Loading [MathJax]/jax/element/mml/optable/BasicLatin.js
Research article Special Issues

A posteriori error estimates of mixed discontinuous Galerkin method for a class of Stokes eigenvalue problems

  • For a class of Stokes eigenvalue problems including the classical Stokes eigenvalue problem and the magnetohydrodynamic Stokes eigenvalue problem a residual type a posteriori error estimate of the mixed discontinuous Galerkin finite element method using PkPk1 element (k1) is studied in this paper. The a posteriori error estimators for approximate eigenpairs are given. The reliability and efficiency of the posteriori error estimator for the eigenfunction are proved and the reliability of the estimator for the eigenvalue is also analyzed. The numerical results are provided to confirm the theoretical predictions and indicate that the method considered in this paper can reach the optimal convergence order O(dof2kd).

    Citation: Lingling Sun, Hai Bi, Yidu Yang. A posteriori error estimates of mixed discontinuous Galerkin method for a class of Stokes eigenvalue problems[J]. AIMS Mathematics, 2023, 8(9): 21270-21297. doi: 10.3934/math.20231084

    Related Papers:

    [1] Liangkun Xu, Hai Bi . A multigrid discretization scheme of discontinuous Galerkin method for the Steklov-Lamé eigenproblem. AIMS Mathematics, 2023, 8(6): 14207-14231. doi: 10.3934/math.2023727
    [2] Jinhua Feng, Shixi Wang, Hai Bi, Yidu Yang . The a posteriori error estimates of the Ciarlet-Raviart mixed finite element method for the biharmonic eigenvalue problem. AIMS Mathematics, 2024, 9(2): 3332-3348. doi: 10.3934/math.2024163
    [3] Bo Tang, Huasheng Wang . The a posteriori error estimate in fractional differential equations using generalized Jacobi functions. AIMS Mathematics, 2023, 8(12): 29017-29041. doi: 10.3934/math.20231486
    [4] Yasir Nadeem Anjam . The qualitative analysis of solution of the Stokes and Navier-Stokes system in non-smooth domains with weighted Sobolev spaces. AIMS Mathematics, 2021, 6(6): 5647-5674. doi: 10.3934/math.2021334
    [5] Zhongdi Cen, Jian Huang, Aimin Xu . A posteriori mesh method for a system of singularly perturbed initial value problems. AIMS Mathematics, 2022, 7(9): 16719-16732. doi: 10.3934/math.2022917
    [6] Cagnur Corekli . The SIPG method of Dirichlet boundary optimal control problems with weakly imposed boundary conditions. AIMS Mathematics, 2022, 7(4): 6711-6742. doi: 10.3934/math.2022375
    [7] Song Lunji . A High-Order Symmetric Interior Penalty Discontinuous Galerkin Scheme to Simulate Vortex Dominated Incompressible Fluid Flow. AIMS Mathematics, 2016, 1(1): 43-63. doi: 10.3934/Math.2016.1.43
    [8] Shuangbing Guo, Xiliang Lu, Zhiyue Zhang . Finite element method for an eigenvalue optimization problem of the Schrödinger operator. AIMS Mathematics, 2022, 7(4): 5049-5071. doi: 10.3934/math.2022281
    [9] Chunjuan Hou, Zuliang Lu, Xuejiao Chen, Fei Huang . Error estimates of variational discretization for semilinear parabolic optimal control problems. AIMS Mathematics, 2021, 6(1): 772-793. doi: 10.3934/math.2021047
    [10] Zuliang Lu, Xiankui Wu, Fei Huang, Fei Cai, Chunjuan Hou, Yin Yang . Convergence and quasi-optimality based on an adaptive finite element method for the bilinear optimal control problem. AIMS Mathematics, 2021, 6(9): 9510-9535. doi: 10.3934/math.2021553
  • For a class of Stokes eigenvalue problems including the classical Stokes eigenvalue problem and the magnetohydrodynamic Stokes eigenvalue problem a residual type a posteriori error estimate of the mixed discontinuous Galerkin finite element method using PkPk1 element (k1) is studied in this paper. The a posteriori error estimators for approximate eigenpairs are given. The reliability and efficiency of the posteriori error estimator for the eigenfunction are proved and the reliability of the estimator for the eigenvalue is also analyzed. The numerical results are provided to confirm the theoretical predictions and indicate that the method considered in this paper can reach the optimal convergence order O(dof2kd).



    Stokes eigenvalue problem is of great significance because of its role in the stability analysis of fluid mechanics. Therefore, it is of great interest to study efficient numerical methods for solving this problem (e.g. see [1,2,3,4,5,6,7,8]).

    The a posteriori error estimates of the classical Stokes eigenvalue problem based on the velocity-pressure formulation received much attention from scholars. For example, the authors in [9,10,11] studied the low-order conforming mixed method and Jia et al. [12] and Sun et al. [6] explored the low-order non-conforming finite elements. Since the accuracy of low-order elements is not high Gedicke et al. [1] used the Arnold-Winther hybrid finite element method to analyze the a posteriori error estimation based on the stress-velocity formulation in R2 and Gedicke et al. [2] adopted the divergence-conforming discontinuous Galerkin finite element method (DGFEM for short) to discuss the a posteriori error estimate for velocity-pressure formulation on shape-regular rectangular meshes. Lepe et al. [13] proposed a mixed numerical method to study the error estimates for a vorticity-based velocity-stress formulation of the Stokes eigenvalue problem.

    In this paper, for a class of Stokes eigenvalue problems (see (2.1)), including the classical Stokes eigenvalue problem in Rd(d=2,3) and the magnetohydrodynamic (MHD) Stokes eigenvalue problem et al. based on the velocity-pressure formulation we study the residual type a posteriori error estimates of the mixed DGFEM using PkPk1 (k1) element on shape-regular simplex meshes. For the Stokes equations the DGFEM was researched by [14,15,16,17,18,19] which laid a foundation for us to study further the Stokes eigenvalue problem. Among them, Badia et al.[14] proved the well-posedness of discrete DG formulation and studied the a priori error estimate of DGFEM with PkPk1 element without using the discrete inf-sup condition. Our main work is as follows:

    (1) We present the a posteriori error estimators for approximate eigenpairs. Referring to [20,21], using the enriching operator (see [22,23]) and the lifting operator (see [24,25]) we prove the reliability and efficiency of the estimator for eigenfunctions. We establish an identity (see Lemma 3.8) and use it to analyze the reliability of the estimator for eigenvalues. The characteristic of the adaptive DGFEM discussed in this paper is that for the Stokes eigenvalue problem in two and three-dimensional domains due to the usage of high-order elements it can capture smooth solutions and achieve the optimal convergence order for local less smooth solutions (eigenfunctions that have local singularity or local low smoothness) on adaptive locally refined graded meshes.

    (2) We implement adaptive computing and the numerical results confirm our theoretical predictions and show that our method is stable, efficient and can obtain high-accuracy approximate eigenvalues. In the existing literature on the classical Stokes eigenvalue problem, Gedicke et al. [1,2] presented the approximate eigenvalues of 11, 10 and 9 significant digits in the unit square, the L-shaped domain and the slit domain, respectively, which are the most accurate approximations in the existing literature. In this paper, we obtain approximate eigenvalues that have the same accuracy as those in [1,2].

    Note that C in different positions in this article represents different positive constants which is independent of mesh size h. We use ab to represent aCb and ab to represent ab and ba.

    Consider the following class of Stokes eigenvalue problems:

    {μΔu+Au+p=λu,inΩ,divu=0,inΩ,u=0,onΩ, (2.1)

    where ΩRd(d=2,3) is a bounded polyhedral domain, u=(u1,...,ud)T is the velocity of the flow, p is the pressure, μ>0 is the kinematic viscosity parameter of the fluid, λ is the eigenvalue of the problem (2.1) and A is a d×d symmetric semi-definite matrix whose elements belong to L(Ω).

    The problem (2.1) includes the classical Stokes eigenvalue problem in Rd(d=2,3) and the MHD Stokes eigenvalue problem et al. When A is a zero matrix (2.1) is the classical Stokes eigenvalue problem. In the case of d=2 when

    A=(Ha)2(H20000),

    the problem (2.1) is the MHD Stokes eigenvalue problem (see [7,8]) where H0 is the intensity of the externally applied magnetic field on the vertical direction, i.e., magnetic field H=(0,H0,0), and Ha is the Hartmann number (see [7,8]) and when

    A=(Ha)2(000H20),

    the problem (2.1) is the MHD Stokes eigenvalue problem while the magnetic field is applied horizontally.

    For the sake of narrative simplicity we take μ=1 in this paper.

    Let Hρ(Ω) be the Sobolev space on Ω of order ρ0 equipped with the norm ρ,Ω (denoted by ρ for simplicity). H10(Ω)={zH1(Ω),z|Ω=0}. We denote zρ=di=1ziρ for z=(z1,,zd)Hρ(Ω)d. We denote by (,) the inner product in L2(Ω)d which is given by (u,z)=Ωuzdx (d=1) and (u,z)=Ωuzdx (d=2,3). Define X=H10(Ω)d with the norm zX=(z,z)12 and define W=L20(Ω)={ϱL2(Ω):(ϱ,1)=0}.

    The weak formulation of (2.1) is given by: Find (λ,u,p)R×X×W, u0=1 such that

    A(u,z)+B(z,p)=λ(u,z),zX, (2.2)
    B(u,ϱ)=0,ϱW, (2.3)

    where

    A(u,z)=(u,z)+(Au,z),B(z,ϱ)=(divz,ϱ).

    Let Th={τ} be a regular simplex partition of Ω with the mesh diameter h=max where h_{\tau} is the diameter of element \tau . We use \mathcal{E}_{h}^{i} and \mathcal{E}_{h}^{b} to denote the set of interior faces (edges) and the set of faces (edges) on \partial\Omega , respectively. \mathcal{E}_{h} = \mathcal{E}_{h}^{i}\cup\mathcal{E}_{h}^{b} . We use h_{F} to denote the measure of F\in \mathcal{E}_{h} . We denote by (\cdot, \cdot)_{\tau} and (\cdot, \cdot)_{F} the inner product in L^{2}(\tau) and L^{2}(F) , respectively. We use \omega(\tau) to represent the union of all elements which share at least one edge (face) with \tau and use \omega(F) to represent the union of the elements having in common with F .

    The broken Sobolev space is defined by

    H^{1}(\Omega, \mathcal{T}_{h}) = \{z\in L^{2}(\Omega): z|_{\tau}\in H^{1}(\tau),\; \forall \tau\in\mathcal{T}_{h}\}.

    For any F\in\mathcal{E}_{h}^{i} , there are two simplices \tau^{+} and \tau^{-} such that F = \tau^{+}\cap\tau^{-} (e.g. see [14]). Let \mathbf{n}^{+} be the unit normal of F pointing from \tau^{+} to \tau^{-} and let \mathbf{n}^{-} = -\mathbf{n}^{+} .

    For any \varphi\in H^{1}(\Omega, \mathcal{T}_{h}) we denote its jump and mean on F\in\mathcal{E}_{h}^{i} by [\![\varphi]\!] = \varphi^{+}\mathbf{n}^{+}+\varphi^{-}\mathbf{n}^{-} and \mathbf{\{} \pmb{\mathsf{ φ}}\mathbf{\}} = \frac{1}{2}(\varphi^{+}+\varphi^{-}) , respectively, where \varphi^{\pm} = \varphi|_{\tau^{\pm}} . For \boldsymbol{\varphi}\in H^{1}(\Omega, \mathcal{T}_{h})^{d} we denote by [\![\boldsymbol{\varphi}]\!] = \boldsymbol{\varphi}^{+}\cdot\mathbf{n}^{+}+\boldsymbol{\varphi}^{-}\cdot\mathbf{n}^{-} the jump and \{\boldsymbol{\varphi}\} = \frac{1}{2}(\boldsymbol{\varphi}^{+}+\boldsymbol{\varphi}^{-}) the mean of \boldsymbol{\varphi} on F\in\mathcal{E}_{h}^{i} . We also denote by [\![\underline{\boldsymbol{\varphi}}]\!] = \boldsymbol{\varphi}^{+}\otimes \mathbf{n}^{+}+\boldsymbol{\varphi}^{-}\otimes \mathbf{n}^{-} the full jump of \boldsymbol{\varphi} on F\in\mathcal{E}_{h}^{i} , where \boldsymbol{\varphi}\otimes \mathbf{n} = [\varphi_{i}n_{j}]_{1\leq i, j\leq d} , \boldsymbol{\varphi} = (\varphi_{i}), \mathbf{n} = (n_{j}) . For tensors \boldsymbol{\chi}\in H^{1}(\Omega, \mathcal{T}_{h})^{d\times d} we denote by [\![\boldsymbol{\chi}]\!] = \boldsymbol{\chi}^{+}\mathbf{n}^{+}+\boldsymbol{\chi}^{-}\mathbf{n}^{-} the jump and \{\boldsymbol{\chi}\} = \frac{1}{2}(\boldsymbol{\chi}^{+}+\boldsymbol{\chi}^{-}) the mean on F\in\mathcal{E}_{h}^{i} .

    For the sake of simplicity, when F\in\mathcal{E}_{h}^{b} by modifying the above definitions appropriately, we obtain the jump and the mean on \partial\Omega . That is to say, we modify \varphi^{-} = 0 (similarly, \boldsymbol{\varphi}^{-} = 0 and \boldsymbol{\chi}^{-} = 0 ) to obtain the definition of jump on \partial\Omega and modify \varphi^{-} = \varphi^{+} (similarly, \boldsymbol{\varphi}^{-} = \boldsymbol{\varphi}^{+} and \boldsymbol{\chi}^{-} = \boldsymbol{\chi}^{+} ) to obtain the definition of mean on \partial\Omega .

    The discrete velocity and pressure spaces are defined as follows (see [7]):

    \begin{align*} &\mathbb{X}_{h} = \{\mathbf{z}_{h}\in L^{2}(\Omega)^{d}: \mathbf{z}_{h}|_{\tau}\in\mathbb{P}_{k}(\tau)^{d},\; \forall \tau \in\mathcal{T}_{h}\},\\ &\mathbb{W}_{h} = \{\varrho_{h}\in \mathbb{W}:\varrho_{h}|_{\tau}\in \mathbb{P}_{k-1}(\tau),\; \forall \tau \in\mathcal{T}_{h}\}, \end{align*}

    where \mathbb{P}_{k}(\tau) is the space of polynomials of degree less than or equal to k\geq 1 on \tau .

    The DGFEM for the problem (2.1) is to find (\lambda_{h}, \mathbf{u}_{h}, p_{h})\in \mathbb{R}\times\mathbb{X}_{h}\times \mathbb{W}_{h} , \|\mathbf{u}_{h}\|_{0} = 1 such that

    \begin{align} \mathbb{A}_{h}(\mathbf{u}_{h},\mathbf{z}_{h})+\mathbb{B}_{h}(\mathbf{z}_{h},p_{h})& = \lambda_{h}(\mathbf{u}_{h},\mathbf{z}_{h}), \; \; \forall \mathbf{z}_{h}\in\mathbb{X}_{h}, \end{align} (2.4)
    \begin{align} \mathbb{B}_{h}(\mathbf{u}_{h},\varrho_{h})& = 0,\; \; \; \forall \varrho_{h}\in \mathbb{W}_{h}, \end{align} (2.5)

    where

    \begin{align} \mathbb{A}_{h}(\mathbf{u}_{h},\mathbf{z}_{h})& = \sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\nabla\mathbf{u}_{h}:\nabla\mathbf{z}_{h}dx+ \sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}A\mathbf{u}_{h}\cdot \mathbf{z}_{h}dx -\sum\limits_{F\in\mathcal{E}_{h}}\int_{F}\{\nabla\mathbf{u}_{h}\}:[\![\underline{\mathbf{z}_{h}}]\!]ds\\ &\; \; \; -\sum\limits_{F\in\mathcal{E}_{h}}\int_{F}\{\nabla\mathbf{z}_{h}\}:[\![\underline{\mathbf{u}_{h}}]\!]ds +\sum\limits_{F\in\mathcal{E}_{h}}\int_{F}\frac{\gamma}{h_{F}}[\![\underline{\mathbf{u}_{h}}]\!]:[\![\underline{\mathbf{z}_{h}}]\!]ds, \end{align} (2.6)
    \begin{align} \mathbb{B}_{h}(\mathbf{z}_{h},\varrho_{h})& = -\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\varrho_{h}div\mathbf{z}_{h}dx+\sum\limits_{F\in\mathcal{E}_{h}}\int_{F}\{\varrho_{h}\}[\![\mathbf{z}_{h}]\!]ds. \end{align} (2.7)

    Here, \gamma/h_{F} is the interior penalty parameter. We select \gamma = C_{1}k^{2} with C_{1} = 10 in the actual numerical implementations from Remark 2.1 in [26].

    Define the DG-norm as follows:

    \begin{eqnarray} &&\|\mathbf{z}_{h}\|_{h}^{2} = \sum\limits_{\tau\in\mathcal{T}_{h}}\| \mathbf{z}_{h}\|_{1,\tau}^{2}+\sum\limits_{F\in\mathcal{E}_{h}}\int\limits_{F}\frac{\gamma}{h_{F}}[\![\underline{\mathbf{z}_{h}}]\!]^{2}ds,\; \; on\; \; \mathbb{X}_{h}+\mathbb{X}; \end{eqnarray} (2.8)
    \begin{eqnarray} &&|||\mathbf{z_{h}}|||^{2} = \|\mathbf{z}_{h}\|_{h}^{2}+\sum\limits_{F\in \mathcal{E}_{h}}\int_{F}\frac{h_{F}}{\gamma}|\nabla \mathbf{z}_{h}|^{2}ds,\; \; on\; \; \mathbb{X}_{h}+H^{1+s}(\Omega)^{d} \; (s > \frac{1}{2}). \end{eqnarray} (2.9)

    Note that \|\cdot\|_{h} is equivalent to |||\cdot||| on \mathbb{X}_{h} .

    From [27] we know that the continuity and coercivity properties hold:

    \begin{eqnarray*} &&|\mathbb{A}_{h}(\mathbf{u}_{h},\mathbf{z}_{h})|\lesssim |||\mathbf{u}_{h}|||\; |||\mathbf{z}_{h}|||,\; \; \; \forall\; \mathbf{u}_{h}, \mathbf{z}_{h}\in \mathbb{X}_{h}+H^{1+s}(\Omega)^{d}\; (s > \frac{1}{2}), \\ && \|\mathbf{u}_{h}\|^{2}_{h}\lesssim \mathbb{A}_{h}(\mathbf{u}_{h},\mathbf{u}_{h}), \; \; \; \forall \mathbf{u}_{h} \in \mathbb{X}_{h}. \end{eqnarray*}

    We consider the boundary problem : Given \boldsymbol{g}\in (L^{2}(\Omega))^{d} ,

    \begin{equation} \begin{cases} -\Delta \mathbf{u}^{g}+A\mathbf{u}^{g}+\nabla p^{g} = \boldsymbol{g},\; \; \; in\; \; \Omega,\\ div \mathbf{u}^{g} = 0,\; \; \; in \; \; \Omega,\\ \mathbf{u}^{g} = 0,\; \; \; on\; \; \partial\Omega. \end{cases} \end{equation} (2.10)

    From the Lax-Milgram theorem we have the existence and uniqueness of the velocity \mathbf{u} in the space \mathbf{Z} = \{\mathbf{z\in\mathbb{X}}: b(\mathbf{z}, \varrho) = 0, \forall \varrho\in \mathbb{W}\} . From the well-known inf-sup condition (see [28]):

    \begin{align*} \beta\|\varrho\|_{L^{2}(\Omega)}\leq \sup\limits_{\mathbf{z}\in \mathbb{X}, \mathbf{z}\neq 0}\frac{\mathbb{B}(\mathbf{z},\varrho)}{\|\mathbf{z}\|_{\mathbb{X}}},\; \forall\; \varrho\in \mathbb{W}, \end{align*}

    the stability of the pressure holds.

    The weak formulation of (2.10) reads: Find (\mathbf{u}^{g}, p^{g})\in \mathbb{X}\times \mathbb{W} such that

    \begin{align} \mathbb{A}(\mathbf{u}^{g},\mathbf{z})+\mathbb{B}(\mathbf{z},p^{g})& = (\boldsymbol{g},\mathbf{z}),\; \; \; \; \forall\mathbf{z}\in\mathbb{X}, \end{align} (2.11)
    \begin{align} \mathbb{B}(\mathbf{u}^{g},\varrho)& = 0,\; \; \; \; \forall \varrho\in \mathbb{W}, \end{align} (2.12)

    and its DGFEM form reads: Find (\mathbf{u}^{g}_{h}, p^{g}_{h})\in \mathbb{X}_{h}\times \mathbb{W}_{h} such that

    \begin{align} \mathbb{A}_{h}(\mathbf{u}^{g}_{h},\mathbf{z}_{h})+\mathbb{B}_{h}(\mathbf{z}_{h},p^{g}_{h})& = (\boldsymbol{g},\mathbf{z}_{h}), \; \; \; \; \forall\mathbf{z}_{h}\in\mathbb{X}_{h}, \end{align} (2.13)
    \begin{align} \mathbb{B}_{h}(\mathbf{u}^{g}_{h},\varrho_{h})& = 0,\; \; \; \; \forall \varrho_{h}\in \mathbb{W}_{h}. \end{align} (2.14)

    We assume that the following regularity is valid: For any \boldsymbol{g}\in (L^{2}(\Omega))^{d}\; (d = 2, 3) there exists (\mathbf{u}^{g}, p^{g})\in (H^{1+r}(\Omega)^{d}\times H^{r}(\Omega))\cap (W^{2, p}(\Omega)^{d}\times W^{1, p}(\Omega))\; (\frac{1}{2} < r\leq 1, \; p > \frac{2d}{d+1}) satisfying (2.10) and

    \begin{eqnarray} \|\mathbf{u}^{g}\|_{1+r}+\|p^{g}\|_{r}\leq C\|\boldsymbol{g}\|_{0}, \end{eqnarray} (2.15)

    where C is a positive constant independent of \boldsymbol{g} .

    From Lemma 6.5 in [27] we can obtain the consistency of the DGFEM that is to say when (\mathbf{u}^{g}, p^{g}) is the solution of the boundary problem (2.10), there hold the following equations:

    \begin{align} \mathbb{A}_{h}(\mathbf{u}^{g},\mathbf{z}_{h})+\mathbb{B}_{h}(\mathbf{z}_{h},p^{g})& = (\boldsymbol{g},\mathbf{z}_{h}), \; \; \forall \mathbf{z}_{h}\in\mathbb{X}_{h}, \end{align} (2.16)
    \begin{align} \mathbb{B}_{h}(\mathbf{u}^{g},\varrho_{h})& = 0,\; \; \; \forall \varrho_{h}\in \mathbb{W}_{h}. \end{align} (2.17)

    From (2.13), (2.14), (2.16) and (2.17) we have

    \begin{align} \mathbb{A}_{h}(\mathbf{u}^{g}-\mathbf{u}^{g}_{h},\mathbf{z}_{h})+\mathbb{B}_{h}(\mathbf{z}_{h}, p^{g}-p^{g}_{h})& = 0,\; \; \; \forall\; \mathbf{z}_{h}\in \mathbb{X}_{h}, \end{align} (2.18)
    \begin{align} \mathbb{B}_{h}(\mathbf{u}^{g}-\mathbf{u}^{g}_{h},\varrho_{h})& = 0,\; \; \; \forall \; \varrho_{h}\in \mathbb{W}_{h}. \end{align} (2.19)

    Badia et al. [14], Cockburn et al. [29], Hansbo et al. [30] and Schötzau et al. [25] proved that (2.13) and (2.14) are well defined and gave the a priori error estimate. From [14] we obtain the following lemma.

    Lemma 2.1. Assume that (\mathbf{u}^{g}, p^{g})\in H^{1+s}(\Omega)^{d}\times H^{s}(\Omega) ( r\leq s\leq k ) with \boldsymbol{g}\in H^{l}(\Omega)^{d} ( 0\leq l\leq k+1 ). Then,

    \begin{eqnarray} \|\mathbf{u}^{g}-\mathbf{u}^{g}_{h}\|_{h}+\|p^{g}-p^{g}_{h}\|_{0}\lesssim h^{s}(\|\mathbf{u}^{g}\|_{1+s}+\|p^{g}\|_{s})+h^{1+l}\|\boldsymbol{g}\|_{l}. \end{eqnarray} (2.20)

    Proof. Since the bilinear form \mathbb{A}(\cdot, \cdot) is coercive on \mathbb{X} , \mathbb{A}_{h}(\cdot, \cdot) is also coercive on \mathbb{X}_{h} . Using the proof method of Theorem 4.1 in [14] we can obtain (2.20).

    Let I_{h}: \mathbb{X}\cap C^{0}(\overline{\Omega})^{d}\to \mathbb{X}_{h}\cap\mathbb{X} be the conforming element interpolation operator and let \vartheta_{h}:\; H^{s}(\Omega)\rightarrow \mathbb{W}_{h} be the local L^{2} projection operator satisfying \vartheta_{h}p|_{\tau}\in \mathbb{P}_{k-1}(\tau) and

    \begin{eqnarray*} \int_{\tau}(p-\vartheta_{h}p)zdx = 0,\; \; \forall z\in \mathbb{P}_{k-1}(\tau),\; \; \; \forall \tau\in\mathcal{T}_{h}. \end{eqnarray*}

    We introduce the following auxiliary problem before estimating the error of velocity in the sense of L^{2} norm:

    \begin{align} \mathbb{A}(\boldsymbol{\omega},\mathbf{z})+\mathbb{B}(\mathbf{z},\varrho)& = (\mathbf{u}^{g}-\mathbf{u}^{g}_{h},\mathbf{z}),\; \; \; \; \forall\mathbf{z}\in\mathbb{X}, \end{align} (2.21)
    \begin{align} \mathbb{B}(\boldsymbol{\omega},v)& = 0,\; \; \; \; \forall v\in \mathbb{W}. \end{align} (2.22)

    Using (2.15), we have

    \begin{align} \|\boldsymbol{\omega}\|_{1+r}+\|\varrho\|_{r}\lesssim\|\mathbf{u}^{g}-\mathbf{u}^{g}_{h}\|_{0}. \end{align} (2.23)

    From Theorem 6.12 in [27] using the Nitsche's technique can prove the following lemma.

    Lemma 2.2. Suppose that the conditions of Lemma 2.1 and (2.15) hold. Then,

    \begin{align} \|\mathbf{u}^{g}-\mathbf{u}^{g}_{h}\|_{0}\lesssim h^{r}(|||\mathbf{u}^{g}-\mathbf{u}^{g}_{h}|||+\|p^{g}-p^{g}_{h}\|_{0}). \end{align} (2.24)

    Proof. From (2.18) and (2.19) we can derive

    \begin{align} \|\mathbf{u}^{g}-\mathbf{u}^{g}_{h}\|_{0}^{2}& = \mathbb{A}_{h}(\boldsymbol{\omega},\mathbf{u}^{g}-\mathbf{u}^{g}_{h})+\mathbb{B}_{h}(\mathbf{u}^{g}-\mathbf{u}^{g}_{h},\varrho)\\ & = \mathbb{A}_{h}(\mathbf{u}^{g}-\mathbf{u}^{g}_{h}, \boldsymbol{\omega}-I_{h}\boldsymbol{\omega})-\mathbb{A}_{h}(\mathbf{u}^{g}-\mathbf{u}^{g}_{h},I_{h}\boldsymbol{\omega}) +\mathbb{B}_{h}( \mathbf{u}^{g}-\mathbf{u}^{g}_{h}, \varrho-\vartheta_{h}\varrho) +\mathbb{B}_{h}( \mathbf{u}^{g}-\mathbf{u}^{g}_{h}, \vartheta_{h}\varrho)\\ & = \mathbb{A}_{h}(\mathbf{u}^{g}-\mathbf{u}^{g}_{h}, \boldsymbol{\omega}-I_{h}\boldsymbol{\omega})-\mathbb{B}_{h}(I_{h}\boldsymbol{\omega}, p^{g}-p^{g}_{h}) +\mathbb{B}_{h}(\mathbf{u}^{g}-\mathbf{u}^{g}_{h},\varrho-\vartheta_{h}\varrho) \\ &\equiv E_{1}+E_{2}+E_{3}. \end{align} (2.25)

    Using the continuity of \mathbb{A}_{h}(\cdot, \cdot) and the approximation property of I_{h}\boldsymbol{\omega} we obtain

    \begin{align*} |E_{1}|\lesssim|||\mathbf{u}^{g}-\mathbf{u}^{g}_{h}|||\; |||\boldsymbol{\omega}-I_{h}\boldsymbol{\omega}|||\lesssim h^{r}|||\mathbf{u}^{g}-\mathbf{u}^{g}_{h}|||\; \|\boldsymbol{\omega}\|_{1+r}. \end{align*}

    Since div \boldsymbol{\omega} = 0 , \boldsymbol{\omega}\in[H_{0}^{1}(\Omega)]^{d} , [\boldsymbol{\omega}] = 0 and [I_{h}\boldsymbol{\omega}] = 0, \forall F\in \mathcal{E}_{h} , we have

    \begin{align*} |E_{2}|& = |-\mathbb{B}_{h}(I_{h}\boldsymbol{\omega}, p^{g}-p^{g}_{h})| = |\mathbb{B}_{h}(\boldsymbol{\omega}-I_{h}\boldsymbol{\omega}, p^{g}-p^{g}_{h})|\\ & = |-( p^{g}-p_{h}^{g}, div(\boldsymbol{\omega}-I_{h}\boldsymbol{\omega}))+\sum\limits_{F\in \mathcal{E}_{h}}\int_{F}\{p^{g}-p^{g}_{h}\}[\![\boldsymbol{\omega}-I_{h}\boldsymbol{\omega}]\!]ds|\\ &\lesssim h^{r}\|p^{g}-p_{h}^{g}\|_{0}\|\boldsymbol{\omega}\|_{1+r}. \end{align*}

    Using the approximation property of \vartheta_{h}\varrho we can get

    \begin{align*} |E_{3}|&\leq \|\varrho-\vartheta_{h}\varrho\|_{0}\|div(\mathbf{u}^{g}-\mathbf{u}^{g}_{h})\|_{0}+\left(\sum\limits_{F\in\mathcal{E}_{h}}\int_{F}\frac{h_{F}}{\gamma}|\{\varrho-\vartheta_{h}\varrho\}|^{2}ds\right)^{\frac{1}{2}} \|\mathbf{u}^{g}-\mathbf{u}^{g}_{h}\|_{h}\\ &\lesssim h^{r}\|\mathbf{u}^{g}-\mathbf{u}^{g}_{h}\|_{h}\|\varrho\|_{r}. \end{align*}

    Substituting E_{1}, E_{2}, E_{3} into (2.25) and using (2.23) we obtain (2.24). The proof is completed.

    From the inf-sup condition and [14] we know that (2.11) and (2.12) are uniquely solvable and stable. Then, we define

    \begin{eqnarray*} && \mathbb{T}: L^{2}(\Omega)^{d}\rightarrow \mathbb{X}, \; \; \; \; \; \; \; \mathbb{T}\boldsymbol{g} = \mathbf{u}^{g},\\ && \mathbb{S}: L^{2}(\Omega)^{d}\rightarrow \mathbb{W}, \; \; \; \; \; \; \; \mathbb{S}\boldsymbol{g} = p^{g}, \end{eqnarray*}

    and it is valid that

    \begin{align} \|\mathbb{T}\boldsymbol{g}\|_{1}+\|\mathbb{S}\boldsymbol{g}\|_{0}\lesssim\|\boldsymbol{g}\|_{0}.\; \; \; \end{align} (2.26)

    From [14] we also know that (2.13) and (2.14) are uniquely solvable and stable and we define

    \begin{eqnarray*} && \mathbb{T}_{h}: L^{2}(\Omega)^{d}\rightarrow \mathbb{X}_{h}, \; \; \; \; \mathbb{T}_{h}\boldsymbol{g} = \mathbf{u}^{g}_{h},\\ && \mathbb{S}_{h}: L^{2}(\Omega)^{d}\rightarrow \mathbb{W}_{h}, \; \; \; \; \mathbb{S}_{h}\boldsymbol{g} = p^{g}_{h}. \end{eqnarray*}

    Hence,

    \begin{align} ||\mathbb{T}_{h}\boldsymbol{g}||_{h}+\|\mathbb{S}_{h}\boldsymbol{g}\|_{0}\lesssim\|\boldsymbol{g}\|_{0}.\; \; \; \end{align} (2.27)

    Thus, (2.2), (2.3) and (2.4), (2.5) have the following equivalent operator forms, respectively:

    \begin{align} & \lambda \mathbb{T} \mathbf{u} = \mathbf{u},\; \; \; \; \; \; \; \; \; \; \mathbb{S} (\lambda \mathbf{u}) = p, \end{align} (2.28)
    \begin{align} & \lambda_{h} \mathbb{T}_{h} \mathbf{u}_{h} = \mathbf{u}_{h},\; \; \; \; \mathbb{S}_{h} (\lambda_{h} \mathbf{u}_{h}) = p_{h}. \end{align} (2.29)

    Next, we will derive the error estimates for the eigenvalue problem.

    From (2.24), (2.20) and (2.15) we have

    \begin{eqnarray} \|\mathbb{T}_{h}-\mathbb{T}\|_{0}\rightarrow 0,\; \; (h\rightarrow 0). \end{eqnarray} (2.30)

    Thus, we can obtain the following Lemma 2.3 (see Lemma 2.3 in [31]) from the Babuška-Osborn spectral approximation theory [32,33].

    From (2.8) and (2.9) we know that |||\cdot||| is a norm stronger than \|\cdot\|_{h} , i.e., \|z\|_{h}\lesssim|||z||| . Additionally, we have

    \begin{eqnarray} |||\mathbf{u}-\mathbf{u}_{h}|||^{2} \lesssim\|\mathbf{u}-\mathbf{u}_{h}\|_{h}^{2}+\sum\limits_{\tau\in\mathcal{T}_{h}}h_{\tau}^{2r}|\mathbf{u}-I_{h}\mathbf{u}|_{1+r,\tau}^{2}. \end{eqnarray} (2.31)

    To show the validity of (2.31), using the trace theorem on the reference element and the scaling argument we have for any \tau\in \mathcal{T}_{h} that

    \begin{align} \|w\|_{0,\partial\tau}\lesssim h_{\tau}^{-\frac{1}{2}}\|w\|_{0,\tau}+h_{\tau}^{r-\frac{1}{2}}|w|_{r,\tau},\; \; \forall w\in H^{r}(\tau),\; r\in (\frac{1}{2},1], \end{align} (2.32)

    and from the inverse inequality and the interpolation estimate and taking w = \nabla (\mathbf{u}-I_{h}\mathbf{u}) in (2.32) we deduce

    \begin{align*} \sum \limits_{F\in\mathcal{E}_{h}}h_{F}\|\nabla (\mathbf{u}-\mathbf{u}_{h})\|^{2}_{0,F} &\lesssim\sum \limits_{F\in\mathcal{E}_{h}}h_{F}\|\nabla (I_{h}\mathbf{u}-\mathbf{u}_{h})\|^{2}_{0,F} +\sum \limits_{F\in\mathcal{E}_{h}}h_{F}\|\nabla (\mathbf{u}-I_{h}\mathbf{u})\|^{2}_{0,F}\nonumber\\ &\lesssim\sum \limits_{\tau\in\mathcal{T}_{h}}\|\nabla (I_{h}\mathbf{u}-\mathbf{u}_{h})\|^{2}_{0,\tau} +\sum\limits_{\tau\in\mathcal{T}_{h}}(\|\nabla (\mathbf{u}-I_{h}\mathbf{u})\|^{2}_{0,\tau}+h_{\tau}^{2r}|\mathbf{u}-I_{h}\mathbf{u}|_{1+r,\tau}^{2}) \nonumber\\ &\lesssim \|\mathbf{u}-\mathbf{u}_{h}\|_{h}^{2}+\sum\limits_{\tau\in\mathcal{T}_{h}}h_{\tau}^{2r}|\mathbf{u}-I_{h}\mathbf{u}|_{1+r,\tau}^{2}. \end{align*}

    By the above inequality and (2.9) we obtain (2.31).

    Theorem 2.1. Let (\lambda, \mathbf{u}, p) and (\lambda_{h}, \mathbf{u}_{h}, p_{h}) be the j th eigenpair of (2.2), (2.3) and (2.4), (2.5), respectively. Assume that the regularity estimate (2.15) is valid and (\mathbf{u}, p)\in H^{1+s}(\Omega)^{d}\times H^{s}(\Omega) for some s\in[r, k] . Then,

    \begin{eqnarray} &&\|\mathbf{u}_{h}-\mathbf{u}\|_{0}\lesssim h^{r}(\|\mathbf{u}-\mathbf{u}_{h}\|_{h}+\|p-p_{h}\|_{0}), \end{eqnarray} (2.33)
    \begin{eqnarray} &&|\lambda_{h}-\lambda|\lesssim h^{2s}, \end{eqnarray} (2.34)
    \begin{eqnarray} &&\|\mathbf{u}-\mathbf{u}_{h}\|_{h}+\|p-p_{h}\|_{0}\lesssim h^{s}(\|\mathbf{u}\|_{1+s}+\|p\|_{s}) . \end{eqnarray} (2.35)

    Proof. In (2.11)–(2.14), we take \boldsymbol{g} = \lambda\mathbf{u} then we obtain \mathbf{u}^{g} = \lambda \mathbb{T}\mathbf{u} , \mathbf{u}^{g}_{h} = \lambda \mathbb{T}_{h}\mathbf{u} , p^{g} = \lambda \mathbb{S}\mathbf{u} and p^{g}_{h} = \lambda \mathbb{S}_{h}\mathbf{u} . Hence, using (2.20) we deduce

    \begin{eqnarray} \|\lambda \mathbb{T}\mathbf{u}-\lambda \mathbb{T}_{h}\mathbf{u}\|_{h}+\|\lambda \mathbb{S}\mathbf{u}-\lambda \mathbb{S}_{h}\mathbf{u}\|_{0} \lesssim h^{s}(\|\mathbf{u}\|_{1+s}+\|p\|_{s}). \end{eqnarray} (2.36)

    By using (2.16), (2.18), (2.19) and (2.36) we obtain

    \begin{align} ((\mathbb{T}-\mathbb{T}_{h})\mathbf{u}, \mathbf{u})& = \mathbb{A}_{h}((\mathbb{T}-\mathbb{T}_{h})\mathbf{u}, \mathbb{T}\mathbf{u})+\mathbb{B}_{h}((\mathbb{T}-\mathbb{T}_{h})\mathbf{u}, \mathbb{S}\mathbf{u})\\ & = \mathbb{A}_{h}(\mathbb{T}\mathbf{u}-\mathbb{T}_{h}\mathbf{u}, \mathbb{T}\mathbf{u}-\mathbb{T}_{h}\mathbf{u})+\mathbb{A}_{h}(\mathbb{T}\mathbf{u}-\mathbb{T}_{h}\mathbf{u}, \mathbb{T}_{h}\mathbf{u})\\ &\; \; \; +2\mathbb{B}_{h}(\mathbb{T}\mathbf{u}-\mathbb{T}_{h}\mathbf{u}, \mathbb{S}\mathbf{u}-\mathbb{S}_{h}\mathbf{u})+\mathbb{B}_{h}(\mathbb{T}\mathbf{u}-\mathbb{T}_{h}\mathbf{u}, 2\mathbb{S}_{h}\mathbf{u})-\mathbb{B}_{h}(\mathbb{T}\mathbf{u}-\mathbb{T}_{h}\mathbf{u}, \mathbb{S}\mathbf{u})\\ & = \mathbb{A}_{h}(\mathbb{T}\mathbf{u}-\mathbb{T}_{h}\mathbf{u}, \mathbb{T}\mathbf{u}-\mathbb{T}_{h}\mathbf{u})+2\mathbb{B}_{h}(\mathbb{T}\mathbf{u}-\mathbb{T}_{h}\mathbf{u}, \mathbb{S}\mathbf{u}-\mathbb{S}_{h}\mathbf{u}) \\ &\; \; \; + (\mathbb{A}_{h}(\mathbb{T}\mathbf{u}-\mathbb{T}_{h}\mathbf{u}, \mathbb{T}_{h}\mathbf{u})+\mathbb{B}_{h}(\mathbb{T}_{h}\mathbf{u}, \mathbb{S}\mathbf{u}-\mathbb{S}_{h}\mathbf{u})) \\ &\; \; \; +\mathbb{B}_{h}(2\mathbb{T}\mathbf{u}-\mathbb{T}_{h}\mathbf{u}, \mathbb{S}_{h}\mathbf{u})-\mathbb{B}_{h}(\mathbb{T}\mathbf{u}, \mathbb{S}\mathbf{u}) \\ & = \mathbb{A}_{h}(\mathbb{T}\mathbf{u}-\mathbb{T}_{h}\mathbf{u}, \mathbb{T}\mathbf{u}-\mathbb{T}_{h}\mathbf{u})+2\mathbb{B}_{h}(\mathbb{T}\mathbf{u}-\mathbb{T}_{h}\mathbf{u}, \mathbb{S}\mathbf{u}-\mathbb{S}_{h}\mathbf{u})\\ &\lesssim h^{2s}(\|\mathbf{u}\|_{1+s}+\|p\|_{s})^{2}. \end{align} (2.37)

    From Lemma 2.3 in [31] we know that

    \begin{align} \|\mathbf{u}_{h}-\mathbf{u}\|_{0}&\lesssim\|(\mathbb{T}-\mathbb{T}_{h})\mathbf{u}\|_{0}, \end{align} (2.38)
    \begin{align} |\lambda_{h}-\lambda|&\lesssim\lambda^{2}((\mathbb{T}-\mathbb{T}_{h})\mathbf{u}, \mathbf{u})+\|(\mathbb{T}-\mathbb{T}_{h})\mathbf{u}\|_{0}^{2}. \end{align} (2.39)

    Substituting (2.37) and (2.24) into (2.39) we obtain (2.34).

    Applying the triangle inequality and (2.27) we get

    \begin{align} &|\; ||\mathbf{u}-\mathbf{u}_{h}||_{h}-||\lambda \mathbb{T}\mathbf{u}-\lambda \mathbb{T}_{h}\mathbf{u}||_{h}\; | \leq \|\lambda_{h}\mathbb{T}_{h}\mathbf{u}_{h}-\lambda \mathbb{T}_{h}\mathbf{u}\|_{h} \lesssim \|\lambda_{h}\mathbf{u}_{h}-\lambda\mathbf{u}\|_{0}, \end{align} (2.40)
    \begin{align} &|\; \|p-p_{h}\|_{0}-\|\lambda \mathbb{S}\mathbf{u}-\lambda \mathbb{S}_{h}\mathbf{u}\|_{0}\; | \leq \|\lambda_{h}\mathbb{S}_{h}\mathbf{u}_{h}-\lambda \mathbb{S}_{h}\mathbf{u}\|_{0}\lesssim \|\lambda_{h}\mathbf{u}_{h}-\lambda\mathbf{u}\|_{0}. \end{align} (2.41)

    From (2.24), (2.38) and (2.39) we deduce

    \begin{align} \|\lambda_{h}\mathbf{u}_{h}-\lambda\mathbf{u}\|_{0}&\lesssim |\lambda_{h}-\lambda|+\|\mathbf{u}_{h}-\mathbf{u}\|_{0}\lesssim \|\lambda \mathbb{T}\mathbf{u}-\lambda \mathbb{T}_{h}\mathbf{u}\|_{0}\\ &\lesssim h^{r}(\|\lambda \mathbb{T}\mathbf{u}-\lambda \mathbb{T}_{h}\mathbf{u}\|_{h}+\|\lambda \mathbb{S}\mathbf{u}-\lambda \mathbb{S}_{h}\mathbf{u}\|_{0}). \end{align} (2.42)

    Then, from (2.40)–(2.42) we obtain

    \begin{align} ||\mathbf{u}-\mathbf{u}_{h}||_{h}+\|p-p_{h}\|_{0}\simeq ||\lambda \mathbb{T}\mathbf{u}-\lambda \mathbb{T}_{h}\mathbf{u}||_{h}+\|\lambda \mathbb{S}\mathbf{u}-\lambda \mathbb{S}_{h}\mathbf{u}\|_{0}. \end{align} (2.43)

    Thus, we get (2.33).

    Combining (2.43) with (2.36) we get (2.35).

    Let (\lambda_{h}, \mathbf{u}_{h}, p_{h})\in \mathbb{R}^{+}\times\mathbb{X}_{h}\times \mathbb{W}_{h} be an approximate eigenpair. First, for each element \tau\in\mathcal{T}_{h} we introduce the residuals:

    \begin{align*} &\eta^{2}_{R_{\tau}} = h_{\tau}^{2}\|\lambda_{h}\mathbf{u}_{h}+\Delta\mathbf{u}_{h}-A\mathbf{u}_{h}-\nabla p_{h}\|^{2}_{0,\tau}+\|div\mathbf{u}_{h}\|_{0,\tau}^{2},\\ &\eta^{2}_{F_{\tau}} = \frac{1}{2}\sum\limits_{F\subset\partial\tau\setminus\partial\Omega}h_{F}\|[\![(p_{h}\mathbf{I}- \nabla\mathbf{u}_{h})]\!]\|^{2}_{0,F}, \end{align*}

    where \mathbf{I} denotes the d\times d (d = 2, 3) identity matrix. Next, we introduce the following estimator \eta_{J_{\tau}} to measure the jump of the approximate solution \mathbf{u}_{h} :

    \begin{align*} \eta^{2}_{J_{\tau}}& = \sum\limits_{F\subset\partial\tau, F\in\mathcal{E}^{i}_{h}}\gamma h_{F}^{-1}||[\![\underline{\mathbf{u}_{h}}]\!]||_{0,F}^{2}+ \sum\limits_{F\subset\partial\tau, F\in\mathcal{E}^{b}_{h}}\gamma h_{F}^{-1}||\mathbf{u}_{h}\otimes\mathbf{n}||_{0,F}^{2}. \end{align*}

    The local error indictor is defined as

    \begin{align*} \eta^{2}_{\tau} = \eta^{2}_{R_{\tau}}+\eta^{2}_{F_{\tau}}+\eta^{2}_{J_{\tau}}. \end{align*}

    Then, the global a posteriori error estimator is defined as

    \begin{align*} \eta_{h} = (\sum\limits_{\tau\in\mathcal{T}_{h}}\eta^{2}_{\tau})^{\frac{1}{2}}. \end{align*}

    We denote \theta_{\tau} = int \{\bigcup\limits_{\overline{\tau}_{i}\cap \overline{\tau}\not = \emptyset}\bar{\tau}_{i}, \tau_{i}\in\mathcal{T}_{h}\} for \tau\in\mathcal{T}_{h} and use \theta_{F} to represent the set of all elements which share at least one node with face F . We denote by \mathbf{z}^{I} the Scott-Zhang interpolation function [34], then \mathbf{z}^{I}\in \mathbb{X}\cap\mathbb{X}_{h} and

    \begin{eqnarray} &&\|\mathbf{z}-\mathbf{z}^{I}\|_{0,\tau}+h_{\tau}\|\nabla(\mathbf{z}-\mathbf{z}^{I})\|_{0,\tau}\lesssim h_{\tau}|\mathbf{z}|_{1,\theta_{\tau}},\; \; \; \; \; \; \; \forall \tau\in\mathcal{T}_{h}, \end{eqnarray} (3.1)
    \begin{eqnarray} &&\|\mathbf{z}-\mathbf{z}^{I}\|_{0,F} \lesssim h_{F}^{\frac{1}{2}}|\mathbf{z}|_{1,\theta_{F}},\; \; \; \; \; \forall F\subset\partial\tau. \end{eqnarray} (3.2)

    Denote

    \begin{eqnarray*} \underline{\sum}_{h} = \{\boldsymbol{\underline{o}}\in L^{2}(\Omega)^{d\times d}: \boldsymbol{\underline{o}}|_{\tau}\in \mathbb{P}_{k}(\tau)^{d\times d}, \tau\in\mathcal{T}_{h}\}. \end{eqnarray*}

    The lifting operator \mathcal{L}:\mathbb{X}(h)\rightarrow \underline{\sum}_{h} is defined by

    \begin{align} \int_{\Omega}\mathcal{L}(\mathbf{z}): \boldsymbol{\underline{o}}dx = \sum\limits_{F\in\mathcal{E}_{h}^{i}}\int_{F}[\![\underline{\mathbf{z}}]\!]:\boldsymbol{\{\underline{o}\}}ds,\; \; \forall \; \boldsymbol{\underline{o}}\in \underline{\sum}_{h}, \end{align} (3.3)

    and has the following property (see [24,25]):

    \begin{align} \|\mathcal{L}(\mathbf{z})\|_{0}^{2}\lesssim \sum\limits_{F\in\mathcal{E}_{h}^{i}}\|h_{F}^{-\frac{1}{2}}[\![\underline{\mathbf{z}}]\!]\|_{0,F}^{2},\; \; \; \forall \; \mathbf{z}\in \mathbb{X}+\mathbb{X}_{h}. \end{align} (3.4)

    Using the lifting operator, we define the following form:

    \begin{align} \widetilde{\mathbb{A}}_{h}(\cdot, \cdot): (\mathbb{X}+\mathbb{X}_{h})\times (\mathbb{X}+\mathbb{X}_{h})\rightarrow \mathbb{R} \end{align} (3.5)

    by

    \begin{align} \widetilde{\mathbb{A}}_{h}(\mathbf{w},\mathbf{z}) = &\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\nabla\mathbf{w}:\nabla\mathbf{z}dx +\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}A\mathbf{w}\cdot\mathbf{z}dx -\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathcal{L}(\mathbf{z}):\nabla\mathbf{w}dx\\ &\; \; \; -\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathcal{L}(\mathbf{w}):\nabla\mathbf{z}dx +\sum\limits_{F\in\mathcal{E}_{h}}\int_{F}\frac{\gamma}{h_{F}}[\![\underline{\mathbf{w}}]\!]:[\![\underline{\mathbf{z}}]\!]ds,\; \; \; \forall\; \mathbf{z}\in \mathbb{X}+\mathbb{X}_{h}. \end{align} (3.6)

    Note that \widetilde{\mathbb{A}}_{h} = \mathbb{A}_{h} on \mathbb{X}_{h}\times \mathbb{X}_{h} and \mathbb{A} = \widetilde{\mathbb{A}}_{h} on \mathbb{X}\times \mathbb{X} . The DGFEM presented in (2.4) and (2.5) is equivalent to finding (\lambda_{h}, \mathbf{u}_{h}, p_{h})\in \mathbb{R}^{+}\times \mathbb{X}_{h}\times \mathbb{W}_{h} and satisfying

    \begin{align} \widetilde{\mathbb{A}}_{h}(\mathbf{u}_{h}, \mathbf{z}_{h})+\mathbb{B}_{h}(\mathbf{z}_{h}, p_{h})& = \lambda_{h}(\mathbf{u}_{h}, \mathbf{z}_{h}), \; \; \; \forall \mathbf{z}_{h}\in\mathbb{X}_{h}, \\ \mathbb{B}_{h}(\mathbf{u}_{h},\varrho_{h})& = 0,\; \; \; \forall \varrho_{h}\in \mathbb{W}_{h}. \end{align} (3.7)

    Lemma 3.1. Let (\mathbf{u}^{g}, p^{g}) and (\mathbf{u}^{g}_{h}, p^{g}_{h}) be the solutions of (2.11), (2.12) and (2.13), (2.14), respectively. Then,

    \begin{align} \|\mathbf{u}^{g}-\mathbf{u}^{g}_{h}\|_{h}+\|p^{g}-p^{g}_{h}\|_{0}\backsimeq \sup\limits_{0\not = \mathbf{z}\in \mathbb{X}}\frac{|(\boldsymbol{g},\mathbf{z})-\widetilde{\mathbb{A}}_{h}(\mathbf{u}^{g}_{h}, \mathbf{z})-\mathbb{B}_{h}(\mathbf{z}, p^{g}_{h})|}{\|\mathbf{z}\|_{h}}+\inf\limits_{\mathbf{z}\in \mathbb{X}}\|\mathbf{u}^{g}_{h}-\mathbf{z}\|_{h}. \end{align} (3.8)

    Proof. For \forall \mathbf{\bar{u}}\in \mathbb{X} , from (2.11) we have

    \begin{eqnarray*} \widetilde{\mathbb{A}}_{h}(\mathbf{u}^{g}-\mathbf{\bar{u}}, \mathbf{u}^{g}-\mathbf{\bar{u}})& = &\widetilde{\mathbb{A}}_{h}(\mathbf{u}^{g}, \mathbf{u}^{g}-\mathbf{\bar{u}})-\widetilde{\mathbb{A}}_{h}(\mathbf{\bar{u}}, \mathbf{u}^{g}-\mathbf{\bar{u}})\nonumber\\ & = &(\boldsymbol{g}, \mathbf{u}^{g}-\mathbf{\bar{u}})-\mathbb{B}( \mathbf{u}^{g}-\mathbf{\bar{u}}, p^{g})-\widetilde{\mathbb{A}}_{h}(\mathbf{u}^{g}_{h}, \mathbf{u}^{g}-\mathbf{\bar{u}})+\widetilde{\mathbb{A}}_{h}(\mathbf{u}^{g}_{h}-\mathbf{\bar{u}}, \mathbf{u}^{g}-\mathbf{\bar{u}}). \end{eqnarray*}

    For \forall \mathbf{\bar{u}}\in \mathbb{X} , \overline{p}\in \mathbb{W} we have

    \begin{eqnarray*} \mathbb{B}_{h}(\mathbf{u}^{g}-\mathbf{\bar{u}}, p^{g}-\bar{p}) = \mathbb{B}_{h}(\mathbf{u}^{g}-\mathbf{\bar{u}}, p^{g})-\mathbb{B}_{h}(\mathbf{u}^{g}-\mathbf{\bar{u}}, p^{g}_{h})-\mathbb{B}_{h}(\mathbf{u}^{g}-\mathbf{\bar{u}},\bar{p}- p^{g}_{h}). \end{eqnarray*}

    Combining the above two equations and taking \mathbf{z} = \mathbf{u}^{g}-\mathbf{\bar{u}} we obtain

    \begin{align} &\|\mathbf{u}^{g}-\mathbf{\bar{u}}\|_{h}\|\mathbf{z}\|_{h}+\mathbb{B}_{h}(\mathbf{z}, p^{g}-\bar{p})\\ = &(\boldsymbol{g}, \mathbf{z})-\mathbb{B}_{h}(\mathbf{z}, p^{g})-\widetilde{\mathbb{A}}_{h}(\mathbf{u}^{g}_{h}, \mathbf{z})+\widetilde{\mathbb{A}}_{h}(\mathbf{u}^{g}_{h}-\mathbf{\bar{u}}, \mathbf{z})+ \mathbb{B}_{h}(\mathbf{z}, p^{g})-\mathbb{B}_{h}(\mathbf{z}, p^{g}_{h})-\mathbb{B}_{h}(\mathbf{z}, \bar{p}- p^{g}_{h})\\ = &(\boldsymbol{g}, \mathbf{z})-\widetilde{\mathbb{A}}_{h}(\mathbf{u}^{g}_{h}, \mathbf{z}) -\mathbb{B}_{h}(\mathbf{z}, p^{g}_{h})+\widetilde{\mathbb{A}}_{h}(\mathbf{u}^{g}_{h}-\mathbf{\bar{u}},\mathbf{z}) -\mathbb{B}_{h}(\mathbf{z},\bar{p}- p^{g}_{h}). \end{align} (3.9)

    From the well-known inf-sup condition we obtain

    \sup\limits_{\mathbf{z}\in \mathbb{X}}\frac{\mathbb{B}_{h}(\mathbf{z}, p^{g}-\bar{p})}{\|\mathbf{z}\|_{h}}\gtrsim\|p^{g}-\bar{p}\|_{0}.

    Dividing both sides of (3.9) by \|\mathbf{z}\|_{h} and taking supremum for \mathbf{z}\in \mathbb{X} we get

    \begin{align} \|\mathbf{u}^{g}-\mathbf{\bar{u}}\|_{h}+\| p^{g}-\bar{p}\|_{0}\lesssim\sup\limits_{\mathbf{z}\in\mathbb{X}}\frac{(\boldsymbol{g}, \mathbf{z})-\widetilde{\mathbb{A}}_{h}(\mathbf{u}^{g}_{h}, \mathbf{z}) -\mathbb{B}_{h}(\mathbf{z}, p^{g}_{h})}{\|\mathbf{z}\|_{h}} +\|\mathbf{u}^{g}_{h}-\mathbf{\bar{u}}\|_{h}+\|\bar{p}- p^{g}_{h}\|_{0},\; \forall (\mathbf{\bar{u}}, \bar{p})\in \mathbb{X}\times \mathbb{W}. \end{align} (3.10)

    Using the triangle inequality we obtain

    \begin{eqnarray} &&\|\mathbf{u}^{g}-\mathbf{u}^{g}_{h}\|_{h}+\| p^{g}-p^{g}_{h}\|_{0}\\ &\lesssim&\sup\limits_{\mathbf{z}\in\mathbb{X}}\frac{(\boldsymbol{g}, \mathbf{z})-\widetilde{\mathbb{A}}_{h}(\mathbf{u}^{g}_{h}, \mathbf{z}) -\mathbb{B}_{h}(\mathbf{z}, p_{h}^{g})}{\|\mathbf{z}\|_{h}}+\|\mathbf{u}^{g}_{h}-\mathbf{\bar{u}}\|_{h}+\|\bar{p}- p^{g}_{h}\|_{0},\; \; \; \forall (\mathbf{\bar{u}}, \bar{p})\in \mathbb{X}\times \mathbb{W}. \end{eqnarray} (3.11)

    Since (\mathbf{\bar{u}}, \bar{p}) is arbitrary and \inf\limits_{\bar{p}\in \mathbb{W}}\|\bar{p}- p^{g}_{h}\|_{0} = 0 , the part \lesssim in (3.8) is valid. The other part \gtrsim in (3.8) is obvious.

    Lemma 3.1 can be extended to the eigenvalue problem.

    Theorem 3.1. Let (\lambda, \mathbf{u}, p) and (\lambda_{h}, \mathbf{u}_{h}, p_{h}) be the j th eigenpair of (2.2), (2.3) and (2.4), (2.5), respectively. Then,

    \begin{align} \|\mathbf{u}-\mathbf{u}_{h}\|_{h}+\|p-p_{h}\|_{0}\backsimeq & \sup\limits_{0\not = \mathbf{z}\in \mathbb{X}}\frac{|\widetilde{\mathbb{A}}_{h}(\mathbf{u}-\mathbf{u}_{h},\mathbf{z})+\mathbb{B}_{h}(\mathbf{z},p-p_{h})|}{\|\mathbf{z}\|_{h}} +\inf\limits_{\mathbf{z}\in \mathbb{X}}\|\mathbf{u}_{h}-\mathbf{z}\|_{h}+\|\lambda_{h}\mathbf{u}_{h}-\lambda \mathbf{u}\|_{0}. \end{align} (3.12)

    Proof. Using (2.26) and (2.27) we can obtain

    \begin{eqnarray} &&\|\mathbf{u}-\mathbf{u}_{h}\|_{h}+\|p-p_{h}\|_{0}\\ & = & \|\lambda \mathbb{T}\mathbf{u}-\lambda_{h} \mathbb{T}\mathbf{u}_{h }+\lambda_{h} \mathbb{T}\mathbf{u}_{h }-\lambda_{h}\mathbb{T}_{h}\mathbf{u}_{h}\|_{h} +\|\lambda \mathbb{S}\mathbf{u}-\lambda_{h} \mathbb{S}\mathbf{u}_{h }+\lambda_{h} \mathbb{S}\mathbf{u}_{h }-\lambda_{h}\mathbb{S}_{h}\mathbf{u}_{h}\|_{h}\\ & \leq& \|\lambda_{h} \mathbb{T}\mathbf{u}_{h }-\lambda_{h}\mathbb{T}_{h}\mathbf{u}_{h}\|_{h} +\|\lambda_{h} \mathbb{S}\mathbf{u}_{h }-\lambda_{h}\mathbb{S}_{h}\mathbf{u}_{h}\|_{h} +\|\lambda \mathbf{u}-\lambda_{h} \mathbf{u}_{h }\|_{0}. \end{eqnarray} (3.13)

    In (2.11)–(2.14) we take \boldsymbol{g} = \lambda_{h}\mathbf{u}_{h} and obtain \mathbf{u}^{g} = \lambda_{h} \mathbb{T}\mathbf{u}_{h } , \mathbf{u}^{g}_{h} = \lambda_{h}\mathbb{T}_{h}\mathbf{u}_{h} , p^{g} = \lambda_{h}\mathbb{S}\mathbf{u}_{h } and p^{g}_{h} = \lambda_{h}\mathbb{S}_{h}\mathbf{u}_{h} . Therefore, from (3.8) we have

    \begin{align} &\; \; \; \|\lambda_{h} \mathbb{T}\mathbf{u}_{h }-\lambda_{h}\mathbb{T}_{h}\mathbf{u}_{h}\|_{h}+\|\lambda_{h} \mathbb{S}\mathbf{u}_{h }-\lambda_{h}\mathbb{S}_{h}\mathbf{u}_{h}\|_{0}\\ &\lesssim \sup\limits_{0\not = \mathbf{z}\in \mathbb{X}}\frac{|(\mathbf{ \pmb{\mathsf{ λ}}_{h}}\mathbf{u}_{h},\mathbf{z})-\widetilde{\mathbb{A}}_{h}(\lambda_{h}\mathbb{T}_{h}\mathbf{u}_{h}, \mathbf{z})-\mathbb{B}_{h}(\mathbf{z}, \mathbb{S}_{h}(\lambda_{h}\mathbf{u}_{h}) )|}{\|\mathbf{z}\|_{h}} +\inf\limits_{\mathbf{z}\in \mathbb{X}}\|\mathbf{u}_{h}-\mathbf{z}\|_{h}. \end{align} (3.14)

    From (2.11) with \boldsymbol{g} = \lambda_{h}\mathbf{u}_{h} , (2.26) and (2.27) we deduce

    \begin{eqnarray} &&\; \; \; |(\mathbf{ \pmb{\mathsf{ λ}}_{h}u_{h}},\mathbf{z})-\widetilde{\mathbb{A}}_{h}(\lambda_{h}\mathbb{T}_{h}\mathbf{u}_{h}, \mathbf{z})-\mathbb{B}_{h}(\mathbf{z}, \mathbb{S}_{h}(\lambda_{h}\mathbf{u}_{h}) )|\\ && = |\widetilde{\mathbb{A}}_{h}(\lambda_{h}\mathbb{T}\mathbf{u}_{h}, \mathbf{z})+\mathbb{B}_{h}(\mathbf{z}, \mathbb{S}(\lambda_{h}\mathbf{u}_{h}))-\widetilde{\mathbb{A}}_{h}(\lambda_{h}\mathbb{T}_{h}\mathbf{u}_{h}, \mathbf{z})-\mathbb{B}_{h}(\mathbf{z}, \mathbb{S}_{h}(\lambda_{h}\mathbf{u}_{h}))|\\ && = |\widetilde{\mathbb{A}}_{h}(\lambda_{h}\mathbb{T}\mathbf{u}_{h}-\lambda_{h}\mathbb{T}_{h}\mathbf{u}_{h}, \mathbf{z})+\mathbb{B}_{h}(\mathbf{z},\mathbb{S}(\lambda_{h}\mathbf{u}_{h})- \mathbb{S}_{h}(\lambda_{h}\mathbf{u}_{h}) )|\\ && = |\widetilde{\mathbb{A}}_{h}(\lambda_{h}\mathbb{T}\mathbf{u}_{h}-\lambda \mathbb{T}\mathbf{u}+\mathbf{u}-\mathbf{u}_{h}, \mathbf{z})+\mathbb{B}_{h}(\mathbf{z},\mathbb{S}(\lambda_{h}\mathbf{u}_{h})-\mathbb{S}(\lambda\mathbf{u})+p- p_{h})|\\ &&\leq |\widetilde{\mathbb{A}}_{h}(\mathbf{u}-\mathbf{u}_{h}, \mathbf{z})+\mathbb{B}_{h}(\mathbf{z},p- p_{h}) )| +C \|\lambda_{h}\mathbf{u}_{h}-\lambda\mathbf{u}\|_{0}\|\mathbf{z}\|_{h}. \end{eqnarray} (3.15)

    Substituting (3.15) into (3.14) gives us

    \begin{align} &\|\lambda_{h} \mathbb{T}\mathbf{u}_{h }-\lambda_{h}\mathbb{T}_{h}\mathbf{u}_{h}\|_{h}+\|\lambda_{h} \mathbb{S}\mathbf{u}_{h }-\lambda_{h}\mathbb{S}_{h}\mathbf{u}_{h}\|_{0}\\ \lesssim &\sup\limits_{0\not = \mathbf{z}\in \mathbb{X}}\frac{|\widetilde{\mathbb{A}}_{h}(\mathbf{u}-\mathbf{u}_{h},\mathbf{z})+\mathbb{B}_{h}(\mathbf{z},p- p_{h}) |}{\|\mathbf{z}\|_{h}}+C( \|\lambda_{h}\mathbf{u}_{h}-\lambda\mathbf{u}\|_{0} +\inf\limits_{\mathbf{z}\in \mathbb{X}}\|\mathbf{u}_{h}-\mathbf{z}\|_{h}). \end{align} (3.16)

    Theorem 2.1 indicates that \|\lambda_{h}\mathbf{u}_{h}-\lambda\mathbf{u}\|_{0} is a small quantity of higher order compared with \|\mathbf{u}-\mathbf{u}_{h}\|_{h}+\|p-p_{h}\|_{0} . From (3.16) and (3.14) the side \lesssim in (3.12) is true. The other side \gtrsim in (3.12) is obvious.

    Lemma 3.2. Under the conditions of Theorem 2.1,

    \begin{align} \widetilde{\mathbb{A}}_{h}(\mathbf{u}-\mathbf{u}_{h}, \mathbf{z})+\mathbb{B}_{h}(\mathbf{z}, p-p_{h}) \lesssim& \sum\limits_{\tau\in\mathcal{T}_{h}}\left(\eta_{R_{\tau}}+\eta_{F_{\tau}}+\eta_{J_{\tau}}\right)\|\mathbf{z}\|_{h} +\|\lambda \mathbf{u}-\lambda_{h} \mathbf{u}_{h}\|_{0}\|\mathbf{z}\|_{h}, \; \; \; \forall z\in \mathbb{X}. \end{align} (3.17)

    Proof. Using (2.7), (3.6), (3.7) and the Green's formula we deduce that

    \begin{align} &\widetilde{\mathbb{A}}_{h}(\mathbf{u}-\mathbf{u}_{h},\mathbf{z}) +\mathbb{B}_{h}(\mathbf{z},p-p_{h}))\\ = &\widetilde{\mathbb{A}}_{h}(\mathbf{u},\mathbf{z})-\widetilde{\mathbb{A}}_{h}(\mathbf{u}_{h},\mathbf{z}) +\mathbb{B}_{h}(\mathbf{z},p)-\mathbb{B}_{h}(\mathbf{z},p_{h})\\ = &\lambda (\mathbf{u},\mathbf{z})-\widetilde{\mathbb{A}}_{h}(\mathbf{u}_{h},\mathbf{z})-\mathbb{B}_{h}(\mathbf{z},p_{h})\\ = &\lambda\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathbf{u}\cdot\mathbf{z}dx-\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\nabla\mathbf{u}_{h}:\nabla\mathbf{z}dx -\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}A\mathbf{u}_{h}\cdot\mathbf{z}dx +\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathcal{L}(\mathbf{z}):\nabla\mathbf{u}_{h}dx\\ &+\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathcal{L}(\mathbf{u}_{h}):\nabla\mathbf{z}dx +\sum\limits_{F\in\mathcal{E}_{h}}\int_{F}\frac{\gamma}{h_{F}}[\![\underline{\mathbf{u}_{h}}]\!]:[\![\underline{\mathbf{z}}]\!]ds+\sum\limits_{\tau\in\mathcal{T}_{h}} \int_{\tau}div\mathbf{z}p_{h}dx\\ = &\lambda\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathbf{u}\cdot\mathbf{z}dx+\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\Delta\mathbf{u}_{h}\cdot\mathbf{z}dx - \sum\limits_{\tau\in\mathcal{T}_{h}}\sum\limits_{F\subset\partial\tau}\int_{F}\frac{\partial\mathbf{u}_{h}}{\partial\mathbf{n}}\cdot\mathbf{z}ds -\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}A\mathbf{u}_{h}\cdot\mathbf{z}dx \\ &+\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathcal{L}(\mathbf{z}):\nabla\mathbf{u}_{h}dx +\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathcal{L}(\mathbf{u}_{h}):\nabla\mathbf{z}dx -\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\nabla p_{h}\cdot\mathbf{z}dx\\ & +\sum\limits_{\tau\in\mathcal{T}_{h}}\sum\limits_{F\subset\partial\tau}\int_{F}p_{_h}\mathbf{z}\cdot\mathbf{n}ds. \end{align} (3.18)

    By \mathbf{z}^{I}\in \mathbb{X}\cap\mathbb{X}_{h} and (2.2)–(2.5) we obtain

    \begin{align*} \widetilde{\mathbb{A}}_{h}(\mathbf{u}-\mathbf{u}_{h}, \mathbf{z})+\mathbb{B}_{h}(\mathbf{z}, p-p_{h}) = \widetilde{\mathbb{A}}_{h}(\mathbf{u}-\mathbf{u}_{h}, \mathbf{z}-\mathbf{z}^{I})+\mathbb{B}_{h}(\mathbf{z}-\mathbf{z}^{I}, p-p_{h}) +\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}(\lambda \mathbf{u}-\lambda_{h}\mathbf{u}_{h})\cdot\mathbf{z}^{I}dx. \end{align*}

    Using (3.6), the Cauchy-Schwartz inequality, (3.1) and (3.2), (3.18) can be written as follows:

    \begin{align} &\; \; \; \widetilde{\mathbb{A}}_{h}(\mathbf{u}-\mathbf{u}_{h},\mathbf{z})+\mathbb{B}_{h}(\mathbf{z},p-p_{h})\\ & = \lambda\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathbf{u}\cdot(\mathbf{z}-\mathbf{z}^{I})dx+\sum\limits_{\tau\in\mathcal{T}_{h}} \int_{\tau}\Delta\mathbf{u}_{h}\cdot(\mathbf{z}-\mathbf{z}^{I})dx- \sum\limits_{\tau\in\mathcal{T}_{h}}\sum\limits_{F\subset\partial\tau}\int_{F}\frac{\partial\mathbf{u}_{h}}{\partial\mathbf{n}}\cdot(\mathbf{z}-\mathbf{z}^{I})ds \\ &\; \; \; -\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}A\mathbf{u}_{h}\cdot(\mathbf{z}-\mathbf{z}^{I})dx +\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathcal{L}(\mathbf{z}-\mathbf{z}^{I}):\nabla\mathbf{u}_{h}dx +\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathcal{L}(\mathbf{u}_{h}):\nabla(\mathbf{z}-\mathbf{z}^{I})dx\\ &\; \; \; -\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\nabla p_{h}\cdot(\mathbf{z}-\mathbf{z}^{I})dx +\sum\limits_{\tau\in\mathcal{T}_{h}}\sum\limits_{F\subset\partial\tau}\int_{F}p_{_h}(\mathbf{z}-\mathbf{z}^{I})\cdot\mathbf{n}ds+\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}(\lambda \mathbf{u}-\lambda_{h}\mathbf{u}_{h})\cdot\mathbf{z}^{I}dx\\ & = \sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}(\Delta\mathbf{u}_{h}+\lambda_{h}\mathbf{u}_{h}-A\mathbf{u}_{h}-\nabla p_{h})\cdot(\mathbf{z}-\mathbf{z}^{I})dx- \sum\limits_{\tau\in\mathcal{T}_{h}}\sum\limits_{F\subset\partial\tau}\int_{F}\frac{\partial\mathbf{u}_{h}}{\partial\mathbf{n}}\cdot(\mathbf{z}-\mathbf{z}^{I})ds \\ &\; \; \; +\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathcal{L}(\mathbf{z}-\mathbf{z}^{I}):\nabla\mathbf{u}_{h}dx+\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathcal{L}(\mathbf{u}_{h}):\nabla(\mathbf{z}-\mathbf{z}^{I})dx \\ &\; \; \; +\sum\limits_{\tau\in\mathcal{T}_{h}}\sum\limits_{F\subset\partial\tau}\int_{F}p_{_h}(\mathbf{z}-\mathbf{z}^{I})\cdot\mathbf{n}ds +\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}(\lambda \mathbf{u}-\lambda_{h}\mathbf{u}_{h})\cdot\mathbf{z}dx\\ & = \sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}(\Delta\mathbf{u}_{h}+\lambda_{h}\mathbf{u}_{h}-A\mathbf{u}_{h}-\nabla p_{h})\cdot(\mathbf{z}-\mathbf{z}^{I})dx+\sum\limits_{\tau\in\mathcal{T}_{h}}\sum\limits_{F\subset\partial\tau\backslash\partial\Omega }\int_{F}(p_{h}\mathbf{I}-\nabla\mathbf{u}_{h})\mathbf{n}\cdot(\mathbf{z}-\mathbf{z}^{I})ds \\ &\; \; \; +\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathcal{L}(\mathbf{z}-\mathbf{z}^{I}):\nabla\mathbf{u}_{h}dx +\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathcal{L}(\mathbf{u}_{h}):\nabla(\mathbf{z}-\mathbf{z}^{I})dx +\sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}(\lambda \mathbf{u}-\lambda_{h}\mathbf{u}_{h})\cdot\mathbf{z}dx\\ &\equiv B_{1}+B_{2}+B_{3}+B_{4}+B_{5}. \end{align} (3.19)

    Next, we will analyze each item on the right-hand side of (3.19). Using the Cauchy-Schwartz inequality and the approximation property (3.1) and (3.2), we have

    \begin{align*} |B_{1}|&\leq \sum\limits_{\tau\in\mathcal{T}_{h}}\|\Delta\mathbf{u}_{h}+\lambda_{h}\mathbf{u}_{h}-A\mathbf{u}_{h}-\nabla p_{h}\|_{0,\tau} \|\mathbf{z}-\mathbf{z}^{I} \|_{0,\tau}\\ &\lesssim\sum\limits_{\tau\in\mathcal{T}_{h}}h_{\tau}\|\Delta\mathbf{u}_{h}+\lambda_{h}\mathbf{u}_{h}-A\mathbf{u}_{h}-\nabla p_{h}\|_{0,\tau} \|\mathbf{z}\|^{2}_{1,\theta_{\tau}}\\ &\lesssim\left(\sum\limits_{\tau\in\mathcal{T}_{h}}h_{\tau}^{2}\|\Delta\mathbf{u}_{h}+\lambda_{h}\mathbf{u}_{h}-A\mathbf{u}_{h}-\nabla p_{h}\|_{0,\tau}^{2}\right)^{\frac{1}{2}} \|\mathbf{z}\|_{h}. \end{align*}

    From (3.2) we deduce

    \begin{align*} |B_{2}|& = |\frac{1}{2}\sum\limits_{\tau\in\mathcal{T}_{h}}\sum\limits_{F\subset\partial\tau\backslash\partial\Omega} \int_{F}[\![p_{h}\mathbf{I}-\nabla\mathbf{u}_{h}]\!]\cdot(\mathbf{z}-\mathbf{z}^{I})ds|\\ &\leq \sum\limits_{\tau\in\mathcal{T}_{h}}\sum\limits_{F\subset\partial\tau\backslash\partial\Omega} \|[\![p_{h}\mathbf{I}-\nabla\mathbf{u}_{h}]\!]\|_{0,F} Ch_{F}^{\frac{1}{2}}\|\mathbf{z}\|_{1,\theta_{F}}\\ &\lesssim\left(\sum\limits_{\tau\in\mathcal{T}_{h}}\sum\limits_{F\subset\partial\tau\backslash\partial\Omega}( h_{F}^{\frac{1}{2}}\|[\![p_{h}\mathbf{I}-\nabla\mathbf{u}_{h}]\!]\|_{0,F})^{2} \right)^{\frac{1}{2}} \|\mathbf{z}\|_{h}. \end{align*}

    For the third term, by the properties of the interpolation function \mathbf{z}^{I} we know [\![\mathbf{z}-\mathbf{z}^{I}]\!] = 0 . Therefore, from the definition of lifting operation \mathcal{L} we have

    \begin{align*} B_{3} = \sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}\mathcal{L}(\mathbf{z}-\mathbf{z}^{I}):\nabla\mathbf{u}_{h}ds = \sum\limits_{F\in\mathcal{E}_{h}}\int_{F}\{\underline{\nabla\mathbf{u}_{h}}\}:[\![\underline{\mathbf{z}-\mathbf{z}^{I}}]\!]ds = 0. \end{align*}

    By the Cauchy-Schwartz inequality, (3.4) and (3.1) we obtain

    \begin{align*} |B_{4}|&\leq \left(\sum\limits_{\tau\in\mathcal{T}_{h}}\|\mathcal{L}(\mathbf{u}_{h})\|_{0,\tau}^{2}\right) ^{\frac{1}{2}} \left(\sum\limits_{\tau\in\mathcal{T}_{h}}\|\nabla(\mathbf{z}-\mathbf{z}^{I})\|_{0,\tau}^{2}\right) ^{\frac{1}{2}}\\ &\lesssim\left(\sum\limits_{F\in\mathcal{E}_{h}^{i}}\|h_{F}^{-\frac{1}{2}}[\![\underline{\mathbf{u}_{h}}]\!]\|_{0,F}^{2} \right)^{\frac{1}{2}} \left(\sum\limits_{\tau\in\mathcal{T}_{h}} \|\nabla(\mathbf{z}-\mathbf{z}^{I})\|_{0,\tau}^{2}\right) ^{\frac{1}{2}}\\ &\lesssim\left(\sum\limits_{F\in\mathcal{E}_{h}^{i}}\|h_{F}^{-\frac{1}{2}}[\![\underline{\mathbf{u}_{h}]\!]}\|_{0,F}^{2} \right)^{\frac{1}{2}}\|\mathbf{z}\|_{h}. \end{align*}

    For the last term of (3.19) we get

    \begin{align*} B_{5} = \sum\limits_{\tau\in\mathcal{T}_{h}}\int_{\tau}(\lambda \mathbf{u}-\lambda_{h}\mathbf{u}_{h})\cdot\mathbf{z}dx \leq \|\lambda\mathbf{u}-\lambda_{h}\mathbf{u}_{h}\|_{0}\|\mathbf{z}\|_{0}. \end{align*}

    Substituting B_{1} B_{5} into (3.19) results in (3.17).

    In [22,23], the authors constructed the enriching operator E_{h}:\mathbb{X}_{h}\to \mathbb{X}_{h}\cap\mathbb{X} by averaging and proved the following lemma.

    Lemma 3.3. It is valid the following estimate:

    \begin{eqnarray} \|\mathbf{u}_{h}-E_{h}\mathbf{u}_{h}\|_{h}\lesssim\sum\limits_{F\in\mathcal{E}_{h}^{i}}\gamma h_{F}^{-1}|[\![\underline{\mathbf{u}_{h}}]\!]|_{0,F}^{2}+ \sum\limits_{F\in\mathcal{E}_{h}^{b}}\gamma h_{F}^{-1}|\mathbf{u}_{h}\otimes\mathbf{n}|_{0,F}^{2}. \end{eqnarray} (3.20)

    Theorem 3.2. Suppose that the conditions of Theorem 2.1 hold. Then,

    \begin{eqnarray} \; \; \; \; \; \; \|\mathbf{u}-\mathbf{u}_{h}\|_{h}+\|p-p_{h}\|_{0} \lesssim\eta_{h}+\|\lambda_{h}\mathbf{u}_{h}-\lambda \mathbf{u}\|_{0}. \end{eqnarray} (3.21)

    Proof. Substituting (3.17) and (3.20) into (3.12), we obtain (3.21).

    Let b_{\tau} and b_{F} be the standard bubble function on element \tau and face F ( d = 3 ) or edge F ( d = 2 ) of \tau , respectively. Then, from [20,21,35] we obtain the following lemma.

    Lemma 3.4. For any vector-valued polynomial function \mathbf{z}_{h} on \tau ,

    \begin{align} \|\mathbf{z}_{h}\|_{0,\tau}&\lesssim \|b_{\tau}^{1/2}\mathbf{z}_{h}\|_{0,\tau}, \end{align} (3.22)
    \begin{align} \|b_{\tau}\mathbf{z}_{h}\|_{0,\tau}&\lesssim \|\mathbf{z}_{h}\|_{0,\tau}, \end{align} (3.23)
    \begin{align} \|\nabla (b_{\tau}\mathbf{z}_{h})\|_{0,\tau}&\lesssim h_{\tau}^{-1}\|\mathbf{z}_{h}\|_{0,\tau}. \end{align} (3.24)

    For any vector-valued polynomial function \sigma on F it is valid that

    \begin{align} \|b_{E}\sigma\|_{0,F}&\lesssim \|\sigma\|_{0,F}, \end{align} (3.25)
    \begin{align} \|\sigma\|_{0,F}&\lesssim \|b_{F}^{1/2}\sigma\|_{0,F}. \end{align} (3.26)

    Furthermore, for each b_{F}\sigma there exists an extension \sigma_{b}\in H_{0}^{1}(\omega(F)) satisfying \sigma_{b}|_{F} = b_{F}\sigma and

    \begin{align} \|\sigma_{b}\|_{0,\tau}&\lesssim h_{F}^{1/2}\|\sigma\|_{0,F},\; \; \; \forall \tau\in\omega(F), \end{align} (3.27)
    \begin{align} \|\nabla \sigma_{b}\|_{0,\tau}&\lesssim h_{F}^{-1/2}\|\sigma\|_{0,F},\; \; \; \forall \tau\in\omega(F). \end{align} (3.28)

    Using the standard arguments (see, e.g., Lemma 3.13 in [36]) and Lemmas 7 and 8 in [2], we can deduce the following local bounds.

    Lemma 3.5. Under the conditions of Theorem 2.1,

    \begin{align} &\eta_{R_{\tau}}\lesssim \|\nabla (\mathbf{u}-\mathbf{u}_{h})\|_{0,\tau}+\|p-p_{h}\|_{0,\tau} +h_{\tau}\|\lambda_{h}\mathbf{u}_{h}-\lambda \mathbf{u}\|_{0,\tau}, \end{align} (3.29)
    \begin{align} &\eta_{F_{\tau}} \lesssim \|\nabla (\mathbf{u}-\mathbf{u}_{h})\|_{0,\omega(\tau)}+\|p-p_{h}\|_{0,\omega(\tau)} +\left( \sum\limits_{\tau\in\omega(\tau)}h_{\tau}^{2} \|\lambda\mathbf{u}-\lambda_{h}\mathbf{u}_{h}\|_{0,\tau}^{2} \right)^{\frac{1}{2}}, \end{align} (3.30)
    \begin{align} &\eta^{2}_{J_{\tau}} = \sum\limits_{F\subset\partial\tau, F\in\mathcal{E}^{i}_{h}}\gamma h_{F}^{-1}|[\![\underline{\mathbf{u}_{h}-\mathbf{u}}]\!]|_{0,F}^{2}+ \sum\limits_{F\subset\partial\tau, F\in\mathcal{E}^{b}_{h}}\gamma h_{F}^{-1}|(\mathbf{u}_{h}-\mathbf{u})\otimes\mathbf{n}|_{0,F}^{2}. \end{align} (3.31)

    Proof. For any \tau\in \mathcal{T}_{h} define the function R and K locally by

    \begin{align*} R|_{\tau} = \lambda_{h}\mathbf{u}_{h}+\triangle\mathbf{u}_{h}-A\mathbf{u}_{h}-\nabla p_{h}\; \; \; {\rm {and}}\; \; \; K|_{\tau} = h_{\tau}^{2}Rb_{\tau}. \end{align*}

    From (3.22) and using \lambda\mathbf{u}+\Delta \mathbf{u}-A\mathbf{u}-\nabla p = 0 , we have

    \begin{align*} h_{\tau}^{2}\|R\|_{0,\tau}^{2}&\lesssim \int_{\tau}R\cdot (h_{\tau}^{2}Rb_{\tau})dx \\ & = \int_{\tau}(\lambda_{h}\mathbf{u}_{h}+\Delta \mathbf{u}_{h}-A\mathbf{u}_{h}-\nabla p_{h})\cdot Kdx\\ & = \int_{\tau}(\lambda_{h}\mathbf{u}_{h}+\Delta \mathbf{u}_{h}-A\mathbf{u}_{h}-\nabla p_{h}-(\lambda\mathbf{u}+\Delta \mathbf{u}-A\mathbf{u}-\nabla p))\cdot Kdx\\ & = \int_{\tau}\Delta(\mathbf{u}_{h}-\mathbf{u})\cdot Kdx-\int_{\tau}\nabla(p_{h}-p)\cdot Kdx+\int_{\tau}(\lambda_{h}\mathbf{u}_{h}-\lambda\mathbf{u})\cdot Kdx -\int_{\tau}A(\mathbf{u}_{h}-\mathbf{u})\cdot Kdx. \end{align*}

    Using integration by parts and K|_{\partial\tau} = 0 , we obtain

    \begin{align*} h_{\tau}^{2}\|R\|_{0,\tau}^{2} \lesssim \int_{\tau}\nabla (\mathbf{u}-\mathbf{u}_{h})\cdot \nabla Kdx+\int_{\tau}(p_{h}-p) div Kdx+\int_{\tau}(\lambda_{h}\mathbf{u}_{h}-\lambda\mathbf{u})\cdot Kdx +\int_{\tau}A(\mathbf{u}_{h}-\mathbf{u})\cdot Kdx. \end{align*}

    Applying the Cauchy-Schwartz inequality yields

    \begin{align} h_{\tau}^{2}\|R\|_{0,\tau}^{2} \lesssim\left( \|\nabla (\mathbf{u}-\mathbf{u}_{h})\|_{0,\tau}+\|p-p_{h}\|_{0,\tau} +h_{\tau}\|\lambda_{h}\mathbf{u}_{h}-\lambda \mathbf{u}\|_{0,\tau}\right. \left.+h_{\tau}\|A\mathbf{u}_{h}-A\mathbf{u}\|_{0,\tau}\right)\left(\|\nabla K\|_{0,\tau}+h_{\tau}^{-1}\|K\|_{0,\tau}\right). \end{align} (3.32)

    From (3.23) and (3.24) we get

    \begin{align*} \|\nabla K\|_{0,\tau}+h_{\tau}^{-1}\|K\|_{0,\tau}\lesssim h_{\tau}\|R\|_{0,\tau}. \end{align*}

    Dividing (3.32) by h_{\tau}\|R\|_{0, \tau} and noting \|\nabla\cdot \mathbf{u}_{h}\|_{0} = \|\nabla\cdot (\mathbf{u}_{h}-\mathbf{u})\|_{0} , we obtain (3.29).

    For any interior edge F\in\mathcal{E}_{h}^{i} let the functions R and \Theta be such that

    \begin{align*} R|_{F} = [\![p_{h}\mathbf{I}-\nabla\mathbf{u}_{h} ]\!]|_{F}\; \; {\rm{and}}\; \; \Theta = {\it h_{F}Rb_{F}}. \end{align*}

    Using (3.26) and [\![p\mathbf{I}-\nabla\mathbf{u}]\!]|_{F} = 0 we get

    \begin{align*} h_{F}\|R\|_{0,F}^{2}\lesssim \int_{F}R\cdot (h_{F}Rb_{F})ds = \int_{F}[\![(p_{h}-p)\mathbf{I}-\nabla(\mathbf{u}_{h}-\mathbf{u})]\!]\cdot\Theta ds. \end{align*}

    Applying the Green's formula over each element of \omega(F) we derive

    \begin{align*} h_{E}\|R\|_{0,F}^{2}&\lesssim\int_{F}[\![((p_{h}-p)\mathbf{I}-\nabla(\mathbf{u}_{h}-\mathbf{u}))]\!]\cdot\Theta ds\\ & = C (\sum\limits_{\tau\in\omega(F)}\int_{\tau}(-\Delta (\mathbf{u}-\mathbf{u}_{h})+\nabla (p-p_{h}))\cdot\Theta dx -\sum\limits_{\tau\in\omega(F)}\int_{\tau}(\nabla (\mathbf{u}-\mathbf{u}_{h})-(p-p_{h})\mathbf{I}):\nabla\Theta dx). \end{align*}

    Using \lambda\mathbf{u}+\Delta\mathbf{u}-A\mathbf{u}-\nabla p = 0 we deduce

    \begin{align} h_{F}\|R\|_{0,F}^{2}&\lesssim \sum\limits_{\tau\in\omega(F)}\int_{\tau}(\lambda_{h}\mathbf{u}_{h}+\Delta\mathbf{u}_{h}-A\mathbf{u}_{h}-\nabla p_{h})\cdot\Theta dx+\sum\limits_{\tau\in\omega(F)}\int_{\tau}(\lambda\mathbf{u}-\lambda_{h}\mathbf{u}_{h})\cdot\Theta dx\\ &\; \; \; +\sum\limits_{\tau\in\omega(F)}\int_{\tau}(-\nabla (u-u_{h})+(p-p_{h})\mathbf{I}):\nabla\Theta dx +\sum\limits_{\tau\in\omega (F)}\int_{\tau}(A\mathbf{u}-A\mathbf{u}_{h})\cdot\Theta dx\\ &\equiv T_{1}+T_{2}+T_{3}+T_{4}. \end{align} (3.33)

    Using the Cauchy-Schwartz inequality, (3.27) and (3.28) yieids

    \begin{align*} &T_{1}\lesssim \left(\sum\limits_{\tau\in\omega(F)}\eta^{2}_{R_{\tau}}\right)^{1/2}\left(\sum\limits_{\tau\in\omega(F)}h^{-2}_{\tau}\|\Theta\|_{0,\tau}^{2} \right)^{1/2}\lesssim \left(\sum\limits_{\tau\in\omega(F)}\eta^{2}_{R_{\tau}}\right)^{1/2}h_{F}^{1/2}\|R\|_{0,F},\\ &T_{2}\lesssim \left( \sum\limits_{\tau\in\omega(F)}\left(h^{2}_{\tau}\|\lambda\mathbf{u}-\lambda_{h}\mathbf{u}_{h} \|^{2}_{0,\tau}\right)\right)^{1/2}h_{F}^{1/2}\|R\|_{0,F},\\ &T_{3}\lesssim \left(\sum\limits_{\tau\in\omega(F)}(\|\nabla (\mathbf{u}-\mathbf{u}_{h}) \|^{2}_{0,\tau}+\|p-p_{h}\| ^{2}_{0,\tau}) \right)^{1/2}h_{F}^{1/2}\|R\|_{0,F},\\ &T_{4}\lesssim \left( \sum\limits_{\tau\in\omega(F)}\left(h^{2}_{\tau}\|A\mathbf{u}-A\mathbf{u}_{h} \|^{2}_{0,\tau}\right)\right)^{1/2}h_{F}^{1/2}\|R\|_{0,F}. \end{align*}

    Combing the above estimates of T_{1} , T_{2} , T_{3} and T_{4} , dividing (3.33) by h_{F}^{1/2}\|R\|_{0, F} and summing over all interior edges of \tau gives us (3.30).

    For any F\in \mathcal{E}_{h}^{i}(\Omega) , [\![\underline{\mathbf{u}}]\!] = 0 and for any F\in \mathcal{E}_{h}\cap \partial\Omega , \mathbf{u}\otimes\mathbf{n} = 0 . Therefore, we obtain (3.31) and finish the proof.

    Theorem 3.3. Suppose that the conditions of Theorem 2.1 hold. Then, the a posteriori error estimator \eta_{h} is efficient:

    \begin{eqnarray} &&\eta^{2}_{\tau}\lesssim\sum\limits_{\tau\in\omega(\tau)}( \|\mathbf{u}-\mathbf{u}_{h}\|^{2}_{0,\tau}+\|p-p_{h}\|^{2}_{0,\tau}+h_{\tau}^{2} \|\lambda\mathbf{u}-\lambda_{h}\mathbf{u}_{h}\|^{2}_{0,\tau}), \end{eqnarray} (3.34)
    \begin{eqnarray} &&\eta^{2}_{h}\lesssim \|\mathbf{u}-\mathbf{u}_{h}\|^{2}_{h}+\|p-p_{h}\|^{2}+\sum\limits_{\tau\in\mathcal{T}_{h}}h_{\tau}^{2}\|\lambda\mathbf{u}-\lambda_{h}\mathbf{u}_{h}\|^{2}_{0,\tau} . \end{eqnarray} (3.35)

    Lemma 3.6. Let (\lambda, \mathbf{u}, p) and (\lambda_{h}, \mathbf{u}_{h}, p_{h}) be the eigenpairs of (2.2), (2.3) and (2.4), (2.5), respectively. Then,

    \begin{eqnarray} \lambda_{h}-\lambda = \mathbb{A}_{h}(\mathbf{u}-\mathbf{u}_{h}, \mathbf{u}-\mathbf{u}_{h})+2\mathbb{B}_{h}(\mathbf{u}-\mathbf{u}_{h},p-p_{h})-\lambda(\mathbf{u}-\mathbf{u}_{h}, \mathbf{u}-\mathbf{u}_{h}). \end{eqnarray} (3.36)

    Proof. By using (2.16) and (2.17) we get

    \begin{align} \mathbb{A}_{h}(\mathbf{u},\mathbf{z}_{h})+\mathbb{B}_{h}(\mathbf{z}_{h},p)& = \lambda(\mathbf{u},\mathbf{z}_{h}),\; \; \; \; \forall \mathbf{z}_{h}\in\mathbb{X}_{h}, \end{align} (3.37)
    \begin{align} \mathbb{B}_{h}(\mathbf{u},\varrho_{h})& = 0,\; \; \; \; \forall \varrho_{h}\in \mathbb{W}_{h}. \end{align} (3.38)

    From (2.2) and (2.3) with (\mathbf{z}, \varrho) = (\mathbf{u}, p) , (2.4) and (2.5) with (\mathbf{z}_{h}, \varrho_{h}) = (\mathbf{u_{h}}, p_{h}) and (3.37), (3.38) we deduce

    \begin{eqnarray*} &&\mathbb{A}_{h}(\mathbf{u}-\mathbf{u}_{h}, \mathbf{u}-\mathbf{u}_{h})+2\mathbb{B}_{h}(\mathbf{u}-\mathbf{u}_{h},p-p_{h})-\lambda(\mathbf{u}-\mathbf{u}_{h}, \mathbf{u}-\mathbf{u}_{h})\nonumber\\ & = &\mathbb{A}_{h}(\mathbf{u}, \mathbf{u})-2\mathbb{B}_{h}(\mathbf{u},\mathbf{u}_{h})+\mathbb{A}_{h}(\mathbf{u}_{h},\mathbf{u}_{h}) +2\mathbb{B}_{h}(\mathbf{u}, p)-2\mathbb{B}_{h}(\mathbf{u}_{h},p)\nonumber\\ && -2\mathbb{B}_{h}(\mathbf{u},p_{h})+2\mathbb{B}_{h}(\mathbf{u}_{h},p_{h})-\lambda(\mathbf{u}, \mathbf{u})+2\lambda(\mathbf{u},\mathbf{u}_{h})-\lambda(\mathbf{u}_{h},\mathbf{u}_{h})\nonumber\\ & = &\lambda_{h}(\mathbf{u}_{h},\mathbf{u}_{h})-\lambda(\mathbf{u}_{h},\mathbf{u}_{h}) = \lambda_{h}-\lambda. \end{eqnarray*}

    We complete the proof.

    Theorem 3.4. Under the conditions of Theorem 2.1,

    \begin{eqnarray} |\lambda-\lambda_{h}|\lesssim \eta_{h}^{2}+\sum\limits_{\tau\in\mathcal{T}_{h}}h_{\tau}^{2r}(|\mathbf{u}-I_{h}\mathbf{u}|_{1+r,\tau}^{2} +\|p-\vartheta_{h}p\|_{r}^{2}). \end{eqnarray} (3.39)

    Proof. Theorem 2.1 indicates that \|\mathbf{u}-\mathbf{u}_{h}\|_{0} is a term of higher order than |||\mathbf{u}-\mathbf{u}_{h}|||+\|p-p_{h}\|_{0} . Hence, from (3.36) and (3.21), we obtain

    \begin{eqnarray*} |\lambda-\lambda_{h}|\lesssim |||\mathbf{u}-\mathbf{u}_{h}|||^{2}+\|p-p_{h}\|_{0}^{2}+\sum\limits_{F\in\mathcal{E}_{h}}h_{F}\|p-p_{h}\|_{0,F}^{2}. \end{eqnarray*}

    Thus, from (2.39) and (3.21) we obtain (3.39).

    Remark 3.1. Theorem 2.1 indicates that \|\lambda_{h}\mathbf{u}_{h}-\lambda \mathbf{u}\|_{0} is a small quantity of higher order than \|\mathbf{u}-\mathbf{u}_{h}\|_{h}+\|p-p_{h}\|_{0} . Theorems 3.2 and 3.3 show that the estimator \eta_{h} for the eigenfunction error \|\mathbf{u}-\mathbf{u}_{h}\|_{h}+\|p-p_{h}\|_{0} is reliable and efficient up to data oscillation. Therefore, a good graded mesh is generated by the adaptive algorithm for the estimator, which makes the eigenfunction error \|\mathbf{u}-\mathbf{u}_{h}\|_{h}+\|p-p_{h}\|_{0} reach the optimal convergence rate O(dof^{-\frac{2k}{d}}) . Hence, from [37,38,39] we can look forward to getting \sum\limits_{\tau\in\mathcal{T}_{h}}h_{\tau}^{2r}(|\mathbf{u}-I_{h}\mathbf{u}|_{1+r, \tau}^{2} +\|p-\vartheta_{h}p\|_{r}^{2})\lesssim dof^{-\frac{2k}{d}} . Thereby from (3.39) we have |\lambda-\lambda_{h}|\lesssim dof^{-\frac{2k}{d}} . Therefore, we think \eta_{h}^{2} can be regarded as the error estimator of \lambda_{h} .

    Remark 3.2. Based on [17], for the problem (2.1) all analysis and conclusions in this paper are valid for the mixed DGFEM using the \mathbb{Q}_{k}-\mathbb{Q}_{k-1} element.

    When \lambda is a multiple eigenvalue the exact eigenfunction approximated by the discrete eigenfunction will change with the change of mesh diameter. In order to implement the adaptive algorithm better, we will conduct our numerical experiments on simple eigenpairs (multiplicity 1).

    Based on [40,41,42], we design an adaptive DGFEM algorithm (ADGFEM) by adopting the standard adaptive loop with the steps solve, estimate, mark and refine with the a posteriori error estimator given in Section 3. We compile our program with the help of the iFEM package [43] and solve the matrix eigenvalue problem by means of the command 'eigs' in MATLAB.

    We adapt the following symbols in our tables:

    l : the l th iteration.

    \lambda_{k, h_{l}} : the k th approximate eigenvalue at the l th iteration.

    dof : the degrees of freedom at the l th iteration.

    The experiment is conducted in three two-dimensional domains the slit domain \Omega_{slit} = (-1, 1)^{2}\setminus \{{0\leq x\leq 1, y = 0}\} , the L-shaped domain \Omega_{L} = (-1, 1)^{2}\setminus [0, 1]\times [-1, 0] and the unit square domain \Omega_{square} = (0, 1)^{2} . In the step mark we select the parameter \theta = 0.5 , and the initial mesh \pi_{h_{0}} with h_{0} = \frac{\sqrt{2}}{16} for the above three two-dimensional domains.

    The reference values for the first eigenvalue of the classical Stokes eigenvalue problem are \lambda_{1, slit} = 29.9168629 , \lambda_{1, L} = 32.13269465 and \lambda_{1, square} = 52.344691168 for \Omega_{slit} , \Omega_{L} and \Omega_{square} , respectively (see [1,2]) and the reference value for the fourth eigenvalue reads \lambda_{4, square} = 128.209584313 in \Omega_{square} (see [2]). We choose the values \lambda_{4, slit} = 40.1527333966 and \lambda_{4, L} = 48.9835839778 as the reference values for the \Omega_{slit} and \Omega_{L} respectively, which are obtained by adaptive procedure using \mathbb{P}_{3}-\mathbb{P}_{2} element with as much degrees of freedom as possible.

    The error curves for the first eigenvalue of the classical Stokes eigenvalue problem are shown in Figures 13 and the fourth eigenvalue are shown in Figures 46. The adaptive refined meshes for the first eigenvalue of the classical Stokes eigenvalue problem by the ADGFEM are shown in Figure 7.

    Figure 1.  The error curves of the first eigenvalue by the ADGFEM using \mathbb{P}_{2}-\mathbb{P}_{1} element (left) and \mathbb{P}_{3}-\mathbb{P}_{2} element (right) for the classical Stokes eigenvalue problem in \Omega_{slit} .
    Figure 2.  The error curves of the first eigenvalue by the ADGFEM using \mathbb{P}_{2}-\mathbb{P}_{1} element (left) and \mathbb{P}_{3}-\mathbb{P}_{2} element (right) for the classical Stokes eigenvalue problem in \Omega_{L} .
    Figure 3.  The error curves of the first eigenvalue by the ADGFEM using \mathbb{P}_{2}-\mathbb{P}_{1} element (left) and \mathbb{P}_{3}-\mathbb{P}_{2} element (right) for the classical Stokes eigenvalue problem in \Omega_{square} .
    Figure 4.  The error curves of the fourth eigenvalue by the ADGFEM using \mathbb{P}_{2}-\mathbb{P}_{1} element (left) and \mathbb{P}_{3}-\mathbb{P}_{2} element (right) for the classical Stokes eigenvalue problem in \Omega_{slit} .
    Figure 5.  The error curves of the fourth eigenvalue by the ADGFEM using \mathbb{P}_{2}-\mathbb{P}_{1} element (left) and \mathbb{P}_{3}-\mathbb{P}_{2} element (right) for the classical Stokes eigenvalue problem in \Omega_{L} .
    Figure 6.  The error curves of the fourth eigenvalue by the ADGFEM using \mathbb{P}_{2}-\mathbb{P}_{1} element (left) and \mathbb{P}_{3}-\mathbb{P}_{2} element (right) for the classical Stokes eigenvalue problem in \Omega_{square} .
    Figure 7.  The adaptive meshes for the first eigenvalue of the classical Stokes eigenvalue problem by the ADGFEM at l = 25 refinement times using \mathbb{P}_{3}-\mathbb{P}_{2} element in \Omega_{slit} (left) and \Omega_{L} (right).

    We observe from Figures 16 that the error curves and error estimators curves for ADGFEM are both almost parallel to the straight line with a slope of -k which indicates that the error estimators are reliable and efficient and the adaptive algorithm can achieve the optimal convergence order. This is consistent with our theoretical results. We also observe from the error curves that under the same dof the approximations obtained by the ADGFEM are more accurate than those computed on uniform meshes.

    The approximations of the first eigenvalue obtained by \mathbb{P}_{3}-\mathbb{P}_{2} element in \Omega_{slit} , \Omega_{L} and \Omega_{square} are listed in Tables 13, respectively. These eigenvalues have the same accuracy as those in [1,2] which achieve 9, 10 and 11 significant digits in \Omega_{slit} , \Omega_{L} and \Omega_{square} , respectively. Furthermore, it shows that our method is effective. The approximations of the fourth eigenvalue obtained by \mathbb{P}_{3}-\mathbb{P}_{2} element in \Omega_{slit} , \Omega_{L} and \Omega_{square} are listed in Tables 46, respectively.

    Table 1.  The approximation of the first eigenvalue of the classical Stokes eigenvalue problem in \Omega_{slit} obtained by the ADGFEM using \mathbb{P}_{3}-\mathbb{P}_{2} element.
    l dof \lambda_{1, h_{l}} l dof \lambda_{1, h_{l}}
    1 53248 29.950023991 26 63206 29.916921865
    5 53560 29.917626784 30 73424 29.916878484
    10 54028 29.917180037 35 110630 29.916865373
    15 54756 29.917731006 40 175812 29.916863378
    25 61412 29.916940636 50 537862 29.916862882

     | Show Table
    DownLoad: CSV
    Table 2.  The approximation of the first eigenvalue of the classical Stokes eigenvalue problem in \Omega_{L} obtained by the ADGFEM using \mathbb{P}_{3}-\mathbb{P}_{2} element.
    l dof \lambda_{1, h_{l}} l dof \lambda_{1, h_{l}}
    1 39936 32.155997914 27 53612 32.132716405
    5 40248 32.139031080 31 75140 32.132699385
    15 41288 32.134171324 41 229424 32.132694780
    24 48880 32.132752576 50 703092 32.132694653
    25 50128 32.132737367 51 796276 32.132694652
    26 51688 32.132725042

     | Show Table
    DownLoad: CSV
    Table 3.  The approximation of the first eigenvalue of the classical Stokes eigenvalue problem in \Omega_{square} obtained by the ADGFEM using \mathbb{P}_{3}-\mathbb{P}_{2} element.
    l dof \lambda_{1, h_{l}} l dof \lambda_{1, h_{l}}
    1 53248 52.3446926681 10 273780 52.3446911721
    5 95316 52.3446912380 14 610376 52.3446911684
    8 170612 52.3446911794 17 1186328 52.3446911679
    9 220324 52.3446911751

     | Show Table
    DownLoad: CSV
    Table 4.  The approximation of the fourth eigenvalue of the classical Stokes eigenvalue problem in \Omega_{slit} obtained by the ADGFEM using \mathbb{P}_{3}-\mathbb{P}_{2} element.
    l dof \lambda_{4, h_{l}} l dof \lambda_{4, h_{l}}
    1 53248 40.1565119894 27 106860 40.1527357945
    15 55068 40.1528559112 41 457002 40.1527334275
    25 91416 40.1527379227 51 1421836 40.1527333968
    26 99788 40.1527369150 52 1618838 40.1527333966

     | Show Table
    DownLoad: CSV
    Table 5.  The approximation of the fourth eigenvalue of the classical Stokes eigenvalue problem in \Omega_{L} obtained by the ADGFEM using \mathbb{P}_{3}-\mathbb{P}_{2} element.
    l dof \lambda_{4, h_{l}} l dof \lambda_{4, h_{l}}
    1 39936 48.9840225306 15 147472 48.9835843422
    2 39988 48.9836097839 16 180648 48.9835842150
    5 40560 48.9836170562 19 313248 48.9835840155
    10 73112 48.9835869187 24 807768 48.9835839802
    13 101920 48.9835847163 27 1414244 48.9835839778
    14 120952 48.9835845197

     | Show Table
    DownLoad: CSV
    Table 6.  The approximation of the fourth eigenvalue of the classical Stokes eigenvalue problem in \Omega_{square} obtained by the ADGFEM using \mathbb{P}_{3}-\mathbb{P}_{2} element.
    l dof \lambda_{4, h_{l}} l dof \lambda_{4, h_{l}}
    1 53248 128.2096127378 10 299052 128.2095843960
    2 64376 128.2095967670 11 365664 128.2095843475
    7 143312 128.2095846555 16 1057888 128.2095843150
    8 187824 128.2095845453 17 1302184 128.2095843141
    9 240084 128.2095844591

     | Show Table
    DownLoad: CSV

    The experiment is conducted in two three-dimensional domains: \Omega_{1} = (0, 1)^{3}\setminus \{0\leq x\leq 0.5, 0\leq y\leq 0.5, 0.5\leq z\leq 1\} and \Omega_{2} = (0, 1)^{3} . In computation we select the initial mesh \pi_{h_{0}} with h_{0} = \frac{\sqrt{3}}{8} and \theta = 0.25 .

    The reference values for the first eigenvalue of the classical Stokes eigenvalue problem are \lambda_{\Omega_{1}} = 70.98560 and \lambda_{\Omega_{2}} = 62.17341 for the domains \Omega_{1} and \Omega_{2} respectively, which are calculated by adaptive procedure with as much degrees of freedom as possible.

    The adaptive refined meshes and the error curves are shown in Figures 8 and 9. We observe from Figures 8 and 9 that the error estimators are reliable and efficient and the adaptive algorithm achieve the optimal convergence order.

    Figure 8.  Adaptive mesh after l = 12 refinement times (left) and the error curves (right) of the first eigenvalue by the ADGFEM using \mathbb{P}_{3}-\mathbb{P}_{2} element for the classical Stokes eigenvalue problem in \Omega_{1} .
    Figure 9.  Adaptive mesh after l = 5 refinement times (left) and the error curves (right) of the first eigenvalue by the ADGFEM using \mathbb{P}_{3}-\mathbb{P}_{2} element for the classical Stokes eigenvalue problem in \Omega_{2} .

    We conduct experiments in \Omega_{L} and \Omega_{square} . We select \theta = 0.5 and the initial mesh \pi_{h_{0}} with h_{0} = \frac{\sqrt{2}}{16} for the above two two-dimensional domains.

    For the MHD Stokes eigenvalue problem with Ha = 5 , we choose the values \lambda_{1, square} = 64.68920947 and \lambda_{1, L} = 40.2764915 as the reference values for \Omega_{square} and \Omega_{L} , respectively, and while Ha = 30 , we choose the values \lambda_{1, square} = 234.34458093 and \lambda_{1, L} = 125.24247135 as the reference values for \Omega_{square} and \Omega_{L} respectively. These reference values are obtained by the ADGFEM using \mathbb{P}_{3}-\mathbb{P}_{2} element with as much degrees of freedom as possible.

    The error curves for the first eigenvalue are shown in Figures 10 and 11 and the adaptive refined meshes for the first eigenvalue of the MHD Stokes eigenvalue problem by the ADGFEM are shown in Figure 12. We observe from Figures 10 and 11 that the error curves and error estimators curves are both approximately parallel to the line with slope -k , which indicates that the error estimators are reliable and efficient and the adaptive algorithm can achieve the optimal convergence order.

    Figure 10.  The error curves of the first eigenvalue by the ADGFEM using \mathbb{P}_{3}-\mathbb{P}_{2} element for the MHD Stokes eigenvalue problem with Ha = 5 in \Omega_{square} (left) and \Omega_{L} (right).
    Figure 11.  The error curves of the first eigenvalue by the ADGFEM using \mathbb{P}_{3}-\mathbb{P}_{2} element for the MHD Stokes eigenvalue problem with Ha = 30 in \Omega_{square} (left) and \Omega_{L} (right).
    Figure 12.  The adaptive meshes for the first eigenvalue of the MHD Stokes eigenvalue problem when Ha = 5 by the ADGFEM at l = 25 refinement times using \mathbb{P}_{3}-\mathbb{P}_{2} element in \Omega_{L} (left) and at l = 8 refinement times using \mathbb{P}_{3}-\mathbb{P}_{2} element in \Omega_{square} (right).

    The approximations of the first eigenvalue for the MHD Stokes eigenvalue problem in \Omega_{square} using \mathbb{P}_{3}-\mathbb{P}_{2} element are listed in Tables 7 and 8, from which we can see that the approximate eigenvalues also has high accuracy.

    Table 7.  The approximation of the first eigenvalue of the MHD Stokes eigenvalue problem with Ha = 5 in \Omega_{square} obtained by the ADGFEM using \mathbb{P}_{3}-\mathbb{P}_{2} element.
    l dof \lambda_{1, h_{l}} l dof \lambda_{1, h_{l}}
    1 53248 64.689210998 5 155844 64.689209484
    2 66664 64.689209879 6 238108 64.689209476
    3 89180 64.689209585 7 343720 64.689209472
    4 114036 64.689209503 8 482144 64.689209470

     | Show Table
    DownLoad: CSV
    Table 8.  The approximation of the first eigenvalue of the MHD Stokes eigenvalue problem with Ha = 30 in \Omega_{square} obtained by the ADGFEM using \mathbb{P}_{3}-\mathbb{P}_{2} element.
    l dof \lambda_{1, h_{l}} l dof \lambda_{1, h_{l}}
    1 53248 234.34471492 6 225472 234.34458119
    2 57720 234.34462502 7 332124 234.34458104
    3 73944 234.34459345 8 494832 234.34458097
    4 106600 234.34458500 9 738192 234.34458094
    5 158964 234.34458219 10 994656 234.34458093

     | Show Table
    DownLoad: CSV

    We also use the ADGFEM with \mathbb{P}_{k}-\mathbb{P}_{k-1}(k = 1, 2) element to calculate the classical Stokes eigenvalue problem and the MHD Stokes eigenvalue problem. The numerical results indicate that the discrete formulations are stable and effective. Due to article length limitations, these results are not listed in the paper.

    In this paper, for a class of Stokes eigenvalue problems including the classical Stokes eigenvalue problem in R^{d}\; (d = 2, 3) and the MHD Stokes eigenvalue problem et al, based on the velocity-pressure formulation we studied the residual type a posteriori error estimates of the mixed DGFEM using \mathbb{P}_{k}-\mathbb{P}_{k-1} \; (k\geq1) element on shape-regular simplex meshes. We proposed the a posteriori error estimator for approximate eigenpairs and proved the reliability and efficiency of the estimator for eigenfunctions and also analyzed their reliability for eigenvalues. The characteristic of the adaptive DGFEM is that it can use high-order elements and capture local low smooth solutions and can achieve the optimal convergence order O(dof^{-\frac{2k}{d}}) in two and three-dimensional domains. Our method is easy to implement on existing software packages. The numerical results confirmed our theoretical predictions and showed that our method is stable, efficient and can obtain high-accuracy approximate eigenvalues.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    This work was supported by the National Natural Science Foundation of China (Nos. 11561014, 11761022), and Science and Technology Planning Project of Guizhou Province (Guizhou Kehe fundamental research-ZK[2022] No.324).

    The authors declare that this work does not have any conflicts of interest.



    [1] J. Gedicke, A. Khan, Arnold-Winther mixed finite elements for Stokes eigenvalue problems, SIAM J. Sci. Comput., 40 (2018), A3449–A3469. https://doi.org/10.1137/17M1162032 doi: 10.1137/17M1162032
    [2] J. Gedicke, A. Khan, Divergence-conforming discontinuous Galerkin finite elements for Stokes eigenvalue problems, Numer. Math., 144 (2020), 585–614. https://doi.org/10.1007/s00211-019-01095-x doi: 10.1007/s00211-019-01095-x
    [3] F. Lepea, D. Mora, Symmetric and nonsymmetric discontinuous Galerkin methods for a pseudostress formulation of the Stokes spectral problem, SIAM J. Sci. Comput., 42 (2020), A698–A722. https://doi.org/10.1137/19M1259535 doi: 10.1137/19M1259535
    [4] F. Lepe, G. Rivera, A virtual element approximation for the pseudostress formulation of the Stokes eigenvalue problem, Comput. Methods Appl. Mech. Eng., 379 (2021), 113753. https://doi.org/10.1016/j.cma.2021.113753 doi: 10.1016/j.cma.2021.113753
    [5] B. Mercier, J. Osborn, J. Rappaz, P. A. Raviart, Eigenvalue approximation by mixed and hybrid methods, Math. Comp., 36 (1981), 427–453. https://doi.org/10.2307/2007651 doi: 10.2307/2007651
    [6] L. L. Sun, Y. D. Yang, The a posteriori error estimates and adaptive computation of nonconforming mixed finite elements for the Stokes eigenvalue problem, Appl. Math. Comput., 421 (2022), 126951. https://doi.org/10.1016/j.amc.2022.126951 doi: 10.1016/j.amc.2022.126951
    [7] L. L. Sun, H. Bi, Y. D. Yang, The discontinuous Galerkin and the nonconforming ECR element approximations for an MHD Stokes eigenvalue problem, Math. Methods Appl. Sci., 46 (2022), 6154–6176. https://doi.org/10.1002/mma.8897 doi: 10.1002/mma.8897
    [8] Ö. Türk, An MHD Stokes eigenvalue problem and its approximation by a spectral collocation method, Comput. Math. Appl., 80 (2020), 2045–2056. https://doi.org/10.1016/j.camwa.2020.09.002 doi: 10.1016/j.camwa.2020.09.002
    [9] M. G. Armentano, V. Moreno, A posteriori error estimates of stabilized low-order mixed finite elements for the Stokes eigenvalue problem, J. Comput. Appl. Math., 269 (2014), 132–149. https://doi.org/10.1016/j.cam.2014.03.027 doi: 10.1016/j.cam.2014.03.027
    [10] H. P. Liu, W. Gong, S. H. Wang, N. N. Yan, Superconvergence and a posteriori error estimates for the Stokes eigenvalue problems, BIT Numer. Math., 53 (2013), 665–687. https://doi.org/10.1007/s10543-013-0422-8 doi: 10.1007/s10543-013-0422-8
    [11] C. Lovadina, M. Lyly, R. Stenberg, A posteriori estimates for the Stokes eigenvalue problem, Numer. Methods Partial Differ. Equ., 25 (2009), 244–257. https://doi.org/10.1002/num.20342 doi: 10.1002/num.20342
    [12] S. H. Jia, F. S. Lu, H. H. Xie, A Posterior error analysis for the nonconforming discretization of Stokes eigenvalue problem, Acta Math. Sin. Engl. Ser., 30 (2014), 949–967. https://doi.org/10.1007/s10114-014-3121-8 doi: 10.1007/s10114-014-3121-8
    [13] F. Lepe, G. Rivera, J. Vellojin, Error estimates for a vorticity-based velocity-stress formulation of the Stokes eigenvalue problem, J. Comput. Appl. Math., 420 (2023), 114798. https://doi.org/10.1016/j.cam.2022.114798 doi: 10.1016/j.cam.2022.114798
    [14] S. Badia, R. Codina, T. Gudi, J. Guzmán, Error analysis of discontinuous Galerkin methods for the Stokes problem under minimal regularity, IMA. J. Numer. Anal., 34 (2014), 800–819. https://doi.org/10.1093/imanum/drt022 doi: 10.1093/imanum/drt022
    [15] B. Cockburn, G. E. Karniadakis, C. W. Shu, Discontinuous Galerkin methods: theory, computation and applications, Berlin, Heidelberg: Springer, 2000. https://doi.org/10.1007/978-3-642-59721-3
    [16] A. Ern, J. Guermond, Finite elements Ⅱ: Galerkin approximation, elliptic and mixed PDEs, Cham: Springer, 2021. https://doi.org/10.1007/978-3-030-56923-5
    [17] P. Houston, D. Schötzau, T. P. Wihler, Energy norm a posteriori error estimation for mixed discontinuous Galerkin approximations of the Stokes problem, J. Sci. Comput., 22 (2005), 347–370. https://doi.org/10.1007/s10915-004-4143-7 doi: 10.1007/s10915-004-4143-7
    [18] P. Houston, D. Schötzau, T. P. Wihler, hp-adaptive discontinuous Galerkin finite element methods for the Stokes problem, European Congress on Computational Methods in Applied Sciences and Engineering (ECCOMAS), 2004, 1–20.
    [19] D. A. Di Pietro, A. Ern, Discrete functional analysis tools for discontinuous Galerkin methods with application to the incompressible Navier-Stokes equations, Math. Comp., 79 (2010), 1303–1330. https://doi.org/10.1090/s0025-5718-10-02333-1 doi: 10.1090/s0025-5718-10-02333-1
    [20] M. Ainsworth, J. T. Oden, A posteriori error estimation in finite element analysis, New York: Wiley, 2011.
    [21] R. Verfürth, A posteriori error estimation techniques, New York: Oxford University Press, 1996.
    [22] S. C. Brenner, Poincaré-Friedrichs inequalities for piecewise H^{1} functions, SIAM J. Numer. Anal., 41 (2003), 306–324. https://doi.org/10.1137/S0036142902401311 doi: 10.1137/S0036142902401311
    [23] O. A. Karakashian, F. Pascal, A posteriori error estimates for a discontinuous Galerkin approximation of second-order elliptic problems, SIAM J. Numer. Anal., 41 (2003), 2374–2399. https://doi.org/10.1137/s0036142902405217 doi: 10.1137/s0036142902405217
    [24] I. Perugia, D. Schötzau, The hp-local discontinuous Galerkin method for low-frequency time-harmonic Maxwell equations, Math. Comp., 72 (2003), 1179–1214. https://doi.org/10.1090/s0025-5718-02-01471-0 doi: 10.1090/s0025-5718-02-01471-0
    [25] D. Schötzau, C. Schwab, A. Toselli, Mixed hp-DGFEM for incompressible flows, SIAM J. Numer. Anal., 40 (2002), 2171–2194. https://doi.org/10.1137/s0036142901399124 doi: 10.1137/s0036142901399124
    [26] P. Houston, I. Perugia, D. Schötzau, An a posteriori error indicator for discontinuous Galerkin discretizations of H(curl)-elliptic partial differential equations, IMA J. Numer. Anal., 27 (2007), 122–150. https://doi.org/10.1093/imanum/drl012 doi: 10.1093/imanum/drl012
    [27] B. Rivière, Discontinuous Galerkin methods for solving elliptic and parabolic equations: theory and implementation, SIAM, 2008.
    [28] V. Girault, P. A. Raviart, Finite element approximation of the Navier-Stokes equations, Berlin, Heidelberg: Springer, 1979.
    [29] B. Cockburn, G. Kanschat, D. Schötzau, C. Schwab, Local discontinuous Galerkin methods for the Stokes system, SIAM J. Numer. Anal., 40 (2002), 319–343. https://doi.org/10.1137/S0036142900380121 doi: 10.1137/S0036142900380121
    [30] P. Hansbo, M. G. Larson, Discontinuous Galerkin methods for incompressible and nearly incompressible elasticity by Nitsche's method, Comput. Methods Appl. Mech. Eng., 191 (2002), 1895–1908. https://doi.org/10.1016/S0045-7825(01)00358-9 doi: 10.1016/S0045-7825(01)00358-9
    [31] Y. D. Yang, Z. M. Zhang, F. B. Lin, Eigenvalue approximation from below using non-conforming finite elements, Sci. China Ser. A Math., 53 (2010), 137–150. https://doi.org/10.1007/s11425-009-0198-0 doi: 10.1007/s11425-009-0198-0
    [32] I. Babuska, J. Osborn, Eigenvalue problems, In: Handbook of numerical analysis, Elsevier, 1991,641–787. https://doi.org/10.1016/S1570-8659(05)80042-0
    [33] D. Boffi, Finite element approximation of eigenvalue problems, Acta Numer., 19 (2010), 1–120. https://doi.org/10.1017/S0962492910000012 doi: 10.1017/S0962492910000012
    [34] L. R. Scott, S. Y. Zhang, Finite element interpolation of nonsmooth functions satisfying boundary conditions, Math. Comp., 54 (1990), 483–493. https://doi.org/10.1090/S0025-5718-1990-1011446-7 doi: 10.1090/S0025-5718-1990-1011446-7
    [35] G. Kanschat, D. Schötzau, Energy norm a posteriori error estimation for divergence-free discontinuous Galerkin approximations of the Navier-Stokes equations, Int. J. Numer. Methods Fluids, 57 (2008), 1093–1113. https://doi.org/10.1002/fld.1795 doi: 10.1002/fld.1795
    [36] Y. P. Zeng, F. Wang, A posteriori error estimates for a discontinuous Galerkin approximation of Steklov eigenvalue problems, Appl. Math., 62 (2017), 243–267. https://doi.org/10.21136/AM.2017.0115-16 doi: 10.21136/AM.2017.0115-16
    [37] H. J. Wu, Z. M. Zhang, Can we have superconvergent gradient recovery under adaptive meshes? SIAM J. Numer. Anal., 45 (2007), 1701–1722. https://doi.org/10.1137/060661430 doi: 10.1137/060661430
    [38] Y. D. Yang, Y. Zhang, H. Bi, A type of adaptive C^{0} non-conforming finite element method for the Helmholtz transmission eigenvalue problem, Comput. Methods Appl. Mech. Eng., 360 (2020), 112697. https://doi.org/10.1016/j.cma.2019.112697 doi: 10.1016/j.cma.2019.112697
    [39] Y. J. Li, H. Bi, Y. D. Yang, The a priori and a posteriori error estimates of DG method for the Steklov eigenvalue problem in inverse scattering, J. Sci. Comput., 91 (2022), 20. https://doi.org/10.1007/s10915-022-01787-x doi: 10.1007/s10915-022-01787-x
    [40] X. Y. Dai, J. C. Xu, A. H. Zhou, Convergence and optimal complexity of adaptive finite element eigenvalue computations, Numer. Math., 110 (2008), 313–355. https://doi.org/10.1007/s00211-008-0169-3 doi: 10.1007/s00211-008-0169-3
    [41] W. Dörfler, A convergent adaptive algorithm for Poisson's equation, SIAM J. Numer. Anal., 33 (1996), 1106–1124. https://doi.org/10.1137/0733054 doi: 10.1137/0733054
    [42] P. Morin, R. H. Nochetto, K. G. Siebert, Convergence of adaptive finite element methods, SIAM Rev., 44 (2002), 631–658. https://doi.org/10.1137/s0036144502409093 doi: 10.1137/s0036144502409093
    [43] L. Chen, iFEM: An integrated finite element method package in Matlab, Technical Report, University of California at Irvine, 2009.
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1603) PDF downloads(91) Cited by(0)

Figures and Tables

Figures(12)  /  Tables(8)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog