An arbitrary order mixed finite element method with boundary value correction for the Darcy flow on curved domains
Abstract
We propose a boundary value correction method for the Brezzi-Douglas-Marini mixed finite element discretization of the Darcy flow with non-homogeneous Neumann boundary condition on 2D curved domains. The discretization is defined on a body-fitted triangular mesh, i.e. the boundary nodes of the mesh lie on the curved physical boundary. However, the boundary edges of the triangular mesh, which are straight, may not coincide with the curved physical boundary. A boundary value correction technique is then designed to transform the Neumann boundary condition from the physical boundary to the boundary of the triangular mesh. One advantage of the boundary value correction method is that it avoids using curved mesh elements and thus reduces the complexity of implementation. We prove that the proposed method reaches optimal convergence for arbitrary order discretizations. Supporting numerical results are presented.
Key words: mixed finite element method, Neumann boundary condition, curved domain, boundary value correction method
1 Introduction
Many practical problems arising in science and engineering are posed on the curved domain . Classical finite element methods defined on a polygonal approximation domain often suffer from an additional geometric error due to the difference between and . This leads to a loss of accuracy for higher-order discretizations [29, 30]. Moreover, when the problem is equipped with an essential boundary condition, effective ways to transfer the boundary condition from to must be designed. Different strategies have been proposed to solve this problem, for example, the isoparametric finite element method [20, 22] and the isogeometric analysis [21, 17]. Another strategy is the boundary value correction method [8], which shifts the essential boundary condition from to , and results in a modified variational formulation. One advantage of the boundary value correction method is that it avoids using curved mesh elements and thus reduces the complexity of implementation. Recently, there are many works utilizing the boundary correction strategy, including the discontinuous Galerkin method via extensions from subdomains [15, 16], transferring technique based on the path integral [26], the shifted boundary method [24, 1], the cutFEM [12], the boundary-corrected virtual element method [2] and the boundary-corrected weak Galerkin method [23], etc.
For the mixed finite element method (MFEM), the Neumann boundary condition becomes essential. Again, higher-order MFEM suffers from a loss of accuracy on curved domains. In [3, 4, 5], the authors studied a parametric Raviart-Thomas MFEM on curved domains, which is a generalization of the isoparametric method to the MFEM. In [27], a cutFEM method is proposed in the curved domain, which is based on a primal mixed formulation [7]. The authors of [27] show that the cutFEM method reaches suboptimal convergence.
In this paper, we propose a new boundary corrected MFEM based on the primal mixed formulation [7], and prove that it has optimal convergence rate. Similarly to [18], we weakly impose the Neumann boundary condition. Then, a boundary value correction technique is designed to pull the Neumann boundary data from to . However, unlike the boundary correction in [8, 1, 24], which considers the Dirichlet boundary condition, the Neumann boundary condition involves the outward normal vector on the boundary, which is different in and . This poses extra difficulty in the design and analysis of our scheme. Finally, following [14], we added a term to the discrete formulation to increase stability.
The paper is organized as follows. In Section 2, we introduce some notation and settings. In Section 3, we describe the model problem and introduce the boundary value correction method. In Section 4, the discrete space and the discrete form are defined and analyzed. In Section 5, we prove the optimal error estimate. In Section 6, numerical results are presented. Finally, we draw our conclusions in Section 7.
2 Notations and preliminaries
Let be a bounded open set in with Lipschitz continuous and piecewise boundary . Let be a body-fitted triangulation of , i.e., all boundary vertices of lie on . We assume that is geometrically conforming, shape regular, and quasi-uniform. Denote by the polygonal region occupied by the triangular mesh , and by the boundary of . When has a curved boundary, the polygonal region does not coincide with , as shown in Fig. 1. Denote by the set of all boundary edges in and by all mesh elements that contain at least one edge in . For each , denote by the curved edge cut by two end points of (see Fig. 1). We further assume that each is continuous, i.e. cannot cross the connection point of two continuous pieces of . Note that may coincide with when part of is flat.
We use to denote a crescent-shaped region surrounded by and , as shown in Fig. 1. There are three possibilities: (a) ; (b) ; (c) and . If is convex, we have . But when is not convex, there is no inclusion relationship between and .
Throughout the paper, we assume that the mesh is geometrically conforming, shape regular, and quasi-uniform [10]. For each triangle , denote by the diameter of . Let . Similarly, for each edge , denote by the length of . Denote by the space of polynomials on with degree less than or equal to . We use the standard notation for Sobolev spaces [10]. Let , for and be the usual Sobolev space equipped with the norm and the seminorm . When , the space coincides with the square integrable space . Denote by the subspace of consisting of functions with mean value zero. The above notation extends to a portion or . For example, is the Sobolev norm on . We also use the notation to indicate the inner-product on , and to indicate the duality pair on or .
All of the above notations extend to vector-valued functions, following the usual rule of product spaces. We use bold face letters to distinguish the vector-valued function spaces from the scalar-valued ones. For example, and are the spaces of vector-valued functions with each component in and , respectively. Define
The norm in is defined by
Throughout the paper, we use to denote less than or equal up to a constant, and the analogous notation to denote greater than or equal up to a constant.
Finally, we introduce some well-known inequalities.
Lemma 2.1.
(Trace Inequality [10]). For , we have
Lemma 2.2.
(Inverse Inequality [10]). Given integers . For , we have
Lemma 2.3.
(Trace-inverse Inequality [10]). Given an integer . For and be an edge of , we have
Lemma 2.4.
(Bramble-Hilbert Lemma [10]). Given an integer . For and integers , one has
3 Model problem and boundary value correction method
In this section, we briefly describe the model problem and the boundary value correction method [8]. Consider the Darcy’s equation with the Neumann boundary condition
(3.1) | ||||||
with and . Here, denotes the unit outward normal on . Problem (3.1) admits a unique weak solution as long as the following compatibility condition holds:
(3.2) |
The weak formulation of Problem (3.1) can be written as: Find satisfying on , such that
(3.3) |
where . The existence and uniqueness of a weak solution to the mixed problem (3.3) can be found in [10].
We use the Brezzi-Douglas-Marini (BDM) finite element method [11] on triangulation to discretize Equation (3.3). The discretization is defined on rather than on . Note that in (3.3), the Neumann boundary condition on becomes essential. This causes the main difficulty in its finite element discretization when is curved and hence not equal to . Another difficulty is that the compatibility condition (3.2) holds only on but not on . We shall take care of these issues one by one. First, we use a boundary correction technique [8] to transform the boundary data from to .
Assume that there exists a map such that
(3.4) |
as shown in Fig. 2, where is the unit vector pointing from to , and . For convenience, we denote . Define , i.e. the unit outward normal vector on is pulled to the corresponding . Denote by the unit outward normal vector on . Since is piecewise continuous, there exists a map satisfying [12]
(3.5) |
Remark 3.1.
In this paper, we do not specify the map . We only require that the distance function satisfies (3.5). For body-fitted meshes, this surely holds by simply setting . In fact, (3.5) can be true on certain unfitted meshes [12]. The method and analysis in this paper extend to unfitted meshes that satisfy (3.5), with a little modification.
For a sufficiently smooth vector-valued function defined between and , consider its -th order Taylor expansion along
where is the -th order directional derivative and the remainder term satisfies
For simplicity of notation, denote
(3.6) |
Both and are functions defined on . Denote by the pullback of from to . Then clearly
(3.7) |
Assume that (3.5) holds, we have the following two lemmas.
Lemma 3.1.
(cf. [23]). Given an integer . For and , one has
Lemma 3.2.
4 The finite element discretization
For a given integer , define the BDM finite element space on by
and denote .
For , define the local interpolation , by
(4.1) |
where is a bubble function, and , are the barycentric coordinates associated with . Denote the global interpolation by .
Denote by and the orthogonal projections onto and , respectively. Define the global projection .
Lemma 4.1.
Lemma 4.2.
For , the interpolation satisfies, for ,
(4.3) | ||||
(4.4) |
Proof.
Lemma 4.3.
Given a positive integer . For any nonnegative integers and any , one has
(4.5) | |||||
(4.6) | |||||
(4.7) |
Finally, we recall the extension property of Sobolev spaces:
Lemma 4.4.
(cf. [28]). Given a Lipschitz domain in and , there exists an extension operator such that
In Lemma 4.4, the hidden constant in may depend on the shape of . Throughout the paper, we shall only use Lemma 4.4 on but not on , in order to ensure that the hidden constant does not depend on . For brevity, we also denote the extension by . The extension and its notation can be extended to vector-valued functions.
4.1 Discretization with boundary correction
Let be the solution to (3.1) and be their extension to . Testing with in gives
(4.8) |
We point out that although in , the term is not in general equal to zero in . Next, using (3.7), it holds
where is a pull-back of the Neumann boundary data from to . Equation (4.8) can then be rewritten into
(4.9) | ||||
Again, note that vanishes in but not necessarily vanishes in .
Similarly, for any , one has
(4.10) |
We then introduce a discretization for (3.1), using (4.9)-(4.10). First, for that are smooth enough so that is well-defined, and for , define bilinear forms and linear forms
(4.11) | ||||
Define the discrete formulation: Find such that
(4.12) |
In the rest of this section, we shall prove the well-posedness of (4.12), and discuss its “compatibility” condition.
4.2 Well-posedness of (4.12)
We first define a mesh-dependent norm in as follows
Note that the norm is also meaningful for functions in smooth enough so that is well-defined.
Lemma 4.5.
Under the assumption (3.5), for we have
(4.13) | ||||
(4.14) |
Proof.
Lemma 4.6.
Under the assumption (3.5), for , we have
Proof.
Lemma 4.7.
For , and , we have
Proof.
From the definition, it is obvious that . The upper bounds of and follow immediately from the Schwarz inequality.
Lemma 4.7 gives the boundedness of , , , and the coercivity of . We then prove the - condition for and . Note that a function is mean-value free on but not on . This poses additional difficulty in the analysis. To this end, we first introduce the following notation to distinguish between two types of crescent-shaped regions , for each , as shown in Fig. 3,

Lemma 4.8.
(Inf-Sup condition). When is sufficiently small, there exist positive constants and independent of such that, for ,
Proof.
Each function can be naturally extended to , by filling the gap region with the same polynomial on neighboring mesh element. For simplicity, we still denote the extension by . Define . Since is mean value free in , we have
(4.15) |
Similar to [23], one gets
(4.16) |
It is well-known (see, e.g., [10]) that there exists a satisfying on and , where the hidden constant in may depend on the shape of but not on . Extend to by setting its value to be outside . The extension is still denoted by , which shall not be mistaken for the extension defined in Lemma 4.4. It is clear that, after the zero-extension, one has and hence the interpolation is well-defined. By the commutative property (4.2), the definition of and (4.16), one has
(4.17) | ||||
We then estimate the right-hand side of (4.17) one by one. Note that on . By (4.16), (4.15) and Lemma 3.1, one has
By the trace-inverse inequality in Lemma 2.3, one gets
Note that and on . By (3.11) and Lemma 4.4, one has
Combining the above, for sufficiently small, one gets
A similar argument gives
Then, for , one obtains
where is the interpolation of the zero-extension of , as explained before.
Finally, we prove , which combined with the above inequality will give the desired inf-sup condition. By Lemma 4.3 and the property (4.2), we have
and
By Lemma 4.5, 4.3 and definitions of and , it follows that
By (3.8) and Lemma 3.1 and Inequality (4.4), it holds
Note that , using lemmas 2.1, 3.1 and inequalities (3.9), (4.6), one gets
Combining the above, we obtain . This completes the proof of the lemma. ∎
Theorem 4.1.
4.3 Compatibility of (4.12) and an equivalent discrete form
Recall that a compatibility condition (3.2) is needed for the well-posedness of the continuous problem (3.1), as well as its weak formulation (3.3). We point out that the compatibility mechanism for the discrete problem works differently. In fact, the essential boundary condition on has been weakened in (4.12), and thus no relation is required between and . In Section 4.2, we have proved that the discrete system (4.12) admits a unique solution, and hence is “compatible”.
However, in practical implementation, the space is not commonly used. It is more convenient to construct a set of basis for than for . Hence we introduce an equivalent discrete problem: Find such that
(4.19) |
where and .
Lemma 4.9.
Proof.
As long as (4.12) is well-posed, problem (4.19) is solvable, i.e., the underlying discrete linear system is “compatible”. However, its solution is not unique since is also a solution. There are various ways to deal with this case, for example, using a Krylov subspace iterative solver, or adding a constraint to the linear system to ensure that the solution stays in .
5 Error analysis
We first obtain an error equation by subtracting equations (4.9)-(4.10) from the discrete problem (4.12)
(5.1) | ||||
Proof.
Lemma 5.2.
(Consistency Error). Assume that and satisfy Equation (3.1), we have
Lemma 5.3.
(Interpolation Error). Let with and . Let be the interpolation of , and be the orthogonal projection of . Then
Proof.
Lemma 5.4.
Proof.
Using the above lemmas and the triangle inequality, we get the main theorem of this section:
Theorem 5.1.
Under the same assumption of Lemma 5.4, one has
(5.5) | ||||
Moreover, the terms and vanish when is convex.
Remark 5.1.
(An implementation trick). When , one has for all , which greatly simplifies the implementation as one no longer needs to compute the Taylor expansion.
6 Numerical experiment
In this section, we present numerical results for problem (3.1) on two types of curved domains, as shown in Fig. 4. Body-fitted, unstructured triangular meshes are used in the computation. We always set , which simplifies the implementation according to Remark 5.1. For convenience, denote
If no boundary correction technique is used, then is just the norm.


.
The domain is a unit disk . The exact solution is
which satisfies a homogeneous Neumann boundary condition on , and . Note that , and are smooth, i.e. they satisfy the regularity requirement of Cor. 5.1. We solve Example on meshes illustrated in Fig. 4 (a), for various size of .
We first test Example using the primal mixed formulation [7] with discretization, without boundary correction. A homogeneous Neumann (essential) boundary condition is imposed on . We report in Tab. 1, which only reaches for , i.e. there is a loss of accuracy as expected.
order | order | order | ||||
---|---|---|---|---|---|---|
1/8 | 4.89e-01 | – | 3.48e-01 | – | 3.44e-01 | – |
1/16 | 3.06e-01 | 0.68 | 2.53e-01 | 0.46 | 2.51e-01 | 0.45 |
1/32 | 1.96e-01 | 0.65 | 1.77e-01 | 0.52 | 1.76e-01 | 0.51 |
1/64 | 1.29e-01 | 0.60 | 1.22e-02 | 0.54 | 1.22e-02 | 0.53 |
We then solve the same problem using the proposed discretization (4.12) with boundary correction. We first set and present the result in Tab. 2. An optimal convergence is observed, which agrees well with the theoretical result in Cor. 5.1.
Cor. 5.1 also gives a lower bound in order for the scheme to reach optimal convergence. For , the smallest that satisfies the lower bound is , respectively. We test the scheme (4.12) with these values, and present the results in Tab. 3. Again, optimal convergence is observed, as predicted in Cor. 5.1.
order | order | order | ||||
---|---|---|---|---|---|---|
1/8 | 3.84e-01 | – | 3.54e-03 | – | 6.54e-05 | – |
1/16 | 1.94e-01 | 0.98 | 7.61e-04 | 2.22 | 7.45e-06 | 3.13 |
1/32 | 9.47e-02 | 1.03 | 1.67e-04 | 2.18 | 7.85e-07 | 3.25 |
1/64 | 4.64e-02 | 1.03 | 3.86e-05 | 2.12 | 8.73e-08 | 3.17 |
order | order | order | ||||
---|---|---|---|---|---|---|
1/8 | 3.86e-01 | – | 3.57e-03 | – | 6.47e-05 | – |
1/16 | 1.95e-01 | 0.99 | 7.71e-04 | 2.21 | 7.42e-06 | 3.13 |
1/32 | 9.48e-02 | 1.04 | 1.69e-04 | 2.19 | 7.82e-07 | 3.25 |
1/64 | 4.65e-02 | 1.03 | 3.89e-05 | 2.12 | 8.68e-08 | 3.17 |
We repeat the above tests in a ring domain , which is not convex. The exact solution is
which satisfies a non-homogeneous Neumann boundary condition on and . This time, we did not perform the test without boundary correction, since there is no effective way to impose the nonhomogeneous Neumann boundary condition on .
We test Example on meshes illustrated in Fig. 4 (b), for various sizes of , using the scheme (4.12) with boundary correction. The results are reported in Tabs. 4 and 5, both indicating optimal convergence. This verifies that Cor. 5.1 holds on nonconvex domains.
order | order | order | ||||
---|---|---|---|---|---|---|
1/8 | 1.36e+01 | – | 1.66e+00 | – | 1.50e-01 | – |
1/16 | 6.86e+00 | 0.99 | 4.20e-01 | 1.98 | 1.89e-02 | 2.99 |
1/32 | 3.44e+00 | 1.00 | 1.05e-01 | 2.00 | 2.37e-03 | 3.00 |
1/64 | 1.72e+00 | 1.00 | 2.64e-02 | 2.00 | 2.96e-04 | 3.00 |
order | order | order | ||||
---|---|---|---|---|---|---|
1/8 | 1.36e+01 | – | 1.66e-00 | – | 1.50e-01 | – |
1/16 | 6.86e+00 | 0.99 | 4.20e-01 | 1.98 | 1.89e-02 | 2.99 |
1/32 | 3.44e+00 | 1.00 | 1.05e-01 | 2.00 | 2.37e-03 | 3.00 |
1/64 | 1.72e+00 | 1.00 | 2.64e-02 | 2.00 | 2.96e-04 | 3.00 |
7 Conclusion
In this paper, we consider the mixed finite element method for second-order elliptic equations on domains with curved boundaries. A boundary value correction technique is proposed to compensate for the loss of accuracy in the discretization. The proposed scheme achieves optimal convergence. Numerical experiments further validate the theoretical results.
References
- [1] N. M. Atallah, C. Canuto, and G. Scovazzi, The high-order shifted boundary method and its analysis, Computer Methods in Applied Mechanics and Engineering, 394:114885, 2022.
- [2] S. Bertoluzza, M. Pennacchio, and D. Prada, Weakly imposed Dirichlet boundary conditions for 2D and 3D virtual elements, Computer Methods in Applied Mechanics and Engineering, 400:115454, 2022.
- [3] F. Bertrand, S. Munzenmaier, and G. Starke. First-order system least squares on curved boundaries: Lowest-order Raviart-Thomas elements, SIAM Journal on Numerical Analysis, 52:880-894, 2014.
- [4] F. Bertrand, S. Munzenmaier, and G. Starke, First-order system least squares on curved boundaries: Higher-order Raviart-Thomas elements, SIAM Journal on Numerical Analysis, 52:3165-3180, 2014.
- [5] F. Bertrand and G. Starke, Parametric Raviart-Thomas elements for mixed methods on domains with curved surfaces, SIAM Journal on Numerical Analysis, 54:3648-3667, 2016.
- [6] D. Boffi, F. Brezzi, and M. Fortin, Mixed finite element methods and applications. Springer, 2013.
- [7] D. Braess, Finite elements: theory, fast solvers, and applications in solid mechanics, Cambridge University Press: Cambridge, UK, 2007.
- [8] J. H. Bramble, T. Dupont, and V. Thomée, Projection methods for Dirichlet’s problem in approximating polygonal domains with boundary-value corrections, Mathematics of Computation, 26:869-879, 1972.
- [9] J. H. Bramble and J. T. King, A robust finite element method for nonhomogeneous Dirichlet problems in domains with curved boundaries, Mathematics of Computation, 63:1-17, 1994.
- [10] S. C. Brenner and L. R. Scott, The mathematical theory of finite element methods, Springer, New York, 2008.
- [11] F. Brezzi, J. Douglas, Jr., and L. D. Marini, Two families of mixed finite elements for second order elliptic problems, Numerische Mathematik, 47(2):217-235, 1985.
- [12] E. Burman, P. Hansbo, and M. G. Larson, A cut finite element method with boundary value correction, Mathematics of Computation, 87:633-657, 2018.
- [13] E. Burman and R. Puppi, Two mixed finite element formulations for the weak imposition of the Neumann boundary conditions for the Darcy flow, Journal of Numerical Mathematics, 30:141-162, 2022.
- [14] P. Cao, J. Chen, and F. Wang, An extended mixed finite element method for elliptic interface problems, Computers Mathematics with Applications, 113:148-159, 2022.
- [15] B. Cockburn, D. Gupta, and F. Reitich, Boundary-conforming discontinuous Galerkin methods via extensions from subdomains, Journal of Scientific Computing, 42:144-184, 2009.
- [16] B. Cockburn and M. Solano, Solving Dirichlet boundary-value problems on curved domains by extensions from subdomains, SIAM Journal on Scientific Computing, 34:A497-A519, 2012.
- [17] J. A. Cottrell, T. J. R. Hughes, and Y. Bazilevs, Isogeometric Analysis: Toward Integration of CAD and FEA, John Wiley Sons, Ltd., Chichester, 2009.
- [18] C. D’Angelo and A. Scotti, A mixed finite element method for Darcy flow in fractured porous media with non-matching grids, ESAIM: Mathematical Modelling and Numerical Analysis, 46: 465-489, 2012.
- [19] R. Durán, Mixed finite element methods, vol. 1939 of Lecture Notes in Mathematics, Springer-Verlag, Berlin, 1-44, 2008, lectures given at the C.I.M.E. Summer School held in Cetraro, June 26-July 1, 2006. Edited by D. Boffi and L. Gastaldi.
- [20] I. Ergatoudis, B. Irons, and O. Zienkiewicz, Curved, isoparametric, quadrilateral elements for finite element analysis, International Journal of Solids and Structures, 4:31-42, 1968.
- [21] T. J. R. Hughes, J. A. Cottrell, and Y. Bazilevs, Isogeometric analysis: CAD, finite elements, NURBS, exact geometry and mesh refinement, Computer Methods in Applied Mechanics and Engineering, 194:4135-4195, 2005.
- [22] M. Lenoir, Optimal isoparametric finite elements and error estimates for domains involving curved boundaries, SIAM Journal on Numerical Analysis, 23(3):562-580, 1986.
- [23] Y. Liu, W. Chen, and Y. Wang, A weak Galerkin mixed finite element method for second order elliptic equations on 2D curved domains, Communications in Computational Physics, 32:1094-1128, 2022.
- [24] A. Main and G. Scovazzi, The shifted boundary method for embedded domain computations. Part I: Poisson and Stokes problems, Journal of Computational Physics, 372:972-995, 2018.
- [25] R. A. Nicolaides, Existence, uniqueness and approximation for generalized saddle point problems, SIAM ournal on Numerical Analysis, 19:349-357, 1982.
- [26] R. Oyarzúa, M. Solano, and P. Zúñiga, A high order mixed-FEM for diffusion problems on curved domains, Journal of Scientific Computing, 79:49-78, 2019.
- [27] R. Puppi, A cut finite element method for the Darcy problem, Preprint arXiv:2111.09922 [math. NA].
- [28] E. M. Stein, Singular integrals and differentiability properties of functions, Princeton University Press, 2, 1970.
- [29] G. Strang and A. E. Berger, The change in solution due to change in domain, Partial Differential Equations (D. C. Spencer, ed.), Proc. Sympos. Pure Math., vol. 23, Amer. Math. Soc, Providence, RI, 199-205, 1973.
- [30] V. Thomée, Polygonal domain approximation in Dirichlet’s problem, Ima Journal of Applied Mathematics, 11:33-44, 1973.