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

An exponential integrator multicontinuum homogenization method for fractional diffusion problem with multiscale coefficients

Yifei Gao, Yating Wang, Wing Tat Leung, Zhengya Yang

An exponential integrator multicontinuum homogenization method for fractional diffusion problem with multiscale coefficients

Yifei Gao, Yating Wang, Wing Tat Leung, Zhengya Yang
Abstract

In this paper, we present a robust and fully discretized method for solving the time fractional diffusion equation with high-contrast multiscale coefficients. We establish the homogenized equation in a coarse mesh using a multicontinuum approach and employ the exponential integrator method for time discretization. The multicontinuum upscaled model captures the physical characteristics of the solution for the high-contrast multiscale problem, including averages and gradient effects in each continuum at the coarse scale. We use the exponential integration method to address the nonlocality induced by the time fractional derivative and the stiffness from the multiscale coefficients in the semi-discretized problem. Convergence analysis of the numerical scheme is provided, along with illustrative numerical examples. Our results demonstrate the accuracy, efficiency, and improved stability for varying order of fractional derivatives.

1 Introduction

Time fractional diffusion equations are extensively utilized to describe the subdiffusion phenomena in porous media. The fractional operators can be a useful way to include memory in the dynamical process. It is modeled through fractional order derivatives carries information about its present as well as past states. Due to the highly heterogeneous nature of the underlying media, the multiscale coefficients in these equations often exhibit high-contrast and nonlocality characteristics. Such features are prevalent in various applications, including composite materials, groundwater modeling, and reservoir simulations. It is essential to consider the heterogeneities of porous media and the phenomena of subdiffusion and we could achieve more accurate modeling in practical applications. However, the time fractional derivative and multiscale feature of the semilinear problems pose significant challenges in simulation.

Indeed, the presence of high-contrast/multicontinuum fields leads to the lack of smoothness in the exact solution, and the fractional order of the time derivative results in long-persistent memory. Thus making it difficult to construct accurate and computationally efficient numerical schemes. To capture all scale information of the solution, small mesh size for the spatial discretization is usually needed to ensure accuracy. And the stiffness arising from the multiscale coefficients requires the use of very small time steps to meet stability requirements using conventional explicit temporal discretization methods. When the order of the time fractional derivative reaches zero, some numerical schemes turns out to be unstable due to the singularity. The nonlocality of the fractional derivative also makes long-time simulation prohibitive.

In previous works, many developed multiscale models have been proposed. These multiscale approaches include multisacle finite element methods (MsFEM) [1, 2, 3], mixed MsFEM [4], numerical homogenized methods [5, 6, 7], numerical upsacling methods [8, 9, 10], generalized multiscale finite element methods (GMsFEM) [11, 12, 13], localized orthogonal decomposition (LOD) [14], constraint energy minimizing GMsFEM (CEM-GMsFEM) [15] and nonlocal multicontinuum method (NLMC) [16], to name a few. Recently, the multicontinuum model is proposed based on the framework of multicontinuum homogenization [17, 18]. It can capture the property of the solution with multiple quantities. The main idea is to represent the solution via averages and its gradient effects in each continua. For diffusion equations, some previous works have been developed [19, 20, 21]. Researchers have developed many methods to discretize the time fractional derivative. The L1 scheme is one of the well known time stepping schemes that uses piecewise linear approximation on each time step [22, 23, 24, 25]. To overcome the difficulties caused by the stiffness and semilinear term, there has been a renewed interest in using exponential integrator methods [26, 27, 28, 29, 19].

In this work, we propose a robust and accurate scheme to solve the high-contrast time fractional diffusion equation. We derive the semi-discretized coarse scale model based on the multicontinuum homogenization method. Then we adopt the exponential integrator method for temporal discretization to ensure stability and efficiency for the semilinear problem. The convergence of our proposed approach is presented. The main contributions of this work are the following.

  • We derive time fractional multicontinuum upscaled model and analyze its convergence.

  • We then adopt exponential integrator for temporal discretization and compare it with L1 scheme for semilinear time fractional diffusion equation. We present the convergence result.

  • We perform some numerical experiments to show the accuracy and stability of our methods. We observe that the proposed method is more stable than L1 scheme when the fractional order is small for the high-contrast subdiffusion equation.

The paper is organized as follows. In Section 2, we will introduce the problem setting and some preliminary definitions for the latter discussion. In Section 3, we will discuss the spatial discretization and present the analysis of semi-discretized problems. We discuss the temporal discretization and present the analysis of full-discretized scheme in Section 4. We present numerical results in Section 5 and give concluding remarks in the final section.

2 Problem setup

We focus on the following time fractional diffusion equation

{Dtα0Cudiv(κu)=f(u)inΩ×(0,T],u(t,x)=0onΩ×(0,T],u(0,x)=u0(x)inΩ,\left\{\begin{aligned} {}^{C}_{0}{D}_{t}^{\alpha}u-\mathrm{div}(\kappa\nabla u)&=f(u)\ &\mathrm{in}\,&\Omega\times\bigl{(}0,T\bigr{]},\\ u(t,x)&=0\ &\mathrm{on}\,&\partial\Omega\times\bigl{(}0,T\bigr{]},\\ u(0,x)&=u_{0}(x)\ &\mathrm{in}\,&\Omega,\end{aligned}\right. (1)

where Dtα0Cu{}^{C}_{0}{D}_{t}^{\alpha}u is defined as the Caputo fractional derivatives of order α(0,1)\alpha\in(0,1) [24]

Dtα0Cu(x,t)=1Γ(1α)0tu(x,s)s1(ts)αds.{}^{C}_{0}{D}_{t}^{\alpha}u(x,t)=\frac{1}{\Gamma(1-\alpha)}\int_{0}^{t}\frac{\partial u(x,s)}{\partial s}\frac{1}{(t-s)^{\alpha}}\,\mathrm{d}s. (2)

We assume that κ\kappa is a high-contrast heterogeneous permeability field. It is uniformly positive and bounded in Ω,i.e. 0<κ0<κ1<\Omega,\ \,\mathrm{i}.\mathrm{e}.\,\ \exists\,0<\kappa_{0}<\kappa_{1}<\infty such that κ0κκ1,inΩ¯\kappa_{0}\leq\kappa\leq\kappa_{1},\ \mathrm{in}\,\overline{\Omega}.

There are three distinct scales which is highly related to our proposed homogenization method:

  • The physical microscopic scale ε\varepsilon, which represents the behavior and microscopic structure with high-contrast;

  • The computational upscaling scale HH, where the homogenized problem is posed;

  • The observing scale hh, where the multicontinuum quantities, such as local averages of the multiscale solution, are defined.

These scale parameter satisfies εh<H\varepsilon\leq h<H.

We start with the weak form of (1): find uH01(Ω)u\in H_{0}^{1}(\Omega) satisfying

(0CDtαu,v)+aε(u,v)=(f,v)vH01(Ω),\bigl{(}\,_{0}^{C}D_{t}^{\alpha}u,v\,\bigr{)}+a_{\varepsilon}(u,v)=(f,v)\quad\forall\,v\in H_{0}^{1}(\Omega), (3)

where (,)(\cdot,\cdot) represent the standard L2L^{2} inner product and aε(,)a_{\varepsilon}(\cdot,\cdot) is defined as aε(u,v)=Ωκuva_{\varepsilon}(u,v)=\int_{\Omega}\kappa\nabla u\cdot\nabla v. Let ωi\omega_{i} be disjoint and Ω=iωi\Omega=\bigcup\limits_{i}\omega_{i}. We then extent the definition of the bilinear operator aε(,)a_{\varepsilon}(\cdot,\cdot) to H1(Ω)H^{1}(\Omega) as follows

aε(u,v)=iωiκu|ωiv|ωi.a_{\varepsilon}(u,v)=\sum\limits_{i}\int_{\omega_{i}}\kappa\nabla u|_{\omega_{i}}\cdot\nabla v|_{\omega_{i}}.

Moreover, we define ua(ω)(ωκ|u|2)12\bigl{\lVert}\,u\,\bigr{\rVert}_{a(\omega)}\coloneqq\bigl{(}\,\int_{\omega}\kappa\bigl{\lvert}\,\nabla u\,\bigr{\rvert}^{2}\,\bigr{)}^{\frac{1}{2}} and denote a=a(Ω)\bigl{\lVert}\,\cdot\,\bigr{\rVert}_{a}=\bigl{\lVert}\,\cdot\,\bigr{\rVert}_{a(\Omega)}. Let \bigl{\lVert}\,\cdot\,\bigr{\rVert} be the standard L2L^{2} norm of the domain Ω\Omega. We also assume the domain Ω\Omega can be divided into two subregions Ω=Ω0Ω1\Omega=\Omega_{0}\cup\Omega_{1}. The subdomains Ω0\Omega_{0} and Ω1\Omega_{1} represent the region with high and low conductivity respectively.

3 Time fractional multicontinuum upscaling model

In this section, we would like to derive the multicontinuum upscaling model. We start with the construction of local basis and present the related convergence analysis. The computational domain Ω\Omega can be partitioned with respect to different scales. Given z[0,H]dz\in[0,H]^{d} (d=2,3d=2,3), we define a rectangular partition 𝒯H(z)\mathcal{T}_{H}(z) with mesh size HH

𝒯H(z)={KΩK=xi+(H2,H2)dΩ,xizHd},i=1,2,\mathcal{T}_{H}(z)=\bigl{\{}\,K\subset\Omega\mid K=x_{i}+(-\frac{H}{2},\frac{H}{2})^{d}\cap\Omega,\frac{x_{i}-z}{H}\in\mathbb{Z}^{d}\,\bigr{\}},\ i=1,2,\cdots

For each coarse block KH(xi)=Kz,H(xi)xi+(H2,H2)dΩ𝒯H(z)K_{H}(x_{i})=K_{z,H}(x_{i})\coloneqq x_{i}+(-\frac{H}{2},\frac{H}{2})^{d}\cap\Omega\in\mathcal{T}_{H}(z), we can separate the element into two subregions, KH,0(xi)=KH(xi)Ω0,KH,1(xi)=KH(xi)Ω1K_{H,0}(x_{i})=K_{H}(x_{i})\cap\Omega_{0},\ K_{H,1}(x_{i})=K_{H}(x_{i})\cap\Omega_{1}. The index set is defined as IH=Iz,H={xiKH(xi)𝒯H(z)}I_{H}=I_{z,H}=\bigl{\{}\,x_{i}\mid K_{H}(x_{i})\in\mathcal{T}_{H}(z)\,\bigr{\}}, it contains all of the center points of the coarse elements in the partition 𝒯H(z)\mathcal{T}_{H}(z). Similarly, we can define the other partition 𝒯h\mathcal{T}_{h} and its index set IhI_{h} with the mesh size hh. An illustration of the mesh grid is shown in Figure 1.

Refer to caption
Fig. 1: The fine grid, coarse grid KH(xi)K_{H}(x_{i}), oversampling domain KH(xi)+K_{H}(x_{i})^{+}.

3.1 Construction of local problems

In this subsection, we construct the local problems for the multicontinuum upscaling model. We define the auxiliary basis ψi,j=χKH(xi)=1incontinuumi\psi_{i,j}=\chi_{K_{H}(x_{i})}=1\quad\mathrm{in\ continuum\ }i and 0otherwise\ 0\quad\mathrm{otherwise}. Based on CEM-GMsFEM/NLMC, the global basis functions φl,jH01(Ω)j=0,1\varphi_{l,j}\in H_{0}^{1}(\Omega)\quad j=0,1 can be obtained by solving the constrained energy minimization problems

aε(φl,i,v)+xkIz,hj=0,1ηk,j(ψk,j,v)\displaystyle a_{\varepsilon}(\varphi_{l,i},v)+\sum\limits_{x_{k}\in I_{z,h}}\sum\limits_{j=0,1}\eta^{k,j}(\psi_{k,j},v) =0,\displaystyle=0, (4)
(φl,i,ψk,j)\displaystyle(\varphi_{l,i},\psi_{k,j}) =δijδlkΩψk,j,\displaystyle=\delta_{ij}\delta_{lk}\int_{\Omega}\psi_{k,j},

where ηk,j\eta^{k,j} represents the energy constraint constant. We set the global multiscale space as Vglo,h=spanl,j{φl,j}={vH01(Ω)v=l,jvl,jφl,j}V_{glo,h}=\mathop{\mathrm{span}}_{l,j}\bigl{\{}\,\varphi_{l,j}\,\bigr{\}}=\bigl{\{}\,v\in H_{0}^{1}(\Omega)\mid v=\sum_{l,j}v_{l,j}\varphi_{l,j}\,\bigr{\}}. The convergence of the CEM-GMsFEM method with global basis is as follows [16], and it will be used in our latter discussions.

Lemma 3.1.

Let fL2(Ω)f\in L^{2}(\Omega) and uglo,hVglo,hu_{glo,h}\in V_{glo,h} be the solution of aε(uglo,h,v)=(f,v)vVglo,ha_{\varepsilon}(u_{glo,h},v)=(f,v)\quad\forall\,v\in V_{glo,h}. We have Ω(uglo,hu)ψk,j=0k,j\int_{\Omega}(u_{glo,h}-u)\psi_{k,j}=0\quad\forall\,k,j and uglo,hua=O(h)\bigl{\lVert}\,u_{glo,h}-u\,\bigr{\rVert}_{a}=O(h).

We can denote uglo,h=xlIzUl,iφl,iu_{glo,h}=\sum_{x_{l}\in I_{z}}U_{l,i}\,\varphi_{l,i} where the coefficients Ul,i=(u,ψl,i)Ωψl,iU_{l,i}=\dfrac{(u,\psi_{l,i})}{\int_{\Omega}\psi_{l,i}} are the local average of the solution uu in the coarse elements KH(xl)K_{H}(x_{l}).

In this work, we would like to introduce a time fractional multicontinuum model based on [17, 18], where the local problems correspond to the average and the gradient effect of the solution. To be more specific, to derive the homogenized equation on coarse scale, we first need to find ϕi\phi_{i} and ϕim\phi_{i}^{m} in ii-th coarse region from the following two cell problems with constraints

{KH+(x)κ(y)yϕi(x,y)yv=k=0,1xjIHKH(x)+djk(ψk(xj,y),v),(ϕi(x,y),ψk(xj,y))=δkiKH(x)+ψk(xj,y)xjIHKH(x)+,\left\{\begin{aligned} \int_{K_{H}^{+}(x)}\kappa(y)\nabla_{y}\phi_{i}(x,y)\nabla_{y}v&=\sum\limits_{k=0,1}\sum\limits_{x_{j}\in I_{H}\cap K_{H}(x)^{+}}d_{jk}\bigl{(}\,\psi_{k}(x_{j},y),v\,\bigr{)},\\ \bigl{(}\,\phi_{i}(x,y),\psi_{k}(x_{j},y)\,\bigr{)}&=\delta_{ki}\int_{K_{H}(x)^{+}}\psi_{k}(x_{j},y)\quad\forall\,x_{j}\in I_{H}\cap K_{H}(x)^{+},\end{aligned}\right. (5)
{KH+(x)κ(y)yϕim(x,y)yv=k=0,1xjIHKH(x)+djk(ψk(xj,y),v),(ϕim(x,y),ψk(xj,y))=δkiKH(x)+(ymcx(l))ψk(xj,y)xjIHKH(x)+,\left\{\begin{aligned} \int_{K_{H}^{+}(x)}\kappa(y)\nabla_{y}\phi_{i}^{m}(x,y)\nabla_{y}v&=\sum\limits_{k=0,1}\sum\limits_{x_{j}\in I_{H}\cap K_{H}(x)^{+}}d_{jk}\bigl{(}\,\psi_{k}(x_{j},y),v\,\bigr{)},\\ \bigl{(}\,\phi_{i}^{m}(x,y),\psi_{k}(x_{j},y)\,\bigr{)}&=\delta_{ki}\int_{K_{H}(x)^{+}}(y^{m}-c_{x}^{(l)})\psi_{k}(x_{j},y)\quad\forall\,x_{j}\in I_{H}\cap K_{H}(x)^{+},\end{aligned}\right. (6)

where KH(x)+K_{H}(x)^{+} is the oversampling domain which enrich KH(x)K_{H}(x) by khk\ h-layer, y(m)y^{(m)} is the ll-th coordinate of ydy\in\mathbb{R}^{d} and cx(m)c_{x}^{(m)} are some constants such that KH(x)(y(m)cx(m))=0\int_{K_{H}(x)}(y^{(m)}-c_{x}^{(m)})=0. The solution ϕi\phi_{i} of the cell problem (5) describe the averages in each continua while the solution ϕim\phi_{i}^{m} of (6) represents the linear basis function. The upsacling model with the following scaling holds

ϕi=O(1),ϕi=O(1ε),ϕim=O(ε),ϕim=O(1).\bigl{\lVert}\,\phi_{i}\,\bigr{\rVert}=O(1),\ \bigl{\lVert}\,\nabla\phi_{i}\,\bigr{\rVert}=O(\frac{1}{\varepsilon}),\ \bigl{\lVert}\,\phi_{i}^{m}\,\bigr{\rVert}=O(\varepsilon),\ \bigl{\lVert}\,\nabla\phi_{i}^{m}\,\bigr{\rVert}=O(1). (7)

3.2 Multi-continuum upscaling model

Now we present the construction of the multiscale upscaling model. We assume that in each coarse block KH(x)K_{H}(x), the solution uu could be approximated by u=ϕiUi+ϕimmUiu=\phi_{i}U_{i}+\phi_{i}^{m}\nabla_{m}U_{i}, where UiU_{i} can be view as a limit of KH(x)uψiKH(x)ψi\frac{\int_{K_{H}(x)}u\psi_{i}}{\int_{K_{H}(x)}\psi_{i}} as the block size goes to zero. It represents the average solution in ii-th continua. With this expression of uu, we can write

KH(x)κuv=KH(x)κ(ϕiUi)v+KH(x)κ(ϕimmUi)v.\int_{K_{H}(x)}\kappa\nabla u\cdot\nabla v=\int_{K_{H}(x)}\kappa\nabla(\phi_{i}U_{i})\cdot\nabla v+\int_{K_{H}(x)}\kappa\nabla(\phi_{i}^{m}\nabla_{m}U_{i})\cdot\nabla v. (8)

Approximately, we have

KH(x)κ(ϕiUi)v\displaystyle\int_{K_{H}(x)}\kappa\nabla(\phi_{i}U_{i})\cdot\nabla v =KH(x)κ(ϕi)Uiv=Ui(x)KH(x)κϕiv,\displaystyle=\int_{K_{H}(x)}\kappa(\nabla\phi_{i})U_{i}\cdot\nabla v=U_{i}(x)\int_{K_{H}(x)}\kappa\nabla\phi_{i}\cdot\nabla v,
KH(x)κ(ϕimmUi)v\displaystyle\int_{K_{H}(x)}\kappa\nabla(\phi_{i}^{m}\nabla_{m}U_{i})\cdot\nabla v =KH(x)κ(ϕim)mUiv=mUi(x)KH(x)κϕimv.\displaystyle=\int_{K_{H}(x)}\kappa\nabla(\phi_{i}^{m})\nabla_{m}U_{i}\cdot\nabla v=\nabla_{m}U_{i}(x)\int_{K_{H}(x)}\kappa\nabla\phi_{i}^{m}\cdot\nabla v.

We take v=ϕjVj+ϕjnnVjv=\phi_{j}V_{j}+\phi_{j}^{n}\nabla_{n}V_{j}. Substitute the above equations into (8), we obtain

KH(x)κuv=KH(x)κ(ϕiUi)v+KH(x)κ(ϕimmUi)v\displaystyle\int_{K_{H}(x)}\kappa\nabla u\cdot\nabla v=\int_{K_{H}(x)}\kappa\nabla(\phi_{i}U_{i})\cdot\nabla v+\int_{K_{H}(x)}\kappa\nabla(\phi_{i}^{m}\nabla_{m}U_{i})\cdot\nabla v
=\displaystyle= mUi(x)nVj(x)KH(x)κϕimϕjn+mUi(x)Vj(x)KH(x)κϕimϕj\displaystyle\nabla_{m}U_{i}(x)\nabla_{n}V_{j}(x)\int_{K_{H}(x)}\kappa\nabla\phi_{i}^{m}\cdot\nabla\phi_{j}^{n}+\nabla_{m}U_{i}(x)V_{j}(x)\int_{K_{H}(x)}\kappa\nabla\phi_{i}^{m}\cdot\nabla\phi_{j}
+Ui(x)nVj(x)KH(x)κϕiϕjn+Ui(x)Vj(x)KH(x)κϕiϕj.\displaystyle+U_{i}(x)\nabla_{n}V_{j}(x)\int_{K_{H}(x)}\kappa\nabla\phi_{i}\cdot\nabla\phi_{j}^{n}+U_{i}(x)V_{j}(x)\int_{K_{H}(x)}\kappa\nabla\phi_{i}\cdot\nabla\phi_{j}.

Denote by αijmn=KH(x)κϕimϕjn,βijm=KH(x)κϕimϕj,γij=KH(x)κϕiϕj\alpha_{ij}^{mn}=\int_{K_{H}(x)}\kappa\nabla\phi_{i}^{m}\cdot\nabla\phi_{j}^{n},\ \beta_{ij}^{m}=\int_{K_{H}(x)}\kappa\nabla\phi_{i}^{m}\cdot\nabla\phi_{j},\ \gamma_{ij}=\int_{K_{H}(x)}\kappa\nabla\phi_{i}\cdot\nabla\phi_{j}. The above equations can be simplified as

KH(x)κuv=mUiαijmnnVj+mUiβijmVj+UiβjinnVj+UiγijVj.\int_{K_{H}(x)}\kappa\nabla u\cdot\nabla v=\nabla_{m}U_{i}\alpha_{ij}^{mn}\nabla_{n}V_{j}+\nabla_{m}U_{i}\beta_{ij}^{m}V_{j}+U_{i}\beta_{ji}^{n}\nabla_{n}V_{j}+U_{i}\gamma_{ij}V_{j}.

Assume the cell problems do not depend on time tt, for the time fractional problem (3), we have Dtα0Cu=ϕiDtα0CUi+ϕimDtα0CmUi{}^{C}_{0}{D}_{t}^{\alpha}u=\phi_{i}\,{}_{0}^{C}D_{t}^{\alpha}U_{i}+\phi_{i}^{m}\,{}_{0}^{C}D_{t}^{\alpha}\nabla_{m}U_{i}

KH(x)Dtα0Cuv=\displaystyle\int_{K_{H}(x)}\,{}_{0}^{C}D_{t}^{\alpha}u\,v= Dtα0CUiVjKH(x)ϕiϕj+0CDtαmUiVjKH(x)ϕimϕj\displaystyle{}_{0}^{C}D_{t}^{\alpha}U_{i}V_{j}\int_{K_{H}(x)}\phi_{i}\phi_{j}+\,_{0}^{C}D_{t}^{\alpha}\nabla_{m}U_{i}V_{j}\int_{K_{H}(x)}\phi_{i}^{m}\phi_{j}
+0CDtαUinVjKH(x)ϕiϕjn+0CDtαmUinVjKH(x)ϕimϕjn.\displaystyle+_{0}^{C}D_{t}^{\alpha}U_{i}\nabla_{n}V_{j}\int_{K_{H}(x)}\phi_{i}\phi_{j}^{n}+\,_{0}^{C}D_{t}^{\alpha}\nabla_{m}U_{i}\nabla_{n}V_{j}\int_{K_{H}(x)}\phi_{i}^{m}\phi_{j}^{n}.

We neglect the higher order terms based on scalings (7). Let mij=KH(x)ϕiϕjm_{ij}=\int_{K_{H}(x)}\phi_{i}\phi_{j}, we derive the time fractional derivative term as

KH(x)Dtα0Cuv=0CDtαUimijVj.\int_{K_{H}(x)}\,{}_{0}^{C}D_{t}^{\alpha}u\,v=\,_{0}^{C}D_{t}^{\alpha}U_{i}m_{ij}V_{j}.

We then define rescaled quantities as follows

αijmn^=1KH(x)αijmn,βijm^=εKH(x)βijm,γij^=ε2KH(x)γij,mij^=1KH(x)mij.\widehat{\alpha_{ij}^{mn}}=\frac{1}{\bigl{\lVert}\,K_{H}(x)\,\bigr{\rVert}}\alpha_{ij}^{mn},\ \widehat{\beta_{ij}^{m}}=\frac{\varepsilon}{\bigl{\lVert}\,K_{H}(x)\,\bigr{\rVert}}\beta_{ij}^{m},\ \widehat{\gamma_{ij}}=\frac{\varepsilon^{2}}{\bigl{\lVert}\,K_{H}(x)\,\bigr{\rVert}}\gamma_{ij},\ \widehat{m_{ij}}=\frac{1}{\bigl{\lVert}\,K_{H}(x)\,\bigr{\rVert}}m_{ij}.

Thus, the weak form of the time fractional homogenized problem can be wirtten as

Ωmij^0CDtαUiVj\displaystyle\int_{\Omega}\widehat{m_{ij}}\,_{0}^{C}D_{t}^{\alpha}U_{i}V_{j} (9)
+\displaystyle+ Ωαijmn^mUinVj+1εΩβijm^mUiVj+1εΩβjin^nVjUi+1ε2Ωγij^UiVj\displaystyle\int_{\Omega}\widehat{\alpha_{ij}^{mn}}\nabla_{m}U_{i}\nabla_{n}V_{j}+\frac{1}{\varepsilon}\int_{\Omega}\widehat{\beta_{ij}^{m}}\nabla_{m}U_{i}V_{j}+\frac{1}{\varepsilon}\int_{\Omega}\widehat{\beta_{ji}^{n}}\nabla_{n}V_{j}U_{i}+\frac{1}{\varepsilon^{2}}\int_{\Omega}\widehat{\gamma_{ij}}U_{i}V_{j}
=\displaystyle= ΩfVj.\displaystyle\int_{\Omega}fV_{j}.

Let

Mij=Ωmij^ϕiϕj,Aij=Ωαijmn^mϕimnϕjn+1εβijm^mϕimϕj+1εβjin^nϕjnϕi+1ε2γij^ϕiϕj,M_{ij}=\int_{\Omega}\widehat{m_{ij}}\phi_{i}\phi_{j},\ A_{ij}=\int_{\Omega}\widehat{\alpha_{ij}^{mn}}\nabla_{m}\phi_{i}^{m}\nabla_{n}\phi_{j}^{n}+\frac{1}{\varepsilon}\widehat{\beta_{ij}^{m}}\nabla_{m}\phi_{i}^{m}\phi_{j}+\frac{1}{\varepsilon}\widehat{\beta_{ji}^{n}}\nabla_{n}\phi_{j}^{n}\phi_{i}+\frac{1}{\varepsilon^{2}}\widehat{\gamma_{ij}}\phi_{i}\phi_{j},

The matrix form of (9) is

M0CDtαU(t)+AU(t)=f(t),M\,_{0}^{C}D_{t}^{\alpha}U(t)+AU(t)=f(t), (10)

where M=(Mij),A=(Aij)M=(M_{ij}),\ A=(A_{ij}).

3.3 Convergence of multi-continuum upscaling model

In this subsection, we will present the convergence of the semi-discretized problem. Firstly, we define the nonlocal multicontinuum (NLMC) downscaling operators

Ph,H:(C1(Ω))2\displaystyle P_{h,H}\colon\bigl{(}C^{1}(\Omega)\bigr{)}^{2} H1(𝒯H(z)),\displaystyle\to H^{1}(\mathcal{T}_{H}(z)), (11)
(U0,U1)\displaystyle(U_{0},U_{1}) i=0,1xlIHχKH(xl)(ϕiUi(xl)+ϕimmUi(xl)),\displaystyle\mapsto\sum\limits_{i=0,1}\sum\limits_{x_{l}\in I_{H}}\chi_{K_{H}(x_{l})}\bigl{(}\,\phi_{i}U_{i}(x_{l})+\phi_{i}^{m}\nabla_{m}U_{i}(x_{l})\,\bigr{)},

where H1(𝒯H(z)){uL2(Ω)u|KH1(K),KΩ}H^{1}(\mathcal{T}_{H}(z))\coloneqq\bigl{\{}\,u\in L^{2}(\Omega)\mid u|_{K}\in H^{1}(K),\ \forall\,K\subset\Omega\,\bigr{\}}. We also define another downscaling operator on the space formed by CEM-GMsFEMs’ global basis as follows

Ph:(C1(Ω))2\displaystyle P_{h}\colon\bigl{(}C^{1}(\Omega)\bigr{)}^{2} H1(Ω),\displaystyle\to H^{1}(\Omega), (12)
(U0,U1)\displaystyle(U_{0},U_{1}) i=0,1xlIhUi(xl)φl,iz.\displaystyle\mapsto\sum\limits_{i=0,1}\sum\limits_{x_{l}\in I_{h}}U_{i}(x_{l})\varphi_{l,i}^{z}.

Define the L2L^{2} projection operator Πi,h\Pi_{i,h} on the auxiliary space Vaux,i=spanl{ψl,i}V_{\mathrm{aux},i}=\mathop{\mathrm{span}}_{l}\bigl{\{}\,\psi_{l,i}\,\bigr{\}} as

Πi,h(v)=xiIz,hKz,h(xl)ΩivKz,h(xl)Ωi1ψl,izi=0,1.\Pi_{i,h}(v)=\sum\limits_{x_{i}\in I_{z,h}}\frac{\int_{K_{z,h}(x_{l})\cap\Omega_{i}}v}{\int_{K_{z,h}(x_{l})\cap\Omega_{i}}1}\psi_{l,i}^{z}\ i=0,1. (13)

The L2L^{2} projection of the exact solution uu might lack smoothness, we assume that Πi,h(u)\Pi_{i,h}(u) can be extended to a function that belongs to C1(Ω)C^{1}(\Omega) with h0h\to 0, and use the notation Πi,h(u)\Pi_{i,h}(u) to represent a smooth function thereafter. Then, we have Ph(Π0,hu,Π1,hu)=uglo,hP_{h}\bigl{(}\,\Pi_{0,h}u,\Pi_{1,h}u\,\bigr{)}=u_{glo,h}. Notice that the local average of the summation on the domain Ω\Omega with respect to the point zz can be represented as follows: for any gC1(Ω)g\in C^{1}(\Omega), we have

[0,H]dxlIz,Hg(xl)=Ωg(x)dx.\int_{[0,H]^{d}}\sum\limits_{x_{l}\in I_{z,H}}g(x_{l})\,=\int_{\Omega}g(x)\,\mathrm{d}x.

Then a bilinear operator a~glo,h:(C1(Ω))2×(C1(Ω))2\widetilde{a}_{glo,h}\colon\bigl{(}C^{1}(\Omega)\bigr{)}^{2}\times\bigl{(}C^{1}(\Omega)\bigr{)}^{2}\to\mathbb{R} can be defined by

a~glo,h((U0,U1),(V0,V1))aε(Ph(U0,U1),Ph(V0,V1)).\widetilde{a}_{glo,h}((U_{0},U_{1}),(V_{0},V_{1}))\coloneqq a_{\varepsilon}\bigl{(}\,P_{h}(U_{0},U_{1}),P_{h}(V_{0},V_{1})\,\bigr{)}.

By Lemma 3.1, we obtain

a~glo,h((Π0,hu,Π1,hu),(Π0,hv,Π1,hv))=(f,(Π0,hv,Π1,hv))vH01(Ω).\widetilde{a}_{glo,h}\bigl{(}\,(\Pi_{0,h}u,\Pi_{1,h}u),(\Pi_{0,h}v,\Pi_{1,h}v)\,\bigr{)}=\bigl{(}\,f,(\Pi_{0,h}v,\Pi_{1,h}v)\,\bigr{)}\quad\forall\,v\in H_{0}^{1}(\Omega). (14)

Furthermore, a homogenized bilinear operator a~h,H:(H1(Ω))2×(H1(Ω))2\widetilde{a}_{h,H}\colon\bigl{(}H^{1}(\Omega)\bigr{)}^{2}\times\bigl{(}H^{1}(\Omega)\bigr{)}^{2}\to\mathbb{R} can be defined as

a~h,H((U0,U1),(V0,V1))Ω1HdKH(x)\displaystyle\widetilde{a}_{h,H}((U_{0},U_{1}),(V_{0},V_{1}))\coloneqq\int_{\Omega}\frac{1}{H^{d}}\int_{K_{H}(x)} κ(y)yi(Ui(x)ϕi(x,y)+mUi(x)ϕim(x,y))\displaystyle\kappa(y)\nabla_{y}\sum\limits_{i}\bigl{(}U_{i}(x)\phi_{i}(x,y)+\nabla_{m}U_{i}(x)\phi_{i}^{m}(x,y)\bigr{)}\cdot (15)
jy(Vi(x)ϕi(x,y)+mVi(x)ϕim(x,y))dydx\displaystyle\sum\limits_{j}\nabla_{y}\bigl{(}V_{i}(x)\phi_{i}(x,y)+\nabla_{m}V_{i}(x)\phi_{i}^{m}(x,y)\bigr{)}\,\mathrm{d}y\,\mathrm{d}x
=Ωi,j(αijmn^mUin\displaystyle=\int_{\Omega}\sum\limits_{i,j}\bigl{(}\widehat{\alpha_{ij}^{mn}}\nabla_{m}U_{i}\nabla_{n} Vj+βijm^mUiVj+βijn^UinVj+γij^UiVj).\displaystyle V_{j}+\widehat{\beta_{ij}^{m}}\nabla_{m}U_{i}V_{j}+\widehat{\beta_{ij}^{n}}U_{i}\nabla_{n}V_{j}+\widehat{\gamma_{ij}}U_{i}V_{j}\bigr{)}.

The following Lemma describes the property of the homogenized operator [18].

Lemma 3.2.

The operator a~h,H\widetilde{a}_{h,H} is an averaging of the multiscsale operator aεa_{\varepsilon} in sense of

a~h,H((U0,U1),(V0,V1))=1Hd[0,H]daε(Ph,H(U0,U1),Ph,H(V0,V1)).\widetilde{a}_{h,H}((U_{0},U_{1}),(V_{0},V_{1}))=\frac{1}{H^{d}}\int_{[0,H]^{d}}a_{\varepsilon}(P_{h,H}(U_{0},U_{1}),P_{h,H}(V_{0},V_{1}))\,. (16)

We consider (U0,h,H(t),U1,h,H(t))(H1(Ω))2\bigl{(}\,U_{0,h,H}(t),U_{1,h,H}(t)\,\bigr{)}\in\bigl{(}H^{1}(\Omega)\bigr{)}^{2} to be the solution of the following upscaled problem satisfying

1Hd[0,H]d(0CDtαPh,H((U0,h,H(t),U1,h,H(t))),Ph,H(V0,V1))\displaystyle\frac{1}{H^{d}}\int_{[0,H]^{d}}\bigl{(}\,_{0}^{C}D_{t}^{\alpha}P_{h,H}((U_{0,h,H}(t),U_{1,h,H}(t))),P_{h,H}(V_{0},V_{1})\,\bigr{)} (17)
+\displaystyle+ a~h,H((U0,h,H(t),U1,h,H(t)),(V0,V1))\displaystyle\widetilde{a}_{h,H}\bigl{(}\,(U_{0,h,H}(t),U_{1,h,H}(t)),(V_{0},V_{1})\,\bigr{)}
=\displaystyle= 1Hd[0,H]d(f,Ph,H(V0,V1))(V0,V1)(H1(Ω))2.\displaystyle\frac{1}{H^{d}}\int_{[0,H]^{d}}(f,P_{h,H}(V_{0},V_{1}))\quad\forall\,(V_{0},V_{1})\in\bigl{(}H^{1}(\Omega)\bigr{)}^{2}.

For a given s>0s>0, we define a norm a,1+s\bigl{\lVert}\,\cdot\,\bigr{\rVert}_{a,1+s} for the space (H1+s(Ω))2\bigl{(}H^{1+s}(\Omega)\bigr{)}^{2} where

(U0,U1)a,1+s2=U0H1+s2+κmaxU1H1+s2.\bigl{\lVert}\,(U_{0},U_{1})\,\bigr{\rVert}_{a,1+s}^{2}=\bigl{\lVert}\,U_{0}\,\bigr{\rVert}_{H^{1+s}}^{2}+\kappa_{\max}\bigl{\lVert}\,U_{1}\,\bigr{\rVert}_{H^{1+s}}^{2}.

The following lemma demonstrate the approximation of the two downscaling operators with the scale hHh\ll H if U0,U1U_{0},U_{1} is smooth sufficiently. We notice that the global basis φl,iz\varphi_{l,i}^{z} shows exponential decaying property [15].

Lemma 3.3.

Given U0,U1H1+s(Ω)U_{0},U_{1}\in H^{1+s}(\Omega) with 0<s10<s\leq 1. If the number of the oversampling layer kk satisfies 2kh<H=O(kh)2kh<H=O(kh). For some constant C0k=O(log(H1))C_{0}\geq k=O(\log(H^{-1})), we have [18]

Ph,H(U0,U1)Ph(U0,U1)Clog(1H)H1+s(U0,U1).\bigl{\lVert}\,P_{h,H}(U_{0},U_{1})-P_{h}(U_{0},U_{1})\,\bigr{\rVert}\leq C\log(\frac{1}{H})H^{1+s}\bigl{\lVert}\,(U_{0},U_{1})\,\bigr{\rVert}. (18)

The relationship between the solution uu and the L2L^{2} projection Ph(Πi,hu)P_{h}(\Pi_{i,h}u) can be described as

aε(u1Hd[0,H]dPh(Π0,hu,Π1,hu),v)=0vVglo,h.a_{\varepsilon}\bigl{(}\,u-\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h}(\Pi_{0,h}u,\Pi_{1,h}u)\,,v\,\bigr{)}=0\quad\forall\,v\in V_{glo,h}. (19)

Moreover, for the time fractional derivative, we have the folloiwng lemma.

Lemma 3.4.

For any function v(t)v(t) absolutely continuous on [0,T][0,T], we have [25]

v(t)0CDtαv(t)120CDtαv2(t)0<α<1.v(t)\,_{0}^{C}D_{t}^{\alpha}v(t)\geq\frac{1}{2}\,_{0}^{C}D_{t}^{\alpha}v^{2}(t)\quad 0<\alpha<1. (20)

With the approximation of the two downscaling operators and by the assumption of the smoothness of U0,h,HU_{0,h,H} and U1,h,HU_{1,h,H}, we derive the convergence of semi-discretize problem as following theorem based on above lemmas.

Theorem 3.1.

Let fL2(Ω)f\in L^{2}(\Omega) and uu be the solution to (1). With the same condition in Lemma 3.3, we have following estimates

u(t)1Hd[0,H]dPh,H(U0,h,H(t),U1,h,H(t))\displaystyle\bigl{\lVert}\,u(t)-\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h,H}\bigl{(}\,U_{0,h,H}(t),U_{1,h,H}(t)\,\bigr{)}\,\bigr{\rVert} (21)
\displaystyle\leq Ch2log(1H)H1+s(f+0t1(ts)1α0CDtα+2u(s)2ds)12.\displaystyle Ch^{2}\log(\frac{1}{H})H^{1+s}\biggl{(}\,\bigl{\lVert}\,f\,\bigr{\rVert}+\int_{0}^{t}\frac{1}{(t-s)^{1-\alpha}}\bigl{\lVert}\,_{0}^{C}D_{t}^{\alpha+2}u(s)\,\bigr{\rVert}^{2}\,\mathrm{d}s\,\biggr{)}^{\tfrac{1}{2}}.
Proof.

We estimate the error by three parts

u1Hd[0,H]dPh,H(U0,h,H,U1,h,H)\displaystyle\bigl{\lVert}\,u-\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h,H}\bigl{(}\,U_{0,h,H},U_{1,h,H}\,\bigr{)}\,\bigr{\rVert}
\displaystyle\leq u1Hd[0,H]dPh(Π0,hu,Π1,hu)\displaystyle\bigl{\lVert}\,u-\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h}(\,\Pi_{0,h}u,\Pi_{1,h}u\,)\,\bigr{\rVert}
+1Hd[0,H]d(Ph,HPh)(Π0,hu,Π1,hu)\displaystyle+\bigl{\lVert}\,\frac{1}{H^{d}}\int_{[0,H]^{d}}(P_{h,H}-P_{h})(\,\Pi_{0,h}u,\Pi_{1,h}u\,)\,\bigr{\rVert}
+1Hd[0,H]dPh,H((Π0,hu,Π1,hu)(U0,h,H,U1,h,H))\displaystyle+\bigl{\lVert}\,\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h,H}\bigl{(}\,(\Pi_{0,h}u,\Pi_{1,h}u)-(U_{0,h,H},U_{1,h,H})\,\bigr{)}\,\bigr{\rVert}
=\displaystyle\ = e1+e2+e3,\displaystyle\bigl{\lVert}\,e_{1}\,\bigr{\rVert}+\bigl{\lVert}\,e_{2}\,\bigr{\rVert}+\bigl{\lVert}\,e_{3}\,\bigr{\rVert},

where e1e_{1} represents the error between the exact solution and its projection on global CEM-GMsFEM space, e2e_{2} measures the difference between two downscaling operators. We have already obtained these estimates in Lemma 3.1. We will then focus on estimating the remaining error e3e_{3}. By upscaled problems (17), we have

1Hd[0,H]d(0CDtαe3,Ph,H(V0,V1))+1Hd[0,H]daε(e3,Ph,H(V0,V1))\displaystyle\frac{1}{H^{d}}\int_{[0,H]^{d}}\bigl{(}\,_{0}^{C}D_{t}^{\alpha}e_{3},P_{h,H}(V_{0},V_{1})\,\bigr{)}+\frac{1}{H^{d}}\int_{[0,H]^{d}}a_{\varepsilon}\bigl{(}\,e_{3},P_{h,H}(V_{0},V_{1})\,\bigr{)}
=\displaystyle= 1Hd[0,H]d(0CDtαe2,Ph,H(V0,V1))+1Hd[0,H]daε(e2,Ph,H(V0,V1))\displaystyle\frac{1}{H^{d}}\int_{[0,H]^{d}}\bigl{(}\,_{0}^{C}D_{t}^{\alpha}e_{2},P_{h,H}(V_{0},V_{1})\,\bigr{)}+\frac{1}{H^{d}}\int_{[0,H]^{d}}a_{\varepsilon}\bigl{(}\,e_{2},P_{h,H}(V_{0},V_{1})\,\bigr{)}
+1Hd[0,H]d(0CDtα(Ph(Π0,hu,Π1,hu)Ph,H(U0,h,H,U1,h,H)),Ph,H(V0,V1))\displaystyle+\frac{1}{H^{d}}\int_{[0,H]^{d}}\bigl{(}\,_{0}^{C}D_{t}^{\alpha}(\,P_{h}(\Pi_{0,h}u,\Pi_{1,h}u\,)-P_{h,H}(U_{0,h,H},U_{1,h,H})\,),P_{h,H}(V_{0},V_{1})\,\bigr{)}
+1Hd[0,H]daε((Ph(Π0,hu,Π1,hu)Ph,H(U0,h,H,U1,h,H)),Ph,H(V0,V1)).\displaystyle+\frac{1}{H^{d}}\int_{[0,H]^{d}}a_{\varepsilon}\bigl{(}\,(P_{h}(\,\Pi_{0,h}u,\Pi_{1,h}u\,)-P_{h,H}(U_{0,h,H},U_{1,h,H})),P_{h,H}(V_{0},V_{1})\,\bigr{)}.

According to the definition of downscaling operators (11) and weak form (3), we have

1Hd[0,H]d(0CDtα(Ph(Π0,hu,Π1,hu)Ph,H(U0,h,H,U1,h,H)),Ph,H(V0,V1))\displaystyle\frac{1}{H^{d}}\int_{[0,H]^{d}}\bigl{(}\,_{0}^{C}D_{t}^{\alpha}(\,P_{h}(\Pi_{0,h}u,\Pi_{1,h}u\,)-P_{h,H}(U_{0,h,H},U_{1,h,H})\,),P_{h,H}(V_{0},V_{1})\,\bigr{)}
+1Hd[0,H]daε((Ph(Π0,hu,Π1,hu)Ph,H(U0,h,H,U1,h,H)),Ph,H(V0,V1))\displaystyle+\frac{1}{H^{d}}\int_{[0,H]^{d}}a_{\varepsilon}\bigl{(}\,(P_{h}(\,\Pi_{0,h}u,\Pi_{1,h}u\,)-P_{h,H}(U_{0,h,H},U_{1,h,H})),P_{h,H}(V_{0},V_{1})\,\bigr{)}
=\displaystyle= 1Hd[0,H]d(0CDtα(Ph(Π0,hu,Π1,hu)),(Ph,HPh)(V0,V1))\displaystyle\frac{1}{H^{d}}\int_{[0,H]^{d}}\bigl{(}\,_{0}^{C}D_{t}^{\alpha}(P_{h}(\,\Pi_{0,h}u,\Pi_{1,h}u)\,),(P_{h,H}-P_{h})(V_{0},V_{1})\,\bigr{)}
+1Hd[0,H]daε(Ph(Π0,hu,Π1,hu),(Ph,HPh)(V0,V1))\displaystyle+\frac{1}{H^{d}}\int_{[0,H]^{d}}a_{\varepsilon}\bigl{(}\,P_{h}(\,\Pi_{0,h}u,\Pi_{1,h}u),(P_{h,H}-P_{h})(V_{0},V_{1})\,\bigr{)}
1Hd[0,H]d(f,(Ph,HPh)(V0,V1))\displaystyle-\frac{1}{H^{d}}\int_{[0,H]^{d}}\bigl{(}\,f,(P_{h,H}-P_{h})(V_{0},V_{1})\,\bigr{)}
=\displaystyle= 1Hd[0,H]d(0CDtα(Ph(Π0,hu,Π1,hu)),(Ph,HPh)(V0,V1))\displaystyle\frac{1}{H^{d}}\int_{[0,H]^{d}}\bigl{(}\,_{0}^{C}D_{t}^{\alpha}(P_{h}(\,\Pi_{0,h}u,\Pi_{1,h}u)\,),(P_{h,H}-P_{h})(V_{0},V_{1})\,\bigr{)}
+1Hd[0,H]daε(Ph(Π0,hu,Π1,hu),Ph,H(V0,V1))aε(u,1Hd[0,H]dPh(V0,V1))\displaystyle+\frac{1}{H^{d}}\int_{[0,H]^{d}}a_{\varepsilon}\bigl{(}\,P_{h}(\,\Pi_{0,h}u,\Pi_{1,h}u),P_{h,H}(V_{0},V_{1})\,\bigr{)}\,-a_{\varepsilon}(u,\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h}(V_{0},V_{1}))
(f,1Hd[0,H]d(Ph,HPh)(V0,V1)).\displaystyle-\bigl{(}\,f,\frac{1}{H^{d}}\int_{[0,H]^{d}}(P_{h,H}-P_{h})(V_{0},V_{1})\,\bigr{)}.

By (19)

(0CDtαe3,1Hd[0,H]dPh,H(V0,V1))+aε(e3,1Hd[0,H]dPh,H(V0,V1))\displaystyle\bigl{(}\,_{0}^{C}D_{t}^{\alpha}e_{3},\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h,H}(V_{0},V_{1})\,\bigr{)}+a_{\varepsilon}(e_{3},\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h,H}(V_{0},V_{1}))
=\displaystyle= (0CDtαe1,1Hd[0,H]dPh(V0,V1))\displaystyle-\bigl{(}\,_{0}^{C}D_{t}^{\alpha}e_{1},\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h}(V_{0},V_{1})\,\bigr{)}
=\displaystyle= (0CDtαe1,1Hd[0,H]dPh,H(V0,V1))+(0CDtαe1,1Hd[0,H]d(Ph,HPh)(V0,V1)).\displaystyle-\bigl{(}\,_{0}^{C}D_{t}^{\alpha}e_{1},\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h,H}(V_{0},V_{1})\,\bigr{)}+\bigl{(}\,_{0}^{C}D_{t}^{\alpha}e_{1},\frac{1}{H^{d}}\int_{[0,H]^{d}}(P_{h,H}-P_{h})(V_{0},V_{1})\,\bigr{)}.

Take (V0,V1)=(Π0,hu,Π1,hu)(U0,h,H,U1,h,H)(V_{0},V_{1})=(\,\Pi_{0,h}u,\Pi_{1,h}u\,)-(U_{0,h,H},U_{1,h,H}), then we obtain

(0CDtαe3,e3)+aε(e3,e3)\displaystyle\bigl{(}\,_{0}^{C}D_{t}^{\alpha}e_{3},e_{3}\,\bigr{)}+a_{\varepsilon}(e_{3},e_{3})
=\displaystyle= (0CDtαe1,e3)\displaystyle-\bigl{(}\,_{0}^{C}D_{t}^{\alpha}e_{1},e_{3}\,\bigr{)}
+(0CDtαe1,1Hd[0,H]d(Ph,HPh)((Π0,hu,Π1,hu)(U0,h,H,U1,h,H))).\displaystyle+\bigl{(}\,_{0}^{C}D_{t}^{\alpha}e_{1},\frac{1}{H^{d}}\int_{[0,H]^{d}}(P_{h,H}-P_{h})(\,(\Pi_{0,h}u,\Pi_{1,h}u)-(U_{0,h,H},U_{1,h,H}))\,\bigr{)}.

By the Young’s inequality and Lemma 3.3

120CDtαe32+e3a2\displaystyle\frac{1}{2}\,_{0}^{C}D_{t}^{\alpha}\bigl{\lVert}\,e_{3}\,\bigr{\rVert}^{2}+\bigl{\lVert}\,e_{3}\,\bigr{\rVert}_{a}^{2} 0CDtαe1e3+Clog(1H)H1+s0CDtαe1e3\displaystyle\leq\bigl{\lVert}\,_{0}^{C}D_{t}^{\alpha}e_{1}\,\bigr{\rVert}\cdot\bigl{\lVert}\,e_{3}\,\bigr{\rVert}+C\log(\frac{1}{H})H^{1+s}\bigl{\lVert}\,_{0}^{C}D_{t}^{\alpha}e_{1}\,\bigr{\rVert}\cdot\bigl{\lVert}\,e_{3}\,\bigr{\rVert}
e32+C(log(1H))2H2+2s0CDtαe12.\displaystyle\leq\bigl{\lVert}\,e_{3}\,\bigr{\rVert}^{2}+C\bigl{(}\,\log(\frac{1}{H})\,\bigr{)}^{2}H^{2+2s}\bigl{\lVert}\,_{0}^{C}D_{t}^{\alpha}e_{1}\,\bigr{\rVert}^{2}.

The following inequality can be derived by Poincaré inequality

Dtα0Ce32C(log(1H))2H2+2s0CDtαe12.\,{}_{0}^{C}D_{t}^{\alpha}\bigl{\lVert}\,e_{3}\,\bigr{\rVert}^{2}\leq C\bigl{(}\,\log(\frac{1}{H})\,\bigr{)}^{2}H^{2+2s}\bigl{\lVert}\,_{0}^{C}D_{t}^{\alpha}e_{1}\,\bigr{\rVert}^{2}.

We conduct Riemann-Liouville fractional integration and square both sides and obtain

e3Ch2log(1H)H1+s(0t1(ts)1α0CDtα+2u(t)2ds)12.\bigl{\lVert}\,e_{3}\,\bigr{\rVert}\leq Ch^{2}\log(\frac{1}{H})H^{1+s}\biggl{(}\,\int_{0}^{t}\frac{1}{(t-s)^{1-\alpha}}\bigl{\lVert}\,_{0}^{C}D_{t}^{\alpha+2}u(t)\,\bigr{\rVert}^{2}\,\mathrm{d}s\,\biggr{)}^{\frac{1}{2}}.

This completes the proof.

4 Time discretization

In this section, we focus on the approximation of the time fractional derivative Dtα0CU(t){}^{C}_{0}{D}_{t}^{\alpha}U(t). We divide the time interval [0,T][0,T] into NN equidistant parts and denote tn=nτ,n=0,1,,Nt_{n}=n\tau,\ n=0,1,\cdots,N, where τ=TN\tau=\frac{T}{N} is the time step size. One of the common way to approximate the time fractional derivative Dtα0CU(t){}^{C}_{0}{D}_{t}^{\alpha}U(t) is the L1 scheme [22]

Dtα0CU(tn+1)=1Γ(2α)j=0nU(tj+1)U(tj)τbj,n=0,1,,N,{}^{C}_{0}{D}_{t}^{\alpha}U(t_{n+1})=\frac{1}{\Gamma(2-\alpha)}\sum\limits_{j=0}^{n}\frac{U(t_{j+1})-U(t_{j})}{\tau}b_{j},\ n=0,1,\cdots,N, (22)

where bj=(j+1)1αj1α,j=0,1,,Nb_{j}=(j+1)^{1-\alpha}-j^{1-\alpha},j=0,1,\cdots,N. For all n=0,1,,Nn=0,1,\cdots,N, the full discretized scheme of the matrix form (10) based on the above approximation can be described as

(M+α0A)Un+1=α0fn+1+(1b1)MUn+Mj=1n1(bjbj+1)Unj+bnMU0,(M+\alpha_{0}A)U^{n+1}=\alpha_{0}f^{n+1}+(1-b_{1})MU^{n}+M\sum\limits_{j=1}^{n-1}(b_{j}-b_{j+1})U^{n-j}+b_{n}MU^{0}, (23)

where α0=Γ(2α)τα\alpha_{0}=\Gamma(2-\alpha)\tau^{\alpha} and UnU^{n} is the approximation of U(tn)U(t_{n}).

4.1 Exponential Integrator

Instead of classic L1 scheme, we introduce the exponential integrator approach in this work. Exponential integrator method has shown to be stable for stiff problems [28] and more efficient for semilinear problems. Thus is suitable for our semi-discretized equation (9). We rewrite the semilinear problem (10) as Dtα0CU(t)+KU(t)=F(U(t)){}^{C}_{0}{D}_{t}^{\alpha}U(t)+KU(t)=F(U(t)) where K=M1AK=M^{-1}A and F(U(t))=M1f(t)F(U(t))=M^{-1}f(t), and let f^\widehat{f} be the Laplace transform of a function ff

f^(s)=0f(t)estdt.\widehat{f}(s)=\int_{0}^{\infty}f(t)\mathrm{e}^{-st}\,\mathrm{d}t.

Perform the Laplace transform on the matrix equation, we have

Dtα0CU^(s)+KU^(s)=F(U)^(s).\widehat{{}^{C}_{0}{D}_{t}^{\alpha}U}(s)+K\widehat{U}(s)=\widehat{F(U)}(s).

The Laplace transform of the Caputo fractional derivative can be written as Dtα0CU^(s)=sαU^(s)sα1U^(0)\widehat{{}^{C}_{0}{D}_{t}^{\alpha}U}(s)=s^{\alpha}\widehat{U}(s)-s^{\alpha-1}\widehat{U}(0) [24]. Thus, we have

U^(s)=(sαI+K)1(F(U)^(s)+sα1U^(0)).\widehat{U}(s)=(s^{\alpha}I+K)^{-1}\bigl{(}\,\widehat{F(U)}(s)+s^{\alpha-1}\widehat{U}(0)\,\bigr{)}.

Utilizing the inverse Laplace transform, we obtain

U(t)\displaystyle U(t) =eα,1(t,K)U(0)+0teα,α(tr,K)F(U(r))dr\displaystyle=e_{\alpha,1}(t,K)U(0)+\int_{0}^{t}e_{\alpha,\alpha}(t-r,K)F(U(r))\,\mathrm{d}r (24)
=eα,1(t,K)U(0)+j=0n1tjtj+1eα,α(tr,K)F(U(r))dr,\displaystyle=e_{\alpha,1}(t,K)U(0)+\sum\limits_{j=0}^{n-1}\int_{t_{j}}^{t_{j+1}}e_{\alpha,\alpha}(t-r,K)F(U(r))\,\mathrm{d}r,

where the function eα,β(t,λ)e_{\alpha,\beta}(t,\lambda) denotes the generalization of the inverse of the Laplace transform (sα+λ)1sαβ(s^{\alpha}+\lambda)^{-1}s^{\alpha-\beta} to the matrix argument KK. Note that eα,β(t,λ)e_{\alpha,\beta}(t,\lambda) can be computed by eα,β(t,λ)=tβ1Eα,β(tαλ)e_{\alpha,\beta}(t,\lambda)=t^{\beta-1}E_{\alpha,\beta}(-t^{\alpha}\lambda), where Eα,β(z)E_{\alpha,\beta}(z) is the Mittag-Leffler function [24]

Eα,β(z)=k=0zkΓ(αk+β).E_{\alpha,\beta}(z)=\sum\limits_{k=0}^{\infty}\frac{z^{k}}{\Gamma(\alpha k+\beta)}.

We approximate F(U(r))F(U(r)) by the constant F(Uj)F(U_{j}), then the integral can be approximated by

tjtj+1eα,α(tnr,K)F(U(r))dr\displaystyle\int_{t_{j}}^{t_{j+1}}e_{\alpha,\alpha}(t_{n}-r,K)F(U(r))\,\mathrm{d}r
=\displaystyle= F(Uj)tjteα,α(tnr,K)dr+F(Uj+1)ttj+1eα,α(tnr,K)dr\displaystyle F(U_{j})\int_{t_{j}}^{t}e_{\alpha,\alpha}(t_{n}-r,K)\,\mathrm{d}r+F(U_{j+1})\int_{t}^{t_{j+1}}e_{\alpha,\alpha}(t_{n}-r,K)\,\mathrm{d}r
=\displaystyle= F(Uj)eα,α+1(tntj,K)F(Uj+1)eα,α+1(tntj+1,K).\displaystyle F(U_{j})e_{\alpha,\alpha+1}(t_{n}-t_{j},K)-F(U_{j+1})e_{\alpha,\alpha+1}(t_{n}-t_{j+1},K).

This leads to the numerical scheme

Un=eα,1(tn,K)U0+j=0n1Wn,jFj,U^{n}=e_{\alpha,1}(t_{n},K)U^{0}+\sum\limits_{j=0}^{n-1}W_{n,j}F^{j}, (25)

where the weights Wn,j=eα,α+1(tntj,K)eα,α+1(tntj+1,K)W_{n,j}=e_{\alpha,\alpha+1}(t_{n}-t_{j},K)-e_{\alpha,\alpha+1}(t_{n}-t_{j+1},K) and Fj=F(Uj)F^{j}=F(U^{j}) represent the approximation of the semilinear term F(U(tj))F(U(t_{j})).

To compute the eα,β(t,K)e_{\alpha,\beta}(t,K), we take the eigenvalue of the matrix K=Q1ΛQK=Q^{-1}\Lambda Q for eα,β(t,K)=Qeα,β(t,Λ)Q1e_{\alpha,\beta}(t,K)=Qe_{\alpha,\beta}(t,\Lambda)Q^{-1}, where Λ\Lambda is a diagonal matrix with the characteristic value of KK. Thus, the discretized scheme with exponential integrator (EI) can be described as

Q1Un=eα,1(nτ,Λ)QU0+eα,α+1(nτ,Λ)QF0+j=1n1eα,α+1((nj)τ,Λ)Q(FjFj1).Q^{-1}{U}^{n}=e_{\alpha,1}(n\tau,\Lambda)Q{U}^{0}+e_{\alpha,\alpha+1}(n\tau,\Lambda)Q{F}^{0}+\sum\limits_{j=1}^{n-1}e_{\alpha,\alpha+1}\bigl{(}(n-j)\tau,\Lambda\bigr{)}Q({F}^{j}-{F}^{j-1}). (26)

4.2 Error estimate for the discretized scheme

In this subsection, we discuss the convergence of the numerical scheme (26). The error between the numerical solution and the exact solution mainly comes from the multicontinuum model and the exponential integrator scheme. We have already obtained the former in Section 3 and the latter, i.e.\,\mathrm{i}.\mathrm{e}.\, temporal discretization error, results from the numerical integration. We first give the estimate of the weight Wn,jW_{n,j} of (25) in the following Lemma by the property of Mittag-Leffler function.

Lemma 4.1.

There exists a constant CC^{\prime} only depending on α,K\alpha,\ K such that for any n1n\geq 1 [28]

Wn,jCτα(nj)α1j=1,2,,n1.\bigl{\lVert}\,W_{n,j}\,\bigr{\rVert}\leq C^{\prime}\tau^{\alpha}(n-j)^{\alpha-1}\ \ j=1,2,\cdots,n-1.

Assume the homogenized solution (U0,h,H(x,t),U1,h,H(x,t))\bigl{(}\,U_{0,h,H}(x,t),U_{1,h,H}(x,t)\,\bigr{)} is sufficiently smooth at time tnt_{n}. We could obtain the convergence of the exponential integrator scheme (26) from the following theorem.

Theorem 4.1.

Let u(x,t)u(x,t) be the solution to (1) and (U0n,U1n)(U_{0}^{n},U_{1}^{n}) be the approximation of homogenized solution (U0,h,H(x,tn),U1,h,H(x,tn))(U_{0,h,H}(x,t_{n}),U_{1,h,H}(x,t_{n})). If the number of the oversampling layer kk satisfies 2kh<H=O(kh)2kh<H=O(kh) and let ff satisfies the Lipschitz condition, we have following error estimates

u(x,tn)1Hd[0,H]dPh,H(U0n,U1n)\displaystyle\bigl{\lVert}\,u(x,t_{n})-\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h,H}(U_{0}^{n},U_{1}^{n})\,\bigr{\rVert} (27)
\displaystyle\leq Ch2log(1H)H1+s(f+0t1(ts)1α0CDtα+2u(t)2ds)\displaystyle Ch^{2}\log(\frac{1}{H})H^{1+s}\biggl{(}\,\bigl{\lVert}\,f\,\bigr{\rVert}+\int_{0}^{t}\frac{1}{(t-s)^{1-\alpha}}\bigl{\lVert}\,_{0}^{C}D_{t}^{\alpha+2}u(t)\,\bigr{\rVert}^{2}\,\mathrm{d}s\,\biggr{)}
+C(τ+tnα1τ1+α).\displaystyle+C\bigl{(}\,\tau+t_{n}^{\alpha-1}\tau^{1+\alpha}\,\bigr{)}.
Proof.

Since

u(x,tn)1Hd[0,H]dPh,H(U0n,U1n)\displaystyle\bigl{\lVert}\,u(x,t_{n})-\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h,H}(U_{0}^{n},U_{1}^{n})\,\bigr{\rVert}
\displaystyle\leq u(tn)1Hd[0,H]dPh,H(U0,h,H(tn),U1,h,H(tn))\displaystyle\bigl{\lVert}\,u(t_{n})-\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h,H}(U_{0,h,H}(t_{n}),U_{1,h,H}(t_{n}))\,\bigr{\rVert}
+1Hd[0,H]dPh,H((U0,h,H(tn),U1,h,H(tn))(U0n,U1n)),\displaystyle+\bigl{\lVert}\,\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h,H}\bigl{(}\,(U_{0,h,H}(t_{n}),U_{1,h,H}(t_{n}))-(U_{0}^{n},U_{1}^{n})\,\bigr{)}\,\bigr{\rVert},

and the estimate of the first term has been proved in Theorem 3.1. We denote U(t)=(U0,h,H(t),U1,h,H(t))U(t)=\bigl{(}\,U_{0,h,H}(t),U_{1,h,H}(t)\,\bigr{)} and Un=(U0n,U1n)U^{n}=(U_{0}^{n},U_{1}^{n}). For the second term, by the definition of the NLMC downscaling operator Ph,HP_{h,H}, we derive that

1Hd[0,H]dPh,H((U0,h,H(tn),U1,h,H(tn))(U0n,U1n))\displaystyle\bigl{\lVert}\,\frac{1}{H^{d}}\int_{[0,H]^{d}}P_{h,H}\bigl{(}\,(U_{0,h,H}(t_{n}),U_{1,h,H}(t_{n}))-(U_{0}^{n},U_{1}^{n})\,\bigr{)}\,\,\bigr{\rVert}
\displaystyle\leq C(U0,h,H(tn),U1,h,H(tn))(U0n,U1n)=CU(tn)Un.\displaystyle C\bigl{\lVert}\,(U_{0,h,H}(t_{n}),U_{1,h,H}(t_{n}))-(U_{0}^{n},U_{1}^{n})\,\bigr{\rVert}=C\bigl{\lVert}\,U(t_{n})-U^{n}\,\bigr{\rVert}.

By (24), we have

U(t)=eα,1(t,K)U(0)+j=0n1Wn,jF(U(tj))+En,U(t)=e_{\alpha,1}(t,K)U(0)+\sum\limits_{j=0}^{n-1}W_{n,j}F(U(t_{j}))+E_{n},

where EnE_{n} is the error included by numerical integration

En=j=0n1tjtj+1eα,α(tna;K)(F(U(r))F(U(tj)))dr,E_{n}=\sum\limits_{j=0}^{n-1}\int_{t_{j}}^{t_{j+1}}e_{\alpha,\alpha}(t_{n}-a;K)\bigl{(}\,F(U(r))-F(U(t_{j}))\,\bigr{)}\,\mathrm{d}r,

and

En\displaystyle\bigl{\lVert}\,E_{n}\,\bigr{\rVert} Lcond(Q)j=0n1tjtj+1eα,α(tnr;Λ)U(r)U(tj)dr\displaystyle\leq L\,\mathrm{cond}(Q)\sum\limits_{j=0}^{n-1}\int_{t_{j}}^{t_{j+1}}\bigl{\lVert}\,e_{\alpha,\alpha}(t_{n}-r;\Lambda)\,\bigr{\rVert}\,\cdot\,\bigl{\lVert}\,U(r)-U(t_{j})\,\bigr{\rVert}\,\mathrm{d}r
C1j=0n1tjtj+1(tnr)α1U(r)U(tj)dr,\displaystyle\leq C_{1}\sum\limits_{j=0}^{n-1}\int_{t_{j}}^{t_{j+1}}(t_{n}-r)^{\alpha-1}\bigl{\lVert}\,U(r)-U(t_{j})\,\bigr{\rVert}\,\mathrm{d}r,

where LL represents the Lipschitz constant. Assume the exact solution U(t)U(t) can be expanded in mix powers of integer and fractional order [6] according to U(t)=C0,0+C0,1tα+C0,2t2α+C1,0t+C1,1t1+α+C1,2t1+2α+U(t)=C_{0,0}+C_{0,1}t^{\alpha}+C_{0,2}t^{2\alpha}+C_{1,0}t+C_{1,1}t^{1+\alpha}+C_{1,2}t^{1+2\alpha}+\cdots. This indicates that U(tj+1)U(tj)\bigl{\lVert}\,U(t_{j+1})-U(t_{j})\,\bigr{\rVert} can be bounded by the order τα\tau^{\alpha} in [0,T][0,T]. Then we have

EnC1τ1+αtnα1+C2τ.\bigl{\lVert}\,E_{n}\,\bigr{\rVert}\leq C_{1}\tau^{1+\alpha}t_{n}^{\alpha-1}+C_{2}\tau.

With the help of Lemma 4.1 and the discrete Gronwall inequality [5]

U(tn)Un\displaystyle\bigl{\lVert}\,U(t_{n})-U^{n}\,\bigr{\rVert} En+j=0n1Wn,jF(U(tj))F(Uj)\displaystyle\leq\bigl{\lVert}\,E_{n}\,\bigr{\rVert}+\sum\limits_{j=0}^{n-1}\bigl{\lVert}\,W_{n,j}\,\bigr{\rVert}\,\bigl{\lVert}\,F(U(t_{j}))-F(U^{j})\,\bigr{\rVert}
En+LCταj=0n1(nj)α1U(tj)Uj\displaystyle\leq\bigl{\lVert}\,E_{n}\,\bigr{\rVert}+LC^{\prime}\tau^{\alpha}\sum\limits_{j=0}^{n-1}(n-j)^{\alpha-1}\bigl{\lVert}\,U(t_{j})-U^{j}\,\bigr{\rVert}
CEnC(τ+tnα1τ1+α).\displaystyle\leq C\bigl{\lVert}\,E_{n}\,\bigr{\rVert}\leq C\bigl{(}\,\tau+t_{n}^{\alpha-1}\tau^{1+\alpha}\,\bigr{)}.

Combining all the results above, we give the final estimate (27).

5 Numerical Experiments

In this section, we will show some numerical tests for the time fractional parabolic equation with the multiscale permeability field κ(x)\kappa(x). We take the domain Ω=[0,1]×[0,1]\Omega=[0,1]\times[0,1]. For the reference solution, we utilize the L1 scheme where the time intevral [0,T][0,T] is divided into uniform parts N^=5N\widehat{N}=5N for time discretization, and finite element method with mesh size 1400\frac{1}{400} for spatial discretization. We will discuss the linear problem with source term f(x)f(x) and semilinear source term f(x;u(x))f(x;u(x)) with two different media respectively. In our examples, we take

κ(x)=ε105inΩ0,κ(x)=1100εinΩ1.\kappa(x)=\frac{\varepsilon}{10^{5}}\quad\mathrm{in}\ \Omega_{0},\quad\kappa(x)=\frac{1}{100\varepsilon}\quad\mathrm{in}\ \Omega_{1}.

In all examples, we set the coarse mesh grid size as H=120H=\frac{1}{20}, T=0.001T=0.001 and N=100N=100. The relative errors between the reference solutions and numerical solutions at each time t=tnt=t_{n} are defined as follows

ein=Πi,hu(x,tn)Uin(x)Πi,hu(x,tn),i=0,1e_{i}^{n}=\frac{\bigl{\lVert}\,\Pi_{i,h}u(x,t_{n})-U_{i}^{n}(x)\,\bigr{\rVert}}{\bigl{\lVert}\,\Pi_{i,h}u(x,t_{n})\,\bigr{\rVert}},\quad i=0,1 (28)

where Πi,hu(x,tn)\Pi_{i,h}u(x,t_{n}) is the average of reference solutions u(x,tn)u(x,t_{n}) (solved by finite element method in the fine grid) while UinU_{i}^{n} denotes the approximated solution from our proposed method at tnt_{n} for a specific continua ii. A logarithmic scale has been used for error-axis in all relative error figures.

5.1 Smooth source term

We first take a smooth source term f(x1,x2)=exp(50((x10.5)2+(x20.5)2))f(x_{1},x_{2})=\exp\bigl{(}\,-50((x_{1}-0.5)^{2}+(x_{2}-0.5)^{2})\,\bigr{)} and the initial condition u0(x1,x2)=5×103sin(2πx1)sin(πx2)u_{0}(x_{1},x_{2})=5\times 10^{-3}\sin(2\pi x_{1})\sin(\pi x_{2}). In this example, we set α=0.9,0.6,0.3\alpha=0.9,0.6,0.3 and ε=110\varepsilon=\frac{1}{10}. The permeability field κ\kappa is depicted in the left of Figure 2, which is a periodic field. The blue region represent the low conductivity region Ω0\Omega_{0} while the yellow region represent the high conductivity region Ω1\Omega_{1}. We plot the reference solution snapshots at t=0,t=T2,t=Tt=0,\ t=\frac{T}{2},\ t=T respectively in Figure 3. In Figure 4, we show the upscaled solutions and the corresponding averaged reference solutions for α=0.9\alpha=0.9 at the final time. We note that our proposed approach provides an accurate approximation of the averaged reference solution for different α\alpha. Figure 5 shows the error history in the temporal direction corresponding to different α\alpha. Using same time step size, we compare the approximate solution obtained using explicit L1 and our proposed method. The relative errors at t=Tt=T are shown in Table 1. It can be seen that for both methods, the approximation performs better when α\alpha is closer to 11. However, we observe that the explicit L1 scheme does not converge to the reference solutions when α\alpha goes to 0, while our method still gives good results, which demonstrates the effectiveness and stability of our proposed method.

Refer to caption
Refer to caption
Fig. 2: Left: κ\kappa; Right: ff.
Refer to caption
Fig. 3: The solution snapshots to linear problems with α=0.9\alpha=0.9, Left: t=0t=0; Middle: t=T2t=\frac{T}{2}; Right: t=Tt=T.
Refer to caption
((a))
Refer to caption
((b))
Refer to caption
((c))
Refer to caption
((d))
Refer to caption
((e))
Refer to caption
((f))
Fig. 4: α=0.9\alpha=0.9 and t=Tt=T. Left: average of reference solutions; Middle: numerical solutions by L1 implicit scheme; Right: numerical solutions by EI method. The first row: U0U_{0}, the second row: U1U_{1}.
α\alpha U0U_{0} U1U_{1}
Explicit L1 EI method Explicit L1 EI method
0.9 0.0105 0.0105 0.0083 0.0083
0.6 0.0133 0.0122 0.0115 0.0130
0.3 NaN 0.0462 NaN 0.0460
Table 1: Relative errors at t=Tt=T.
Refer to caption
((a))
Refer to caption
((b))
Refer to caption
((c))
Refer to caption
((d))
Refer to caption
((e))
Refer to caption
((f))
Fig. 5: The first row: relative errors by explicit L1 scheme; The second row: relative errors by EI method. For each column, we take α=0.3,0.6,0.9\alpha=0.3,0.6,0.9 form left to right.

5.2 Semilinear problems

In this numerical test, we consider the semilinear problems and use the same media as in the first experiment. The results of the semilinear problems are also computed with different α\alpha by our proposed method. We take f=uu3+f0(x1,x2)f=u-u^{3}+f_{0}(x_{1},x_{2}) with the initial condition u0(x1,x2)=x1(1x1)x2(1x2)u_{0}(x_{1},x_{2})=-x_{1}(1-x_{1})x_{2}(1-x_{2}), where f0(x1,x2)=exp(50((x10.4)2+(x20.6)2))f_{0}(x_{1},x_{2})=\exp\bigl{(}\,-50\bigl{(}\,(x_{1}-0.4)^{2}+(x_{2}-0.6)^{2}\,\bigr{)}\,\bigr{)} is the smooth source term shown in the right of Figure 9. We present the relative errors for different α=0.9,0.6,0.3\alpha=0.9,0.6,0.3. When α\alpha tends to 0, the explicit L1 scheme turns to be unstable while the EI method performs well. Moreover, for the semilinear problems, the L1 scheme needs iterative methods at each time step while the exponential integrator method does not. The reference solution snapshots are depicted in Figure 6. The Figure 7 illustrates the numerical solutions by L1 scheme and exponential integrator method compared to the average of the reference solutions. We figure out that our proposed method also works well for semilinear diffusion problems. We represent the relative errors in Figure 8 and Table 2 with different α\alpha.

Refer to caption
Fig. 6: The solution snapshots to semilinear problem with α=0.6\alpha=0.6, Left: t=0t=0; Middle: t=T2t=\frac{T}{2}; Right: t=Tt=T.
Refer to caption
((a))
Refer to caption
((b))
Refer to caption
((c))
Refer to caption
((d))
Refer to caption
((e))
Refer to caption
((f))
Fig. 7: α=0.6\alpha=0.6 and t=Tt=T. Left: average of reference solutions; Middle: numerical solutions by L1 implicit scheme; Right: numerical solutions by EI method. The first row: U0U_{0}, the second row: U1U_{1}.
α\alpha U0U_{0} U1U_{1}
Explicit L1 EI method Explicit L1 EI method
0.9 0.0165 0.0046 0.0035 0.0176
0.6 0.2125 0.0070 0.2203 0.0019
0.3 NaN 0.0607 NaN 0.0106
Table 2: Relative errors at t=Tt=T.
Refer to caption
((a))
Refer to caption
((b))
Refer to caption
((c))
Refer to caption
((d))
Refer to caption
((e))
Refer to caption
((f))
Fig. 8: The first row: relative errors by explicit L1 scheme; The second row: relative errors by EI method. For each column, we take α=0.3,0.6,0.9\alpha=0.3,0.6,0.9 form left to right.

5.3 A more complicated media

In our final example, we take nonperiodic media κ\kappa as shown in the left of Figure 9. The other settings are the same as the previous semilinear case. We also represent the reference solution snapshots in Figure 10 with α=0.6\alpha=0.6. By the relative errors in Figure 11 and Table 3, the similar findings has been observed as in the previous cases. The errors of the numerical solutions is slightly larger compared to the previous examples, but they still converge to the reference solution.

Refer to caption
Refer to caption
Fig. 9: Left: κ\kappa; Right: f0f_{0}.
Refer to caption
Fig. 10: The solution snapshots to semi-linear problem with α=0.6\alpha=0.6, Left: t=0t=0; Middle: t=T2t=\frac{T}{2}; Right: t=Tt=T.
α\alpha U0U_{0} U1U_{1}
Explicit L1 EI method Explicit L1 EI method
0.9 0.0201 0.0122 0.0345 0.0308
0.6 0.2133 0.0145 0.2137 0.0374
0.3 NaN 0.1318 NaN 0.1173
Table 3: Relative errors at t=Tt=T.
Refer to caption
((a))
Refer to caption
((b))
Refer to caption
((c))
Refer to caption
((d))
Refer to caption
((e))
Refer to caption
((f))
Fig. 11: The first row: relative errors by explicit L1 scheme; The second row: relative errors by EI method. For each column, we take α=0.3,0.6,0.9\alpha=0.3,0.6,0.9 form left to right.

6 Conclusions

In this paper, we present a time fractional multicontinuum upscaled model for solving the high-contrast fractional diffusion equation. The multicontinuum homogenization method is used to derive a coarse scale model, where the solutions are approximated through averages and accounts for gradient effects within each continuum, while incorporating microscale information in the upscaled quantities. We conduct a semi-discretized error analysis. We demonstrate that the nonlocal multicontinuum downscaling (NLMC) operator can be effectively approximated by the CEM-GMsFEM downscaling operator. Additionally, we utilize an exponential integrator approach for the time fractional derivative term and prove the convergence of the full discrete scheme. Numerical examples show that our method exhibits greater stability compared to the L1 scheme, while maintaining a similar convergence order. We explore various high-contrast heterogeneous media and source terms in our numerical experiments. It is noted that our proposed approach can be extended to scenarios involving two continua under appropriate assumptions. Future research would consist of addressing the singularity of the time fractional derivative and extended current approach to the high-order exponential integrator.

Acknowledgments

Y. Wang’s work is partially supported by the NSFC grant 12301559. W.T. Leung is partially supported by the Hong Kong RGC Early Career Scheme 21307223.

References

  • [1] Yalchin Efendiev and Thomas Y Hou. Multiscale finite element methods: theory and applications, volume 4. Springer Science & Business Media, 2009.
  • [2] Thomas Y Hou and Xiao-Hui Wu. A multiscale finite element method for elliptic problems in composite materials and porous media. Journal of computational physics, 134(1):169–189, 1997.
  • [3] Thomas Hou, Xiao-Hui Wu, and Zhiqiang Cai. Convergence of a multiscale finite element method for elliptic problems with rapidly oscillating coefficients. Mathematics of computation, 68(227):913–943, 1999.
  • [4] Zhiming Chen and Thomas Hou. A mixed multiscale finite element method for elliptic problems with oscillating coefficients. Mathematics of Computation, 72(242):541–576, 2003.
  • [5] Jennifer Dixon. On the order of the error in discretization methods for weakly singular second kind non-smooth solutions. BIT Numerical Mathematics, 25:623–634, 1985.
  • [6] Ch Lubich. Runge-kutta theory for volterra integrodifferential equations. Numerische Mathematik, 40:119–135, 1982.
  • [7] Alain Bourgeat. Homogenized behavior of two-phase flows in naturally fractured reservoirs with uniform fractures distribution. Computer Methods in Applied Mechanics and Engineering, 47(1-2):205–216, 1984.
  • [8] Louis J Durlofsky. Numerical calculation of equivalent grid block permeability tensors for heterogeneous porous media. Water resources research, 27(5):699–708, 1991.
  • [9] Yuguang Chen, Louis J Durlofsky, M Gerritsen, and Xian-Huan Wen. A coupled local–global upscaling approach for simulating flow in highly heterogeneous formations. Advances in water resources, 26(10):1041–1060, 2003.
  • [10] Xiao-Hui Wu, Yalchin Efendiev, and Thomas Y Hou. Analysis of upscaling absolute permeability. Discrete and Continuous Dynamical Systems Series B, 2(2):185–204, 2002.
  • [11] Yalchin Efendiev, Juan Galvis, and Thomas Y Hou. Generalized multiscale finite element methods (gmsfem). Journal of computational physics, 251:116–135, 2013.
  • [12] Yalchin Efendiev, Juan Galvis, and Xiao-Hui Wu. Multiscale finite element methods for high-contrast problems using local spectral basis functions. Journal of Computational Physics, 230(4):937–955, 2011.
  • [13] Eric T Chung, Yalchin Efendiev, and Wing Tat Leung. Residual-driven online generalized multiscale finite element methods. Journal of Computational Physics, 302:176–190, 2015.
  • [14] Axel Målqvist and Daniel Peterseim. Localization of elliptic multiscale problems. Mathematics of Computation, 83(290):2583–2603, 2014.
  • [15] Eric T Chung, Yalchin Efendiev, and Wing Tat Leung. Constraint energy minimizing generalized multiscale finite element method. Computer Methods in Applied Mechanics and Engineering, 339:298–319, 2018.
  • [16] Eric T Chung, Yalchin Efendiev, Wing Tat Leung, Maria Vasilyeva, and Yating Wang. Non-local multi-continua upscaling for flows in heterogeneous fractured media. Journal of Computational Physics, 372:22–34, 2018.
  • [17] Yalchin Efendiev and Wing Tat Leung. Multicontinuum homogenization and its relation to nonlocal multicontinuum theories. Journal of Computational Physics, 474:111761, 2023.
  • [18] Wing Tat Leung. Some convergence analysis for multicontinuum homogenization. arXiv preprint arXiv:2401.12799, 2024.
  • [19] Luis F Contreras, David Pardo, Eduardo Abreu, Judit Muñoz-Matute, Ciro Díaz, and Juan Galvis. An exponential integration generalized multiscale finite element method for parabolic problems. Journal of Computational Physics, 479:112014, 2023.
  • [20] Bangti Jin, Raytcho Lazarov, and Zhi Zhou. Numerical methods for time-fractional evolution equations with nonsmooth data: a concise overview. Computer Methods in Applied Mechanics and Engineering, 346:332–358, 2019.
  • [21] Guanglian Li. Wavelet-based edge multiscale parareal algorithm for subdiffusion equations with heterogeneous coefficients in a large time domain. Journal of Computational and Applied Mathematics, 440:115608, 2024.
  • [22] Yumin Lin and Chuanju Xu. Finite difference/spectral approximations for the time-fractional diffusion equation. Journal of computational physics, 225(2):1533–1552, 2007.
  • [23] Mengnan Li, Eric Chung, and Lijian Jiang. A constraint energy minimizing generalized multiscale finite element method for parabolic equations. Multiscale Modeling & Simulation, 17(3):996–1018, 2019.
  • [24] Igor Podlubny. Fractional differential equations: an introduction to fractional derivatives, fractional differential equations, to methods of their solution and some of their applications. elsevier, 1998.
  • [25] Zhengya Yang, Xuejuan Chen, Yanping Chen, and Jing Wang. Accurate numerical simulations for fractional diffusion equations using spectral deferred correction methods. Computers & Mathematics with Applications, 153:123–129, 2024.
  • [26] Steven M Cox and Paul C Matthews. Exponential time differencing for stiff systems. Journal of Computational Physics, 176(2):430–455, 2002.
  • [27] Roberto Garrappa and Marina Popolizio. Generalized exponential time differencing methods for fractional order problems. Computers & Mathematics with Applications, 62(3):876–890, 2011.
  • [28] Roberto Garrappa. Exponential integrators for time–fractional partial differential equations. The European Physical Journal Special Topics, 222(8):1915–1927, 2013.
  • [29] S Mohammed Hosseini and Zohreh Asgari. Solution of stochastic nonlinear time fractional pdes using polynomial chaos expansion combined with an exponential integrator. Computers & Mathematics with Applications, 73(6):997–1007, 2017.