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

A combined multiscale finite element method based on the LOD technique for the multiscale elliptic problems with singularities

Kuokuo Zhang11footnotemark: 1 dg1721016@smail.nju.edu.cn Weibing Deng222The work of this author was partially supported by the NSF of China grant 12090023 and 12171237, and by the Ministry of Science and Technology of China grant 2020YFA0713800. wbdeng@nju.edu.cn Haijun Wu333The work of this author was partially supported by the NSF of China grants 11071116, 91130004. hjw@nju.edu.cn Department of Mathematics, Nanjing University, Nanjing 210093, People’s Republic of China
Abstract

In this paper, we construct a combined multiscale finite element method (MsFEM) using the Local Orthogonal Decomposition (LOD) technique to solve the multiscale problems which may have singularities in some special portions of the computational domain. For example, in the simulation of steady flow transporting through highly heterogeneous porous media driven by extraction wells, the singularities lie in the near-well regions. The basic idea of the combined method is to utilize the traditional finite element method (FEM) directly on a fine mesh of the problematic part of the domain and using the LOD-based MsFEM on a coarse mesh of the other part. The key point is how to define local correctors for the basis functions of the elements near the coarse and fine mesh interface, which require meticulous treatment. The proposed method takes advantages of the traditional FEM and the LOD-based MsFEM, which uses much less DOFs than the standard FEM and may be more accurate than the LOD-based MsFEM for problems with singularities. The error analysis is carried out for highly varying coefficients, without any assumptions on scale separation or periodicity. Numerical examples with periodic and random highly oscillating coefficients, as well as the multiscale problems on the L-shaped domain, and multiscale problems with high-contrast channels or well-singularities are presented to demonstrate the efficiency and accuracy of the proposed method.

keywords:
Multiscale problems, non-periodic , LOD , well-singularity , high-contrast channel.
MSC:
[2021] 34E13, 65N12 , 65N30
journal: Journal of Computational Physics

1 Introduction

In this paper we consider the elliptic problems with rapidly varying (non-periodic) coefficients, which involve many spatial scales. Such problems are typically referred to as multiscale problems and often arisen in composite materials and flows in porous media. Any meaningful numerical simulation of these problems such as standard finite element method (FEM) has to account for the highly heterogeneous fine-scale structures in the whole computational domain. This means that the underlying computational mesh has to be sufficiently fine and hence requires an enormous computational demand.

In order to overcome this difficulty, many kinds of methods have been developed in recent decades to solve such multiscale problems. Roughly speaking, from the perspective of final approximation solution, these numerical methods can be categorized into two classes. One is to solve the original problem in the constructed coarse-grid multiscale basis function space hence obtains a good approximation of the original-problem solution; the other is to solve a macro model equivalent to the original problem on the coarse grid mesh hence grasps the macro behavior of the multiscale solution. See, for example, the generalized FEM (GFEM) [1, 2, 3], the multiscale FEM (MsFEM) [4, 5, 6], the variational multiscale methods (VMM) or residual-free bubbles method (RFBM) [7, 8, 9, 10], the heterogeneous multiscale methods (HMM) [11, 12, 13], the multiscale finite-volume method [14], the multigrid numerical homogenization techniques [15, 16], the mortar multiscale methods [17, 18], the localized orthogonal decomposition methods (LODM) [19, 20, 21], the equation-free approaches [22, 23], the generalized MsFEM (GMsFEM) [24, 25], the multiscale-spectral GFEM (MS-GEFM) [26, 27], the constraint energy minimizing GMsFEM (CEM-GMsFEM)[28, 29, 30], some numerical homogenization methods or upscaling methods [31, 32, 33, 34, 35], and so on.

Most of the above mentioned multiscale methods consist of two parts, one is the macro solver on coarse mesh such as various finite element or finite volume methods, and the other is the cell problems solving on the coarse grid or oversampling elements. The multiscale algorithm captures the fine-scale information of the solution by solving the cell problems, and then uses the solutions of the cell problems to form an equivalent macro model or a low dimensional multiscale approximation space of the solution. The definition of the cell problem is mainly based on the differential operator of the original multiscale problem, such as the elliptic operator, so that the variability of the multiscale coefficients can be brought into the final solution model through the solution of the cell problems.

In this paper, we are concerned with a special kind of multiscale problems – those with singularities. For example, the one in L-region has singularity near the corner; while the problem with high-contrast channel that connects the boundaries of coarse-grid blocks has singularity at the edge of the channel [36, 37, 38, 32]; furthermore, the problem with steady flow transporting through highly heterogeneous porous media driven by extraction wells has singularity near the well [39]. The traditional multiscale methods on coarse grids may be inefficient when dealing with singularity. This is mainly because the local singularity of the solution is hard to be grasped effectively at the coarse grid level. To solve this kind of singular problems, some numerical methods have been proposed in the literature. See, for instance, the adaptive GMsFEM used to solve the high contrast problem [40, 41], the MsFEM used to solve the high contrast interface and channel problems [42, 36], the complete multiscale coarse grid algorithm by using the Green functions for solving steady flow problem involving well singularities in heterogeneous porous medium [39], the CEM-GMsFEM used to solve the high contrast problem [28, 29], the LODM used to solve high contrast and complex geometric boundary problems [21, 43, 44], the combined MsFEM used to solve high contrast channel and well–singularity problems [45, 46], and some generalized finite element methods and numerical homogenization methods used to solve high contrast problems [16, 26, 27, 31, 32], and so on. Among them, most of the multiscale methods capture the small scale information of the original-problem solution through the solution of the cell problems. Moreover, for the problems with singularities, such as the problem with high-contrast and narrow channels, in order to grasp the singularities, it needs to construct multiscale finite element approximation space via solving special cell problems. For example, the CEM-GMsFEM first needs to construct the auxiliary multiscale functions by solving the local spectral problem. Consequently, the auxiliary function space is constructed by selecting the eigenfunctions corresponding to small eigenvalues, which correspond to high contrast channels. Finally, the online multiscale basis functions are constructed based on constrained energy minimization in the auxiliary function space.

However, it is difficult to define the corresponding subproblems to construct the required approximation space for the problems with source term singularity, such as the porous medium flow problem with well singularities. In [45, 46], the authors combined the standard FEM with the oversampling MsFEM and Petrov-Galerkin MsFEM to solve the multiscale problem with singularity. The standard FEM is used on a fine mesh of the problematic part of the domain and the oversampling MsFEM or Petrov-Galerkin MsFEM is used on a coarse mesh of the other part. The transmission condition on the interface between coarse and fine meshes is dealt with the penalty technique. The proposed methods take the advantages of the standard FEM and the MsFEM, and maintain the accuracy of the two methods. It is shown [45, 46] that the combined multiscale methods can solve the multiscale elliptic problems with fine and long-ranged high contrast channels and the well singularities very efficiently. But, the error analysis of the methods is still based on the classical homogenization theory, which requires the assumption that the diffusion coefficient is periodic. Therefore, how to improve the algorithm so that the optimal error estimate can be obtained for any diffusion coefficient needs further study. We remark that in the past decade, there are many nice multiscale methods dealing with arbitrary oscillating coefficients, such as the LODM, CEM-GMsFEM, MS-GFEM, and some numerical homogenization methods mentioned above [16, 19, 20, 21, 26, 27, 28, 29, 30, 31, 32].

In this paper, we focus on the LODM which was originally introduced in [19] and could be derived from the VMM framework [20, 21]. The orthogonal decomposition method starts from two finite element spaces, a coarse space VHV_{H} and a very high dimensional space VhV_{h} which can approximate the multiscale solution well. Further, the decomposition can be described in three steps: (1) define a quasi-interpolation operator IH:VhVHI_{H}:V_{h}\rightarrow V_{H}, (2) define a high dimensional space of negligible information by the kernel of the operator IHI_{H}, i.e. WhW_{h}:= kern(IHI_{H}), and (3) find the orthogonal complement of WhW_{h} in VhV_{h} with respect to the energy scalar product. With this strategy, it is possible to split the space VhV_{h} into the orthogonal direct sum of a low dimensional multiscale space VHmsV_{H}^{ms} and a high dimensional remainder space WhW_{h}. The multiscale problem is solved in the low dimensional space VHmsV_{H}^{ms} and is therefore cheap. However, the construction of the exact splitting of VhV_{h} = VHmsWhV_{H}^{ms}\oplus W_{h} is unpractical since it needs to define the correction operator in the whole domain which is computationally expensive. We call the method as an ideal one whose solution is referred as ideal solution. To reduce the computational complexity, several localization strategies were proposed and analyzed in [19, 20, 21]. In fact, the computation of the orthogonal decomposition is localized to the patches of the elements, which we introduced as LODM. The reason which makes localization successful is that outside of the support of the coarse finite element basis functions of VHV_{H}, the canonical basis functions of the multiscale space VHmsV_{H}^{ms} have the property of exponential decay. We remark that the LODM often use the fine-scale solution in VhV_{h} for comparison, which is referred to as reference solution.

The essence of the LODM is to construct a low-dimensional solution space (with a locally supported basis functions) that has very accurate approximation properties with respect to the exact solution. So far, the idea of LOD has been generalized to several kinds of discretization techniques such as discontinuous Galerkin [47], Petrov-Galerkin formulations [48] and mesh-free methods [49]. Moreover, the method has been successfully applied to many kinds of problems such as semi-linear elliptic problems [50], eigenvalue problems [51, 52], problems on complicated geometries [43], and so on. We refer the reader to [53, 54] and references therein for more works about LODM. The attractive point of this method is that it does not rely on the classical homogenization theory and does not need the scale separation assumption.

Based on the above observation, we will use the LOD technique to improve the combined MsFEM and make it suitable for general multiscale problems. Note that the traditional FEM has many excellences to deal with the singularities, such as, refining the mesh or enlarging the polynomial order of the finite element space. Thus, in order to take advantages of both methods, we introduce a combined FE and LOD method (FE–LODM) to solve the multiscale problems with singularities. The idea of this approach is to utilize the traditional FEM directly on a fine mesh of the problematic part of the domain and use the LODM on a coarse mesh of the other part. Comparing to the implement of LODM, there are two key issues of the FE–LODM to consider. The first one is how to define the corresponding quasi-interpolation operator in the subdomain using fine mesh. Here we just choose the L2L^{2} projection Πh\Pi_{h}, which has the property that Πhuh=uh\Pi_{h}u_{h}=u_{h} for uhu_{h} belongs to the fine mesh linear FEM space. This property is very important in our later error analysis, which yields a very useful result that the ideal solution is equals to the reference solution in the subdomain using fine mesh. The second one is how to define the correction operators near the interface between the coarse and fine mesh. A delicate treatment should be done for the elements who have an edge or face in the interface of coarse–fine mesh. For the introduced FE-LODM, we carry out a rigorous and careful analysis for the elliptic equation with arbitrary diffusion coefficient to show both the energy and L2L^{2} errors of the method have the optimal convergence rate. The numerical results also show that the proposed FE-LODM is very efficient for multiscale problems with random generated coefficients and singularities.

The rest of this paper is organized as follows. In Section 2, we give the model problem and define a fine-scale reference problem. Section 3 is devoted to deriving the FE-LODM. In Section 4, we present the error analysis of the approach. In Section 5, we provide some numerical results to demonstrate the efficiency of our method. Conclusions are draw in the last section.

Throughout this paper, standard notations for Lebesgue and Sobolev spaces are employed, and CC denotes the generic constant, which depends on neither the mesh size nor the diffusion coefficient. We also use the shorthand notation aba\lesssim b and bab\gtrsim a for the inequality aCba\leq Cb and bCab\geq Ca. In addition, the shorthand notation aba\eqsim b represents that aba\lesssim b and bab\lesssim a.

2 Model problem and reference approximations

In this section, we first present the multiscale model problem, then introduce its interior penalty continuous-discontinuous Galerkin (IPCDG) discretization on fine meshes and discuss the approximation errors of the IPCDG method. The IPCDG solution will be used as a fine-scale reference solution to estimate the error of the FE-LODM. Note that the IPCDG method was first introduced in [55] for the Helmholtz equation.

2.1 Model Problem

In this paper, we consider a second order elliptic problem with highly varying diffusion coefficient. Let Ωd\Omega\subset\mathbb{R}^{d}, dd=2,3, be a polygonal/polyhedral domain, and the elliptic equation reads as

{(Au)=finΩ,u=0onΩ,\left\{\begin{aligned} &-\nabla\cdot(A\nabla u)=f\quad\mathrm{in}\,\,\Omega,\\ &\qquad\qquad\quad u=0\quad\mathrm{on}\,\,\partial\Omega,\end{aligned}\right. (2.1)

where we assume that fL2(Ω)f\in L^{2}(\Omega), and the diffusion matrix AL(Ω,d×d)A\in L^{\infty}(\Omega,{\mathbb{R}^{d\times d}}) is a symmetric matrix with uniform spectral bounds βα>0\beta\geq\alpha>0, i.e.

σ(A(x))[α,β]xΩ.\displaystyle\sigma(A(x))\subset[\alpha,\beta]\quad\forall\,x\,\in\,\Omega. (2.2)

The weak formulation of problem (2.1) is to find uH01(Ω)u\in H^{1}_{0}(\Omega) such that

ΩAuvdx=ΩfvdxvH01(Ω).\displaystyle\int_{\Omega}A\nabla u\cdot\nabla v\,\mathrm{d}x=\int_{\Omega}fv\,\mathrm{d}x\quad\forall\,v\in\,H^{1}_{0}(\Omega). (2.3)

Clearly, the Lax-Milgram lemma [56] implies that (2.3) has a unique solution.

In order to deal with the multiscale problem that has singularities, we decompose the research domain Ω\Omega into two parts, Ω1\Omega_{1} and Ω2\Omega_{2}, where Ω1\Omega_{1} consists of some subdomain(s) containing the singularities and Ω2=Ω\Ω¯1\Omega_{2}=\Omega\backslash\overline{\Omega}_{1} (see Figure 1 for an illustration). Let Γ=Ω1Ω2\Gamma\,=\partial\Omega_{1}\cap\partial\Omega_{2} be the interface between Ω1\Omega_{1} and Ω2\Omega_{2}. We assume that the length/area of Γ\Gamma satisfies |Γ|=O(1)|\Gamma|=O(1), and Γ\Gamma is Lipschitz continuous.

Refer to caption

Refer to caption

Figure 1: A decomposition of Ω\Omega into Ω1\Omega_{1} with singularities and Ω2=Ω\Ω1\Omega_{2}=\Omega\backslash\Omega_{1}, where the green lines represent the interface Γ\Gamma. Left: A fine-scale mesh for the reference problem; Right: A mesh for the combined multiscale methods.

For any subdomain ωΩ\omega\subseteq\Omega, we denote by (u,v)ω=ωuv(u,v)_{\omega}=\int_{\omega}uv. For any segment/patch γΓ\gamma\subseteq\Gamma, denote by u,vγ\left\langle u,v\right\rangle_{\gamma} =γuv=\int_{\gamma}uv. For brevity, let (u,v)=(u,v)Ω(u,v)=(u,v)_{\Omega}, (u,v)=(u,v)Ω(\nabla u,\nabla v)=(\nabla u,\nabla v)_{\Omega} and u,v=u,vΓ\left\langle u,v\right\rangle=\left\langle u,v\right\rangle_{\Gamma}.

2.2 Reference problem

In this subsection, we introduce the IPCDG method which will be used as the fine-scale reference problem to estimate the error of our FE-LODM.

Let h,Ω1\mathcal{M}_{h,\Omega_{1}} and h,Ω2\mathcal{M}_{h,\Omega_{2}} be regular and quasi-uniform triangulations of Ω1\Omega_{1} and Ω2\Omega_{2}, respectively. Denote by h,h:=h,Ω1h,Ω2\mathcal{M}_{h,h}:=\mathcal{M}_{h,\Omega_{1}}\cup\mathcal{M}_{h,\Omega_{2}} the resulted triangulation of Ω\Omega. Note that any element in the triangulation is considered closed by convention. For any Th,hT\in\mathcal{M}_{h,h}, let hT:=diamTh_{T}:={\rm diam}\,T. Denote by h:=maxTh,hhTh:=\max_{T\in\mathcal{M}_{h,h}}h_{T}. Let 𝐧\mathbf{n} be the unit normal vector of Γ\Gamma that points from Ω1\Omega_{1} to Ω2\Omega_{2}. We define the jump and average of a function vv across Γ\Gamma by [v]:=v|Ω1v|Ω2[v]:=v|_{\Omega_{1}}-v|_{\Omega_{2}} and {v}:=(v|Ω1+v|Ω2)/2\{v\}:=(v|_{\Omega_{1}}+v|_{\Omega_{2}})/2, respectively. Moreover, we denote by h\nabla_{h} the piecewise gradient on Ω1Ω2\Omega_{1}\cup\Omega_{2}, that is, hv|Ωi=(v|Ωi),i=1,2\nabla_{h}v|_{\Omega_{i}}=\nabla(v|_{\Omega_{i}}),i=1,2.

Denote by Γi=ΩΩi\Gamma_{i}=\partial\Omega\cap\partial\Omega_{i} and

HΓi1(Ωi):={vH1(Ωi):v|Γi=0 in the sense of trace},i=1,2.\displaystyle H_{\Gamma_{i}}^{1}(\Omega_{i}):=\left\{v\in H^{1}(\Omega_{i}):\;v|_{\Gamma_{i}}=0\text{ in the sense of trace}\right\},i=1,2.

Let Vh,ΩiV_{h,\Omega_{i}} be the continuous linear Lagrange finite element space on h,Ωi,i=1,2\mathcal{M}_{h,\Omega_{i}},i=1,2, respectively, i.e.

Vh,Ωi:={vhHΓi1(Ωi):vh|TP1,Th,Ωi},\displaystyle V_{h,\Omega_{i}}:=\left\{v_{h}\in H_{\Gamma_{i}}^{1}(\Omega_{i}):\;v_{h}|_{T}\in P_{1},\;\forall\,T\in\mathcal{M}_{h,\Omega_{i}}\right\},

where P1P_{1} is the set of polynomials with total degree 1\leq 1. Then the approximation space of the fine-scale reference problem is defined by

Vh,h:={vh:vh|ΩiVh,Ωi,i=1,2}.\displaystyle V_{h,h}:=\left\{v_{h}:\;v_{h}|_{\Omega_{i}}\in V_{h,\Omega_{i}},\;i=1,2\right\}. (2.4)

Note that a discrete function in Vh,hV_{h,h} is continuous on each Ωi,i=1,2\Omega_{i},i=1,2, but may be discontinuous across the interface Γ\Gamma. Given some positive penalty parameters γ0>0\gamma_{0}>0, for any subset ωΩ\omega\subseteq\Omega, we define a symmetric bilinear form aω(,)a_{\omega}(\cdot,\cdot) as follows:

aω(u,v):=\displaystyle a_{\omega}(u,v):= (Ahu,hv)ω({Ahu𝐧},[v]Γω+[u],{Ahv𝐧}Γω)\displaystyle(A\nabla_{h}u,\nabla_{h}v)_{\omega}-\big{(}\left\langle\{A\nabla_{h}u\cdot\mathbf{n}\},[v]\right\rangle_{\Gamma\cap\omega}+\left\langle[u],\{A\nabla_{h}v\cdot\mathbf{n}\}\right\rangle_{\Gamma\cap\omega}\big{)}
+Jω(u,v),\displaystyle+J_{\omega}(u,v), (2.5)
Jω(u,v):=\displaystyle J_{\omega}(u,v):= γ0h[u],[v]Γω.\displaystyle\left\langle\frac{\gamma_{0}}{h}[u],[v]\right\rangle_{\Gamma\cap\omega}. (2.6)

Then the IPCDG method (cf. [55]) reads as: find uh,hVh,hu_{h,h}\in V_{h,h}, such that

aΩ(uh,h,vh,h)=(f,vh,h)vh,hVh,h.a_{\Omega}(u_{h,h},v_{h,h})=(f,v_{h,h})\quad\forall\,v_{h,h}\in V_{h,h}. (2.7)
Remark 1.

(1) Noticed that uh,hu_{h,h} is continuous in Ω2\Omega_{2} and subdomain(s) in Ω1\Omega_{1} and that the discontinuities across the interface Γ\Gamma is treated by the interior penalty technique from the IPDG methods [57], so we call this method (2.7) the IPCDG method.

(2) It is easy to verify that the IPCDG method is consistent with the multiscale problem (2.1), that is

aΩ(uuh,h,vh,h)=0vh,hVh,h.\displaystyle a_{\Omega}(u-u_{h,h},v_{h,h})=0\quad\forall\,v_{h,h}\in V_{h,h}.

Introduce the discrete energy norm

|v|h,h=(A12hv0,Ω2+γ0h[v]Γ2+hγ0{Ahv𝐧}Γ2)12.\displaystyle\left\|\hskip-0.8pt\left|v\right|\hskip-0.8pt\right\|_{h,h}=\Big{(}\|A^{\frac{1}{2}}\nabla_{h}v\|^{2}_{0,\Omega}+\frac{\gamma_{0}}{h}\|[v]\|^{2}_{\Gamma}+\frac{h}{\gamma_{0}}\|\{A\nabla_{h}v\cdot\mathbf{n}\}\|_{\Gamma}^{2}\Big{)}^{\frac{1}{2}}.

Clearly, the bilinear form aΩa_{\Omega} is continuous on V×VV\times V where V:={v:v|ΩiH2(Ωi)HΓi1(Ωi),i=1,2}V:=\{v:\;v|_{\Omega_{i}}\in H^{2}(\Omega_{i})\cap H^{1}_{\Gamma_{i}}(\Omega_{i}),i=1,2\}, i.e.

|aΩ(u,v)||u|h,h|v|h,h.\displaystyle|a_{\Omega}(u,v)|\lesssim\left\|\hskip-0.8pt\left|u\right|\hskip-0.8pt\right\|_{h,h}\left\|\hskip-0.8pt\left|v\right|\hskip-0.8pt\right\|_{h,h}. (2.8)

Following [55, 57], it may be proved that there exists a positive constant α0\alpha_{0} such that

aΩ(vh,h,vh,h)|vh,h|h,hvh,hVh,h,if γ0α0,\displaystyle a_{\Omega}(v_{h,h},v_{h,h})\gtrsim\left\|\hskip-0.8pt\left|v_{h,h}\right|\hskip-0.8pt\right\|_{h,h}\quad v_{h,h}\in V_{h,h},\quad\text{if }\gamma_{0}\geq\alpha_{0}, (2.9)

and hence the following Céa’s lemma and the well-posedness hold for the IPCDG method (2.7) if the penalty parameter γ0α0\gamma_{0}\geq\alpha_{0}. We omitted the details.

Lemma 2.1.

Suppose γ0α0\gamma_{0}\geq\alpha_{0}. Then the following error estimate holds:

|uuh,h|h,hinfvh,hVh,h|uvh,h|h,h.\displaystyle\left\|\hskip-0.8pt\left|u-u_{h,h}\right|\hskip-0.8pt\right\|_{h,h}\lesssim\inf_{v_{h,h}\in V_{h,h}}\left\|\hskip-0.8pt\left|u-v_{h,h}\right|\hskip-0.8pt\right\|_{h,h}.

Then the error estimate of the IPCDG method may be obtained by combining the above Céa’s lemma and the interpolation error estimates. We omitted the details and just assume that the IPCDG solution uh,hu_{h,h} is a good approximation of the exact solution uu. We will use uh,hu_{h,h} as a reference solution to estimate the error of our FE-LODM.

For further error analysis, we introduce the following norm on the restriction of the space Vh,hV_{h,h} onto a subdomain ωΩ\omega\subseteq\Omega:

vh,h,ω=(A12hv0,ω2+γ0h[v]Γω2)12\|v\|_{h,h,\omega}=\Big{(}\|A^{\frac{1}{2}}\nabla_{h}v\|^{2}_{0,\omega}+\frac{\gamma_{0}}{h}\|[v]\|^{2}_{\Gamma\cap\omega}\Big{)}^{\frac{1}{2}} (2.10)

and denote by h,h:=h,h,Ω\|\cdot\|_{h,h}:=\|\cdot\|_{h,h,\Omega} the norm on the whole domain Ω\Omega. Noting that the norm h,h\|\cdot\|_{h,h} is just the norm ||h,h\left\|\hskip-0.8pt\left|\cdot\right|\hskip-0.8pt\right\|_{h,h} with the third term dropped, by using the trace and inverse inequalities [56], it is easy to show that the two norms are equivalent on the fine-scale approximation space Vh,hV_{h,h}.

3 FE-LODM formulation

In this section, we will present the FE-LODM which uses FEM in the domain Ω1\Omega_{1} containing singularities and LODM in Ω2\Omega_{2} where the solution is smooth but highly oscillating and the two methods are joint at the interface Γ\Gamma by using the interior penalty technique. To do this, we first introduce coarse meshes on Ω2\Omega_{2} and coarse-scale finite element spaces, secondly state the multiscale decomposition of the fine-scale space Vh,hV_{h,h} and the approximation space for the FE-LODM, then present the ideal combined multiscale method, and finally formulate the localized combined multiscale method, i.e., FE-LODM.

3.1 Coarse-scale FE spaces

Let H,Ω2\mathcal{M}_{H,\Omega_{2}} a shape-regular coarse triangulation of Ω2\Omega_{2} such that the fine reference mesh h,Ω2\mathcal{M}_{h,\Omega_{2}} is a refinement of it. Denote by HH the maximum diameter of elements in H,Ω2\mathcal{M}_{H,\Omega_{2}}. Clearly, h<Hh<H. Let Γh\mathcal{M}_{\Gamma_{h}} and ΓH\mathcal{M}_{\Gamma_{H}} be the set of interface elements in h,Ω1\mathcal{M}_{h,\Omega_{1}} and H,Ω2\mathcal{M}_{H,\Omega_{2}}, respectively, i.e.

Γh:={Th,Ω1:|TΓ|0}andΓH:={TH,Ω2:|TΓ|0}.\displaystyle\mathcal{M}_{\Gamma_{h}}:=\left\{T\in\mathcal{M}_{h,\Omega_{1}}:\;|T\cap\Gamma|\neq 0\right\}\;\text{and}\;\mathcal{M}_{\Gamma_{H}}:=\left\{T\in\mathcal{M}_{H,\Omega_{2}}:\;|T\cap\Gamma|\neq 0\right\}.

Denote by Γh:={TΓ:TΓh}\Gamma_{h}:=\left\{T\cap\Gamma:\;T\in\mathcal{M}_{\Gamma_{h}}\right\} and ΓH:={TΓ:TΓH}\Gamma_{H}:=\left\{T\cap\Gamma:\;T\in\mathcal{M}_{\Gamma_{H}}\right\} the two partitions of the interface Γ\Gamma induced by h,Ω1\mathcal{M}_{h,\Omega_{1}} and H,Ω2\mathcal{M}_{H,\Omega_{2}}, respectively. In addition, we assume that h,Ω1\mathcal{M}_{h,\Omega_{1}} and H,Ω2\mathcal{M}_{H,\Omega_{2}} satisfy the matching condition that Γh\Gamma_{h} is a refinement of ΓH\Gamma_{H}. Introduce the coarse-scale finite element space on the coarse mesh H,Ω2\mathcal{M}_{H,\Omega_{2}}:

VH,Ω2:={vHHΓ21(Ω2):vH|TP1,TH,Ω2}.\displaystyle V_{H,\Omega_{2}}:=\left\{v_{H}\in H_{\Gamma_{2}}^{1}(\Omega_{2}):\;v_{H}|_{T}\in P_{1},\;\forall\,T\in\mathcal{M}_{H,\Omega_{2}}\right\}.

Let

Vh,H:={vh,H:vh,H|Ω1Vh,Ω1,vh,H|Ω2VH,Ω2}.\displaystyle V_{h,H}:=\left\{v_{h,H}:\;v_{h,H}|_{\Omega_{1}}\in V_{h,\Omega_{1}},\;v_{h,H}|_{\Omega_{2}}\in V_{H,\Omega_{2}}\right\}.

Moreover, we denote by

V0,h:={vh,hVh,h:vh,h|Ω1=0}andV0,H:={vh,HVh,H:vh,H|Ω1=0}.\displaystyle V_{0,h}:=\{v_{h,h}\in V_{h,h}:\;v_{h,h}|_{\Omega_{1}}=0\}\;\text{and}\;V_{0,H}:=\{v_{h,H}\in V_{h,H}:\;v_{h,H}|_{\Omega_{1}}=0\}.

3.2 Multiscale Decomposition

First, we need to define a quasi-interpolation operator from the fine-scale approximation space to the coarse-scale space. For this, we first introduce a weighted Clément-type quasi-interpolation operator 𝒞H\mathcal{C}_{H} defined on the region Ω2\Omega_{2} (see [58, 59]). Let 𝒩H\mathcal{N}_{H} be the set of vertices of elements in H,Ω2\mathcal{M}_{H,\Omega_{2}} and let 𝒩̊H:=𝒩H\Γ2\mathring{\mathcal{N}}_{H}:=\mathcal{N}_{H}\backslash\Gamma_{2}. For any node z𝒩Hz\in\mathcal{N}_{H}, let Φz\Phi_{z} \in VH,Ω2V_{H,\Omega_{2}} be the nodal basis function at zz. The Clément-type quasi-interpolation operator 𝒞H\mathcal{C}_{H}: HΓ21(Ω2)H_{\Gamma_{2}}^{1}(\Omega_{2}) \mapsto VH,Ω2V_{H,\Omega_{2}} is given by:

𝒞Hu:=z𝒩̊HuzΦzwithuz=(u,Φz)Ω2(1,Φz)Ω2uHΓ21(Ω2).\mathcal{C}_{H}u:=\sum\limits_{{z}\in\mathring{\mathcal{N}}_{H}}u_{z}\Phi_{z}\quad with\,\,u_{z}=\frac{(u,\Phi_{z})_{\Omega_{2}}}{(1,\Phi_{z})_{\Omega_{2}}}\qquad\forall\,u\,\in H_{\Gamma_{2}}^{1}(\Omega_{2}). (3.1)

Further, let Πh:L2(Ω1)Vh,Ω1\Pi_{h}:L^{2}(\Omega_{1})\mapsto V_{h,\Omega_{1}} be the L2L^{2}-projection operator. Clearly,

Πhvh=vhvhVh,Ω1.\Pi_{h}v_{h}=v_{h}\qquad\forall\,v_{h}\,\in V_{h,\Omega_{1}}. (3.2)

Then the quasi-interpolation operator 𝒞h,H\mathcal{C}_{h,H}:Vh,hV_{h,h} \mapsto Vh,HV_{h,H} can be defined as

{𝒞h,Hvh,h|Ω1:=Πh(vh,h|Ω1);𝒞h,Hvh,h|Ω2:=𝒞H(vh,h|Ω2),for any vh,hVh,h.\begin{cases}\mathcal{C}_{h,H}v_{h,h}|_{\Omega_{1}}:=\Pi_{h}(v_{h,h}|_{\Omega_{1}});\\ \mathcal{C}_{h,H}v_{h,h}|_{\Omega_{2}}:=\mathcal{C}_{H}(v_{h,h}|_{\Omega_{2}}),\end{cases}\quad\text{for~{}any~{}}v_{h,h}\in V_{h,h}. (3.3)

By the operator 𝒞h,H\mathcal{C}_{h,H}, we can define its kernel space Wh,h:={vh,hVh,h𝒞h,Hvh,h=0}W_{h,h}:=\{v_{h,h}\in V_{h,h}\mid\mathcal{C}_{h,H}v_{h,h}=0\}, and use it to construct a splitting of the space Vh,hV_{h,h} into the direct sum

Vh,h=Vh,HWh,h.V_{h,h}=V_{h,H}\oplus W_{h,h}. (3.4)

Notice that, for any wh,hw_{h,h} \in Wh,hW_{h,h}, from (3.2) it follows that wh,h|Ω1=0w_{h,h}|_{\Omega_{1}}=0. Hence, in the following, we change the notation Wh,hW_{h,h} into W0,hW_{0,h} for emphasis.

Note that the subspace W0,hW_{0,h} is a fine-scale remainder space (high-dimensional space), which contains the fine-scale features of V0,hV_{0,h} that can not be expressed in the low-dimensional space V0,HV_{0,H}. Following the idea of LODM ([21, 50]), we look for a splitting Vh,h=Vh,HmsW0,hV_{h,h}=V^{ms}_{h,H}\oplus W_{0,h} such that the space Vh,HmsV^{ms}_{h,H} has good H1H^{1} approximation properties to the solution of the multiscale problem. It is obvious that Vh,HmsV^{ms}_{h,H} is a low-dimensional space that it has the same dimension as Vh,HV_{h,H}. In order to explicitly construct such a splitting, we look for the orthogonal complement of W0,hW_{0,h} in Vh,hV_{h,h} with respect to the scalar product aΩ(,)a_{\Omega}{(\cdot,\cdot)}.

The corresponding fine-scale projection QhQ_{h}: Vh,hW0,hV_{h,h}\mapsto W_{0,h} is given by: for vh,hv_{h,h} \in Vh,hV_{h,h}, find Qhvh,h{Q_{h}v_{h,h}} \in W0,hW_{0,h}, such that

aΩ(Qhvh,h,w0,h)=aΩ(vh,h,w0,h)w0,hW0,h.a_{\Omega}({Q_{h}v_{h,h}},w_{0,h})=a_{\Omega}(v_{h,h},w_{0,h})\quad\forall\,w_{0,h}\in W_{0,h}. (3.5)

Using the fine-scale projection, we can define the approximation space on the whole domain Ω\Omega for the ideal combined multiscale method by

Vh,Hms:=(IQh)Vh,H.\displaystyle V^{ms}_{h,H}:=({I}-Q_{h})V_{h,H}. (3.6)

3.3 The ideal combined multiscale method

Next, we define the ideal combined multiscale method for the problem (2.7) as follows: for all vh,HmsVh,Hmsv^{ms}_{h,H}\in V^{ms}_{h,H}, find uh,HmsVh,Hmsu^{ms}_{h,H}\in V^{ms}_{h,H}, such that

aΩ(uh,Hms,vh,Hms)=(f,vh,Hms).a_{\Omega}(u^{ms}_{h,H},v^{ms}_{h,H})=(f,v^{ms}_{h,H}). (3.7)

With above definition of the ideal method, we can see that, by taking vh,Hms=uh,Hmsv^{ms}_{h,H}=u^{ms}_{h,H} in (3.7) and using the coercivity (2.9) of aΩa_{\Omega} on Vh,hV_{h,h}, we have the following stability for the ideal solution uh,Hmsu^{ms}_{h,H}: if γ0α0\gamma_{0}\geq\alpha_{0},

uh,Hmsh,hfL2(Ω).\displaystyle\|u^{ms}_{h,H}\|_{h,h}\lesssim\|f\|_{L^{2}(\Omega)}. (3.8)
Remark 2.

It can be proved that 𝒞h,H\mathcal{C}_{h,H} is an isomorphism on Vh,HV_{h,H} (see Lemma 4.4). Thus we can split uh,Hmsu^{ms}_{h,H} into

uh,Hms=(𝒞h,H|Vh,H)1𝒞h,Huh,HmsVh,H((𝒞h,H|Vh,H)1𝒞h,Huh,Hmsuh,Hms)W0,h.u^{ms}_{h,H}=\underbrace{(\mathcal{C}_{h,H}|_{V_{h,H}})^{-1}\mathcal{C}_{h,H}u_{h,H}^{ms}}_{\in V_{h,H}}-\underbrace{\left((\mathcal{C}_{h,H}|_{V_{h,H}})^{-1}\mathcal{C}_{h,H}u_{h,H}^{ms}-u^{ms}_{h,H}\right)}_{\in W_{0,h}}.

Moreover, it is easy to check that

(𝒞h,H|Vh,H)1𝒞h,Huh,Hmsuh,Hms=Qh((𝒞h,H|Vh,H)1𝒞h,Huh,Hms).(\mathcal{C}_{h,H}|_{V_{h,H}})^{-1}\mathcal{C}_{h,H}u_{h,H}^{ms}-u^{ms}_{h,H}=Q_{h}((\mathcal{C}_{h,H}|_{V_{h,H}})^{-1}\mathcal{C}_{h,H}u_{h,H}^{ms}).

Therefore, we have the splitting that obeys (3.6)

uh,Hms=uh,HQh(uh,H),where uh,H:=(𝒞h,H|Vh,H)1𝒞h,Huh,Hms.\displaystyle u^{ms}_{h,H}=u_{h,H}-Q_{h}(u_{h,H}),\quad\text{where }u_{h,H}:=(\mathcal{C}_{h,H}|_{V_{h,H}})^{-1}\mathcal{C}_{h,H}u_{h,H}^{ms}. (3.9)

It is clear that 𝒞h,Huh,Hms=𝒞h,Huh,H\mathcal{C}_{h,H}u^{ms}_{h,H}=\mathcal{C}_{h,H}u_{h,H}.

Before closing this subsection, we present an interesting result in the following proposition about the ideal method (3.7) which says that there is no difference between the ideal solution and the reference solution in the subdomain Ω1\Omega_{1}.

Proposition 1.

Let uh,hu_{h,h} and uh,Hmsu_{h,H}^{ms} be the solutions to (2.7) and (3.7), respectively. Then we have

uh,huh,HmsW0,hand𝒞h,Huh,h=𝒞h,Huh,H,u_{h,h}-u_{h,H}^{ms}\in W_{0,h}\quad\text{and}\quad\mathcal{C}_{h,H}u_{h,h}=\mathcal{C}_{h,H}u_{h,H}, (3.10)

where uh,Hu_{h,H} is defined in Remark 2. Especially, uh,h|Ω1=uh,Hms|Ω1u_{h,h}|_{\Omega_{1}}=u_{h,H}^{ms}|_{\Omega_{1}}.

Proof.

Form (2.7) and (3.7), for all vh,HmsVh,Hmsv_{h,H}^{ms}\in V_{h,H}^{ms}, it follows that

aΩ(uh,h,vh,Hms)=(f,vh,Hms),\displaystyle a_{\Omega}(u_{h,h},v_{h,H}^{ms})=(f,v_{h,H}^{ms}),
aΩ(uh,Hms,vh,Hms)=(f,vh,Hms).\displaystyle a_{\Omega}(u_{h,H}^{ms},v_{h,H}^{ms})=(f,v_{h,H}^{ms}).

Thus

aΩ(uh,huh,Hms,vh,Hms)=0,a_{\Omega}(u_{h,h}-u_{h,H}^{ms},v_{h,H}^{ms})=0, (3.11)

which means uh,huh,HmsW0,hu_{h,h}-u_{h,H}^{ms}\in W_{0,h}. Hence

𝒞h,Huh,h=𝒞h,Huh,Hms=𝒞h,Huh,H,\mathcal{C}_{h,H}u_{h,h}=\mathcal{C}_{h,H}u_{h,H}^{ms}=\mathcal{C}_{h,H}u_{h,H},

which yields the results immediately. ∎

3.4 Formulation of FE-LODM

Note that the fine-scale projection QhQ_{h} in (3.5) is defined globally onto W0,hW_{0,h}, and consequently, in order to calculating the basis functions of the discrete space Vh,HmsV^{ms}_{h,H}, one has to solve many large equations with dim(W0,h){\rm dim}(W_{0,h}) unknowns. Therefore, for practical application, we have to localize the definition of QhQ_{h} to obtain an approximation of the ideal combined multiscale method, i.e., FE-LODM. To do so, we first decompose QhQ_{h} by restrict (3.5) to each elements in h,H\mathcal{M}_{h,H} and then localize the restrictions.

For each TH,Ω2T\in\mathcal{M}_{H,\Omega_{2}} we associate it with a point set T~T\tilde{T}\supseteq T defined as follows. If TH,Ω2\ΓHT\in\mathcal{M}_{H,\Omega_{2}}\backslash\mathcal{M}_{\Gamma_{H}}, we just let T~=T\tilde{T}=T. While, for TΓHT\in\mathcal{M}_{\Gamma_{H}}, we let

T~:=T{tΓh:(tΓ)(TΓ)}\tilde{T}:=T\cup\left\{t\in\mathcal{M}_{\Gamma_{h}}:\;(t\cap\Gamma)\subset(T\cap\Gamma)\right\}

be the union of TT and those interface elements in Γh\mathcal{M}_{\Gamma_{h}} whose intersections with the interface Γ\Gamma are contained in T\partial T (see Figure 2 for an illustration).

Refer to caption

Figure 2: An illustration of the interface combined element T~\tilde{T} for TΓHT\in\mathcal{M}_{\Gamma_{H}}.

Then we define the restrictions of QhQ_{h} for any TH,Ω2T\in\mathcal{M}_{H,\Omega_{2}} as: QhTvh,h{Q_{h}^{T}}v_{h,h} \in W0,hW_{0,h} such that

aΩ(QhTvh,h,w0,h)=aT~(vh,h,w0,h)w0,hW0,h.a_{\Omega}(Q_{h}^{T}v_{h,h},w_{0,h})=a_{\tilde{T}}(v_{h,h},w_{0,h})\quad\forall\,w_{0,h}\in W_{0,h}. (3.12)
Remark 3.

For vh,hVh,h,w0,hW0,hv_{h,h}\in V_{h,h},w_{0,h}\in W_{0,h}, aΩ(vh,h,w0,h)a_{\Omega}(v_{h,h},w_{0,h}) has nonzero terms on the interface Γh\Gamma_{h}. We integrate them into the definition of QhTvh,h{Q_{h}^{T}}v_{h,h} for the elements TΓHT\in\mathcal{M}_{\Gamma_{H}} by restricting the bilinear form aa onto T~\tilde{T}.

Noting that any function in W0,hW_{0,h} vanishes in Ω1\Omega_{1}, from (2.5) we have

aΩ1\(TΓhT)(vh,h,w0,h)=0vh,hVh,h,w0,hW0,h.\displaystyle a_{\Omega_{1}\backslash(\cup_{T\in\mathcal{M}_{\Gamma_{h}}}T)}(v_{h,h},w_{0,h})=0\quad\forall\,v_{h,h}\in V_{h,h},w_{0,h}\in W_{0,h}.

Therefore it follows from (3.5) that

aΩ(Qhvh,h,w0,h)=TH,Ω2aT~(vh,h,w0,h)w0,hW0,h,\displaystyle a_{\Omega}(Q_{h}v_{h,h},w_{0,h})=\sum_{T\in\mathcal{M}_{H,\Omega_{2}}}a_{\tilde{T}}(v_{h,h},w_{0,h})\quad\forall\,w_{0,h}\in W_{0,h},

and hence we have the following decomposition:

Qh=TH,Ω2QhT.\displaystyle Q_{h}=\sum_{T\in\mathcal{M}_{H,\Omega_{2}}}Q_{h}^{T}. (3.13)

Although the definitions of QhTvh,HQ_{h}^{T}v_{h,H} are independent of each other and can be computed in parallel, they are still global and have to be solved on the whole fine-scale space W0,hW_{0,h}.

Next we introduce local versions of the correction operators QhTQ_{h}^{T} and give error estimates between them. To this end, we need the definitions of element patches (c.f. [21]).

Definition 1 (element patches).

Given TH,Ω2T\in\mathcal{M}_{H,\Omega_{2}}, the patches TLT_{L} are defined recursively as follows:

T0:=T,\displaystyle T_{0}:=T,
TL:={TH,Ω2|TTL1}L=1,2,.\displaystyle T_{L}:=\left\{T^{\prime}\in\mathcal{M}_{H,\Omega_{2}}|\,\,T^{\prime}\cap{T_{L-1}}\neq\emptyset\right\}\quad L=1,2,\cdots.

The restriction of the fine-scale correction space W0,hW_{0,h} to the element patch TLT_{L} is defined by

W0,h(TL):={vh,hW0,h:vh,h=0 in Ω\TL}.\displaystyle W_{0,h}(T_{L}):=\{v_{h,h}\in W_{0,h}:\,\,v_{h,h}=0\,\,\text{ in }\,\,\Omega\backslash T_{L}\}.

The localized approximation of the correction operator QhTQ_{h}^{T} is defined as follows.

Definition 2.

For TH,Ω2{T}\in\mathcal{M}_{H,\Omega_{2}} and the patch TLT_{L}, the local correction operator QhT,LQ_{h}^{T,L}: Vh,hV_{h,h} \mapsto W0,h(TL)W_{0,h}(T_{L}) is defined as follows: given vh,hv_{h,h} \in Vh,hV_{h,h}, find QhT,Lvh,hQ_{h}^{T,L}v_{h,h} \in W0,h(TL)W_{0,h}(T_{L}) such that

aΩ(QhT,Lvh,h,w0,h)=aT~(vh,h,w0,h)w0,hW0,h(TL).a_{\Omega}(Q_{h}^{T,L}v_{h,h},w_{0,h})=a_{\tilde{T}}(v_{h,h},w_{0,h})\quad\forall\,w_{0,h}\in W_{0,h}(T_{L}). (3.14)

According to the decomposition (3.13) and the above definition, the global corrector of level LL is given by

QhL:=TH,Ω2QhT,L.{Q^{L}_{h}:=\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}Q_{h}^{T,L}.} (3.15)

Further, we define the localized multiscale approximation space as follows:

Vh,Hms,L:=(IQhL)Vh,H={vh,HQhLvh,H:vh,HVh,H}.\displaystyle V^{ms,L}_{h,H}:=\big{(}{I}-Q_{h}^{L}\big{)}V_{h,H}{=\left\{v_{h,H}-Q^{L}_{h}v_{h,H}:\;v_{h,H}\in V_{h,H}\right\}}. (3.16)

Then, the FE-LODM reads as: find uh,Hms,LVh,Hms,Lu^{ms,L}_{h,H}\in V^{ms,L}_{h,H}, such that

aΩ(uh,Hms,L,vh,Hms,L)=(f,vh,Hms,L)vh,Hms,LVh,Hms,L.a_{\Omega}\big{(}u^{ms,L}_{h,H},v^{ms,L}_{h,H}\big{)}=\big{(}f,v^{ms,L}_{h,H}\big{)}\quad\forall\,v^{ms,L}_{h,H}\in V^{ms,L}_{h,H}. (3.17)
Remark 4.

(1) It is clear that Vh,Hms,LVh,hV^{ms,L}_{h,H}\subset V_{h,h} as a consequence of the assumption that h,Ω2\mathcal{M}_{h,\Omega_{2}} is a refinement of H,Ω2\mathcal{M}_{H,\Omega_{2}}. Therefore the FE-LODM inherits the well-posedness from the reference problem (2.7).

(2) Unlike Vh,HmsV^{ms}_{h,H} whose multiscale basis functions supported globally on Ω2\Omega_{2}, the multiscale basis functions of Vh,Hms,LV^{ms,L}_{h,H} locally support on small patches of size O(LH)O(LH), and hence the computational cost for assembling the global system of the FE-LODM (3.17) is usually much less than that for the ideal combined method (3.7).

4 Error estimates for the FE-LODM

In this section, we derive the H1H^{1} and L2L^{2} error estimates for the proposed FE-LODM.

We first recall two local trace inequalities which will be used in this paper frequently. Here we omitted the proof since it is a direct consequence of the standard trace inequality (cf. [56, Theorem 1.6.6, p.39]) and the scaling argument (cf. [60]).

Lemma 4.1.

Let TT be an element in the triangulation h,Ω1\mathcal{M}_{h,\Omega_{1}}, h,Ω2\mathcal{M}_{h,\Omega_{2}}, or H,Ω2\mathcal{M}_{H,\Omega_{2}}. Then, we have

v0,T\displaystyle\|v\|_{0,\partial T} diam(T)12v0,T+v0,T12v0,T12vH1(T),\displaystyle\lesssim{\rm diam}(T)^{-\frac{1}{2}}\|v\|_{0,T}+\|v\|^{\frac{1}{2}}_{0,T}\|\nabla v\|^{\frac{1}{2}}_{0,T}\quad\forall\,v\,\in\,H^{1}(T),
v0,T\displaystyle\|v\|_{0,\partial T} diam(T)12v0,TvP1(T),\displaystyle\lesssim{\rm diam}(T)^{-\frac{1}{2}}\|v\|_{0,T}\quad\forall\,v\,\in\,P_{1}(T),

where the invisible constants depend only on the regularity of the element TT.

We shall also make use of the following norm on the restriction of the space Vh,HV_{h,H} onto a subdomain ωΩ\omega\subseteq\Omega:

vh,H,ω=(A12hv0,ω2+γ0H[v]Γω2)12,\|v\|_{h,H,\omega}=\Big{(}\|A^{\frac{1}{2}}\nabla_{h}v\|^{2}_{0,\omega}+\frac{\gamma_{0}}{H}\|[v]\|^{2}_{\Gamma\cap\omega}\Big{)}^{\frac{1}{2}}, (4.1)

and denote by h,H:=h,H,Ω\|\cdot\|_{h,H}:=\|\cdot\|_{h,H,\Omega}. Note that the norm h,H,ω\|\cdot\|_{h,H,\omega} is almost the same as the previous one h,h,ω\|\cdot\|_{h,h,\omega} in (2.10) except replacing γ0h\frac{\gamma_{0}}{h} there by γ0H\frac{\gamma_{0}}{H}.

4.1 Properties of the operator 𝒞h,H\mathcal{C}_{h,H}

In this subsection, we state two lemmas on the quasi-interpolation operator 𝒞h,H\mathcal{C}_{h,H} given in (3.3).

First we recall some stability and error estimates for the operator 𝒞H\mathcal{C}_{H}, whose proof can be found in [58, 59].

Lemma 4.2.

For any TH,Ω2T\in\mathcal{M}_{H,\Omega_{2}} and vhVh,Ω2{v}_{h}\in V_{h,\Omega_{2}}, there hold following estimates

h𝒞Hvh0,T\displaystyle\|\nabla_{h}\mathcal{C}_{H}v_{h}\|_{0,T} hvh0,T^,\displaystyle\lesssim\|\nabla_{h}v_{h}\|_{0,\hat{T}}, (4.2)
vh𝒞Hvh0,T\displaystyle\|v_{h}-\mathcal{C}_{H}v_{h}\|_{0,T} +Hh(vh𝒞Hvh)0,THhvh0,T^,\displaystyle+H\|\nabla_{h}(v_{h}-\mathcal{C}_{H}v_{h})\|_{0,T}\lesssim H\|\nabla_{h}v_{h}\|_{0,\hat{T}}, (4.3)

where T^={TH,Ω2:TT}.\hat{T}=\cup\{T^{{}^{\prime}}\in\mathcal{M}_{H,\Omega_{2}}:\;T^{{}^{\prime}}\cap{T}\neq\emptyset\}.

Using Lemma 4.2, we may prove the following stability result for 𝒞h,H\mathcal{C}_{h,H}.

Lemma 4.3.

For any vh,h{v_{h,h}} \in Vh,hV_{h,h}, it holds that

𝒞h,Hvh,hh,H,ωvh,hh,H,ω^,\|\mathcal{C}_{h,H}v_{h,h}\|_{h,H,\omega}\lesssim\|v_{h,h}\|_{h,H,\hat{\omega}}, (4.4)

where ω^:=({T^:Tω,TH,Ω2}{T:Tω,Th,Ω1})\hat{\omega}:=\bigcup\big{(}\{\hat{T}:\;T\cap\omega\neq\emptyset,T\in\mathcal{M}_{H,\Omega_{2}}\}\cup\{T:\;T\cap\omega\neq\emptyset,T\in\mathcal{M}_{h,\Omega_{1}}\}\big{)}.

Proof.

From the definitions (2.10) and (4.1) of the norms, (3.2), (3.3), and (4.2), it is easy to see that

𝒞h,Hvh,hh,H,ω2\displaystyle\|\mathcal{C}_{h,H}v_{h,h}\|^{2}_{h,H,\omega}\lesssim{} TH,Ω2ωA12hvh,h0,T^2\displaystyle\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}\cap\omega}\|A^{\frac{1}{2}}\nabla_{h}v_{h,h}\|^{2}_{0,\hat{T}}
+A12hvh,h0,Ω1ω2+γ0H[𝒞h,Hvh,h]Γω2\displaystyle+\|A^{\frac{1}{2}}\nabla_{h}v_{h,h}\|^{2}_{0,\Omega_{1}\cap\omega}+\frac{\gamma_{0}}{H}\|[\mathcal{C}_{h,H}v_{h,h}]\|^{2}_{{\Gamma\cap\omega}}
\displaystyle\lesssim{} A12hvh,h0,Ωω^2+γ0H[𝒞h,Hvh,h]Γω2.\displaystyle{\|A^{\frac{1}{2}}\nabla_{h}v_{h,h}\|^{2}_{0,\Omega\cap\hat{\omega}}+\frac{\gamma_{0}}{H}\|[\mathcal{C}_{h,H}v_{h,h}]\|^{2}_{\Gamma\cap\omega}}. (4.5)

For the second term on the right hand side, using the triangle inequality, we have

γ0H[𝒞h,Hvh,h]Γω2\displaystyle\frac{\gamma_{0}}{H}\|[\mathcal{C}_{h,H}v_{h,h}]\|^{2}_{\Gamma\cap\omega}\lesssim{} γ0H[𝒞h,Hvh,hvh,h]Γω2+γ0H[vh,h]Γω2\displaystyle{\frac{\gamma_{0}}{H}\|[\mathcal{C}_{h,H}v_{h,h}-v_{h,h}]\|^{2}_{\Gamma\cap\omega}+\frac{\gamma_{0}}{H}\|[v_{h,h}]\|^{2}_{\Gamma\cap\omega}}
\displaystyle\lesssim{} EΓHEωγ0H𝒞Hvh,Ω2vh,Ω2E2+γ0H[vh,h]Γω2\displaystyle\sum\limits_{\begin{subarray}{c}E\in\Gamma_{H}\\ E\cap\omega\neq\emptyset\end{subarray}}\frac{\gamma_{0}}{H}\|\mathcal{C}_{H}{v_{h,\Omega_{2}}}-v_{h,\Omega_{2}}\|^{2}_{E}+\frac{\gamma_{0}}{H}\|[v_{h,h}]\|^{2}_{\Gamma\cap\omega} (4.6)
:=\displaystyle:={} I+γ0H[vh,h]Γω2,\displaystyle\mathrm{I}+\frac{\gamma_{0}}{H}\|[v_{h,h}]\|^{2}_{\Gamma\cap\omega},

where vh,Ω2v_{h,\Omega_{2}}: =vh,h|Ω2v_{h,h}|_{\Omega_{2}}, and we have used the fact that 𝒞h,Hvh,h|Ω1=vh,h|Ω1\mathcal{C}_{h,H}v_{h,h}|_{\Omega_{1}}=v_{h,h}|_{\Omega_{1}}.

Further, from Lemma 4.1 and (4.3), it follows that

I\displaystyle\mathrm{I} TΓHTωγ0H(H1𝒞Hvh,Ω2vh,Ω20,T2+Hh(𝒞Hvh,Ω2vh,Ω2)0,T2)\displaystyle\lesssim\sum\limits_{\begin{subarray}{c}T\in\mathcal{M}_{\Gamma_{H}}\\ T\cap\omega\neq\emptyset\end{subarray}}\frac{\gamma_{0}}{H}\left(H^{-1}\|\mathcal{C}_{H}v_{h,\Omega_{2}}-v_{h,\Omega_{2}}\|^{2}_{0,T}+H\|\nabla_{h}(\mathcal{C}_{H}v_{h,\Omega_{2}}-v_{h,\Omega_{2}})\|^{2}_{0,T}\right)
TΓHTωhvh,Ω20,T^2,\displaystyle\lesssim\sum_{\begin{subarray}{c}T\in\mathcal{M}_{\Gamma_{H}}\\ T\cap\omega\neq\emptyset\end{subarray}}\|\nabla_{h}v_{h,\Omega_{2}}\|^{2}_{0,\hat{T}},

which together with (4.1) and (4.6) yields the result immediately. ∎

The following lemma gives a stability estimate of 𝒞h,H|Vh,H\mathcal{C}_{h,H}|_{V_{h,H}}, whose proof is arranged in A for the convenience of the reader.

Lemma 4.4.

𝒞h,H\mathcal{C}_{h,H} is an isomorphism on Vh,HV_{h,H} and satisfies the following estimate

(𝒞h,H|Vh,H)1vh,Hh,Hvh,Hh,Hvh,HVh,H.\displaystyle\big{\|}\big{(}\mathcal{C}_{h,H}|_{V_{h,H}}\big{)}^{-1}v_{h,H}\big{\|}_{h,H}\lesssim\left\|v_{h,H}\right\|_{h,H}\quad\forall\,v_{h,H}\in V_{h,H}.

The following lemma is crucial for the error analysis, which can be proved by following the proof of [19, Lemma 2.1] or [49, Lemma 1]. We omit the details.

Lemma 4.5.

For each v0,H{v}_{0,H} \in V0,HV_{0,H}, there exists a v0,hv_{0,h} \in V0,hV_{0,h}, such that 𝒞h,Hv0,h=v0,H\mathcal{C}_{h,H}v_{0,h}=v_{0,H}, v0,hh,h\|v_{0,h}\|_{h,h} \lesssim v0,Hh,H\|v_{0,H}\|_{h,H} and suppv0,hsupp(𝒞h,Hv0,H){{\rm supp}\,v_{0,h}}\subseteq{\rm supp}\,(\mathcal{C}_{h,H}v_{0,H}).

We emphasize that the above result holds for any function in V0,HV_{0,H}, not Vh,HV_{h,H}, which is sufficient for the later analysis. In fact, we have tried to use Vh,HV_{h,H} instead of V0,HV_{0,H}, but the error estimate has become worse, multiplying by an additional factor H/hH/h.

4.2 Error estimate for the ideal combined multiscale method

The following theorem gives an error bound for the ideal multiscale method (3.7), where the correctors for the basis functions have to be solved globally (see (3.5) and (3.6)). The proposed ideal combined multiscale method preserves the common linear order convergence O(H)O(H) for the H1H^{1}-error without suffering from preasymptotic effects due to the highly varying diffusion coefficient.

Theorem 4.1.

If uh,hu_{h,h} and uh,Hmsu^{ms}_{h,H} are the solutions of the reference problem (2.7) and the approximation problem (3.7) respectively, then it holds that

uh,huh,Hmsh,hHfL2(Ω).\|u_{h,h}-u^{ms}_{h,H}\|_{h,h}\lesssim H\|f\|_{L^{2}(\Omega)}. (4.7)
Proof.

Let eh=uh,huh,Hmse_{h}=u_{h,h}-u^{ms}_{h,H}. From Proposition 1 we have ehe_{h} \in W0,hW_{0,h}. Hence Ch,Heh=0{C}_{h,H}e_{h}=0. Thus, using the Cauchy-Schwarz inequality and the coercivity (2.9) of aΩa_{\Omega} on Vh,hV_{h,h}, we have

ehh,h2\displaystyle\|e_{h}\|^{2}_{h,h} aΩ(eh,eh)=(f,eh)=(f,eh𝒞h,Heh)\displaystyle\lesssim a_{\Omega}(e_{h},e_{h})=(f,e_{h})=(f,e_{h}-\mathcal{C}_{h,H}e_{h})
fL2(Ω)eh𝒞h,HehL2(Ω)\displaystyle\leq\|f\|_{L^{2}(\Omega)}\|e_{h}-\mathcal{C}_{h,H}e_{h}\|_{L^{2}(\Omega)}
=fL2(Ω)TH,Ω2eh𝒞Heh0,T.\displaystyle=\|f\|_{L^{2}(\Omega)}\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}\|e_{h}-\mathcal{C}_{H}e_{h}\|_{0,T}.

Further, from (4.3), it follows that

TH,Ω2eh𝒞Heh0,T\displaystyle\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}\|e_{h}-\mathcal{C}_{H}e_{h}\|_{0,T} TH,Ω2Heh0,T^Hehh,h,\displaystyle\lesssim\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}H\|\nabla e_{h}\|_{0,\hat{T}}\lesssim H\|e_{h}\|_{h,h},

which combines the above estimate yields the result immediately. ∎

4.3 Error estimates for the localized method

In this subsection we first estimate the errors between the correctors QhTQ_{h}^{T} and QhT,LQ_{h}^{T,L} due to the truncations to local patches. Then we provide H1H^{1}- and L2L^{2}- error bounds for the FE-LODM.

We will frequently make use of the following cut-off functions on element patches: for each TT \in MH,Ω2M_{H,\Omega_{2}} and l1<l2l_{1}<l_{2} \in \mathbb{N}, the cut-off functions ηTl1,l2\eta^{l_{1},l_{2}}_{T} \in Vh,HV_{h,H} satisfy

ηTl1,l2|Tl1=1,\displaystyle\eta^{l_{1},l_{2}}_{T}|_{T_{l_{1}}}=1, (4.8)
ηTl1,l2|Ω\Tl2=0,\displaystyle\eta^{l_{1},l_{2}}_{T}|_{\Omega\backslash T_{l_{2}}}=0, (4.9)
h\displaystyle\|\nabla_{h} ηTl1,l2L(Ω)1(l2l1)HT.\displaystyle\eta^{l_{1},l_{2}}_{T}\|_{L^{\infty}(\Omega)}\lesssim\frac{1}{(l_{2}-l_{1})H_{T}}. (4.10)

Let Ih,hI_{h,h} ( Ih,h|Ωi:HΓi1(Ωi)C(Ω¯i)Vh,Ωi,i=1,2I_{h,h}|_{\Omega_{i}}:H_{\Gamma_{i}}^{1}(\Omega_{i})\cap C(\bar{\Omega}_{i})\mapsto V_{h,\Omega_{i}},i=1,2 ) be the linear Lagrange interpolation operator with respect to h,h\mathcal{M}_{h,h}. The following lemma provides a stability estimate of the operator Ih,hI_{h,h}.

Lemma 4.6.

For TH,Ω2T\in\mathcal{M}_{H,\Omega_{2}}, assume that ηTs,n,n>s>0\eta_{T}^{s,n},n>s>0\in\mathbb{N} is the cut-off function which satisfies (4.8)–(4.10). Then for wW0,hw\in W_{0,h}, the following estimates hold

Ih,h(ηTs,nw)h,h\displaystyle\|I_{h,h}(\eta_{T}^{s,n}w)\|_{h,h} wh,h,Tn+1,\displaystyle\lesssim\|w\|_{h,h,T_{n+1}}, (4.11)
ηTs,nwIh,h(ηTs,nw)h,h\displaystyle\|\eta_{T}^{s,n}w-I_{h,h}(\eta_{T}^{s,n}w)\|_{h,h} wh,h,Tn+1\Ts1,\displaystyle\lesssim\|w\|_{h,h,T_{n+1}\backslash T_{s-1}}, (4.12)
Ih,h(ηTs,nw)h,h,Tn\Ts\displaystyle\|I_{h,h}(\eta_{T}^{s,n}w)\|_{h,h,T_{n}\backslash T_{s}} wh,h,Tn+1\Ts1,\displaystyle\lesssim\|w\|_{h,h,T_{n+1}\backslash T_{s-1}}, (4.13)
Ih,h(1ηTs,n)wh,h\displaystyle\|I_{h,h}(1-\eta_{T}^{s,n})w\|_{h,h} wh,h,Ω\Ts1.\displaystyle\lesssim\|w\|_{h,h,\Omega\backslash T_{s-1}}. (4.14)

The proof of this lemma is similar to that of [21, Lemma A.2], except that we have to deal with the elements near the interface between coarse and fine meshes. For convenience of the reader, we arrange it in B.

The following key lemma says that the errors of the localized correction problems decay exponentially with respect to the number of truncation layers LL.

Lemma 4.7.

Let uh,hu_{h,h} be the reference solution to (2.7) and uh,HmsVh,Hmsu_{h,H}^{ms}\in V_{h,H}^{ms} be the ideal solution to (3.7), respectively. Denote by uh,H=(𝒞h,H|Vh,H)1𝒞h,Huh,HmsVh,H.u_{h,H}=(\mathcal{C}_{h,H}|_{V_{h,H}})^{-1}\mathcal{C}_{h,H}u_{h,H}^{ms}\in V_{h,H}. Further, for TH,Ω2T\in\mathcal{M}_{H,\Omega_{2}} and its element patch TLT_{L}, let qhT=QhT(uh,H)q^{T}_{h}=Q_{h}^{T}(u_{h,H}) and qhT,L=QhT,L(uh,H)q_{h}^{T,L}=Q_{h}^{T,L}(u_{h,H}) be the global and local multiscale-corrected solution obtained in (3.12) and (3.14), respectively. Then there exists a constant 0<θ<10<\theta<1 independent of L,h,HL,h,H, and TT, such that

qhTqhT,Lh,hθLuh,Hh,h,T~,\|q^{T}_{h}-q_{h}^{T,L}\|_{h,h}\lesssim\theta^{L}\|u_{h,H}\|_{h,h,\tilde{T}},

where T~\tilde{T} is defined in Section 3.4.

The proof of this lemma is similar to that given in [19] and [21], but with some special details related to the interface elements need to be accounted for. To make the error analysis clearer, we arrange the proof of this lemma in C.

The following theorem gives the H1H^{1}-error estimate for the FE-LODM. Using this theorem, we can quantify how many truncation layers in the localization patches can ensure the linear convergence of O(H)O(H).

Theorem 4.2.

Suppose γ0α0\gamma_{0}\geq\alpha_{0}. Let uh,hu_{h,h} and uh,Hms,Lu^{ms,L}_{h,H} be the reference solution to (2.7) and the solution to the FE-LODM (3.17), respectively. Then we have

uh,huh,Hms,Lh,hHfL2(Ω)+(Hh)12Ld2θLfL2(Ω),\|u_{h,h}-u^{ms,L}_{h,H}\|_{h,h}\lesssim H\|f\|_{L^{2}(\Omega)}+\Big{(}\frac{H}{h}\Big{)}^{\frac{1}{2}}L^{\frac{d}{2}}\theta^{L}\|f\|_{L^{2}(\Omega)}, (4.15)

where 0<θ<10<\theta<1 is given in Lemma 4.7. Moreover, there exists a positive constant L0L_{0} such that when LL0|log(Hh)12|L\geq L_{0}|\log(Hh)^{\frac{1}{2}}|, we have the following estimate, which is of the same order as the ideal multiscale method,

uh,huh,Hms,Lh,hHfL2(Ω).\displaystyle\|u_{h,h}-u^{ms,L}_{h,H}\|_{h,h}\lesssim H\|f\|_{L^{2}(\Omega)}. (4.16)
Proof.

Let uh,HmsVh,Hmsu_{h,H}^{ms}\in V_{h,H}^{ms} be the solution to the ideal method (3.7) using the global basis. From (3.9) and (3.13), the ideal solution can be rewritten as follows:

uh,Hms=uh,HTH,Ω2QhT(uh,H),\displaystyle u_{h,H}^{ms}=u_{h,H}-\sum_{T\in\mathcal{M}_{H,\Omega_{2}}}Q_{h}^{T}(u_{h,H}),

where uh,H=(𝒞h,H|Vh,H)1𝒞h,Huh,HmsVh,Hu_{h,H}=(\mathcal{C}_{h,H}|_{V_{h,H}})^{-1}\mathcal{C}_{h,H}u_{h,H}^{ms}\in V_{h,H}. Denote by

u~h,Hms,L\displaystyle\widetilde{u}^{ms,L}_{h,H} :=uh,HTH,Ω2QhT,L(uh,H),\displaystyle:=u_{h,H}-\sum_{T\in\mathcal{M}_{H,\Omega_{2}}}Q_{h}^{T,L}(u_{h,H}),
z\displaystyle z :=TH,Ω2(QhT(uh,H)QhT,L(uh,H)).\displaystyle:=\sum_{T\in\mathcal{M}_{H,\Omega_{2}}}\big{(}Q_{h}^{T}(u_{h,H})-Q_{h}^{T,L}(u_{h,H})\big{)}.

It is easy to see that

uh,Hmsu~h,Hms,Lh,h=zh,h.\|u^{ms}_{h,H}-\widetilde{u}^{ms,L}_{h,H}\|_{h,h}=\|z\|_{h,h}. (4.17)

According to Lemma 4.5 with v0,H=𝒞h,HIh,h(zηTL+2,L+3z)v_{0,H}=\mathcal{C}_{h,H}I_{h,h}\big{(}z-\eta_{T}^{L+2,L+3}z\big{)}, there exists a function bV0,hb\in V_{0,h} such that

𝒞h,H(b)=𝒞h,HIh,h(z\displaystyle\mathcal{C}_{h,H}(b)=\mathcal{C}_{h,H}I_{h,h}\big{(}z- ηTL+2,L+3z),\displaystyle\eta_{T}^{L+2,L+3}z\big{)}, (4.18)
bh,h𝒞h,HIh,h(zηTL+2,L+3z)h,H\displaystyle\|b\|_{h,h}\lesssim\big{\|}\mathcal{C}_{h,H}I_{h,h}\big{(}z-\eta_{T}^{L+2,L+3}z\big{)}\big{\|}_{h,H} =𝒞h,HIh,h(ηTL+2,L+3z)h,H,\displaystyle=\big{\|}\mathcal{C}_{h,H}I_{h,h}\big{(}\eta_{T}^{L+2,L+3}z\big{)}\big{\|}_{h,H}, (4.19)

where we have used 𝒞h,HIh,hz=0\mathcal{C}_{h,H}I_{h,h}z=0 to derive the last equality, which is a consequence of the fact that zW0,hz\in W_{0,h}. From (4.9), we have

𝒞h,HIh,h(zηTL+2,L+3z)=𝒞h,HIh,h(ηTL+2,L+3z)=0inΩ\TL+4,\mathcal{C}_{h,H}I_{h,h}\big{(}z-\eta_{T}^{L+2,L+3}z\big{)}=-\mathcal{C}_{h,H}I_{h,h}\big{(}\eta_{T}^{L+2,L+3}z\big{)}=0\quad\text{in}\quad\Omega\backslash T_{L+4},

which together with Lemma 4.5, implies that

supp(b)supp(𝒞h,H2Ih,h(zηTL+2,L+3z))TL+5\TL.{\mathrm{supp}}(b)\subseteq{\mathrm{supp}}\big{(}{\mathcal{C}_{h,H}^{2}}I_{h,h}\big{(}z-\eta_{T}^{L+2,L+3}z\big{)}\big{)}\subseteq T_{L+5}\backslash T_{L}.

Therefore, from (4.18), we have Ih,h(zηTL+2,L+3z)bW0,hI_{h,h}\big{(}{z-}\eta_{T}^{L+2,L+3}z\big{)}-b\in W_{0,h} . Hence, from (3.12) it follows that

aΩ(QhTuh,H,Ih,h(zηTL3,L2z)b)=aT~(uh,H,Ih,h(zηTL+2,L+3z)b)=0.a_{\Omega}\big{(}Q_{h}^{T}u_{h,H},I_{h,h}(z-\eta_{T}^{L-3,L-2}z)-b\big{)}=a_{\tilde{T}}\big{(}u_{h,H},I_{h,h}(z-\eta_{T}^{L+2,L+3}z)-b\big{)}=0. (4.20)

Further, from supp(Ih,h(zηTL+2,L+3z)b)Ω2\TL\mathrm{supp}(I_{h,h}\big{(}{z-}\eta_{T}^{L+2,L+3}z\big{)}-b)\subset\Omega_{2}\backslash T_{L}, it follows that

aΩ(QhT,Luh,H,Ih,h(zηTL+2,L+3z)b)=0,a_{\Omega}\big{(}Q_{h}^{T,L}u_{h,H},I_{h,h}(z-\eta_{T}^{L+2,L+3}z)-b\big{)}=0,

which combines with (4.20) yields

aΩ((QhTQhT,L)uh,H,zIh,h(ηTL+2,L+3z)b)=0.a_{\Omega}\big{(}(Q_{h}^{T}-Q_{h}^{T,L})u_{h,H},z-I_{h,h}(\eta_{T}^{L+2,L+3}z)-b\big{)}=0. (4.21)

Therefore, from (2.9), it follows that

zh,h2\displaystyle\|z\|^{2}_{h,h} aΩ(z,z)\displaystyle\lesssim a_{\Omega}(z,z)
=TH,Ω2aΩ(QhTuh,HQhT,Luh,H,z)\displaystyle=\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}a_{\Omega}\big{(}{Q_{h}^{T}u_{h,H}-Q_{h}^{T,L}u_{h,H},z}\big{)}
=TH,Ω2aΩ((QhTQhT,L)uh,H,Ih,h(ηTL+2,L+3z)+b).\displaystyle=\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}a_{\Omega}\big{(}(Q_{h}^{T}-Q_{h}^{T,L})u_{h,H},I_{h,h}(\eta_{T}^{L+2,L+3}z)+b\big{)}.

Further, using (4.19) and Lemmas 4.3 and 4.6, we obtain

zh,h2\displaystyle\|z\|^{2}_{h,h} TH,Ω2(QhTQhT,L)uh,Hh,h(Ih,h(ηTL+2,L+3z)h,h+bh,h)\displaystyle\lesssim\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}\|(Q_{h}^{T}-Q_{h}^{T,L})u_{h,H}\|_{h,h}\big{(}\|I_{h,h}(\eta_{T}^{L+2,L+3}z)\|_{h,h}+\|b\|_{h,h}\big{)}
TH,Ω2(QhTQhT,L)uh,Hh,hIh,h(ηTL+2,L+3z)h,h\displaystyle{\lesssim\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}\|(Q_{h}^{T}-Q_{h}^{T,L})u_{h,H}\|_{h,h}\|I_{h,h}(\eta_{T}^{L+2,L+3}z)\|_{h,h}}
TH,Ω2(QhTQhT,L)uh,Hh,hzh,h,TL+4.\displaystyle\lesssim\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}\|(Q_{h}^{T}-Q_{h}^{T,L})u_{h,H}\|_{h,h}\|z\|_{h,h,T_{L+4}}.

Thus, by use of the Cauchy-Schwarz inequality, we have

zh,h2\displaystyle\|z\|^{2}_{h,h} (TH,Ω2(QhTQhT,L)uh,Hh,h2)12(TH,Ω2zh,h,TL+42)12\displaystyle\lesssim\bigg{(}\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}\|(Q_{h}^{T}-Q_{h}^{T,L})u_{h,H}\|_{h,h}^{2}\bigg{)}^{\frac{1}{2}}\bigg{(}\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}\|z\|_{h,h,T_{L+4}}^{2}\bigg{)}^{\frac{1}{2}}
(TH,Ω2(QhTQhT,L)uh,Hh,h2)12(Ld2zh,h),\displaystyle\lesssim\bigg{(}\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}\|(Q_{h}^{T}-Q_{h}^{T,L})u_{h,H}\|_{h,h}^{2}\bigg{)}^{\frac{1}{2}}\left(L^{\frac{d}{2}}\|z\|_{h,h}\right),

which yields

zh,h2LdTH,Ω2(QhTQhT,L)uh,Hh,h2.\|z\|_{h,h}^{2}\lesssim L^{d}\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}\|(Q_{h}^{T}-Q_{h}^{T,L})u_{h,H}\|_{h,h}^{2}.

According to Lemma 4.7, we have

zh,h2\displaystyle\|z\|_{h,h}^{2} Ldθ2LTH,Ω2uh,Hh,h,T~2Ldθ2Luh,Hh,h2.\displaystyle\lesssim L^{d}\theta^{2L}\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}\|u_{h,H}\|_{h,h,\tilde{T}}^{2}\lesssim L^{d}\theta^{2L}\|u_{h,H}\|_{h,h}^{2}.

Moreover, from Lemmas 4.4 and 4.3 and the stability estimate (3.8) of uh,Hmsu_{h,H}^{ms}, we have

uh,Hh,h2\displaystyle\|u_{h,H}\|_{h,h}^{2} Hhuh,Hh,H2=Hh(𝒞h,H|Vh,H)1𝒞h,Huh,Hmsh,H2\displaystyle\lesssim\frac{H}{h}\|u_{h,H}\|_{h,H}^{2}=\frac{H}{h}\|(\mathcal{C}_{h,H}|_{V_{h,H}})^{-1}\mathcal{C}_{h,H}u_{h,H}^{ms}\|_{h,H}^{2}
Hh𝒞h,Huh,Hmsh,H2Hhuh,Hmsh,HHhfL2(Ω).\displaystyle\lesssim\frac{H}{h}\|\mathcal{C}_{h,H}u_{h,H}^{ms}\|_{h,H}^{2}\lesssim\frac{H}{h}\|u_{h,H}^{ms}\|_{h,H}\lesssim\frac{H}{h}\|f\|_{L^{2}(\Omega)}.

Thus,

zh,h(Hh)12Ld2θLfL2(Ω).\|z\|_{h,h}\lesssim\Big{(}\frac{H}{h}\Big{)}^{\frac{1}{2}}L^{\frac{d}{2}}\theta^{L}\|f\|_{L^{2}(\Omega)}. (4.22)

Noting that Vh,Hms,LVh,hV_{h,H}^{ms,L}\subseteq V_{h,h}, from the continuity and coercivity of aΩa_{\Omega} (2.8)–(2.9), we have the following estimate of Céa lemma type:

uh,huh,Hms,Lh,hinfvh,Hms,LVh,Hms,Luh,hvh,Hms,Lh,h,\displaystyle\|u_{h,h}-u^{ms,L}_{h,H}\|_{h,h}\lesssim\inf\limits_{v^{ms,L}_{h,H}\in V_{h,H}^{ms,L}}\|u_{h,h}-v^{ms,L}_{h,H}\|_{h,h},

which implies that

uh,huh,Hms,Lh,huh,hu~h,Hms,Lh,h.\displaystyle\|u_{h,h}-u^{ms,L}_{h,H}\|_{h,h}\lesssim\|u_{h,h}-\widetilde{u}^{ms,L}_{h,H}\|_{h,h}.

Thus, we have

uh,huh,Hms,Lh,huh,huh,Hmsh,h+uh,Hmsu~h,Hms,Lh,h=uh,huh,Hmsh,h+zh,h,\displaystyle\begin{split}\|u_{h,h}-u^{ms,L}_{h,H}\|_{h,h}&\leq\|u_{h,h}-u^{ms}_{h,H}\|_{h,h}+\|u^{ms}_{h,H}-\widetilde{u}^{ms,L}_{h,H}\|_{h,h}\\ &=\|u_{h,h}-u^{ms}_{h,H}\|_{h,h}+\|z\|_{h,h},\end{split} (4.23)

which combines (4.7) and (4.22) yields the estimate (4.15) immediately.

It remains to prove (4.16). Let θ1=1+θ2\theta_{1}=\frac{1+\theta}{2}. Noting that Ld2θLθ1LL^{\frac{d}{2}}\theta^{L}\lesssim\theta_{1}^{L}, we have

(Hh)12Ld2θL(Hh)12θ1LH,if L|log(Hh)12||logθ1|,\displaystyle\Big{(}\frac{H}{h}\Big{)}^{\frac{1}{2}}L^{\frac{d}{2}}\theta^{L}\lesssim\Big{(}\frac{H}{h}\Big{)}^{\frac{1}{2}}\theta_{1}^{L}\lesssim H,\quad\text{if }L\geq\frac{|\log(Hh)^{\frac{1}{2}}|}{|\log\theta_{1}|},

which implies that (4.16) holds. This completes the proof of the theorem. ∎

Remark 5.

(1) If h=Hmh=H^{m} for some constant m>1m>1, then |log(Hh)12||logH||\log(Hh)^{\frac{1}{2}}|\eqsim|\log H|, and hence the condition LL0|log(Hh)12|L\geq L_{0}|\log(Hh)^{\frac{1}{2}}| becomes LL0|logH|L\geq L_{0}^{\prime}|\log H| for some sufficiently large constant L0L_{0}^{\prime}. Note that this is a standard condition for LOD type methods [19, 21, 47].

(2) In the case where h=H3h=H^{3}, it is easy to see that Hh=H1\sqrt{\frac{H}{h}}=H^{-1}. Hence (4.15) becomes

uh,huh,Hms,Lh,hHfL2(Ω)+H1Ld2θLfL2(Ω),\|u_{h,h}-u^{ms,L}_{h,H}\|_{h,h}\lesssim H\|f\|_{L^{2}(\Omega)}+H^{-1}L^{\frac{d}{2}}\theta^{L}\|f\|_{L^{2}(\Omega)}, (4.24)

which is the same result as those of the methods mentioned in [19, 48, 47]. We emphasize that in our later numerical experiments, we choose h=Hmh=H^{m} for some constant 1<m<31<m<3, which follows that Hh<H1\sqrt{\frac{H}{h}}<H^{-1}. This means in this case the estimate (4.15) is better than (4.24).

The following theorem gives the L2L^{2} error estimate of the proposed FE-LODM.

Theorem 4.3.

Suppose γ0α0\gamma_{0}\geq\alpha_{0}. Then we have the following estimate:

uh,huh,Hms,LL2(Ω)(H+(Hh)12Ld2θL)2fL2(Ω).\|u_{h,h}-u^{ms,L}_{h,H}\|_{L^{2}(\Omega)}\lesssim\Big{(}H+\Big{(}\frac{H}{h}\Big{)}^{\frac{1}{2}}L^{\frac{d}{2}}\theta^{L}\Big{)}^{2}\|f\|_{L^{2}(\Omega)}. (4.25)

Moreover, there exists a positive constant L0L_{0} such that when LL0|log(Hh)12|L\geq L_{0}|\log(Hh)^{\frac{1}{2}}|, we have the following estimate,

uh,huh,Hms,LL2(Ω)H2fL2(Ω).\displaystyle\|u_{h,h}-u^{ms,L}_{h,H}\|_{L^{2}(\Omega)}\lesssim H^{2}\|f\|_{L^{2}(\Omega)}. (4.26)
Proof.

It suffices to prove (4.25). Let eh=uh,huh,Hms,L.e_{h}=u_{h,h}-u^{ms,L}_{h,H}. We consider the dual problem

{(Aw)=ehinΩ,w=0onΩ.\left\{\begin{aligned} -\nabla\cdot(A\nabla{w})&={e_{h}}\quad\mathrm{in}\,\,\Omega,\\ w&=0\quad\,\,\mathrm{on}\,\,\partial\Omega.\end{aligned}\right.

Let wh,hVh,hw_{h,h}\in V_{h,h} be the IPCDG approximation of ww:

aΩ(wh,h,ϕh,h)=(eh,ϕh,h)ϕh,hVh,h,a_{\Omega}(w_{h,h},\phi_{h,h})=(e_{h},\phi_{h,h})\quad\forall\,\phi_{h,h}\in V_{h,h},

and let wh,Hms,LVh,Hms,Lw^{ms,L}_{h,H}\in V^{ms,L}_{h,H} be the FE-LOD approximation of ww:

aΩ(wh,Hms,L,vh,Hms,L)=(eh,vh,Hms,L)vh,Hms,LVh,Hms,L.a_{\Omega}\big{(}w^{ms,L}_{h,H},v^{ms,L}_{h,H}\big{)}=\big{(}e_{h},v^{ms,L}_{h,H}\big{)}\quad\forall\,v^{ms,L}_{h,H}\in V^{ms,L}_{h,H}.

Further, from (2.7) and (3.17), it follows that

aΩ(eh,wh,Hms,L)=0.a_{\Omega}\big{(}e_{h},w^{ms,L}_{h,H}\big{)}=0.

Thus from Theorem 4.2 we have

ehL2(Ω)2\displaystyle\|e_{h}\|^{2}_{L^{2}(\Omega)} =aΩ(wh,h,eh)=aΩ(eh,wh,hwh,Hms,L)\displaystyle=a_{\Omega}(w_{h,h},e_{h})=a_{\Omega}\big{(}e_{h},w_{h,h}-w^{ms,L}_{h,H}\big{)}
ehh,hwh,hwh,Hms,Lh,h\displaystyle\lesssim\|e_{h}\|_{h,h}\big{\|}w_{h,h}-w^{ms,L}_{h,H}\big{\|}_{h,h}
ehh,h(H+(Hh)12Ld2θL)ehL2(Ω),\displaystyle\lesssim\|e_{h}\|_{h,h}\Big{(}H+\Big{(}\frac{H}{h}\Big{)}^{\frac{1}{2}}L^{\frac{d}{2}}\theta^{L}\Big{)}\|e_{h}\|_{L^{2}(\Omega)},

which yields

ehL2(Ω)ehh,h(H+(Hh)12Ld2θL)(H+(Hh)12Ld2θL)2fL2(Ω).\|e_{h}\|_{L^{2}(\Omega)}\lesssim\|e_{h}\|_{h,h}\Big{(}H+\Big{(}\frac{H}{h}\Big{)}^{\frac{1}{2}}L^{\frac{d}{2}}\theta^{L}\Big{)}\lesssim\Big{(}H+\Big{(}\frac{H}{h}\Big{)}^{\frac{1}{2}}L^{\frac{d}{2}}\theta^{L}\Big{)}^{2}\|f\|_{L^{2}(\Omega)}.

This completes the proof of the theorem. ∎

5 Numerical Tests

In this section, we first numerically study how the size of element patches affects the errors, and then illustrate the ability of the proposed FE-LODM to deal with singularities by solving multiscale elliptic problems with a corner singularity and high-contrast channels and steady flow transporting through highly heterogeneous porous media driven by extraction wells, respectively. For comparison, we also present results of the local orthogonal decomposition method (LODM) in [53] and the combined multiscale finite element method (FE-OMsPGM) introduced in [46]. We use the IPCDG solution uh,hu_{h,h} to (2.7) on a very fine mesh as a reference solution. Denote the energy norm by E:=A120,Ω1Ω2\|\cdot\|_{E}:=\|\nabla A^{\frac{1}{2}}\cdot\|_{0,\Omega_{1}\cup\Omega_{2}}. We measure the relative errors of an approximate solution UhU_{h} in the L2L^{2}, LL^{\infty} and energy norms respectively as follows:

Uhuh,hL2uh,hL2,Uhuh,hLuh,hL,Uhuh,hEuh,hE.\frac{\|{U_{h}}-u_{h,h}\|_{L^{2}}}{\|u_{h,h}\|_{L^{2}}},\quad\frac{\|U_{h}-u_{h,h}\|_{L^{\infty}}}{\|u_{h,h}\|_{L^{\infty}}},\quad\frac{\|U_{h}-u_{h,h}\|_{E}}{\|u_{h,h}\|_{E}}.

5.1 Effect of the size of the element patches

In this subsection we study how the size of element patches affects the errors by simulating the following example.

Example 1.

Consider the model problem (2.1) on the unit square Ω=(0,1)×(0,1)\Omega=(0,1)\times(0,1) with the source term f1f\equiv 1 and different diffusion coefficients to be specified below. And we set Ω1=(14,38)×(14,38)\Omega_{1}=(\frac{1}{4},\frac{3}{8})\times(\frac{1}{4},\frac{3}{8}) as shown in Figure 3.

Refer to caption
Figure 3: An illustration of the separated domain used in Example 1.

First we test the dependence of the error on the size of the element patches. Consider the following diffusion coefficient:

A(x1,x2)=2+1.8sin(2πx1/ϵ)2+1.8cos(2πx2/ϵ)+2+1.8sin(2πx2/ϵ)2+1.8sin(2πx1/ϵ){A}(x_{1},x_{2})=\frac{2+1.8\sin(2\pi x_{1}/\epsilon)}{2+1.8\cos(2\pi x_{2}/\epsilon)}+\frac{2+1.8\sin(2\pi x_{2}/\epsilon)}{2+1.8\sin(2\pi x_{1}/\epsilon)} (5.1)

with ϵ=1/5\epsilon=1/5. We fix HH=232^{-3}, h=27h=2^{-7}, and let the size of element patches vary. Table 1 shows the relative errors in the energy and L2L^{2} norms on Ω\Omega and Ω1\Omega_{1} between the reference solution uh,hu_{h,h} and the FE-LOD solution uh,Hms,Lu_{h,H}^{ms,L} with L=1,2,3,6,10L=1,2,3,6,10 and the ideal solution uh,Hmsu_{h,H}^{ms} as well, respectively.

Table 1: Example 1: Relative errors for different LL, h=27h=2^{-7}, H=23H=2^{-3}, γ0\gamma_{0}=10.
  LL Error Error in Ω\Omega Error in Ω1\Omega_{1}
Energy L2L^{2} Energy L2L^{2}
1 0.1360e-00 0.2929e-01 0.2580e-01 0.1429e-01
2 0.7361e-01 0.1127e-01 0.5881e-02 0.1371e-02
3 0.5712e-01 0.8685e-02 0.1453e-02 0.1712e-03
6 0.5534e-01 0.8625e-02 0.2586e-04 0.6106e-05
10 0.5509e-01 0.8570e-02 0.5213e-06 0.1604e-06
ideal solution 0.5509e-01 0.8569e-02 0.5984e-13 0.2713e-13

It is observed that the larger the parameter LL, the smaller the relative errors in the energy and L2L^{2} norms on Ω\Omega, and tends the errors of the ideal solution, respectively. This observation verifies the estimate in Theorem 4.2. We also notice that the errors of the FE-LOD solution on Ω1\Omega_{1} decrease very quickly as LL increases. Especially, in the ideal case, the errors on Ω1\Omega_{1} are almost equal to zero, which is coincided with the result stated in Proposition 1.

Secondly, we study how to choose the size (LL) of the element patches to achieve the satisfied approximation behaviour for different coarse-fine grid elements. Recall that in Theorem 4.2, to balance the error between the terms on the right-hand side of (4.15) , it is required that the localization parameter LL satisfies LL0|log(Hh)12|L\geq L_{0}|\log(Hh)^{\frac{1}{2}}| for some positive constant L0L_{0}. Hence, in the following experiments, we choose L=L0|log(Hh)12|L=\lceil L_{0}|\log(Hh)^{\frac{1}{2}}|\rceil for different constants L0L_{0}. We adopt uniform coarse meshes with sizes H=2iH=2^{-i}, i=2,3,4,5i=2,3,4,5, and choose the fine scale reference mesh with size h=29h=2^{-9}, which can resolve the multiscale feature of AA. The first test is done for the periodic diffusion coefficient defined in (5.1) with ϵ=1/20\epsilon=1/20, which is denoted by A1A_{1} for convenience.

Refer to caption

Refer to caption

Figure 4: Example 1 with diffusion coefficient A1(x)A_{1}(x): Relative errors in energy-norm (left) and L2L^{2}-norm (right) against the size of coarse mesh for L0=1/2,1,L_{0}=1/2,1, and 3/23/2, respectively.

Figure 4 shows the log-log plots of the relative errors in energy-norm (left) and L2L^{2}-norm (right) against the size of coarse mesh (HH) with different constants L0=1/2,1,3/2L_{0}=1/2,1,3/2, respectively. It is observed that the method with L0=1/2L_{0}=1/2 does not performs well for large mesh size HH, while when L0L_{0} is taken as 11 or 3/23/2, the error between the terms on the right-hand side of estimate in Theorem 4.2 seems to be balanced, and the convergence rate of the energy-norm error maintains as well as that of the L2L^{2}-norm error.

Note that for larger localization parameter LL, it costs more computational effort to compute the corrector functions and cause reduced sparseness in the coarse scale stiffness matrix. Therefore in the remaining numerical experiments we use L0=1L_{0}=1. In order to further illustrate the reasonability of choosing L0=1L_{0}=1, we show the relative error results in Figure 5 for three different diffusion coefficients AA: A1A_{1} is defined as above; A2A_{2} is taken as the background medium in Figure 7, which is a piecewise constant function on a Cartesian grid of size 292^{-9} and is periodic in both the xx- and yy-directions; A3A_{3} is a randomly generated diffusion coefficient using the moving ellipse average technique in [61] with the parameters described in Example 2 below. It can be seen that taking L|log(Hh)12|L\geq|\log(Hh)^{\frac{1}{2}}| (i.e. L0=1L_{0}=1) in the FE-LODM can give the optimal convergence rates in both energy- and L2L^{2}- norms for all cases, which are the same as those of the ideal combined multiscale method (see Theorems 4.14.3).

Refer to caption

Refer to caption

Figure 5: Example 1: Relative errors in energy-norm (left) and L2L^{2}-norm (right) against the size of coarse mesh for L0=1L_{0}=1 and A=AiA=A_{i}, i=1,2,3,i=1,2,3, respectively.

5.2 Application to the multiscale elliptic problem with corner singularity

In this subsection, we consider the following L-shaped domain problem

Example 2.

The multiscale problem (2.1) on the L-shaped domain Ω=((0,1)×(0,1))\((12,1)×(0,12))\Omega=\big{(}(0,1)\times(0,1)\big{)}\backslash\big{(}(\frac{1}{2},1)\times(0,\frac{1}{2})\big{)} with the random log-normal permeability field AA, which is generated by using the moving ellipse average technique [61] with the variance of the logarithm of the permeability σ2=1.5\sigma^{2}=1.5, and the correlation lengths l1=l2=0.01l_{1}=l_{2}=0.01 (isotropic heterogeneity) in x1x_{1} and x2x_{2} directions, respectively. One realization of the resulting permeability field in our numerical experiments is depicted in Figure 6.

Refer to caption
Figure 6: Example 2: The random log-normal permeability field AA. AmaxAmin\frac{A_{max}}{A_{min}}=2.9642e+03.

.

In this example, we set the refined subdomain Ω1=((38,58)×(38,58))\((12,58)×(38,12))\Omega_{1}=\big{(}(\frac{3}{8},\frac{5}{8})\times(\frac{3}{8},\frac{5}{8})\big{)}\backslash\big{(}(\frac{1}{2},\frac{5}{8})\times(\frac{3}{8},\frac{1}{2})\big{)} to capture the singularity at the reentrant corner, fix H=25H=2^{-5}, h=210h=2^{-10}, and choose the parameter L=log|Hh|=3L=\lceil\log\sqrt{|Hh|}\rceil=3. We compare the relative errors of the FE-LOD solution in the L2L^{2}, LL^{\infty}, and energy norms with those of the LOD solution and FE-OMsPG [46] solution in the whole domain as well as in the refined region Ω1\Omega_{1}. The errors are listed in Table 2. We observe that the introduced FE-LODM gives a better approximation than the LOD and FE-OMsPG methods. In particular, in the refined region Ω1\Omega_{1}, our method gives much better results than the LOD method, which is very useful if one needs high-accuracy solution in the problematic area.

Table 2: Example 2: Relative errors for the model problem on the L-shaped domain. h=210h=2^{-10}, H=25H=2^{-5}, γ0\gamma_{0}=10.
  Method Error Energy norm L2L^{2} LL^{\infty}
LODM 0.2834e-01 0.1553e-02 0.6536e-02
FE-LODM 0.2628e-01 0.1449e-02 0.6526e-02
FE-OMsPGM 0.1045e-00 0.7596e-02 0.2424e-01
LODM (error in Ω1\Omega_{1}) 0.1507e-02 0.1695e-02 0.5583e-02
FE-LODM (error in Ω1\Omega_{1} ) 0.4257e-03 0.4043e-03 0.5018e-03

5.3 Application to the multiscale problem with high-contrast channels

In this subsection we use the FE-LOD method to solve the elliptic multiscale problem with high–contrast channels.

Example 3.

The oscillating coefficient is set as that of [46]. Namely, as shown in Figure 7, we set the high-permeability channels and inclusions with permeability values equal to 10510^{5} and 8×1048\times 10^{4} respectively, and set the other values as 11.

Refer to caption
Figure 7: Example 3: Permeability field A=105A=10^{5} in two channels consisting of dark small rectangles; A=8×104A=8\times 10^{4} in small square inclusions; A=1A=1 otherwise.

We set Ω1\Omega_{1} be the union of two layers of coarse-grid elements which contain the channels, fix h=210h=2^{-10}, H=25H=2^{-5}, and choose the parameter L=log|Hh|=3L=\lceil\log\sqrt{|Hh|}\rceil=3 for this example. The results are listed in Table 3. It is observed that the FE-LOD method performs much better than the other two methods.

Table 3: Example 3: Relative errors for the model problem with the coefficient given by Figure 7 . h=210h=2^{-10} ,  H=25H=2^{-5} ,  γ0=10\gamma_{0}=10.
Relative error Energy norm L2L^{2} LL^{\infty}
LODM 0.4938e-01 0.4882e-02 0.1889e-01
FE-LODM 0.2169e-01 0.7238e-03 0.1248e-02
FE-OMsPGM 0.1063e-00 0.5564e-02 0.2580e-00

5.4 Application to the multiscale problem with Dirac singularities

In this subsection, we consider the multiscale problem with singular source terms inside the domain, which originates from the simulation of steady flow transporting through highly heterogeneous porous media driven by extraction wells. This kind of well-singularity problem is of great importance in hydrology, petroleum reservoir engineering, and soil venting techniques.

Denote by d(P,r)d({P},r) the disk centered at point PP with radius r>0r>0. We let Ω\Omega=(0,1)×(0,1)(0,1)\times(0,1) and consider two wells dj=d(Pj,r),j=1,2d_{j}=d({P}_{j},r),j=1,2 with P1(14,34)P_{1}(\frac{1}{4},\frac{3}{4}), P2(34,14)P_{2}(\frac{3}{4},\frac{1}{4}), and r=105r=10^{-5}. Since the size of the well (radius rr) is negligible in situations, we make an approximation to the original single phase pressure equation by the multiscale problem (2.1) with source term f=j=12qjδPjf=\sum\limits_{j=1}^{2}q_{j}\delta_{P_{j}} (c.f. [39]), where qjq_{j} is the well flow rate and δPj\delta_{P_{j}} is the Dirac measure at PjP_{j}. On the well boundary dj\partial d_{j}, two quantities are of particular importance in practical applications: the well bore pressure (WBP) and the well flow rate. Here we fix the well flow rate qjq_{j} and try to find the well bore pressure. In the computations we always take q1=1q_{1}=-1 and q2=1q_{2}=1, which corresponds to the situation that the well d1d_{1} is an extraction well and d2d_{2} is an injection well.

In the following two examples, we set Ω1\Omega_{1} be the union of two small squares with edge size of 116\frac{1}{16} centered at points Pi,i=1,2P_{i},i=1,2. And similarly, we choose the localization parameter L=log|Hh|=3L=\lceil\log\sqrt{|Hh|}\rceil=3.

Example 4.

Let the oscillating coefficient AA be given by

A(x1,x2)=1(2+1.5sin2πx1ϵ)(2+1.5sin2πx2ϵ),A(x_{1},x_{2})=\frac{1}{(2+1.5\sin\frac{2\pi x_{1}}{\epsilon})(2+1.5\sin\frac{2\pi x_{2}}{\epsilon})}, (5.2)

where we fix ϵ=1/64\epsilon=1/64.

Since the exact WBPs are unknown, we use the method introduced in [39] to compute them based on the well-resolved solutions obtained on a uniform 2048×20482048\times 2048 mesh. Then we can get the “exact” WBP α1=5.3884973\alpha_{1}=-5.3884973 in the first well and α2=5.3884973\alpha_{2}=5.3884973 in the second well (see [39, Example 7.1]).

In addition, we implement three other methods for comparison, including the LODM, the MsFEM introduced in [39, Algorithm 7.1] (referred as G-MsFEM) and the FE-OMsPGM introduced in [46]. The G-MsFEM needs to compute the discrete Green functions in a very fine mesh and it uses the developed new Peaceman method to compute the WBPs (see [39, Section 6]). We also use the new Peaceman method to calculate the WBP on each well for the LOD and FE-LOD method. For FE-OMsPGM, we use the Peaceman model [62] to compute the WBPs since the bilinear form of FE-OMsPG method is nonsymmetric. The results are listed in Table 4. We can see that our FE-LODM provides a better approximation of the WBP than G-MsFEM and FE-OMsPGM, and a much better approximation than the LODM in this example.

Table 4: Example 4: WBPs and relative errors at two wells. h=211h=2^{-11} ,  H=26H=2^{-6} ,  γ0\gamma_{0}=10.
Methods Well no. WBP Relative error
G-MsFEM 1 -5.3838442 0.8635e-03
FE-OMsPGM 1 -5.3843102 0.7770e-03
LODM 1 -5.6792672 0.5396e-01
FE-LODM 1 -5.3876085 0.1649e-03
G-MsFEM 2 5.3739254 0.2704e-02
FE-OMsPGM 2 5.3843102 0.7770e-03
LODM 2 5.6792672 0.5396e-01
FE-LODM 2 5.3876085 0.1649e-03
Example 5.

We generate the random permeability field AA on a uniform 1024×10241024\times 1024 mesh by using the technique in [61]. Figure 8 shows a realization of the random permeability field.

Refer to caption
Figure 8: Example 5: The random permeability field AA, the ratio of AmaxAmin\frac{A_{max}}{A_{min}}=6.06629e+0036.06629e+003.

Using the same method as above, we can get the “exact” well bore pressures α1=0.9860407\alpha_{1}=-0.9860407 and α2=4.6507306\alpha_{2}=4.6507306 by using the fine-grid solution on the 1024×10241024\times 1024 mesh. The results are presented in Table 5. We observe that the proposed FE-LODM gives much better approximation than the other three methods, which may be due to the fact that the FE-LODM has more accurate solution near the well than others. This superiority can also be found in the results of FE–OMsPGM, in which the local fine mesh approximation is used in the well-singularity region that is same as that of FE-LODM.

Table 5: Example 5: WBPs and relative errors at two wells. h=210h=2^{-10} ,  H=26H=2^{-6} ,  γ0\gamma_{0}=10.
Methods Well no. WBP Relative error
G-MsFEM 1 -0.9478413 0.3874e-01
FE-OMsPGM 1 -0.9701813 0.1608e-01
LODM 1 -0.7536890 0.2356e-00
FE-LODM 1 -0.9864050 0.3695e-03
G-MsFEM 2 1.1477417 0.7532e-00
FE-OMsPGM 2 4.6391148 0.2498e-02
LODM 2 2.8657275 0.3838e-00
FE-LODM 2 4.6591764 0.1816e-02

6 Conclusions

In this paper, we have proposed a new combined multiscale method to solve the multiscale elliptic problems which may have singularities. In order to get a good approximation of the solution in the problematic region, we use the traditional FEM directly on a very fine mesh of this subdomain, while in other region where we have a highly oscillating coefficients, we use the multiscale LODM. The key point of implementing this idea is how to define the corrected basis function in the near interface elements. To this end, we introduce a special definition of the cell problems for the elements near the interface. The error analysis is carried out for highly varying coefficients, without any assumption on scale separation or periodicity. Our theoretical and numerical results show that the proposed method is very attractive for multiscale problems with singularities.

Appendix A Proof of Lemma 4.4

Given vH=j𝒩̊HvjΦjVH,Ω2,v_{H}=\sum\limits_{j\in\mathring{\mathcal{N}}_{H}}v_{j}\Phi_{j}\in V_{H,\Omega_{2}}, let wH:=𝒞HvHw_{H}:=\mathcal{C}_{H}v_{H}. From (3.1) and (3.3), it follows that

wH=i𝒩̊HwiΦiwithwi=j𝒩̊H(Φi,Φj)Ω2(1,Φi)Ω2vj.\displaystyle w_{H}=\sum\limits_{i\in\mathring{\mathcal{N}}_{H}}w_{i}\Phi_{i}\quad\text{with}\,w_{i}=\sum\limits_{j\in\mathring{\mathcal{N}}_{H}}\frac{(\Phi_{i},\Phi_{j})_{\Omega_{2}}}{(1,\Phi_{i})_{\Omega_{2}}}v_{j}.

Denote by D=diag((1,Φ1)Ω2,(1,Φ2)Ω2,,(1,ΦN)Ω2)D=\text{diag}\big{(}(1,\Phi_{1})_{\Omega_{2}},(1,\Phi_{2})_{\Omega_{2}},\cdots,(1,\Phi_{N})_{\Omega_{2}}\big{)}, M=((Φi,Φj)Ω2)N×NM=\big{(}(\Phi_{i},\Phi_{j})_{\Omega_{2}}\big{)}_{N\times N}, V=(vi)N×1,W=(wi)N×1,V=(v_{i})_{N\times 1},W=(w_{i})_{N\times 1}, where NN represents the numbers of vertices in 𝒩̊H\mathring{\mathcal{N}}_{H}. Thus we have D1MV=WD^{-1}MV=W, which yields

VTMV=VTDW.V^{T}MV=V^{T}DW.

Hence, by use of the above equation, it is follows that

vHL2(Ω)2\displaystyle\|v_{H}\|^{2}_{L^{2}(\Omega)} =VTMV|V||DW|Hd2vHL2(Ω)Hd|W|\displaystyle=V^{T}MV\leq|V||DW|\lesssim H^{-\frac{d}{2}}\|v_{H}\|_{L^{2}(\Omega)}H^{d}|W|
vHL2(Ω)wHL2(Ω),\displaystyle\lesssim\|v_{H}\|_{L^{2}(\Omega)}\|w_{H}\|_{L^{2}(\Omega)},

which yields

vHL2(Ω)\displaystyle\|v_{H}\|_{L^{2}(\Omega)} wHL2(Ω).\displaystyle\lesssim\|w_{H}\|_{L^{2}(\Omega)}.

Therefore 𝒞H\mathcal{C}_{H} is an isomorphism on VH,Ω2V_{H,\Omega_{2}} and it holds that

(𝒞H|VH,Ω2)1wH0,Ω2\displaystyle\|(\mathcal{C}_{H}|_{V_{H,\Omega_{2}}})^{-1}w_{H}\|_{0,\Omega_{2}} wH0,Ω2.\displaystyle\lesssim\|w_{H}\|_{0,\Omega_{2}}. (A.1)

Thus, it follows from (3.2) that 𝒞h,H\mathcal{C}_{h,H} is an isomorphism on Vh,HV_{h,H}. For the sake of simplicity, we denote (𝒞h,H|Vh,H)1(\mathcal{C}_{h,H}|_{V_{h,H}})^{-1} by 𝒞h,H1\mathcal{C}_{h,H}^{-1} in the following. From (3.2), it is clear that 𝒞h,Hvh,Hvh,HV0,H\mathcal{C}_{h,H}v_{h,H}-v_{h,H}\in V_{0,H}. Using the inverse inequality and Lemma 4.1, we have

vh,H𝒞h,H1vh,Hh,H2=\displaystyle\|v_{h,H}-\mathcal{C}_{h,H}^{-1}v_{h,H}\|_{h,H}^{2}={} 𝒞h,H1(𝒞h,Hvh,Hvh,H)h,H2\displaystyle\|\mathcal{C}_{h,H}^{-1}(\mathcal{C}_{h,H}v_{h,H}-v_{h,H})\|^{2}_{h,H}
=\displaystyle={} A12((𝒞H|VH,Ω2)1(𝒞h,Hvh,Hvh,H))0,Ω22\displaystyle\|A^{\frac{1}{2}}\nabla\big{(}(\mathcal{C}_{H}|_{V_{H,\Omega_{2}}})^{-1}(\mathcal{C}_{h,H}v_{h,H}-v_{h,H})\big{)}\|^{2}_{0,\Omega_{2}}
+γ0H(𝒞H|VH,Ω2)1(𝒞h,Hvh,Hvh,H)Γ2\displaystyle+\frac{\gamma_{0}}{H}\|(\mathcal{C}_{H}|_{V_{H,\Omega_{2}}})^{-1}(\mathcal{C}_{h,H}v_{h,H}-v_{h,H})\|^{2}_{\Gamma}
\displaystyle\lesssim{} H2(𝒞H|VH,Ω2)1(𝒞h,Hvh,Hvh,H)0,Ω2.2\displaystyle H^{-2}\|(\mathcal{C}_{H}|_{V_{H,\Omega_{2}}})^{-1}(\mathcal{C}_{h,H}v_{h,H}-v_{h,H})\|^{2}_{0,\Omega_{2}.}

Further, from (A.1) and (4.3), we have

vh,H𝒞h,H1vh,Hh,H2\displaystyle\|v_{h,H}-\mathcal{C}_{h,H}^{-1}v_{h,H}\|_{h,H}^{2} H2𝒞h,Hvh,Hvh,H0,Ω2\displaystyle\lesssim H^{-2}\|\mathcal{C}_{h,H}v_{h,H}-v_{h,H}\|^{2}_{0,\Omega}
=H2TH,Ω2𝒞h,Hvh,Hvh,H0,T2\displaystyle=H^{-2}\sum\limits_{T\in{\mathcal{M}}_{H,\Omega_{2}}}\|\mathcal{C}_{h,H}v_{h,H}-v_{h,H}\|^{2}_{0,T}
vh,Hh,H2.\displaystyle\lesssim\|v_{h,H}\|^{2}_{h,H}.

Finally, using triangle inequality, we conclude that

𝒞h,H1vh,Hh,H\displaystyle\|\mathcal{C}_{h,H}^{-1}v_{h,H}\|_{h,H} 𝒞h,H1vh,Hvh,Hh,H+vh,Hh,H\displaystyle\leq\|\mathcal{C}_{h,H}^{-1}v_{h,H}-v_{h,H}\|_{h,H}+\|v_{h,H}\|_{h,H}
vh,Hh,H.\displaystyle\lesssim\|v_{h,H}\|_{h,H}.

This completes the proof of the lemma. ∎

Appendix B Proof of Lemma 4.6

We first prove the inequality (4.11). From the interpolation error estimates and the inverse inequality, we have

Ih,h(ηTs,nw)0,Ω2\displaystyle\|\nabla I_{h,h}(\eta_{T}^{s,n}w)\|_{0,\Omega_{2}} Ih,h(ηTs,nw)(ηTs,nw)0,Ω2+(ηTs,nw)0,Ω2\displaystyle\leq\|\nabla I_{h,h}(\eta_{T}^{s,n}w)-\nabla(\eta_{T}^{s,n}w)\|_{0,\Omega_{2}}+\|\nabla(\eta_{T}^{s,n}w)\|_{0,\Omega_{2}}
(Th,Ω2hT2|ηTs,nw|2,T2)12+(ηTs,nw)0,Ω2\displaystyle\lesssim\bigg{(}\sum_{T\in\mathcal{M}_{h,\Omega_{2}}}h_{T}^{2}|\eta_{T}^{s,n}w|_{2,T}^{2}\bigg{)}^{\frac{1}{2}}+\|\nabla(\eta_{T}^{s,n}w)\|_{0,\Omega_{2}}
(ηTs,nw)0,Ω2.\displaystyle\lesssim\|\nabla{(\eta_{T}^{s,n}w)}\|_{0,\Omega_{2}}. (B.1)

Using the triangle inequality, we obtain

Ih,h(ηTs,nw)h,h2=\displaystyle\|I_{h,h}(\eta_{T}^{s,n}w)\|^{2}_{h,h}={} A12Ih,h(ηTs,nw)0,Ω22+eΓhγ0hIh,h(ηTs,nw)e2\displaystyle\|A^{\frac{1}{2}}\nabla I_{h,h}(\eta_{T}^{s,n}w)\|^{2}_{0,\Omega_{2}}+\sum\limits_{e\in\Gamma_{h}}\frac{\gamma_{0}}{h}\|I_{h,h}(\eta_{T}^{s,n}w)\|^{2}_{e}
\displaystyle\lesssim{} A12(ηTs,nw)0,Ω22+eΓhγ0hηTs,nwIh,h(ηTs,nw)e2\displaystyle\|A^{\frac{1}{2}}\nabla(\eta_{T}^{s,n}w)\|^{2}_{0,\Omega_{2}}+\sum\limits_{e\in\Gamma_{h}}\frac{\gamma_{0}}{h}\|\eta_{T}^{s,n}w-I_{h,h}(\eta_{T}^{s,n}w)\|^{2}_{e}
+eΓhγ0hηTs,nwe2:=R1+R2+R3.\displaystyle+\sum\limits_{e\in\Gamma_{h}}\frac{\gamma_{0}}{h}\|\eta_{T}^{s,n}w\|^{2}_{e}:=\mathrm{R}_{1}+\mathrm{R}_{2}+\mathrm{R}_{3}.

Using the fact that 𝒞h,Hw\mathcal{C}_{h,H}w=0, from (4.3) and (4.10), we have

R1\displaystyle\mathrm{R}_{1}\lesssim TTn\Ts(w𝒞h,Hw)ηTs,n0,T2+A12w0,Tn2\displaystyle\sum\limits_{T\in T_{n}\backslash T_{s}}\|{(w-\mathcal{C}_{h,H}w)}\nabla\eta_{T}^{s,n}\|^{2}_{0,T}+\|A^{\frac{1}{2}}\nabla w\|^{2}_{0,T_{n}}
\displaystyle\lesssim HηTs,nL(Ω)2w0,Tn+1\Ts12+A12w0,Tn2\displaystyle\|H\nabla\eta_{T}^{s,n}\|^{2}_{L^{\infty}(\Omega)}\|\nabla w\|^{2}_{0,T_{n+1}\backslash T_{s-1}}+\|A^{\frac{1}{2}}\nabla w\|^{2}_{0,T_{n}}
\displaystyle\lesssim A12w0,Tn+12.\displaystyle\|A^{\frac{1}{2}}\nabla w\|^{2}_{0,T_{n+1}}.

Further, from Lemma 4.1, it follows that

R2\displaystyle\mathrm{R}_{2}\lesssim Th,Ω2γ0h(h1ηTs,nwIh,h(ηTs,nw)0,T2\displaystyle\sum_{T\in\mathcal{M}_{h,\Omega_{2}}}\frac{\gamma_{0}}{h}\Big{(}h^{-1}\|\eta_{T}^{s,n}w-I_{h,h}(\eta_{T}^{s,n}w)\|^{2}_{0,T}
+ηTs,nwIh,h(ηTs,nw)0,T(ηTs,nwIh,h(ηTs,nw))0,T)\displaystyle\quad+\|\eta_{T}^{s,n}w-I_{h,h}(\eta_{T}^{s,n}w)\|_{0,T}\|\nabla(\eta_{T}^{s,n}w-I_{h,h}(\eta_{T}^{s,n}w))\|_{0,T}\Big{)}
\displaystyle\lesssim A12(ηTs,nw)0,Ω22=R1.\displaystyle\|A^{\frac{1}{2}}\nabla(\eta_{T}^{s,n}w)\|^{2}_{0,\Omega_{2}}=\mathrm{R}_{1}.

For R3\mathrm{R}_{3}, it is easy to see

R3eΓh,eTnγ0hwe2.\mathrm{R}_{3}\lesssim\sum_{\begin{subarray}{c}e\in\Gamma_{h},e\subset T_{n}\end{subarray}}\frac{\gamma_{0}}{h}\|w\|^{2}_{e}.

Combining the above estimates of R1,R2\mathrm{R}_{1},\mathrm{R}_{2} and R3\mathrm{R}_{3}, we have

Ih,h(ηTs,nw)h,h2wh,h,Tn+12,\|I_{h,h}(\eta_{T}^{s,n}w)\|^{2}_{h,h}\lesssim\|w\|^{2}_{h,h,T_{n+1}},

which yields (4.11) immediately.

Next, we give the proof of the (4.12). Noting that wW0,hw\in W_{0,h}, and ηTs,n|Ts1\eta_{T}^{s,n}|_{T_{s}}\equiv 1, ηTs,n|Ω\Tn0\eta_{T}^{s,n}|_{\Omega\backslash T_{n}}\equiv 0, it is obvious that

ηTs,nwIh,h(ηTs,nw)h,h2=A12(ηTs,nwIh,h(ηTs,nw))0,Tn\Ts2+eΓhγ0hηTs,nwIh,h(ηTs,nw)e2:=I1+I2.\begin{split}\|\eta_{T}^{s,n}w-I_{h,h}(\eta_{T}^{s,n}w)\|^{2}_{h,h}=&\|A^{\frac{1}{2}}\nabla(\eta_{T}^{s,n}w-I_{h,h}(\eta_{T}^{s,n}w))\|^{2}_{0,T_{n}\backslash T_{s}}\\ &+\sum\limits_{e\in\Gamma_{h}}\frac{\gamma_{0}}{h}\|\eta_{T}^{s,n}w-I_{h,h}(\eta_{T}^{s,n}w)\|^{2}_{e}:=\mathrm{I}_{1}+\mathrm{I}_{2}.\end{split}

Similar to the estimate of R2\mathrm{R}_{2}, from Lemma 4.1, it follows that

I2I1.\mathrm{I}_{2}\lesssim\mathrm{I}_{1}.

Further, by a similar argument to (B) and using (4.3) and (4.10), we have

I1\displaystyle\mathrm{I}_{1} A12(ηTs,nw)0,Tn\Ts2\displaystyle\lesssim\|A^{\frac{1}{2}}\nabla(\eta_{T}^{s,n}w)\|^{2}_{0,T_{n}\backslash T_{s}}
TTn\Ts(w𝒞h,Hw)ηTs,n0,T2+A12w0,Tn\Ts2\displaystyle\lesssim\sum\limits_{T\in T_{n}\backslash T_{s}}\|(w-\mathcal{C}_{h,H}w)\nabla\eta_{T}^{s,n}\|^{2}_{0,T}+\|A^{\frac{1}{2}}\nabla w\|^{2}_{0,T_{n}\backslash T_{s}}
HηTs,nL(Ω)2w0,Tn+1\Ts12+A12w0,Tn\Ts2\displaystyle\lesssim\|H\nabla\eta_{T}^{s,n}\|^{2}_{L^{\infty}(\Omega)}\|\nabla w\|^{2}_{0,T_{n+1}\backslash T_{s-1}}+\|A^{\frac{1}{2}}\nabla w\|^{2}_{0,T_{n}\backslash T_{s}}
A12w0,Tn+1\Ts12,\displaystyle\lesssim\|A^{\frac{1}{2}}\nabla w\|^{2}_{0,T_{n+1}\backslash T_{s-1}},

which follows (4.12) immediately. The proof of (4.13) and (4.14) is similar to the above inequality. ∎

Appendix C Proof of Lemma 4.7

The proof is divided into four steps.

Step   1. In this step we prove the following estimate for L5L\geq 5:

qhTqhT,Lh,hqhTh,h,Ω\TL5.\|q^{T}_{h}-q_{h}^{T,L}\|_{h,h}\lesssim\|q^{T}_{h}\|_{h,h,\Omega\backslash T_{L-5}}. (C.1)

From (3.12) and (3.14), qhTW0,hq^{T}_{h}\in W_{0,h} and qhT,LW0,h(TL)q_{h}^{T,L}\in W_{0,h}(T_{L}) satisfy

aΩ(qhT,w)=aT~(uh,H,w)wW0,h,\displaystyle a_{\Omega}(q^{T}_{h},w)=a_{\tilde{T}}(u_{h,H},w)\quad\forall\,w\in W_{0,h},
aΩ(qhT,L,w)=aT~(uh,H,w)wW0,h(TL).\displaystyle a_{\Omega}(q_{h}^{T,L},w)=a_{\tilde{T}}(u_{h,H},w)\quad\forall\,w\in W_{0,h}(T_{L}).

Subtracting the above two equations yields

aΩ(qhTqhT,L,w)=0wW0,h(TL).a_{\Omega}(q^{T}_{h}-q_{h}^{T,L},w)=0\quad\forall\,w\in W_{0,h}(T_{L}). (C.2)

Denote by e:=qhTqhT,Le:=q^{T}_{h}-q_{h}^{T,L}. Using the coercivity and continuity of aΩa_{\Omega}, from (C.2), for any wW0,h(TL)w\in W_{0,h}(T_{L}) it follows that

eh,h2\displaystyle\|e\|^{2}_{h,h} aΩ(e,qhTqhT,L)=aΩ(e,qhTw)\displaystyle\lesssim a_{\Omega}(e,q^{T}_{h}-q_{h}^{T,L})=a_{\Omega}(e,q^{T}_{h}-w)
eh,hqhTwh,h,\displaystyle\lesssim\|e\|_{h,h}\|q^{T}_{h}-w\|_{h,h},

which yields

qhTqhT,Lh,hinfwW0,h(TL)qhTwh,h.\|q^{T}_{h}-q_{h}^{T,L}\|_{h,h}\lesssim{\inf_{w\in W_{0,h}(T_{L})}}\|q^{T}_{h}-w\|_{h,h}. (C.3)

For the element TT, let ηTL3,L2\eta_{T}^{L-3,L-2} be the cut off function defined in (4.8)–(4.10) (with l1=L3,l2=L2l_{1}=L-3,l_{2}=L-2). Since ηTL3,L21\eta_{T}^{L-3,L-2}\equiv 1 on TL3T_{L-3} and ηTL3,L20\eta_{T}^{L-3,L-2}\equiv 0 on Ω\TL2\Omega\backslash T_{L-2}, it is easy to check that:

𝒞h,HIh,h(ηTL3,L2qhT)=𝒞h,HqhT=0on TL4,\displaystyle\mathcal{C}_{h,H}I_{h,h}(\eta_{T}^{L-3,L-2}q^{T}_{h})=\mathcal{C}_{h,H}q^{T}_{h}=0\quad\text{on }T_{L-4},

and hence

supp(𝒞h,HIh,h(ηTL3,L2qhT))\displaystyle{\mathrm{supp}}(\mathcal{C}_{h,H}I_{h,h}(\eta_{T}^{L-3,L-2}q^{T}_{h})) TL1\TL4,\displaystyle\subseteq T_{L-1}\backslash T_{L-4}, (C.4)
supp(𝒞h,H2Ih,h(ηTL3,L2qhT))\displaystyle{\mathrm{supp}}(\mathcal{C}_{h,H}^{2}I_{h,h}(\eta_{T}^{L-3,L-2}q^{T}_{h})) TL\TL5.\displaystyle\subseteq T_{L}\backslash T_{L-5}. (C.5)

Using Lemma 4.5, for 𝒞h,HIh,h(ηTL3,L2qhT)\mathcal{C}_{h,H}I_{h,h}(\eta_{T}^{L-3,L-2}q^{T}_{h}), there is a μV0,h\mu\in V_{0,h} such that

𝒞h,Hμ=𝒞h,HIh,h(ηTL3,L2qhT),\displaystyle\mathcal{C}_{h,H}\mu=\mathcal{C}_{h,H}I_{h,h}(\eta_{T}^{L-3,L-2}q^{T}_{h}),\quad μh,h𝒞h,HIh,h(ηTL3,L2qhT)h,H,\displaystyle{\|\mu\|_{h,h}\lesssim\|\mathcal{C}_{h,H}I_{h,h}(\eta_{T}^{L-3,L-2}q^{T}_{h})\|_{h,H},}
andsupp(μ)TL\TL5.\displaystyle\text{and}\quad{\mathrm{supp}}(\mu)\subseteq T_{L}\backslash T_{L-5}.

Further, using (C.4), Lemmas 4.3 and 4.6, we have

μh,h𝒞h,HIh,h(ηTL3,L2qhT)h,H,TL1\TL4Ih,h(ηTL3,L2qhT)h,h,TL\TL5=Ih,h(ηTL3,L2qhT)h,h,TL2\TL3+qhTh,h,TL3\TL5qhTh,h,TL1\TL5.\begin{split}\|\mu\|_{h,h}&\lesssim\|\mathcal{C}_{h,H}I_{h,h}(\eta_{T}^{L-3,L-2}q^{T}_{h})\|_{h,H,T_{L-1}\backslash T_{L-4}}\\ &\lesssim\|I_{h,h}(\eta_{T}^{L-3,L-2}q^{T}_{h})\|_{h,h,T_{L}\backslash T_{L-5}}\\ &=\|I_{h,h}(\eta_{T}^{L-3,L-2}q^{T}_{h})\|_{h,h,T_{{L-2}}\backslash T_{L-3}}+\|q^{T}_{h}\|_{h,h,T_{L-3}\backslash T_{L-5}}\\ &\lesssim\|q^{T}_{h}\|_{h,h,T_{L-1}\backslash T_{L-5}}.\end{split} (C.6)

Hence taking w=Ih,h(ηTL3,L2qhT)μW0,h(TL)w=I_{h,h}(\eta_{T}^{L-3,L-2}q^{T}_{h})-\mu\in W_{0,h}(T_{L}) in (C.3) and using (C.6) and Lemma 4.6, we have

qhTqhT,Lh,h\displaystyle\|q^{T}_{h}-q_{h}^{T,L}\|_{h,h} Ih,h(1ηTL3,L2)qhTh,h+μh,h\displaystyle\lesssim\|I_{h,h}(1-\eta_{T}^{L-3,L-2})q^{T}_{h}\|_{h,h}+\|\mu\|_{h,h}
qhTh,h,Ω\TL4+qhTh,h,TL1\TL5,\displaystyle\lesssim\|q^{T}_{h}\|_{h,h,\Omega\backslash T_{L-4}}+\|q^{T}_{h}\|_{h,h,T_{L-1}\backslash T_{L-5}},

which implies that (C.1) holds.

Step   2. Suppose we can prove the following recursive inequality

qhTh,h,Ω\TMθ0qhTh,h,Ω\Tmm=M50,\|q^{T}_{h}\|_{h,h,\Omega\backslash T_{M}}\leq\theta_{0}\|q^{T}_{h}\|_{h,h,\Omega\backslash T_{m}}\quad\forall\,m=M-5\geq 0, (C.7)

where 0<θ0<10<\theta_{0}<1 is a constant independent of MM and qhTq^{T}_{h}.

For L=5k+jL=5k+j with integers k1k\geq 1 and 0j40\leq j\leq 4, setting θ=θ015\theta=\theta_{0}^{\frac{1}{5}} and using (C.7) repeatedly, we conclude that

qhTh,h,Ω\TL5\displaystyle\|q^{T}_{h}\|_{h,h,\Omega\backslash T_{L-5}} θ0k1qhTh,h,Ω\TjθLj5qhTh,h\displaystyle\lesssim\theta_{0}^{k-1}\|q^{T}_{h}\|_{h,h,\Omega\backslash T_{j}}\lesssim\theta^{L-j-5}\|q^{T}_{h}\|_{h,h}
θLqhTh,h.\displaystyle\lesssim\theta^{L}\|q^{T}_{h}\|_{h,h}. (C.8)

Clearly, the above estimate also holds for 5L95\leq L\leq 9 and hence (C.8) holds for L5L\geq 5.

Step   3. In this step we prove (C.7). Let ε=1ηTm+2,M2\varepsilon=1-\eta_{T}^{m+2,M-2} satisfying ε1\varepsilon\equiv 1 in Ω\TM2\Omega\backslash T_{M-2}, ε0\varepsilon\equiv 0 in Tm+2T_{m+2}, and 0<ε<10<\varepsilon<1 otherwise. It is easy to see that

qhTh,h,Ω\TM2εqhTh,h2aΩ(εqhT,εqhT)=aΩ(qhT,ε2qhT)+(AhεqhT,hεqhT).\begin{split}\|q^{T}_{h}\|^{2}_{h,h,\Omega\backslash T_{M}}&\leq\|\varepsilon q^{T}_{h}\|^{2}_{h,h}\lesssim a_{\Omega}(\varepsilon q^{T}_{h},\varepsilon q^{T}_{h})\\ &=a_{\Omega}(q^{T}_{h},\varepsilon^{2}q^{T}_{h})+(A{\nabla_{h}}\varepsilon\cdot q^{T}_{h},\nabla_{h}\varepsilon\cdot q^{T}_{h}).\end{split} (C.9)

For the function 𝒞h,HIh,h(ε2qhT)\mathcal{C}_{h,H}I_{h,h}(\varepsilon^{2}q^{T}_{h}), using Lemma 4.5, there exists a γV0,h\gamma\in V_{0,h} such that 𝒞h,Hγ=𝒞h,HIh,h(ε2qhT)\mathcal{C}_{h,H}\gamma=\mathcal{C}_{h,H}I_{h,h}(\varepsilon^{2}q^{T}_{h}), and

supp(𝒞h,HIh,h(ε2qhT))TM1\Tm+1,\displaystyle{\mathrm{supp}}(\mathcal{C}_{h,H}I_{h,h}(\varepsilon^{2}q^{T}_{h}))\subseteq T_{M-1}\backslash T_{m+1},
supp(γ)supp(𝒞h,H2Ih,h(ε2qhT))TM\Tm.\displaystyle{\mathrm{supp}}(\gamma)\subseteq{\mathrm{supp}}(\mathcal{C}_{h,H}^{2}I_{h,h}(\varepsilon^{2}q^{T}_{h}))\subseteq T_{M}\backslash T_{m}.

In addition, it holds that

γh,h𝒞h,HIh,h(ε2qhT)h,H.\|\gamma\|_{h,h}\lesssim\|\mathcal{C}_{h,H}I_{h,h}(\varepsilon^{2}q^{T}_{h})\|_{h,H}. (C.10)

Since Ih,h(ε2qhT)γW0,h(Ω2\Tm)I_{h,h}(\varepsilon^{2}q^{T}_{h})-\gamma\in W_{0,h}(\Omega_{2}\backslash T_{m}), from (3.12), it follows that

aΩ(qhT,Ih,h(ε2qhT)γ)=aT~(uh,H,Ih,h(ε2qhT)γ)=0,a_{\Omega}(q^{T}_{h},I_{h,h}(\varepsilon^{2}q^{T}_{h})-\gamma)=a_{\tilde{T}}(u_{h,H},I_{h,h}(\varepsilon^{2}q^{T}_{h})-\gamma)=0,

which combines (C.9) yields

qhTh,h,Ω\TM2\displaystyle\|q^{T}_{h}\|^{2}_{h,h,\Omega\backslash T_{M}} aΩ(qhT,ε2qhTIh,h(ε2qhT))+aΩ(qhT,γ)\displaystyle\lesssim a_{\Omega}(q^{T}_{h},\varepsilon^{2}q^{T}_{h}-I_{h,h}(\varepsilon^{2}q^{T}_{h}))+a_{\Omega}(q^{T}_{h},\gamma)
+(AεqhT,εqhT):=T1+T2+T3.\displaystyle\quad+(A\nabla\varepsilon\cdot q^{T}_{h},\nabla\varepsilon\cdot q^{T}_{h}):=\mathrm{T}_{1}+\mathrm{T}_{2}+\mathrm{T}_{3}.

Using the same argument as that of (4.12), we obtain

T1\displaystyle\mathrm{T}_{1} qhTh,h,TM2\Tm+2ε2qhTIh,h(ε2qhT)h,h,TM2\Tm+2\displaystyle\lesssim\|q^{T}_{h}\|_{h,h,T_{M-2}\backslash T_{m+2}}\|\varepsilon^{2}q^{T}_{h}-I_{h,h}(\varepsilon^{2}q^{T}_{h})\|_{h,h,T_{M-2}\backslash T_{m+2}}
qhTh,h,TM1\Tm+12.\displaystyle\lesssim\|q^{T}_{h}\|^{2}_{h,h,T_{M-1}\backslash T_{m+1}}.

Further, for T2\mathrm{T}_{2}, using the same argument as that of (4.11), from (C.10), we have

T2\displaystyle\mathrm{T}_{2} qhTh,h,TM\Tmγh,h\displaystyle\lesssim\|q^{T}_{h}\|_{h,h,T_{M}\backslash T_{m}}\|\gamma\|_{h,h}
qhTh,h,TM\Tm𝒞h,HIh,h(ε2qhT)h,H,TM1\Tm+1\displaystyle\lesssim\|q^{T}_{h}\|_{h,h,T_{M}\backslash T_{m}}\|\mathcal{C}_{h,H}I_{h,h}(\varepsilon^{2}q^{T}_{h})\|_{h,H,T_{M-1}\backslash T_{m+1}}
qhTh,h,TM\TmIh,h(ε2qhT)h,h,TM\Tm\displaystyle\lesssim\|q^{T}_{h}\|_{h,h,T_{M}\backslash T_{m}}\|I_{h,h}(\varepsilon^{2}q^{T}_{h})\|_{h,h,T_{M}\backslash T_{m}}
=qhTh,h,TM\Tm(qhTh,h,TM\TM2+Ih,h(ε2qhT)h,h,TM2\Tm+2)\displaystyle=\|q^{T}_{h}\|_{h,h,T_{M}\backslash T_{m}}\left(\|q^{T}_{h}\|_{h,h,T_{M}\backslash T_{M-2}}+\|I_{h,h}(\varepsilon^{2}q^{T}_{h})\|_{h,h,T_{M-2}\backslash T_{m+2}}\right)
qhTh,h,TM\Tm2.\displaystyle\lesssim\|q^{T}_{h}\|^{2}_{h,h,T_{M}\backslash T_{m}}.

To estimate T3\mathrm{T}_{3}, by use of the assumption (4.8)–(4.10) and (4.3), it follows that

T3\displaystyle\mathrm{T}_{3} =A12εqhT0,Ω2=TH,Ω2A12εqhT0,T2\displaystyle=\|A^{\frac{1}{2}}\nabla\varepsilon\cdot q^{T}_{h}\|^{2}_{0,\Omega}=\sum\limits_{T\in\mathcal{M}_{H,\Omega_{2}}}\|A^{\frac{1}{2}}\nabla\varepsilon\cdot q^{T}_{h}\|^{2}_{0,T}
TTM2\Tm+2εL(Ω)2qhT𝒞HqhT0,T2\displaystyle\lesssim\sum\limits_{T\in T_{M-2}\backslash T_{m+2}}\|\nabla\varepsilon\|_{L^{\infty}(\Omega)}^{2}\|q^{T}_{h}-\mathcal{C}_{H}q^{T}_{h}\|^{2}_{0,T}
qhTh,h,TM1\Tm+12.\displaystyle\lesssim\|q^{T}_{h}\|^{2}_{h,h,T_{M-1}\backslash T_{m+1}}.

Thus, by using the above estimates of T1,T2,\mathrm{T}_{1},\mathrm{T}_{2}, and T3\mathrm{T}_{3}, we have, for some positive constant C0C_{0},

qhTh,h,Ω\TM2\displaystyle\|q^{T}_{h}\|^{2}_{h,h,\Omega\backslash T_{M}} C0qhTh,h,TM\Tm2=C0qhTh,h,Ω\Tm2C0qhTh,h,Ω\TM2,\displaystyle\leq C_{0}\|q^{T}_{h}\|^{2}_{h,h,T_{M}\backslash T_{m}}=C_{0}\|q^{T}_{h}\|^{2}_{h,h,\Omega\backslash T_{m}}-C_{0}\|q^{T}_{h}\|^{2}_{h,h,\Omega\backslash T_{M}},

which implies that (C.7) holds with θ0:=(C0C0+1)12.\theta_{0}:=\big{(}\frac{C_{0}}{C_{0}+1}\big{)}^{\frac{1}{2}}.

Step   4. Next, we estimate qhTh,h2\|q^{T}_{h}\|_{h,h}^{2} in (C.8).

qhTh,h2\displaystyle\|q^{T}_{h}\|_{h,h}^{2} aΩ(qhT,qhT)=aT~(uh,H,qhT)uh,Hh,h,T~qhTh,h,\displaystyle\lesssim a_{\Omega}(q^{T}_{h},q^{T}_{h})=a_{\tilde{T}}(u_{h,H},q^{T}_{h})\lesssim\|u_{h,H}\|_{h,h,\tilde{T}}\|q^{T}_{h}\|_{h,h},

where T~\tilde{T} is defined in Section 3.4, which together with (C.1) and (C.8) completes the proof of the lemma. ∎

References

References

  • [1] I. Babuška, G. Caloz, J. Osborn, Special finite element methods for a class of second order elliptic problems with rough coefficients, SIAM J. Numer. Anal. 31 (1994) 945–981.
  • [2] I. Babuška, U. Banerjee, J. E. Osborn, Survey of meshless and generalized finite element methods: a unified approach, Acta Numer. 12 (2003) 1–125.
  • [3] T. Strouboulis, I. Babuška, K. Copps, The design and analysis of the generalized finite element method, Comput. Methods Appl. Mech. Engrg. 181 (1-3) (2000) 43–69.
  • [4] T. Y. Hou, X. H. Wu, A multiscale finite element method for elliptic problems in composite materials and porous media, J. Comput. Phys. 134 (1997) 169–189.
  • [5] Y. Efendiev, T. Y. Hou, X. H. Wu, Convergence of a nonconforming multiscale finite element method, SIAM J. Numer. Anal. 37 (2000) 888–910.
  • [6] Z. Chen, T. Y. Hou, A mixed multisclae finite method for elliptic problems with oscillating coefficients, Math. Comp. 72 (2002) 541–576.
  • [7] T. Hughes, Multiscale phenomena: Green’s functions, the dirichlet to neumann formulation, subgrid scale models, bubbles and the origin of stabilized methods, Comput. Meth. Appl. Mech. Eng. 127 (1995) 387–401.
  • [8] T. J. R. Hughes, G. R. Feijóo, L. Mazzei, J.-B. Quincy, The variational multiscale method—a paradigm for computational mechanics, Comput. Methods Appl. Mech. Engrg. 166 (1-2) (1998) 3–24.
  • [9] F. Brezzi, L. P. Franca, T. J. R. Hughes, A. Russo, b=gb=\int g, Comput. Methods Appl. Mech. Engrg. 145 (1997) 329–339.
  • [10] J. Fish, Z. Yuan, Multiscale enrichment based on partition of unity, Inter. J. Numer. Meth. Eng. 62 (2005) 1341–1359.
  • [11] W. E, B. Engquist, The heterogeneous multiscale methods, Commun. Math. Sci. 1 (2003) 87–132.
  • [12] W. E, P. Ming, P. Zhang, Analysis of the heterogeneous multiscale method for elliptic homogenization problems, J. Am. Math. Soc. 18 (2005) 121–156.
  • [13] A. Abdulle, W. E, B. Engquist, E. Vanden-Eijnden, The heterogeneous multiscale method, Acta Numer. 21 (2012) 1–87.
  • [14] P. Jenny, S. Lee, H. Tchelepi, Multi-scale finite-volume method for elliptic problems in subsurface flow simulation, J. Comput. Phys. 187 (1) (2003) 47–67.
  • [15] J. Fish, V. Belsky, Multigrid method for periodic heterogeneous media. I. Convergence studies for one-dimensional case, Comput. Methods Appl. Mech. Engrg. 126 (1-2) (1995) 1–16.
  • [16] H. Owhadi, Multigrid with rough coefficients and multiresolution operator decomposition from hierarchical information games, SIAM Rev. 59 (1) (2017) 99–149.
  • [17] M. F. Wheeler, Mortar upscaling for multiphase flow in porous media, Computational Geoences 6 (1) (2002) 73–100.
  • [18] T. Arbogast, G. Pencheva, M. F. Wheeler, I. Yotov, A multiscale mortar mixed finite element method, Multiscale Model. Simul. 6 (1) (2007) 319–346.
  • [19] A. Målqvist, D. Peterseim, Localization of elliptic multiscale problems, Math. Comp. 83 (290) (2014) 2583–2603.
  • [20] P. Henning, D. Peterseim, Oversampling for the multiscale finite element method, Multiscale Model. Simul. 11 (4) (2013) 1149–1175.
  • [21] P. Henning, A. Målqvist, Localized orthogonal decomposition techniques for boundary value problems, SIAM J. Sci. Comput. 36 (4) (2014) A1609–A1634.
  • [22] A. J. Roberts, I. G. Kevrekidis, General tooth boundary conditions for equation free modeling, SIAM J. Sci. Comput. 29 (4) (2007) 1495–1510.
  • [23] A. Papavasiliou, I. G. Kevrekidis, Variance reduction for the equation-free simulation of multiscale stochastic systems, Multiscale Model. Simul. 6 (1) (2007) 70–89.
  • [24] Y. Efendiev, J. Galvis, T. Y. Hou, Generalized multiscale finite element methods (GMsFEM), J. Comput. Phys. 251 (2013) 116–135.
  • [25] E. T. Chung, Y. Efendiev, T. Y. Hou, Adaptive multiscale model reduction with generalized multiscale finite element methods, J. Comput. Phys. 320 (2016) 69–95.
  • [26] I. Babuška, R. Lipton, P. Sinz, M. Stuebner, Multiscale-spectral GFEM and optimal oversampling, Comput. Methods Appl. Mech. Engrg. 364 (2020) 112960, 28.
  • [27] I. Babuska, R. Lipton, Optimal local approximation spaces for generalized finite element methods with application to multiscale problems, Multiscale Model. Simul. 9 (1) (2011) 373–406.
  • [28] E. T. Chung, Y. Efendiev, W. T. Leung, Constraint energy minimizing generalized multiscale finite element method, Comput. Methods Appl. Mech. Engrg. 339 (2018) 298–319.
  • [29] E. T. Chung, Y. Efendiev, W. T. Leung, Generalized multiscale finite element methods with energy minimizing oversampling, Internat. J. Numer. Methods Engrg. 117 (3) (2019) 316–343.
  • [30] M. Li, E. T. Chung, L. Jiang, A constraint energy minimizing generalized multiscale finite element method for parabolic equations, Multiscale Model. Simul. 17 (3) (2019) 996–1018.
  • [31] H. Owhadi, L. Zhang, Metric-based upscaling, Comm. Pure Appl. Math. 60 (5) (2007) 675–723.
  • [32] H. Owhadi, L. Zhang, Localized bases for finite-dimensional homogenization approximations with nonseparated scales and high contrast, Multiscale Model. Simul. 9 (2011) 1373–1398.
  • [33] L. Durlofsky, Numerical calculation of equivalent grid block permeability tensors for heterogeneous porous media, Water Resources Research 27 (1991) 699–708.
  • [34] T. Arbogast, Numerical subgrid upscaling of two-phase flow in porous media, in: Z. Chen, R. E. Ewing, Z. C. Shi (Eds.), Numerical Treatment of Multiphase Flows in Porous Media, Vol. 552 of Lect. Notes Phys., Springer-Verlag, New York, 2000, pp. 35–49.
  • [35] X. H. Wu, Y. Efendiev, T. Y. Hou, Analysis of upscaling absolute permeability, Discrete and Continuous Dynamical Systems-series B (2002) 185–204.
  • [36] Y. Efendiev, J. Galvis, X. H. Wu, Multiscale finite element methods for high-contrast problems using local spectral basis functions, J. Comput. Phys. 230 (4) (2011) 937–955.
  • [37] J. Galvis, Y. Efendiev, Domain decomposition preconditioners for multiscale flows in high-contrast media, Multiscale Model. Simul. 8 (2010) 1461–1483.
  • [38] J. Galvis, Y. Efendiev, Domain decomposition preconditioners for multiscale flows in high-contrast media: Reduced dimension coarse spaces, Multiscale Model. Simul. 8 (2010) 1621–1644.
  • [39] Z. Chen, X. Y. Yue, Numerical homogenization of well singularities in the flow transport through heterogeneous porous media, Multiscale Model. Simul. 1 (2003) 260–303.
  • [40] E. T. Chung, Y. Efendiev, G. Li, An adaptive GMsFEM for high-contrast flow problems, J. Comput. Phys. 273 (2014) 54–73.
  • [41] E. T. Chung, Y. Efendiev, W. T. Leung, An adaptive generalized multiscale discontinuous Galerkin method for high-contrast flow problems, Multiscale Model. Simul. 16 (3) (2018) 1227–1257.
  • [42] C. C. Chu, I. G. Graham, T. Y. Hou, A new multiscale finite element method for high-contrast elliptic interface problems, Math. Comp. 79 (272) (2010) 1915–1955.
  • [43] D. Elfverson, M. G. Larson, A. Målqvist, Multiscale methods for problems with complex geometry, Comput. Methods Appl. Mech. Engrg. 321 (2017) 103–123.
  • [44] F. Hellman, A. Målqvist, Contrast independent localization of multiscale problems, Multiscale Model. Simul. 15 (4) (2017) 1325–1355.
  • [45] W. Deng, H. Wu, A combined finite element and multiscale finite element method for the multisclae elliptic problems, Multiscale Model. Simul. 12 (4) (2014) 1424–1457.
  • [46] F. Song, W. Deng, H. Wu, A combined finite element and oversampling multiscale Petrov-Galerkin method for the multisclae elliptic problems with singularities, J. Comput. Phys. 305 (2016) 722–743.
  • [47] D. Elfverson, E. H. Georgoulis, A. Målqvist, D. Peterseim, Convergence of a discontinuous Galerkin multiscale method, SIAM J. Numer. Anal. 51 (6) (2013) 3351–3372.
  • [48] D. Elfverson, V. Ginting, P. Henning, On multiscale methods in petrov-galerkin formulation, Numerische Mathematik (2015) 1–40.
  • [49] P. Henning, P. Morgenstern, D. Peterseim, Multiscale partition of unity, in: Meshfree methods for partial differential equations VII, Vol. 100 of Lect. Notes Comput. Sci. Eng., Springer, Cham, 2015, pp. 185–204.
  • [50] P. Henning, A. Målqvist, D. Peterseim, A localized orthogonal decomposition method for semi-linear elliptic problems, ESAIM Math. Model. Numer. Anal 48 (5) (2014) 1331–1349.
  • [51] A. Målqvist, D. Peterseim, Computation of eigenvalues by numerical upscaling, Numer. Math. 130 (2) (2014) 337–361.
  • [52] P. Henning, A. Målqvist, D. Peterseim, Two-level discretization techniques for ground state computations of Bose-Einstein condensates, SIAM J. Numer. Anal. 52 (4) (2014) 1525–1550.
  • [53] C. Engwer, P. Henning, A. Målqvist, D. Peterseim, Efficient implementation of the localized orthogonal decomposition method, Comput. Methods Appl. Mech. Engrg. 350 (2019) 123–153.
  • [54] D. Peterseim, Variational multiscale stabilization and the exponential decay of fine-scale correctors, in: Building bridges: connections and challenges in modern approaches to numerical partial differential equations, Vol. 114 of Lect. Notes Comput. Sci. Eng., Springer, [Cham], 2016, pp. 341–367.
  • [55] H. Cao, H. Wu, IPCDGM and multiscale IPDPGM for the Helmholtz problem with large wave number, J. Comp. Appl. Math. 369 (2020) 112590.
  • [56] S. C. Brenner, L. R. Scott, The mathematical theory of finite element methods, Springer-Verlag, New York, 2002.
  • [57] D. Arnold, An interior penalty finite element method with discontinuous elements, SIAM J. Numer. Anal. 19 (1982) 742–760.
  • [58] C. Carstensen, Quasi-interpolation and a posteriori error analysis in finite element methods, M2AN Math. Model. Numer. Anal. 33 (1999) 1187–1202.
  • [59] C. Carstensen, R. Verfürth, Edge residuals dominate a posteriori error estimates for low order finite element methods, SIAM J. Numer. Anal 36 (1999) 1571–1587.
  • [60] P. Ciarlet, The Finite Element Method for Elliptic Problems, Amsterdam, North Holland, 1978.
  • [61] L. Durlofsky, Numerical calculation of equivalent grid block permeability tensors for heterogeneous porous media, Water Resources Res. 27 (1991) 699–708.
  • [62] D. W. Peaceman, Interpretation of well-block pressures in numerical reservoir simulations with non-square grid blocks and anisotropic permeability, Soc. Pet. Eng. J. 23 (1983) 531–543.