A mixed multiscale finite element method for elliptic problems with oscillating coefficients

By Zhiming Chen and Thomas Y. Hou

Abstract

The recently introduced multiscale finite element method for solving elliptic equations with oscillating coefficients is designed to capture the large-scale structure of the solutions without resolving all the fine-scale structures. Motivated by the numerical simulation of flow transport in highly heterogeneous porous media, we propose a mixed multiscale finite element method with an over-sampling technique for solving second order elliptic equations with rapidly oscillating coefficients. The multiscale finite element bases are constructed by locally solving Neumann boundary value problems. We provide a detailed convergence analysis of the method under the assumption that the oscillating coefficients are locally periodic. While such a simplifying assumption is not required by our method, it allows us to use homogenization theory to obtain the asymptotic structure of the solutions. Numerical experiments are carried out for flow transport in a porous medium with a random log-normal relative permeability to demonstrate the efficiency and accuracy of the proposed method.

1. Introduction

Many problems of fundamental and practical importance in science and engineering have multiple-scale solutions. Typical examples include composite materials with fine micro-structures and highly heterogeneous porous media. The direct numerical simulation of problems involving multiscale solutions is difficult, due to the requisite of tremendous amount of computer memory and CPU time, which can easily exceed the limit of today’s computer resources. On the other hand, in practice, it is often sufficient to predict the large scale solutions to a certain accuracy. Thus, various methods of upscaling or homogenization have been developed which replace the governing equations with multiscale solutions by the homogenized equations, whose solutions can be resolved on a coarse-scale mesh.

The recently developed multiscale finite element method Reference 18, Reference 19, Reference 13 provides an effective way to capture the large-scale structures of the solutions on a coarse mesh. The central idea of the method is to incorporate the local small-scale information of the leading-order differential operator into the finite element bases. It is through these multiscale bases and the finite element formulation that the effect of small scales on the large scales is correctly captured. Apart from the computational advantages, such as saving in computer memory and good parallel efficiency Reference 18 of the multiscale finite element method, we stress that the real significance of the method lies in its ability to solve the problems in coarse meshes. This is particularly advantageous when multiple simulations or realizations are necessary due to changes of boundary conditions or source functions for given fine microstructures of composite materials or highly heterogeneous permeability of porous media. Other relevant works on constructing special finite element bases can be found in Reference 3 for layered microstructures and in Reference 6 for convection-dominated diffusion problems.

The motivation of this paper is to explore the possibility of applying the multiscale finite element method to numerical computation of flow transport in highly heterogeneous porous media. In its simplest form, neglecting the effect of gravity, compressibility, capillary pressure, and considering constant porosity and unit mobility, the governing equations for the flow transport can be described by the following partial differential equations (see Reference 21, Reference 31, and Reference 12):

where is the pressure, is the water saturation, is the relative permeability tensor, and is the Darcy velocity. The highly heterogeneous properties of the medium are built into the permeability tensor , which is generated through the use of sophisticated geological and geostatistical modeling tools. The detailed structure of the permeability coefficients makes the direct simulation of the above model infeasible. For example, it is common in real simulations to use millions of grid blocks, with each block having a dimension of tens of meters, whereas the permeability measured from cores is at a scale of centimeters Reference 24. This gives more than degrees of freedom per spatial dimension in the computation. This makes a direct simulation to resolve all small scales prohibitive even with today’s most powerful supercomputers. On the other hand, from an engineering perspective, it is often sufficient to predict the macroscopic properties of the solutions. Thus it is highly desirable to derive effective coarse grid models to capture the correct large solution without resolving the small-scale features. Numerical upscaling is one of the commonly used approaches in practice. There are extensive studies in the literature on the numerical upscaling of two-phase flows through porous media, see, e.g., Reference 1, Reference 21, Reference 11, Reference 14, Reference 30.

Mixed finite element approximations for second order elliptic problems, which approximate the source variable and flux simultaneously, have been studied by many authors (cf., e.g., Reference 27 and the book Reference 5). The local conservation of velocity flux is an important property in the mixed finite element methods. The violation of this local conservation property will lead to leakage of velocity flux. This will deteriorate the accuracy of the numerical solution for long-time computations. This is the reason why mixed finite element methods are attractive for porous medium simulations Reference 28 (see also the numerical results in §5.1 of this paper). In this paper, we first propose and analyze a mixed multiscale finite element method with an over-sampling technique for solving elliptic equations with oscillating coefficients, and then apply the method to compute the above flow model Equation 1.1-Equation 1.2. The use of the over-sampling technique is crucial in eliminating large resonance errors from the element boundaries. This point has already been well demonstrated in the previous study Reference 18Reference 13 for the displacement multiscale finite element method. We will demonstrate that this technique applies also to the mixed multiscale finite element method. Our computational results show that the mixed multiscale method with the over-sampling technique gives more accurate results than the corresponding method without such technique.

We have performed careful numerical experiments to validate our mixed multiscale finite element method. To test how well our method works in a realistic application, we apply our method to a flow in porous media with a random log-normal relative permeability tensor. This is one of the practical benchmark test cases for oil recovery problems. Our computational results demonstrate convincingly the importance of the local conservation property in the flow simulation. When the local conservation property is not satisfied, the fractional flow curve deviates significantly from that obtained from the well-resolved mixed finite element calculation after a short time. The mixed multiscale finite element method, on the other hand, provides an accurate numerical approximation even on a coarse grid, with accuracy comparable to that obtained using the standard fine-mesh mixed finite element method. Finally, we compare the performance of the mixed multiscale method studied in this paper with the displacement multiscale finite element method studied in Reference 18, Reference 19, Reference 13 when applied to the flow transport problem in heterogeneous porous media. The two methods give similar results at an early stage. But the fractional flow curve obtained by the displacement multiscale method deviates significantly from that obtained by the well-resolved calculation after a short time, clearly suffering from the violation of the local conservation property. The mixed multiscale method gives a more accurate solution than the displacement multiscale method for long-time calculations, due to the inherent local conservation property of the mixed multiscale method.

The ultimate goal of this study is to produce an effective coarse grid model for the two-phase flow with heterogeneous porous media. To this end, we need to upscale the saturation equation. Without capillary pressure, the saturation equation is hyperbolic. The effective equation is difficult to derive and has a nonlocal memory effect Reference 29. Using an upscaling technique recently developed in Reference 14 for the saturation equation, we show how the proposed mixed multiscale finite element method leads to a complete coarse grid algorithm. Our numerical results demonstrate convincingly that the fractional flow curve obtained from the resulting coarse grid model gives a very good approximation to that obtained from the fine grid calculation. Typically, many realizations on the same microstructure (permeability field) are made due to changes in boundary conditions and source fields. In such a case, the multiscale finite element bases are only constructed once, and can be used in the subsequent computations. Thus, the coarse grid model offers substantial saving in both memory and computing time. The saving could be as large as a factor of 10,000 if one can scale up by a factor of 10 in each space dimension (three space dimensions plus time).

The outline of the paper is as follows. In §2 we introduce the mixed multiscale finite element methods and present the main convergence results. In §3 we review the homogenization results for elliptic problems with Neumann boundary conditions. These results are the basis of our convergence analysis. In §4 we prove the error estimate for the mixed multiscale finite element method with over-sampling introduced in §2.3. The analysis depends on an abstract formulation for nonconforming mixed finite element methods, the homogenization theory, and a technique of “freezing coefficients” to deal with the local periodic coefficients. In §5 we apply our method to simulate the flow transport model Equation 1.1-Equation 1.2 for a practical random log-normal permeability.

2. Mixed multiscale finite element methods

2.1. Notation and background

Let be a polyhedral domain in with boundary whose unit outer normal is denoted by . Let be a domain in with the Lipschitz boundary . For each integer and , we denote by the standard Sobolev space of real functions having all their weak derivatives of order up to in the Lebesgue space . The norm and the seminorm of will be denoted by and , respectively. As usual, when , is denoted by with the norm and the seminorm . In the following, we let denote the subspace of whose functions have zero average over .

We consider the following second order elliptic equations with locally periodic coefficients Reference 4:

In this paper the usual Einstein summation convention for repeated indices is used. Here, is assumed to be a small parameter, and is a symmetric matrix which satisfies the uniform ellipticity condition:

for some positive constant . Furthermore, we assume that , where stands for the collection of all periodic functions with respect to the unit cube .

Let be the solution of the homogenized problem of Equation 2.1-Equation 2.2:

where with

and is the periodic solution of the cell problem

Here is the Kronecker delta, i.e., for and for . Note that in Equation 2.7 plays the role of a parameter. However, since is differentiable in , we can easily show that is also differentiable in . The convergence property of to will be studied in detail in the next section.

Throughout the paper we impose the following assumptions on the data:

(H1) , on for some .

(H2) Compatibility: .

We remark that the assumption on the boundary value is not very restrictive. For example, in the case when is a convex polygon in with sides , and , such a flux can be constructed through with being the solution of the equation in with the Neumann boundary condition on . The regularity results in Reference 17 then ensure . We also stress that the explicit formula of such a flux is not required in the definition of our methods.

Denote by the subspace of whose functions have zero average over , and let be the subspace of given by

The norm of will be denoted by . Let and the inverse matrix of ; then . The mixed formulation to Equation 2.1-Equation 2.2 then reads as follows: find a pair such that on and

Here stands for the inner product of or . The existence of a unique solution of problem Equation 2.8-Equation 2.9 and its equivalence to Equation 2.1-Equation 2.2 follow directly from the abstract Babuska-Brezzi theory Reference 5, Reference 27.

Let us assume that is a regular and quasi-uniform partition of into simplices. For any , let be its diameter, its Lebesgue measure, the unit outer normal to , and the surfaces or edges of with being the measure of . The mixed multiscale finite element method that we are going to introduce is closely related to the Raviart-Thomas elements Reference 27. For any , we denote by the lowest order Raviart-Thomas approximation of :

where is the constant element space. Let be the basis of which satisfies

Since is constant in , it follows easily from Green’s formula that . Let be the lowest order Raviart-Thomas finite element space. It is well-known that there exists an interpolation operator such that satisfies the relations

and the error estimate

Let be the standard piecewise constant finite element space for approximating . Then the standard mixed finite element method using the lowest order Raviart-Thomas element leads to the following discrete problem: find a pair such that on and

The following well-known error estimate can be obtained by using the Babuska-Brezzi theory Reference 27, Reference 5

Note that since Reference 19, the error estimate Equation 2.12 implies that the mesh size must satisfy to obtain accurate approximations. The purpose of the mixed multiscale finite element methods to be introduced next is to remove such a strong requirement on the mesh size . This is achieved by building the local small-scale information into the finite element bases.

2.2. Mixed multiscale finite element method

Recalling that denotes the subspace of whose functions have zero average over , we define as the solution of the following Neumann problem over , for :

Equation Equation 2.13 is the weak formulation of the following boundary value problem

Now let and denote by the multiscale finite element space spanned by . Recall that . We have

Moreover, for any , the relations

imply in . The degrees of freedom for can be chosen as the zeroth order moments of on the sides or faces of . In practical applications, the base functions of will be approximated by solving Equation 2.13 on a triangulation of with a mesh size resolving using the lowest-order Raviart-Thomas mixed finite element method.

Let , , and . We define as the following multiscale finite element space for approximating the flux :

Let . To introduce an approximation of the boundary data , we let be the collection of all sides or faces of the triangulation which lie on the boundary . For any such that , we let be the corresponding multiscale basis and define on with given by

We can now introduce the following discretization of Equation 2.8-Equation 2.9: find a pair such that on and

We have the following theorem concerning the convergence of this discrete problem.

Theorem 2.1.

The discrete problem Equation 2.14-Equation 2.15 has a unique solution such that on . Moreover, if the homogenized solution and the assumptions (H1)-(H2) are satisfied, then there exists a constant , independent of and , such that

The proof of this theorem is simpler than that of Theorem 2.2 below for the over-sampling mixed multiscale finite element method, and will be omitted. We remark that the error estimate Equation 2.16 is uniform as , which is in strong contrast to the error estimate Equation 2.12, which blows up as . We also observe that the error estimate Equation 2.16 deteriorates when is of the same order as the mesh size . This is due to the boundary layer effect of the multiscale bases, that is, the mismatch of the oscillating structure of and the linear behavior of the finite element solution on the boundaries , . This phenomenon is called resonant error in Reference 18. The resonant error can be significantly reduced by an over-sampling technique introduced in Reference 18. The over-sampling technique is based on the observation that the width of the boundary layer is of order . Thus, if one first constructs some intermediate base functions on a larger sample domain (with size larger than ) and uses only the interior information from these intermediate base functions to construct the actual computational bases, then the boundary layer effect can be substantially reduced, resulting in more accurate approximations.

2.3. Over-sampling technique

In this subsection we will use the over-sampling technique Reference 18 to construct new multiscale finite element bases for which the boundary layer effect of the bases , is reduced.

For any , we denote by a macro-element which contains and satisfies following condition:

(H3)

and for some positive constants independent of . The minimum angle of is bounded below by some positive constant independent of .

Let be the basis of which satisfies

Recall that stands for the subspace of whose functions have zero average over . Let be the multiscale finite element basis of defined as in Equation 2.13, i.e., , where is the solution of the following problem over , :

Then we define the over-sampling finite element basis over by

where the constants are chosen such that

The existence of the constants is guaranteed because also forms the basis of . Since , we know that

Here again, in practical applications, the base functions of will be further approximated by solving Equation 2.17 on a triangulation of with a mesh size resolving using the lowest order Raviart-Thomas finite element method. When applying the multiscale finite element method to solve Equation 1.1-Equation 1.2, these base functions can be precomputed initially to generate the coarse grid operator. So this will be only a small overhead of the overall computations. This is also a common practice in many upscaling methods Reference 10, Reference 11.

We define with norm . Let and be the projection

Let be the finite element space

and define through the relation for any , . The over-sampling mixed multiscale finite element space for approximating the flux is then defined as

We remark that, in general, . Here, we require to impose some intrinsic continuity of the zeroth order moment of the normal components of across the interelement boundaries of the triangulation . Finally, let be the closed subspace of defined by

The over-sampling mixed finite element approximation of Equation 2.14-Equation 2.15 can be formulated as follows: find a pair such that on and

Here is the approximation of the boundary value with given by

for any such that , and is the corresponding over-sampling multiscale basis.

We have the following theorem about the convergence of the over-sampling mixed multiscale finite element method. The assumption (H4) will be made precise in §4, see equation Equation 4.8. This is an assumption on the maximum bound of a boundary corrector term (in the multiple scale expansion) in terms of the gradient of the homogenized solution. The validity of this assumption can be easily proved for the corresponding Dirichlet boundary value problem, due to the maximum principle.

Theorem 2.2.

The discrete problem Equation 2.21-Equation 2.22 has a unique solution such that on . Moreover, if the homogenized solution and the assumptions (H1)-(H4) are satisfied, then there exist constants and , independent of and , such that if for all , the following error estimate holds:

The proof of this theorem, which will be given in §4, depends on an abstract formulation for nonconforming mixed finite element methods, the homogenization results for elliptic equations with locally periodic coefficients in §3, and a technique of “freezing coefficients” to deal with the local periodicity of the coefficients. This new technique can also be used to extend the convergence analysis in Reference 19, Reference 13 for displacement multiscale finite element methods to elliptic equations with locally periodic coefficients. We remark that locally periodic coefficients model a wide class of microstructures which include, in particular, the layered structures. Previous numerical experiments in Reference 13 and our numerical experiments indicate that the over-sampling technique leads to significant improvement over the direct mixed multiscale finite element method Equation 2.14-Equation 2.15.

3. Homogenization results

In this section we summarize some homogenization results for the Neumann problem in the form to be used in this paper.

Let be a domain in with Lipschitz boundary . Given and which satisfy the compatibility condition

we consider the following Neumann problem with rapidly oscillating coefficients: find such that

It is well-known that this problem admits a unique solution in due to the compatibility condition Equation 3.1. Let be the unique solution of the homogenized problem

where is given by Equation 2.6.

In the following we will assume that . This regularity is valid, for example, when , on for some , where , and is a smooth domain Reference 15 or a convex polyhedral domain in or Reference 17, Reference 23.

Now we set

where is the solution of the cell problem Equation 2.7. Then by simple calculations we get

where

Due to Equation 2.6 and Equation 2.7 we know that and . Thus there exist skew-symmetric matrices such that Reference 20, p.6

With this notation, we can rewrite

We define as the solution of the following problem:

Note that . This problem can be regarded as the weak formulation of the elliptic problem

where is the unit outer normal to . Hence plays the role of a boundary corrector in the asymptotic expansion for . The following theorem generalizes similar results in Reference 20, §1.4 for periodic coefficients.

Theorem 3.1.

Assume that . Then there exists a constant , independent of , the domain , and the data and , such that

Moreover, the boundary corrector satisfies the estimate

where stands for the length of the boundary if , and the surface area of if .

Proof.

For the sake of completeness, we sketch the proof here in order to track down the exact dependence of the constant in Equation 3.8 and Equation 3.9. First, we deduce from Equation 3.2, Equation 3.3, Equation 3.4-Equation 3.6, and Equation 3.7 that

This implies Equation 3.8 after taking and using the ellipticity condition Equation 2.3.

To show Equation 3.9, we introduce a cutoff function , in , outside the -neighborhood of the boundary , and in with independent of and . Then, we have

Note that the first term is divergence free since the matrix is skew-symmetric and with compact support. We have from Equation 3.7 that

Denote ; then and thus

This completes the proof.

An important open question is to show that the boundary corrector is of first order, i.e., for some constant independent of . Similar results for the boundary corrector for Dirichlet problem are known (cf., e.g., Reference 4, Reference 2).

4. Convergence analysis

The purpose of this section is to prove Theorem 2.2. We first recall the abstract formulation for the nonconforming mixed finite element method in §4.1, then establish the discrete inf-sup condition in §4.2, and finally prove the error estimate in §4.3.

4.1. Abstract results

The abstract results in this section modify slightly the standard Babuška-Brezzi theory Reference 5, Reference 16 in that we take into account the discretization of inhomogeneous boundary conditions. Another way to deal with the approximation of boundary conditions is considered in Reference 7, which, however, is not applicable here due to the oscillatory nature of the coefficient in Equation 2.8-Equation 2.9. Let and be two Hilbert spaces with norms and respectively. Denote by and the corresponding dual spaces, and by and the duality pairings. Let and be two continuous bilinear forms. Let be a closed subspace of . Given , , and , we consider the following variational problem: find a pair such that and

Define . The following result concerning the existence and uniqueness of solutions of the abstract problem Equation 4.1-Equation 4.2 can be easily proved by rewriting Equation 4.1-Equation 4.2 in terms of and using the standard theory in Reference 5, §II.1, Reference 16, §I.4.

Lemma 4.1.

Assume that the bilinear form is -elliptic, i.e., there exists a constant such that

and the bilinear form satisfies the following inf-sup condition for some constant :

Then, the abstract problem Equation 4.1-Equation 4.2 has a unique solution such that .

We consider now nonconforming approximations of the mixed problem Equation 4.1-Equation 4.2. Let be a Hilbert space with the norm such that . We denote by and continuous bilinear forms which are approximations of the bilinear forms and , respectively. The index will eventually refer to the mesh for which these approximations are derived. Furthermore, we assume that , where is the dual space of . Let and be finite dimensional spaces. Denote by a closed subspace of . Let be an approximation of . We consider the following approximation of Equation 4.1-Equation 4.2: find a pair such that and

Note that we have not assumed , so Equation 4.3-Equation 4.4 is a nonconforming discretization of the variational problem Equation 4.1-Equation 4.2. Let . We have the following abstract result, which accounts for the inhomogeneous boundary conditions and can be proved by modifying the argument in Reference 5, §II.2.6.

Lemma 4.2.

Assume that is uniformly coercive in , i.e., there exists a constant , independent of , such that

and satisfies the following discrete inf-sup condition for some constant independent of :

Then, the discrete problem Equation 4.3-Equation 4.4 has a unique solution such that . Moreover, the solution fulfills the following error estimate:

where the constant depends only on and the operator norms , .

The last term at the right-hand side of the above abstract error estimate, which vanishes if , represents the nonconforming error.

4.2. The discrete inf-sup condition

We will use the abstract framework in §4.1 to study the discrete problem Equation 2.21-Equation 2.22. Thus we define with the norm , and introduce the bilinear forms and as

It is clear that the operator norms satisfy and , where is the constant in Equation 2.3.

The over-sampling mixed finite element approximation Equation 2.21-Equation 2.22 is then equi-valent to the following form: find a pair such that on and

To study the properties of this discrete problem, we need the following lemma, which is crucial to the derivation of the discrete inf-sup condition for the bilinear form .

Lemma 4.3.

There exists a linear operator such that for any and we have in and for some constant independent of .

Proof.

For any and , we have for some constants We are going to define the desired operator through the relation for any with the local form of the operator defined by

This implies that in . Thus, , i.e., the operator is well-defined. Moreover, by Equation 2.20, we have in .

It remains to show the stability estimate in this lemma. To do so, we let , where is defined in Equation 2.17; then we have . Define and in . Then from the definition of , we have in and on . Now, for any , we have, after integration by parts,

By taking in above equality, we get . To conclude the proof we only need to notice that

by the assumption (H3) and Equation 2.19. This completes the proof.

Lemma 4.4.

The discrete problem Equation 4.5-Equation 4.6 has a unique solution such that . Moreover, the following error estimate holds:

Proof.

We apply the abstract Lemma 4.2. First, we note that, for any and , by Equation 2.20, in . Therefore, since on for any . Now, since is constant over , we have

which implies clearly the ellipticity of the bilinear form after using Equation 2.3.

Next, we check the discrete inf-sup condition. It follows from Equation 2.10-Equation 2.11 that the discrete inf-sup condition is true for the Raviart-Thomas element Reference 27, Reference 5:

where and is a constant independent of . We now use Lemma 4.3 to conclude that, for any ,

This completes the proof.

4.3. Proof of Theorem 2.2

The error analysis in this subsection depends crucially on the behavior of the boundary corrector in the homogenization theory. Let and be the respective solutions of problems Equation 3.2 and Equation 3.3 in §3. Let be the boundary corrector defined in Equation 3.7. We make the following assumption on the behavior of :

(H4)

There exists a constant , independent of and , such that the following estimate is valid:

This assumption, which is evident from our numerical experiments, is the basis of the over-sampling technique to reduce the resonant error. We notice that here is the boundary corrector for the homogenization problem with Neumann boundary conditions. Such an estimate has been obtained for the homogenization problem with Dirichlet boundary conditions, for which the maximum principle can be used. However, this estimate is yet to be proved for the homogenization problem with Neumann boundary conditions.

Now for any subdomain with , the interior estimate in Reference 2 yields

Thus, the assumption (H4) implies that there exists a constant , independent of , and the domains and , such that

The starting point of our error analysis is the error estimate in Lemma 4.4. We are going to estimate the three terms on the right-hand side of Equation 4.7 in order.

Lemma 4.5.

There exists a constant , independent of and such that

Proof.

It is easy to see from Equation 2.1-Equation 2.2 that

For , let us define a function

whose average over vanishes because , and thus . Now standard finite element interpolation estimates imply that

This completes the proof.

Lemma 4.6.

Under the assumptions (H1)-(H4) there exists a constant , independent of and such that

To prove this lemma, we let and define on each as

By the definition of we know that , where the operator is defined in §2. Thus, , which implies . Moreover, we have . The proof of Lemma 4.6 is then reduced to estimating .

First, we obtain from Equation 2.20 that

which, along with Equation 2.4, yields

Thus, since by Equation 2.1, we get

It remains to estimate , which we are going to do locally. For any , we denote by the extension of the linear function to defined as follows. Let ; then with , where is given in Equation 2.19. Now let in ; then , where is defined in Equation 2.10. Therefore, from Equation 2.17 we know that , where is the solution of the problem

Let be the solution of the corresponding perturbed homogenized problem

where is the local average of over , defined as

Since and is a constant symmetric matrix, it is easy to check that

Hence, there exists a unique function such that , which implies in by the uniqueness of the solution of the Neumann problem Equation 4.12 in . We remark that this technique of “freezing coefficients” is crucial in dealing with the locally periodic coefficients.

Lemma 4.7.

The solution of the problem Equation 4.11 belongs to and satisfies the following stability estimates:

Moreover, the following error estimate is valid:

Proof.

Since and , it is obvious that . By Equation 2.11 and the assumption , we have

and

Next, let ; then by an inverse estimate and Equation 2.11 we obtain that

This completes the proof.

Lemma 4.8.

Let be the solution of Equation 4.11 and . Then there exists a constant , independent of and , such that

Proof.

Note that in the definition of the homogenized problem Equation 4.12 we have used the constant coefficient instead of . Thus, the homogenization result in Theorem 3.1 cannot be directly used. However, if we let be the solution of the problem

then, using an argument similar to the proof of Theorem 3.1, we obtain from Equation 4.11-Equation 4.12 that

From the definition of in Equation 4.13 we know that . Hence

Next, since , , , and are bounded functions in , we deduce that

By taking in Equation 4.14 and using Equation 4.15, Equation 4.16, we have that

where we have used Lemma 4.7 in the second inequality. Finally, by Equation 4.9 and Lemma 4.7, we have, since from (H3),

This completes the proof.

Now we are ready to complete the proof of Lemma 4.6.

Proof of Lemma 4.6.

Let . Then it is easy to show by Lemma 4.7 that

Thus, we have, by Lemma 4.8,

Finally, by Theorem 3.1 we arrive at

This estimate, together with Equation 4.10, completes the proof.

To estimate the nonconforming error in Equation 4.7, we first recall the following result, whose proof can be found in Reference 13, Lemma 3.2.

Lemma 4.9.

Let be a periodic function with respect to the unit cube and assume that its average over vanishes. Then, for any , , we have

For any and , since , there exist constants such that , and thus by Equation 2.18. Let in ; then we have , where is defined in Equation 2.10. From Equation 2.19 we have in . Now from the definition of the basis function in Equation 2.17 we know that in , where is the solution of the problem

Let be the solution of the corresponding perturbed homogenized problem

where with

Again, the argument before Lemma 4.7 implies that in . By the finite element inverse estimates and the assumption , we easily obtain

Let . Then, using an argument similar to the proof of Lemma 4.8, we deduce that

where we have used inverse estimate in the last inequality. Now we are going to prove the following stability estimate, which plays an important role in the subsequent analysis.

Lemma 4.10.

Under the assumptions (H1)-(H4) there exist constants and , independent of and , such that if and for all , the following stability estimate is valid:

Proof.

From Equation 2.3, Equation 4.20, and Equation 4.19 we have

Next, by simple calculations we get

where is given by

Multiplying the above equality by and then integrating over , we obtain

Now let then is periodic with respect to the unit cube and . Moreover, by finite element interpolation theory, we have

Now by Lemma 4.9 and Equation 4.19 we know that

Moreover, the estimates in Equation 4.19 imply that

Substituting these estimates into Equation 4.23, we conclude that

Now the desired estimate follows from Equation 4.21 under the condition that is sufficiently small. This complete the proof.

From this lemma and Equation 4.19, we deduce that

Moreover, since and in , we have

Proof of Theorem 2.2.

By Lemmas 4.5 and 4.6 we know that we only need to estimate the nonconforming error in Lemma 4.4. Since by definition, we have for any . Thus, since , we get from Equation 2.9 that

For any and , since in and in with the respective solutions of Equation 4.17 and Equation 4.18, we deduce from Equation 4.22 that

Now we are going to estimate the terms , term by term. First, by Equation 4.20 and Equation 4.25-Equation 4.26 we get

Next, by Equation 4.25-Equation 4.26 and Lemma 4.10, we obtain

To estimate the term , we write and derive

By Cauchy’s inequality and Equation 4.25 we obtain

Next, by the argument used in deriving Equation 4.24 and by Lemma 4.9, we have

Since in , integration by parts gives

Thus, again by the argument used in deriving Equation 4.24 and by Lemma 4.9, we get

Finally, by Cauchy’s inequality and Equation 4.25, we have

Combining the above estimates, we conclude that

which, after using Theorem 3.1, yields

This completes the proof.

To conclude this section, we remark that in the proofs of Lemma 4.3 and Theorem 2.2, the choice of the constants through the relation in Equation 2.18 to define the over-sampling multiscale bases plays a crucial role. Another natural method to define the over-sampling multiscale bases is in with the coefficients so chosen that

We have carried out numerical experiments for both methods to define the over-sampling multiscale bases and found that both methods perform quite well. But a rigorous theoretical justification of the choice Equation 4.27, i.e., the derivation of an error estimate like that in Theorem 2.2, is left for future study.

5. Numerical results

In this section we first test the accuracy of the over-sampling mixed multiscale finite element method for solving Equation 1.1 with periodically oscillating coefficients. Then we consider the application of our method to the flow transport problem Equation 1.1-Equation 1.2 (the core-plug model) with a practical random log-normal permeability. In particular, using an upscaling technique recently developed in Reference 14, we show how the proposed mixed multiscale finite element method leads to a complete coarse grid algorithm.

Let be the unit square in . A uniform finite element mesh is constructed by first dividing the domain into subrectangles and then connecting the lower left and the upper right vertices of each subrectangle. In the following such a mesh is referred to the mesh. We divide the finite elements in the triangulation into two groups: the lower right and upper left elements. For each triangle , an over-sampling element is created according to whether is a lower-right element or an upper-left element as shown in Figure 1. Here we assume that the permeability field is known outside the domain when the over-sampling element is extended outside . This assumption is not a restriction for practical applications, because the permeability is generated by a geostatistical method based on the statistical characteristics, which are usually known outside the domain of interest. On the other hand, it is also possible to apply over-sampling only to the portion of the element that is inside the physical domain Reference 18.

5.1. Accuracy of the multiscale finite element method

In this subsection, we test the over-sampling mixed multiscale finite element method by solving Equation 2.1-Equation 2.2 with

where . The exact solution of the test problem is unknown, and so we compare the coarse grid solutions obtained by the over-sampling multiscale method with the resolved solution , where is computed by using the standard conforming linear finite element method on the mesh.

The test problem is computed on the mesh by the over-sampling mixed multiscale method. For each triangle , the over-sampling multiscale base functions on the over-sampling element are computed on a uniform triangular mesh which divides the horizontal and vertical edges of into subintervals, respectively. So the mesh size for solving the base functions is in each space direction.

The results of the computations are shown in Tables 1 and 2. In the tables

is the relative error between the -th component of the resolved flux and the corresponding component of the computed flux using the over-sampling multiscale method.

Recall from Theorem 2.2 that the convergence rate is . In Table 1, we test the convergence of our method when is fixed at 0.5. We can see that the solution still converges as decreases. This shows that the constant in the term in the error estimate is small. We also observe that the convergence rate is better than when the resonant error term is not dominant. On the other hand, the method still suffers from the secondary cell resonance as pointed out in Reference 13. In Table 2, we show the errors for a fixed . We observe that the errors increase as decreases, indicating the existence of the error term in the error estimate. The resonant error is the strongest when . We remark that previous studies Reference 18Reference 13 indicate that the cell resonance error is more visible for periodic coefficients, especially when is a rational number. But the cell resonance error is generically small for random coefficients.

5.2. The core-plug model: local conservation property

Now we consider the application of the over-sampling mixed multiscale finite element method to the core-plug of the flow transport problem Equation 1.1-Equation 1.2. Let be the cross-section of the reservoir through which the flow moves. Let , , and . We impose the following boundary conditions for the pressure equation Equation 1.1:

which represents a flow from to with no-flow boundary conditions on . For the saturation equation Equation 1.2, we impose the following initial condition and boundary condition on the inflow boundary :

Recall that stands for the saturation of water. Denote by the identity matrix in , and assume the relative permeability tensor . Geostatistical models often suggest that the logarithm of the permeability field is weakly or second order stationary in space, so that the mean log permeability is constant and its covariance only depends on the relative distance of two points rather than their actual location Reference 1, Reference 10, Reference 11, Reference 21 and Reference 31. This is to say that the permeability field is generically a log-normal random field. In this section, we will test the performance of our mixed multiscale finite element method for a random log-normal permeability field. Here we generate the random log-normal permeability field by using the moving ellipse average technique Reference 10 with the variance of the logarithm of the permeability , and the correlation lengths and in the and directions, respectively. One realization of the resulting permeability field in our numerical experiments is depicted in Figure 2.

Throughout we are interested in the fractional flow curves. A fractional flow curve at time is defined as the production rate on by the following formula:

Here stands for the saturation of the produced fluid (e.g., oil), and thus represents the percentage of the produced oil over water. We also use the dimensionless time PVI in our computations. One PVI is the time required to fill the whole reservoir by injecting water at . It is equal to Reference 12

where and are the lengths of the cross-section of the reservoir in the and directions ( are scaled to be in our simulations), is the real time and is the average velocity on in the direction. It is clear that in the unit of PVI, the time now is equal to the fraction of the water left in the whole reservoir, due to the heterogeneities of the medium.

The saturation equation Equation 1.2 will be solved by an upwinding scheme in the finite element formulation; see Reference 22 and Reference 8. Given the discrete velocity field , for any with unit outer normal , we divide its boundary into two parts:

Here is the midpoint of the side . For any piecewise constant function over the mesh , we define its upwinding value on as

and assume on Let be the time step and . Then, denoting by the approximation of the water saturation in at time , we use the following upwinding scheme to solve Equation 1.2:

We note that in the scheme Equation 5.1 only the normal components of the discrete velocity on the sides of each element are required. For the mixed finite element methods solving Equation 1.1, such normal components are directly computed as part of the method. In the following simulations, the “exact” fractional flow curve is referred to the one computed by solving Equation 1.1 by the lowest order Raviart-Thomas finite element method and Equation 1.2 by the scheme Equation 5.1 on the fine mesh.

Our first experiment in this subsection is designed to show the importance of the local conservation property in the flow transport simulations. Let be the approximation of the Darcy velocity ; then the local conservation property means

This is an important property, because if it is violated it would introduce unphysical numerical sources or sinks. It is clear that Equation 5.2 is always satisfied when the mixed finite element method is used to solve Equation 1.1. In this experiment we first solve the pressure equation Equation 1.1 by using a standard conforming linear finite element method on the mesh to get , and then we recover the normal components of the velocity by two different methods. For any , in the first method we take as the interior trace of . The resulting fractional flow curve is depicted in Figure 3(a). We note that in this method the normal components of the velocity are in general discontinuous across the interelement boundaries, but the local conservation property Equation 5.2 is satisfied since is constant on each element and so is . In the second method, we take on any side as the average of its values in two adjacent elements. The resulting fractional flow curve is shown in Figure 3(b). We can see clearly that the fractional flow curve obtained from the second method deviates significantly from the solid line obtained from the “exact” solution. Note that in this method the normal components are continuous across the interelement boundaries, but the local conservation property is violated. This shows that the local conservation property is more important than the continuity of the normal components. Moreover, we observe that the fractional flow curve computed by the first method is in good agreement with the “exact” one obtained by using the mixed method on the fine mesh. We remark that the mixed finite element method involves the computation of the solution of a discrete problem which has many more degrees of freedom than solving the pressure equation Equation 1.1 using a linear finite element method. Thus, when one can afford to solve the whole system Equation 1.1-Equation 1.2 on a fine mesh, instead of using the mixed finite element method, one should use the linear finite element method to solve the pressure equation Equation 1.1 and then generate the velocity field by the first method mentioned above. However, we remark that the linear finite element method is a special case. If one uses higher order displacement finite element methods to solve Equation 1.1, there are no obvious ways to recover the velocity field such that the local conservation property Equation 5.2 is satisfied.

Next we compare the performance of the mixed multiscale finite element method and the displacement multiscale finite element method Reference 18. One important feature of the multiscale finite element method is its ability to reconstruct locally the fine-scale velocity field from the coarse grid pressure solver. The key idea is to use the multiscale base functions to interpolate the fine grid solution. To demonstrate that we can accurately capture the correct small-scale velocity field, we first solve the pressure equation Equation 1.1 on a coarse mesh and use the multiscale finite element basis functions to reconstruct the velocity field on the fine mesh. We then use this locally reconstructed fine grid velocity field to solve the saturation equation Equation 1.2 on the fine mesh. In Figure 4(a), the resulting fractional flow curve is compared with the “exact” fractional flow curve which is obtained by solving the pressure and the saturation equations on the fine mesh using the mixed finite element method. As we can see from Figure 4(a), the two fractional flow curves agree very well.

We then solve the pressure equation Equation 1.1 by the displacement over-sampling multiscale finite element method on a coarse mesh and use the displacement multiscale finite element bases to reconstruct the velocity field on the fine mesh. Again the saturation equation Equation 1.2 is solved on the fine mesh. The resulting fractional flow curve and the “exact” fraction curve are shown in Figure 4(b). Due to the violation of the local conservation property, the displacement multiscale finite element method begins to deviate significantly from the “exact” fractional flow curve after some time (beyond ). This numerical example demonstrates that the over-sampling mixed multiscale finite element method provides a better way to locally reconstruct the fine grid velocity field which is suitable for long-time computations of the saturation field. This is an important consideration from a practical viewpoint.

5.3. The core-plug model: a coarse mesh algorithm

Now we describe how the proposed mixed multiscale finite element can be combined with the existing upscaling technique for the saturation equation Equation 1.2 to get a complete coarse grid algorithm for the problem Equation 1.1-Equation 1.2. The numerical upscaling of the saturation equation has been under intensive study in the literature, see Reference 11, Reference 14, Reference 12, and Reference 21. Here, we use the upscaling method proposed in Reference 14 and Reference 12 to design an overall coarse grid model for the problem Equation 1.1-Equation 1.2. The work of Reference 14 for upscaling the saturation equation involves a moment closure argument. The velocity and the saturation are separated into a local mean quantity and a small-scale perturbation with zero mean. For example, the Darcy velocity is expressed as in Equation 1.2, where is the average of the velocity over each coarse element and is the deviation of the fine-scale velocity from its coarse-scale average. After some manipulations, an average equation for the saturation can be derived as follows Reference 14:

where the diffusion coefficients are defined by

stands for the average of over each coarse element and is the length of the coarse grid streamline in the direction which starts at time at point , i.e.,

where is the solution of the following system of ODEs:

Note that the hyperbolic equation Equation 1.2 is now replaced by a convection-diffusion equation. The convection-dominant parabolic equation Equation 5.3 is solved by the characteristic linear finite element method Reference 9, Reference 26 in our simulation. The flow transport model Equation 1.1-Equation 1.2 is solved in the coarse grid as follows:

1.

Solve the pressure equation Equation 1.1 by the over-sampling mixed multiscale finite element method and obtain the fine-scale velocity field using the multiscale basis functions.

2.

Compute the coarse grid average and the fine-scale deviation on the coarse grid.

3.

At each time step, solve the convection-diffusion equation Equation 5.3 by the characteristic linear finite element method on the coarse grid in which the lengths of the streamline are computed for the center of each coarse grid element.

The mixed multiscale finite element method can be readily combined with the above upscaling model for the saturation equation. The local fine grid velocity will be constructed from the multiscale finite element base functions. The main cost in the above algorithm lies in the computation of multiscale bases, which can be done a priori and completely in parallel. This algorithm is particularly attractive when multiple simulations must be carried out due to the change of boundary and source distribution, as is often the case in engineering applications. In such situation, the cost of computing the multiscale base functions is just overhead. Moreover, once these base functions are computed, they can be used for subsequent time integration of the saturation. Because the evolution equation is now solved on a coarse grid, a larger time step can be used. This also offers additional computational saving. For many oil recovery problems, due to the excessively large fine grid data, upscaling is a necessary step before performing many simulations and realizations on the upscaled coarse grid model. If one can coarsen the fine grid by a factor of 10 in each dimension, the computational saving of the coarse grid model over the original fine model could be as large as a factor 10,000 (three space dimensions plus time).

We perform a coarse grid computation of the above algorithm on the coarse mesh. The fractional flow curve using the above algorithm is depicted in Figure 5. It gives excellent agreement with the “exact” fractional flow curve. The contour plots of the saturation at times and are depicted in Figure 6. The contour plots of the saturation on the fine mesh at times and computed by using the “exact” velocity field are displayed in Figure 7. We observe that the contour plots in Figure 6 approximate the “exact” ones in Figure 7 in accuracy, but the sharp oil/water interfaces in Figure 7 are smeared out. This is due to the parabolic nature of the upscaled equation Equation 5.3. We have also performed many other numerical experiments to test the robustness of this combined coarse grid model. We found that for permeability fields with strong layered structure, the above coarse grid model is very robust. The agreement with the fine grid calculations is very good. We are currently working towards a qualitative and quantitative understanding of this upscaling model. It is likely that one can justify this upscaling model for a periodic layered permeability field.

Finally, we remark that the upscaling equation Equation 5.3 uses small-scale information of the velocity field to define the diffusion coefficients. This information can be constructed locally through the mixed multiscale basis functions. This is an important property of our multiscale finite element method. It is clear that solving directly the homogenized pressure equation

will not provide such small-scale information. On the other hand, whenever one can afford to resolve all the small-scale features by a fine grid, one can use fast linear solvers, such as multigrid methods, to solve the pressure equation Equation 1.1 on the fine mesh. From the fine grid computation, one can easily construct the average velocity and its deviation . However, when multiple simulations must be carried out due to the change of boundary conditions, the pressure equation Equation 1.1 will then have to be solved again on the fine mesh. The multiscale finite element method only solves the pressure equation once on a coarse mesh, and the fine grid velocity can be constructed locally through the finite element bases. This is the main advantage of our mixed multiscale finite element method. This process becomes more difficult for a nonlinear two-phase flow, due to the dynamic coupling between the pressure and the saturation. We are now investigating the possibility of upscaling the two-phase flow by using multiscale finite element base functions that are constructed from the one-phase flow (time-independent). In this case, we need to provide corrections to the pressure equation to account for the scale interaction near the oil/water interface.

Acknowledgments

We wish to thank Dr. Yalchin R. Efendiev for many inspiring discussions and for providing us the code generating the log-normal permeability field. We would also like to thank Dr. Hector Ceniceros for his valuable comments on our original manuscript, and the referee for his careful reading and constructive comments.

Figures

Figure 1.

The element and its over-sampling element : lower right elements (left) and upper right elements (right). The length of the horizontal and vertical edges of is four times the corresponding length of the edges of .

Graphic without alt text
Table 1.

Results for the over-sampling method and .

Mraterate
81/1612.6155E-211.4035E-2
161/327.3081E-20.78706.5842E-20.7920
321/644.6645E-20.64784.1563E-20.6637
641/1283.6665E-20.34732.5297E-20.7163
Table 2.

Results for the over-sampling method ().

Mesh
NMraterate
128166.6152E-26.0833E-2
64324.2149E-20.65033.4530E-20.8170
32643.6664E-20.20112.5297E-20.4489
161285.1504E-2-0.49032.8311E-2-0.1624
Figure 2.

The random log-normal permeability field . The ratio of maximum to minimum is .

Graphic without alt text
Figure 3.

Comparison of two different methods recovering the normal components of the velocity from piecewise linear pressure . Each solid line is the “exact” fractional flow curve using the mixed finite element method solving the pressure equation. (a) The dotted line is the fractional flow curve obtained from the first method of local velocity recovery, which satisfies the local conservation property. (b) The dotted line is the fractional flow curve obtained from the second method of local velocity recovery, which violates the local conservation property.

Graphic without alt text
Figure 4.

Comparison of the mixed multiscale finite element method and the displacement multiscale finite element method. Each solid line is the “exact” fractional flow curve using the mixed finite element method to solve the pressure equation. (a) The dotted line is the fractional flow curve using the velocity by the mixed multiscale finite element method. (b) The dotted line is the fractional flow curve using the velocity recovered from the displacement multiscale finite element method.

Graphic without alt text
Figure 5.

The accuracy of the coarse grid algorithm. The solid line is the “exact” fractional flow curve using the mixed finite element method for solving the pressure equation. The slash-dotted line is the fractional flow curve using the above coarse grid algorithm.

Graphic without alt text
Figure 6.

The contour plots of the saturation computed using the upscaled model on a mesh at time (left) and (right)

Graphic without alt text
Figure 7.

The contour plots of the saturation computed on the fine mesh using the “exact” velocity field at time (left) and (right)

Graphic without alt text

Mathematical Fragments

Equations (1.1), (1.2)
Equations (2.1), (2.2)
Equation (2.3)
Equations (2.4), (2.5)
Equation (2.6)
Equation (2.7)
Equations (2.8), (2.9)
Equation (2.10)
Equation (2.11)
Equation (2.12)
Equation (2.13)
Equations (2.14), (2.15)
Theorem 2.1.

The discrete problem Equation 2.14-Equation 2.15 has a unique solution such that on . Moreover, if the homogenized solution and the assumptions (H1)-(H2) are satisfied, then there exists a constant , independent of and , such that

Equation (2.17)
Equation (2.18)
Equation (2.19)
Equation (2.20)
Equations (2.21), (2.22)
Theorem 2.2.

The discrete problem Equation 2.21-Equation 2.22 has a unique solution such that on . Moreover, if the homogenized solution and the assumptions (H1)-(H4) are satisfied, then there exist constants and , independent of and , such that if for all , the following error estimate holds:

Equation (3.1)
Equation (3.2)
Equation (3.3)
Equation (3.4)
Equation (3.6)
Equation (3.7)
Theorem 3.1.

Assume that . Then there exists a constant , independent of , the domain , and the data and , such that

Moreover, the boundary corrector satisfies the estimate

where stands for the length of the boundary if , and the surface area of if .

Equations (4.1), (4.2)
Equations (4.3), (4.4)
Lemma 4.2.

Assume that is uniformly coercive in , i.e., there exists a constant , independent of , such that

and satisfies the following discrete inf-sup condition for some constant independent of :

Then, the discrete problem Equation 4.3-Equation 4.4 has a unique solution such that . Moreover, the solution fulfills the following error estimate:

where the constant depends only on and the operator norms , .

Equations (4.5), (4.6)
Lemma 4.3.

There exists a linear operator such that for any and we have in and for some constant independent of .

Lemma 4.4.

The discrete problem Equation 4.5-Equation 4.6 has a unique solution such that . Moreover, the following error estimate holds:

Equation (4.8)
Equation (4.9)
Lemma 4.5.

There exists a constant , independent of and such that

Lemma 4.6.

Under the assumptions (H1)-(H4) there exists a constant , independent of and such that

Equation (4.10)
Equation (4.11)
Equation (4.12)
Equation (4.13)
Lemma 4.7.

The solution of the problem Equation 4.11 belongs to and satisfies the following stability estimates:

Moreover, the following error estimate is valid:

Lemma 4.8.

Let be the solution of Equation 4.11 and . Then there exists a constant , independent of and , such that

Equation (4.14)
Equation (4.15)
Equation (4.16)
Lemma 4.9.

Let be a periodic function with respect to the unit cube and assume that its average over vanishes. Then, for any , , we have

Equation (4.17)
Equation (4.18)
Equation (4.19)
Equation (4.20)
Lemma 4.10.

Under the assumptions (H1)-(H4) there exist constants and , independent of and , such that if and for all , the following stability estimate is valid:

Equation (4.21)
Equation (4.22)
Equation (4.23)
Equation (4.24)
Equation (4.25)
Equation (4.26)
Equation (4.27)
Equation (5.1)
Equation (5.2)
Equation (5.3)

References

Reference [1]
T. Arbogast, Numerical subgrid upscaling of two-phase flow in porous media, TICAM Report 99-30, University of Texas at Austin, 1999.
Reference [2]
M. Avellaneda and F.-H. Lin, Compactness methods in the theory of homogenization, Comm. Pure Appl. Math. 40 (1987), 803–847. MR 88i:35019
Reference [3]
I. Babuska, G. Caloz, and E. Osborn, Special finite element methods for a class of second order elliptic problems with rough coefficients, SIAM J. Numer. Anal. 31 (1994), 945–981. MR 95g:65146
Reference [4]
A. Bensoussan, J.L. Lions, and G. Papanicolaou, Asymptotic Analysis for Periodic Structures, North-Holland, Amsterdam, 1978. MR 82h:35001
Reference [5]
F. Brezzi and M. Fortin, Mixed and Hybrid Finite Element Methods, Springer, New York, 1991. MR 92d:65187
Reference [6]
F. Brezzi, L.P. Franca, T.J.R. Hughes, and A. Russo, , Comput. Methods Appl. Mech. Engrg. 145 (1997), 329–339. MR 98g:65086
Reference [7]
Z. Chen, Mixed finite element methods for a dynamical Ginzburg-Landau model in superconductivity, Numer. Math. 76 (1997), 323–353. MR 98d:65139
Reference [8]
Z. Chen and Q. Du, An upwinding mixed finite element method for a mean field model of superconducting vortices, RARIO Math. Model. Numer. Anal. 34 (2000), 687–706. MR 2001f:65133
Reference [9]
J. Douglas, Jr. and T.F. Russell, Numerical methods for convection-dominated diffusion problem based on combining the method of characteristics with finite element or finite difference procedures, SIAM J. Numer. Anal. 19 (1982), 871–885. MR 84b:65093
Reference [10]
L.J. Durlofsky, Numerical calculation of equivalent grid block permeability tensors for heterogeneous porous media, Water Resources Research 27 (1991), 699–708.
Reference [11]
L.J. Durlofsky, R.C. Jones, and W.J. Milliken, A nonuniform coarsening approach for the scale-up of displacement processes in heterogeneous porous media, Adv. Water Resources, 20 (1997), 335–347.
Reference [12]
Y.R. Efendiev, The Multiscale Finite Element Method and its Applications, Ph.D. thesis, California Institute of Technology, 1999.
Reference [13]
Y.R. Efendiev, T.Y. Hou, and X.H. Wu, The convergence of nonconforming multiscale finite element methods, SIAM J. Numer. Anal. 37 (2000), 888–910. MR 2002a:65176
Reference [14]
Y.R. Efendiev, L.J. Durlofsky, and S.H. Lee, Modeling of subgrid effects in coarse scale simulations of transport in heterogeneous porous media, Water Resources Research 36 (2000), 2031–2041.
Reference [15]
D. Gilbarg and N.S. Trudinger, Elliptic Partial Differential Equations of Second Order, Springer-Verlag, Berlin, 1983. MR 86c:35035
Reference [16]
V. Girault and P.-A. Raviart, Finite Element Methods for Navier-Stokes Equations, Springer, Berlin, 1986. MR 88b:65129
Reference [17]
P. Grisvard, Elliptic Problems on Nonsmooth Domains, Pitman, Boston, 1985. MR 86m:35044
Reference [18]
T.Y. Hou and X.H. Wu, A multiscale finite element method for elliptic problems in composite materials and porous media, J. Comput. Phys. 134 (1997), 169–189. MR 98e:73132
Reference [19]
T.Y. Hou, X.H. Wu, and Z. Cai, Convergence of a multiscale finite element method for elliptic problems with rapidly oscillating coefficients, Math. Comp. 68 (1999), 913–943. MR 99i:65126
Reference [20]
V.V. Jikov, S.M. Kozlov, and O.A. Oleinik, Homogenization of Differential Operators and Integral Functionals, Springer, Berlin, 1994. MR 96h:35003b
Reference [21]
P. Langlo and M.S. Espedal, Macrodispersion for two-phase, immiscible flow in porous media, Adv. Water Resources 17 (1994), 297–316.
Reference [22]
P. Lesaint and P.-A. Raviart, “On a Finite Element Method for Solving the Neutron Transport equation”, in Mathematical Aspects of the Finite Element Method in Partial Differential Equations, ed. by C. de Boor, Academic Press, New York, 1974. MR 58:31918
Reference [23]
G.M. Lieberman, Oblique derivative problems in Lipschitz domains II. Discontinuous boundary data, J. Reine Angew. Math. 389 (1988), 1–21. MR 89h:35094
Reference [24]
J.F. McCarthy, Comparison of fast algorithms for estimating large-scale permeabilities of heterogeneous media, Transport in Porous Media, 19 (1995), 123–137.
[25]
S. Moskow and M. Vogelius, First order corrections to the homogenized eigenvalues of a periodic composite medium. A convergence proof, Proc. Royal Soc. Edinburgh A 127 (1997), 1263–1299. MR 99g:35018
Reference [26]
O. Pironneau, On the transport-diffusion algorithm and its application to the Navier-Stokes equations, Numer. Math. 38 (1982), 309–332. MR 83d:65258
Reference [27]
P. Raviart and J. Thomas, “A mixed element method for 2nd order elliptic problems”, in Mathematical Aspects of the Finite Element Method, Lecture Notes on Mathematics 606, Springer, Berlin, 1977. MR 58:3547
Reference [28]
T.F. Russell and M.F. Wheeler, “Finite element and finite difference methods for continuous flows in porous media”, in The Mathematics of Reservoir Simulation, R.E. Ewing, ed., SIAM, Philadelphia, 1983.
Reference [29]
L. Tartar, “Nonlocal Effect Induced by Homogenization”, in PDEs and Calculus of Variations, F. Columbini, ed., Birkhäuser Publ., Boston, 1989.
Reference [30]
T.C. Wallstrom, S.L. Hou, M.A. Christie, L.J. Durlofsky and D.H. Sharp, Accurate scale up of two phase flow using renormalization and nonuniform coarsening, Computational Geosciences 3 (1999), 69–87.
Reference [31]
S. Verdiere and M.H. Vignal, Numerical and theoretical study of a dual mesh method using finite volume schemes for two-phase flow problems in porous media, Numer. Math. 80 (1998), 601–639. MR 99g:76117
[32]
P.M. De Zeeuw, Matrix-dependent prolongation and restrictions in a blackbox multigrid solver, J. Comput. Appl. Math. 33 (1990), 1–27. MR 92e:65152

Article Information

MSC 2000
Primary: 65F10 (Iterative methods for linear systems), 65F30 (Other matrix algorithms)
Author Information
Zhiming Chen
LSEC, Institute of Computational Mathematics, Chinese Academy of Sciences, Beijing 100080, Peoples Republic of China
zmchen@lsec.cc.ac.cn
Thomas Y. Hou
Applied Mathematics, California Institute of Technology, Pasadena, California 91125.
hou@acm.caltech.edu
Additional Notes

The first author was supported in part by China NSF under the grants 19771080 and 10025102 and by China MOS under the grant G1999032804.

The second author was supported in part by NSF under the grant DMS-0073916 and by ARO under the grant DAAD19-99-1-0141.

Journal Information
Mathematics of Computation, Volume 72, Issue 242, ISSN 1088-6842, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , , and published on .
Copyright Information
Copyright 2002 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/S0025-5718-02-01441-2
  • MathSciNet Review: 1954956
  • Show rawAMSref \bib{1954956}{article}{ author={Chen, Zhiming}, author={Hou, Thomas}, title={A mixed multiscale finite element method for elliptic problems with oscillating coefficients}, journal={Math. Comp.}, volume={72}, number={242}, date={2003-04}, pages={541-576}, issn={0025-5718}, review={1954956}, doi={10.1090/S0025-5718-02-01441-2}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.