A Robin-type domain decomposition method for a novel mixed-type DG method for the coupled Stokes-Darcy problem
Lina Zhao111Department of Mathematics, City University
of Hong Kong, Kowloon Tong, Hong Kong SAR, China. (linazha@cityu.edu.hk). This author was supported by a grant from City University of Hong Kong (Project No. 7200699).
Abstract
In this paper, we first propose and analyze a novel mixed-type DG method for the coupled Stokes-Darcy problem on simplicial meshes. The proposed formulation is locally conservative. A mixed-type DG method in conjunction with the stress-velocity formulation is employed for the Stokes equations, where the symmetry of stress is strongly imposed. The staggered DG method is exploited to discretize the Darcy equations. As such, the discrete formulation can be easily adapted to account for the Beavers-Joseph-Saffman interface conditions without introducing additional variables. Importantly, the continuity of normal velocity is satisfied exactly at the discrete level. A rigorous convergence analysis is performed for all the variables. Then we devise and analyze a domain decomposition method via the use of Robin-type interface boundary conditions, which allows us to solve the Stokes subproblem and the Darcy subproblem sequentially with low computational costs. The convergence of the proposed iterative method is analyzed rigorously. In particular, the proposed iterative method also works for very small viscosity coefficient. Finally, several numerical experiments are carried out to demonstrate the capabilities and accuracy of the novel mixed-type scheme, and the convergence of the domain decomposition method.
Coupling incompressible flow and porous media flow has drawn great attention over the past years, which has been involved in many practical applications, such as ground water contamination and industrial filtration. This coupled phenomenon can be mathematically expressed by the Stokes-Darcy problem, where the free fluid region is governed by the Stokes equations and the porous media region is described by Darcy’s law, and three transmission conditions are prescribed on the interface (cf. [1, 29]).
The devising of numerical schemes for the coupled Stokes-Darcy problem hinges on a suitable choice of stable pairs for both the Stokes equations and Darcy equations. As it is well known, the standard mixed formulations for the Stokes equations and Darcy equations earn different compatibility conditions, thus a straightforward application of the existing solvers for the Stokes equations and Darcy equations may not be feasible.
To this end, a great amount of effort has been devoted to developing accurate and efficient numerical schemes for the coupled Stokes-Darcy problem, and a non-exhaustive list of these approaches include Lagrange multiplier methods [21, 17, 32, 18], weak Galerkin method [9, 22], strongly conservative methods [20, 16], stabilized mixed finite element method [28, 24], discontinuous Galerkin (DG) methods [26, 34], virtual element method [23, 33], a lowest-order staggered DG method [37] and penalty methods [38]. The coupled Stokes-Darcy problem describes multiphysics phenomena, and involves a Stokes subproblem and a Darcy subproblem, it is thus natural to resort to domain decomposition methods, which allows one to solve the coupled system sequentially with a low computational cost. Various domain decomposition methods have been developed for the coupled Stokes-Darcy problem, see, for example [13, 6, 8, 7, 31, 19], most of which are based on velocity-pressure formulation of the Stokes equations.
The first purpose of this paper is to devise and analyze a novel mixed-type method for the coupled Stokes-Darcy problem. In the proposed formulation, we use a mixed-type DG method in conjunction with stress-velocity formulation for the discretization of the Stokes subproblem, and we use staggered DG method introduced in [10] for the discretization of the Darcy subproblem. Unlike the schemes proposed in [27, 34], we enforce the normal continuity of velocity directly into the formulation of the method without resorting to Lagrange multiplier. This is based on our observation that the degrees of freedom for the Darcy velocity space is defined by using dual edge degrees of freedom and interior degrees of freedom, and the interface can be treated as the union of the primal edges. Therefore, the normal continuity of velocity can be simply imposed into the discrete formulation by replacing the Darcy’s normal velocity by the Stokes’ normal velocity, which in conjunction with a suitable decomposition of the stress variable on the interface yields the resulting formulation. It is worth mentioning that the normal continuity of velocity is satisfied exactly at the discrete level.
The advantages of the proposed formulation are multifold: First, the symmetry of stress is strongly imposed. The stress variable has a physical meaning and its computation is also important from a practical point of view. Second, pressure is eliminated from the equation, which reduces the size of the global system. Third, the continuity of normal velocity can be imposed directly without using Lagrange multiplier. A rigorous convergence analysis is carried out for all the variables. The proof for the optimal convergence of -error of Darcy velocity is non-trivial, to overcome this issue, we exploit a non-standard trace theorem. We remark that the proposed scheme utilizes stress-velocity formulation for the Stokes equations, which is rarely explored for the coupled Stokes-Darcy problem in the existing literature. Our work will shed new insights into the devising and analysis of novel numerical schemes for the coupled Stokes-Darcy problem. We also address that the numerical results demonstrate that the proposed scheme also works well for small values of viscosity.
The proposed global formulation involves four variables: stress, fluid velocity, Darcy velocity and Darcy pressure, which may require high computational costs especially for large scale problems. To reduce the computational costs, we aim to devise a domain decomposition method based on the proposed spatial discretization, where the global formulation is decomposed into the Stokes subproblem and the Darcy subproblem by using newly constructed Robin-type interface conditions. The construction of the novel Robin-type interface condition is not a simple extension of the method introduced in [30], instead it takes advantage of the special features of staggered DG method. Moreover, the compatibility conditions are derived to ensure the equivalence of the modified discrete formulation and the original discrete formulation. The convergence of the Robin-type domain decomposition method is rigorously analyzed with the help of the compatibility conditions. Our convergence analysis indicates that the convergence of the Robin-type domain decomposition method is for , where and are parameters introduced in Section 5. Moreover, when and satisfy certain conditions, then the convergence rate of the Robin-type domain decomposition method is -independent, which is particularly inspiring. Several numerical experiments are carried out to demonstrate the performance of the proposed scheme. We can observe that the proposed domain decomposition method converges as reflected by the theories.
In particular, it also converges for small values of viscosity.
The rest of the paper is organized as follows. In the next section, we describe the model problem and derive the stress-velocity formulation for the Stokes equations. In Section 3, we derive the discrete formulation and prove the unique solvability. The convergence error estimates for all the variables measured in -error are given in Section 4. Then, the Robin-type domain decomposition method is constructed and analyzed in Section 5. Several numerical experiments are carried out in Section 6 to verify the proposed theories. Finally, a concluding remark is given in Section 7.
2 Model problem
Let denote the polygonal domain in , where and represent the fluid domain and porous media domain, respectively. Let denote the interface between and , and let for , see Figure 1 for an illustration of the computational domain. We use to represent the unit normal vector to . Let be an orthonormal system of tangential vectors on .
Figure 1: Profile of the computational domain.
In , the fluid flow is governed by the Stokes equations
(2.1)
(2.2)
(2.3)
(2.4)
where is the viscosity coefficient which is assumed to be a positive constant, is the fluid velocity, is the fluid pressure, is the stress tensor, is the identity matrix and is the deformation tensor defined by . Hereafter, we use to represent the transpose of .
In , the governing equations are Darcy equations
(2.5)
(2.6)
(2.7)
On the interface , we prescribe the following Beavers-Joseph-Saffman conditions (cf. [1, 29])
(2.8)
(2.9)
(2.10)
where is the phenomenological friction coefficient.
Now we will derive a stress-velocity formulation based on (2.1)-(2.3) by eliminating . First, we have
thus, we obtain
Consequently, we can recast (2.2) into the following equivalent formulation
where
Note that is a trace-free tensor called
deviatoric part and .
Then we can rewrite (2.1)-(2.3) as the following equivalent system:
(2.11)
(2.12)
(2.13)
We introduce some notations that will be used throughout the paper. Let . By , we denote the scalar product in . We use the same notation for the scalar product in and in . More precisely, for and for . The associated norm is denoted by .
Given an integer , we denote the scalar-valued Sobolev spaces by with the norm and seminorm . In addition, we use and to denote the vector-valued and tensor-valued Sobolev spaces, respectively. In the following, we use to denote a generic constant independent of the meshsize which may have different values at different occurrences.
3 The new scheme
In this section, we are devoted to the derivation of a novel mixed-type DG method for the coupled Stokes-Darcy system (2.5)-(2.10) and (2.11)-(2.12). To this end, we first introduce the following meshes.
Following [10, 36, 35], we first let be the
initial partition of the domain into non-overlapping triangular meshes as shown in the left panel of Figure 2. We use to denote the set of edges lying on the interface .
We also let be the set of all edges excluding the interface edges in the initial partition and be the
subset of all interior edges of .
For each triangular mesh in the initial partition , we construct the sub-triangulation by connecting an interior point to
all the vertices of . For simplicity, we select as the center point.
We rename the union of these triangles sharing the common point by . Moreover, we will use to denote the set of all the new edges generated by this subdivision process and
use to denote the resulting quasi-uniform triangulation,
on which our basis functions are defined. Here the sub-triangulation satisfies standard mesh regularity assumptions (cf. [12]).
This construction is illustrated in Figure 2,
where the black solid lines are edges in
and the red dotted lines are edges in . For each interior edge , we use to denote the union of the two triangles in sharing the edge ,
and for each boundary edge , we use to denote the triangle in having the edge ,
see Figure 2.
On the other hand, we let be a family of shape-regular triangulations of . For simplicity, we assume that the meshes between the two domains and match at the interface. We use to denote the set of all edges of excluding the edges lying on the interface. We use to represent the subset of , i.e., is the union of interior edges of . In addition, we let and . In what follows, stands for the length of edge . For each triangle
, we let be the diameter of
and
. For each interior edge , we then fix as one of the two possible unit normal vectors on .
When there is no ambiguity,
we use instead of to simplify the notation. In addition, we use to represent the corresponding unit tangent vector.
To simplify the presentation, we only consider triangular meshes in this paper, and the extension to polygonal meshes will be investigated in the future paper.
For , and , we define and as the spaces of polynomials of degree up to order on and , respectively. For a scalar or vector function belonging to the broken Sobolev space, its jump and average on an interior edge are defined as
where and , are the
two triangles in having the edge . For the boundary edges, we simply define and . We can omit the subscript when it is clear which edge we are referring to.
Figure 2: The schematics of the meshes for . Primal meshes (left), dual meshes and simplicial meshes (right). The solid lines represent the primal edges and the dashed lines represent the dual edges. represents the primal element and represents the dual element.
We define the following spaces for the numerical approximations.
and
In the following, we use to represent the element-wise defined
deformation tensor, i.e., for . Hereafter, we use and to represent the gradient and divergence operators defined element by element, respectively.
For later use, we specify the degrees of freedom for as follows.
(SD1)
For , we have
(3.1)
(SD2)
For each , we define
(3.2)
Now we are ready to derive the discrete formulation based on the first order system (2.5)-(2.6) and (2.11)-(2.12). Multiplying (2.1) by a test function and applying integration by parts yield
where the second term can be recast into the following form by using (2.9) and (2.10)
Multiplying (2.6) by a test function and performing integration by parts lead to
Finally we multiply (2.5) by , then using the interface condition (2.8) implies that
Based on the above derivations, we are now in position to define the following bilinear forms, which is instrumental for later use.
and
It follows from integration by parts that
The discrete formulation reads as follows: Find and such that
(3.3)
(3.4)
(3.5)
where is a constant over each edge .
Integration by parts implies the following discrete adjoint property
(3.6)
To facilitate later analysis, we define the following mesh-dependent norm/semi-norm
for any .
Following [10, 11], we have the following inf-sup condition
(3.7)
For later use, we also introduce the following space
where and is the Raviart-Thomas element of index introduced in [25].
Remark 3.1.
The scheme is locally mass conservative. Indeed, if one choses the test function in (3.4) such that on belonging to excluding the elements having the edge lying on and zeros otherwise, we have
In addition, if we take the test function in (3.5) such that on for all and zeros otherwise, we have
Theorem 3.1.
(unique solvability).
There exists a unique solution to (3.3)-(3.4).
Proof.
(3.3)-(3.4) is a square linear system, uniqueness implies existences, thus, it suffices to show the uniqueness.
To prove the uniqueness, we set and . Then taking and in (3.3)-(3.4) and summing up the resulting equations yield
which implies that for any , and .
Since for any , we have from the definition of that
Taking implies that . Then the discrete Korn’s inequality (cf. [3]) gives .
On the other hand we have , which indicates that we can express as
It follows from the inf-sup condition (cf. [2]) that
which implies , thereby .
Finally, we have from the inf-sup condition (3.7), the discrete adjoint property (3.6), the discrete Poincaré inequality (cf. [3]) and (3.4) that
which implies that .
Therefore, the proof is completed.
∎
4 A priori error estimates
In this section we will present the convergence error estimates for all the variables. To begin, we define the following interpolation operators, which play an important role for the subsequent analysis.
We let denote the BDM projection operator (cf. [4]),
which satisfies the following error estimates for (see, e.g., [4, 14])
Let be defined by
and , be defined by
It is easy to see that and are well defined polynomial preserving operators. In addition, the following approximation properties hold for and (cf. [12, 10])
(4.1)
(4.2)
Finally, we let denote the orthogonal projection onto , then it holds
(4.3)
Performing integration by parts on the discrete formulation (3.3)-(3.5) and using the interface conditions (2.8)-(2.10), we can get the following error equations:
(4.4)
(4.5)
(4.6)
which indicates that our discrete formulation is consistent.
Theorem 4.1.
Let and . In addition, let and be the discrete solution of (3.3)-(3.5). Then the following convergence error estimate holds
Proof.
Taking , ,
and in (4.4)-(4.6) and summing up the resulting equations yield
where we use the definitions of and , i.e.,
We can bound the first two terms on the right hand side by the Cauchy-Schwarz inequality
Using the definition of , we can infer that
where we use the fact that for any .
Thereby, we can estimate by
(4.7)
An application of the Cauchy-Schwarz inequality and the definition of leads to
The Cauchy-Schwarz inequality yields
Combining the preceding estimates, Young’s inequality, the interpolation error estimates and the trace inequality implies that
Therefore, the proof is completed by using the interpolation error estimates (4.1)-(4.3).
∎
We can observe from Theorem 4.1 that the convergence rate for -error of is not optimal in terms of the polynomial order. The proof for the optimal convergence rate of -error of is non-trivial and it relies on a non-standard trace theorem, which will be explained later. In order to achieve the optimal convergence order for -error of , we first need to show the error estimates of and . To this end, we assume that the following dual problem holds (cf. [16]):
(4.8)
(4.9)
(4.10)
and
(4.11)
(4.12)
(4.13)
The following interface conditions are prescribed on the interface
Assume that the following regularity estimate holds
(4.14)
Then we can state the convergence error estimate for errors of and .
Theorem 4.2.
Let and . In addition, let and be the discrete solution of (3.3)-(3.5). Then the following convergence error estimate holds
Proof.
Multiplying (4.8) by , (4.9)
by , (4.11) by and (4.12) by and performing integration by parts lead to
The first two terms on the right hand side can be bounded by the Cauchy-Schwarz inequality
We can rewrite the third term as follows
(4.15)
where the first term on the right hand side can be bounded by
The second term on the right hand side of (4.15) can be estimated similarly to (4.7). Indeed, we have
On the other hand, we have from the definitions of , and
The Cauchy-Schwarz inequality implies
The last term can be bounded by the trace inequality and (4.14)
Then combining the preceding estimates with Theorem 4.1 and the interpolation error estimates (4.1)-(4.3) implies the desired estimate.
∎
Proceeding similarly to Lemma 2.3 of [5], we can get the following lemma.
Lemma 4.1.
Let be an edge of and let . Let for and .
Assume that is a given function in . Then, there exists a small and a positive constant independent of such that
The trace inequality introduced in Lemma 4.1 is crucial to obtain the optimal convergence error estimate for -error of . Now, we can state the following theorem.
Theorem 4.3.
Let and . In addition, let and be the discrete solution of (3.3)-(3.5). Then, we have
where we make use of the definitions of the interpolation operators and .
For any and , we have from the inverse inequality and Lemma 4.1 that
where and .
For sufficiently close to , it holds , thereby it follows that
(4.18)
On the other hand, we have from the trace inequality (cf. [15]) and the discrete Poincaré inequality (cf. [3]) that
In addition, it follows from the inf-sup condition (3.7) and (4.5) that
Thereby, summing (4.18) over all the edges in and using the above arguments lead to
which coupled with the interpolation error estimates (4.1)-(4.3) and Theorem 4.2 completes the proof.
∎
Theorem 4.4.
Let be the numerical solution of (3.3)-(3.5), then the interface condition (2.8) is satisfied exactly at the discrete level, i.e.,
Proof.
We can infer from (4.17) and the discrete adjoint property that
thereby, we have
Then we can define using the degrees of freedom defined in (3.1)-(3.2) such that it satisfies
Therefore, we have
We can infer that
Therefore, the proof is completed.
∎
5 Robin type domain decomposition method
In this section, we introduce a novel domain decomposition method based on the Robin-type interface conditions, which can decouple the global system (3.3)-(3.5) into the Stokes subproblem and the Darcy subproblem. To begin with, we define two functions and on the interface such that
(5.1)
(5.2)
where and are two positive constants which will be specified later.
In addition, the following compatibility conditions should be satisfied in order to get the equivalence between the interface conditions (2.8)-(2.9) and (5.1)-(5.2)
Taking in the above equations and summing up the resulting equations, we can obtain
(5.25)
(5.26)
The following identities are satisfied on the interface
Therefore, we can infer from (5.25) and (5.26) that
and
which implies the desired estimates.
∎
Theorem 5.1.
If , then the solution of Algorithm DDM converges to the solution of the Stokes-Darcy system (3.3)-(3.5). Moreover, the following estimate holds
If and , then the solution of Algorithm DDM converges to the solution of the Stokes-Darcy system (3.3)-(3.5). In addition, the following estimate holds
Proof.
Let denote the set of nodes corresponding to the degrees of freedom specified in (SD1)-(SD2) in and let . We define an extension operator , which satisfies
Note that and an application of scaling arguments implies that
(5.27)
Then setting and in (5.21)-(5.22) and summing up the resulting equations, we can get
Therefore, proceeding similarly to (5.36), we can obtain
Therefore, the proof is completed.
∎
6 Numerical experiments
In this section, several numerical experiments will be presented to show the performance of the proposed scheme. First, we test the convergence of the proposed scheme (cf. (3.3)-(3.5)) for different polynomial orders. In addition, the robustness of the scheme with respect to different values of is demonstrated. Then, we show the convergence of the Algorithm DDM with respect to different values of and . Note that we use the uniform criss-cross meshes in the following examples and similar performance can be observed for other types of triangular meshes. In the following examples, we set . The stopping criterion for the iteration of the
algorithm is selected as a fixed tolerance of for the difference between two successive iterative velocities in
-norm, i.e.,
6.1 Example 1
In this example, we take and . The exact solution is given by
which does not satisfy the interface conditions, in this respect, the discrete formulation shall be adapted to account for this situation. The convergence history for various values of for polynomial orders are displayed in Tables 1-2. We can observe that optimal convergence rates for all the variables measured in -error can be obtained. In addition, the accuracy for -error of fluid velocity is slightly influenced by the values of , which demonstrates the robustness of our method with respect to .
Mesh
Error
Order
Error
Order
Error
Order
Error
Order
1
2
2.42e-02
N/A
5.61e-01
N/A
1.69e-02
N/A
6.90e-03
N/A
4
6.70e-03
1.85
3.02e-01
0.89
4.10e-03
2.05
1.70e-03
2.01
8
1.70e-03
1.95
1.56e-01
0.95
9.45e-04
2.03
4.29e-04
2.00
16
4.44e-04
1.97
7.95e-02
0.98
2.46e-04
2.02
1.07e-04
2.00
32
1.12e-04
1.99
4.01e-02
0.99
6.09e-05
2.01
2.68e-05
2.00
2
2
2.80e-03
N/A
8.71e-02
N/A
7.86e-04
N/A
3.00e-04
N/A
4
3.91e-04
2.83
2.35e-02
1.89
8.64e-05
3.18
3.69e-05
3.02
8
5.25e-05
2.90
6.20e-03
1.92
1.02e-05
3.07
4.58e-06
3.00
16
6.79e-06
2.95
1.60e-03
1.96
1.27e-06
3.02
5.73e-07
3.00
32
8.64e-07
2.98
4.04e-04
1.98
1.58e-07
3.00
7.16e-08
3.00
3
2
2.19e-04
N/A
7.60e-03
N/A
4.96e-05
N/A
7.61e-06
N/A
4
1.42e-05
3.95
1.00e-03
2.91
2.33e-06
4.41
4.18e-07
4.18
8
9.00e-07
3.98
1.29e-04
2.96
1.31e-07
4.15
2.58e-08
4.01
16
5.66e-08
3.99
1.64e-05
2.98
7.80e-09
4.06
1.61e-09
4.00
32
3.55e-09
3.99
2.06e-06
2.99
4.77e-010
4.03
1.00e-10
3.99
Table 1: Convergence history with for Example 6.1.
Mesh
Error
Order
Error
Order
Error
Order
Error
Order
1
2
4.99e-02
N/A
1.33e-01
N/A
2.34e-02
N/A
7.00e-03
N/A
4
1.57e-02
1.67
5.72e-02
1.21
3.90e-03
2.58
1.70e-03
2.05
8
3.60e-03
2.11
2.83e-02
1.02
9.49e-04
2.04
4.25e-04
1.99
16
8.79e-04
2.04
1.42e-02
1.00
2.37e-04
2.00
1.06e-04
2.00
32
2.17e-04
2.02
7.10e-03
1.00
5.93e-05
2.00
2.65e-05
2.00
2
2
6.66e-03
N/A
2.35e-02
N/A
1.30e-03
N/A
3.06e-04
N/A
4
8.23e-04
2.99
6.00e-03
1.97
1.22e-04
3.36
3.70e-05
3.04
8
1.15e-04
2.84
1.50e-03
1.99
1.29e-05
3.24
4.57e-06
3.01
16
1.53e-05
2.91
3.75e-04
1.99
1.63e-06
2.98
5.72e-07
3.00
32
1.97e-06
2.96
9.38e-05
2.00
1.95e-07
3.06
7.15e-08
2.99
3
2
7.27e-04
N/A
1.80e-03
N/A
7.91e-05
N/A
7.46e-06
N/A
4
3.37e-05
4.43
2.23e-04
2.99
2.77e-06
4.83
4.03e-07
4.21
8
1.91e-06
4.14
2.80e-05
2.99
1.62e-07
4.09
2.55e-08
3.98
16
1.13e-07
4.07
5.49e-06
3.00
1.06e-08
3.93
1.61e-09
3.99
32
6.87e-09
4.04
4.37e-07
3.00
6.25e-010
4.08
1.00e-10
3.99
Table 2: Convergence history with for Example 6.1.
6.2 Example 2
In this example, we set and . The exact solution is defined by
Here we set and . In addition, we set the polynomial order . We aim to test the convergence of Algorithm DDM for different values of and .
First, we let and . We can observe from Table 3 that the convergence rate of the algorithm is -dependent for the case of ; indeed, more iterations are needed for smaller meshsize. This is consistent with our theoretical results as the converge rate of Algorithm DDM is proved to depend on . In addition, the solution converges faster when is smaller. Next, we set . We let and . We can see from Table 4 that less iterations are required compared to the case of . In addition, the convergence rate is -independent, which is consistent with our analysis given in Theorem 5.1. In both cases, we can achieve the optimal convergence rates for errors of all the variables. Moreover, we also display the velocity errors for both the Stokes and Darcy regions with respect to different choices of and in Figure 3. It shows that Algorithm DDM converges for and it tends to converge faster for smaller ratio of , which is consistent with our theory.
Mesh
Iteration
N
Error
Order
Error
Order
Error
Order
Error
Order
0.5
4
144
3.50e-03
N/A
1.29e-01
N/A
2.07e-02
N/A
4.20e-03
N/A
8
264
8.78e-04
1.99
6.94e-02
0.90
5.10e-03
2.01
1.10e-03
1.99
16
312
2.17e-04
2.01
3.62e-02
0.94
1.30e-03
2.00
2.65e-04
1.99
0.25
4
52
3.50e-03
N/A
1.29e-01
N/A
2.07e-02
N/A
4.20e-03
N/A
8
98
8.78e-04
1.99
6.94e-02
0.90
5.10e-03
2.01
1.11e-03
1.99
16
180
2.17e-04
2.01
3.62e-02
0.94
1.30e-03
2.00
2.65e-04
1.99
0.1
4
22
3.50e-03
N/A
1.29e-01
N/A
2.07e-02
N/A
4.20e-03
N/A
8
42
8.78e-04
1.99
6.94e-02
0.90
5.10e-03
2.01
1.11e-03
1.99
16
76
2.17e-04
2.01
3.62e-02
0.94
1.30e-04
2.00
2.65e-04
1.99
Table 3: The convergence of Algorithm DDM for for Example 6.2.
Iteration
1
1/2
4
28
3.50e-03
1.29e-01
2.07e-02
4.20e-03
8
30
8.78e-04
6.94e-02
5.10e-03
1.10e-03
16
30
2.17e-04
3.62e-02
1.30e-03
2.65e-04
32
30
5.40e-05
1.85e-02
3.21e-04
6.63e-05
1
1/4
4
16
3.50e-03
1.29e-01
2.07e-02
4.20e-03
8
16
8.78e-04
6.94e-02
5.10e-03
1.10e-03
16
16
2.17e-04
3.62e-02
1.30e-03
2.65e-04
32
16
5.40e-05
1.85e-02
3.21e-04
6.63e-05
Table 4: The convergence of Algorithm DDM for for Example 6.2.
Figure 3: The numerical velocity errors for the Stokes (left) and Darcy (right) regions with meshsize for Example 6.2.
6.3 Example 3
In this example, we set and .
We use the exact solution defined by
Here, we take and the interface conditions are satisfied exactly. In this example we attempt to test the convergence of Algorithm DDM with respect to different values of . Figure 4 shows the errors of fluid velocity and Darcy velocity with respect to various values of under the setting . One can clearly observe that Algorithm DDM converges for all the cases, and converges slower for smaller values of . Next, we fix and take various combinations of and . We can observe from Figure 5 that when the ratio gets smaller, Algorithm DDM tends to converge faster, which is consistent with our theory.
Figure 4: The numerical velocity errors for the Stokes (left) and Darcy (right) regions with meshsize for Example 6.3.
Figure 5: The numerical velocity errors for the Stokes (left) and Darcy (right) with meshsize for Example 6.3.
6.4 Example 4
Finally, we use the modified driven cavity flow to test the performance of Algorithm DDM. To this end, we set and . The exact solution is unknown.
The boundary conditions for are defined as
Homogeneous Dirichlet boundary condition is imposed for on .
The source term is taken to be and .
We first display the approximated solution for with in Figure 6. The converge of Algorithm DDM for the Stokes and Darcy regions is shown in Figure 7, where different values of are used. As expected, the algorithm converges for and it tends to converge faster for smaller values of .
Figure 6: Profiles of numerical solution for Example 6.4.
Figure 7: The numerical velocity errors for the Stokes (left) and Darcy (right) regions for Example 6.4.
7 Conclusion
Our contributions for this paper are twofold. First, we devise and analyze a new method for the coupled Stokes-Darcy problem. This method is based on a stress-velocity formulation for the Stokes equations, which is rarely explored for the coupled Stokes-Darcy problem. The stress-velocity formulation is very important and addresses variables of physical interest. The interface conditions are imposed straightforwardly without resorting to additional variables. In addition, the normal continuity of velocity is satisfied exactly at the discrete level. The convergence error estimates for all the variables are provided. Next, we design a domain decomposition method to decouple the global system into the Stokes subproblem and the Darcy subproblem via a suitable application of Robin-type interface conditions. Moreover, the convergence of the proposed iterative method is analyzed.
References
[1]
G. S. Beavers and D. D. Joseph.
Boundary conditions at a naturally permeable wall.
J. Fluid Mech., 30: 197–207, 1967.
[2]
D. Boffi, F. Brezzi, and M. Fortin.
Mixed Finite Element Methods and Applications, volume 1.
Springer, 2013.
[3]
S. C. Brenner.
Poincaré–Friedrichs inequalities for piecewise
functions.
SIAM J. Nume. Anal., 41:306–324, 2003.
[4]
F. Brezzi, J. Douglas, and L. D. Marini.
Two families of mixed finite elements for second order elliptic
problems.
Numer. Math., 47:217–235, 1985.
[5]
Z. Cai, C. He, and S. Zhang.
Discontinuous finite element methods for interface problems: Robust a
priori and a posteriori error estimates.
SIAM J. Numer. Anal., 55:400–418, 2017.
[6]
Y. Cao, M. Gunzburger, X. He, and X. Wang.
Robin-Robin domain decomposition methods for the steady-state
Stokes-Darcy system with the Beavers-Joseph interface condition.
Numer. Math., 117:601–629, 2011.
[7]
Y. Cao, M. Gunzburger, X. He, and X. Wang.
Parallel, non-iterative, multi-physics domain decomposition methods
for time-dependent Stokes-Darcy systems.
Math. Comp., 83:1617–1644, 2014.
[8]
W. Chen, M. Gunzburger, F. Hua, and X. Wang.
A parallel Robin-Robin domain decomposition method for the
Stokes-Darcy system.
SIAM J. Numer. Anal., 49:1064–1084, 2011.
[9]
W. Chen, F. Wang, and Y. Wang.
Weak Galerkin method for the coupled Darcy-Stokes flow.
IMA J. Numer. Anal., 36:897–921, 2015.
[10]
E. T. Chung and B. Engquist.
Optimal discontinuous Galerkin methods for the acoustic wave
equation in higher dimensions.
SIAM J. Numer. Anal., 47:3820–3848, 2009.
[11]
E. T. Chung, H. H. Kim, and O. B. Widlund.
Two-level overlapping Schwarz algorithms for a staggered
discontinuous Galerkin method.
SIAM J. Numer. Anal., 51:47–67, 2013.
[12]
P. G. Ciarlet.
The Finite Element Method for Elliptic Problems.
North-Holland Publishing, Amsterdam, 1978.
[13]
M. Discacciati, A. Quarteroni, and A. Valli.
Robin-Robin domain decomposition methods for the Stokes-Darcy
coupling.
SIAM J. Numer. Anal., 45:1246–1268, 2007.
[14]
R. G. Durán.
Error analysis in , for mixed
finite element methods for linear and quasi-linear elliptic problems.
ESAIM: Math. Model. Numer. Anal, 22:371–387, 1988.
[15]
X. Feng and O. A. Karakashian.
Two-level additive Schwarz methods for a discontinuous Galerkin
approximation of second order elliptic problems.
SIAM J. Numer. Anal., 39:1343–1365, 2001.
[16]
G. Fu and C. Lehrenfeld.
A strongly conservative hybrid DG/mixed FEM for the coupling of
Stokes and Darcy flow.
J. Sci. Comput., 77:1605–1620, 2018.
[17]
G. N. Gatica, S. Meddahi, and R. Oyarzúa.
A conforming mixed finite-element method for the coupling of fluid
flow with porous media flow.
IMA J. Numer. Anal., 29(1):86–108, 2008.
[18]
G. N. Gatica, R. Oyarzúa, and F.-J. Sayas.
Analysis of fully-mixed finite element methods for the
Stokes-Darcy coupled problem.
Math. Comp., 80:1911–1948, 2011.
[19]
X. He, J. Li, Y. Lin, and J. Ming.
A domain decomposition method for the steady-state
Navier–Stokes–Darcy model with Beavers–Joseph interface
condition.
SIAM J. Sci. Comput., 37:S264–S290, 2015.
[20]
G. Kanschat and B. Riviére.
A strongly conservative finite element method for the coupling of
Stokes and Darcy flow.
J. Comput. Phys., 229:5933–5943, 2010.
[21]
W. J. Layton, F. Schieweck, and I. Yotov.
Coupling fluid flow with porous media flow.
SIAM J. Numer. Anal., 40:2195–2218, 2002.
[22]
R. Li, Y. Gao, J. Li, and Z. Chen.
A weak Galerkin finite element method for a coupled
Stokes-Darcy problem on general meshes.
J. Comput. Appl. Math., 334:111–127, 2018.
[23]
X. Liu, R. Li, and Z. Chen.
A virtual element method for the coupled Stokes-Darcy problem
with the Beaver-Joseph-Saffman interface condition.
Calcolo, 56, 2019.
[24]
M. A. A. Mahbub, X. He, N. J. Nasu, C. Qiu, and H. Zheng.
Coupled and decoupled stabilized mixed finite element methods for
nonstationary dual-porosity-Stokes fluid flow model.
Int. J. Numer. Methods Eng., 120:803–833, 2019.
[25]
P. A. Raviart and J. M. Thomas.
A mixed finite element method for second order elliptic
problems, Mathematical Aspects of Finite Element Method, (I. Galligani & E.
Magenes eds).
Lectures Notes in Math. 606. New York: Springer, pp. 292–315., 1977.
[26]
B. Riviére.
Analysis of a discontinuous finite element method for the coupled
Stokes and Darcy problems.
J. Sci. Comput., 22:479–1500, 2005.
[27]
B. Riviére and I. Yotov.
Locally conservative coupling of Stokes and Darcy flows.
SIAM J. Numer. Anal., 42:1959–1977, 2005.
[28]
H. Rui and J. Zhang.
A stabilized mixed finite element method for coupled Stokes and
Darcy flows with transport.
Comput. Methods Appl. Mech. Engrg., 315:169–189, 2017.
[29]
P. G. Saffman.
On the boundary condition at the surface of a porous medium.
Stud. Appl. Math., 50:93–101, 1971.
[30]
Y. Sun, W. Sun, and H. Zheng.
Domain decomposition method for the fully-mixed Stokes-Darcy
coupled problem.
Comput. Methods Appl. Mech. Engrg., 374:113578, 2021.
[31]
D. Vassilev, C. Wang, and I. Yotov.
Domain decomposition for coupled Stokes and Darcy flows.
Comput. Methods Appl. Mech. Engrg., 268:264–283, 2014.
[32]
D. Vassilev and I. Yotov.
Coupling Stokes-Darcy flow with transport.
SIAM J. Sci. Comput., 31:3661–3684, 2009.
[33]
G. Wang, F. Wang, L. Chen, and Y. He.
A divergence free weak virtual element method for the
Stokes-Darcy problem on general meshes.
Comput. Methods Appl. Mech. Engrg., 344:998–1020, 2019.
[34]
J. Wen, J. Su, Y. He, and H. Chen.
A discontinuous Galerkin method for the coupled Stokes and
Darcy problem.
J. Sci. Comput., 85, 2020.
[35]
L. Zhao, E. T. Chung, E.-J. Park, and G. Zhou.
Staggered DG method for coupling of the Stokes and
Darcy–Forchheimer problems.
SIAM J. Numer. Anal., 59:1–31, 2021.
[36]
L. Zhao and E.-J. Park.
A staggered discontinuous Galerkin method of minimal dimension on
quadrilateral and polygonal meshes.
SIAM J. Sci. Comput., 40(4):A2543–A2567, 2018.
[37]
L. Zhao and E.-J. Park.
A lowest-order staggered DG method for the coupled Stokes-Darcy
problem.
IMA J. Numer. Anal., 40(4):2871–2897, 2020.
[38]
G. Zhou, T. Kashiwabara, I. Oikawa, E. T. Chung, and M.-C. Shiue.
An analysis on the penalty and Nitsche’s methods for the
Stokes-Darcy system with a curved interface.
Appl. Numer. Math., 165:83–118, 2021.