Vert ^2_{\scriptscriptstyle {\dot{\mathbf{H}}}^{3\theta -2}}\,\right) ^{\frac{1}{2}} \nonumber \\&\le C\,h^{r\theta }\,\left( \, \,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}}\tfrac{1}{\lambda _i^{2-6\theta }}\,\right) ^{\frac{1}{2}} \quad \forall \,\theta \in [0,1]. \end{aligned}$$
(5.14)

Setting \(\theta =\tfrac{1}{6}-\delta \) with \(\delta \in \left( 0,\tfrac{1}{6}\right] \), we have

$$\begin{aligned} \begin{aligned} {\mathcal {E}}^m_{{\scriptscriptstyle \mathrm SDR}}\le&\,C\,h^{\frac{r}{6}-r\delta } \,\left( \,\sum _{i=1}^{\scriptscriptstyle \mathsf{M}} \tfrac{1}{i^{1+6\delta }}\, \right) ^{\frac{1}{2}}\\ \le&\,C\,h^{\frac{r}{6}-r\delta }\,\left( \,1+\int _1^{\scriptscriptstyle \mathsf{M}}x^{-1-6\delta }\,\mathrm{d}x\,\right) ^{\frac{1}{2}}\\ \le&\,C\,h^{\frac{r}{6}-r\delta }\,\delta ^{-\frac{1}{2}} \,\left( \,1-\mathsf{M}^{-6\delta }\,\right) ^{\frac{1}{2}},\\ \end{aligned} \end{aligned}$$

which obviously yields (5.11) with \(\epsilon =r\delta \). \(\square \)

5.3 Estimating the total error

Theorem 5.3

Let \(r=2\) or 3, \(\mathsf{u}\) be the solution to the problem (1.5), and \((\mathsf{U}_h^m)_{m=0}^{\scriptscriptstyle M}\) be the finite element approximations of \(\mathsf{u}\) constructed by (1.10)–(1.11). Then, there exists a constant \({\widehat{c}}_{\scriptscriptstyle \mathrm{TTL}}>0\), independent of h, \(\Delta \tau \), \(\Delta {t}\) and \(\mathsf{M}\), such that

$$\begin{aligned} \max _{0\le {m}\le {\scriptscriptstyle M}}\left( \,{\mathbb {E}}\left[ \,\Vert \mathsf{U}_h^m-\mathsf{u}^m\Vert _{\scriptscriptstyle 0,D}^2\right] \,\right) ^{\frac{1}{2}} \le \,{\widehat{c}}_{\scriptscriptstyle \mathrm{TTL}}\,\left( \,\epsilon ^{-\frac{1}{2}}_1\,\Delta \tau ^{\frac{1}{8}-\epsilon _1} +\epsilon _2^{-\frac{1}{2}}\,\,\,h^{\frac{r}{6}-\epsilon _2}\,\right) \end{aligned}$$
(5.15)

for all \(\epsilon _1\in \left( 0,\tfrac{1}{8}\right] \) and \(\epsilon _2\in \left( 0,\frac{r}{6}\right] \).

Proof

The error bound (5.15) follows easily from (5.4), (5.11), and (5.3). \(\square \)