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

An Efficient Online-Offline Method for Elliptic Homogenization Problems

Yufang Huang LSEC, Institute of Computational Mathematics and Scientific/Engineering Computing, AMSS, Chinese Academy of Sciences, No. 55, East Road Zhong-Guan-Cun, Beijing 100190, China
and School of Mathematical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China
Cornell University, 425 East 61st Street, New York, 10065, USA [email protected]
Pingbing Ming LSEC, Institute of Computational Mathematics and Scientific/Engineering Computing, AMSS, Chinese Academy of Sciences, No. 55, East Road Zhong-Guan-Cun, Beijing 100190, China
and School of Mathematical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China
[email protected]
 and  Siqi Song LSEC, Institute of Computational Mathematics and Scientific/Engineering Computing, AMSS, Chinese Academy of Sciences, No. 55, East Road Zhong-Guan-Cun, Beijing 100190, China
and School of Mathematical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China
[email protected]
Abstract.

We present a new numerical method for solving the elliptic homogenization problem. The main idea is that the missing effective matrix is reconstructed by solving the local least-squares in an offline stage, which shall be served as the input data for the online computation. The accuracy of the proposed method are analyzed with the aid of the refined estimates of the reconstruction operator. Two dimensional and three dimensional numerical tests confirm the efficiency of the proposed method, and illustrate that this online-offline strategy may significantly reduce the cost without loss of accuracy.

Key words and phrases:
Numerical homogenization, Heterogeneous multiscale method, Online-offline computing, Least-squares reconstruction, Stability
2000 Mathematics Subject Classification:
35B27, 65N30, 74S05, 74Q05
The second author would like to thank Professor Ruo Li for his helpful discussion in the earlier stage of the present work. P. B. Ming and S. Q. Song are supported by National Natural Science Foundation of China through Grant No. 11971467 and Beijing Academy of Artificial Intelligence (BAAI).

1. Introduction

We consider a prototypical elliptic boundary value problem

(1.1) {div(aε(x)uε(x))=f(x),xDd,uε(x)=0,xD,\left\{\begin{aligned} -\operatorname{div}(a^{\;\varepsilon}(x)\nabla u^{\,\varepsilon}(x))&=f(x),\qquad&&x\in D\subset\mathbb{R}^{d},\\ u^{\,\varepsilon}(x)&=0,\qquad&&x\in\partial D,\\ \end{aligned}\right.

where ε\varepsilon is a small parameter that signifies explicitly the multiscale nature of the problem. We assume that the coefficient aεa^{\;\varepsilon}, which is not necessarily symmetric, belongs to a set (α,β,D)\mathcal{M}(\alpha,\beta,D) that is defined by

(1.2) (α,β,D):={[L(D)]d2|\displaystyle\mathcal{M}(\alpha,\beta,D){:}=\{\mathcal{B}\in[L^{\infty}(D)]^{d^{2}}| (ξ,ξ)α|ξ|2,|(x)ξ|β|ξ|,\displaystyle(\mathcal{B}\xi,\xi)\geq\alpha\lvert\xi\rvert^{2},\lvert\mathcal{B}(x)\xi\rvert\leq\beta\lvert\xi\rvert,
for any ξd and a.e. xD},\displaystyle\text{ for any }\xi\in\mathbb{R}^{d}\text{ and a.e. }x\in D\},

where DD is a bounded domain in d\mathbb{R}^{d} and (,)(\cdot,\cdot) denotes the inner product in d\mathbb{R}^{d} , while ||\lvert\cdot\rvert is the corresponding norm.

In the sense of H-convergence [53, 54], for every sequence aε(α,β,D)a^{\;\varepsilon}\in\mathcal{M}(\alpha,\beta,D) and fH1(D)f\in H^{-1}(D), the sequence uεu^{\,\varepsilon} of the solution to (1.1) satisfies

(1.3) {uεu0 weakly in H01(D),aεuε𝒜u0 weakly in [L2(D)]d,as ε0,\left\{\begin{aligned} u^{\,\varepsilon}\rightharpoonup u_{0}\qquad&\text{ weakly in }H_{0}^{1}(D),\\ a^{\;\varepsilon}\nabla u^{\,\varepsilon}\rightharpoonup\mathcal{A}\nabla u_{0}\qquad&\text{ weakly in }[L^{2}(D)]^{d},\end{aligned}\right.\qquad\text{as\quad}\varepsilon\to 0,

where u0u_{0} is the solution of the homogenization problem

(1.4) {div(𝒜(x)u0(x))=f(x),xD,u0(x)=0,xD,\left\{\begin{aligned} -\operatorname{div}(\mathcal{A}(x)\nabla u_{0}(x))&=f(x),\qquad&&x\in D,\\ u_{0}(x)&=0,\qquad&&x\in\partial D,\end{aligned}\right.

and 𝒜(α,β,D)\mathcal{A}\in\mathcal{M}(\alpha,\beta,D). Here H01(D),L2(D)H_{0}^{1}(D),L^{2}(D) and H1(D)H^{-1}(D) are standard Sobolev spaces [4].

The quantities of interest for Problem (1.1) and Problem (1.4) are the homogenized solution u0u_{0} over the whole domain and the solution uεu^{\,\varepsilon} at certain critical local region. The former stands for the information at the large scale, and the later mimics the information at small scale. There are lots of work devoted to efficiently compute such quantities during the last several decades; see, e.g., [6, 20, 17], among many others. Presently we are interested in the efficient way to compute u0u_{0}. A typical way that towards this is provided by the heterogeneous multiscale method (HMM) [18, 3], and the FE2{}^{2}-method [40] commonly used in the engineering community that is also in the same spirit of HMM. The underlying idea of this approach is to extract 𝒜\mathcal{A} by solving the cell problems posed on the sampling points of the macoscopic solver. At each point, one needs to solve dd cell problems with dd the dimensionality. Therefore, the main computational cost comes from solving all these cell problems. The number of the cell problems grows rapidly when higher-order macroscopic solvers are employed. To reduce the cost, certain nonconventional quadrature schemes were proposed in [16] when finite element method is used as the macroscopic solver. The number of the cell problems reduces to one third compared to the standard mid-point quadrature scheme when 2\mathbb{P}_{2} Lagrange finite element method is employed as the macroscopic solver. Unfortunately, it does not seem easy to extend such idea to even higher order macroscopic solvers because the quadrature nodes tend to accumulate in the interior of the element [51, 48, 50].

In [35], the authors presented a local least-squares reconstruction of the effective matrix using the solution of the cell problems posed on the vertices of the triangulation, which was dubbed as HMM-LS. The total number of the cell problems equals to the total number of the interior vertices of the triangulation, which is of 𝒪(hd)\mathcal{O}(h^{-d}) with hh the mesh size of the macroscopic solver. This method achieves higher-order accuracy with almost the same cost of HMM with 1\mathbb{P}_{1} Lagrange finite element method as the macroscopic solver [19]. A drawback of this method is that the number of the cell problems is still quite large when mesh refinement is necessary. Moreover, if the adaptive strategy is used in the macroscopic solver, then one has to solve many cell problems around the regions with mesh refinement.

In this work, we propose an offline-online method to compute u0u_{0} efficiently. The main idea is to separate the microscopic solver from the macroscopic solver. In the offline stage, we firstly solve the cell problems posed on a sampling point set and obtain the effective matrix at all these points, then we reconstruct an effective matrix locally by solving the discrete least-squares. The sampling set is constructed from a triangulation of domain, which is usually coarser than the online triangulation. In the online stage, we solve the macroscopic problem with the effective matrix prepared in the offline stage. Such decoupled strategy brings more flexibilities to reduce the number of the cell problems, we may either refine the offline triangulation mesh or increase the reconstruction order, which is guided by the a priori error estimate. The offline computation bears certain similarity with h-p finite element method [7, 47]. Moreover, the offline computation is no longer linked to the macroscopic solvers, which is particularly attractive to higher order macroscopic solver and three dimensional multiscale problems. With the aid of the theoretical results proved in [36, 35], we study the accuracy and the stability of the reconstruction procedure, which is crucial to prove the optimal error estimate of the proposed method. As illustrated by the numerical tests in § 4, the offline-online method converges with optimal order while the cost is smaller than both HMM and HMM-LS method.

The reduced basis HMM proposed in [1, 2] also employed the offline-online idea. The difference between reduced basis HMM and our method lies in the following points: Firstly they employed reduced basis idea in the offline stage while we construct an independent triangulation upon which the cell problems are solved. Secondly they used the empirical interpolation method [8] while we resort to a local least-squares to reconstruct the effective matrix. Finally, a thorough analysis of the least-squares reconstruction is conducted in our work, which concerns the approximation accuracy and the stability of the reconstruction, such rigorous theoretical results put the method on a firm footing. In addition, the analysis of the least-squares reconstruction is of independent interest for other problems such as the construction of the optimal polynomial admissible meshes [13, 45], the discrete norm for polynomials [46], the approximation of the Fekete points [11] and the discontinuous Galerkin method based on patch reconstruction [36, 37].

The rest of the paper is organized as follows. In § 2, we introduce the offline-online method that is based on a local discrete least-squares reconstruction. We derive the optimal error estimate in § 3, in particular, we prove the discrete least-squares is stable with respect to small perturbation. In § 4, we report numerical examples in two and three dimensions, the coefficient aεa^{\;\varepsilon} may be locally periodic, almost periodic and random checker-board. To demonstrate the efficiency of the offline-online method, we compare it with HMM and HMM-LS method. In addition, we solve a problem posed on L-shape domain with nonsmooth solution. The conclusions are drawn in the last section.

Throughout this paper, we shall use the standard notations for the Sobolev spaces, norms and semi-norms, cf., [4], e.g.,

vH1(D):=vL2(D)+vL2(D),|v|Wm,p(D):=|α|=mαvLp(D).\|\,v\,\|_{H^{1}(D)}{:}=\|\,v\,\|_{L^{2}(D)}+\|\,\nabla v\,\|_{L^{2}(D)},\qquad\lvert v\rvert_{W^{m,p}(D)}{:}=\sum_{\lvert\alpha\rvert=m}\|\,\nabla^{\alpha}v\,\|_{L^{p}(D)}.

For any measurable set EE, we define the mean of an integrable function gg over EE as

gE:=1|E|Eg(x)dx.\left\langle\,g\,\right\rangle_{E}{:}=\dfrac{1}{\lvert E\rvert}\int_{E}g(x)\,\mathrm{d}\,x.

We shall also use the discrete p\ell_{p} norm for any xnx\in\mathbb{R}^{n} as

xp:={(i=1n|xi|p)1/p1p<,max1in|xi|p=.\|\,x\,\|_{\ell_{p}}{:}=\begin{cases}\left(\sum_{i=1}^{n}\lvert x_{i}\rvert^{p}\right)^{1/p}&1\leq p<\infty,\\ \max_{1\leq i\leq n}\lvert x_{i}\rvert&p=\infty.\end{cases}

Throughout the paper the generic constant CC may be different from line to line, while it is independent of ε\varepsilon and the mesh size parameters h,Hh,H.

2. The Offline-online Method

The macroscopic solver is chosen as the standard l\mathbb{P}_{l} Lagrange finite element method (l\mathbb{P}_{l} FEM) [14]. l\mathbb{P}_{l} is defined as the set of polynomials with degree less than ll for the sum of all variables. The finite element space is denoted by VhV_{h} corresponding to the triangulation τh\tau_{h} with mesh size hh that is the maximum of the element size hτh_{\tau} for all elements ττh\tau\in\tau_{h}, where hτh_{\tau} is the diameter of τ\tau. We assume that all the elements τ\tau in τh\tau_{h} satisfy the shape-regular condition in the sense of Ciarlet and Raviart [14], i.e., there exists a constant σ0\sigma_{0} such that hτ/ρτσ0h_{\tau}/\rho_{\tau}\leq\sigma_{0}, where ρτ\rho_{\tau} is the diameter of the smallest ball inscribed into τ\tau, and σ0\sigma_{0} is the so-called chunkiness parameter [12].

The method consists of offline part and online part. In the offline part, we approximate the effective matrix 𝒜\mathcal{A} as follows.

Offline We firstly construct a sampling triangulation 𝒯H\mathcal{T}_{H} with mesh size HH over domain DD. For simplicity, we assume that 𝒯H\mathcal{T}_{H} consists of simplices, and 𝒯H\mathcal{T}_{H} is assumed to be shape-regular with the chunkiness parameter σ\sigma. On each element K𝒯HK\in\mathcal{T}_{H}, the approximation effective matrix AHA_{H} is reconstructed by solving a least-squares: for i,j=1,,di,j=1,\cdots,d,

(2.1) (AH)ij=argminpm(S(K))xK(K)|(AH(xK))ijp(xK)|2.(A_{H})_{ij}=\operatorname{arg}\min\limits_{p\in\mathbb{P}_{m}(S(K))}\sum\limits_{x_{K}\in\mathcal{I}(K)}\left\lvert\,(A_{H}(x_{K}))_{ij}-p(x_{K})\,\right\rvert^{2}.

Here (K)\mathcal{I}(K) is the set of all sampling points that belong to S(K)S(K), where S(K)S(K) is a patch of elements around KK, which usually includes KK. Its precise definition will be given later on. We refer to Fig. 1 for an example of such S(K)S(K).

At each sampling point xKx_{K}, the effective matrix AH(xK)A_{H}(x_{K}) is defined by averaging the flux arising from the cell problems:

(2.2) AH(xK)=(aεv1ϵIδ,,aϵvdϵIδ),A_{H}(x_{K})=\left(\left\langle\,a^{\varepsilon}\nabla v_{1}^{\epsilon}\,\right\rangle_{I_{\delta}},\cdots,\left\langle\,a^{\epsilon}\nabla v_{d}^{\epsilon}\,\right\rangle_{I_{\delta}}\right),

where the cell Iδ(xK):=xK+δYI_{\delta}(x_{K}){:}=x_{K}+\delta Y with Y:=(1/2,1/2)dY{:}=(-1/2,1/2)^{d} and δ\delta the cell size. Here for i=1,,di=1,\cdots,d, viεv_{i}^{\varepsilon} satisfies

(2.3) {(aεviε)=0in Iδ(xK),viε=xion Iδ(xK).\left\{\begin{aligned} -\nabla\cdot(a^{\;\varepsilon}\nabla v_{i}^{\varepsilon})=0&\qquad\text{in\quad}I_{\delta}(x_{K}),\\ v_{i}^{\varepsilon}=x_{i}&\qquad\text{on\quad}\partial I_{\delta}(x_{K}).\end{aligned}\right.

Online Given AHA_{H}, we find uhVhu_{h}\in V_{h} such that

(2.4) DAH(x)uhvdx=Df(x)v(x)dxfor all vVh.\int_{D}A_{H}(x)\nabla u_{h}\cdot\nabla v\,\mathrm{d}\,x=\int_{D}f(x)v(x)\,\mathrm{d}\,x\quad\text{for all\quad}v\in V_{h}.
Refer to caption

(a). Example of S2(K)S_{2}(K) constructed by including all the Moore neighbors.

Refer to caption

(b). Example of S(K)S(K) and the set (K)\mathcal{I}(K) consists of the black dots.

Figure 1. Examples of the element patches and the sample set.
Remark.

The element K𝒯HK\in\mathcal{T}_{H} may not be a simplex, which may be polygons or polytopes, the corresponding shape-regular condition and other mesh conditions may be found in [36]. Under these mesh conditions, the properties of the reconstruction are still valid.

In what follows, we supplement some details in the algorithm. The first thing is the construction of the element patch S(K)S(K) for any element K𝒯HK\in\mathcal{T}_{H}. We start from assigning a threshold value NlowestN_{\text{lowest}} that is used to control the size of S(K)S(K). There are several different ways to find S(K)S(K). One way is to define S(K)S(K) in an recursive way as in [35]: For any tt\in\mathbb{N}, we let

(2.5) S0(K):=K,St(K)={K𝒯HK¯St1(K)¯}.S_{0}(K){:}=K,\quad S_{t}(K)=\{\,K\in\mathcal{T}_{H}\,\mid\,\overline{K}\cap\overline{S_{t-1}(K)}\not=\emptyset\,\}.

Once #St(K)Nlowest\#S_{t}(K)\geq N_{\text{lowest}}, we stop the construction and let S(K)=St(K)S(K)=S_{t}(K). This means that we add the Moore neighbors [52] to S(K)S(K) in a recursive way. Another way is using the Von Neumann neighbor [52], i.e., we include the adjacent edge-neighboring elements into the element patch instead of the Moore neighbor. We refer to [36] and [37, Appendix A] for a detailed description for such construction, while S(K)S(K) in all the tests in § 4 are defined as in (2.5). We denote by t(K)\mathcal{I}_{t}(K) the set containing all the sampling points that belong to St(K)S_{t}(K). In all the tests in § 4, we use the barycentric of each element KK as the sampling point. There are also other choices for construction t(K)\mathcal{I}_{t}(K). However, the vertices are not preferred because the communication cost is quite high for three dimensional problems, though this is a good choice for two dimensional problems; cf., [35]. In what follows, we may drop the subscripts tt in St(K)S_{t}(K) and t(K)\mathcal{I}_{t}(K) when there is no confusion may occur.

We remark that there are many other variants for the definitions of the cell problems and the effective matrix in the literatures; see e.g., [59, 24]. The periodic cell problem will be discussed in § 4. As illustrated by the examples in § 4, the overhead caused by the local least-squares is small compared to the computational cost of solving all the cell problems, though the number of the local least-squares is also the same with the cell problems.

3. Convergence of the Method

To study the convergence of the method, we define

e(MOD):=maxxD𝒜(x)AH(x)F,e(\text{MOD}){:}=\max_{x\in D}\|\,\mathcal{A}(x)-A_{H}(x)\,\|_{F},

where BF\|\,B\,\|_{F} is the Frobenius norm of a dd-byd-d matrix BB.

Similar to [35, Lemma 3.1, Lemma 3.2], the following lemma gives the error estimates of the proposed method.

Lemma 3.1.

Let u0u_{0} be the homogenized solution with the homogenized matrix in (α,β,D)\mathcal{M}(\alpha,\beta,D). If e(MOD)<αe(\text{MOD})<\alpha, then there exists a unique solution uhu_{h} satisfying (2.4).

If e(MOD)<να/(1+ν)e(\text{MOD})<\nu\alpha/(1+\nu) for any ν>0\nu>0, then

(3.1) (u0uh)L2(D)βαinfvVh(u0v)L2(D)+(1+ν)Cpα2fH1(D)e(MOD),\|\,\nabla(u_{0}-u_{h})\,\|_{L^{2}(D)}\leq\dfrac{\beta}{\alpha}\inf_{v\in V_{h}}\|\,\nabla(u_{0}-v)\,\|_{L^{2}(D)}+\dfrac{(1+\nu)C_{p}}{\alpha^{2}}\|\,f\,\|_{H^{-1}(D)}e(\text{MOD}),

where CpC_{p} appears in the discrete Poincaré inequality that only dependends on DD:

vH1(D)CpvL2(D)for all vVh.\|\,v\,\|_{H^{1}(D)}\leq C_{p}\|\,\nabla v\,\|_{L^{2}(D)}\qquad\text{for all\quad}v\in V_{h}.

Moreover, there exists CC depending only on α,β,ν,Cp\alpha,\beta,\nu,C_{p} and fH1(D)\|\,f\,\|_{H^{-1}(D)} such that

u0uhL2(D)\displaystyle\|\,u_{0}-u_{h}\,\|_{L^{2}(D)} C(infvVh(u0v)L2(D)supgL2(D)=1infχVh(ϕgχ)L2(D)+e(MOD)),\displaystyle\leq C\left(\inf_{v\in V_{h}}\|\,\nabla(u_{0}-v)\,\|_{L^{2}(D)}\sup_{\|\,g\,\|_{L^{2}(D)}=1}\inf_{\chi\in V_{h}}\|\,\nabla(\phi_{g}-\chi)\,\|_{L^{2}(D)}+e(\text{MOD})\right),

where ϕgH01(D)\phi_{g}\in H_{0}^{1}(D) is the unique solution of the problem:

(3.2) D𝒜(x)vϕgdx=Dg(x)v(x)dxfor all vH01(D).\int_{D}\mathcal{A}(x)\nabla v\cdot\nabla\phi_{g}\,\mathrm{d}\,x=\int_{D}g(x)v(x)\,\mathrm{d}\,x\qquad\text{for all\quad}v\in H_{0}^{1}(D).

The proof of the above lemma follows the same line of [19, Theorem 1.1] except the explicit constants in (3.1), we omit the proof.

To further elucidate the error structure of the method, we define the reconstruction operator for any piecewise constant function vv defined on 𝒯H\mathcal{T}_{H} as follows. Let Kv\mathcal{R}_{K}v be the solution of the least-squares

(3.3) Kv=argminpm(S(K))xK(K)|v(xK)p(xK)|2.\mathcal{R}_{K}v=\operatorname{arg}\min\limits_{p\in\mathbb{P}_{m}(S(K))}\sum\limits_{x_{K}\in\mathcal{I}(K)}\left\lvert\,v(x_{K})-p(x_{K})\,\right\rvert^{2}.

We imbed K\mathcal{R}_{K} into a global operator as |K=K\mathcal{R}|_{K}=\mathcal{R}_{K}.

Given the reconstruction operator \mathcal{R}, we decompose e(MOD)e(\text{MOD}) as

(3.4) 𝒜(x)AH(x)\displaystyle\mathcal{A}(x)-A_{H}(x) =𝒜(x)𝒜(x)+𝒜(x)AH(x)\displaystyle=\mathcal{A}(x)-\mathcal{R}\mathcal{A}(x)+\mathcal{R}\mathcal{A}(x)-A_{H}(x)
=(𝒜(x)𝒜(x))+(𝒜(x)A¯H),\displaystyle=\left(\mathcal{A}(x)-\mathcal{R}\mathcal{A}(x)\right)+\mathcal{R}\left(\mathcal{A}(x)-\overline{A}_{H}\right),

where A¯H(x)\overline{A}_{H}(x) is a piecewise matrix posed on 𝒯H\mathcal{T}_{H} that is defined by

A¯H|K:=AH(xK),\overline{A}_{H}|_{K}{:}=A_{H}(x_{K}),

with AH(xK)A_{H}(x_{K}) given by (2.2) and (2.3).

The first term in the right-hand side of (3.4) is the reconstruction error while the second one is the so-called estimation error. To quantify these two terms, we need some properties of the reconstruction operator, which is by now well-understood by virtue of [35] and [36]. To state such properties, we make two assumptions on S(K)S(K) and (K)\mathcal{I}(K).

Assumption A For every K𝒯HK\in\mathcal{T}_{H}, there exist constants RR and rr that are independent of KK such that BrS(K)BRB_{r}\subset S(K)\subset B_{R} with R2rR\geq 2r, and S(K)S(K) is star-shaped with respect to BrB_{r}, where BρB_{\rho} is a disk with radius ρ\rho.

This assumption concerns the geometry of S(K)S(K), which is crucial for the uniform boundedness of \mathcal{R}. The motivation for this assumption lies in the following Markov inequality [38]:

(3.5) gL(S(K))4m2Rr2gL(S(K))for allgm(S(K)).\|\,\nabla g\,\|_{L^{\infty}(S(K))}\leq\dfrac{4m^{2}R}{r^{2}}\|\,g\,\|_{L^{\infty}(S(K))}\qquad\text{for all}\quad g\in\mathbb{P}_{m}(S(K)).

Here gL(S(K)):=maxxS(K)¯g(x)2\|\,\nabla g\,\|_{L^{\infty}(S(K))}{:}=\max_{x\in\overline{S(K)}}\|\,\nabla g(x)\,\|_{\ell_{2}}. This inequality is proved in [36, Lemma 5], which is a combination of [57, Proposition 11.6] and the fact that S(K)S(K) satisfies the uniform interior cone condition [4], which is a direct consequence of Assumption A.

In [35], the authors make the following assumption on S(K)S(K).

Assumption A’ S(K)S(K) is a bounded convex domain and there exists RR such that S(K)BRS(K)\subset B_{R}.

By [26, Lemma 1.2.2.2 and Corollary 1.2.2.3], Assumption A’ implies Assumption A. Under Assumption A’, Wilhelmsen [58] proved the following Markov inequality:

(3.6) gL(S(K))4m2w(K)gL(S(K))for allgm(S(K)),\|\,\nabla g\,\|_{L^{\infty}(S(K))}\leq\dfrac{4m^{2}}{w(K)}\|\,g\,\|_{L^{\infty}(S(K))}\qquad\text{for all}\quad g\in\mathbb{P}_{m}(S(K)),

where w(K)w(K) is the width of S(K)S(K), which is the minimum distance between parallel supporting hyperplanes of S(K)S(K).

The above Markov inequalities may be viewed as a type of inverse inequality. By the classical inverse inequality for p-finite element method [47, Theorem 4.76] and a simple scaling argument, we may conclude that there exists CC independent of the diameter of S(K)S(K) but depends on the shape of S(K)S(K) such that for all gm(S(K))g\in\mathbb{P}_{m}(S(K)),

(3.7) gL(S(K))Cm2diamS(K)gL(S(K)).\|\,\nabla g\,\|_{L^{\infty}(S(K))}\leq\dfrac{Cm^{2}}{\text{diam}S(K)}\|\,g\,\|_{L^{\infty}(S(K))}.

The index 22 is sharp if S(K)S(K) is a locally Lipschitz domain. However, for a cuspidal domain, Kroó and Szabodos [34] proved that if S(K)S(K) is a Lipγ\gamma-domain with 0<γ<10<\gamma<1111A typical Lipγ\gamma-domain is the unit γ\ell_{\gamma}-ball, i.e.,{xd|x1|γ++|xd|γ1}\{\,x\in\mathbb{R}^{d}\,\mid\,\lvert x_{1}\rvert^{\gamma}+\cdots+\lvert x_{d}\rvert^{\gamma}\leq 1\,\} with 0<γ<10<\gamma<1., then there exists CC depends on the shape of S(K)S(K) such that

(3.8) gL(S(K))Cm2/γdiamS(K)gL(S(K)),\|\,\nabla g\,\|_{L^{\infty}(S(K))}\leq\dfrac{Cm^{2/\gamma}}{\text{diam}S(K)}\|\,g\,\|_{L^{\infty}(S(K))},

where the index 2/γ2/\gamma is sharp. This would require a larger patch to ensure the reconstruction accuracy, and the reconstruction is less stable than the patch satisfying either Assumption A or Assumption A’. Moreover, the constant CC in (3.7) is only known for S(K)S(K) with special shape in the literature, e.g., S(K)S(K) is an interval [46], and S(K)S(K) is a simplex [56, 33]. All the prefactors are important for us to derive the realistic conditions that ensure the uniform boundedness of Λ(m,(K))\Lambda(m,\mathcal{I}(K)); cf., Lemma 3.3.

On the other hand, The authors in [35] derived an explicit expression of CC that depends on the recursion depth tt and the chunkiness parameter σ\sigma under Assumption A’. It seems that the convexity of the patch is rather restrictive in implementation, particularly for an LL-shape domain. Assumption A is less restrictive and is easy to check in practice.

Assumption B For any K𝒯HK\in\mathcal{T}_{H} and pm(S(K))p\in\mathbb{P}_{m}(S(K)),

p|(K)=0 implies p|S(K)0.p|_{\mathcal{I}(K)}=0\text{\qquad implies\qquad}p|_{S(K)}\equiv 0.

This assumption concerns the cardinality of the sampling set (K)\mathcal{I}(K), which gives the uniqueness and hereby the existence of the solution of the discrete least-squares (2.1). Assumption B requires that the cardinality of (K)\mathcal{I}(K) is at least (m+dd)\binom{m+d}{d} to ensure the unisolvence of the discrete least-squares. A quantitative version of this assumption is

Λ(m,(K))<,\Lambda(m,\mathcal{I}(K))<\infty,

with

Λ(m,(K)):=maxpm(S(K))pL(S(K))p|(K).\Lambda(m,\mathcal{I}(K)){:}=\max_{p\in\mathbb{P}_{m}(S(K))}\dfrac{\|\,p\,\|_{L^{\infty}(S(K))}}{\|\,p|_{\mathcal{I}(K)}\,\|_{\ell_{\infty}}}.

In practical implementation, the positions of the sampling nodes may be slightly perturbed due to the measure error or certain uncertainties [15]. A natural question arises whether the reconstruction is robust with respect to the uncertainties. We shall prove that the reconstruction is stable with respect to small perturbation.

Next we prove some properties of the reconstruction operator K\mathcal{R}_{K}.

Lemma 3.2.

If Assumption B holds, then there exists a unique solution of (3.3) for any K𝒯HK\in\mathcal{T}_{H}. Moreover K\mathcal{R}_{K} satisfies

Kg=gfor all gm(S(K)).\mathcal{R}_{K}g=g\qquad\text{for all\quad}g\in\mathbb{P}_{m}(S(K)).

The stability result is valid for any K𝒯HK\in\mathcal{T}_{H} and gC0(S(K))g\in C^{0}(S(K)), and

(3.9) KgL(K)Λ(m,(K))#(K)g|(K).\|\,\mathcal{R}_{K}g\,\|_{L^{\infty}(K)}\leq\Lambda(m,\mathcal{I}(K))\sqrt{\#\mathcal{I}(K)}\;\|\,g|_{\mathcal{I}(K)}\,\|_{\ell_{\infty}}.

The quasi-optimal approximation property is valid in the sense that

(3.10) gKgL(K)(1+Λ(m,(K)))#(K)infpm(S(K))gpL(S(K)).\|\,g-\mathcal{R}_{K}g\,\|_{L^{\infty}(K)}\leq\left(1+\Lambda(m,\mathcal{I}(K))\right)\sqrt{\#\mathcal{I}(K)}\inf_{p\in\mathbb{P}_{m}(S(K))}\|\,g-p\,\|_{L^{\infty}(S(K))}.

If Assumption A and Assumption B are valid, then for any δ(0,1)\delta\in(0,1), there exists

(3.11) ϵ=δr24Λ(m,(K))m2R,\epsilon=\dfrac{\delta r^{2}}{4\Lambda(m,\mathcal{I}(K))m^{2}R},

such that for the perturbed sampling set I~(K)(K)+Bϵ(0)\widetilde{I}(K)\subset\mathcal{I}(K)+B_{\epsilon}(0), there exists a unique R~Kgm(S(K))\widetilde{R}_{K}g\in\mathbb{P}_{m}(S(K)) satisfying

(3.12) R~KgL(K)Λ(m,(K))1δ#I~(K)g|I~(K),\|\,\widetilde{R}_{K}g\,\|_{L^{\infty}(K)}\leq\dfrac{\Lambda(m,\mathcal{I}(K))}{1-\delta}\sqrt{\#\widetilde{I}(K)}\,\|\,g|_{\widetilde{I}(K)}\,\|_{\ell_{\infty}},

where Bε(0)B_{\varepsilon}(0) is a ball centered at 0 with radius ε\varepsilon.

If Assumption A’ and Assumption B are valid, then the perturbation result (3.12) remains true with

ε=δw(K)4Λ(m,(K))m2.\varepsilon=\dfrac{\delta w(K)}{4\Lambda(m,\mathcal{I}(K))m^{2}}.

The above lemma except the perturbation estimate (3.12) is proved in [35, Theorem 3.3], which is crucial for the accuracy of the reconstruction operator \mathcal{R}, more refined estimates on the accuracy of \mathcal{R} may be found in [36, Lemma 4]. The perturbation estimate (3.12) shows that the set of the sampling nodes (K)\mathcal{I}(K) perturbed a little bit remains a norming set with a slightly bigger upper bound, i.e., for all δ(0,1)\delta\in(0,1),

Λ(m,I~(K))=Λ(m,(K))1δ.\Lambda(m,\widetilde{I}(K))=\dfrac{\Lambda(m,\mathcal{I}(K))}{1-\delta}.

This perturbation estimate has been encapsulated in an abstract form in [36, Lemma 2], while there is no proof for the perturbed reconstruction operator R~\widetilde{R}.

Proof.

For any pm(S(K))p\in\mathbb{P}_{m}(S(K)), we let |p(x)|=p|(K)\lvert p(x^{\ast})\rvert=\|\,p|_{\mathcal{I}(K)}\,\|_{\ell_{\infty}}, then for any yI~(K)(K)+Bϵ(0)y\in\widetilde{I}(K)\subset\mathcal{I}(K)+B_{\epsilon}(0) with ϵ\epsilon given by (3.11), by Taylor’s expansion, we obtain

|p(x)||p(y)|+ϵpL(S(K)).\lvert p(x^{\ast})\rvert\leq\lvert p(y)\rvert+\epsilon\|\,\nabla p\,\|_{L^{\infty}(S(K))}.

By Assumption A, the Markov inequality (3.5) is valid, and we obtain

|p(x)||p(y)|+4m2Rϵr2pL(S(K)).\lvert p(x^{\ast})\rvert\leq\lvert p(y)\rvert+\dfrac{4m^{2}R\epsilon}{r^{2}}\|\,p\,\|_{L^{\infty}(S(K))}.

Using Assumption B and the above two inequalities, we obtain

pL(S(K))\displaystyle\|\,p\,\|_{L^{\infty}(S(K))} Λ(m,(K))p|(K)=Λ(m,(K))|p(x)|\displaystyle\leq\Lambda(m,\mathcal{I}(K))\|\,p|_{\mathcal{I}(K)}\,\|_{\ell_{\infty}}=\Lambda(m,\mathcal{I}(K))\lvert p(x^{\ast})\rvert
Λ(m,(K))|p(y)|+Λ(m,(K))4m2Rr2ϵpL(S(K))\displaystyle\leq\Lambda(m,\mathcal{I}(K))\lvert p(y)\rvert+\Lambda(m,\mathcal{I}(K))\dfrac{4m^{2}R}{r^{2}}\epsilon\|\,p\,\|_{L^{\infty}(S(K))}
Λ(m,(K))p|I~(K)+δpL(S(K)),\displaystyle\leq\Lambda(m,\mathcal{I}(K))\|\,p|_{\widetilde{I}(K)}\,\|_{\ell_{\infty}}+\delta\|\,p\,\|_{L^{\infty}(S(K))},

which immediately implies

(3.13) pL(S(K))Λ(m,(K))1δp|I~(K)for all δ(0,1).\|\,p\,\|_{L^{\infty}(S(K))}\leq\dfrac{\Lambda(m,\mathcal{I}(K))}{1-\delta}\|\,p|_{\widetilde{I}(K)}\,\|_{\ell_{\infty}}\qquad\text{for all\quad}\delta\in(0,1).

Applying the above perturbation estimate to R~Kg\widetilde{R}_{K}g, we obtain

R~KgL(S(K))Λ(m,(K))1δR~Kg|I~(K).\|\,\widetilde{R}_{K}g\,\|_{L^{\infty}(S(K))}\leq\dfrac{\Lambda(m,\mathcal{I}(K))}{1-\delta}\|\,\widetilde{R}_{K}g|_{\widetilde{I}(K)}\,\|_{\ell_{\infty}}.

This also gives the existence and uniqueness of R~Kg\widetilde{R}_{K}g.

By the definition of R~Kg\widetilde{R}_{K}g, we obtain

R~Kg|I~(K)2R~Kg|I~(K)22g|I~(K)22#I~(K)g|I~(K)2.\|\,\widetilde{R}_{K}g|_{\widetilde{I}(K)}\,\|_{\ell_{\infty}}^{2}\leq\|\,\widetilde{R}_{K}g|_{\widetilde{I}(K)}\,\|_{\ell_{2}}^{2}\leq\|\,g|_{\widetilde{I}(K)}\,\|_{\ell_{2}}^{2}\leq\#\widetilde{I}(K)\|\,g|_{\widetilde{I}(K)}\,\|_{\ell_{\infty}}^{2}.

A combination of the above two inequalities and the fact that #I~(K)=#(K)\#\widetilde{I}(K)=\#\mathcal{I}(K) give (3.12).

If Assumption A’ and Assumption B are valid, then we follow exactly the same line that leads to (3.12) except that we use the Markov inequality (3.6) for a convex patch S(K)S(K). ∎

The following lemma ensures the uniform boundedness of Λ(m,(K))\Lambda(m,\mathcal{I}(K)).

Lemma 3.3.

If Assumption A holds, then for any ε>0\varepsilon>0, if r>m2RHK(1+1/ε),r>m\sqrt{2RH_{K}(1+1/\varepsilon)}, then we may take

(3.14) Λ(m,(K))=1+ε.\Lambda(m,\mathcal{I}(K))=1+\varepsilon.

Moreover, if r>2mRHK,r>2m\sqrt{RH_{K}}, then we may take Λ(m,(K))=2.\Lambda(m,\mathcal{I}(K))=2.

If Assumption A’ holds, then for any ε>0\varepsilon>0, if w(K)>2m2HK(1+1/ε),w(K)>2m^{2}H_{K}(1+1/\varepsilon), then we may take

(3.15) Λ(m,(K))=1+ε.\Lambda(m,\mathcal{I}(K))=1+\varepsilon.

The first statement is proved in [36, Lemma 5], we only prove the second statement (3.15), which improves [35, Lemma 3.5].

Proof.

Let xS(K)¯x^{\ast}\in\overline{S(K)} such that |p(x)|=maxxS(K)¯|p(x)|\lvert p(x^{\ast})\rvert=\max_{x\in\overline{S(K)}}\lvert p(x)\rvert, and x(K)x_{\ell}\in\mathcal{I}(K) such that |xx|=miny(K)|yx|\lvert x_{\ell}-x^{\ast}\rvert=\min_{y\in\mathcal{I}(K)}\lvert y-x^{\ast}\rvert. Then

|xx|HK/2.\lvert x_{\ell}-x^{\ast}\rvert\leq H_{K}/2.

By Taylor’s expansion, we have

p(x)=p(x)+(xx)p(ξx),p(x_{\ell})=p(x^{\ast})+(x_{\ell}-x^{\ast})\cdot\nabla p(\xi_{x}),

where ξx\xi_{x} lies on the line with endpoints xx^{\ast} and xx_{\ell}. This gives

|p(x)||p(x)|+HK2pL(S(K)).\lvert p(x_{\ell})\rvert\leq\lvert p(x^{\ast})\rvert+\dfrac{H_{K}}{2}\|\,\nabla p\,\|_{L^{\infty}(S(K))}.

Using the Markov inequality (3.6), we immediately have

pL(S(K))p|(K)+2m2HKw(K)pL(S(K)).\|\,p\,\|_{L^{\infty}(S(K))}\leq\|\,p|_{\mathcal{I}(K)}\,\|_{\ell_{\infty}}+\dfrac{2m^{2}H_{K}}{w(K)}\|\,p\,\|_{L^{\infty}(S(K))}.

This implies (3.15). ∎

If Assumption A holds, then we usually have RtHKR\simeq tH_{K}, the estimate (3.14) suggests that rmtHKr\simeq m\sqrt{t}H_{K} implies the uniform boundedness of Λ(m,(K))\Lambda(m,\mathcal{I}(K)). This means that S(K)S(K) cannot be too narrow in certain directions. If Assumption A’ holds, then the estimate (3.15) shows that w(K)m2HKw(K)\simeq m^{2}H_{K}, which immediately implies tm2t\simeq m^{2}. Both conditions show that a relative large patch is required for the reconstruction. Furthermore, if S(K)S(K) is a cuspidal domain, then we may use (3.8) to prove that if

diamS(K)HKc0mγ(1+1/ε),\dfrac{\text{diam}\,S(K)}{H_{K}}\geq c_{0}m^{\gamma}(1+1/\varepsilon),

with a constant c0c_{0} depending on the shape of S(K)S(K), then the bound (3.14) is also valid. This condition indicates that S(K)S(K) has a recursive depth tmγt\simeq m^{\gamma} with γ>2\gamma>2. An even larger patch is required to ensure the stability. This may occur for domain DD with complicated boundary or rough boundary.

It remains to find an upper bound for #(K)\#\mathcal{I}(K). This is a direct consequence of Assumption A or Assumption A’ and the shape regularity of 𝒯H\mathcal{T}_{H}.

Lemma 3.4.

If 𝒯H\mathcal{T}_{H} is shape-regular and Assumption A or Assumption A’ is valid, then

(3.16) #(K)(σR/HK)d.\#\mathcal{I}(K)\leq\left(\sigma R/H_{K}\right)^{d}.
Proof.

For any element K𝒯HK\in\mathcal{T}_{H}, using Assumption A or Assumption A’ and note that there is only one sample point inside each element KK, we obtain

#(K)|K|Vol Bd(R),\#\mathcal{I}(K)\lvert K\rvert\leq\text{Vol\;}B_{d}(R),

where Vol Bd(R)\text{Vol\;}B_{d}(R) stands for the volume of a dd-dimensional ball with radius RR.

Using the shape regularity of 𝒯H\mathcal{T}_{H}, we obtain

|K|Vol Bd(ρK)σdVol Bd(hK).\lvert K\rvert\geq\text{Vol\;}B_{d}(\rho_{K})\geq\sigma^{-d}\text{Vol\;}B_{d}(h_{K}).

A combination of the above two inequalities yields (3.16). ∎

It is clear that the upper bound (3.16) is independent of the way for construction S(K)S(K). For the two ways based on either the Moore neighbor or the von Neumann neighbor, we have #(K)td\#\mathcal{I}(K)\simeq t^{d} with tt the recursion depth, which is consistent with the corresponding upper bound proved in [35, Lemma 3.4] and [36, Lemma 6], in which the two-dimensional problem has been dealt with. Both Lemma 3.3 and Lemma 3.4 require that #(K)\#\mathcal{I}(K) should be quite large, equivalently, S(K)S(K) is large, so that the uniform boundedness of Λ(m,(K))\Lambda(m,\mathcal{I}(K)) is valid. In numerical tests below, we observe that the method still works quite well even when #(K)\#\mathcal{I}(K) is far less than the theoretical threshold. We refer to [35] for a list of the size of S(K)S(K) and the upper bound of maxK𝒯HΛ(m,(K))\max_{K\in\mathcal{T}_{H}}\Lambda(m,\mathcal{I}(K)).

Based on the above three lemmas, we are ready to estimate (3.4).

Lemma 3.5.

If Assumption A or Assumption A’ and Assumption B are valid and the effective matrix 𝒜ijCm+1(D)\mathcal{A}_{ij}\in C^{m+1}(D), then there exists CC that depends on 𝒜ijCm+1(D),m,R,r,γ\|\,\mathcal{A}_{ij}\,\|_{C^{m+1}(D)},m,R,r,\gamma and tt but independent of HH.

(3.17) e(MOD)C(Hm+1+e1(MOD)),e(\text{MOD})\leq C\left(H^{m+1}+e_{1}(\text{MOD})\right),

where e1(MOD):=maxx(K),K𝒯H(𝒜AH)(x)Fe_{1}(\text{MOD}){:}=\max_{x\in\mathcal{I}(K),K\in\mathcal{T}_{H}}\|\,(\mathcal{A}-A_{H})(x)\,\|_{F}.

It is worthwhile to mention that e1(MOD)e_{1}(\text{MOD}) is the so-called estimating error in HMM [18]. There are many works devoted to bounding e1(MOD)e_{1}(\text{MOD}) and developing new algorithms to improve the estimates; see; e.g., [19, 59, 10, 23, 5, 24, 43] and the references therein.

Proof.

We start from the decomposition (3.4). For each K𝒯HK\in\mathcal{T}_{H}, using Assumption A and Assumption B or Assumption A’ and Assumption B, Lemma 3.3, Lemma 3.4 and  (3.10), we obtain, on each element K𝒯HK\in\mathcal{T}_{H}, there exists CC that depends on t,σ,r,R,dt,\sigma,r,R,d and 𝒜ijCm+1(D)\|\,\mathcal{A}_{ij}\,\|_{C^{m+1}(D)} such that

𝒜ij(K𝒜)ijL(K)Cinfpm(S(K))𝒜ij(x)pL(S(K))CHm+1.\|\,\mathcal{A}_{ij}-(\mathcal{R}_{K}\mathcal{A})_{ij}\,\|_{L^{\infty}(K)}\leq C\inf_{p\in\mathbb{P}_{m}(S(K))}\|\,\mathcal{A}_{ij}(x)-p\,\|_{L^{\infty}(S(K))}\leq CH^{m+1}.

Summing up all i,j=1,,di,j=1,\cdots,d and K𝒯HK\in\mathcal{T}_{H} we obtain

(3.18) maxxD𝒜(x)𝒜(x)FCHm+1.\max_{x\in D}\|\,\mathcal{A}(x)-\mathcal{R}\mathcal{A}(x)\,\|_{F}\leq CH^{m+1}.

Using (3.9), we obtain, for each K𝒯HK\in\mathcal{T}_{H},

K(𝒜ij(A¯H)ij)L(K)Λ(m,(K))#(K)(𝒜ij(AH)ij)|(K).\|\,\mathcal{R}_{K}\left(\mathcal{A}_{ij}-(\overline{A}_{H})_{ij}\right)\,\|_{L^{\infty}(K)}\leq\Lambda(m,\mathcal{I}(K))\sqrt{\#\mathcal{I}(K)}\|\,\left(\mathcal{A}_{ij}-(A_{H})_{ij}\right)|_{\mathcal{I}(K)}\,\|_{\ell_{\infty}}.

Summing up all i,j=1,,di,j=1,\cdots,d, we obtain

maxxKK(𝒜AH)(x)F2Λ2(m,(K))#(K)maxx(K)𝒜(x)AH(x)F2.\max_{x\in K}\|\,\mathcal{R}_{K}(\mathcal{A}-A_{H})(x)\,\|_{F}^{2}\leq\Lambda^{2}(m,\mathcal{I}(K))\#\mathcal{I}(K)\max_{x\in\mathcal{I}(K)}\|\,\mathcal{A}(x)-A_{H}(x)\,\|_{F}^{2}.

Using Assumption A and Assumption B or Assumption A’ and Assumption B, Lemma 3.3 and Lemma 3.4, there exists CC depends on R,r,tR,r,t and σ\sigma such that

maxxD(𝒜AH)FCe1(MOD),\max_{x\in D}\|\,\mathcal{R}(\mathcal{A}-A_{H})\,\|_{F}\leq Ce_{1}(\text{MOD}),

which together with (3.18) gives (3.17) and finishes the proof. ∎

Substituting the estimate (3.17) into Lemma 3.1, we obtain the main result of this part.

Theorem 1.

Let u0u_{0} be the homogenized solution with u0Hl+1(D)<\|\,u_{0}\,\|_{H^{l+1}(D)}<\infty. If Assumption A or Assumption A’ and Assumption B hold, and e1(MOD)<α/4e_{1}(\text{MOD})<\alpha/4, then there exists a unique solution uhu_{h} of Problem (2.4) that satisfying

(3.19) (u0uh)L2(D)C(hl+Hm+1+e1(MOD)).\|\,\nabla(u_{0}-u_{h})\,\|_{L^{2}(D)}\leq C\left(h^{l}+H^{m+1}+e_{1}(\text{MOD})\right).

Moreover, if the solution of the auxiliary problem (3.2) admits a unique solution ϕg\phi_{g} satisfying the regularity estimate ϕgH2(D)CgL2(D)\|\,\phi_{g}\,\|_{H^{2}(D)}\leq C\|\,g\,\|_{L^{2}(D)}, then

(3.20) u0uhL2(D)C(hl+1+Hm+1+e1(MOD)).\|\,u_{0}-u_{h}\,\|_{L^{2}(D)}\leq C\left(h^{l+1}+H^{m+1}+e_{1}(\text{MOD})\right).

In view of the above error estimates (3.19) and (3.20), we may have Hhl/(m+1)H\simeq h^{l/(m+1)} or Hh(l+1)/(m+1)H\simeq h^{(l+1)/(m+1)}. For a fixed hh, one may increase mm to decrease the cost in the offline stage. This suggests that higher order reconstruction is preferred to save cost while without loss of accuracy, which is confirmed by the tests in the next section.

The error estimate for HMM-LS method in [35] reads as

(3.21) {(u0u~h)L2(D)C(hl+hm+1+e1(MOD)),u0u~hL2(D)C(hl+1+hm+1+e1(MOD)),\left\{\begin{aligned} \|\,\nabla(u_{0}-\widetilde{u}_{h})\,\|_{L^{2}(D)}&\leq C\left(h^{l}+h^{m+1}+e_{1}(\text{MOD})\right),\\ \|\,u_{0}-\widetilde{u}_{h}\,\|_{L^{2}(D)}&\leq C\left(h^{l+1}+h^{m+1}+e_{1}(\text{MOD})\right),\end{aligned}\right.

where u~h\widetilde{u}_{h} is the solution of HMM-LS method. For a fixed hh, the above error estimate indicates that mlm\simeq l to balance the error provided that e1(MOD)e_{1}(\text{MOD}) is sufficiently small, the number of the cell problems in this case is the total number of the vertices of the online triangulation τh\tau_{h}. We shall compare these two methods in the next part.

4. Numerical Results

In this section, we report a few numerical examples to test the accuracy and efficiency of the proposed offline-online method. The examples include the problems with locally periodic coefficients, with almost periodic coefficients and with random coefficients. We also test problems posed on L-shape domain in two dimension and three dimension for which the solutions are usually nonsmooth.

In all the tests, the offline triangulations 𝒯H\mathcal{T}_{H} are uniform partitions of DD, and we list the threshold number (NlowestN_{\text{lowest}} is defined in § 2) of the sampling points for different orders of reconstruction in Table 1, the number #(K)\#\mathcal{I}(K) should be slightly larger than the number that lists in the table when KK abuts the boundary of the domain. For d=3d=3, we only report m=3m=3 because this is the only case we test in this part, the examples for the lower order reconstruction may be found in [28].

Table 1. The number of the threshold number for the sampling point in reconstruction.
m=1m=1 m=2m=2 m=3m=3
d=2d=2 5 7 13
d=3d=3 NONE NONE 27

Noted that the theory covers the case when S(K)S(K) consists of polygons or polytopes, we refer to [36] for the demonstration of such patches. It is worth pointing out that the nonuniform 𝒯H\mathcal{T}_{H} is useful when the shape of DD is very complicate. In that case, the sampling points near the boundary has to be dense to ensure the accuracy of the reconstruction. We refer to [28] for the examples of nonuniform patch, which may further reduce the number of the cell problems and makes e(MOD)e(\text{MOD}) more evenly distributed over the whole domain, which is useful for adaptivity.

The finite element solvers except Example 3 are carried on FreeFem++ toolbox [27]222https://freefem.org/, and the test for Example 3 is performed in a parallel hierarchical grid platform (PHG) [60]333http://lsec.cc.ac.cn/phg/. The error of the method is calculated by the relative error measured in H1H^{1} norm and L2L^{2} norm:

(u0uh)L2(D)u0L2(D)andu0uhL2(D)u0L2(D).\dfrac{\|\,\nabla(u_{0}-u_{h})\,\|_{L^{2}(D)}}{\|\,\nabla u_{0}\,\|_{L^{2}(D)}}\qquad\text{and}\qquad\dfrac{\|\,u_{0}-u_{h}\,\|_{L^{2}(D)}}{\|\,u_{0}\,\|_{L^{2}(D)}}.

The exact homogenized solution u0u_{0} is generated by discretization (1.4) with 2\mathbb{P}_{2} FEM over a very refined 500×500500\times 500 mesh in most cases unless otherwise stated. In all the tests we set ε=106\varepsilon=10^{-6}.

4.1. Problems with local periodic coefficient in two dimension

We consider an example with locally periodical coefficient in d=2d=2, which is taken from [41].

Example 1.
{div(aεuϵ(x))=f(x),inD,uϵ(x)=0,onD,\left\{\begin{aligned} -\operatorname{div}(a^{\;\varepsilon}\nabla u^{\epsilon}(x))&=f(x),\quad\quad&&\text{in}\quad D,\\ u^{\epsilon}(x)&=0,\quad&&\text{on}\quad\partial D,\end{aligned}\right.

where D=(0,1)2D=(0,1)^{2}, f(x)=1f(x)=1 and

aε(x)=(2.5+1.5sin(2πx1))(2.5+1.5cos(2πx2))(2.5+1.5sin(2πx1/ε))(2.5+1.5sin(2πx2/ϵ))12×2,a^{\;\varepsilon}(x)=\frac{(2.5+1.5\sin(2\pi x_{1}))(2.5+1.5\cos(2\pi x_{2}))}{(2.5+1.5\sin(2\pi x_{1}/\varepsilon))(2.5+1.5\sin(2\pi x_{2}/\epsilon))}1_{2\times 2},

where 12×21_{2\times 2} denotes the 22-by2-2 identity matrix. The homogenization problem is

(4.1) {div(𝒜(x)u0(x))=f(x),inD,u0(x)=0,onD.\left\{\begin{aligned} -\operatorname{div}(\mathcal{A}(x)\nabla u_{0}(x))&=f(x),\quad\quad&&\text{in}\quad D,\\ u_{0}(x)&=0,\quad\quad&&\text{on}\quad\partial D.\end{aligned}\right.

A direct calculation gives the following analytical expression of 𝒜\mathcal{A}:

(4.2) 𝒜(x1,x2)=15(2.5+1.5sin(2πx1))(2.5+1.5cos(2πx2))12×2.\mathcal{A}(x_{1},x_{2})=\dfrac{1}{5}(2.5+1.5\sin(2\pi x_{1}))(2.5+1.5\cos(2\pi x_{2}))1_{2\times 2}.

The offline triangulation 𝒯H\mathcal{T}_{H} consists of a uniform Q×QQ\times Q squares. To obtain the online triangulation τh\tau_{h}, we firstly triangulate DD into a uniform N×NN\times N squares and secondly divide each square into two sub-triangles along its diagonal with positive slope. To study the effect of the reconstruction, we use the analytical expression (4.2) of 𝒜\mathcal{A} for reconstruction so that e1(MOD)=0e_{1}(\text{MOD})=0. We denote this numerical solution by uh0u_{h}^{0}. Fig. 2 presents the accuracy of the offline computation, which corroborates the estimate (3.17).

Refer to caption

Figure 2. e(MOD)e(\text{MOD}) for different reconstruction orders.

To study the convergence rate of the method, it is more convenient to reshape the error estimates in Theorem 1 in terms of NN and QQ as follows.

(4.3) {(u0uh)L2(D)C(Nl+Q(m+1)),u0uhL2(D)C(N(l+1)+Q(m+1)).\left\{\begin{aligned} \|\,\nabla(u_{0}-u_{h})\,\|_{L^{2}(D)}&\leq C(N^{-l}+Q^{-(m+1)}),\\ \|\,u_{0}-u_{h}\,\|_{L^{2}(D)}&\leq C(N^{-(l+1)}+Q^{-(m+1)}).\end{aligned}\right.

Balancing the discretization error and the reconstruction error, we set QQ as

Q={𝒪(Nlm+1),H1 error,𝒪(Nl+1m+1),L2 error.Q=\left\{\begin{aligned} \mathcal{O}(N^{\frac{l}{m+1}}),&\quad H^{1}\text{ error},\\ \mathcal{O}(N^{\frac{l+1}{m+1}}),&\quad L^{2}\text{ error}.\end{aligned}\right.

Following this refinement strategy, we plot the relative H1H^{1} error and L2L^{2} error in Fig. 3, which are consistent with the estimates (4.3).

Refer to caption

(a). Rate of convergence in H1H^{1}.

Refer to caption

(b). Rate of convergence in L2L^{2}.

Figure 3. Rates of convergence for Example 1.

We may also fix the accuracy of the online solver and compare the effect of the reconstruction with different orders in the offline computation. Particularly we choose the online solver as 2\mathbb{P}_{2} FEM with the meshsize parameter N=100N=100. Fig. 4 clearly shows that the higher-order reconstruction is more accurate with less cost.

Refer to caption

(a). The H1H^{1} error with different reconstruction orders.

Refer to caption

(b). The L2L^{2} error with different reconstruction orders.

Figure 4. The effect of QQ and the reconstruction orders in Example 1; Online solver is 2\mathbb{P}_{2} FEM and N=100N=100.

Next, we also fix the accuracy of the online solver and compare the number of the cell problems and the running time with reconstructions of different orders. The online solver is still 2\mathbb{P}_{2} FEM with the meshsize parameter N=100N=100. Instead of using the analytical expression of 𝒜\mathcal{A} for reconstruction, we solve the periodic cell problems

(4.4) {div(aεviε(x))=0in Iε,viεxi is periodicon Iε,\left\{\begin{aligned} -\operatorname{div}(a^{\;\varepsilon}\nabla v^{\,\varepsilon}_{i}(x))&=0\qquad&&\text{in\quad}I_{\varepsilon},\\ v^{\,\varepsilon}_{i}-x_{i}&\text{\quad is periodic}\qquad&&\text{on\quad}\partial I_{\varepsilon},\end{aligned}\right.

with 2\mathbb{P}_{2} FEM, and employ (2.2) to obtain AHA_{H}. We denote the reconstructed solution as uh1u_{h}^{1}, which is very close to uh0u_{h}^{0} in the sense

uh1uh0L2(D)uh0L2(D)108.\dfrac{\|\,u_{h}^{1}-u_{h}^{0}\,\|_{L^{2}(D)}}{\|\,u_{h}^{0}\,\|_{L^{2}(D)}}\leq 10^{-8}.

This means that the e1(MOD)e_{1}(\text{MOD}) is very small, though nonzero. Table 2 shows that the main cost comes from solving the cell problems, and higher-order reconstruction is more efficient in terms of both the accuracy and the efficiency.

Table 2. Comparison among reconstruction of different orders with the same online solver (N=100N=100, 2\mathbb{P}_{2} FEM).
Q \sharp cell problems Relative H1H^{1} error
Time
(cell problems)
Total Time
m=1m=1 45 2025 6.03e-4 348.65s 363.27s
m=2m=2 25 625 5.92e-4 106.97s 110.76s
m=3m=3 18 324 5.92e-4 54.5s 58.23

In the last test, we compare the offline-online method with HMM-LS method in [35]. The H1H^{1} error bound in (3.21) may be rewritten as

(u0u~h)L2(D)C(Nl+N(m+1)+e1(MOD)).\|\,\nabla(u_{0}-\widetilde{u}_{h})\,\|_{L^{2}(D)}\leq C(N^{-l}+N^{-(m+1)}+e_{1}(\text{MOD})).

We follow the setup in the last test to solve the cell problems so that e1(MOD)e_{1}(\text{MOD}) is negligible in the above estimate. We take l=2l=2 and m=1m=1 to balance the error so that

(u0u~h)L2(D)CN2.\|\,\nabla(u_{0}-\widetilde{u}_{h})\,\|_{L^{2}(D)}\leq C\,N^{-2}.

In the offline-online method, we use 2\mathbb{P}_{2} FEM as the macroscopic solver and choose the reconstruction order m=2m=2 and m=3m=3. To balance the error of  (4.3), we take Q=2N2/3Q=2N^{2/3} for m=2m=2 and Q=2.5N1/2Q=2.5N^{1/2} for m=3m=3 and we obtain (u0uh)L2(D)CN2\|\,\nabla(u_{0}-u_{h})\,\|_{L^{2}(D)}\leq C\,N^{-2}. It seems that the relative H1H^{1} errors for both methods reported in Fig. 5 are comparable, which is consist with the theoretical estimates. The number of the cell problems and the running time plotted in Fig. 5 demonstrate that the offline-online method is more efficient. This is easily understood because the number of the cell problems in the offline-online method is of 𝒪(N4/3)\mathcal{O}(N^{4/3}) for m=2m=2 and is of 𝒪(N)\mathcal{O}(N) for m=3m=3, while the number of the cell problems in HMM-LS is of 𝒪(N2)\mathcal{O}(N^{2}).

Refer to caption

(a). Relative H1H^{1} error.

Refer to caption

(b). Running time and the number of the cell problems.

Figure 5. Comparison between the offline-online method and HMM-LS method in [35].

4.2. Problems posed on L-shape domain in d=2,3d=2,3

In this part, we test the offline-online strategy for a problem posed on L-shape domain. It is well-known that the adaptive strategy has to be used in the macroscopic solver because the solution u0u_{0} is nonsmooth [26].

Example 2.

The boundary value problem is the same with Example 1 except that the domain D=(1,1)2[0,1]×[1,0]D=(-1,1)^{2}\setminus[0,1]\times[-1,0]. We let

u0(x)=(x12+x22)1/3.u_{0}(x)=(x_{1}^{2}+x_{2}^{2})^{1/3}.

The inhomogeneous boundary condition g(x)=u0(x)|xDg(x)=u_{0}(x)|_{x\in\partial D} celland the source term ff is computed from the homogenized problem (4.1)1.

In the offline stage, the periodic cell problems (4.4) are solved by 2\mathbb{P}_{2} FEM over a mesh of size 50×5050\times 50. The third order reconstruction is used to obtain AHA_{H}. In the online stage, 1\mathbb{P}_{1} FEM is used as the macroscopic solver. Then the error estimate roughly reads as

(4.5) (u0uh)L2(D)C(DOF1/2+Q4+e1(MOD)),\|\,\nabla(u_{0}-u_{h})\,\|_{L^{2}(D)}\leq C\left(\text{DOF}^{-1/2}+Q^{-4}+e_{1}(\text{MOD})\right),

where DOF is the total degrees of freedom in the online computation Given the error tolerance (TOL), in the offline stage we need to require

QTOL1/4ande1(MOD)TOL.Q\lesssim\text{TOL}^{-1/4}\qquad\text{and}\qquad e_{1}(\text{MOD})\lesssim\text{TOL}.

Firstly we set TOL=102=10^{-2}. The triangulation 𝒯H\mathcal{T}_{H} is plot in Fig. 6a, we have 972972 sampling points in total. Under these settings, e(MOD)=1.54E02e(\text{MOD})=1.54E-02 in the offline stage. let 𝒯H\mathcal{T}_{H} be the initial mesh of the online computation. The mesh is refined by the strategy taken from [27], and we plot the resulting mesh after 88 iterations in Fig. 6b.

Refer to caption

(a). Offline sampling mesh 𝒯H\mathcal{T}_{H}.

Refer to caption

(b). Online adaptive mesh τh\tau_{h} after 88 iterations.

Figure 6. Offline and online triangulations for L-shaped domain.

The relative H1H^{1} error is plot in Fig. 7a, and we observe that the error decays with optimal rate of convergence 𝒪(DOF1/2)\mathcal{O}(\text{DOF}^{-1/2}).

Refer to caption

(a). Tolerance=102=10^{-2}.

Refer to caption

(b). Tolerance=103=10^{-3}.

Figure 7. The H1H^{1} error of Example 2

Next we decrease the error tolerance TOL to 10310^{-3}, and we also use the third order reconstruction to approximate 𝒜\mathcal{A} and there are 43324332 sampling points in total. The error caused by the reconstruction is e(MOD)=9.48E04e(\text{MOD})=9.48E-04. The relative H1H^{1} error is plot in Fig. 7b, and we observe that the error decays with optimal rate of convergence 𝒪(DOF1/2)\mathcal{O}(\text{DOF}^{-1/2}).

Finally we compare the accuracy and efficiency between our method and HMM. The offline computation is the same as the previous test, while in the online stage, we solve the homogenized problem with linear element over the same mesh as HMM. Such mesh is obtained after 88 iterations with 3439434394 triangles from a uniform mesh dividing each of the uniform squares into two sub-triangles; see Fig. 6b. It is clear that the mesh refinement mostly takes place in the vicinity of the re-entrant corner. It follows from Table 3 that the offline-online method and HMM attain almost the same H1H^{1} error, while the running time of the offline-online method is only around 2.46%2.46\% of HMM. This is easily understood because the number of the cell problems is only 1.67%1.67\% of HMM.

Table 3. Comparison between the offline-online method and HMM.
\sharp cell problems Relative H1H^{1} error Time
Offline-Online 576 3.53e-5 295.95s
HMM 34394 3.62e-5 12020.6s

Next we consider a three-dimensional problem in L-shape domain.

Example 3.

The example is the same with Example 1 except that

aε(x)=(2.5+1.5sin(2πx1))(2.5+1.5cos(2πx2))(2.5+1.5cos(2πx3))(2.5+1.5sin(2πx1/ϵ))(2.5+1.5sin(2πx2/ϵ))(2.5+1.5sin(2πx3/ϵ))13×3×3,a^{\;\varepsilon}(x)=\frac{(2.5+1.5\sin(2\pi x_{1}))(2.5+1.5\cos(2\pi x_{2}))(2.5+1.5\cos(2\pi x_{3}))}{(2.5+1.5\sin(2\pi x_{1}/\epsilon))(2.5+1.5\sin(2\pi x_{2}/\epsilon))(2.5+1.5\sin(2\pi x_{3}/\epsilon))}1_{3\times 3\times 3},

and D=(0,1)3[0,0.5]3D=(0,1)^{3}\setminus[0,0.5]^{3}. The homogenization problem is the same with (4.1) with the effective matrix as

(4.6) 𝒜(x)=15(2.5+1.5sin(2πx1))(2.5+1.5cos(2πx2))(2.5+1.5cos(2πx3))13×3×3.\mathcal{A}(x)=\dfrac{1}{5}(2.5+1.5\sin(2\pi x_{1}))(2.5+1.5\cos(2\pi x_{2}))(2.5+1.5\cos(2\pi x_{3}))1_{3\times 3\times 3}.

We let

u0(x)=((x10.5)2+(x20.5)2+(x30.5)2)1/10,u_{0}(x)=\left((x_{1}-0.5)^{2}+(x_{2}-0.5)^{2}+(x_{3}-0.5)^{2}\right)^{1/10},

and

g(x)=u0(x)|D,f(x)=div(𝒜(x)u0).g(x)=u_{0}(x)|_{\partial D},\qquad\qquad f(x)=-\operatorname{div}(\mathcal{A}(x)\nabla u_{0}).

We write the error estimate roughly as

(4.7) (u0uh)L2(D)C(DOF1/3+Q(m+1)+eβNs).\|\,\nabla(u_{0}-u_{h})\,\|_{L^{2}(D)}\leq C\left(\text{DOF}^{-1/3}+Q^{-(m+1)}+e^{-\beta N_{s}}\right).

where DOF is the total degrees of freedom in the online computation, and the factor eβNse^{-\beta N_{s}} comes from the approximation error caused by using Fourier spectral method [49] to solve the periodic cell problems (4.4), and β\beta is a universal constant and NsN_{s} is the points we used in each cell. The offline refinement strategy reads as

QTOl1m+1and eβNsTOL.Q\lesssim\text{TOl}^{-\frac{1}{m+1}}\qquad\text{and\qquad}e^{-\beta N_{s}}\lesssim\text{TOL}.

We set TOL=3E3=3E-3 and m=3m=3. We firstly divide the domain into 77 subdomain as in the last example, next we triangulate each subdomain by a uniform mesh with Q=33Q=33. There are 76237623 sampling points in total; see Fig. 8a. We choose Ns=25N_{s}=25. The H1H^{1} error is plot in Fig. 8b. We observe that the H1H^{1} error decays with an optimal convergence rate 𝒪(DOF1/3)\mathcal{O}(\text{DOF}^{-1/3}).

Refer to caption

(a). Offline sampling mesh 𝒯H\mathcal{T}_{H}.

Refer to caption

(b). H1 error with tolerance=3E3=3E-3.

Figure 8. Offline triangulation and H1 error for Example 3.

4.3. Example with almost periodic coefficient

Example 4.

The example is the same with Example 1 except that

(4.8) aε(x)=a0(x/ε)a1(x)12×2,a^{\;\varepsilon}(x)=a^{0}(x/\varepsilon)\,a^{1}(x)1_{2\times 2},

where

a0(x)=(6+sin(2πx1)2+sin(22πx1)2006+sin(2πx2)2+sin(22πx2)2),a^{0}(x)=\left(\begin{array}[]{cc}6+\sin\left(2\pi x_{1}\right)^{2}+\sin\left(2\sqrt{2}\pi x_{1}\right)^{2}&0\\ 0&6+\sin\left(2\pi x_{2}\right)^{2}+\sin\left(2\sqrt{2}\pi x_{2}\right)^{2}\\ \end{array}\right),

and

a1(x)=(2.5+1.5sin(2πx1))(2.5+1.5cos(2πx2)).a^{1}(x)=(2.5+1.5\sin(2\pi x_{1}))(2.5+1.5\cos(2\pi x_{2})).

Such coefficient belongs to Kozlov class [32] and the example is adapted from [24]. The coefficients a110(x/ε)a^{0}_{11}(x/\varepsilon) and a11εa^{\;\varepsilon}_{11} are visualized in Fig. 9 with ε=0.1\varepsilon=0.1.

Refer to caption

(a). plot of a110(x/ε)a^{0}_{11}(x/\varepsilon) .

Refer to caption

(a). plot of a11εa^{\;\varepsilon}_{11}.

Figure 9. Plots of the coefficients in Example 4 (ε=0.1\varepsilon=0.1)

The effective matrix is given by

(4.9) 𝒜ij(x)=lim infR+(χj+xj)a0(x)(χi+xi)QRa1(x),i,j=1,2,\mathcal{A}_{ij}(x)=\liminf_{R\rightarrow+\infty}\left\langle\,\nabla(\chi^{j}+x_{j})\cdot a^{0}(x)\nabla(\chi^{i}+x_{i})\,\right\rangle_{Q_{R}}*a_{1}(x),\quad i,j=1,2,

where QR=(R,R)2Q_{R}=(-R,R)^{2} and {χi}i=12\{\chi^{i}\}_{i=1}^{2} are the solutions of

(4.10) (a0(x)χi)=(a0(x)xi)in 2.-\nabla\cdot(a^{0}(x)\nabla\chi^{i})=\nabla\cdot(a^{0}(x)\nabla x_{i})\quad\text{in\quad}\mathbb{R}^{2}.

As oppose to the locally periodic case, there is no explicit expression of 𝒜\mathcal{A}. The naive approach to approximate (4.9) consists in replacing {χi}i=12\{\chi^{i}\}_{i=1}^{2} by {χRi}i=12\{\chi^{i}_{R}\}_{i=1}^{2}which are solutions of a truncated cell problem

{(a0(x)χRi)=(a0(x)xi)in QR,χRi=0on QR,\left\{\begin{aligned} -\nabla\cdot(a^{0}(x)\nabla\chi^{i}_{R})=\nabla\cdot(a^{0}(x)\nabla x_{i})&\quad\text{in\quad}Q_{R},\\ \chi^{i}_{R}=0&\quad\text{on\quad}\partial Q_{R},\end{aligned}\right.

and 𝒜R\mathcal{A}_{R} is defined by

(𝒜R(x))ij:=(χRj+xj)a0(x)(χRi+xi)QRa1(x),i,j=1,2.\left(\mathcal{A}_{R}(x)\right)_{ij}{:}=\left\langle\,\nabla(\chi^{j}_{R}+x_{j})\cdot a^{0}(x)\nabla(\chi^{i}_{R}+x_{i})\,\right\rangle_{Q_{R}}*a_{1}(x),\quad i,j=1,2.

We take R=200R=200 and use 2\mathbb{P}_{2} FEM over a 4000×40004000\times 4000 mesh to solve the above truncated cell problem and obtain

𝒜200=(7.00a1(x)007.00a1(x)).\mathcal{A}_{200}=\begin{pmatrix}7.00*a^{1}(x)&0\\ 0&7.00*a^{1}(x)\end{pmatrix}.

In the offline stage, we choose δ=10ε\delta=10\,\varepsilon and use 2\mathbb{P}_{2} FEM over a mesh with size 120×120120\times 120 to solve the cell problems (2.3). Next we plot the relative H1H^{1} error and the relative L2L^{2} error in Fig. 10, which clearly shows the higher-order reconstruction is more accurate.

Refer to caption

(a). The relative H1H^{1} error with different reconstruction orders.

Refer to caption

(b). The relative L2L^{2} error with different reconstruction orders.

Figure 10. The effect of QQ and the reconstruction orders in Example 4; Online solver is 2\mathbb{P}_{2} FEM and N=100N=100.

Finally we compare the number of the cell problems and the running time among reconstructions of different orders in Table 4.

Table 4. Comparison of the reconstruction of different orders; online solver is 2\mathbb{P}_{2} FEM and N=100N=100.
Q \sharp cell problems Relative H1H^{1} error Time
m=1m=1 20 400 6.08e-3 1083.08s
m=2m=2 14 196 4.95e-3 546.24s

The second order reconstruction takes only about 50%50\% of the time used for the first order reconstruction. It is clear that the higher-order reconstruction is more accurate with less cost.

4.4. Example with random coefficient

Example 5.

The example is the same with Example 1 except that

(4.11) aε(x)=(arandε+a0(x))12×2,a^{\;\varepsilon}(x)=\left(a^{\;\varepsilon}_{\text{rand}}+a_{0}(x)\right)1_{2\times 2},

where arandεa^{\;\varepsilon}_{\text{rand}} is a random checker-board, and a0(x)=(2.5+1.5sin(2πx1))(2.5+1.5cos(2πx2))a_{0}(x)=(2.5+1.5\sin(2\pi x_{1}))(2.5+1.5\cos(2\pi x_{2})). arandεa^{\;\varepsilon}_{\text{rand}} is constructed by partitioning D=(0,1)2D=(0,1)^{2} into uniform square cells of size ε\varepsilon, each of which is randomly designated as k1k_{1} or k2k_{2} with probability p1p_{1} and p2=1p1p_{2}=1-p_{1}, respectively. We visualize one realization of the random coefficients in Fig. 11 with ε=0.02\varepsilon=0.02. Theorem 1 remains true with a minor modification of e1(MOD)e_{1}(\text{MOD}), we refer to [19] for related result.

Refer to caption
Refer to caption
Figure 11. Plot of one realization of a11εa^{\;\varepsilon}_{11} in (4.11). (a) is the contour of a11εa^{\;\varepsilon}_{11}.

As oppose to the standard random checker-board [30], there is no explicit formula for the effective matrix of (4.11). To extract the effective matrix over DD, we take δ=16ε\delta=16\,\varepsilon as the cell size and use 2\mathbb{P}_{2} FEM over a mesh of size 160×160160\times 160 to solve the cell problem (4.4). We set the offline mesh size as Q=20Q=20 and use third order reconstruction to get an approximation effective matrix 𝒜\mathcal{A}, which will be exploited to obtain the reference solution in the following test.

To justify this approach, we choose AHA_{H} at three representative points in DD: A=(1/4,0)A=(1/4,0), which is one of the maximum point of a0a_{0} over DD; B=(3/4,1/2)B=(3/4,1/2), which is one of the minimum point of a0a_{0} over DD, and C=(1,1/4)C=(1,1/4), which is one of the maximum point of a0\nabla a_{0} over DD. We let δ=Lε\delta=L\,\varepsilon and use 2\mathbb{P}_{2} FEM over a mesh of size 10L×10L10L\times 10L to solve the cell problems (4.4) posed over these three points. We denote by 𝒜16n\mathcal{A}_{16}^{n} the approximation effective matrix for nn-th realization with L=16L=16. In addition, we use the empirical average

(4.12) 𝔼(𝒜16)(x):=1Nn=1N𝒜16n(x),\mathbb{E}(\mathcal{A}_{16})(x){:}=\dfrac{1}{N}\sum_{n=1}^{N}\mathcal{A}_{16}^{n}(x),

as the proxy of the expectation 𝔼\mathbb{E}, where NN is the total number of the realization. We take N=1000N=1000 in the simulation, and the results are reported in Table 5.

Table 5. The approximating effective matrix on three points with L=16L=16.
𝔼(𝒜16)\mathbb{E}(\mathcal{A}_{16})
A=(1/4,0)A=(1/4,0) (20.782.58e42.58e420.78)\left(\begin{array}[]{cc}20.78&-2.58e-4\\ -2.58e-4&20.78\\ \end{array}\right)
B=(3/4,1/2)B=(3/4,1/2) (5.209.47e49.47e45.20)\left(\begin{array}[]{cc}5.20&-9.47e-4\\ -9.47e-4&5.20\\ \end{array}\right)
C=(1,1/4)C=(1,1/4) (10.844.86e44.86e410.84)\left(\begin{array}[]{cc}10.84&-4.86e-4\\ -4.86e-4&10.84\\ \end{array}\right)

Secondly, for x=A,B,Cx=A,B,C, we measure the variance as

σLdiag(x):=𝔼((𝒜L(x)11𝔼(𝒜16(x))11)2+(𝒜L(x)22𝔼(𝒜16(x))22)2),\sigma^{\text{diag}}_{L}(x){:}=\sqrt{\mathbb{E}\left(\left(\mathcal{A}_{L}(x)_{11}-\mathbb{E}(\mathcal{A}_{16}(x))_{11}\right)^{2}+\left(\mathcal{A}_{L}(x)_{22}-\mathbb{E}(\mathcal{A}_{16}(x))_{22}\right)^{2}\right)},

and

σL12(x):=𝔼(𝒜L(x)2)12.\sigma^{12}_{L}(x){:}=\sqrt{\mathbb{E}\left(\mathcal{A}_{L}(x)^{2}\right)_{12}}.

Fig. 12a and Fig. 12b suggest that σLdiag(x)\sigma^{\text{diag}}_{L}(x) and σL12(x)\sigma^{12}_{L}(x) decay as 𝒪(L1)\mathcal{O}(L^{-1}), which is consistent with the theoretical predictions [25]. A systematical numerical tests for the variance may be found in a recent work [31].

Refer to caption

(a). Decays of σLdiag\sigma^{\text{diag}}_{L} with respect to the cell size LL.

Refer to caption

(b). Decays of σL12\sigma^{12}_{L} with respect to the cell size LL.

Figure 12. The accuracy for the approximating effective matrix.

Finally we compare the effect of the reconstructions with different orders in the offline computation. We fix the online solver as 2\mathbb{P}_{2} FEM over a mesh with size 100×100100\times 100. In the offline stage, we choose δ=8ε\delta=8\,\varepsilon and use 2\mathbb{P}_{2} FEM over a mesh with size 80×8080\times 80 to solve the cell problems (2.3). We take Q=8, 10, 16Q=8,\,10,\,16 and m=1, 2m=1,\,2 in the tests, and compute the ensemble average of the relative H1H^{1} error and L2L^{2} error

𝔼((u0uh)L2(D))u0L2(D)and𝔼(u0uhL2(D))u0L2(D).\dfrac{\mathbb{E}\left(\|\,\nabla(u_{0}-u_{h})\,\|_{L^{2}(D)}\right)}{\|\,\nabla u_{0}\,\|_{L^{2}(D)}}\quad\text{and}\quad\dfrac{\mathbb{E}\left(\|\,u_{0}-u_{h}\,\|_{L^{2}(D)}\right)}{\|\,u_{0}\,\|_{L^{2}(D)}}.

The expectation 𝔼\mathbb{E} is replaced by the empirical average as that in (4.12) with N=1000N=1000 realizations. The visualization in Fig. 13a and Fig. 13b clearly shows that the second order reconstruction is more accurate than the first order reconstruction.

Refer to caption

(a). Ensemble average of the H1H^{1} error.

Refer to caption

(b). Ensemble average of the L2L^{2} error.

Figure 13. The error between numerical solution and homogenized solution in Example 5.

5. Conclusion

We have proposed a new online-offline method to solve the multiscale elliptic problems. Both theoretical and numerical results show that the method significantly reduces the cost while retains the optimal rate of convergence. Moreover, the strategy is problem independent, and it can be extended to time-dependent problems [42, 21]. The implementation of the present method is mainly based on the a priori error estimate. Adaptive algorithms based on the a priori error estimate should be developed for automatic tuning of the parameters so that the method is more efficient. We shall leave all these issues in the future work.

References

  • [1] A. Abdulle and Y. Bai, Reduced basis finite element heterogeneous multiscale method for high-order discretizations of elliptic homogenization problems, J. Comput. Phys., 231 (2012), 7014-7036.
  • [2] A. Abdulle and Y. Bai, Adaptive reduced basis finite element heterogeneous multiscale method, Comput. Methods Appl. Mech. Engrg., 257 (2013), 203-220.
  • [3] A. Abdulle, W. E, B. Engquist and E. Vanden-Eijnden, The heterogeneous multiscale method, Acta Numer., 21 (2012), 1-87.
  • [4] R. A. Adams and J. J. F. Fournier, Sobolev Spaces, 2nd eds., Elsevier, Amsterdam, 2003.
  • [5] D. Arjmand and O. Runborg, A time dependent approach for removing the cell boundary error in elliptic homogenization problems, J. Comput. Phys., 341 (2016), 206-227.
  • [6] I. Babuška, Homogenization and its applications. mathematical and computational problems. in Numerical Solutions of Partial Differential Equations-III, B. Hubbard, eds., New York, Academic Press, 1976, 89-116.
  • [7] I. Babuška, B. A. Szabo and I. Katz, The p-version of the finite element method, SIAM J. Numer. Anal., 18 (1981), 515-545.
  • [8] M. Barrault, Y. Maday, N. Nguyen and A. Patera, An empirical interpolation method: application to efficient reduced-basis discretization of partial differential equations, C. R. Math. Acad. Sci. Paris, 339 (2004), 667-672.
  • [9] A. Bensoussan, J.-L. Lions and G. Papanicolaou. Asymptotic Analysis for Periodic Structures, North-Holland Publishing Co., Amsterdam-New York-Oxford, 1978.
  • [10] X. Blanc and C. Le Bris, Improving on computation of homogenized coefficients in the periodic and quasi-periodic settings, Netw. Heterog. Media, 5 (2010), 1-29.
  • [11] L. Bos, J.-P. Calvi, N. Levenberg, A. Sommariva and M. Vianello, Geometric weakly admissible meshes, discrete least squares approximations and approximate Fekete points, Math. Comp., 80 (2011), 1623-1638.
  • [12] S. C. Brenner and L. R. Scott, The Mathematical Theory of Finite Element Methods, 3rd eds., Springer Science + Business Media, LLC, 2008.
  • [13] J.-P. Calvi and N. Levenberg, Uniform approximation by discrete least squares polynomials, J. Approx. Theory, 152 (2008), 82-100.
  • [14] P. G. Ciarlet, The Finite Element Method for the Elliptic Problems, North-Holland Publishing Co., Amsterdam-New York-Oxford, 1978.
  • [15] B. DeVolder, J. Glimm, J.W. Grove, Y. Kang, Y. Lee, K. Pao, D.H. Sharp and K. Ye, Uncertainty quantification for multiscale simulations, J. Fluids Eng., 124 (2002), 29-41.
  • [16] R. Du and P. B. Ming, Heterogeneous multiscale finite element method with novel numerical integration schemes, Commun. Math. Sci., 8 (2010), 863-885.
  • [17] W. E, Principles of Multiscale Modeling, Cambridge University Press, Cambridge, 2011.
  • [18] W. E and B. Engquist, The heterogeneous multiscale methods, Commun. Math. Sci., 1 (2003), 87-132.
  • [19] W. E, P. B. Ming and P. W. Zhang, Analysis of the heterogeneous multiscale method for elliptic homogenization problems, J. Amer. Math. Soc., 18 (2005), 121-156.
  • [20] Y. Efendiev and T. Y. Hou, Multiscale Finite Element Methods: Theory and Applications, Springer Science + Business Media, LLC, 2009.
  • [21] B. Engquist, H. Holst and O. Runborg, Multiscale methods for wave propagation in heterogeneous media over long time, in Numerical Analysis of Multiscale Computations, B. Engquist et al. (eds.), Springer-Verlag, Berlin, Heidelberg, 2012, 167-186.
  • [22] B. Engquist and P. E. Souganidis, Asymptotic and numerical homogenization, Acta Numer., 17 (2008), 147-190.
  • [23] A. Gloria, Reduction of the resonance error - part 1: approximation of homogenized coefficients, Math. Models Methods Appl. Sci., 21 (2011), 1601-1630.
  • [24] A. Gloria and M. Habibi, Reduction in the resonance error in numerical homogenization II: correctors and extrapolation, Found. Comput. Math., 16 (2016), 217-296.
  • [25] A. Gloria, S. Neukamm and F. Otto, Quantification of ergodicity in stochastic homogenization: optimal bounds via spectral gap on Glauber dynamics, Invent. Math., 199 (2015), 455-515.
  • [26] P. Grisvard, Elliptic Problems in Nonsmooth Domains, Pitman, Boston, 1985.
  • [27] F. Hecht, New development in FreeFem++, J. Numer. Math., 20 (2012), 251-265.
  • [28] Y. F. Huang, Study on some efficient algorithms for multiscale partial differential equations, PH. D dissertation, Chinese Academy of Sciences, 2018.
  • [29] P. Kanouté, D. P. Boso, J. L. Chaboche and B. A. Schrefler, Multiscale methods for composites: a review, Arch. Comput. Methods Eng., 16 (2009), 31-75.
  • [30] J. B. Keller, A theorem on the conductivity of a composite medium, J. Mathematical Phys., 5 (1964) 548-549.
  • [31] V. Khoromskaia, B. N. Khoromskij and F. Otto, Numerical study in stochastic homogenization for elliptic partial differential equations: convergence rate in the size of representative volume elements, Numer. Linear Algebra Appl., 27 (2020), e2296.
  • [32] S. M. Kozlov, Averaging of differential operators with almost periodic rapidly oscillating coefficients, Math. Sb., 107 (1978), 199-217, 317.
  • [33] A. Kroó and S. Révész, On Bernstein and Markov-type inequalities for multivariate polynomials on convex bodies, J. Approx. Theory, 99 (1999), 134-152.
  • [34] A. Kroó and J. Szabados, Markov-Bernstein type inequalities for multivariate polynomials on sets with cusps, J. Approx. Theory, 102 (2000), 72-95.
  • [35] R. Li, P. B. Ming and F. Y. Tang, An efficient high order heterogeneous multiscale method for elliptic problems, Multiscale Model. Simul., 10 (2012), 259-283.
  • [36] R. Li, P. B. Ming, Z. Y. Sun and Z. J. Yang, An arbitrary-order Discontinuous Galerkin method with one unknown per element, J. Sci. Comput., 80 (2019), 268-288.
  • [37] R. Li and F. Y. Yang, A discontinuous Galerkin method by patch reconstruction for elliptic interface problem on unfitted mesh, SIAM J. Sci. Comput., 42 (2020), A1428-A1457.
  • [38] A. Markov, Sur une question posee par Mendeleieff, Bull. Acad. Sci. St. Petersburg, 62 (1889), 1-24.
  • [39] K. Matouš, M. G. D. Geers, V. G. Kouznestova and A. Gillman, A review of predictive nonlinear theories for multiscale modeling of heterogeneous materials, J. Comput. Phys., 330 (2017), 192-220.
  • [40] J. C. Michel, H. Moulinec and P. Suquet, Effective properties of composite materials with periodic microstructure: a computational approach, Comput. Methods Appl. Mech. Engrg., 172 (1999), 109-143.
  • [41] P. B. Ming and X. Y. Yue, Numerical methods for multiscale elliptic problems, J. Comput. Phys., 214 (2006), 421-445.
  • [42] P. B. Ming and P. W. Zhang, Analysis of the heterogeneous multiscale method for parabolic homogenization problems, Math. Comp., 76 (2007), 153-177.
  • [43] J.-C. Mourrat, Efficient methods for the estimation of homogenized coefficients, Found. Comput. Math., 19 (2019), 435-483.
  • [44] N. C. Nguyen, A multiscale reduced-basis method for parametrized elliptic partial differential equations with multiple scales, J. Comput. Phys., 227 (2008), 9807-9822.
  • [45] F. Piazzon, Optimal polynomial admissible meshes on some classes of compact subsets of d\mathbb{R}^{d}, J. Approx. Theory, 207 (2016), 241-264.
  • [46] E. A. Rakhmanov, Bounds for polynomials with a unit discrete norm, Ann. of Math., 165 (2007), 55-88.
  • [47] C. Schwab, p- and hp-Finite Element Methods, Oxford University Press, New York, 1998.
  • [48] H. J. Schmid, On cubature formulae with a minimum number of knots, Numer. Math., 31 (1978), 281-297.
  • [49] J. Shen, T. Tang and L. L. Wang, Spectral Method: Algorithms, Analysis and Applications, Analysis and Applications, Springer-Verlag, Berlin, Heidelberg, 2011.
  • [50] P. Šolín, K. Segeth and I. Doležel, Higher-order Finite Element Methods, Chapman & Hall/CRC, Boca Raton, FL, 2004.
  • [51] A. H. Stroud, Approximate Calculation of Multiple Integrals, Prentice-Hall, Inc., Englewood Cliffs, 1971.
  • [52] D. O. Sullivan, Exploring spatial process dynamics using irregular cellular automaton models, Geogr. Anal., 33 (2001), 1-18.
  • [53] L. Tartar, H-convergence, Course Peccot, Collége de France. Partially written by F. Murat, Séminaire d’Analyse Fonctionnelle et Numérique de l’Université d’Alger, 1977-1978, March 1977.
  • [54] L. Tartar, The General Theory of Homogenization: A Personalized Introduction, Springer-Verlag, Berlin; UMI, Bologna, 2009.
  • [55] S. Torquato, Random Heterogeneous Materials: Microstructure and Macroscopic Properties, Springer Science+Business Media, New York, 2002.
  • [56] R. Verfürth, On the constants in some inverse inequalities for finite element functions, Technical Report 257, University of Bochum, 1999.
  • [57] H. Wendland, Scattered Data Approximation, Cambridge University Press, Cambridge, 2005.
  • [58] D. R. Wilhelmsen, A Markov inequality in several dimensions, J. Approx. Theory, 11 (1974), 216-220.
  • [59] X. Y. Yue and W. E, The local microscale problem in the multiscale modeling of strongly heterogeneous media: effects of boundary conditions and cell size, J. Comput. Phys., 222 (2007), 556-572.
  • [60] L. B. Zhang, A Parallel algorithm for adaptive local refinement of tetrahedral meshes using bisection, Numer. Math. Theor. Meth. Appl., 2 (2009), 65-89.
  • [61] T. Zohdi and P. Wriggers, An Introduction to Computational Micromechanics, Springer-Verlag, Berlin, Heidelberg, 2005.