An Efficient Online-Offline Method for Elliptic Homogenization Problems
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, Stability2000 Mathematics Subject Classification:
35B27, 65N30, 74S05, 74Q051. Introduction
We consider a prototypical elliptic boundary value problem
(1.1) |
where is a small parameter that signifies explicitly the multiscale nature of the problem. We assume that the coefficient , which is not necessarily symmetric, belongs to a set that is defined by
(1.2) | ||||
where is a bounded domain in and denotes the inner product in , while is the corresponding norm.
In the sense of H-convergence [53, 54], for every sequence and , the sequence of the solution to (1.1) satisfies
(1.3) |
where is the solution of the homogenization problem
(1.4) |
and . Here and are standard Sobolev spaces [4].
The quantities of interest for Problem (1.1) and Problem (1.4) are the homogenized solution over the whole domain and the solution 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 . A typical way that towards this is provided by the heterogeneous multiscale method (HMM) [18, 3], and the FEmethod [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 by solving the cell problems posed on the sampling points of the macoscopic solver. At each point, one needs to solve cell problems with 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 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 with the mesh size of the macroscopic solver. This method achieves higher-order accuracy with almost the same cost of HMM with 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 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 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.,
For any measurable set , we define the mean of an integrable function over as
We shall also use the discrete norm for any as
Throughout the paper the generic constant may be different from line to line, while it is independent of and the mesh size parameters .
2. The Offline-online Method
The macroscopic solver is chosen as the standard Lagrange finite element method ( FEM) [14]. is defined as the set of polynomials with degree less than for the sum of all variables. The finite element space is denoted by corresponding to the triangulation with mesh size that is the maximum of the element size for all elements , where is the diameter of . We assume that all the elements in satisfy the shape-regular condition in the sense of Ciarlet and Raviart [14], i.e., there exists a constant such that , where is the diameter of the smallest ball inscribed into , and 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 as follows.
Offline We firstly construct a sampling triangulation with mesh size over domain . For simplicity, we assume that consists of simplices, and is assumed to be shape-regular with the chunkiness parameter . On each element , the approximation effective matrix is reconstructed by solving a least-squares: for ,
(2.1) |
Here is the set of all sampling points that belong to , where is a patch of elements around , which usually includes . Its precise definition will be given later on. We refer to Fig. 1 for an example of such .
At each sampling point , the effective matrix is defined by averaging the flux arising from the cell problems:
(2.2) |
where the cell with and the cell size. Here for , satisfies
(2.3) |
Online Given , we find such that
(2.4) |

(a). Example of constructed by including all the Moore neighbors.

(b). Example of and the set consists of the black dots.
Remark.
The element 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 for any element . We start from assigning a threshold value that is used to control the size of . There are several different ways to find . One way is to define in an recursive way as in [35]: For any , we let
(2.5) |
Once , we stop the construction and let . This means that we add the Moore neighbors [52] to 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 in all the tests in § 4 are defined as in (2.5). We denote by the set containing all the sampling points that belong to . In all the tests in § 4, we use the barycentric of each element as the sampling point. There are also other choices for construction . 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 in and 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
where is the Frobenius norm of a by matrix .
Similar to [35, Lemma 3.1, Lemma 3.2], the following lemma gives the error estimates of the proposed method.
Lemma 3.1.
Let be the homogenized solution with the homogenized matrix in . If , then there exists a unique solution satisfying (2.4).
If for any , then
(3.1) |
where appears in the discrete Poincaré inequality that only dependends on :
Moreover, there exists depending only on and such that
where is the unique solution of the problem:
(3.2) |
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 defined on as follows. Let be the solution of the least-squares
(3.3) |
We imbed into a global operator as .
Given the reconstruction operator , we decompose as
(3.4) | ||||
where is a piecewise matrix posed on that is defined by
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 and .
Assumption A For every , there exist constants and that are independent of such that with , and is star-shaped with respect to , where is a disk with radius .
This assumption concerns the geometry of , which is crucial for the uniform boundedness of . The motivation for this assumption lies in the following Markov inequality [38]:
(3.5) |
Here . This inequality is proved in [36, Lemma 5], which is a combination of [57, Proposition 11.6] and the fact that satisfies the uniform interior cone condition [4], which is a direct consequence of Assumption A.
In [35], the authors make the following assumption on .
Assumption A’ is a bounded convex domain and there exists such that .
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) |
where is the width of , which is the minimum distance between parallel supporting hyperplanes of .
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 independent of the diameter of but depends on the shape of such that for all ,
(3.7) |
The index is sharp if is a locally Lipschitz domain. However, for a cuspidal domain, Kroó and Szabodos [34] proved that if is a Lip-domain with 111A typical Lip-domain is the unit ball, i.e., with ., then there exists depends on the shape of such that
(3.8) |
where the index 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 in (3.7) is only known for with special shape in the literature, e.g., is an interval [46], and is a simplex [56, 33]. All the prefactors are important for us to derive the realistic conditions that ensure the uniform boundedness of ; cf., Lemma 3.3.
On the other hand, The authors in [35] derived an explicit expression of that depends on the recursion depth and the chunkiness parameter under Assumption A’. It seems that the convexity of the patch is rather restrictive in implementation, particularly for an -shape domain. Assumption A is less restrictive and is easy to check in practice.
Assumption B For any and ,
This assumption concerns the cardinality of the sampling set , 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 is at least to ensure the unisolvence of the discrete least-squares. A quantitative version of this assumption is
with
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 .
Lemma 3.2.
If Assumption B holds, then there exists a unique solution of (3.3) for any . Moreover satisfies
The stability result is valid for any and , and
(3.9) |
The quasi-optimal approximation property is valid in the sense that
(3.10) |
If Assumption A and Assumption B are valid, then for any , there exists
(3.11) |
such that for the perturbed sampling set , there exists a unique satisfying
(3.12) |
where is a ball centered at with radius .
If Assumption A’ and Assumption B are valid, then the perturbation result (3.12) remains true with
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 , more refined estimates on the accuracy of may be found in [36, Lemma 4]. The perturbation estimate (3.12) shows that the set of the sampling nodes perturbed a little bit remains a norming set with a slightly bigger upper bound, i.e., for all ,
This perturbation estimate has been encapsulated in an abstract form in [36, Lemma 2], while there is no proof for the perturbed reconstruction operator .
Proof.
For any , we let , then for any with given by (3.11), by Taylor’s expansion, we obtain
By Assumption A, the Markov inequality (3.5) is valid, and we obtain
Using Assumption B and the above two inequalities, we obtain
which immediately implies
(3.13) |
Applying the above perturbation estimate to , we obtain
This also gives the existence and uniqueness of .
By the definition of , we obtain
A combination of the above two inequalities and the fact that give (3.12).
The following lemma ensures the uniform boundedness of .
Lemma 3.3.
If Assumption A holds, then for any , if then we may take
(3.14) |
Moreover, if then we may take
If Assumption A’ holds, then for any , if then we may take
(3.15) |
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 such that , and such that . Then
If Assumption A holds, then we usually have , the estimate (3.14) suggests that implies the uniform boundedness of . This means that cannot be too narrow in certain directions. If Assumption A’ holds, then the estimate (3.15) shows that , which immediately implies . Both conditions show that a relative large patch is required for the reconstruction. Furthermore, if is a cuspidal domain, then we may use (3.8) to prove that if
with a constant depending on the shape of , then the bound (3.14) is also valid. This condition indicates that has a recursive depth with . An even larger patch is required to ensure the stability. This may occur for domain with complicated boundary or rough boundary.
It remains to find an upper bound for . This is a direct consequence of Assumption A or Assumption A’ and the shape regularity of .
Lemma 3.4.
If is shape-regular and Assumption A or Assumption A’ is valid, then
(3.16) |
Proof.
For any element , using Assumption A or Assumption A’ and note that there is only one sample point inside each element , we obtain
where stands for the volume of a -dimensional ball with radius .
Using the shape regularity of , we obtain
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 . For the two ways based on either the Moore neighbor or the von Neumann neighbor, we have with 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 should be quite large, equivalently, is large, so that the uniform boundedness of is valid. In numerical tests below, we observe that the method still works quite well even when is far less than the theoretical threshold. We refer to [35] for a list of the size of and the upper bound of .
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 , then there exists that depends on and but independent of .
(3.17) |
where .
It is worthwhile to mention that is the so-called estimating error in HMM [18]. There are many works devoted to bounding and developing new algorithms to improve the estimates; see; e.g., [19, 59, 10, 23, 5, 24, 43] and the references therein.
Proof.
Theorem 1.
Let be the homogenized solution with . If Assumption A or Assumption A’ and Assumption B hold, and , then there exists a unique solution of Problem (2.4) that satisfying
(3.19) |
Moreover, if the solution of the auxiliary problem (3.2) admits a unique solution satisfying the regularity estimate , then
(3.20) |
In view of the above error estimates (3.19) and (3.20), we may have or . For a fixed , one may increase 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) |
where is the solution of HMM-LS method. For a fixed , the above error estimate indicates that to balance the error provided that is sufficiently small, the number of the cell problems in this case is the total number of the vertices of the online triangulation . 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 are uniform partitions of , and we list the threshold number ( is defined in § 2) of the sampling points for different orders of reconstruction in Table 1, the number should be slightly larger than the number that lists in the table when abuts the boundary of the domain. For , we only report because this is the only case we test in this part, the examples for the lower order reconstruction may be found in [28].
5 | 7 | 13 | |
NONE | NONE | 27 |
Noted that the theory covers the case when consists of polygons or polytopes, we refer to [36] for the demonstration of such patches. It is worth pointing out that the nonuniform is useful when the shape of 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 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 norm and norm:
The exact homogenized solution is generated by discretization (1.4) with FEM over a very refined mesh in most cases unless otherwise stated. In all the tests we set .
4.1. Problems with local periodic coefficient in two dimension
We consider an example with locally periodical coefficient in , which is taken from [41].
Example 1.
where , and
where denotes the by identity matrix. The homogenization problem is
(4.1) |
A direct calculation gives the following analytical expression of :
(4.2) |
The offline triangulation consists of a uniform squares. To obtain the online triangulation , we firstly triangulate into a uniform 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 for reconstruction so that . We denote this numerical solution by . Fig. 2 presents the accuracy of the offline computation, which corroborates the estimate (3.17).
To study the convergence rate of the method, it is more convenient to reshape the error estimates in Theorem 1 in terms of and as follows.
(4.3) |
Balancing the discretization error and the reconstruction error, we set as
Following this refinement strategy, we plot the relative error and error in Fig. 3, which are consistent with the estimates (4.3).

(a). Rate of convergence in .

(b). Rate of convergence in .
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 FEM with the meshsize parameter . Fig. 4 clearly shows that the higher-order reconstruction is more accurate with less cost.

(a). The error with different reconstruction orders.

(b). The error with different reconstruction orders.
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 FEM with the meshsize parameter . Instead of using the analytical expression of for reconstruction, we solve the periodic cell problems
(4.4) |
with FEM, and employ (2.2) to obtain . We denote the reconstructed solution as , which is very close to in the sense
This means that the 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.
Q | cell problems | Relative error |
|
Total Time | |||
---|---|---|---|---|---|---|---|
45 | 2025 | 6.03e-4 | 348.65s | 363.27s | |||
25 | 625 | 5.92e-4 | 106.97s | 110.76s | |||
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 error bound in (3.21) may be rewritten as
We follow the setup in the last test to solve the cell problems so that is negligible in the above estimate. We take and to balance the error so that
In the offline-online method, we use FEM as the macroscopic solver and choose the reconstruction order and . To balance the error of (4.3), we take for and for and we obtain . It seems that the relative 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 for and is of for , while the number of the cell problems in HMM-LS is of .
(a). Relative error.
(b). Running time and the number of the cell problems.
4.2. Problems posed on L-shape domain in
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 is nonsmooth [26].
Example 2.
In the offline stage, the periodic cell problems (4.4) are solved by FEM over a mesh of size . The third order reconstruction is used to obtain . In the online stage, FEM is used as the macroscopic solver. Then the error estimate roughly reads as
(4.5) |
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
Firstly we set TOL. The triangulation is plot in Fig. 6a, we have sampling points in total. Under these settings, in the offline stage. let 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 iterations in Fig. 6b.
(a). Offline sampling mesh .
(b). Online adaptive mesh after iterations.
The relative error is plot in Fig. 7a, and we observe that the error decays with optimal rate of convergence .
(a). Tolerance.
(b). Tolerance.
Next we decrease the error tolerance TOL to , and we also use the third order reconstruction to approximate and there are sampling points in total. The error caused by the reconstruction is . The relative error is plot in Fig. 7b, and we observe that the error decays with optimal rate of convergence .
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 iterations with 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 error, while the running time of the offline-online method is only around of HMM. This is easily understood because the number of the cell problems is only of HMM.
cell problems | Relative 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.
We write the error estimate roughly as
(4.7) |
where DOF is the total degrees of freedom in the online computation, and the factor comes from the approximation error caused by using Fourier spectral method [49] to solve the periodic cell problems (4.4), and is a universal constant and is the points we used in each cell. The offline refinement strategy reads as
We set TOL and . We firstly divide the domain into subdomain as in the last example, next we triangulate each subdomain by a uniform mesh with . There are sampling points in total; see Fig. 8a. We choose . The error is plot in Fig. 8b. We observe that the error decays with an optimal convergence rate .

(a). Offline sampling mesh .

(b). H1 error with tolerance.
4.3. Example with almost periodic coefficient
Example 4.
Such coefficient belongs to Kozlov class [32] and the example is adapted from [24]. The coefficients and are visualized in Fig. 9 with .
(a). plot of .
(a). plot of .
The effective matrix is given by
(4.9) |
where and are the solutions of
(4.10) |
As oppose to the locally periodic case, there is no explicit expression of . The naive approach to approximate (4.9) consists in replacing by which are solutions of a truncated cell problem
and is defined by
We take and use FEM over a mesh to solve the above truncated cell problem and obtain
In the offline stage, we choose and use FEM over a mesh with size to solve the cell problems (2.3). Next we plot the relative error and the relative error in Fig. 10, which clearly shows the higher-order reconstruction is more accurate.

(a). The relative error with different reconstruction orders.

(b). The relative error with different reconstruction orders.
Finally we compare the number of the cell problems and the running time among reconstructions of different orders in Table 4.
Q | cell problems | Relative error | Time | |
---|---|---|---|---|
20 | 400 | 6.08e-3 | 1083.08s | |
14 | 196 | 4.95e-3 | 546.24s |
The second order reconstruction takes only about 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) |
where is a random checker-board, and . is constructed by partitioning into uniform square cells of size , each of which is randomly designated as or with probability and , respectively. We visualize one realization of the random coefficients in Fig. 11 with . Theorem 1 remains true with a minor modification of , we refer to [19] for related result.


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 , we take as the cell size and use FEM over a mesh of size to solve the cell problem (4.4). We set the offline mesh size as and use third order reconstruction to get an approximation effective matrix , which will be exploited to obtain the reference solution in the following test.
To justify this approach, we choose at three representative points in : , which is one of the maximum point of over ; , which is one of the minimum point of over , and , which is one of the maximum point of over . We let and use FEM over a mesh of size to solve the cell problems (4.4) posed over these three points. We denote by the approximation effective matrix for th realization with . In addition, we use the empirical average
(4.12) |
as the proxy of the expectation , where is the total number of the realization. We take in the simulation, and the results are reported in Table 5.
Secondly, for , we measure the variance as
and
Fig. 12a and Fig. 12b suggest that and decay as , which is consistent with the theoretical predictions [25]. A systematical numerical tests for the variance may be found in a recent work [31].

(a). Decays of with respect to the cell size .

(b). Decays of with respect to the cell size .
Finally we compare the effect of the reconstructions with different orders in the offline computation. We fix the online solver as FEM over a mesh with size . In the offline stage, we choose and use FEM over a mesh with size to solve the cell problems (2.3). We take and in the tests, and compute the ensemble average of the relative error and error
The expectation is replaced by the empirical average as that in (4.12) with realizations. The visualization in Fig. 13a and Fig. 13b clearly shows that the second order reconstruction is more accurate than the first order reconstruction.

(a). Ensemble average of the error.

(b). Ensemble average of the error.
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 , 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.