Abstract
In this paper, we develop an H(div)-conforming finite element method for Biot’s consolidation model in poroelasticity. In our method, the flow variables are discretized by an H(div)-conforming mixed finite elements. For relaxing the H1-conformity of the displacement, we approximate the displacement by using an H(div)-conforming finite element method, in which the tangential components are discretized in the interior penalty discontinuous Galerkin framework. For both the semi-discrete and the fully discrete schemes, we prove the existence and uniqueness theorems of the approximate solutions and derive the optimal convergence rate for each variable.
Keywords: poroelasticity, mixed finite element, H(div)-conforming, discontinuous Galerkin method
1. Introduction
Poroelasticity [3] is of increasing interest because of its vital importance in various science and engineering applications. For example, the mathematical models for carbon sequestration in environment engineering, seismic wave propagation in earthquake prediction, surface subsidence, evolution of fractured reservoirs during gas production, and biomechanical descriptions of tissues and bones are all poroelastic models. These models describe the interactions between a fluid flow and a deformable elastic porous medium which is saturated in the fluid. In this work, we are interested in Biot’s consolidation model. In the model, the motion of fluid in the porous medium is described by Darcy’s law, whereas the deformation of the porous medium is governed by linear elasticity.
Because of the complex nature of Biot’s model and the domain is usually irregular, it is not easy to obtain an analytical solution of this model. Thus, many researchers turn their attention to computational simulations. However, noting that both fluid dynamics and elasticity are involved in Biot’s model, it is very important to design effective numerical methods that can mimic the physical laws involved in. Unfortunately, all numerical difficulties that existing in linear elasticity and fluid mechanics will also arise in numerical approximations of Biot’s model. For the linear elasticity part, continuous Galerkin approximation of the displacement may cause locking or nonphysical pressure oscillations [5,27,30]. For eliminating the locking phenomenon in solving linear elasticity model, one can apply mixed Finite element method [23, 37] or employ discontinuous Galerkin (DG) method [30], or use the nonconforming finite element [36], or use weak Galerkin methods [9,17,32]. In numerical methods for incompressible fluid flow models, the standard Stokes elements (such as Taylor-Hood element and Mini element) have the shortcoming in that they do not satisfy the divergence constraint strongly or globally and therefore are not mass conservative [11,12,18].
In this work, we follow the strategy developed in [11,12,33] and adopt an H(div)-conforming finite element for the displacement. The purpose is to relax the H1-conformity of displacement. The advantages of adopting such a discretization are two-fold: On one hand, the normal components of displacement across elements are continuous and therefore are locally conservative; On the other hand, the tangential components are discretized through an interior penalty discontinuous Galerkin method. As it is discontinuous Galerkin approximation, such a discretization enables us to overcome the locking phenomenon and the pressure oscillation [18,31,38]. We comment here that applying H(div)-conforming finite elements in a DG framework was initially proposed in [11] (see also [12,33]) for solving Stokes equations in fluid mechanics. Later, this method is extended to solve the Darcy-Stokes interface problems [10,19], Brinkman problem [20] and magnetic induction model [8]. In Biot’s model, for the fluid part, we note that the governing equation is Darcy’s law. If the mixed form of Darcy’s law is used, it is natural to apply an H(div)-conforming finite element discretization to approximate the flow variables pressure because such a discretization can guarantee the mass conservation. In this work, we adopt Brezzi-Douglas-Marini (BDMk) space for both the flow variables and the displacement. Moreover, we present a unified treatment of both flow variables and the displacement in our Finite Element method. This work can be regarded as a further development of H(div)-conforming finite element methods for solving Biot’s problem. By using the framework presented in [28,29,36,38], we give a detailed analysis of our method. In particular, for both the semi-discrete and the fully discrete schemes for Biot’s model, we prove the existence and uniqueness theorems of the approximate solutions and derive the optimal convergence rate for each variable.
The rest of this paper is organized as follows. In Section 2, we describe Biot’s consolidation model, the functional spaces and the corresponding weak formulation. A spatial semi-discrete scheme based on H(div)-conforming elements is proposed in Section 3. The existence and uniqueness theorems for the semi-discrete numerical scheme are proved. Moreover, we derive the a priori error estimates of the solution of the semi-discrete scheme. In Section 4, a fully discrete numerical scheme based on the backward Euler time discretization is presented and analyzed. Conclusions are drawn in Section 5.
2. Biot’s consolidation model and its weak formulation
Let be a bounded convex polygonal domain with a Lipschitz boundary ∂Ω. We consider the following Biot’s consolidation model in Ω over a time interval (0,T]:
| (2.1a) |
| (2.1b) |
| (2.1c) |
Here, u(x,t) is the displacement of the solid phase, p(x,t) is the fluid pressure, and q(x,t) is the Darcy volumetric fluid flux,
| (2.2) |
In the above expressions, σ(x,t) is the total stress tensor with λ and μ being the Lamé constants, c0 ≥ 0 is the storage coefficient, and α is the Biot.-Willis constant [3], ψ is a source term, f is the external force, K(x) is a symmetric and uniformly positive definite tensor satisfying
| (2.3) |
Here, ξ is any 2-by-l vector, kmin and kmax are two positive constants.
Denoting Гd and Гt as the Dirichlet. boundary and the traction boundary for the elastic variables, denoting Гp and Гf as the pressure Dirichlet boundary and the fluid normal flux boundary, we assume that and . The boundary conditions and initial conditions for the above Biot. system read as:
| (2.4a) |
| (2.4b) |
| (2.4c) |
| (2.4d) |
| (2.4e) |
| (2.4f) |
Here, n denotes the unit outward normal vector.
Let us introduce some notations. As usual, Hs() denotes the standard Sobolev space of functions with regularity exponent s ≥ 0. The associated norm and the semi-norm are denoted as and . When s = 0, H0() is L2(). For simplicity, when = Ω, the norm is written as . For the space (Hs())2, its norm is still denoted by . A subspace of H1(Ω) with vanishing trace on Γd is denoted by . Furthermore, we define with its graph norm . Two subspaces of H(div; Ω) are and . For the ease of notations, we set ,, and .
Multiplying by test functions and integrating by parts, the standard mixed weak formulation of (2.1) reads as: find (p,q, u) ∈ × × V such that, t ∈ (0,T],
| (2.5a) |
| (2.5b) |
| (2.5c) |
Here and hereafter, denotes the inner product in with being the product of tensors.
In the sequel, we shall deal with functions of time and space. To this end, we introduce the standard Bochner space Lp(0,T; Hs(Ω)), which consists of all functions with norm
for 1 ≤ p < ∞. When p = ∞, the space is endowed with the norm
3. A semi-discrete scheme
In this section, we discuss on how to conduct the spatial discretization and present the corresponding semi-discrete numerical scheme. Let be a shape-regular triangulation of Ω. We denote hK as the diameter of K and . Moreover, we denote as the set of interior edges of elements in , as the set of boundary edges on Γd and as the set of boundary edges on Γt. Set . The length of an edge is denoted by he. Moreover, we introduce the set . The shape-regularity of the mesh implies that there exits an integer N∂ > 0, independent of h, such that
| (3.1) |
This means that the maximum number of edges that are related to K is uniformly bounded (see Lemma 1.41 in [13]). We associate each edge e ∈ ℰh with a fixed unit normal n and ensure that the unit normal for each edge on the boundary ∂Ω is exactly the exterior unit normal n. Let be an interior edge, shared by two elements K1 and K2. For a scalar piecewise smooth function φ with , we define the average and jump by
On a boundary edge ,
Define
| (3.2) |
| (3.3) |
and
| (3.4) |
Here, BDMk(k ≥ 1) is the H(div)-conforming space introduced by Brezzi, Douglas and Marini [6], and Pk(K) denotes the space of polynomials of degree less than or equal to k on K. Let be the BDMk. interpolation [6], and Ph be the L2– projection from L2(Ω) onto . It is well known that the following properties hold true [6]:
| (3.5a) |
| (3.5b) |
| (3.5c) |
| (3.5d) |
| (3.5e) |
Here and in the following, we use C to denote a positive generic constant (may take different values at different occurrences), which is independent of h, △t, and Lamé constants μ and λ.
3.1. An H(div)-conforming element method
Multiplying the equation (2.1c) by any , integrating by parts on every element K, and then summing over all elements in , we obtain
| (3.6) |
Note that in the above equality we have used the fact that is continuous across each interior edge. For an edge e, if n and τ are the unit normal and tangential vectors which form a right-handed coordinate system, there holds the following decomposition,
Applying the above decomposition yields
Noting from the above decomposition, the equality [ab] = [a]{b} + {a}[b], the regularity of the exact solution, and the fact is continuous across each interior edge, one can derive that
Thus (3.6) is reduced to
| (3.7) |
As with the usual interior penalty DG methods [1], adding some stabilized terms in the above equation, and noting that σn = 0 on Γt, our DG approximation of (2.1c) is
| (3.8) |
where
| (3.9) |
From the definitions of functional spaces and ah, we note that the exact solutions of (2.1a), (2.1b) and (2.1c) satisfy
| (3.10a) |
| (3.10b) |
| (3.10c) |
Naturally, the corresponding H(div)-conforming finite element method for (2.1a), (2.1b) and (2.1c) reads as: given the initial conditions and , find such that
| (3.11a) |
| (3.11b) |
| (3.11c) |
3.2. The existence and uniqueness
In order to prove the existence and uniqueness of the solutions of (3.11), we will use the theory of differential-algebraic equations (DAEs) developed in [36].
By introducing the corresponding finite element basis functions, one can represent the solutions qh(x,t), ph(x,t) and uh(x,t) as
Here,, ;,; and . Similarly, we define row vectors and according to the right hand side. Rearranging the above equations, one can rewrite (3.11) as an equivalent system of DAEs:
| (3.12) |
Here,, , and
| (3.13) |
where auu, aqq, app, aup and, aqp denote the matrices corresponding to the bilinear forms ah(uh, v), (K–1qh,z), c0(p,w), α(∇∙uh,w) and (∇∙qh,w) in (3.11), respectively. According to the theory of DAEs, as pointed out in [36], it is sufficient to prove the existence and uniqueness of (3.12) by verifying the existence and uniqueness of the following saddle point problem: find
| (3.14a) |
| (3.14b) |
where
To prove the existence and uniqueness of problem (3.14), by using the theory of saddle point problems [7], it is enough to prove that the above bilinear forms satisfy certain LBB conditions. For the subsequent analysis, we define two mesh-dependent, norm and by
and
Actually, one can define another norm by
Using the discrete version of the Korn’s inequality [4], it can be proved that , and are equivalent on . The details of the proof can be found in [2,15].
Let K be an element with e as an edge. For all , it is well known [4] that there exists a constant C > 0 such that
| (3.15) |
Then, by the shape-regularity of the mesh, there holds [4, 24]
| (3.16) |
Applying the standard inverse inequality to the last term of the above inequality, we see that
| (3.17) |
where Ctr depends only on the polynomial degree k and the shape-regularity of the mesh. Thus, there exists a constant C0 > 0 such that
| (3.18) |
with .
Setting , then we have the following lemma.
Lemma 3.1.
There exists a constant Ccont > 0, independent of μ and λ, such that
| (3.19) |
Furthermore, if the penalty parameter γ is sufficiently large, then there exists a constant Ccoer > 0 such that.
| (3.20) |
Here, Ccoer does not. depend on the Lamé constants μ and λ.
Proof.
The inequality of (3.19) can be easily derived from the Cauchy-Schwarz inequality. It leaves us to prove (3.20). Using Young’s inequality and a trace inequality (3.17), we have, for any ε > 0,
| (3.21) |
Here N∂ and Ctr axe defined in (3.1) and (3.17), respectively. Substituting the above inequality into (3.9) yields
| (3.22) |
Setting ε = 2N∂Ctr in the above inequality and choosing a sufficiently large penalty parameter γ to ensure , we have
| (3.23) |
Here, . Combining (3.23) with (3.18), we have
| (3.24) |
The inequality (3.20) follows by setting . Since Ctr depends only on the polynomial degree k and the shape-regularity of the mesh and C0 = 1 + Ctr, we see that depend on the Lamé constants μ and λ.
Remark 3.1.
In general, as in other interior penalty DG methods, one can choose to obtain (3.20). In fact, in (3.22), setting and , i.e., , we have
| (3.25) |
where . This, together with (3.18), gives (3.20). We further comment here that the constant depends on the polynomial degree k. For two dimensional triangle elements, Ctr scales as k(k + 2). More comments on Ctr can be found in [14, 34] and Remark 1.48 in [13]. In actual computation, one can choose . More discussions on choosing γ can be found in Remark 2.1 in [16].
For the space , we equip it with a discrete norm
Lemma 3.2.
If the penalty parameter γ is sufficiently large, then there exists a constant C > 0 such that
| (3.26) |
Proof.
Lemma 3.3.
There exists a positive constant β > 0 such that
| (3.27) |
Proof.
For any , there exists a such that (cf. Lemma 11.2.3 in [5])
| (3.28) |
From (3.5d), we note that
| (3.29) |
Setting v = 0, by using (3.28) and (3.29), we see that
The lemma follows by setting .
In Lemmas 3.2 and 3.3, we have proved the LBB condition of the saddle point problems (3.14). Noting that the bilinear form is symmetric positive semidefinite, we then obtain the following main result of this subsection.
Theorem 3.1.
The semidiscrete scheme (3.11) has a unique solution.
Remark 3.2.
The author of a recent work [38] has pointed out that if ker , one can remove spurious pressure oscillations which arise when c0 = 0 and K → 0. Since we use standard mixed finite element spaces and for the displacement and pressure variables, there naturally holds . Therefore, there will be no spurious pressure oscillation by using our method.
3.3. Error estimates for the semi-discrete scheme
3.3.1. Error estimates for the case c0 ≥ β0 > 0
Theorem 3.2.
Let and be the solutions of (2.5) and (3.11), respectively. Moreover, we assume that
Then, provided that the penalty parameter is sufficiently large, the following finite element error estimate holds.
| (3.30) |
where .
Proof.
Subtracting (3.10a), (3.10b) and (3.10c) from (3.11a), (3.11b) and (3.11c), respectively, we have
| (3.31a) |
| (3.31b) |
| (3.31c) |
We then split the error p–ph as with and . Similarly, with and with and . Since the estimates for ξp, ξq and ξu can be derived by the interpolation error bounds in (3.5b) and (3.5d), it leaves us to estimate θp, θq and θu. To this end, using (3.5c) and (3.5e), we can rewrite (3.31a), (3.31b) and (3.31c) by
| (3.32a) |
| (3.32b) |
| (3.32c) |
Setting w = θp, Z = θq and v = (θu)t in the above equations and using the chain rule in time and the symmetry of , we obtain
| (3.33) |
| (3.34) |
| (3.35) |
The initial conditions imply that Using this fact, summing equations (3.33)–(3.35), integrating in time from 0 to t(≤T), we obtain
| (3.36) |
Here,
For B1, we can bound it as follows:
| (3.37) |
For B2, integrating by parts, we firstly obtain
| (3.38) |
Then noting that using (3.19) and Young’s inequality, we further have
| (3.39) |
with ε being an arbitrarily small number.
Noting from the above bounds, using (2.3) and (3.20), we have
| (3.40) |
We can choose ε small enough to make be positive. We note that the above inequality still holds if one replaces the left-hand side of (3.40) by Therefore, dividing both sides of the above inequality by Cmin and using Gronwall’s lemma, we have
| (3.41) |
Noting that the above estimate holds for all 0 ≤ t ≤ T, and using some appropriate approximation properties of Ph in (3.5b) and in (3.5d), we obtain
| (3.42) |
where, This estimate can be rewritten by the following equivalent formulation:
| (3.43) |
Combining the above estimate with the interpolation error estimates for ξp, ξq and ξu, and using the triangle inequality, we obtain the assertion (3.30).
3.3.2. Error estimates for the case c0 ≥ 0
Note that the results in Theorem 3.2 in the previous subsection hold under the assumption that c0 > 0. If c0 = 0, the optimal error estimates are derived using the weaker norm. To this end, we need the following lemma [10].
Lemma 3.4.
Let be the solutions of (2.5) and (3.11), respectively. Then, there exists a constant c0 > 0 such that
By using the above result, we can obtain the following main result.
Theorem 3.3.
Under the same assumption as that in Theorem 3.2, the following error estimate holds.
| (3.45) |
Proof.
Squaring both sides of (3.44) and then integrating them in time from 0 to T, we see that Then, the desired result follows from the error bound in (3.30), the interpolation estimates and the triangle inequality.
4. The fully discrete scheme
4.1. The fully discrete scheme
For simplicity, we apply the backward Euler method as the time discretization scheme. Let N be a positive integer and let The fully discrete approximation of (3.11) reads as: given the initial conditions at each time t = tn, find such that
| (4.1a) |
| (4.1b) |
| (4.1c) |
4.2. The existence and uniqueness
In this subsection, we will show the existence and uniqueness of solutions of (4.1) for each time step t = tn, 1 ≤n≤ N. Firstly, the Eqs. (4.1a)–(4.1c) can be transformed into the following equivalent variational formulation.
| (4.2a) |
| (4.2b) |
Here, the bilinear forms are
Similar to the semi-discrete case, to prove the existence and uniqueness of the saddle point problem (4.2), it is sufficient to verify that these bilinear forms satisfy LBB conditions [7]. To this end, we need to define a discrete time-dependent, norm for the space namely,
Lemma 4.1.
If the penalty parameter γ is sufficiently large, then there exists a constant C > 0 such that
| (4.4) |
Proof.
The assertion follows from the definition of in (4.3), (3.20) and (2.3).
Lemma 4.2.
There exists a positive constant β > 0 such that
| (4.5) |
Proof.
For any there exists a such that (cf. Lemma 11.2.3 in [5])
| (4.6) |
From (3.5d), we obtain
| (4.7) |
In view of (4.6) and (4.7), and setting v = 0, we have
The desired result follows by setting
In Lemmas 4.1 and 4.2, we have proved the LBB conditions of the saddle point problem (4.2). Noting that the bilinear form is symmetric positive semidefinite, then we obtain the following main result.
Theorem 4.1.
At each time t = tn( 1 ≤ n ≤ N), the fully discrete numerical scheme (4.1) has a unique solution if the penalty parameter γ is sufficiently large.
4.3. Error estimates for the fully discrete scheme
4.3.1. Error estimates for the case c0 ≥ β0 > 0
For any function g(t, x), at each time we denote By Taylor’s expansion, there hold
| (4.8) |
| (4.9) |
Theorem 4.2.
Let be the solutions of (2.5) and (4.1), respectively. Moreover, we assume that
and that the penalty parameter γ is sufficiently large. Then, the following error estimate holds.
| (4.10) |
Proof.
We note that (3.10) holds for the exact solution at any time t = tn. Using this fact, combining with (4.8) and (4.9), we see that
| (4.11a) |
| (4.11b) |
| (4.11c) |
for any
Subtracting (4.1a), (4.1b) and (4.1c) from (4.11a), (4.11b) and (4.11c), respectively, we obtain
| (4.12b) |
| (4.12c) |
We then split the error Similarly, Since the estimates for can be derived by the interpolation error bounds, it leaves us to estimate To this end, using (3.5c) and (3.5e), we can rewrite (4.12a), (4.12b) and (4.12c) by
| (4.13a) |
| (4.13b) |
| (4.13c) |
Setting in the above equations and adding them together, we obtain
| (4.14) |
To estimate the bounds for the above error equation, we need the following inequalities.
| (4.15) |
and
| (4.16) |
In view of the above inequalities, summing (4.14) from 1 to m(≤ N), and noting that we obtain
| (4.17) |
where
The first term T1 can be bounded by
Since
then we further have
| (4.18) |
Similarly, the second term T2 can be bounded by
| (4.19) |
For the third term T3, it is easy to show that
To bound the last term T4, we need the following equalities.
| (4.21) |
and
| (4.22) |
Then, using (4.21), (4.22), (3.19) and Young’s inequality, and noting that we have
| (4.23) |
where ε is an arbitrarily small number.
Combining the bounds above, and using (2.3) and (3.20), we have
| (4.24) |
We can choose t being small enough to ensure is positive. Then, we note that the above inequality still holds if one replaces the left-hand side of (4.24) by Using the discrete Gronwall’s inequality, some approximation properties, and noting that (4.24) holds for any 1 ≤m < N, we see that
| (4.25) |
Combining the above estimate with the interpolation error estimates for and using the triangle inequality, we obtain the assertion (4.26).
4.3.2. Error estimates for the case c0 ≥ 0
Similar to the semi-discrete scheme, if c0 = 0, one can derive an optimal error bound for the pressure under a weaker norm. Specifically, we have
Theorem 4.3.
Let be the solutions of (2.5) and (4.1), respectively. Under the same assumption as that in Theorem 4.2, the following error estimate holds:
| (4.26) |
5. Concluding remarks
In this work, we propose an H(div)-conforming Finite Element method for solving Biot’s consolidation model. In our method, both the displacement and the fluid velocity are approximated by using BDMk space. As we use H(div)-conforming elements, the normal components of displacement and fluid velocity are continuous across element interfaces. Therefore, our method is locally conservative. Moreover, there is no pressure oscillation of our method because the continuity of the tangential component of elasticity part are imposed by using an interior penalty Discontinuous Galerkin method. After introducing the spatial discretization, we present a semi-discrete scheme and a fully discrete scheme. The existence and uniqueness of solutions of the semi-discrete scheme and fully discrete scheme are proved by analyzing the corresponding differential algebraic equations (DAEs). Then, under some assumptions on the regularities of the solution, we derive the optimal error bound for each variable.
Acknowledgments
The first author’s work is partially supported by Natural Science Foundation of Guangdong Province, China (Grant. No. 2018A030307024), and by National Natural Science Foundation of China (Grant No. 11526097). The second author’s work is partially supported by NIH BUILD grant through Pilot project and Natural Science Foundation (Grant Nos. HRD-1700328 and DMS-1831950). The third author’s work is partially supported by National Natural Science Foundation of China (Grant Nos. 11371199, 11371198, 11871272, 11871281).
References
- [1].Arnold DN, Brezzi F, Cockburn B and Marini LD, Unified analysis of discontinuous Galerkin methods for elliptic problems, SIAM J. Numer. Anal, 39 (2002), 1749–1779. [Google Scholar]
- [2].Ayuso de Dios B, Brezzi F, Marini L, Xu J and Zikatanov L, A simple preconditioner for a discontinuous Galerkin method for the Stokes problem, J. Sci. Comput, 58 (2014), 517–547. [Google Scholar]
- [3].Biot MA, Theory of elasticity and consolidation for a porous anisotropic solid, J. Appl Phys, 26 (1955), 182–185. [Google Scholar]
- [4].Brenner S, Korn’s inequalities for piecewise H1 vector fields. Math. Comput. (2004), 1067–87. [Google Scholar]
- [5].Brenner SC and Scott LR, The Mathematical Theory of Finite Element Methods, 3rd ed., Springer-Verlag, New York, Berlin, Heidelberg, 2008. [Google Scholar]
- [6].Brezzi F, Douglas J and Marini LD, Two families of mixed finite elements for second order elliptic problems, Numer. Math, 47 (1985), 217–235. [Google Scholar]
- [7].Brezzi F and Fortin M, Mixed and Hybrid Fnite Element Methods, Springer-Verlag, 1991. [Google Scholar]
- [8].Cai W, Hu J and Zhang S, High order hierarchical divergence-free constrained transport H(div) finite element method for magnetic induction equation, Numer. Math. Theor. Meth. Appl.,10 (2017), pp. 243–254. [Google Scholar]
- [9].Chen Y, Chen G and Xie X, Weak Galerkin finite element method for Biot’s consolidation problem, J. Comput. Appl. Math, 330 (2018), 398–416. [Google Scholar]
- [10].Chen Y, Huang F and Xie X, H(div) conforming finite element methods for the coupled Stokes and Darcy problem, J. Comput. Appl. Math, 235 (2011), 4337–4349. [Google Scholar]
- [11].Cockburn B, Kanschat G and Schötzau D, A locally conservative LDG method for the incompressible Navier-Stokes equations, Math. Comp, 74 (2005), 1067–1095. [Google Scholar]
- [12].Cockburn B, Kanschat G and Schotzau D, A note on discontinuous Galerkin divergence- free solutions of the Navier-Stokes equations, J. Sci. Comput., 31 (2007), 61–73. [Google Scholar]
- [13].Di Pietro DA and Ern A, Mathematical Aspects of Discontinuous Galerkin Methods, Springer-Verlag, Berlin, 2012. [Google Scholar]
- [14].Hansbo P and Larson MG, Discontinuous Galerkin methods for incompressible and nearly incompressible elasticity by Nitsche’s method, Comput. Methods Appl. Mech. En- grg., 191 (2002), 1895–1908. [Google Scholar]
- [15].Hong Q, Kraus J, Xu J and Zikatanov L, A robust multigrid method for discontinuous Galerkin discretizations of Stokes and linear elasticity equations, Numer. Math, 132 (2016), 23–49. [Google Scholar]
- [16].Houston P, Perugia I and Schotzau D, An a posteriori error indicator for discontinuous Galerkin discretizations of H(curl)-elliptic partial differential equations, IMA J. Numer. Anal, 27 (2007), 122–150. [Google Scholar]
- [17].Hu X, Mu L and Ye X, Weak Galerkin method for the Biot’s consolidation model, Comput. Math. Appl, 75 (2018), 2017–2030. [Google Scholar]
- [18].John V, Linke A, Merdon C, Neilan M, and Rebholz L. On the divergence constraint in mixed finite element methods for incompressible flows. SIAM Rev., 59(3), (2017), 492–544. [Google Scholar]
- [19].Kanschat G and Riviere B, A strongly conservative finite element method for the coupling of Stokes and Darcy flow, J. Comput, Phys, 229 (2010), 5933–5943. [Google Scholar]
- [20].Könnö J and Stenberg R, H(div)-conforming finite elements for the Brinkman problem, Math Models Methods Appl. Sci., 21 (2011), 2227–2248. [Google Scholar]
- [21].Korsawe J and Starke G, A least-squares mixed finite element method for Biot’s consolidation problem in porous media, SIAM J. Numer. Anal, 43 (2005), 318–339. [Google Scholar]
- [22].Korsawe J, Starke G, Wang W and Kolditz O, Finite element analysis of poro-elastic consolidation in porous media: standard and mixed approaches, Comput. Methods Appl. Mech. Engrg., 195 (2006), 1096–1115. [Google Scholar]
- [23].Lee JJ, Robust error analysis of coupled mixed methods for Biot’s consolidation model, J. Sci. Comput, 69 (2016), 610–632. [Google Scholar]
- [24].Mardal K, Winther R, An observation on Korn’s inequality for nonconforming finite element methods. Math. Comput, 75(253) (2006), 1–6. [Google Scholar]
- [25].Murad MA and Loula AFD, Improved accuracy in finite element analysis of Biot’s consolidation problem, Comput. Methods Appl. Mech. Engrg., 95 (1992), 359–382. [Google Scholar]
- [26].Murad MA and Loula AFD, On stability and convergence of finite element approximations of Biot’s consolidation problem, Internat. J. Numer. Methods Engrg, 37 (1994), 645–667. [Google Scholar]
- [27].Murad MA, Thomee V and Loula AFD, Asymptotic behavior of semidiscrete finite- element approximations of Biot’s consolidation problem, SIAM J. Numer. Anal, 33 (1996), 1065–1083. [Google Scholar]
- [28].Phillips PJ and Wheeler MF, A coupling of mixed and continuous Galerkin finite element methods for poroelasticity I: the continuous in time case, Comput. Geosci, 11 (2007), 131–144. [Google Scholar]
- [29].Phillips PJ and Wheeler MF, A coupling of mixed and continuous Galerkin finite element methods for poroelasticity II: the discrete in time case, Comput. Geosci, 11 (2007), 145–158. [Google Scholar]
- [30].Phillips PJ and Wheeler MF, A coupling of mixed and discontinuous Galerkin finite element methods for poroelasticity, Comput. Geosci, 12 (2008), 417–435. [Google Scholar]
- [31].Phillip PJ and Wheeler MF, Overcoming the problem of locking in linear elasticity and poroelasticity: an heuristic approach, Comput. Geosci., 13 (2009), 5–12. [Google Scholar]
- [32].Sun M and Rui H, A coupling of weak Galerkin and mixed finite element methods for poroelasticity, Comput. Math. Appl, 73 (2017), 804–823. [Google Scholar]
- [33].Wang J and Ye X, New finite element methods in computational fluid dynamics by H(div) elements, SIAM J. Numer. Anal, 45 (2007), 1269–1286. [Google Scholar]
- [34].Warburton T and Hesthaven JS, On the constants in hp-finite element trace inverse inequalities, Comput. Methods Appl. Mech. Engrg, 192 (2003), 2765–2773. [Google Scholar]
- [35].Wheeler MF, Xue G and Yotov I, Coupling multipoint flux mixed finite element methods with continuous Galerkin methods for poroelasticity, Comput. Geosci, 18 (2014), 57–45. [Google Scholar]
- [36].Yi SY, A coupling of nonconforming and mixed finite element methods for Biot’s consolidation model, Numer. Methods PDEs, 29 (2013), 1749–1777. [Google Scholar]
- [37].Yi SY, Convergence analysis of a new mixed finite element method for Biot’s consolidation model, Numer. Methods PDEs, 30 (2014), 1189–1210. [Google Scholar]
- [38].Yi SY, A study of two modes of locking in poroelasticity, SIAM J. Numer. Anal, 55 (2017), 1915–1936. [Google Scholar]
