This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Locking free staggered DG method for the Biot system of poroelasticity on general polygonal meshes

Lina Zhao111Department of Mathematics,The Chinese University of Hong Kong, Hong Kong SAR, China. (lzhao@math.cuhk.edu.hk).   Eric Chung222Department of Mathematics, The Chinese University of Hong Kong, Hong Kong SAR, China. (tschung@math.cuhk.edu.hk).   Eun-Jae Park333School of Mathematics and Computing (Computational Science and Engineering), Yonsei University, Seoul 03722, Korea. (ejpark@yonsei.ac.kr).
Abstract

In this paper we propose and analyze a staggered discontinuous Galerkin method for a five-field formulation of the Biot system of poroelasticity on general polygonal meshes. Elasticity is equipped with stress-displacement-rotation formulation with weak stress symmetry for arbitrary polynomial orders, which extends the piecewise constant approximation developed in (L. Zhao and E.-J. Park, SIAM J. Sci. Comput. 42:A2158–A2181,2020). The proposed method is locking free and can handle highly distorted grids possibly including hanging nodes, which is desirable for practical applications. We prove the convergence estimates for the semi-discrete scheme and fully discrete scheme for all the variables in their natural norms. In particular, the stability and convergence analysis do not need a uniformly positive storativity coefficient. Moreover, to reduce the size of the global system, we propose a five-field formulation based fixed stress splitting scheme, where the linear convergence of the scheme is proved. Several numerical experiments are carried out to confirm the optimal convergence rates and the locking-free property of the proposed method.

Keywords: Staggered DG, General polygonal mesh, Locking-free, Fixed stress splitting, Weak symmetry, Biot system, Poroelasticity

1 Introduction

The Biot system of poroelasticity [3, 41] models fluid flow within deformable porous media, and has been extensively studied in the literature due to its broad range of applications such as geosciences, carbon sequestration and biomedical applications. The system is a fully coupled system, consisting of an equilibrium equation for the solid and a mass balance equation for the fluid. A great amount of effort has been devoted to developing efficient numerical schemes for the Biot system. The two-field displacement-pressure formulation is studied by numerous numerical methods such as finite difference method [20], finite volume method [35], finite element methods [31, 32, 33, 40], coupling of HHO method with SWIP method [4], weak Galerkin method [22] and HDG method [19]. The three-field displacement-pressure-Darcy velocity formulation with various couplings of continuous and discontinuous Galerkin methods, and mixed finite element methods is studied in [37, 38, 45, 44, 36, 29, 47, 27]. Very recently, a stabilized hybrid mixed finite element method uniformly stable with respect to the physical parameters is proposed in [34]. Alternatively, fully-mixed formulations of the Biot system has also drawn great attention [24, 46, 26, 1]. Typical difficulty encountered in the design of numerical schemes for the Biot system is to avoid locking and to get estimates independent of the storativity coefficient.

Therefore, the purpose of this paper is to develop and analyze a staggered discontinuous Galerkin (DG) method on general polygonal meshes for the Biot system of poroelasticity based on a five-field formulation. As a new generation of numerical schemes, staggered DG method offers many salient features such as the easy treatment of general polygonal elements possibly including hanging nodes, the local and global mass conservation and superconvergence. All these features make it a good candidate for practical applications, thus it has been extensively studied for a wide range of partial differential equations arising from physical applications [12, 13, 15, 14, 23, 9, 8, 10, 11, 16, 18, 51, 54, 53, 52, 48, 50, 49]. Therefore, it is of great importance to develop a staggered DG method for the Biot system of poroelasticity. To this end, we will first extend staggered DG method developed in [53] with piecewise constant approximations to arbitrary polynomial orders, where the symmetry of stress is weakly imposed via the carefully designed approximation space for rotation. Then it is combined with staggered DG method proposed in [51] to discretize the Biot system of poroelasticity.

The resulting formulation is fully mixed with stress tensor, the fluid flux, the rotation, the displacement and the pressure as the unknowns. At a first sight this appears to increase the computational burden by introducing additional variables compared to methods that only involve displacement and pressure. However, the introduction of these variables enables us to achieve all the variables of physical interest directly without resorting to postprocessing, which is actually more accurate. We emphasize that the rotation variable introduced here is a scalar field, which can reduce the size of the global system. Another advantage of our method is that no further stabilization or penalty term is needed in contrast to other DG methods. In addition, local conservation is preserved on the dual mesh. It is worth mentioning that the mass matrix is block diagonal, which allows local elimination of the stress, velocity and rotation. As such, we can exploit local elimination to reduce the global saddle point system to displacement-pressure system. Since this procedure is similar to that proposed in [1], where block diagonal matrices can be generated thanks to the vertex quadrature rule utilized, we will not repeat here. Instead, we will introduce fixed stress splitting scheme to reduce the size of the global system. Fixed stress splitting scheme is a popular method for iteratively solving the Biot system, but has not been addressed for the formulation we exploited. Thus, we will propose a fixed stress splitting scheme for our formulation and prove the linear convergence of the given scheme. The main ingredient is to decompose the global system into two subproblem, then we can solve the flow problem and the mechanics problem separately. In particular, as mentioned before, the mass matrix is block diagonal, thus we can also apply local elimination for each problem, consequently, we can obtain a system which is solely based on displacement and pressure.

We analyze the unique solvability, stability and convergence estimates for the semidiscrete continuous-in-time and the fully discrete methods. Note that a different approach from [28] is pursued to prove the inf-sup condition for the elasticity equation. The optimal convergence rates for all the variables in their natural norms are analyzed. In addition, the error estimates are independent of the anisotropy of the permeability of the porous media. More importantly, the stability and convergence estimates are independent of the storativity coefficient c0c_{0} and are valid for c0=0c_{0}=0. Moreover, like the linear elasticity [53], the error bounds are robust for nearly incompressible materials. The numerical experiments including cantilever bracket benchmark problem presented verify the optimal convergence rates and locking free property of the proposed method.

The rest of the paper is organized as follows. In the next section, we describe the model problem and the associated weak formulation. Then in section 3, we derive the discrete formulation for the Biot system of poroelasticity and prove the unique solvability. Convergence analysis for semi-discrete and fully discrete methods is presented in section 4. In section 5, a fixed stress splitting scheme is introduced and analyzed. Several numerical experiments are carried out in section 6 to confirm the proposed theories. Finally, a conclusion is given.

2 Model problem and weak formulation

In this section we introduce the poroelasticity system and its fully mixed weak formulation, which lays foundation for the derivation of our staggered DG scheme. Let Ω2\Omega\subset\mathbb{R}^{2} be a simply connected bounded domain, occupied by a poroelastic media saturated with fluid. Then the stress-strain constitutive relationship for the poroelastic body is given by

Aσe=ϵ(u),\displaystyle A\sigma_{e}=\epsilon(u), (2.1)

where ϵ(u)=12(u+uT)\epsilon(u)=\frac{1}{2}(\nabla u+\nabla u^{T}) and σe=2μϵ(u)+λuI\sigma_{e}=2\mu\epsilon(u)+\lambda\nabla\cdot uI. Here AA is a symmetric, bounded and uniformly positive definite linear operator representing the compliance tensor, σe\sigma_{e} is the elastic strain, uu is the solid displacement.

In the case of a homogeneous and isotropic body, we have

Aσ=12μ(σλ2μ+2λtr(σ)I),\displaystyle A\sigma=\frac{1}{2\mu}(\sigma-\frac{\lambda}{2\mu+2\lambda}tr(\sigma)I),

where II is the 2×22\times 2 identity matrix and μ>0\mu>0, λ0\lambda\geq 0 are the Lamé coefficients. In this case the elastic stress is σe=2μϵ(u)+λuI\sigma_{e}=2\mu\epsilon(u)+\lambda\nabla\cdot uI. The poroelastic stress, which includes the effect of the fluid pressure pp, is given as

σ=σeαpI,\displaystyle\sigma=\sigma_{e}-\alpha pI, (2.2)

where 0<α10<\alpha\leq 1 is the Biot-Willis constant.

Given a vector field ff representing the body forces and a source term qq, the quasi-static Biot system that governs the fluid flow within the poroelastic media is given as follows:

σ=finΩ×[0,T],K1z+p=0inΩ×[0,T],t(c0p+αu)+z=qinΩ×[0,T],\displaystyle\begin{aligned} -\nabla\cdot\sigma&=f\quad&&{\rm in~{}}\Omega\times[0,T],\\ K^{-1}z+\nabla p&=0\quad&&{\rm in~{}}\Omega\times[0,T],\\ \frac{\partial}{\partial t}(c_{0}p+\alpha\nabla\cdot u)+\nabla\cdot z&=q\quad&&{\rm in~{}}\Omega\times[0,T],\end{aligned} (2.3)

where zz is the Darcy velocity, c00c_{0}\geq 0 is a mass storativity coefficient, T>0T>0 is a finite time and KK is a symmetric and positive definite tensor representing the permeability of the porous media divided by the fluid viscosity, i.e., there exist two strictly positive real numbers K1K_{1} and K2K_{2} satisfying for almost every xΩx\in\Omega and all ξ2\xi\in\mathbb{R}^{2} such that |ξ|=1|\xi|=1

0<K1K(x)ξξK2.0<K_{1}\leq K(x)\xi\cdot\xi\leq K_{2}.

Further, we introduce the global anisotropy ratio

ϱB=K2K1.\displaystyle\varrho_{B}=\frac{K_{2}}{K_{1}}. (2.4)

For the sake of simplicity, we assume that the system satisfies homogeneous Dirichlet boundary conditions for uu and pp, i.e., u=0,p=0u=0,p=0 on Ω\partial\Omega and the initial condition p(x,0)=p0p(x,0)=p^{0}. Now we are ready to derive the weak formulation, we have from (2.1) and (2.2)

u=tr(ϵ(u))=tr(Aσe)=trA(σ+αpI),\displaystyle\nabla\cdot u=tr(\epsilon(u))=tr(A\sigma_{e})=trA(\sigma+\alpha pI),

which can be substituted in the third equation of (2.3) to get

t(c0p+αtrA(σ+αpI))+z=q.\displaystyle\partial_{t}(c_{0}p+\alpha trA(\sigma+\alpha pI))+\nabla\cdot z=q.

In the weakly symmetric stress formulation, we allow for σ\sigma to be non-symmetric and introduce the Lagrange multiplier γ=rot(u)/2\gamma=\mbox{rot}(u)/2, rot(u)=u1y+u2x\mbox{rot}(u)=-\frac{\partial u_{1}}{\partial y}+\frac{\partial u_{2}}{\partial x}. The constitutive equation (2.1) can be rewritten as

A(σ+αpI)=uγχ,\displaystyle A(\sigma+\alpha pI)=\nabla u-\gamma\chi,

where

χ=(0110).\displaystyle\chi=\left(\begin{array}[]{cc}0&-1\\ 1&0\\ \end{array}\right).

By the preceding arguments, we can propose the following mixed weak formulation for the Biot system of poroelasticity: Find (σ,u,γ,z,p)L2(Ω)2×2×H01(Ω)2×L2(Ω)×L2(Ω)2×H01(Ω)(\sigma,u,\gamma,z,p)\in L^{2}(\Omega)^{2\times 2}\times H^{1}_{0}(\Omega)^{2}\times L^{2}(\Omega)\times L^{2}(\Omega)^{2}\times H^{1}_{0}(\Omega) such that

(A(σ+αpI),ψ)(u,ψ)+(γ,as(ψ))\displaystyle(A(\sigma+\alpha pI),\psi)-(\nabla u,\psi)+(\gamma,as(\psi)) =0ψL2(Ω)2×2,\displaystyle=0\hskip 25.6073pt\forall\psi\in L^{2}(\Omega)^{2\times 2}, (2.5)
(σ,v)\displaystyle(\sigma,\nabla v) =(f,v)vH01(Ω)2,\displaystyle=(f,v)\quad\forall v\in H^{1}_{0}(\Omega)^{2}, (2.6)
(as(σ),η)\displaystyle(as(\sigma),\eta) =0ηL2(Ω),\displaystyle=0\hskip 28.45274pt\forall\eta\in L^{2}(\Omega), (2.7)
(K1z,ξ)+(p,ξ)\displaystyle(K^{-1}z,\xi)+(\nabla p,\xi) =0ξL2(Ω)2,\displaystyle=0\hskip 28.45274pt\forall\xi\in L^{2}(\Omega)^{2}, (2.8)
c0(tp,w)+α(tA(σ+αpI),wI)(z,w)\displaystyle c_{0}(\partial_{t}p,w)+\alpha(\partial_{t}A(\sigma+\alpha pI),wI)-(z,\nabla w) =(q,w)wH01(Ω),\displaystyle=(q,w)\quad\forall w\in H^{1}_{0}(\Omega), (2.9)

where we use (trAψ,w)=(Aψ,wI)(trA\psi,w)=(A\psi,wI) and as(ψ)=ψ12+ψ21as(\psi)=-\psi_{12}+\psi_{21} for ψL2(Ω)2×2\psi\in L^{2}(\Omega)^{2\times 2}.

For the initial condition we can set z0=Kp0z^{0}=-K\nabla p^{0} which naturally satisfies (2.8). We can also determine (σ0,u0,γ0)(\sigma^{0},u^{0},\gamma^{0}) by solving the elasticity problem (2.5)-(2.7) with p0p^{0} given as initial data. We refer to the initial data obtained by this procedure as compatible initial data.

Before closing this section, we also introduce some notations that will be used later. Let D2D\subset\mathbb{R}^{2}. By (,)D(\cdot,\cdot)_{D}, we denote the scalar product in L2(D):(p,q)D:=Dpq𝑑xL^{2}(D):(p,q)_{D}:=\int_{D}p\;q\;dx. When DD coincides with Ω\Omega, the subscript Ω\Omega will be dropped. We use the same notation for the scalar product in L2(D)2L^{2}(D)^{2} and in L2(D)2×2L^{2}(D)^{2\times 2}. More precisely, (ξ,w)D:=i=12(ξi,wi)(\xi,w)_{D}:=\sum_{i=1}^{2}(\xi^{i},w^{i}) for ξ,wL2(D)2\xi,w\in L^{2}(D)^{2} and (ψ,ζ)D:=i=12j=12(ψi,j,ζi,j)D(\psi,\zeta)_{D}:=\sum_{i=1}^{2}\sum_{j=1}^{2}(\psi^{i,j},\zeta^{i,j})_{D} for ψ,ζL2(D)2×2\psi,\zeta\in L^{2}(D)^{2\times 2}. The associated norm is denoted by 0,D\|\cdot\|_{0,D}. Given an integer m0m\geq 0 and n1n\geq 1, Wm,n(D)W^{m,n}(D) and W0m,n(D)W_{0}^{m,n}(D) denote the usual Sobolev space provided the norm and semi-norm vWm,n(D)={||mDvLn(D)n}1/n\|v\|_{W^{m,n}(D)}=\{\sum_{|\ell|\leq m}\|D^{\ell}v\|^{n}_{L^{n}(D)}\}^{1/n}, |v|Wm,n(D)={||=mDvLn(D)n}1/n|v|_{W^{m,n}(D)}=\{\sum_{|\ell|=m}\|D^{\ell}v\|^{n}_{L^{n}(D)}\}^{1/n}. If n=2n=2 we usually write Hm(D)=Wm,2(D)H^{m}(D)=W^{m,2}(D) and H0m(D)=W0m,2(D)H_{0}^{m}(D)=W_{0}^{m,2}(D), vHm(D)=vWm,2(D)\|v\|_{H^{m}(D)}=\|v\|_{W^{m,2}(D)} and |v|Hm(D)=|v|Wm,2(D)|v|_{H^{m}(D)}=|v|_{W^{m,2}(D)}. In addition, for an integer l0l\geq 0 we define spaces of vector valued functions such as Hl(0,T;Hm(Ω))H^{l}(0,T;H^{m}(\Omega)) and L(0,T;Hm(Ω))L^{\infty}(0,T;H^{m}(\Omega)) with the norms

ψHl(0,T;Hm(Ω))=i=0l(0Tdtiψ(t)Hm(Ω)2𝑑t)1/2,ψL(0,T;Hm(Ω))=max0tTψ(t)Hm(Ω).\displaystyle\|\psi\|_{H^{l}(0,T;H^{m}(\Omega))}=\sum_{i=0}^{l}\Big{(}\int_{0}^{T}\|d_{t}^{i}\psi(t)\|_{H^{m}(\Omega)}^{2}\;dt\Big{)}^{1/2},\quad\|\psi\|_{L^{\infty}(0,T;H^{m}(\Omega))}=\max_{0\leq t\leq T}\|\psi(t)\|_{H^{m}(\Omega)}.

In the sequel, we use CC to denote a positive constant independent of the meshsize and of the anisotropy ratio ϱB\varrho_{B} in (2.4), which may have different values at different occurrences. In addition, for any ψL2(Ω)2×2\psi\in L^{2}(\Omega)^{2\times 2}, (Aψ,ψ)1/2(A\psi,\psi)^{1/2} is a norm equivalent to ψ0\|\psi\|_{0}, which will be denoted by A1/2ψ0\|A^{1/2}\psi\|_{0} throughout the paper.

3 Staggered DG discretization

In this section, we present the staggered DG discretization for the Biot system of poroelasticity. Specifically, a fully mixed formulation will be used, and the symmetry of stress is imposed weakly via the introduction of a lagrange multiplier.

To begin with, we introduce the construction of staggered meshes, which lays foundation for the construction of finite element spaces suitable for our setting. Let 𝒯u\mathcal{T}_{u} be a shape-regular family of matching meshes covering Ω\Omega exactly, where hanging nodes are allowed. Note that the shape-regularity for general polygonal meshes can be described as follows: (1) Every element M𝒯uM\in\mathcal{T}_{u} is star-shaped with respect to a ball of radius ρBhE\geq\rho_{B}h_{E}, where ρB\rho_{B} is a positive constant and hEh_{E} is the diameter of MM; (2) For every element M𝒯uM\in\mathcal{T}_{u} and every edge eMe\subset\partial M, it satisfies heρehMh_{e}\geq\rho_{e}h_{M}, where ρe\rho_{e} is a positive constant and heh_{e} represents the length of edge ee. For an element M𝒯uM\in\mathcal{T}_{u}, we select an interior point ν\nu and connect it to all the vertices of MM. The resulting sub-triangulation is denoted as 𝒯h\mathcal{T}_{h}. We remark that ν\nu is an arbitrary point interior to MM and for simplicity we choose it as center point. The union of all the edges in the primal partition 𝒯u\mathcal{T}_{u} is denoted as u\mathcal{F}_{u}. We use u0\mathcal{F}_{u}^{0} to represent the subset of u\mathcal{F}_{u}, that is the set of edges in u\mathcal{F}_{u} that does not lie on Ω\partial\Omega. In addition, the edges generated during the subdivision process due to the connection of the interior point ν\nu to all the vertices are called dual edges, which is denoted as p\mathcal{F}_{p}. In addition, we define =up\mathcal{F}=\mathcal{F}_{u}\cap\mathcal{F}_{p} and =u0p\mathcal{F}=\mathcal{F}_{u}^{0}\cap\mathcal{F}_{p}. For each triangle τ𝒯h\tau\in\mathcal{T}_{h}, we let hτh_{\tau} be the diameter of τ\tau and h=max{hτ,τ𝒯h}h=\max\{h_{\tau},\tau\in\mathcal{T}_{h}\}. Next, we introduce the dual mesh. For each interior edge eu0e\in\mathcal{F}_{u}^{0}, we use D(e)D(e) to represent the dual mesh, which is the union of the two triangles in 𝒯h\mathcal{T}_{h} sharing the common edge ee. For each edge eu\u0e\in\mathcal{F}_{u}\backslash\mathcal{F}_{u}^{0}, we use D(e)D(e) to denote the triangle in 𝒯h\mathcal{T}_{h} having the edge ee, see Figure 1 for an illustration. For each edge ee, we define a unit normal vector nen_{e} as follows: If ee is an interior edge, then nen_{e} is the unit normal vector of ee pointing towards the outside of Ω\Omega. If ee is an interior edge, we then fix nen_{e} as one of the two possible unit normal vectors on ee. When there is no ambiguity, we use nn instead of nen_{e} to simplify the notation. Let k0k\geq 0 be the order of approximation, for every τ𝒯h\tau\in\mathcal{T}_{h} and ee\in\mathcal{F}, we define Pk(τ)P^{k}(\tau) and Pk(e)P^{k}(e) as the spaces of polynomials of degree less than or equal to kk on τ\tau and ee, respectively.

Refer to caption
Figure 1: Schematic of the primal mesh and the dual mesh. Solid lines represent edges in u\mathcal{F}_{u} and dashed lines represent edges in p\mathcal{F}_{p}.

For a scalar, vector, or tensor function vv that is double-valued on an interior edge e0e\in\mathcal{F}^{0}, its jump and average on ee are defined as

[v]=vτ1vτ2,{v}=12(vτ1+vτ2),\displaystyle[v]=v\mid_{\tau_{1}}-v\mid_{\tau_{2}},\quad\{v\}=\frac{1}{2}(v\mid_{\tau_{1}}+v\mid_{\tau_{2}}),

where τ1\tau_{1} and τ2\tau_{2} are the two triangles belonging to 𝒯h\mathcal{T}_{h} which share the common edge ee. We set [v]=vτ1[v]=v\mid_{\tau_{1}} and {v}=vτ1\{v\}=v\mid_{\tau_{1}} on boundary edges.

Now we are ready to introduce the finite element spaces that will be used for the construction of our method. First, the locally H1(Ω)H^{1}(\Omega)-conforming SDG space ShS_{h} is defined by

Sh:={wL2(Ω):wτPk(τ),[w]e=0eu;wΩ=0},S_{h}:=\{w\in L^{2}(\Omega):w\mid_{\tau}\in P^{k}(\tau),[w]\mid_{e}=0\;\forall e\in\mathcal{F}_{u};w\mid_{\partial\Omega}=0\},

which is equipped by the following norm

wZ2=τ𝒯hw0,τ2+ephe1[w]0,e2.\displaystyle\|w\|_{Z}^{2}=\sum_{\tau\in\mathcal{T}_{h}}\|\nabla w\|_{0,\tau}^{2}+\sum_{e\in\mathcal{F}_{p}}h_{e}^{-1}\|[w]\|_{0,e}^{2}.

For a function v=(v1,v2)[Sh]2v=(v_{1},v_{2})\in[S_{h}]^{2} we define vh2=v1Z2+v2Z2\|v\|_{h}^{2}=\|v_{1}\|_{Z}^{2}+\|v_{2}\|_{Z}^{2}.

Next, the locally H(div;Ω)H(\mbox{div};\Omega)-conforming SDG space Σh\Sigma_{h} is defined by

Σh:={ξL2(Ω)2:ξτPk(τ)2τ𝒯h,[ξn]e=0ep}.\Sigma_{h}:=\{\xi\in L^{2}(\Omega)^{2}:\xi\mid_{\tau}\in P^{k}(\tau)^{2}\;\forall\tau\in\mathcal{T}_{h},[\xi\cdot n]\mid_{e}=0\;\forall e\in\mathcal{F}_{p}\}.

Finally, the locally H1(Ω)H^{1}(\Omega)-conforming finite element space is defined as

Mh:={ηL2(Ω):ητPk(τ)τ𝒯h,[η]e=0ep}.M_{h}:=\{\eta\in L^{2}(\Omega):\eta\mid_{\tau}\in P^{k}(\tau)\;\forall\tau\in\mathcal{T}_{h},[\eta]\mid_{e}=0\;\forall e\in\mathcal{F}_{p}\}.

We can derive the staggered DG discretization for the Biot system of poroelasticity by following [51, 53], which reads as follows: Find (σh,uh,γh,zh,ph)[Σh]2×[Sh]2×Mh×Σh×Sh(\sigma_{h},u_{h},\gamma_{h},z_{h},p_{h})\in[\Sigma_{h}]^{2}\times[S_{h}]^{2}\times M_{h}\times\Sigma_{h}\times S_{h} such that

(A(σh+αphI),ψ)Bh(uh,ψ)+(γh,as(ψ))\displaystyle(A(\sigma_{h}+\alpha p_{h}I),\psi)-B_{h}^{*}(u_{h},\psi)+(\gamma_{h},as(\psi)) =0ψ[Σh]2,\displaystyle=0\hskip 25.6073pt\forall\psi\in[\Sigma_{h}]^{2}, (3.1)
Bh(σh,v)\displaystyle B_{h}(\sigma_{h},v) =(f,v)v[Sh]2,\displaystyle=(f,v)\quad\forall v\in[S_{h}]^{2}, (3.2)
(as(σh),η)\displaystyle(as(\sigma_{h}),\eta) =0ηMh,\displaystyle=0\hskip 28.45274pt\forall\eta\in M_{h}, (3.3)
(K1zh,ξ)+bh(ph,ξ)\displaystyle(K^{-1}z_{h},\xi)+b_{h}^{*}(p_{h},\xi) =0ξΣh,\displaystyle=0\hskip 29.87547pt\forall\xi\in\Sigma_{h}, (3.4)
c0(tph,w)+α(tA(σh+αphI),wI)bh(zh,w)\displaystyle c_{0}(\partial_{t}p_{h},w)+\alpha(\partial_{t}A(\sigma_{h}+\alpha p_{h}I),wI)-b_{h}(z_{h},w) =(q,w)wSh,\displaystyle=(q,w)\quad\forall w\in S_{h}, (3.5)

where

Bh(ψ,v)\displaystyle B_{h}(\psi,v) =ep(ψn,[v])e+(ψ,v)(ψ,v)[Σh]2×[Sh]2,\displaystyle=-\sum_{e\in\mathcal{F}_{p}}(\psi n,[v])_{e}+(\psi,\nabla v)\quad\forall(\psi,v)\in[\Sigma_{h}]^{2}\times[S_{h}]^{2},
Bh(v,ψ)\displaystyle B_{h}^{*}(v,\psi) =eu0(v,[ψn])e(v,ψ)(ψ,v)[Σh]2×[Sh]2\displaystyle=\sum_{e\in\mathcal{F}_{u}^{0}}(v,[\psi n])_{e}-(v,\nabla\cdot\psi)\quad\forall(\psi,v)\in[\Sigma_{h}]^{2}\times[S_{h}]^{2}

and

bh(ξ,w)\displaystyle b_{h}(\xi,w) =ep(ξn,[w])e+(ξ,w)(ξ,w)Σh×Sh,\displaystyle=-\sum_{e\in\mathcal{F}_{p}}(\xi\cdot n,[w])_{e}+(\xi,\nabla w)\quad\forall(\xi,w)\in\Sigma_{h}\times S_{h},
bh(w,ξ)\displaystyle b_{h}^{*}(w,\xi) =eu0(w,[ξn])e(w,ξ)(ξ,w)Σh×Sh.\displaystyle=\sum_{e\in\mathcal{F}_{u}^{0}}(w,[\xi\cdot n])_{e}-(w,\nabla\cdot\xi)\quad\forall(\xi,w)\in\Sigma_{h}\times S_{h}.

The bilinear forms defined above satisfy the following adjoint properties

bh(ξ,w)\displaystyle b_{h}(\xi,w) =bh(w,ξ)(w,ξ)Sh×Σh,\displaystyle=b_{h}^{*}(w,\xi)\quad\forall(w,\xi)\in S_{h}\times\Sigma_{h},
Bh(ψ,v)\displaystyle B_{h}(\psi,v) =Bh(v,ψ)(ψ,v)[Σh]2×[Sh]2.\displaystyle=B_{h}^{*}(v,\psi)\quad\forall(\psi,v)\in[\Sigma_{h}]^{2}\times[S_{h}]^{2}.

In addition, the following inf-sup condition holds; cf. [13].

wZCsupξΣhbh(w,ξ)ξ0wSh\displaystyle\|w\|_{Z}\leq C\sup_{\xi\in\Sigma_{h}}\frac{b_{h}^{*}(w,\xi)}{\|\xi\|_{0}}\quad\forall w\in S_{h} (3.6)

and

vhCsupψ[Σh]2Bh(v,ψ)ψ0v[Sh]2.\displaystyle\|v\|_{h}\leq C\sup_{\psi\in[\Sigma_{h}]^{2}}\frac{-B_{h}^{*}(v,\psi)}{\|\psi\|_{0}}\quad\forall v\in[S_{h}]^{2}. (3.7)
Remark 3.1.

(local conservation on the dual mesh). Taking v=1v=1 in (3.2) to be identically one over the dual mesh D(e),euD(e),\forall e\in\mathcal{F}_{u}, we have

(σhnD,1)D(e)=(f,1)D(e),\displaystyle-(\sigma_{h}n_{D},1)_{\partial D(e)}=(f,1)_{D(e)},

where nDn_{D} is the outward unit normal vector of D(e)D(e). This property is the crux for a posteriori estimates based on equilibrated fluxes; cf. [51, 54].

Now we prove the inf-sup condition, which is the key to proving the stability and convergence estimates of the proposed method. The methodology exploited here is different from that of [28]. Indeed the norm used here bypasses the construction of divergence free functions in proving the convergence estimates.

Lemma 3.1.

The following inf-sup condition holds

supψh[Σh]2Bh(vh,ψh)+(ηh,as(ψh))ψh0C(vhh2+ηh02)12,(vh,ηh)[Sh]2×Mh.\displaystyle\sup_{\psi_{h}\in[\Sigma_{h}]^{2}}\frac{-B_{h}^{*}(v_{h},\psi_{h})+(\eta_{h},as(\psi_{h}))}{\|\psi_{h}\|_{0}}\geq C(\|v_{h}\|_{h}^{2}+\|\eta_{h}\|_{0}^{2})^{\frac{1}{2}},\quad\forall(v_{h},\eta_{h})\in[S_{h}]^{2}\times M_{h}. (3.8)
Proof.

To prove the inf-sup condition (3.8), we apply Fortin interpolation operator (cf. [5]), namely, we need to build an interpolation operator Jh:H1(Ω)2×2[Σh]2J_{h}:H^{1}(\Omega)^{2\times 2}\rightarrow[\Sigma_{h}]^{2} such that it satisfies for any ψH1(Ω)2×2\psi\in H^{1}(\Omega)^{2\times 2}

Bh(vh,ψJhψ)+(ηh,as(ψJhψ))\displaystyle-B_{h}^{*}(v_{h},\psi-J_{h}\psi)+(\eta_{h},as(\psi-J_{h}\psi)) =0(vh,ηh)[Sh]2×Mh,\displaystyle=0\quad\forall(v_{h},\eta_{h})\in[S_{h}]^{2}\times M_{h},
Jhψ0\displaystyle\|J_{h}\psi\|_{0} Cψ0.\displaystyle\leq C\|\psi\|_{0}.

To this end we will first build ψ1[Σh]2\psi_{1}\in[\Sigma_{h}]^{2} such that it satisfies

Bh(vh,ψ1ψ)=0vh[Sh]2,ψ10Cψ0,\begin{split}B_{h}^{*}(v_{h},\psi_{1}-\psi)&=0\quad\forall v_{h}\in[S_{h}]^{2},\\ \|\psi_{1}\|_{0}&\leq C\|\psi\|_{0},\end{split} (3.9)

then we correct ψ1\psi_{1} by using ψ2[Σh]2\psi_{2}\in[\Sigma_{h}]^{2} which satisfies Bh(vh,ψ2)=0,vh[Sh]2B_{h}^{*}(v_{h},\psi_{2})=0,\;\forall v_{h}\in[S_{h}]^{2}, in addition, there holds

(as(ψψ2),ηh)\displaystyle(as(\psi-\psi_{2}),\eta_{h}) =(as(ψ1),ηh)ηhMh,\displaystyle=(as(\psi_{1}),\eta_{h})\quad\forall\eta_{h}\in M_{h}, (3.10)
ψ20\displaystyle\|\psi_{2}\|_{0} Cψψ10.\displaystyle\leq C\|\psi-\psi_{1}\|_{0}. (3.11)

Since Bh(,)B_{h}^{*}(\cdot,\cdot) satisfies the inf-sup condition (3.7), which means there exists ψ1[Σh]2\psi_{1}\in[\Sigma_{h}]^{2} that satisfies

Bh(vh,ψψ1)=0vh[Sh]2,ψ10Cψ0.\begin{split}B_{h}^{*}(v_{h},\psi-\psi_{1})&=0\quad\forall v_{h}\in[S_{h}]^{2},\\ \|\psi_{1}\|_{0}&\leq C\|\psi\|_{0}.\end{split}

Next, we correct ψ1\psi_{1} in the following way. We define the following spaces for each primal element M𝒯uM\in\mathcal{T}_{u}:

Vhk(M)\displaystyle V_{h}^{k}(M) ={ςMH01(M)2:ςτPk(τ)2,τM},\displaystyle=\{\varsigma\mid_{M}\in H^{1}_{0}(M)^{2}:\varsigma\mid_{\tau}\in P^{k}(\tau)^{2},\tau\subset M\},
Qhk(M)\displaystyle Q_{h}^{k}(M) ={qML2(M):qτPk(τ),τM,Mqdx=0}.\displaystyle=\{q\mid_{M}\in L^{2}(M):q\mid_{\tau}\in P^{k}(\tau),\tau\subset M,\int_{M}q\;dx=0\}.

For each primal element M𝒯uM\in\mathcal{T}_{u}, we consider the stable approximation Vhk+1(M)×Qhk(M)V_{h}^{k+1}(M)\times Q_{h}^{k}(M) for the Stokes element such that

curlVhk+1(M)Σh(M),\displaystyle\mbox{curl}V_{h}^{k+1}(M)\subset\Sigma_{h}(M),

where Σh(M)\Sigma_{h}(M) denotes Σh\Sigma_{h} restricted to the primal element MM. We then solve for (ψhM,phM)Vhk+1(M)×Qhk(M)(\psi_{h}^{*}\mid_{M},p_{h}^{*}\mid_{M})\in V_{h}^{k+1}(M)\times Q_{h}^{k}(M) such that

(2μϵ(ψh),ϵ(ϕh))M+(ph,ϕh)M\displaystyle(2\mu\epsilon(\psi_{h}^{*}),\epsilon(\phi_{h}))_{M}+(p_{h}^{*},\nabla\cdot\phi_{h})_{M} =(f,ϕh)MϕhVhk+1(M),\displaystyle=(f,\phi_{h})_{M}\quad\forall\phi_{h}\in V_{h}^{k+1}(M),
(qh,ψh)M\displaystyle(q_{h},\nabla\cdot\psi_{h}^{*})_{M} =(as(ψψ1),qh)MqhQhk(M).\displaystyle=(as(\psi-\psi_{1}),q_{h})_{M}\quad\forall q_{h}\in Q_{h}^{k}(M).

Since ψh=0\psi_{h}^{*}=0 on M\partial M it is easy to check that the second equation also satisfies for a constant function. Here we remark that the existing stable pair that fits into the above spaces can be Taylor-Hood element (cf. [7]).

Let ψ2M=curlψhM\psi_{2}\mid_{M}=-\mbox{curl}\psi_{h}^{*}\mid_{M}, then we can check that

(as(ψψ2),ηh)M\displaystyle(as(\psi-\psi_{2}),\eta_{h})_{M} =(as(ψ1),ηh)ηhMh(M),\displaystyle=(as(\psi_{1}),\eta_{h})\quad\forall\eta_{h}\in M_{h}(M),
ψ20,M\displaystyle\|\psi_{2}\|_{0,M} Cψψ10,M,\displaystyle\leq C\|\psi-\psi_{1}\|_{0,M},

which satisfies (3.10) by summing over all the primal elements. Here Mh(M)M_{h}(M) denotes the space MhM_{h} restricted to each primal element M𝒯uM\in\mathcal{T}_{u}. In addition, it also holds

Bh(vh,ψ2)=0vh[Sh]2.\displaystyle B_{h}^{*}(v_{h},\psi_{2})=0\quad\forall v_{h}\in[S_{h}]^{2}.

Therefore, we have built ψ1\psi_{1} and ψ2\psi_{2} which satisfy (3.9)-(3.11). Hence, by the well-known arguments from [5] we can infer that the existence of ψ1\psi_{1} and ψ2\psi_{2} leads to the inf-sup condition.

Remark 3.2.

In the above proof, we employ Taylor-Hood stable pair, in fact we can also exploit Scott-Vogelius element. In which case, we need to be careful about the singular nodes, indeed if singular nodes exists then one needs to enforce additional restriction on those nodes (cf. [21]), which is naturally satisfied for our choice of space MhM_{h}.

We also emphasize that we choose locally H1H^{1}-conforming space for γh\gamma_{h} to enforce the symmetry of stress, which is different from that of [28]. In fact, we can also exploit fully discontinuous space for γh\gamma_{h}, which then leads to strongly symmetric stress. In this case, the final system will be singular for rectangular mesh if one chooses the center point as ν\nu. One can perturb the position of ν\nu to avoid singularity. We want to develop a numerical scheme which is stable regardless of the position of ν\nu, thus we choose locally H1H^{1}-conforming space for γh\gamma_{h}.

In order to prove the well-posedness of (3.1)-(3.5), we will make use of the following theorem; cf. [43].

Theorem 3.1.

Let the linear, symmetric and monotone operators 𝒩\mathcal{N} be given for the real vector space EE to its dual algebraic dual EE^{*}, and let EbE_{b}^{\prime} be the Hilbert space which is the dual of EE with the seminorm

|x|b=(𝒩x(x))12,xE.\displaystyle|x|_{b}=(\mathcal{N}x(x))^{\frac{1}{2}},\quad\forall x\in E.

Let E×Eb\mathcal{M}\subset E\times E_{b}^{\prime} be a relation with domain D={xE:(x)0}D=\{x\in E:\mathcal{M}(x)\neq 0\}. Assume \mathcal{M} is monotone and Rg(𝒩+)=EbRg(\mathcal{N}+\mathcal{M})=E_{b}^{\prime}. Then for each x0Dx_{0}\in D and for each W1,1(0,T;Eb)\mathcal{F}\in W^{1,1}(0,T;E_{b}^{\prime}), there is a solution xx of

ddt(𝒩x(t))+(x(t))(t), 0<t<T\displaystyle\frac{d}{dt}(\mathcal{N}x(t))+\mathcal{M}(x(t))\ni\mathcal{F}(t),\quad\forall\;0<t<T (3.12)

with

𝒩xW1,(0,T;Eb),x(t)D, 0tT,and𝒩x(0)=𝒩x0.\displaystyle\mathcal{N}x\in W^{1,\infty}(0,T;E_{b}^{\prime}),\;x(t)\in D,\;\forall\;0\leq t\leq T,\quad\mbox{and}\quad\mathcal{N}x(0)=\mathcal{N}x_{0}.
Theorem 3.2.

There exists a unique solution to (3.1)-(3.5).

Proof.

In order to fit (3.1)-(3.5) in the form of Theorem 3.1, we consider a slightly modified formulation, with (3.1) differentiated in time and the new variables u˙h\dot{u}_{h} and γh˙\dot{\gamma_{h}} representing dtuhd_{t}u_{h} and dtγhd_{t}\gamma_{h}, respectively:

(tA(σh+αphI),ψ)Bh(u˙h,ψ)+(γ˙h,as(ψ))\displaystyle(\partial_{t}A(\sigma_{h}+\alpha p_{h}I),\psi)-B_{h}^{*}(\dot{u}_{h},\psi)+(\dot{\gamma}_{h},as(\psi)) =0ψ[Σh]2.\displaystyle=0\quad\forall\psi\in[\Sigma_{h}]^{2}. (3.13)

We now introduce the operators

(Aσσσh,ψ)=(Aσh,ψ),(Aσpσh,w)=(Aσh,αwI),(Aσuσh,v)=Bh(σh,v),\displaystyle(A_{\sigma\sigma}\sigma_{h},\psi)=(A\sigma_{h},\psi),\quad(A_{\sigma p}\sigma_{h},w)=(A\sigma_{h},\alpha wI),\quad(A_{\sigma u}\sigma_{h},v)=B_{h}(\sigma_{h},v),
(Aσγσh,η)=(as(σh),η),(Azzzh,ξ)=(K1zh,ξ),(Azpzh,w)=bh(zh,w),\displaystyle(A_{\sigma\gamma}\sigma_{h},\eta)=(as(\sigma_{h}),\eta),\quad(A_{zz}z_{h},\xi)=(K^{-1}z_{h},\xi),\quad(A_{zp}z_{h},w)=b_{h}(z_{h},w),
(Appph,w)=c0(ph,w)+α(AαphI,wI).\displaystyle(A_{pp}p_{h},w)=c_{0}(p_{h},w)+\alpha(A\alpha p_{h}I,wI).

We have a system in the form of (3.12), where

x˙=(σhu˙hγ˙hzhph),𝒩=(Aσσ000AσpT000000000000000Aσp000App),=(0AσuTAσγT00Aσu0000Aσγ0000000AzzAzpT000Azp0)\displaystyle\dot{x}=\left(\begin{array}[]{c}\sigma_{h}\\ \dot{u}_{h}\\ \dot{\gamma}_{h}\\ z_{h}\\ p_{h}\\ \end{array}\right),\;\mathcal{N}=\left(\begin{array}[]{ccccc}A_{\sigma\sigma}&0&0&0&A_{\sigma p}^{T}\\ 0&0&0&0&0\\ 0&0&0&0&0\\ 0&0&0&0&0\\ A_{\sigma p}&0&0&0&A_{pp}\\ \end{array}\right),\;\mathcal{M}=\left(\begin{array}[]{ccccc}0&-A_{\sigma u}^{T}&A_{\sigma\gamma}^{T}&0&0\\ A_{\sigma u}&0&0&0&0\\ A_{\sigma\gamma}&0&0&0&0\\ 0&0&0&A_{zz}&A_{zp}^{T}\\ 0&0&0&-A_{zp}&0\\ \end{array}\right)

and =(0,f,0,0,q)T\mathcal{F}=(0,f,0,0,q)^{T}.

The dual space EbE_{b}^{\prime} is L2(Ω)2×2×0×0×0×L2(Ω)L^{2}(\Omega)^{2\times 2}\times 0\times 0\times 0\times L^{2}(\Omega), and the condition W1,1(0,T;Eb)\mathcal{F}\in W^{1,1}(0,T;E_{b}^{\prime}) in Theorem 3.1 allows for non-zero source terms only in the equations with time derivatives, which means f=0f=0. We can reduce our problem to a system with f=0f=0 by solving for each t(0,T]t\in(0,T] an elasticity problem with a source term ff, cf. [42].

(Aσhf,ψ)Bh(u˙hf,ψ)+(γ˙hf,as(ψ))=0ψ[Σh]2,Bh(σhf,v)=(f,v)v[Sh]2,(as(σhf),η)=0ηMh.\displaystyle\begin{aligned} (A\sigma_{h}^{f},\psi)-B_{h}^{*}(\dot{u}_{h}^{f},\psi)+(\dot{\gamma}^{f}_{h},as(\psi))&=0\quad&&\forall\psi\in[\Sigma_{h}]^{2},\\ B_{h}(\sigma_{h}^{f},v)&=(f,v)\quad&&\forall v\in[S_{h}]^{2},\\ (as(\sigma_{h}^{f}),\eta)&=0\quad&&\forall\eta\in M_{h}.\end{aligned}

Subtracting the above equations from (3.1)-(3.5), we can obtain

(A((σhσ˙hf)+αphI),ψ)Bh(uhu˙hf,ψ)+(γhγ˙hf,as(ψ))\displaystyle(A((\sigma_{h}-\dot{\sigma}_{h}^{f})+\alpha p_{h}I),\psi)-B_{h}^{*}(u_{h}-\dot{u}_{h}^{f},\psi)+(\gamma_{h}-\dot{\gamma}_{h}^{f},as(\psi)) =(A(σhftσhf),ψ),\displaystyle=(A(\sigma_{h}^{f}-\partial_{t}\sigma_{h}^{f}),\psi), (3.14)
Bh(σhσhf,v)\displaystyle B_{h}(\sigma_{h}-\sigma_{h}^{f},v) =0,\displaystyle=0, (3.15)
(as(σhσhf),η)\displaystyle(as(\sigma_{h}-\sigma_{h}^{f}),\eta) =0,\displaystyle=0, (3.16)
(K1zh,ξ)+bh(ph,ξ)\displaystyle(K^{-1}z_{h},\xi)+b_{h}^{*}(p_{h},\xi) =0,\displaystyle=0, (3.17)
c0(tph,w)+α(tA((σhσhf)+αphI),wI)bh(zh,w)\displaystyle c_{0}(\partial_{t}p_{h},w)+\alpha(\partial_{t}A((\sigma_{h}-\sigma_{h}^{f})+\alpha p_{h}I),wI)-b_{h}(z_{h},w) =(q,w)(αtAσhf,wI),\displaystyle=(q,w)-(\alpha\partial_{t}A\sigma_{h}^{f},wI), (3.18)

which results in a problem with a modified right hand side =(Aσσ(σhftσhf),0,0,0,qAσptσhf)T\mathcal{F}=(A_{\sigma\sigma}(\sigma_{h}^{f}-\partial_{t}\sigma_{h}^{f}),0,0,0,q-A_{\sigma p}\partial_{t}\sigma_{h}^{f})^{T}.

The range condition Rg(𝒩+)=EbRg(\mathcal{N}+\mathcal{M})=E_{b}^{\prime} can be verified by showing that the square finite dimensional homogeneous system: Find (σ^h,u^h,γ^h,z^h,p^h)[Σh]2×[Sh]2×Mh×Σh×Sh(\hat{\sigma}_{h},\hat{u}_{h},\hat{\gamma}_{h},\hat{z}_{h},\hat{p}_{h})\in[\Sigma_{h}]^{2}\times[S_{h}]^{2}\times M_{h}\times\Sigma_{h}\times S_{h} such that

(A(σ^h+αp^hI),ψ)Bh(u^h,ψ)+(γ^h,as(ψ))\displaystyle(A(\hat{\sigma}_{h}+\alpha\hat{p}_{h}I),\psi)-B_{h}^{*}(\hat{u}_{h},\psi)+(\hat{\gamma}_{h},as(\psi)) =0ψ[Σh]2,\displaystyle=0\quad\forall\psi\in[\Sigma_{h}]^{2}, (3.19)
Bh(σ^h,v)\displaystyle B_{h}(\hat{\sigma}_{h},v) =0v[Sh]2,\displaystyle=0\quad\forall v\in[S_{h}]^{2}, (3.20)
(as(σ^h),η)\displaystyle(as(\hat{\sigma}_{h}),\eta) =0ηMh,\displaystyle=0\quad\forall\eta\in M_{h}, (3.21)
(K1z^h,ξ)+bh(p^h,ξ)\displaystyle(K^{-1}\hat{z}_{h},\xi)+b_{h}^{*}(\hat{p}_{h},\xi) =0ξΣh,\displaystyle=0\quad\forall\xi\in\Sigma_{h}, (3.22)
c0(tp^h,w)+α(tA(σ^h+αp^hI),wI)bh(z^h,w)\displaystyle c_{0}(\partial_{t}\hat{p}_{h},w)+\alpha(\partial_{t}A(\hat{\sigma}_{h}+\alpha\hat{p}_{h}I),wI)-b_{h}(\hat{z}_{h},w) =0wSh,\displaystyle=0\quad\forall w\in S_{h}, (3.23)

has only zero solution. Taking (ψ,v,η,ξ,w)=(σ^h,u^h,γ^h,z^h,p^h)(\psi,v,\eta,\xi,w)=(\hat{\sigma}_{h},\hat{u}_{h},\hat{\gamma}_{h},\hat{z}_{h},\hat{p}_{h}) in (3.19)-(3.23) and summing up the resulting equations yield

A12(σ^h+αp^hI)02+c0p^h02+K12z^h02=0,\displaystyle\|A^{\frac{1}{2}}(\hat{\sigma}_{h}+\alpha\hat{p}_{h}I)\|_{0}^{2}+c_{0}\|\hat{p}_{h}\|_{0}^{2}+\|K^{-\frac{1}{2}}\hat{z}_{h}\|_{0}^{2}=0,

which gives σ^h+αp^hI=0\hat{\sigma}_{h}+\alpha\hat{p}_{h}I=0 and z^h=0\hat{z}_{h}=0. By the inf-sup condition (3.6), we have

p^hZCz^h0,\displaystyle\|\hat{p}_{h}\|_{Z}\leq C\|\hat{z}_{h}\|_{0},

thereby p^h=0\hat{p}_{h}=0 by using the discrete Poincaré-Friedrichs inequality (cf. [6]), consequently, σ^h=0\hat{\sigma}_{h}=0. The inf-sup condition (3.8) and (3.19) lead to

u^hh+γ^h0CA12(σ^h+αp^hI)0,\displaystyle\|\hat{u}_{h}\|_{h}+\|\hat{\gamma}_{h}\|_{0}\leq C\|A^{\frac{1}{2}}(\hat{\sigma}_{h}+\alpha\hat{p}_{h}I)\|_{0},

which yields u^h=0\hat{u}_{h}=0 and γ^h=0\hat{\gamma}_{h}=0.

Finally, we need compatible initial data x˙0D\dot{x}_{0}\in D, i.e., x˙0Eb\mathcal{M}\dot{x}_{0}\in E_{b}^{\prime}. Let us first consider the initial data x0=(σh0,uh0,γh0,zh0,ph0)x^{0}=(\sigma_{h}^{0},u_{h}^{0},\gamma_{h}^{0},z_{h}^{0},p_{h}^{0}) for the discrete formulation (3.1)-(3.5). We take x0x^{0} to be the elliptic projection of the initial data x~0=(σ0,u0,γ0,z0,p0)\tilde{x}_{0}=(\sigma^{0},u^{0},\gamma^{0},z^{0},p^{0}) for the weak formulation (2.5)-(2.9), which is constructed from p0p^{0} by the procedure described at the end of section 2. With the reduction to a problem with f=0f=0, the construction satisfies (𝒩+)x~0Eb(\mathcal{N}+\mathcal{M})\tilde{x}^{0}\in E_{b}^{\prime}. Then by

(𝒩+)x0=(𝒩+)x~0,\displaystyle(\mathcal{N}+\mathcal{M})x^{0}=(\mathcal{N}+\mathcal{M})\tilde{x}^{0},

we have x0=(𝒩+)x~0𝒩x0Eb\mathcal{M}x^{0}=(\mathcal{N}+\mathcal{M})\tilde{x}^{0}-\mathcal{N}x^{0}\in E_{b}^{\prime}. For the initial data of the differentiated problem (3.13), (3.2)-(3.5), we simply take x˙0=(σh0,0,0,zh0,ph0)\dot{x}^{0}=(\sigma_{h}^{0},0,0,z_{h}^{0},p_{h}^{0}), which also satisfies x˙0Eb\mathcal{M}\dot{x}^{0}\in E_{b}^{\prime}. Here uh0u_{h}^{0} and γh0\gamma_{h}^{0} are not needed for the differentiated problem, but will be used to recover the solution of the original problem.

Now all conditions of Theorem 3.1 are satisfied and we conclude the existence of a solution of (3.13), (3.2)-(3.5) with σh(0)=σh0\sigma_{h}(0)=\sigma_{h}^{0} and ph(0)=ph0p_{h}(0)=p_{h}^{0}. By taking t0t\rightarrow 0 in (3.4) and using that zh0z_{h}^{0} and ph0p_{h}^{0} satisfy (3.4) at t=0t=0, we also infer that zh(0)=zh0z_{h}(0)=z_{h}^{0}.

Next, we recover the solution of the original problem. Let us define

uh(t)=uh0+0tu˙h𝑑s,γh(t)=γh0+0tγ˙h𝑑st(0,T].\displaystyle u_{h}(t)=u_{h}^{0}+\int_{0}^{t}\dot{u}_{h}\;ds,\quad\gamma_{h}(t)=\gamma_{h}^{0}+\int_{0}^{t}\dot{\gamma}_{h}\;ds\quad\forall t\in(0,T].

By construction, uh(0)=uh0u_{h}(0)=u_{h}^{0} and γh(0)=γh0\gamma_{h}(0)=\gamma_{h}^{0}. Integrating (3.13) from 0 to any t(0,T]t\in(0,T] and using that σh0,uh0,γh0\sigma_{h}^{0},u_{h}^{0},\gamma_{h}^{0} satisfy (3.1) at t=0t=0, we conclude that (3.1) holds for all tt. This completes the existence proof. Uniqueness follows from the stability bound given in Theorem 4.1 in the next section.

Before closing this section, we introduce some interpolation operators which will play an important role for later analysis. There exists the interpolation operator Ih:H1(Ω)ShI_{h}:H^{1}(\Omega)\rightarrow S_{h} satisfying for any ωH1(Ω)\omega\in H^{1}(\Omega)

(Ihωω,ψ)e=0ψPk(e),eu,(Ihωω,ψ)τ=0ψPk1(τ),τ𝒯h.\begin{split}(I_{h}\omega-\omega,\psi)_{e}&=0\quad\forall\psi\in P^{k}(e),e\in\mathcal{F}_{u},\\ (I_{h}\omega-\omega,\psi)_{\tau}&=0\quad\forall\psi\in P^{k-1}(\tau),\tau\in\mathcal{T}_{h}.\end{split} (3.24)

Similarly, there also exists an operator Πh:[Hδ(Ω)]2Σh,δ>1/2\Pi_{h}:[H^{\delta}(\Omega)]^{2}\rightarrow\Sigma_{h},\delta>1/2 such that for any φ[Hδ(Ω)]2\varphi\in[H^{\delta}(\Omega)]^{2}

((Πhφφ)n,ϕ)e=0ϕPk(e),ep,(Πhφφ,ϕ)τ=0ϕPk1(τ)2,τ𝒯h.\begin{split}((\Pi_{h}\varphi-\varphi)\cdot n,\phi)_{e}&=0\quad\forall\phi\in P^{k}(e),e\in\mathcal{F}_{p},\\ (\Pi_{h}\varphi-\varphi,\phi)_{\tau}&=0\quad\forall\phi\in P^{k-1}(\tau)^{2},\tau\in\mathcal{T}_{h}.\end{split} (3.25)

The definition of the interpolation operators implies that

bh(zΠhz,w)\displaystyle b_{h}(z-\Pi_{h}z,w) =0wSh,\displaystyle=0\quad\forall w\in S_{h}, (3.26)
bh(pIhp,v)\displaystyle b_{h}^{*}(p-I_{h}p,v) =0vΣh.\displaystyle=0\quad\forall v\in\Sigma_{h}. (3.27)

In addition for Πhσ=(Πhσ1,Πhσ2)[Σh]2\Pi_{h}\sigma=(\Pi_{h}\sigma_{1},\Pi_{h}\sigma_{2})\in[\Sigma_{h}]^{2} and Ihu=(Ihu1,Ihu2)[Sh]2I_{h}u=(I_{h}u_{1},I_{h}u_{2})\in[S_{h}]^{2}, we also have

Bh(σΠhσ,v)\displaystyle B_{h}(\sigma-\Pi_{h}\sigma,v) =0v[Sh]2,\displaystyle=0\quad\forall v\in[S_{h}]^{2}, (3.28)
Bh(uIhu,ψ)\displaystyle B_{h}^{*}(u-I_{h}u,\psi) =0ψ[Σh]2.\displaystyle=0\quad\forall\psi\in[\Sigma_{h}]^{2}. (3.29)

It is easy to check that IhI_{h} and Πh\Pi_{h} are polynomial preserving operators, which satisfy the following interpolation error estimates (cf. [17, 13])

φΠhφ0Chk+1φHk+1(Ω)φHk+1(Ω)2,ωIhω0Chk+1ωHk+1(Ω)ωHk+1(Ω).\begin{split}\|\varphi-\Pi_{h}\varphi\|_{0}&\leq Ch^{k+1}\|\varphi\|_{H^{k+1}(\Omega)}\quad\forall\varphi\in H^{k+1}(\Omega)^{2},\\ \|\omega-I_{h}\omega\|_{0}&\leq Ch^{k+1}\|\omega\|_{H^{k+1}(\Omega)}\quad\forall\omega\in H^{k+1}(\Omega).\end{split} (3.30)

In addition, we take πh\pi_{h} to be the standard nodal interpolation operator, which satisfies

πhζζ0Chk+1ζHk+1(Ω)ζHk+1(Ω).\displaystyle\|\pi_{h}\zeta-\zeta\|_{0}\leq Ch^{k+1}\|\zeta\|_{H^{k+1}(\Omega)}\quad\forall\zeta\in H^{k+1}(\Omega). (3.31)

4 Error analysis

In this section, we prove the convergence estimates for the semi-discrete scheme and fully discrete scheme with backward Euler in time. In particular we will show that the stability and convergence error estimates are independent of the anisotropy ratio ϱB\varrho_{B} and of the storativity coefficient c0c_{0}, and are valid for c0=0c_{0}=0.

4.1 Error analysis for semi-discrete scheme

In this subsection we prove the stability and convergence estimates for the semi-discrete scheme. To begin, we consider the following stability estimate.

Theorem 4.1.

(stability). Let (σh,uh,γh,zh,ph)(\sigma_{h},u_{h},\gamma_{h},z_{h},p_{h}) be the numerical solution of (3.1)-(3.5), then we have

σhL(0,T;L2(Ω))2+phL(0,T;L2(Ω))2+uhL(0,T;L2(Ω))2+K12zhL(0,T;L2(Ω))2+γhL(0,T;L2(Ω))2+K12zhL2(0,T;L2(Ω))2+uhL2(0,T;L2(Ω))2+phL2(0,T;L2(Ω))2+σhL2(0,T;L2(Ω))2+γhL2(0,T;L2(Ω))2C(fL(0,T;L2(Ω))2+fH1(0,T;L2(Ω))2+qH1(0,T;L2(Ω))2+qL(0,T;L2(Ω))2+p0H1(Ω)2).\begin{split}&\|\sigma_{h}\|_{L^{\infty}(0,T;L^{2}(\Omega))}^{2}+\|p_{h}\|_{L^{\infty}(0,T;L^{2}(\Omega))}^{2}+\|u_{h}\|_{L^{\infty}(0,T;L^{2}(\Omega))}^{2}+\|K^{-\frac{1}{2}}z_{h}\|_{L^{\infty}(0,T;L^{2}(\Omega))}^{2}\\ &\;+\|\gamma_{h}\|_{L^{\infty}(0,T;L^{2}(\Omega))}^{2}+\|K^{-\frac{1}{2}}z_{h}\|_{L^{2}(0,T;L^{2}(\Omega))}^{2}+\|u_{h}\|_{L^{2}(0,T;L^{2}(\Omega))}^{2}+\|p_{h}\|_{L^{2}(0,T;L^{2}(\Omega))}^{2}\\ &\;+\|\sigma_{h}\|_{L^{2}(0,T;L^{2}(\Omega))}^{2}+\|\gamma_{h}\|_{L^{2}(0,T;L^{2}(\Omega))}^{2}\leq C\Big{(}\|f\|_{L^{\infty}(0,T;L^{2}(\Omega))}^{2}\\ &\;+\|f\|_{H^{1}(0,T;L^{2}(\Omega))}^{2}+\|q\|_{H^{1}(0,T;L^{2}(\Omega))}^{2}+\|q\|_{L^{\infty}(0,T;L^{2}(\Omega))}^{2}+\|p^{0}\|_{H^{1}(\Omega)}^{2}\Big{)}.\end{split}
Proof.

Differentiating (3.1) in time, taking ψ=σh,v=tuh,η=tγh,ξ=zh,w=ph\psi=\sigma_{h},v=\partial_{t}u_{h},\eta=\partial_{t}\gamma_{h},\xi=z_{h},w=p_{h} in (3.1)-(3.5) and summing up the resulting equations yield

12ddt(A12(σh+αphI)02+c0ph02)+K12zh02=(f,tuh)+(q,ph).\displaystyle\frac{1}{2}\frac{d}{dt}\Big{(}\|A^{\frac{1}{2}}(\sigma_{h}+\alpha p_{h}I)\|_{0}^{2}+c_{0}\|p_{h}\|_{0}^{2}\Big{)}+\|K^{-\frac{1}{2}}z_{h}\|_{0}^{2}=(f,\partial_{t}u_{h})+(q,p_{h}).

Integrating over time implies

12(A12(σh+αphI)(t)02+c0ph(t)02)+0tK12zh02𝑑s=0t(tf,uh)𝑑s\displaystyle\frac{1}{2}\Big{(}\|A^{\frac{1}{2}}(\sigma_{h}+\alpha p_{h}I)(t)\|_{0}^{2}+c_{0}\|p_{h}(t)\|_{0}^{2}\Big{)}+\int_{0}^{t}\|K^{-\frac{1}{2}}z_{h}\|_{0}^{2}\;ds=-\int_{0}^{t}(\partial_{t}f,u_{h})\;ds
+0t(q,ph)𝑑s+12(A12(σh+αphI)(0)02+c0ph002)+(f,uh)(t)(f,uh)(0).\displaystyle\;+\int_{0}^{t}(q,p_{h})\;ds+\frac{1}{2}\Big{(}\|A^{\frac{1}{2}}(\sigma_{h}+\alpha p_{h}I)(0)\|_{0}^{2}+c_{0}\|p_{h}^{0}\|_{0}^{2}\Big{)}+(f,u_{h})(t)-(f,u_{h})(0).

The Cauchy-Schwarz inequality and Young’s inequality lead to

A12(σh+αphI)(t)02+c0ph(t)02+20tK12zh02dsϵ1(uh(t)02+0t(ph02+uh02)ds)+12ϵ1(f(t)02+0t(q02+tf02)ds)+A12(σh+αphI)(0)02+c0ph002+uh002+f(0)02.\begin{split}&\|A^{\frac{1}{2}}(\sigma_{h}+\alpha p_{h}I)(t)\|_{0}^{2}+c_{0}\|p_{h}(t)\|_{0}^{2}+2\int_{0}^{t}\|K^{-\frac{1}{2}}z_{h}\|_{0}^{2}\;ds\leq\epsilon_{1}\Big{(}\|u_{h}(t)\|_{0}^{2}\\ &\;+\int_{0}^{t}(\|p_{h}\|_{0}^{2}+\|u_{h}\|_{0}^{2})\;ds\Big{)}+\frac{1}{2\epsilon_{1}}\Big{(}\|f(t)\|_{0}^{2}+\int_{0}^{t}(\|q\|_{0}^{2}+\|\partial_{t}f\|_{0}^{2})\;ds\Big{)}\\ &\;+\|A^{\frac{1}{2}}(\sigma_{h}+\alpha p_{h}I)(0)\|_{0}^{2}+c_{0}\|p_{h}^{0}\|_{0}^{2}+\|u_{h}^{0}\|_{0}^{2}+\|f(0)\|_{0}^{2}.\end{split} (4.1)

By (3.1) and the inf-sup condition (3.8), we have

uhh+γh0Csupψ[Σh]2Bh(uh,ψ)+(as(ψ),γh)ψ0CA12(σh+αphI)0.\displaystyle\|u_{h}\|_{h}+\|\gamma_{h}\|_{0}\leq C\sup_{\psi\in[\Sigma_{h}]^{2}}\frac{-B_{h}^{*}(u_{h},\psi)+(as(\psi),\gamma_{h})}{\|\psi\|_{0}}\leq C\|A^{\frac{1}{2}}(\sigma_{h}+\alpha p_{h}I)\|_{0}.

Then an appeal to the discrete Poincaré-Friedrichs inequalities gives

uh0CA12(σh+αphI)0.\displaystyle\|u_{h}\|_{0}\leq C\|A^{\frac{1}{2}}(\sigma_{h}+\alpha p_{h}I)\|_{0}. (4.2)

Taking ψ=σh\psi=\sigma_{h}, v=uhv=u_{h} and η=γh\eta=\gamma_{h} in (3.1)-(3.3), and summing up the resulting equations, we deduce that

σh02C(ph02+ϵ2uh02+1ϵ2f02).\displaystyle\|\sigma_{h}\|_{0}^{2}\leq C(\|p_{h}\|_{0}^{2}+\epsilon_{2}\|u_{h}\|_{0}^{2}+\frac{1}{\epsilon_{2}}\|f\|_{0}^{2}). (4.3)

Thus, we can infer from (4.2) and (4.3) that

uh02CA12(σh+αphI)02C(σh02+ph02)C(ph02+ϵ2uh02+1ϵ2f02).\displaystyle\|u_{h}\|_{0}^{2}\leq C\|A^{\frac{1}{2}}(\sigma_{h}+\alpha p_{h}I)\|_{0}^{2}\leq C(\|\sigma_{h}\|_{0}^{2}+\|p_{h}\|_{0}^{2})\leq C\Big{(}\|p_{h}\|_{0}^{2}+\epsilon_{2}\|u_{h}\|_{0}^{2}+\frac{1}{\epsilon_{2}}\|f\|_{0}^{2}\Big{)}.

Choosing ϵ2\epsilon_{2} small enough yields

uh02C(ph02+1ϵ2f02).\displaystyle\|u_{h}\|_{0}^{2}\leq C(\|p_{h}\|_{0}^{2}+\frac{1}{\epsilon_{2}}\|f\|_{0}^{2}). (4.4)

On the other hand, we have from (3.4) and the inf-sup condition (3.6)

phZCsupξΣhbh(ph,ξ)ξ0=CsupξΣh(K1zh,ξ)ξ0CK1zh0.\displaystyle\|p_{h}\|_{Z}\leq C\sup_{\xi\in\Sigma_{h}}\frac{b_{h}^{*}(p_{h},\xi)}{\|\xi\|_{0}}=C\sup_{\xi\in\Sigma_{h}}\frac{(K^{-1}z_{h},\xi)}{\|\xi\|_{0}}\leq C\|K^{-1}z_{h}\|_{0}. (4.5)

Combining the discrete Poincaré-Friedrichs inequalities, (4.4) and (4.5) lead to

0t(uh02+ph02)𝑑sC0t(K1zh02+f02)𝑑s.\displaystyle\int_{0}^{t}(\|u_{h}\|_{0}^{2}+\|p_{h}\|_{0}^{2})\;ds\leq C\int_{0}^{t}(\|K^{-1}z_{h}\|_{0}^{2}+\|f\|_{0}^{2})\;ds. (4.6)

Therefore, choosing ϵ1\epsilon_{1} small enough, we can infer from (4.1), (4.5) and (4.6) that

A12(σh+αphI)(t)02+c0ph(t)02+uh(t)02+γh(t)02+0t(K12zh02+uh02+ph02+σh02+γh02)dsC(f(t)02+0t(f02+q02+tf02)ds+ph002+uh002+f(0)02+σh002).\begin{split}&\|A^{\frac{1}{2}}(\sigma_{h}+\alpha p_{h}I)(t)\|_{0}^{2}+c_{0}\|p_{h}(t)\|_{0}^{2}+\|u_{h}(t)\|_{0}^{2}+\|\gamma_{h}(t)\|_{0}^{2}\\ &\;+\int_{0}^{t}\Big{(}\|K^{-\frac{1}{2}}z_{h}\|_{0}^{2}+\|u_{h}\|_{0}^{2}+\|p_{h}\|_{0}^{2}+\|\sigma_{h}\|_{0}^{2}+\|\gamma_{h}\|_{0}^{2}\Big{)}\;ds\leq C\Big{(}\|f(t)\|_{0}^{2}\\ &\;+\int_{0}^{t}(\|f\|_{0}^{2}+\|q\|_{0}^{2}+\|\partial_{t}f\|_{0}^{2})\;ds+\|p_{h}^{0}\|_{0}^{2}+\|u_{h}^{0}\|_{0}^{2}+\|f(0)\|_{0}^{2}+\|\sigma_{h}^{0}\|_{0}^{2}\Big{)}.\end{split} (4.7)

Next, differentiating (3.1)-(3.4) in time and combining them with (3.5) with the choices (ψ,v,η,ξ,w)=(tσh,tuh,tγh,zh,tph)(\psi,v,\eta,\xi,w)=(\partial_{t}\sigma_{h},\partial_{t}u_{h},\partial_{t}\gamma_{h},z_{h},\partial_{t}p_{h}), we can obtain

20t(tA12(σh+αphI)02+c0tph02)𝑑s+K12zh(t)02ϵ1(ph(t)02+0ttuh02)+1ϵ1(q(t)02+0ttf02𝑑s)+0t(ph02+tq02)𝑑s+K12zh002+ph002+q(0)02.\begin{split}&2\int_{0}^{t}\Big{(}\|\partial_{t}A^{\frac{1}{2}}(\sigma_{h}+\alpha p_{h}I)\|_{0}^{2}+c_{0}\|\partial_{t}p_{h}\|_{0}^{2}\Big{)}\;ds+\|K^{-\frac{1}{2}}z_{h}(t)\|_{0}^{2}\leq\epsilon_{1}\Big{(}\|p_{h}(t)\|_{0}^{2}+\int_{0}^{t}\|\partial_{t}u_{h}\|_{0}^{2}\Big{)}\\ &+\frac{1}{\epsilon_{1}}\Big{(}\|q(t)\|_{0}^{2}+\int_{0}^{t}\|\partial_{t}f\|_{0}^{2}\;ds\Big{)}+\int_{0}^{t}(\|p_{h}\|_{0}^{2}+\|\partial_{t}q\|_{0}^{2})\;ds+\|K^{-\frac{1}{2}}z_{h}^{0}\|_{0}^{2}+\|p_{h}^{0}\|_{0}^{2}+\|q(0)\|_{0}^{2}.\end{split} (4.8)

Then by inf-sup condition (3.8) and (3.1) differentiated in time, we have

tuhh+tγh0CtA12(σh+αphI)0.\displaystyle\|\partial_{t}u_{h}\|_{h}+\|\partial_{t}\gamma_{h}\|_{0}\leq C\|\partial_{t}A^{\frac{1}{2}}(\sigma_{h}+\alpha p_{h}I)\|_{0}. (4.9)

Choosing ϵ1\epsilon_{1} in (4.8) small enough yields

0t(tA12(σh+αphI)02+c0tph02)𝑑s+K12zh(t)02+ph(t)02\displaystyle\int_{0}^{t}\Big{(}\|\partial_{t}A^{\frac{1}{2}}(\sigma_{h}+\alpha p_{h}I)\|_{0}^{2}+c_{0}\|\partial_{t}p_{h}\|_{0}^{2}\Big{)}\;ds+\|K^{-\frac{1}{2}}z_{h}(t)\|_{0}^{2}+\|p_{h}(t)\|_{0}^{2}
C(0t(ph02+tq02+tf02)𝑑s+q(t)02+K12zh002+ph002+q(0)02),\displaystyle\leq C\Big{(}\int_{0}^{t}(\|p_{h}\|_{0}^{2}+\|\partial_{t}q\|_{0}^{2}+\|\partial_{t}f\|_{0}^{2})\;ds+\|q(t)\|_{0}^{2}+\|K^{-\frac{1}{2}}z_{h}^{0}\|_{0}^{2}+\|p_{h}^{0}\|_{0}^{2}+\|q(0)\|_{0}^{2}\Big{)},

which coupling with (4.7) leads to

σh(t)02+ph(t)02+uh(t)02+K12zh(t)02+γh(t)02+0t(K12zh02+uh02+ph02+σh02+γh02)dsC(f(t)02+0t(f02+q02+tf02+tq02)ds+q(t)02+ph002+uh002+f(0)02+σh002+K12zh002+q(0)02).\begin{split}&\|\sigma_{h}(t)\|_{0}^{2}+\|p_{h}(t)\|_{0}^{2}+\|u_{h}(t)\|_{0}^{2}+\|K^{-\frac{1}{2}}z_{h}(t)\|_{0}^{2}+\|\gamma_{h}(t)\|_{0}^{2}\\ &\;+\int_{0}^{t}\Big{(}\|K^{-\frac{1}{2}}z_{h}\|_{0}^{2}+\|u_{h}\|_{0}^{2}+\|p_{h}\|_{0}^{2}+\|\sigma_{h}\|_{0}^{2}+\|\gamma_{h}\|_{0}^{2}\Big{)}\;ds\leq C\Big{(}\|f(t)\|_{0}^{2}+\int_{0}^{t}\Big{(}\|f\|_{0}^{2}+\|q\|_{0}^{2}\\ &\;+\|\partial_{t}f\|_{0}^{2}+\|\partial_{t}q\|_{0}^{2}\Big{)}\;ds+\|q(t)\|_{0}^{2}+\|p_{h}^{0}\|_{0}^{2}+\|u_{h}^{0}\|_{0}^{2}+\|f(0)\|_{0}^{2}+\|\sigma_{h}^{0}\|_{0}^{2}+\|K^{-\frac{1}{2}}z_{h}^{0}\|_{0}^{2}+\|q(0)\|_{0}^{2}\Big{)}.\end{split}

Since the discrete initial data (σh0,uh0,γh0,zh0,ph0)(\sigma_{h}^{0},u_{h}^{0},\gamma_{h}^{0},z_{h}^{0},p_{h}^{0}) is the elliptic projection of the initial data (σ0,u0,γ0,z0,p0)(\sigma^{0},u^{0},\gamma^{0},z^{0},p^{0}), we have

σh00+uh0+γh00+zh00+ph00C(p0H1(Ω)+f(0)0).\displaystyle\|\sigma_{h}^{0}\|_{0}+\|u_{h}^{0}\|+\|\gamma_{h}^{0}\|_{0}+\|z_{h}^{0}\|_{0}+\|p_{h}^{0}\|_{0}\leq C(\|p^{0}\|_{H^{1}(\Omega)}+\|f(0)\|_{0}). (4.10)

Therefore, the proof is completed.

Next, we will prove the convergence estimate. To simplify the notation, we let

σσh\displaystyle\sigma-\sigma_{h} =(σΠhσ)+(Πhσσh):=Cσ+Dσ,\displaystyle=(\sigma-\Pi_{h}\sigma)+(\Pi_{h}\sigma-\sigma_{h}):=C_{\sigma}+D_{\sigma},
uuh\displaystyle u-u_{h} =(uIhu)+(Ihuuh):=Cu+Du\displaystyle=(u-I_{h}u)+(I_{h}u-u_{h}):=C_{u}+D_{u}
γγh\displaystyle\gamma-\gamma_{h} =(γπhγ)+(πhγγh):=Cγ+Dγ,\displaystyle=(\gamma-\pi_{h}\gamma)+(\pi_{h}\gamma-\gamma_{h}):=C_{\gamma}+D_{\gamma},
zzh\displaystyle z-z_{h} =(zΠhz)+(Πhzzh):=Cz+Dz,\displaystyle=(z-\Pi_{h}z)+(\Pi_{h}z-z_{h}):=C_{z}+D_{z},
pph\displaystyle p-p_{h} =(pIhp)+(Ihpph):=Cp+Dp.\displaystyle=(p-I_{h}p)+(I_{h}p-p_{h}):=C_{p}+D_{p}.
Theorem 4.2.

Let (σh,uh,γh,zh,ph)(\sigma_{h},u_{h},\gamma_{h},z_{h},p_{h}) be the numerical solution of (3.1)-(3.5) and assume that the solution of (2.5)-(2.9) is sufficiently smooth, then we have the following error estimate

σσhL(0,T;L2(Ω))2+pphL(0,T;L2(Ω))2+uuhL(0,T;L2(Ω))2+γγhL(0,T;L2(Ω))2\displaystyle\|\sigma-\sigma_{h}\|_{L^{\infty}(0,T;L^{2}(\Omega))}^{2}+\|p-p_{h}\|_{L^{\infty}(0,T;L^{2}(\Omega))}^{2}+\|u-u_{h}\|_{L^{\infty}(0,T;L^{2}(\Omega))}^{2}+\|\gamma-\gamma_{h}\|_{L^{\infty}(0,T;L^{2}(\Omega))}^{2}
+K12(zzh)L(0,T;L2(Ω))2+σσhL2(0,T;L2(Ω))2+pphL2(0,T;L2(Ω))2\displaystyle\;+\|K^{-\frac{1}{2}}(z-z_{h})\|_{L^{\infty}(0,T;L^{2}(\Omega))}^{2}+\|\sigma-\sigma_{h}\|_{L^{2}(0,T;L^{2}(\Omega))}^{2}+\|p-p_{h}\|_{L^{2}(0,T;L^{2}(\Omega))}^{2}
+uuhL2(0,T;L2(Ω))2+σσhL2(0,T;L2(Ω))2+K12(zzh)L2(0,T;L2(Ω))2\displaystyle\;+\|u-u_{h}\|_{L^{2}(0,T;L^{2}(\Omega))}^{2}+\|\sigma-\sigma_{h}\|_{L^{2}(0,T;L^{2}(\Omega))}^{2}+\|K^{-\frac{1}{2}}(z-z_{h})\|_{L^{2}(0,T;L^{2}(\Omega))}^{2}
Ch2(k+1)(uL2(0,T;Hk+1(Ω))2+γH1(0,T;Hk+1(Ω))2+pH1(0,T;Hk+1(Ω))2+σH1(0,T;Hk+1(Ω))2\displaystyle\leq Ch^{2(k+1)}\Big{(}\|u\|_{L^{2}(0,T;H^{k+1}(\Omega))}^{2}+\|\gamma\|_{H^{1}(0,T;H^{k+1}(\Omega))}^{2}+\|p\|_{H^{1}(0,T;H^{k+1}(\Omega))}^{2}+\|\sigma\|_{H^{1}(0,T;H^{k+1}(\Omega))}^{2}
+K12zH1(0,T;Hk+1(Ω))2+γL(0,T;Hk+1(Ω))2+K12zL(0,T;Hk+1(Ω))2+σL(0,T;Hk+1(Ω))2\displaystyle\;+\|K^{-\frac{1}{2}}z\|_{H^{1}(0,T;H^{k+1}(\Omega))}^{2}+\|\gamma\|_{L^{\infty}(0,T;H^{k+1}(\Omega))}^{2}+\|K^{-\frac{1}{2}}z\|_{L^{\infty}(0,T;H^{k+1}(\Omega))}^{2}+\|\sigma\|_{L^{\infty}(0,T;H^{k+1}(\Omega))}^{2}
+pL(0,T;Hk+1(Ω))2+uL(0,T;Hk+1(Ω))2).\displaystyle\;+\|p\|_{L^{\infty}(0,T;H^{k+1}(\Omega))}^{2}+\|u\|_{L^{\infty}(0,T;H^{k+1}(\Omega))}^{2}\Big{)}.
Proof.

We obtain the following error equations by replacing (σ,u,γ,z,p)(\sigma,u,\gamma,z,p) by (σh,uh,γh,zh,ph)(\sigma_{h},u_{h},\gamma_{h},z_{h},p_{h}) and performing integration by parts on (3.1)-(3.5)

(A(Dσ+αDpI),ψ)Bh(Du,ψ)+(Dγ,as(ψ))\displaystyle(A(D_{\sigma}+\alpha D_{p}I),\psi)-B_{h}^{*}(D_{u},\psi)+(D_{\gamma},as(\psi)) =(A(Cσ+αCpI),ψ)\displaystyle=-(A(C_{\sigma}+\alpha C_{p}I),\psi)
+Bh(Cu,ψ)(Cγ,as(ψ)),\displaystyle+B_{h}^{*}(C_{u},\psi)-(C_{\gamma},as(\psi)), (4.11)
Bh(Dσ,v)\displaystyle B_{h}(D_{\sigma},v) =Bh(Cσ,v),\displaystyle=-B_{h}(C_{\sigma},v), (4.12)
(as(Dσ),η)\displaystyle(as(D_{\sigma}),\eta) =(as(Cσ),η),\displaystyle=-(as(C_{\sigma}),\eta), (4.13)
(K1Dz,ξ)+bh(Dp,ξ)\displaystyle(K^{-1}D_{z},\xi)+b_{h}^{*}(D_{p},\xi) =(K1Cz,ξ)bh(Cp,ξ),\displaystyle=-(K^{-1}C_{z},\xi)-b_{h}^{*}(C_{p},\xi), (4.14)
c0(tDp,w)+α(tA(Dσ+αDpI),wI)bh(Dz,w)\displaystyle c_{0}(\partial_{t}D_{p},w)+\alpha(\partial_{t}A(D_{\sigma}+\alpha D_{p}I),wI)-b_{h}(D_{z},w) =c0(tCp,w)\displaystyle=-c_{0}(\partial_{t}C_{p},w)
α(tA(Cσ+αDpI),wI)+bh(Cz,w)\displaystyle-\alpha(\partial_{t}A(C_{\sigma}+\alpha D_{p}I),wI)+b_{h}(C_{z},w) (4.15)

for (ψ,v,η,ξ,w)[Σh]2×[Sh]2×Mh×Σh×Sh(\psi,v,\eta,\xi,w)\in[\Sigma_{h}]^{2}\times[S_{h}]^{2}\times M_{h}\times\Sigma_{h}\times S_{h}.

Differentiating (4.11) with time, taking ψ=Dσ,v=tDu,η=tDγ,ξ=Dz,w=Dp\psi=D_{\sigma},v=\partial_{t}D_{u},\eta=\partial_{t}D_{\gamma},\xi=D_{z},w=D_{p}, and summing up the resulting equations yield

12t(A12(Dσ+αDpI)02+c0Dp02)+K12Dz02=(tA(Cσ+αCpI),Dσ)+Bh(tCu,Dσ)(tCγ,as(Dσ))Bh(Cσ,tDu)+(as(Cσ),tDγ)(K1Cz,Dz)bh(Cp,Dz)c0(tCp,Dp)α(tA(Cσ+αCpI),DpI)+bh(Cz,Dp).\begin{split}&\frac{1}{2}\partial_{t}\Big{(}\|A^{\frac{1}{2}}(D_{\sigma}+\alpha D_{p}I)\|_{0}^{2}+c_{0}\|D_{p}\|_{0}^{2}\Big{)}+\|K^{-\frac{1}{2}}D_{z}\|_{0}^{2}=-(\partial_{t}A(C_{\sigma}+\alpha C_{p}I),D_{\sigma})\\ &\;+B_{h}^{*}(\partial_{t}C_{u},D_{\sigma})-(\partial_{t}C_{\gamma},as(D_{\sigma}))-B_{h}(C_{\sigma},\partial_{t}D_{u})+(as(C_{\sigma}),\partial_{t}D_{\gamma})-(K^{-1}C_{z},D_{z})\\ &\;-b_{h}^{*}(C_{p},D_{z})-c_{0}(\partial_{t}C_{p},D_{p})-\alpha(\partial_{t}A(C_{\sigma}+\alpha C_{p}I),D_{p}I)+b_{h}(C_{z},D_{p}).\end{split} (4.16)

It follows from (3.26) and (3.27) that

Bh(tCu,Dσ)=0,Bh(Cσ,tDu)=0,bh(Cp,Dz)=0,bh(Cz,Dp)=0.\displaystyle B_{h}^{*}(\partial_{t}C_{u},D_{\sigma})=0,\quad-B_{h}(C_{\sigma},\partial_{t}D_{u})=0,\quad b_{h}^{*}(C_{p},D_{z})=0,\quad b_{h}(C_{z},D_{p})=0.

In addition, the Cauchy-Schwarz inequality yields

|(tA(Cσ+αCpI),Dσ)|\displaystyle|-(\partial_{t}A(C_{\sigma}+\alpha C_{p}I),D_{\sigma})| tA12(Cσ+αCpI0A12Dσ0,\displaystyle\leq\|\partial_{t}A^{\frac{1}{2}}(C_{\sigma}+\alpha C_{p}I\|_{0}\|A^{\frac{1}{2}}D_{\sigma}\|_{0},
|(tCγ,as(Dσ))|\displaystyle|(\partial_{t}C_{\gamma},as(D_{\sigma}))| tCγ0as(Dσ)0,\displaystyle\leq\|\partial_{t}C_{\gamma}\|_{0}\|as(D_{\sigma})\|_{0},
|(K1Cz,Dz)|\displaystyle|(K^{-1}C_{z},D_{z})| K12Cz0K12Dz0,\displaystyle\leq\|K^{-\frac{1}{2}}C_{z}\|_{0}\|K^{-\frac{1}{2}}D_{z}\|_{0},
|c0(tCp,Dp)|\displaystyle|c_{0}(\partial_{t}C_{p},D_{p})| c0tCp0Dp0,\displaystyle\leq c_{0}\|\partial_{t}C_{p}\|_{0}\|D_{p}\|_{0},
|(tA(Cσ+αCpI),DpI)|\displaystyle|(\partial_{t}A(C_{\sigma}+\alpha C_{p}I),D_{p}I)| tA12(Cσ+αCpI)0Dp0.\displaystyle\leq\|\partial_{t}A^{\frac{1}{2}}(C_{\sigma}+\alpha C_{p}I)\|_{0}\|D_{p}\|_{0}.

Therefore, we have from (4.16) that

12t(A12(Dσ+DpI)02+c0Dp02)+K12Dz02tA12(Cσ+αCpI)0A12Dσ0+tCγ0Dσ0\displaystyle\frac{1}{2}\partial_{t}\Big{(}\|A^{\frac{1}{2}}(D_{\sigma}+D_{p}I)\|_{0}^{2}+c_{0}\|D_{p}\|_{0}^{2}\Big{)}+\|K^{-\frac{1}{2}}D_{z}\|_{0}^{2}\leq\|\partial_{t}A^{\frac{1}{2}}(C_{\sigma}+\alpha C_{p}I)\|_{0}\|A^{\frac{1}{2}}D_{\sigma}\|_{0}+\|\partial_{t}C_{\gamma}\|_{0}\|D_{\sigma}\|_{0}
+(as(Cσ),tDγ)+K12Cz0K12Dz0+c0tCp0Dp0+tA12(Cσ+αCpI)0Dp0.\displaystyle\;+(as(C_{\sigma}),\partial_{t}D_{\gamma})+\|K^{-\frac{1}{2}}C_{z}\|_{0}\|K^{-\frac{1}{2}}D_{z}\|_{0}+c_{0}\|\partial_{t}C_{p}\|_{0}\|D_{p}\|_{0}+\|\partial_{t}A^{\frac{1}{2}}(C_{\sigma}+\alpha C_{p}I)\|_{0}\|D_{p}\|_{0}.

Integrating in time from 0 to an arbitrary t(0,T]t\in(0,T] results in

12(A12(Dσ+DpI)(t)02+c0Dp(t)02)+0tK12Dz02ds12(A12(Dσ+DpI)(0)02+c0Dp(0)02)0t(tas(Cσ),Dγ)ds+1ϵ2Cσ(t)02+ϵ2Dγ(t)02(as(Cσ),Dγ)(0)+ϵ10t(Dp02+K12Dz02+Dσ02)ds+1ϵ10t(tCγ02+K12Cz02+tCp02+tA12(Cσ+αCpI)02)ds.\begin{split}&\frac{1}{2}\Big{(}\|A^{\frac{1}{2}}(D_{\sigma}+D_{p}I)(t)\|_{0}^{2}+c_{0}\|D_{p}(t)\|_{0}^{2}\Big{)}+\int_{0}^{t}\|K^{-\frac{1}{2}}D_{z}\|_{0}^{2}\;ds\leq\frac{1}{2}\Big{(}\|A^{\frac{1}{2}}(D_{\sigma}+D_{p}I)(0)\|_{0}^{2}\\ &\;+c_{0}\|D_{p}(0)\|_{0}^{2}\Big{)}-\int_{0}^{t}(\partial_{t}as(C_{\sigma}),D_{\gamma})\;ds+\frac{1}{\epsilon_{2}}\|C_{\sigma}(t)\|_{0}^{2}+\epsilon_{2}\|D_{\gamma}(t)\|_{0}^{2}-(as(C_{\sigma}),D_{\gamma})(0)\\ &\;+\epsilon_{1}\int_{0}^{t}\Big{(}\|D_{p}\|_{0}^{2}+\|K^{-\frac{1}{2}}D_{z}\|_{0}^{2}+\|D_{\sigma}\|_{0}^{2}\Big{)}\;ds+\frac{1}{\epsilon_{1}}\int_{0}^{t}\Big{(}\|\partial_{t}C_{\gamma}\|_{0}^{2}+\|K^{-\frac{1}{2}}C_{z}\|_{0}^{2}+\|\partial_{t}C_{p}\|_{0}^{2}\\ &\;+\|\partial_{t}A^{\frac{1}{2}}(C_{\sigma}+\alpha C_{p}I)\|_{0}^{2}\Big{)}\;ds.\end{split} (4.17)

It follows from the inf-sup condition (3.8) and (4.11) that

Duh+Dγ0Csupψ[Σh]2Bh(Du,ψ)+(as(ψ),Dγ)ψ0=Csupψ[Σh]2(A(Dσ+αDpI),ψ)(A(Cσ+αCpI),ψ)(Cγ,as(ψ))ψ0C(A12(Dσ+αDpI)0+A12(Cσ+αCpI)0+Cγ0).\begin{split}\|D_{u}\|_{h}+\|D_{\gamma}\|_{0}&\leq C\sup_{\psi\in[\Sigma_{h}]^{2}}\frac{-B_{h}^{*}(D_{u},\psi)+(as(\psi),D_{\gamma})}{\|\psi\|_{0}}\\ &=C\sup_{\psi\in[\Sigma_{h}]^{2}}\frac{-(A(D_{\sigma}+\alpha D_{p}I),\psi)-(A(C_{\sigma}+\alpha C_{p}I),\psi)-(C_{\gamma},as(\psi))}{\|\psi\|_{0}}\\ &\leq C\Big{(}\|A^{\frac{1}{2}}(D_{\sigma}+\alpha D_{p}I)\|_{0}+\|A^{\frac{1}{2}}(C_{\sigma}+\alpha C_{p}I)\|_{0}+\|C_{\gamma}\|_{0}\Big{)}.\end{split} (4.18)

Taking ψ=Dσ,v=Du,η=Dγ\psi=D_{\sigma},v=D_{u},\eta=D_{\gamma} in (4.11)-(4.13), and summing up the resulting equations yield

(A(Dσ+αDpI),Dσ)=(A(Cσ+αCpI),Dσ)(Cγ,as(Dσ))+(as(Cσ),Dγ),\displaystyle(A(D_{\sigma}+\alpha D_{p}I),D_{\sigma})=-(A(C_{\sigma}+\alpha C_{p}I),D_{\sigma})-(C_{\gamma},as(D_{\sigma}))+(as(C_{\sigma}),D_{\gamma}),

which together with Young’s inequality implies

A12Dσ02C(Dp02+Cσ02+Cp02+Cγ02+1ϵ0Cσ02+ϵ0Dγ02).\displaystyle\|A^{\frac{1}{2}}D_{\sigma}\|_{0}^{2}\leq C\Big{(}\|D_{p}\|_{0}^{2}+\|C_{\sigma}\|_{0}^{2}+\|C_{p}\|_{0}^{2}+\|C_{\gamma}\|_{0}^{2}+\frac{1}{\epsilon_{0}}\|C_{\sigma}\|_{0}^{2}+\epsilon_{0}\|D_{\gamma}\|_{0}^{2}\Big{)}.

Taking ϵ0\epsilon_{0} small enough and using (4.18), we can obtain

Dσ02C(Cp02+Dp02+Cσ02+Cγ02).\displaystyle\|D_{\sigma}\|_{0}^{2}\leq C\Big{(}\|C_{p}\|_{0}^{2}+\|D_{p}\|_{0}^{2}+\|C_{\sigma}\|_{0}^{2}+\|C_{\gamma}\|_{0}^{2}\Big{)}. (4.19)

It follows from (3.6) and (4.14) that

DpZCsupξΣhbh(Dp,ξ)ξ0=CsupξΣh(K1Dz,ξ)(K1Cz,ξ)ξ0C(K1Dz0+K1Cz0),\displaystyle\|D_{p}\|_{Z}\leq C\sup_{\xi\in\Sigma_{h}}\frac{b_{h}^{*}(D_{p},\xi)}{\|\xi\|_{0}}=C\sup_{\xi\in\Sigma_{h}}\frac{-(K^{-1}D_{z},\xi)-(K^{-1}C_{z},\xi)}{\|\xi\|_{0}}\leq C(\|K^{-1}D_{z}\|_{0}+\|K^{-1}C_{z}\|_{0}), (4.20)

which together with the discrete Poincaré-Friedrichs inequalities implies

0tDp0𝑑s0tDpZ𝑑s0t(K1Dz0+K1Cz0)𝑑s.\displaystyle\int_{0}^{t}\|D_{p}\|_{0}\;ds\leq\int_{0}^{t}\|D_{p}\|_{Z}\;ds\leq\int_{0}^{t}(\|K^{-1}D_{z}\|_{0}+\|K^{-1}C_{z}\|_{0})\;ds.

An application of (4.19) and (4.20) leads to

0tDσ0𝑑s\displaystyle\int_{0}^{t}\|D_{\sigma}\|_{0}\;ds C0t(K1Dz0+K1Cz0+Cσ0+Cγ0+Cp0)𝑑s.\displaystyle\leq C\int_{0}^{t}\Big{(}\|K^{-1}D_{z}\|_{0}+\|K^{-1}C_{z}\|_{0}+\|C_{\sigma}\|_{0}+\|C_{\gamma}\|_{0}+\|C_{p}\|_{0}\Big{)}\;ds.

Besides, we can infer from (4.18) and (4.19) that

Duh+Dγ0C(K1Dz0+K1Cz0+Cσ0+Cγ0+Cp0).\displaystyle\|D_{u}\|_{h}+\|D_{\gamma}\|_{0}\leq C\Big{(}\|K^{-1}D_{z}\|_{0}+\|K^{-1}C_{z}\|_{0}+\|C_{\sigma}\|_{0}+\|C_{\gamma}\|_{0}+\|C_{p}\|_{0}\Big{)}.

Choosing ϵ1\epsilon_{1} and ϵ2\epsilon_{2} in (4.17) small enough, we can infer from the preceding arguments

A12(Dσ+DpI)(t)02+c0Dp(t)02+0t(K12Dz02+Dσ02+Dp02+Dγ02+Du02)ds12(A12(Dσ+DpI)(0)02+c0Dp(0)02)+Cσ(0)02+Dγ(0)02+Cσ(t)02+0t(Cγ02+K12Cz02+tCp02+Cσ02+tCσ02+Cp02+tCγ02)𝑑s.\begin{split}&\|A^{\frac{1}{2}}(D_{\sigma}+D_{p}I)(t)\|_{0}^{2}+c_{0}\|D_{p}(t)\|_{0}^{2}+\int_{0}^{t}\Big{(}\|K^{-\frac{1}{2}}D_{z}\|_{0}^{2}+\|D_{\sigma}\|_{0}^{2}+\|D_{p}\|_{0}^{2}+\|D_{\gamma}\|_{0}^{2}\\ &\;+\|D_{u}\|_{0}^{2}\Big{)}\;ds\leq\frac{1}{2}\Big{(}\|A^{\frac{1}{2}}(D_{\sigma}+D_{p}I)(0)\|_{0}^{2}+c_{0}\|D_{p}(0)\|_{0}^{2}\Big{)}+\|C_{\sigma}(0)\|_{0}^{2}+\|D_{\gamma}(0)\|_{0}^{2}+\|C_{\sigma}(t)\|_{0}^{2}\\ &+\int_{0}^{t}\Big{(}\|C_{\gamma}\|_{0}^{2}+\|K^{-\frac{1}{2}}C_{z}\|_{0}^{2}+\|\partial_{t}C_{p}\|_{0}^{2}+\|C_{\sigma}\|_{0}^{2}+\|\partial_{t}C_{\sigma}\|_{0}^{2}+\|C_{p}\|_{0}^{2}+\|\partial_{t}C_{\gamma}\|_{0}^{2}\Big{)}\;ds.\end{split} (4.21)

Differentiating (4.11)-(4.14) in time and setting (ψ,v,η,ξ,w)=(tDσ,tDu,tDγ,Dz,tDp)(\psi,v,\eta,\xi,w)=(\partial_{t}D_{\sigma},\partial_{t}D_{u},\partial_{t}D_{\gamma},D_{z},\partial_{t}D_{p}), we can obtain

tA12(Dσ+αDpI)02+c0tDp02+12tK12Dz02=(tA(Cσ+αCpI),tDσ+αtDpI)\displaystyle\|\partial_{t}A^{\frac{1}{2}}(D_{\sigma}+\alpha D_{p}I)\|_{0}^{2}+c_{0}\|\partial_{t}D_{p}\|_{0}^{2}+\frac{1}{2}\partial_{t}\|K^{-\frac{1}{2}}D_{z}\|_{0}^{2}=-(\partial_{t}A(C_{\sigma}+\alpha C_{p}I),\partial_{t}D_{\sigma}+\alpha\partial_{t}D_{p}I)
(tCγ,as(tDσ))+(tas(Cσ),tDγ)(tK1Cz,Dz)c0(tCp,tDp).\displaystyle-(\partial_{t}C_{\gamma},as(\partial_{t}D_{\sigma}))+(\partial_{t}as(C_{\sigma}),\partial_{t}D_{\gamma})-(\partial_{t}K^{-1}C_{z},D_{z})-c_{0}(\partial_{t}C_{p},\partial_{t}D_{p}).

Integrating over time and using as(tDσ)=as(t(Dσ+αDpI))as(\partial_{t}D_{\sigma})=as(\partial_{t}(D_{\sigma}+\alpha D_{p}I)) yield

0ttA12(Dσ+αDpI)02𝑑s+c00ttDp02𝑑s+12K12Dz(t)02ϵ10ttA12(Dσ+αDpI)02𝑑s+14ϵ1(0ttA12(Cσ+αCpI)02𝑑s+0ttCγ02𝑑s)+14ϵ20ttCσ02𝑑s+ϵ20ttDγ02𝑑s+0tK12tCz02𝑑s+0tK12Dz02𝑑s+c020ttCp02𝑑s+c020ttDp02𝑑s+12K12Dz(0)02.\begin{split}&\int_{0}^{t}\|\partial_{t}A^{\frac{1}{2}}(D_{\sigma}+\alpha D_{p}I)\|_{0}^{2}\;ds+c_{0}\int_{0}^{t}\|\partial_{t}D_{p}\|_{0}^{2}\;ds+\frac{1}{2}\|K^{-\frac{1}{2}}D_{z}(t)\|_{0}^{2}\\ &\leq\epsilon_{1}\int_{0}^{t}\|\partial_{t}A^{\frac{1}{2}}(D_{\sigma}+\alpha D_{p}I)\|_{0}^{2}\;ds+\frac{1}{4\epsilon_{1}}\Big{(}\int_{0}^{t}\|\partial_{t}A^{\frac{1}{2}}(C_{\sigma}+\alpha C_{p}I)\|_{0}^{2}\;ds+\int_{0}^{t}\|\partial_{t}C_{\gamma}\|_{0}^{2}\;ds\Big{)}\\ &\;+\frac{1}{4\epsilon_{2}}\int_{0}^{t}\|\partial_{t}C_{\sigma}\|_{0}^{2}\;ds+\epsilon_{2}\int_{0}^{t}\|\partial_{t}D_{\gamma}\|_{0}^{2}\;ds+\int_{0}^{t}\|K^{-\frac{1}{2}}\partial_{t}C_{z}\|_{0}^{2}\;ds+\int_{0}^{t}\|K^{-\frac{1}{2}}D_{z}\|_{0}^{2}\;ds\\ &\;+\frac{c_{0}}{2}\int_{0}^{t}\|\partial_{t}C_{p}\|_{0}^{2}\;ds+\frac{c_{0}}{2}\int_{0}^{t}\|\partial_{t}D_{p}\|_{0}^{2}\;ds+\frac{1}{2}\|K^{-\frac{1}{2}}D_{z}(0)\|_{0}^{2}.\end{split} (4.22)

By inf-sup condition (3.8) and (4.11) differentiated in time, we can obtain

tDuh+tDγ0CtA12(Dσ+αDpI)0.\displaystyle\|\partial_{t}D_{u}\|_{h}+\|\partial_{t}D_{\gamma}\|_{0}\leq C\|\partial_{t}A^{\frac{1}{2}}(D_{\sigma}+\alpha D_{p}I)\|_{0}.

Choosing ϵ1\epsilon_{1} and ϵ2\epsilon_{2} in (4.22) small enough leads to

0ttA12(Dσ+αDpI)02𝑑s+c00ttDp02𝑑s+K12Dz(t)02\displaystyle\int_{0}^{t}\|\partial_{t}A^{\frac{1}{2}}(D_{\sigma}+\alpha D_{p}I)\|_{0}^{2}\;ds+c_{0}\int_{0}^{t}\|\partial_{t}D_{p}\|_{0}^{2}\;ds+\|K^{-\frac{1}{2}}D_{z}(t)\|_{0}^{2}
C(0ttA12(Cσ+αCpI)02ds+0ttCγ02ds+0ttCσ02ds+0tK12tCz02ds\displaystyle\;\leq C\Big{(}\int_{0}^{t}\|\partial_{t}A^{\frac{1}{2}}(C_{\sigma}+\alpha C_{p}I)\|_{0}^{2}\;ds+\int_{0}^{t}\|\partial_{t}C_{\gamma}\|_{0}^{2}\;ds+\int_{0}^{t}\|\partial_{t}C_{\sigma}\|_{0}^{2}\;ds+\int_{0}^{t}\|K^{-\frac{1}{2}}\partial_{t}C_{z}\|_{0}^{2}\;ds
+c00ttCp02ds+0tK12Dz02ds+K12Dz(0)02),\displaystyle\;+c_{0}\int_{0}^{t}\|\partial_{t}C_{p}\|_{0}^{2}\;ds+\int_{0}^{t}\|K^{-\frac{1}{2}}D_{z}\|_{0}^{2}\;ds+\|K^{-\frac{1}{2}}D_{z}(0)\|_{0}^{2}\Big{)},

which can be combined with (4.18), (4.19), (4.20) and (4.21) implies

Dσ(t)02+Dp(t)02+Du(t)02+Dγ(t)02+K12Dz(t)02+0t(K12Dz02+Dσ02+Dp02+Dγ02+Du02)𝑑sC(A12(Dσ+DpI)(0)02+c0Dp(0)02+0t(Cγ02+Cz02+Cσ02+Cp02+tCγ02+tCp02+tA12(Cσ+αCpI)02+tCσ02)ds+K12Dz(0)02+Cσ(0)02+Dγ(0)02+K12tCz(t)02+Cσ(t)02+Cp(t)02+Cγ(t)02).\begin{split}&\|D_{\sigma}(t)\|_{0}^{2}+\|D_{p}(t)\|_{0}^{2}+\|D_{u}(t)\|_{0}^{2}+\|D_{\gamma}(t)\|_{0}^{2}+\|K^{-\frac{1}{2}}D_{z}(t)\|_{0}^{2}\\ &\;+\int_{0}^{t}\Big{(}\|K^{-\frac{1}{2}}D_{z}\|_{0}^{2}+\|D_{\sigma}\|_{0}^{2}+\|D_{p}\|_{0}^{2}+\|D_{\gamma}\|_{0}^{2}+\|D_{u}\|_{0}^{2}\Big{)}\;ds\\ &\leq C\Big{(}\|A^{\frac{1}{2}}(D_{\sigma}+D_{p}I)(0)\|_{0}^{2}+c_{0}\|D_{p}(0)\|_{0}^{2}+\int_{0}^{t}\Big{(}\|C_{\gamma}\|_{0}^{2}+\|C_{z}\|_{0}^{2}+\|C_{\sigma}\|_{0}^{2}+\|C_{p}\|_{0}^{2}\\ &\;+\|\partial_{t}C_{\gamma}\|_{0}^{2}+\|\partial_{t}C_{p}\|_{0}^{2}+\|\partial_{t}A^{\frac{1}{2}}(C_{\sigma}+\alpha C_{p}I)\|_{0}^{2}+\|\partial_{t}C_{\sigma}\|_{0}^{2}\Big{)}\;ds+\|K^{-\frac{1}{2}}D_{z}(0)\|_{0}^{2}\\ &\;+\|C_{\sigma}(0)\|_{0}^{2}+\|D_{\gamma}(0)\|_{0}^{2}+\|K^{-\frac{1}{2}}\partial_{t}C_{z}(t)\|_{0}^{2}+\|C_{\sigma}(t)\|_{0}^{2}+\|C_{p}(t)\|_{0}^{2}+\|C_{\gamma}(t)\|_{0}^{2}\Big{)}.\end{split}

The initial data satisfies

Dσ(0)0+Dγ(0)0+Dp(0)0+K12Dz(0)0C(Cσ(0)0+Cγ(0)0+Cp(0)0+K12Cz(0)0).\begin{split}&\|D_{\sigma}(0)\|_{0}+\|D_{\gamma}(0)\|_{0}+\|D_{p}(0)\|_{0}+\|K^{-\frac{1}{2}}D_{z}(0)\|_{0}\\ &\leq C\Big{(}\|C_{\sigma}(0)\|_{0}+\|C_{\gamma}(0)\|_{0}+\|C_{p}(0)\|_{0}+\|K^{-\frac{1}{2}}C_{z}(0)\|_{0}\Big{)}.\end{split} (4.23)

Therefore, the proof is completed by combining the preceding arguments and the interpolation error estimates (3.30)-(3.31).

Remark 4.1.

(robustness of error estimates for nearly incompressible materials). We can observe from the above proof that we employ the coercivity of AA, while the coercivity of AA on L2(Ω)2×2L^{2}(\Omega)^{2\times 2} is not uniform in λ\lambda. Indeed,

cλψ02A12ψ02ψL2(Ω)2×2\displaystyle c_{\lambda}\|\psi\|_{0}^{2}\leq\|A^{\frac{1}{2}}\psi\|_{0}^{2}\quad\forall\psi\in L^{2}(\Omega)^{2\times 2}

holds with a constant cλ>0c_{\lambda}>0 but cλ0c_{\lambda}\rightarrow 0 as λ+\lambda\rightarrow+\infty. It means that our error estimates obtained using the coercivity of AA may have error bounds growing unboundedly as λ+\lambda\rightarrow+\infty. In order to get a uniform bound, we can follow the framework given in [28, 53], which is omitted for simplicity. In fact, proceeding analogously to [28, 53] we can check that our proposed method has uniformly bounded estimate.

4.2 Error analysis for fully discrete scheme

In this subsection we analyze the convergence estimates for the fully discrete scheme. To this end we introduce a partition of the time interval [0,T][0,T] into subintervals [tn1,tn],1nN(Nis an integer)[t_{n-1},t_{n}],1\leq n\leq N(N\;\mbox{is an integer}) and denote the time step size by Δt=TN\Delta t=\frac{T}{N}. Using backward Euler scheme in time, we get the fully discrete staggered DG method as follows: Find (σh,uh,γh,zh,ph)[Σh]2×[Sh]2×Mh×Σh×Sh(\sigma_{h},u_{h},\gamma_{h},z_{h},p_{h})\in[\Sigma_{h}]^{2}\times[S_{h}]^{2}\times M_{h}\times\Sigma_{h}\times S_{h} such that

(A(σhn+αphnI),ψ)Bh(uhn,ψ)+(γhn,as(ψ))\displaystyle(A(\sigma_{h}^{n}+\alpha p_{h}^{n}I),\psi)-B_{h}^{*}(u_{h}^{n},\psi)+(\gamma_{h}^{n},as(\psi)) =0,\displaystyle=0, (4.24)
Bh(σhn,v)\displaystyle B_{h}(\sigma_{h}^{n},v) =(fn,v),\displaystyle=(f^{n},v), (4.25)
(as(σhn),η)\displaystyle(as(\sigma_{h}^{n}),\eta) =0,\displaystyle=0, (4.26)
(K1zhn,ξ)+bh(phn,ξ)\displaystyle(K^{-1}z_{h}^{n},\xi)+b_{h}^{*}(p_{h}^{n},\xi) =0,\displaystyle=0, (4.27)
c0(phnphn1Δt,w)+α(A(σhn+αphnI)A(σhn1+αphn1I)Δt,wI)bh(zhn,w)\displaystyle c_{0}(\frac{p_{h}^{n}-p_{h}^{n-1}}{\Delta t},w)+\alpha(\frac{A(\sigma_{h}^{n}+\alpha p_{h}^{n}I)-A(\sigma_{h}^{n-1}+\alpha p_{h}^{n-1}I)}{\Delta t},wI)-b_{h}(z_{h}^{n},w) =(qn,w)\displaystyle=(q^{n},w) (4.28)

for (ψ,v,η,ξ,w)[Σh]2×[Sh]2×Mh×Σh×Sh(\psi,v,\eta,\xi,w)\in[\Sigma_{h}]^{2}\times[S_{h}]^{2}\times M_{h}\times\Sigma_{h}\times S_{h}.

For a Sobolev space HH on Ω\Omega with a norm H\|\cdot\|_{H}, we define the discrete in time norms by

ψl2(0,T;H):=(n=1NΔtψH2)1/2,ψl(0,T;H):=max0nNψH\displaystyle\|\psi\|_{l^{2}(0,T;H)}:=\Big{(}\sum_{n=1}^{N}\Delta t\|\psi\|_{H}^{2}\Big{)}^{1/2},\quad\|\psi\|_{l^{\infty}(0,T;H)}:=\max_{0\leq n\leq N}\|\psi\|_{H}

and we let

σnσhn\displaystyle\sigma^{n}-\sigma_{h}^{n} =(σnΠhσn)+(Πhσnσhn):=Cσn+Dσn,\displaystyle=(\sigma^{n}-\Pi_{h}\sigma^{n})+(\Pi_{h}\sigma^{n}-\sigma_{h}^{n}):=C_{\sigma}^{n}+D_{\sigma}^{n},
unuhn\displaystyle u^{n}-u_{h}^{n} =(unIhun)+(Ihunuhn):=Cun+Dun\displaystyle=(u^{n}-I_{h}u^{n})+(I_{h}u^{n}-u_{h}^{n}):=C_{u}^{n}+D_{u}^{n}
γnγhn\displaystyle\gamma^{n}-\gamma_{h}^{n} =(γnπhγn)+(πhγnγhn):=Cγn+Dγn,\displaystyle=(\gamma^{n}-\pi_{h}\gamma^{n})+(\pi_{h}\gamma^{n}-\gamma_{h}^{n}):=C_{\gamma}^{n}+D_{\gamma}^{n},
znzhn\displaystyle z^{n}-z_{h}^{n} =(znΠhzn)+(Πhznzhn):=Czn+Dzn,\displaystyle=(z^{n}-\Pi_{h}z^{n})+(\Pi_{h}z^{n}-z_{h}^{n}):=C_{z}^{n}+D_{z}^{n},
pnphn\displaystyle p^{n}-p_{h}^{n} =(pnIhpn)+(Ihpnphn):=Cpn+Dpn.\displaystyle=(p^{n}-I_{h}p^{n})+(I_{h}p^{n}-p_{h}^{n}):=C_{p}^{n}+D_{p}^{n}.
Theorem 4.3.

(existence and uniqueness). There exists a unique solution to the fully discrete scheme (4.24)-(4.28).

Proof.

Since (4.24)-(4.24) is a square linear system, uniqueness implies existence. It suffices to show the uniqueness. To this end, suppose fn,qnf^{n},q^{n} and the (n1)(n-1)th time step solution vanishes. Then we have from (4.24)-(4.28)

(A(σhn+αphnI),ψ)Bh(uhn,ψ)+(γhn,as(ψ))\displaystyle(A(\sigma_{h}^{n}+\alpha p_{h}^{n}I),\psi)-B_{h}^{*}(u_{h}^{n},\psi)+(\gamma_{h}^{n},as(\psi)) =0ψ[Σh]2,\displaystyle=0\quad\forall\psi\in[\Sigma_{h}]^{2}, (4.29)
Bh(σhn,v)\displaystyle B_{h}(\sigma_{h}^{n},v) =0v[Sh]2,\displaystyle=0\quad\forall v\in[S_{h}]^{2}, (4.30)
(as(σhn),η)\displaystyle(as(\sigma_{h}^{n}),\eta) =0ηMh,\displaystyle=0\quad\forall\eta\in M_{h}, (4.31)
(K1zhn,ξ)+bh(phn,ξ)\displaystyle(K^{-1}z_{h}^{n},\xi)+b_{h}^{*}(p_{h}^{n},\xi) =0ξΣh,\displaystyle=0\quad\forall\xi\in\Sigma_{h}, (4.32)
c0(phn,w)+α(A(σhn+αphnI),wI)Δtbh(zhn,w)\displaystyle c_{0}(p_{h}^{n},w)+\alpha(A(\sigma_{h}^{n}+\alpha p_{h}^{n}I),wI)-\Delta tb_{h}(z_{h}^{n},w) =0wSh.\displaystyle=0\quad\forall w\in S_{h}. (4.33)

Taking ψ=σhn,v=uhn,η=γhn,ξ=zhn,w=phn\psi=\sigma_{h}^{n},v=u_{h}^{n},\eta=\gamma_{h}^{n},\xi=z_{h}^{n},w=p_{h}^{n} in (4.29)-(4.33) and summing up the resulting equations yield

A12(σhn+αphnI)02+c0phn02+ΔtK12zhn02=0,\displaystyle\|A^{\frac{1}{2}}(\sigma_{h}^{n}+\alpha p_{h}^{n}I)\|_{0}^{2}+c_{0}\|p_{h}^{n}\|_{0}^{2}+\Delta t\|K^{-\frac{1}{2}}z_{h}^{n}\|_{0}^{2}=0,

thereby σhn+αphnI=0\sigma_{h}^{n}+\alpha p_{h}^{n}I=0 and zhn=0z_{h}^{n}=0.

We can infer from inf-sup condition (3.6) and (4.32) that

phnZCK1zhn0,\displaystyle\|p_{h}^{n}\|_{Z}\leq C\|K^{-1}z_{h}^{n}\|_{0},

which implies phn=0p_{h}^{n}=0. Hence σhn=0\sigma_{h}^{n}=0 because of σhn+αphnI=0\sigma_{h}^{n}+\alpha p_{h}^{n}I=0.

It follows from (3.8) and (4.29) that

uhnh+γhn0CA12(σhn+αphnI)0,\displaystyle\|u_{h}^{n}\|_{h}+\|\gamma_{h}^{n}\|_{0}\leq C\|A^{\frac{1}{2}}(\sigma_{h}^{n}+\alpha p_{h}^{n}I)\|_{0},

which gives uhn=0u_{h}^{n}=0 and γhn=0\gamma_{h}^{n}=0. This completes the proof.

Theorem 4.4.

Let (σh,uh,γh,zh,ph)[Σh]2×[Sh]2×Mh×Σh×Sh(\sigma_{h},u_{h},\gamma_{h},z_{h},p_{h})\in[\Sigma_{h}]^{2}\times[S_{h}]^{2}\times M_{h}\times\Sigma_{h}\times S_{h} be the numerical solution of the fully discrete scheme (4.24)-(4.28) and assume that the solution of (2.5)-(2.9) is sufficiently smooth, then we have the following convergence estimate

K12(zzh)l2(0,T;L2(Ω))2+σσhl2(0,T;L2(Ω))2+zzhl2(0,T;L2(Ω))2+uuhl2(0,T;L2(Ω))2+σσhl2(0,T;L2(Ω))2+pphl(0,T;L2(Ω))2+K12(zzh)l(0,T;L2(Ω))2+σσhl(0,T;L2(Ω))2+γγhl(0,T;L2(Ω))2+uuhl(0,T;L2(Ω))2C(h2(k+1)(pH1(0,T;Hk+1(Ω))2+σH1(0,T;Hk+1(Ω))2+K12zH1(0,T;Hk+1(Ω))2+γH1(0,T;Hk+1(Ω))2+σL(0,T;Hk+1(Ω))2+pL(0,T;Hk+1(Ω))2+uL2(0,T;Hk+1(Ω))2+uL(0,T;Hk+1(Ω))2+γL(0,T;Hk+1(Ω))2+K12zL(0,T;Hk+1(Ω))2)+Δt2(pH2(0,T;L2(Ω))2+σH2(0,T;L2(Ω))2)).\begin{split}&\|K^{-\frac{1}{2}}(z-z_{h})\|_{l^{2}(0,T;L^{2}(\Omega))}^{2}+\|\sigma-\sigma_{h}\|_{l^{2}(0,T;L^{2}(\Omega))}^{2}+\|z-z_{h}\|_{l^{2}(0,T;L^{2}(\Omega))}^{2}+\|u-u_{h}\|_{l^{2}(0,T;L^{2}(\Omega))}^{2}\\ &\;+\|\sigma-\sigma_{h}\|_{l^{2}(0,T;L^{2}(\Omega))}^{2}+\|p-p_{h}\|_{l^{\infty}(0,T;L^{2}(\Omega))}^{2}+\|K^{-\frac{1}{2}}(z-z_{h})\|_{l^{\infty}(0,T;L^{2}(\Omega))}^{2}+\|\sigma-\sigma_{h}\|_{l^{\infty}(0,T;L^{2}(\Omega))}^{2}\\ &\;+\|\gamma-\gamma_{h}\|_{l^{\infty}(0,T;L^{2}(\Omega))}^{2}+\|u-u_{h}\|_{l^{\infty}(0,T;L^{2}(\Omega))}^{2}\leq C\Big{(}h^{2(k+1)}\Big{(}\|p\|_{H^{1}(0,T;H^{k+1}(\Omega))}^{2}\\ &\;+\|\sigma\|_{H^{1}(0,T;H^{k+1}(\Omega))}^{2}+\|K^{-\frac{1}{2}}z\|_{H^{1}(0,T;H^{k+1}(\Omega))}^{2}+\|\gamma\|_{H^{1}(0,T;H^{k+1}(\Omega))}^{2}+\|\sigma\|_{L^{\infty}(0,T;H^{k+1}(\Omega))}^{2}\\ &\;+\|p\|_{L^{\infty}(0,T;H^{k+1}(\Omega))}^{2}+\|u\|_{L^{2}(0,T;H^{k+1}(\Omega))}^{2}+\|u\|_{L^{\infty}(0,T;H^{k+1}(\Omega))}^{2}+\|\gamma\|_{L^{\infty}(0,T;H^{k+1}(\Omega))}^{2}\\ &\;+\|K^{-\frac{1}{2}}z\|_{L^{\infty}(0,T;H^{k+1}(\Omega))}^{2}\Big{)}+\Delta t^{2}\Big{(}\|p\|_{H^{2}(0,T;L^{2}(\Omega))}^{2}+\|\sigma\|_{H^{2}(0,T;L^{2}(\Omega))}^{2}\Big{)}\Big{)}.\end{split}
Proof.

We have the following error equations by performing integration by parts on (4.24)-(4.28)

(A(σnσhn+α(pnphn)I),ψ)Bh(unuhn,ψ)+(γnγhn,as(ψ))=0,\displaystyle(A(\sigma^{n}-\sigma_{h}^{n}+\alpha(p^{n}-p_{h}^{n})I),\psi)-B_{h}^{*}(u^{n}-u_{h}^{n},\psi)+(\gamma^{n}-\gamma_{h}^{n},as(\psi))=0, (4.34)
Bh(σnσhn,v)=0,\displaystyle B_{h}(\sigma^{n}-\sigma_{h}^{n},v)=0, (4.35)
(as(σnσhn),η)=0,\displaystyle(as(\sigma^{n}-\sigma_{h}^{n}),\eta)=0, (4.36)
(K1(znzhn),ξ)+bh(pnphn,ξ)=0,\displaystyle(K^{-1}(z^{n}-z_{h}^{n}),\xi)+b_{h}^{*}(p^{n}-p_{h}^{n},\xi)=0, (4.37)
αΔt(A(σnσhn+αpnphnI)A(σn1σhn1+α(pn1phn1I)),wI)\displaystyle\frac{\alpha}{\Delta t}(A(\sigma^{n}-\sigma_{h}^{n}+\alpha p^{n}-p_{h}^{n}I)-A(\sigma^{n-1}-\sigma_{h}^{n-1}+\alpha(p^{n-1}-p_{h}^{n-1}I)),wI)
+c0Δt(pnphn(pn1phn1),w)bh(znzhn,w)=c0(pnpn1Δtpt(:,tn),w)\displaystyle\;+\frac{c_{0}}{\Delta t}(p^{n}-p_{h}^{n}-(p^{n-1}-p_{h}^{n-1}),w)-b_{h}(z^{n}-z_{h}^{n},w)=c_{0}(\frac{p^{n}-p^{n-1}}{\Delta t}-p_{t}(:,t_{n}),w)
+αΔt(A(σnσn1+α(pnpn1)I)ΔtA(σt(:,tn)+αpt(:,tn)),wI)\displaystyle\;+\frac{\alpha}{\Delta t}(A(\sigma^{n}-\sigma^{n-1}+\alpha(p^{n}-p^{n-1})I)-\Delta tA(\sigma_{t}(:,t_{n})+\alpha p_{t}(:,t_{n})),wI) (4.38)

for (ψ,v,η,ξ,w)[Σh]2×[Sh]2×Mh×Σh×Sh(\psi,v,\eta,\xi,w)\in[\Sigma_{h}]^{2}\times[S_{h}]^{2}\times M_{h}\times\Sigma_{h}\times S_{h}.

Consider (4.34) at iterations nn and n1n-1 and taking the difference, then setting ψ=Dσn,v=Dun,η=DγnDγn1,ξ=Dzn,w=Dpn\psi=D_{\sigma}^{n},v=D_{u}^{n},\eta=D_{\gamma}^{n}-D_{\gamma}^{n-1},\xi=D_{z}^{n},w=D_{p}^{n} in the difference equation and (4.35)-(4.38), and summing up the resulting equations, we have

(K1(znzhn),Dzn)+c0Δt(pnphn(pn1phn1),Dpn)+1Δt(A(σnσhn+α(pnphn)I)A(σn1σhn1+α(pn1phn1)I),Dσn+αDpnI)=c0Δt(pnpn1Δtpt(:,tn),Dpn)+1Δt(A(σnσn1+α(pnpn1)I)ΔtA(σt(:,tn)+αpt(:,tn)),αDpnI)1Δt(CγnCγn1,as(Dσn))+1Δt(DγnDγn1,as(Cσn)).\begin{split}&(K^{-1}(z^{n}-z_{h}^{n}),D_{z}^{n})+\frac{c_{0}}{\Delta t}(p^{n}-p_{h}^{n}-(p^{n-1}-p_{h}^{n-1}),D_{p}^{n})\\ &\;+\frac{1}{\Delta t}(A(\sigma^{n}-\sigma_{h}^{n}+\alpha(p^{n}-p_{h}^{n})I)-A(\sigma^{n-1}-\sigma_{h}^{n-1}+\alpha(p^{n-1}-p_{h}^{n-1})I),D_{\sigma}^{n}+\alpha D_{p}^{n}I)\\ &=\frac{c_{0}}{\Delta t}(p^{n}-p^{n-1}-\Delta tp_{t}(:,t_{n}),D_{p}^{n})+\frac{1}{\Delta t}(A(\sigma^{n}-\sigma^{n-1}+\alpha(p^{n}-p^{n-1})I)\\ &\;-\Delta tA(\sigma_{t}(:,t_{n})+\alpha p_{t}(:,t_{n})),\alpha D_{p}^{n}I)-\frac{1}{\Delta t}(C_{\gamma}^{n}-C_{\gamma}^{n-1},as(D_{\sigma}^{n}))+\frac{1}{\Delta t}(D_{\gamma}^{n}-D_{\gamma}^{n-1},as(C_{\sigma}^{n})).\end{split}

Using the inequality (ab,a)|a|2|b|22(a-b,a)\geq\frac{|a|^{2}-|b|^{2}}{2} and the Cauchy-Schwarz inequality, we can obtain

K12Dzn02+c02Δt(Dpn02Dpn102+DpnDpn102)+12Δt(A12(Dσn+αDpn)02A12(Dσn1+αDpn1I)02)C(K12Czn0K12Dzn0+c0ΔtCpnCpn1Dpn0+1Δt(DγnDγn1,as(Cσn))+1ΔtA(Cσn+αCpnI)A(Cσn1+αCpn1I)0(Dσn0+Dpn0)+1ΔtA(σnσn1+α(pnpn1)I)ΔtA(σt(:,tn)+αpt(:,tn)I)0Dpn0+1ΔtCγnCγn10Dσn0+c0Δtpnpn1Δtpt(:,tn)0Dpn0).\begin{split}&\|K^{-\frac{1}{2}}D_{z}^{n}\|_{0}^{2}+\frac{c_{0}}{2\Delta t}\Big{(}\|D_{p}^{n}\|_{0}^{2}-\|D_{p}^{n-1}\|_{0}^{2}+\|D_{p}^{n}-D_{p}^{n-1}\|_{0}^{2}\Big{)}\\ &\;+\frac{1}{2\Delta t}\Big{(}\|A^{\frac{1}{2}}(D_{\sigma}^{n}+\alpha D_{p}^{n})\|_{0}^{2}-\|A^{\frac{1}{2}}(D_{\sigma}^{n-1}+\alpha D_{p}^{n-1}I)\|_{0}^{2}\Big{)}\\ &\leq C\Big{(}\|K^{-\frac{1}{2}}C_{z}^{n}\|_{0}\|K^{-\frac{1}{2}}D_{z}^{n}\|_{0}+\frac{c_{0}}{\Delta t}\|C_{p}^{n}-C_{p}^{n-1}\|\|D_{p}^{n}\|_{0}+\frac{1}{\Delta t}(D_{\gamma}^{n}-D_{\gamma}^{n-1},as(C_{\sigma}^{n}))\\ &\;+\frac{1}{\Delta t}\|A(C_{\sigma}^{n}+\alpha C_{p}^{n}I)-A(C_{\sigma}^{n-1}+\alpha C_{p}^{n-1}I)\|_{0}\Big{(}\|D_{\sigma}^{n}\|_{0}+\|D_{p}^{n}\|_{0}\Big{)}\\ &\;+\frac{1}{\Delta t}\|A(\sigma^{n}-\sigma^{n-1}+\alpha(p^{n}-p^{n-1})I)-\Delta tA(\sigma_{t}(:,t_{n})+\alpha p_{t}(:,t_{n})I)\|_{0}\|D_{p}^{n}\|_{0}\\ &\;+\frac{1}{\Delta t}\|C_{\gamma}^{n}-C_{\gamma}^{n-1}\|_{0}\|D_{\sigma}^{n}\|_{0}+\frac{c_{0}}{\Delta t}\|p^{n}-p^{n-1}-\Delta tp_{t}(:,t_{n})\|_{0}\|D_{p}^{n}\|_{0}\Big{)}.\end{split} (4.39)

It follows from (3.6), (4.37) and the discrete Poincaré-Friedrichs inequality that

Dpn0CDpnZCsupξΣhbh(phnIhpn,ξ)ξ0CK1(znzhn)0.\displaystyle\|D_{p}^{n}\|_{0}\leq C\|D_{p}^{n}\|_{Z}\leq C\sup_{\xi\in\Sigma_{h}}\frac{b_{h}^{*}(p_{h}^{n}-I_{h}p^{n},\xi)}{\|\xi\|_{0}}\leq C\|K^{-1}(z^{n}-z_{h}^{n})\|_{0}. (4.40)

Similarly, we can infer from (3.8) and (4.34) that

Dunh+Dγn0C(A12(σnσhn+α(pnphn)I)0+Cγn0).\displaystyle\|D_{u}^{n}\|_{h}+\|D_{\gamma}^{n}\|_{0}\leq C(\|A^{\frac{1}{2}}(\sigma^{n}-\sigma_{h}^{n}+\alpha(p^{n}-p_{h}^{n})I)\|_{0}+\|C_{\gamma}^{n}\|_{0}). (4.41)

Besides, taking (ψ,v,η)=(Dσn,Dun,Dγn)(\psi,v,\eta)=(D_{\sigma}^{n},D_{u}^{n},D_{\gamma}^{n}) in (4.34)-(4.36) and summing up the resulting equations yield

(A(σnσhn+α(pnphn)I),Dσn)+(Cγn,as(Dσn))(as(Cσn),Dγn)=0,\displaystyle(A(\sigma^{n}-\sigma_{h}^{n}+\alpha(p^{n}-p_{h}^{n})I),D_{\sigma}^{n})+(C_{\gamma}^{n},as(D_{\sigma}^{n}))-(as(C_{\sigma}^{n}),D_{\gamma}^{n})=0,

then an application of Young’s inequality yields

Dσn0C(Cpn0+Dpn0+Cσn0+Cγn0+ϵ1Dγn0+1ϵ1Cσn0).\begin{split}\|D_{\sigma}^{n}\|_{0}&\leq C\Big{(}\|C_{p}^{n}\|_{0}+\|D_{p}^{n}\|_{0}+\|C_{\sigma}^{n}\|_{0}+\|C_{\gamma}^{n}\|_{0}+\epsilon_{1}\|D_{\gamma}^{n}\|_{0}+\frac{1}{\epsilon_{1}}\|C_{\sigma}^{n}\|_{0}\Big{)}.\end{split} (4.42)

Choosing ϵ1\epsilon_{1} small enough and combining with (4.40), (4.41) give

Dσn0C(Cpn0+Dpn0+Cσn0+Cγn0)C(pnIhpn0+K1(znzhn)0+Πhσnσn0+γnπhγn0).\begin{split}\|D_{\sigma}^{n}\|_{0}&\leq C\Big{(}\|C_{p}^{n}\|_{0}+\|D_{p}^{n}\|_{0}+\|C_{\sigma}^{n}\|_{0}+\|C_{\gamma}^{n}\|_{0}\Big{)}\\ &\leq C\Big{(}\|p^{n}-I_{h}p^{n}\|_{0}+\|K^{-1}(z^{n}-z_{h}^{n})\|_{0}+\|\Pi_{h}\sigma^{n}-\sigma^{n}\|_{0}+\|\gamma^{n}-\pi_{h}\gamma^{n}\|_{0}\Big{)}.\end{split} (4.43)

On the other hand, we have from Taylor’s expansion

pnpn1Δtpt(:,tn)=tn1tn(ttn1)ptt(:,tn)𝑑s,\displaystyle p^{n}-p^{n-1}-\Delta tp_{t}(:,t_{n})=-\int_{t_{n-1}}^{t_{n}}(t-t_{n-1})p_{tt}(:,t_{n})\;ds,

where the right hand side can be estimated by the Cauchy-Schwarz inequality

1Δttn1tn(ttn1)ptt(:,tn)𝑑s02Δt3tn1tnptt02𝑑s.\displaystyle\|\frac{1}{\Delta t}\int_{t_{n-1}}^{t_{n}}(t-t_{n-1})p_{tt}(:,t_{n})\;ds\|_{0}^{2}\leq\frac{\Delta t}{3}\int_{t_{n-1}}^{t_{n}}\|p_{tt}\|_{0}^{2}\;ds.

Similarly, we have

1ΔtA(σnσn1Δtσt(:,tn))02\displaystyle\|\frac{1}{\Delta t}A(\sigma^{n}-\sigma^{n-1}-\Delta t\sigma_{t}(:,t_{n}))\|_{0}^{2} CΔt3tn1tnσtt02𝑑s,\displaystyle\leq\frac{C\Delta t}{3}\int_{t_{n-1}}^{t_{n}}\|\sigma_{tt}\|_{0}^{2}\;ds,
1ΔtA((pnpn1)IΔtpt(:,tn)I)02\displaystyle\|\frac{1}{\Delta t}A((p^{n}-p^{n-1})I-\Delta tp_{t}(:,t_{n})I)\|_{0}^{2} CΔt3tn1tnptt02𝑑s.\displaystyle\leq\frac{C\Delta t}{3}\int_{t_{n-1}}^{t_{n}}\|p_{tt}\|_{0}^{2}\;ds.

Therefore, we can infer from (4.39) and Young’s inequalities that

K12Dzn02+c02Δt(Dpn02Dpn102+DpnDpn102)+12Δt(A12(Dσn+αDpnI)02A12(Dσn1+αDpn1I)02)C(K12Czn02+Cσn02+Cγn02+Cpn02+1(Δt)2(CpnCpn102+A12(CσnCσn1)02+CγnCγn102)+Δttn1tn(ptt02+σtt02)ds+1Δt(DγnDγn1,as(Cσn))).\begin{split}&\|K^{-\frac{1}{2}}D_{z}^{n}\|_{0}^{2}+\frac{c_{0}}{2\Delta t}\Big{(}\|D_{p}^{n}\|_{0}^{2}-\|D_{p}^{n-1}\|_{0}^{2}+\|D_{p}^{n}-D_{p}^{n-1}\|_{0}^{2}\Big{)}+\frac{1}{2\Delta t}\Big{(}\|A^{\frac{1}{2}}(D_{\sigma}^{n}+\alpha D_{p}^{n}I)\|_{0}^{2}\\ &\;-\|A^{\frac{1}{2}}(D_{\sigma}^{n-1}+\alpha D_{p}^{n-1}I)\|_{0}^{2}\Big{)}\leq C\Big{(}\|K^{-\frac{1}{2}}C_{z}^{n}\|_{0}^{2}+\|C_{\sigma}^{n}\|_{0}^{2}+\|C_{\gamma}^{n}\|_{0}^{2}+\|C_{p}^{n}\|_{0}^{2}\\ &\;+\frac{1}{(\Delta t)^{2}}(\|C_{p}^{n}-C_{p}^{n-1}\|_{0}^{2}+\|A^{\frac{1}{2}}(C_{\sigma}^{n}-C_{\sigma}^{n-1})\|_{0}^{2}+\|C_{\gamma}^{n}-C_{\gamma}^{n-1}\|_{0}^{2})\\ &\;+\Delta t\int_{t_{n-1}}^{t_{n}}\Big{(}\|p_{tt}\|_{0}^{2}+\|\sigma_{tt}\|_{0}^{2}\Big{)}\;ds+\frac{1}{\Delta t}(D_{\gamma}^{n}-D_{\gamma}^{n-1},as(C_{\sigma}^{n}))\Big{)}.\end{split} (4.44)

Changing nn to jj in (4.39) and summing over j=1,,nj=1,\cdots,n, we can obtain

2Δtj=1nK12Dzj02+c0(Dpn02Dp002+j=1nDpnDpn102)+A12(Dσn+αDpnI)02C(j=1nΔt(K12Czj02+Cσj02+Cγj02+Cpj02)+j=1n1Δt(CpjCpj102+CσjCσj102+CγjCγj102)+Δt2j=1ntj1tj(ptt02+σtt02)ds+A12(Dσ0+αDp0I)02+j=1n(DγjDγj1,as(Cσj))).\begin{split}&2\Delta t\sum_{j=1}^{n}\|K^{-\frac{1}{2}}D_{z}^{j}\|_{0}^{2}+c_{0}\Big{(}\|D_{p}^{n}\|_{0}^{2}-\|D_{p}^{0}\|_{0}^{2}+\sum_{j=1}^{n}\|D_{p}^{n}-D_{p}^{n-1}\|_{0}^{2}\Big{)}+\|A^{\frac{1}{2}}(D_{\sigma}^{n}+\alpha D_{p}^{n}I)\|_{0}^{2}\\ &\leq C\Big{(}\sum_{j=1}^{n}\Delta t(\|K^{-\frac{1}{2}}C_{z}^{j}\|_{0}^{2}+\|C_{\sigma}^{j}\|_{0}^{2}+\|C_{\gamma}^{j}\|_{0}^{2}+\|C_{p}^{j}\|_{0}^{2})+\sum_{j=1}^{n}\frac{1}{\Delta t}\Big{(}\|C_{p}^{j}-C_{p}^{j-1}\|_{0}^{2}\\ &\;+\|C_{\sigma}^{j}-C_{\sigma}^{j-1}\|_{0}^{2}+\|C_{\gamma}^{j}-C_{\gamma}^{j-1}\|_{0}^{2}\Big{)}+\Delta t^{2}\sum_{j=1}^{n}\int_{t_{j-1}}^{t_{j}}(\|p_{tt}\|_{0}^{2}+\|\sigma_{tt}\|_{0}^{2})\;ds\\ &\;+\|A^{\frac{1}{2}}(D_{\sigma}^{0}+\alpha D_{p}^{0}I)\|_{0}^{2}+\sum_{j=1}^{n}(D_{\gamma}^{j}-D_{\gamma}^{j-1},as(C_{\sigma}^{j}))\Big{)}.\end{split} (4.45)

It follows from the interpolation error estimates (3.30)-(3.31) that

Czj0\displaystyle\|C_{z}^{j}\|_{0} Chk+1zjHk+1(Ω),Cγj0Chk+1γjHk+1(Ω),\displaystyle\leq Ch^{k+1}\|z^{j}\|_{H^{k+1}(\Omega)},\quad\|C_{\gamma}^{j}\|_{0}\leq Ch^{k+1}\|\gamma^{j}\|_{H^{k+1}(\Omega)},
Cσj0\displaystyle\|C_{\sigma}^{j}\|_{0} Chk+1σjHk+1(Ω),Cpj0Chk+1pjHk+1(Ω).\displaystyle\leq Ch^{k+1}\|\sigma^{j}\|_{H^{k+1}(\Omega)},\quad\|C_{p}^{j}\|_{0}\leq Ch^{k+1}\|p^{j}\|_{H^{k+1}(\Omega)}.

The Cauchy-Schwarz inequality and the interpolation error estimates (3.30) imply

1ΔtCpjCpj102=1Δt(pjIhpj)(pj1Ihpj1)02=1Δttj1tj(ptIhpt)02Ch2(k+1)tj1tjptHk+1(Ω)2,1ΔtCγjCγj102Ch2(k+1)tj1tjγtHk+1(Ω)2.\begin{split}\frac{1}{\Delta t}\|C_{p}^{j}-C_{p}^{j-1}\|_{0}^{2}&=\frac{1}{\Delta t}\|(p^{j}-I_{h}p^{j})-(p^{j-1}-I_{h}p^{j-1})\|_{0}^{2}=\frac{1}{\Delta t}\|\int_{t_{j-1}}^{t_{j}}(p_{t}-I_{h}p_{t})\|_{0}^{2}\\ &\leq Ch^{2(k+1)}\int_{t_{j-1}}^{t_{j}}\|p_{t}\|_{H^{k+1}(\Omega)}^{2},\\ \frac{1}{\Delta t}\|C_{\gamma}^{j}-C_{\gamma}^{j-1}\|_{0}^{2}&\leq Ch^{2(k+1)}\int_{t_{j-1}}^{t_{j}}\|\gamma_{t}\|_{H^{k+1}(\Omega)}^{2}.\end{split} (4.46)

Discrete integration by parts in time yields

j=1n(DγjDγj1,as(Cσj))=DγnCσnDγ0Cσ0j=1n(Dγj1,CσjCσj1),\displaystyle\sum_{j=1}^{n}(D_{\gamma}^{j}-D_{\gamma}^{j-1},as(C_{\sigma}^{j}))=D_{\gamma}^{n}C_{\sigma}^{n}-D_{\gamma}^{0}C_{\sigma}^{0}-\sum_{j=1}^{n}(D_{\gamma}^{j-1},C_{\sigma}^{j}-C_{\sigma}^{j-1}),

where the last term can be estimated by

CσjCσj102CΔth2(k+1)tj1tjσtHk+1(Ω)2.\displaystyle\|C_{\sigma}^{j}-C_{\sigma}^{j-1}\|_{0}^{2}\leq C\Delta th^{2(k+1)}\int_{t_{j-1}}^{t_{j}}\|\sigma_{t}\|_{H^{k+1}(\Omega)}^{2}.

Thus, the preceding arguments lead to

Δtj=1nK12Dzj02+c0(Dpn02Dp002+j=1nDpnDpn102)+A12(Dσn+αDpnI)02C(h2(k+1)(γH1(0,T;Hk+1(Ω))2+pH1(0,T;Hk+1(Ω))2+zL2(0,T;Hk+1(Ω))+σH1(0,T;Hk+1(Ω))2+σL(0,T;Hk+1(Ω))2)+Δt2j=1ntj1tj(ptt02+σtt02)ds+A12(Dσ0+αDp0I)02+Cσ002+Dγ002).\begin{split}&\Delta t\sum_{j=1}^{n}\|K^{-\frac{1}{2}}D_{z}^{j}\|_{0}^{2}+c_{0}\Big{(}\|D_{p}^{n}\|_{0}^{2}-\|D_{p}^{0}\|_{0}^{2}+\sum_{j=1}^{n}\|D_{p}^{n}-D_{p}^{n-1}\|_{0}^{2}\Big{)}+\|A^{\frac{1}{2}}(D_{\sigma}^{n}+\alpha D_{p}^{n}I)\|_{0}^{2}\\ &\leq C\Big{(}h^{2(k+1)}(\|\gamma\|_{H^{1}(0,T;H^{k+1}(\Omega))}^{2}+\|p\|_{H^{1}(0,T;H^{k+1}(\Omega))}^{2}+\|z\|_{L^{2}(0,T;H^{k+1}(\Omega))}+\|\sigma\|_{H^{1}(0,T;H^{k+1}(\Omega))}^{2}\\ &\;+\|\sigma\|_{L^{\infty}(0,T;H^{k+1}(\Omega))}^{2})+\Delta t^{2}\sum_{j=1}^{n}\int_{t_{j-1}}^{t_{j}}(\|p_{tt}\|_{0}^{2}+\|\sigma_{tt}\|_{0}^{2})\;ds+\|A^{\frac{1}{2}}(D_{\sigma}^{0}+\alpha D_{p}^{0}I)\|_{0}^{2}+\|C_{\sigma}^{0}\|_{0}^{2}+\|D_{\gamma}^{0}\|_{0}^{2}\Big{)}.\end{split}

Considering (4.34)-(4.37) at iterations nn and n1n-1, and taking the difference, setting ψ=DσnDσn1,v=DunDun1,η=DγnDγn1,ξ=Dzn,w=DpnDpn1\psi=D_{\sigma}^{n}-D_{\sigma}^{n-1},v=D_{u}^{n}-D_{u}^{n-1},\eta=D_{\gamma}^{n}-D_{\gamma}^{n-1},\xi=D_{z}^{n},w=D_{p}^{n}-D_{p}^{n-1} and summing up the resulting equations yield

1ΔtA12(Dσn+αDpnI)A12(Dσn1+αDpn1I)02\displaystyle\frac{1}{\Delta t}\|A^{\frac{1}{2}}(D_{\sigma}^{n}+\alpha D_{p}^{n}I)-A^{\frac{1}{2}}(D_{\sigma}^{n-1}+\alpha D_{p}^{n-1}I)\|_{0}^{2}
+(K1(DznDzn1),Dzn)+c0Δt(pnphn(pn1phn1),DpnDpn1)\displaystyle\;+(K^{-1}(D_{z}^{n}-D_{z}^{n-1}),D_{z}^{n})+\frac{c_{0}}{\Delta t}(p^{n}-p_{h}^{n}-(p^{n-1}-p_{h}^{n-1}),D_{p}^{n}-D_{p}^{n-1})
=c0(pnpn1Δtpt(:,tn),DpnDpn1)(K1(CznCzn1),Dzn)\displaystyle=c_{0}(\frac{p^{n}-p^{n-1}}{\Delta t}-p_{t}(:,t_{n}),D_{p}^{n}-D_{p}^{n-1})-(K^{-1}(C_{z}^{n}-C_{z}^{n-1}),D_{z}^{n})
+1Δt(A(σnσn1+α(pnpn1)I)ΔtA(σt(:,tn)+αpt(:,tn)I),α(DpnDpn1)I)\displaystyle\;+\frac{1}{\Delta t}(A(\sigma^{n}-\sigma^{n-1}+\alpha(p^{n}-p^{n-1})I)-\Delta tA(\sigma_{t}(:,t_{n})+\alpha p_{t}(:,t_{n})I),\alpha(D_{p}^{n}-D_{p}^{n-1})I)
1Δt(A(Cσn+αCpnI)A(Cσn1+αCpn1I),Dσn+αDpnI(Dσn1+αDpn1I))\displaystyle\;-\frac{1}{\Delta t}(A(C_{\sigma}^{n}+\alpha C_{p}^{n}I)-A(C_{\sigma}^{n-1}+\alpha C_{p}^{n-1}I),D_{\sigma}^{n}+\alpha D_{p}^{n}I-(D_{\sigma}^{n-1}+\alpha D_{p}^{n-1}I))
1Δt(CγnCγn1),as(DσnDσn1))+1Δt(as(CσnCσn1),DγnDγn1).\displaystyle\;-\frac{1}{\Delta t}(C_{\gamma}^{n}-C_{\gamma}^{n-1}),as(D_{\sigma}^{n}-D_{\sigma}^{n-1}))+\frac{1}{\Delta t}(as(C_{\sigma}^{n}-C_{\sigma}^{n-1}),D_{\gamma}^{n}-D_{\gamma}^{n-1}).

We can infer from inf-sup condition (3.6) and (3.8) that

Ihpnphn(Ihpn1phn1)0\displaystyle\|I_{h}p^{n}-p_{h}^{n}-(I_{h}p^{n-1}-p_{h}^{n-1})\|_{0} CK1(znzhn(zn1zhn1))0,\displaystyle\leq C\|K^{-1}(z^{n}-z_{h}^{n}-(z^{n-1}-z_{h}^{n-1}))\|_{0},
πhγnγhn(πhγn1γhn1)0\displaystyle\|\pi_{h}\gamma^{n}-\gamma_{h}^{n}-(\pi_{h}\gamma^{n-1}-\gamma_{h}^{n-1})\|_{0} CA12(Dσn+αDpnI)A12(Dσn1+αDpn1I)0.\displaystyle\leq C\|A^{\frac{1}{2}}(D_{\sigma}^{n}+\alpha D_{p}^{n}I)-A^{\frac{1}{2}}(D_{\sigma}^{n-1}+\alpha D_{p}^{n-1}I)\|_{0}.

Therefore, we have

1ΔtA12(Dσn+αDpnI)A12(Dσn1+αDpn1I)02+12(K12Dzn02\displaystyle\frac{1}{\Delta t}\|A^{\frac{1}{2}}(D_{\sigma}^{n}+\alpha D_{p}^{n}I)-A^{\frac{1}{2}}(D_{\sigma}^{n-1}+\alpha D_{p}^{n-1}I)\|_{0}^{2}+\frac{1}{2}\Big{(}\|K^{-\frac{1}{2}}D_{z}^{n}\|_{0}^{2}
K12Dzn102+K12DznK12Dzn102)+c0ΔtDpnDpn102\displaystyle\;-\|K^{-\frac{1}{2}}D_{z}^{n-1}\|_{0}^{2}+\|K^{-\frac{1}{2}}D_{z}^{n}-K^{-\frac{1}{2}}D_{z}^{n-1}\|_{0}^{2}\Big{)}+\frac{c_{0}}{\Delta t}\|D_{p}^{n}-D_{p}^{n-1}\|_{0}^{2}
C(Δt3tn1tn(ptt02+σtt02)ds+1Δt(CσnCσn102\displaystyle\leq C\Big{(}\frac{\Delta t}{3}\int_{t_{n-1}}^{t_{n}}(\|p_{tt}\|_{0}^{2}+\|\sigma_{tt}\|_{0}^{2})\;ds+\frac{1}{\Delta t}\Big{(}\|C_{\sigma}^{n}-C_{\sigma}^{n-1}\|_{0}^{2}
+CpnCpn102+CγnCγn102+K12(CznCzn1)02)+ΔtDzn02).\displaystyle\;+\|C_{p}^{n}-C_{p}^{n-1}\|_{0}^{2}+\|C_{\gamma}^{n}-C_{\gamma}^{n-1}\|_{0}^{2}+\|K^{-\frac{1}{2}}(C_{z}^{n}-C_{z}^{n-1})\|_{0}^{2}\Big{)}+\Delta t\|D_{z}^{n}\|_{0}^{2}\Big{)}.

Changing nn to jj and summing over j=1,,nj=1,\cdots,n, we can obtain

j=1n1ΔtA12(Dσj+αDpjI)A12(Dσj1+αDpj1I)02+K12Dzn02+j=1nc0ΔtDpjDpj102\displaystyle\sum_{j=1}^{n}\frac{1}{\Delta t}\|A^{\frac{1}{2}}(D_{\sigma}^{j}+\alpha D_{p}^{j}I)-A^{\frac{1}{2}}(D_{\sigma}^{j-1}+\alpha D_{p}^{j-1}I)\|_{0}^{2}+\|K^{-\frac{1}{2}}D_{z}^{n}\|_{0}^{2}+\sum_{j=1}^{n}\frac{c_{0}}{\Delta t}\|D_{p}^{j}-D_{p}^{j-1}\|_{0}^{2}
C(j=1n1Δt(CσjCσj102+CpjCpj102+CγjCγj102+K12(CzjCzj1)02)\displaystyle\leq C\Big{(}\sum_{j=1}^{n}\frac{1}{\Delta t}\Big{(}\|C_{\sigma}^{j}-C_{\sigma}^{j-1}\|_{0}^{2}+\|C_{p}^{j}-C_{p}^{j-1}\|_{0}^{2}+\|C_{\gamma}^{j}-C_{\gamma}^{j-1}\|_{0}^{2}+\|K^{-\frac{1}{2}}(C_{z}^{j}-C_{z}^{j-1})\|_{0}^{2}\Big{)}
+K12(Ihz0zh0)02+Δt30tn(ptt02+σtt02)ds+j=1nΔtK12Dzj02).\displaystyle+\|K^{-\frac{1}{2}}(I_{h}z^{0}-z_{h}^{0})\|_{0}^{2}+\frac{\Delta t}{3}\int_{0}^{t_{n}}(\|p_{tt}\|_{0}^{2}+\|\sigma_{tt}\|_{0}^{2})\;ds+\sum_{j=1}^{n}\Delta t\|K^{-\frac{1}{2}}D_{z}^{j}\|_{0}^{2}\Big{)}.

Proceeding analogously to (4.46), we can get

1ΔtCσjCσj102\displaystyle\frac{1}{\Delta t}\|C_{\sigma}^{j}-C_{\sigma}^{j-1}\|_{0}^{2} Ch2(k+1)tj1tjσtHk+1(Ω)2,\displaystyle\leq Ch^{2(k+1)}\int_{t_{j-1}}^{t_{j}}\|\sigma_{t}\|_{H^{k+1}(\Omega)}^{2},
1ΔtCzjCzj102\displaystyle\frac{1}{\Delta t}\|C_{z}^{j}-C_{z}^{j-1}\|_{0}^{2} Ch2(k+1)tj1tjK12ztHk+1(Ω)2.\displaystyle\leq Ch^{2(k+1)}\int_{t_{j-1}}^{t_{j}}\|K^{-\frac{1}{2}}z_{t}\|_{H^{k+1}(\Omega)}^{2}.

Combining the above estimates with the interpolation error estimates (3.30)-(3.31) completes the proof.

5 Fixed stress splitting scheme

The global system (4.24)-(4.28) consists of five variables which is relatively large. In order to reduce the computational costs, we propose the following fixed stress splitting scheme inspired by [30]. This includes two steps: First, the flow problem is solved independently. Second, the mechanics problem is solved using updated pressure and flux. For fixed n,in,i\in\mathbb{N}, the detailed splitting scheme reads as follows:

Step 1: Given (zhn,i1,phn,i1,σhn,i1)(z_{h}^{n,i-1},p_{h}^{n,i-1},\sigma_{h}^{n,i-1}), find (zhn,i,phn,i)(z_{h}^{n,i},p_{h}^{n,i}) such that

(K1zhn,i,ξ)+bh(phn,i,ξ)=0ξΣh,\displaystyle(K^{-1}z_{h}^{n,i},\xi)+b_{h}^{*}(p_{h}^{n,i},\xi)=0\quad\forall\xi\in\Sigma_{h}, (5.1)
((c0+β)phn,iΔt,w)+α(A(αphn,iI)Δt,wI)bh(zhn,i,w)=(qn,w)+(c0phn1Δt,w)\displaystyle((c_{0}+\beta)\frac{p_{h}^{n,i}}{\Delta t},w)+\alpha(\frac{A(\alpha p_{h}^{n,i}I)}{\Delta t},wI)-b_{h}(z_{h}^{n,i},w)=(q^{n},w)+(c_{0}\frac{p_{h}^{n-1}}{\Delta t},w)
+(βphn,i1Δt,w)α(A(σhn,i1)A(σhn1+αphn1I)Δt,wI)wSh.\displaystyle+(\beta\frac{p_{h}^{n,i-1}}{\Delta t},w)\;-\alpha(\frac{A(\sigma_{h}^{n,i-1})-A(\sigma_{h}^{n-1}+\alpha p_{h}^{n-1}I)}{\Delta t},wI)\quad\forall w\in S_{h}. (5.2)

Step 2: Given phn,ip_{h}^{n,i}, find (σhn,i,uhn,i,γhn,i)(\sigma_{h}^{n,i},u_{h}^{n,i},\gamma_{h}^{n,i}) such that

(A(σhn,i),ψ)Bh(uhn,i,ψ)+(γhn,i,as(ψ))\displaystyle(A(\sigma_{h}^{n,i}),\psi)-B_{h}^{*}(u_{h}^{n,i},\psi)+(\gamma_{h}^{n,i},as(\psi)) =(A(αphn,iI),ψ)ψ[Σh]2,\displaystyle=-(A(\alpha p_{h}^{n,i}I),\psi)\quad\forall\psi\in[\Sigma_{h}]^{2}, (5.3)
Bh(σhn,i,v)\displaystyle B_{h}(\sigma_{h}^{n,i},v) =(fn,v)v[Sh]2,\displaystyle=(f^{n},v)\quad\forall v\in[S_{h}]^{2}, (5.4)
(as(σhn,i),η)\displaystyle(as(\sigma_{h}^{n,i}),\eta) =0ηMh.\displaystyle=0\quad\forall\eta\in M_{h}. (5.5)

The initial guess for the iterations is chosen to be the solution at the last time step, i.e., (σhn,0,phn,0)=(σhn1,phn1)(\sigma_{h}^{n,0},p_{h}^{n,0})=(\sigma_{h}^{n-1},p_{h}^{n-1}).

Remark 5.1.

Since the mass matrix is block diagonal, we can apply local elimination for (5.1)-(5.2) and (5.3)-(5.5), respectively. In this way, we can get a reduced system which solely depends on the displacement and pressure.

Theorem 5.1.

(linear convergence of fixed stress splitting). Let (σhn,uhn,γhn,zhn,phn)(\sigma_{h}^{n},u_{h}^{n},\gamma_{h}^{n},z_{h}^{n},p_{h}^{n}) be the solution of (4.24)-(4.28) and let (σhn,i,uhn,i,γhn,i,zhn,i,phn,i)(\sigma_{h}^{n,i},u_{h}^{n,i},\gamma_{h}^{n,i},z_{h}^{n,i},p_{h}^{n,i}) be the solution of (5.1)-(5.5). Then for all β\beta satisfying

βα22(μ+λ).\displaystyle\beta\geq\frac{\alpha^{2}}{2(\mu+\lambda)}.

We have

phnphn,i02β2c0+β2+α2μ+λ+ΔtCp1Cinf1K112phnphn,i102,i1,\displaystyle\|p_{h}^{n}-p_{h}^{n,i}\|_{0}^{2}\leq\frac{\frac{\beta}{2}}{c_{0}+\frac{\beta}{2}+\frac{\alpha^{2}}{\mu+\lambda}+\Delta tC_{p}^{-1}C_{inf}^{-1}K_{1}^{\frac{1}{2}}}\|p_{h}^{n}-p_{h}^{n,i-1}\|_{0}^{2},\quad\forall i\geq 1, (5.6)

where CpC_{p} is the Poincaré constant and CinfC_{inf} is the inf-sup constant in (3.6).

Proof.

Let eui=uhn,i+uhne_{u}^{i}=-u_{h}^{n,i}+u_{h}^{n}, eσi=σhn,i+σhne_{\sigma}^{i}=-\sigma_{h}^{n,i}+\sigma_{h}^{n}, eγi=γhn,i+γhne_{\gamma}^{i}=-\gamma_{h}^{n,i}+\gamma_{h}^{n}, epi=phn,i+phne_{p}^{i}=-p_{h}^{n,i}+p_{h}^{n} and ezi=zhn,i+zhne_{z}^{i}=-z_{h}^{n,i}+z_{h}^{n}. Subtracting (5.1)-(5.5) from (4.24)-(4.28), we obtain

(A(eσi+αepiI),ψ)Bh(eui,ψ)+(eγi,as(ψ))\displaystyle(A(e_{\sigma}^{i}+\alpha e_{p}^{i}I),\psi)-B_{h}^{*}(e_{u}^{i},\psi)+(e_{\gamma}^{i},as(\psi)) =0,\displaystyle=0,
Bh(eσi,v)\displaystyle B_{h}(e_{\sigma}^{i},v) =0,\displaystyle=0,
as(eσi,η)\displaystyle as(e_{\sigma}^{i},\eta) =0,\displaystyle=0,
(K1ezi,ξ)+bh(epi,ξ)\displaystyle(K^{-1}e_{z}^{i},\xi)+b_{h}^{*}(e_{p}^{i},\xi) =0,\displaystyle=0,
c0Δt(epi,w)βΔt(phn,iphn,i1,w)+αΔt(A(eσi1+αepiI),wI)bh(ezi,w)\displaystyle\frac{c_{0}}{\Delta t}(e_{p}^{i},w)-\frac{\beta}{\Delta t}(p_{h}^{n,i}-p_{h}^{n,i-1},w)+\frac{\alpha}{\Delta t}(A(e_{\sigma}^{i-1}+\alpha e_{p}^{i}I),wI)-b_{h}(e_{z}^{i},w) =0\displaystyle=0

for (ψ,v,η,ξ,w)[Σh]2×[Sh]2×Mh×Σh×Sh(\psi,v,\eta,\xi,w)\in[\Sigma_{h}]^{2}\times[S_{h}]^{2}\times M_{h}\times\Sigma_{h}\times S_{h}.

Taking ψ=eσi1,v=eui,η=eγi\psi=e_{\sigma}^{i-1},v=e_{u}^{i},\eta=e_{\gamma}^{i}, ξ=ezi,w=epi\xi=e_{z}^{i},w=e_{p}^{i} in the above equations, and summing up the resulting equations, we can infer that

c0epi02+β(epiepi1,epi)+α2(A(epiI),epiI)+(A(eσi1),eσi)+ΔtK12ezi02=0.\displaystyle c_{0}\|e_{p}^{i}\|_{0}^{2}+\beta(e_{p}^{i}-e_{p}^{i-1},e_{p}^{i})+\alpha^{2}(A(e_{p}^{i}I),e_{p}^{i}I)+(A(e_{\sigma}^{i-1}),e_{\sigma}^{i})+\Delta t\|K^{-\frac{1}{2}}e_{z}^{i}\|_{0}^{2}=0. (5.7)

Consider (5.3)-(5.4) at iterations ii and i1i-1 and take the difference, setting ψ=σhn,iσhn,i1\psi=\sigma_{h}^{n,i}-\sigma_{h}^{n,i-1} and summing up the resulting equations yield

(A(σhn,iσhn,i1),σhn,iσhn,i1)=α(A((phn,iphn,i1)I),σhn,iσhn,i1),\displaystyle(A(\sigma_{h}^{n,i}-\sigma_{h}^{n,i-1}),\sigma_{h}^{n,i}-\sigma_{h}^{n,i-1})=-\alpha(A((p_{h}^{n,i}-p_{h}^{n,i-1})I),\sigma_{h}^{n,i}-\sigma_{h}^{n,i-1}),

which leads to

A12(σhn,iσhn,i1)02\displaystyle\|A^{\frac{1}{2}}(\sigma_{h}^{n,i}-\sigma_{h}^{n,i-1})\|_{0}^{2} αA12((phn,iphn,i1)I)0A12(σhn,iσhn,i1)0\displaystyle\leq\alpha\|A^{\frac{1}{2}}((p_{h}^{n,i}-p_{h}^{n,i-1})I)\|_{0}\|A^{\frac{1}{2}}(\sigma_{h}^{n,i}-\sigma_{h}^{n,i-1})\|_{0}
(α22A12((phn,iphn,i1)I)02+12A12(σhn,iσhn,i1)02).\displaystyle\leq\Big{(}\frac{\alpha^{2}}{2}\|A^{\frac{1}{2}}((p_{h}^{n,i}-p_{h}^{n,i-1})I)\|_{0}^{2}+\frac{1}{2}\|A^{\frac{1}{2}}(\sigma_{h}^{n,i}-\sigma_{h}^{n,i-1})\|_{0}^{2}\Big{)}.

Thus, we have

A1/2(eσieσi1)0αA1/2((phn,iphn,i1)I)0αA1/2((epiepi1)I)0α(λ+μ)1/2epiepi10.\begin{split}\|A^{1/2}(e_{\sigma}^{i}-e_{\sigma}^{i-1})\|_{0}&\leq\alpha\|A^{1/2}((p_{h}^{n,i}-p_{h}^{n,i-1})I)\|_{0}\leq\alpha\|A^{1/2}((e_{p}^{i}-e_{p}^{i-1})I)\|_{0}\\ &\leq\frac{\alpha}{(\lambda+\mu)^{1/2}}\|e_{p}^{i}-e_{p}^{i-1}\|_{0}.\end{split} (5.8)

We can infer from (5.7), the equality (ab,a)=a2b2+(ab)22(a-b,a)=\frac{a^{2}-b^{2}+(a-b)^{2}}{2} and the identity (A(eσi1),eσi)=14(A12(eσi+eσi1)02A12(eσieσi1)02)(A(e_{\sigma}^{i-1}),e_{\sigma}^{i})=\frac{1}{4}\Big{(}\|A^{\frac{1}{2}}(e_{\sigma}^{i}+e_{\sigma}^{i-1})\|_{0}^{2}-\|A^{\frac{1}{2}}(e_{\sigma}^{i}-e_{\sigma}^{i-1})\|_{0}^{2}\Big{)} that

c0epi02+β2(epi02epi102+epiepi102)+α2(A(epiI),epiI)\displaystyle c_{0}\|e_{p}^{i}\|_{0}^{2}+\frac{\beta}{2}(\|e_{p}^{i}\|_{0}^{2}-\|e_{p}^{i-1}\|_{0}^{2}+\|e_{p}^{i}-e_{p}^{i-1}\|_{0}^{2})+\alpha^{2}(A(e_{p}^{i}I),e_{p}^{i}I)
+14(A12(eσi+eσi1)02A12(eσieσi1)02)+ΔtK12ezi02=0.\displaystyle\;+\frac{1}{4}\Big{(}\|A^{\frac{1}{2}}(e_{\sigma}^{i}+e_{\sigma}^{i-1})\|_{0}^{2}-\|A^{\frac{1}{2}}(e_{\sigma}^{i}-e_{\sigma}^{i-1})\|_{0}^{2}\Big{)}+\Delta t\|K^{-\frac{1}{2}}e_{z}^{i}\|_{0}^{2}=0.

Therefore, an appeal to (5.8) yields

c0epi02+β2(epi02epi102)+(β2α24(λ+μ))epiepi102+α2(A(epiI),epiI)\displaystyle c_{0}\|e_{p}^{i}\|_{0}^{2}+\frac{\beta}{2}(\|e_{p}^{i}\|_{0}^{2}-\|e_{p}^{i-1}\|_{0}^{2})+(\frac{\beta}{2}-\frac{\alpha^{2}}{4(\lambda+\mu)})\|e_{p}^{i}-e_{p}^{i-1}\|_{0}^{2}+\alpha^{2}(A(e_{p}^{i}I),e_{p}^{i}I)
+14A12(eσi+eσi1)02+ΔtK12ezi020.\displaystyle\;+\frac{1}{4}\|A^{\frac{1}{2}}(e_{\sigma}^{i}+e_{\sigma}^{i-1})\|_{0}^{2}+\Delta t\|K^{-\frac{1}{2}}e_{z}^{i}\|_{0}^{2}\leq 0.

It follows from the inf-sup condition (3.6) and the discrete Poincaré-Friedrichs inequality that

epi0CpCinfK1ezi0,\displaystyle\|e_{p}^{i}\|_{0}\leq C_{p}C_{inf}\|K^{-1}e_{z}^{i}\|_{0},

where CpC_{p} is the Poincaré constant and CinfC_{inf} is the inf-sup constant in (3.6). Thereby, if we require

βα22(λ+μ),\displaystyle\beta\geq\frac{\alpha^{2}}{2(\lambda+\mu)},

then there holds

c0epi02+β2(epi02epi102)+α2(A(epiI),epiI)+ΔtCp1Cinf1K112epi02+14A12(eσi+eσi1)020,\begin{split}&c_{0}\|e_{p}^{i}\|_{0}^{2}+\frac{\beta}{2}(\|e_{p}^{i}\|_{0}^{2}-\|e_{p}^{i-1}\|_{0}^{2})+\alpha^{2}(A(e_{p}^{i}I),e_{p}^{i}I)\\ &\;+\Delta tC_{p}^{-1}C_{inf}^{-1}K_{1}^{\frac{1}{2}}\|e_{p}^{i}\|_{0}^{2}+\frac{1}{4}\|A^{\frac{1}{2}}(e_{\sigma}^{i}+e_{\sigma}^{i-1})\|_{0}^{2}\leq 0,\end{split} (5.9)

which leads to the desired estimate (5.6).

6 Numerical experiments

In this section we present several numerical experiments to verify the theoretical convergence rates and illustrate the behavior of the proposed method. We also present one benchmark example showing the locking-free property of the method. In the simulation presented below, we use the polynomial order k=1k=1.

6.1 Smooth solution test

In this example, we test the convergence of our method. To this end, we set Ω=(0,1)2\Omega=(0,1)^{2}, where the displacement and pressure are given by

u=(sin(2πt)sin(2πx)sin(2πy)xcos(t)),p=exp(t)(sin(πx)cos(πy)+10).\displaystyle u=\left(\begin{array}[]{c}\sin(2\pi t)\sin(2\pi x)\sin(2\pi y)\\ x\cos(t)\\ \end{array}\right),\quad p=\mbox{exp}(t)(\sin(\pi x)\cos(\pi y)+10).

The corresponding ff and qq can be calculated by (2.1)-(2.3). We set α=1,λ=1,μ=1\alpha=1,\lambda=1,\mu=1 and the simulation time is T=0.01T=0.01 with time steps Δt=h2\Delta t=h^{2}. We remark that Δt\Delta t should be chosen small enough so that the time discretization error will not influence the convergence rates. For this example, we consider two values of KK, i.e., K=IK=I and

K=(150001)\displaystyle K=\left(\begin{array}[]{cc}\frac{1}{50}&0\\ 0&1\\ \end{array}\right) (6.3)

We exploit square meshes and trapezoidal meshes for this example, see Figure 2 for an illustration. The convergence history against the number of degrees of freedom for rectangular meshes and trapezoidal meshes with various values of c0c_{0} is depicted in Figure 3. Optimal convergence rates matching the theoretical results can be obtained for various values of c0c_{0}. In addition, with different values of c0c_{0}, the accuracy for all the errors remain almost the same and the method is still valid for c0=0c_{0}=0. Further, the accuracy for L2L^{2} errors of σ,u,γ,z,p\sigma,u,\gamma,z,p on rectangular meshes and trapezoidal meshes is slightly different. Figure 4 shows the convergence history on rectangular meshes for anisotropic KK defined in (6.3) and similar performances can be observed, which illustrates the robustness of our method to the anisotropy ratio.

Refer to caption
Refer to caption
Figure 2: Example 6.1. Square mesh (left) and trapezoidal mesh (right).
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3: Example 6.1. Top plots: convergence history on square meshes for c0=1c_{0}=1 (left), c0=103c_{0}=10^{-3} (middle) and c0=0c_{0}=0 (right) for K=1K=1. Bottom plots: convergence history on trapezoidal meshes for c0=1c_{0}=1 (left), c0=103c_{0}=10^{-3} (middle) and c0=0c_{0}=0 (right) for K=1K=1.
Refer to caption
Refer to caption
Refer to caption
Figure 4: Example 6.1. Convergence history on square meshes for c0=1c_{0}=1 (left), c0=103c_{0}=10^{-3} (middle) and c0=0c_{0}=0 (right) for anisotropic KK.

6.2 Cantilever bracket benchmark problem

The computational domain is the unit square. We impose no-flow boundary condition along all sides. The deformation is fixed at the left edge, and a downward traction is applied along the top. The bottom and right sides are traction-free, see Figure 5 for an illustration of the boundary condition. The time step for this example is Δt=0.001\Delta t=0.001.

We use the same physical parameters as in [39] as they typically induce locking:

E=104,ν=0.4,α=0.93,c0=0,K=106,\displaystyle E=10^{4},\quad\nu=0.4,\quad\alpha=0.93,\quad c_{0}=0,\quad K=10^{-6},

where EE is the Young’s modulus and ν\nu is the Poisson ratio, and the Lamé parameters are computed by

λ=Eν(1+ν)(12ν),μ=E2(1+ν).\displaystyle\lambda=\frac{E\nu}{(1+\nu)(1-2\nu)},\quad\mu=\frac{E}{2(1+\nu)}.

Figure 6 shows that our proposed method yields a smooth pressure field, in contrast to the non-physical checkboard pattern that one obtains with continuous elasticity elements at the early time steps, see [39]. In addition, Figure 6 also indicates that the pressure solution along different xx-lines at time T=0.005T=0.005. We can observe that it is free of oscillations and our solution agrees with the one obtained by DG-mixed discretizations [39].

Refer to caption
Figure 5: Example 6.2. Profile of boundary condition.
Refer to caption
Refer to caption
Figure 6: Example 6.2. Pressure field at T=0.001T=0.001 (left) and pressure at different xx-lines, T=0.005T=0.005 (right).

6.3 Anisotropic meshes

The computational domain is taken to be Ω=(0,1)2\Omega=(0,1)^{2} and we set α=1,λ=1,μ=1,K=I\alpha=1,\lambda=1,\mu=1,K=I. The simulation time is T=0.01T=0.01 with time steps Δt=h2\Delta t=h^{2}. The exact displacement and pressure are given as

u=(sin(2πt)sin(2πy)exp(x/ϑ)sin(2πt)cos(2πy)exp(x/ϑ)),p=exp(t) exp(x/ϑ)x(1x)2y(1y)2.\displaystyle u=\left(\begin{array}[]{c}\sin(2\pi t)\sin(2\pi y)\mbox{exp}(-x/\vartheta)\\ \sin(2\pi t)\cos(2\pi y)\mbox{exp}(-x/\vartheta)\\ \end{array}\right),\quad p=\mbox{exp}(t)\mbox{ exp}(-x/\vartheta)x(1-x)^{2}y(1-y)^{2}.

The used meshes are of Shishkin type (cf. [2]), see the example in Figure 8. For a parameter N2N\geq 2 they are constructed by choosing a transition point parameter δ(0,1)\delta\in(0,1) and generating a grid of points (xj,yj)(x^{j},y^{j}) by

xj\displaystyle x^{j} ={j2δN,0jN2,j,δ+(jN2)2(1δ)N,N2<jN,j,\displaystyle=\begin{cases}j\frac{2\delta}{N},\hskip 71.13188pt0\leq j\leq\frac{N}{2},\;j\in\mathbb{N},\\ \delta+(j-\frac{N}{2})\frac{2(1-\delta)}{N},\quad\frac{N}{2}<j\leq N,\;j\in\mathbb{N},\end{cases}
yi\displaystyle y^{i} =iN0iN,i.\displaystyle=\frac{i}{N}\quad 0\leq i\leq N,\;i\in\mathbb{N}.

The transition point parameter δ\delta is chosen as δ=min{12,3ϑ|ln(ϑ)|}\delta=\min\{\frac{1}{2},3\vartheta|\mbox{ln}(\vartheta)|\}. Here we choose ϑ=102\vartheta=10^{-2} and the corresponding exact velocity and pressure at t=0.01t=0.01 are displayed in Figures 7 and 8, where the exponential boundary layer near x=0x=0 is clearly visible. The layer has a width of 𝒪(ϑ)\mathcal{O}(\vartheta) and is present in the velocity and pressure solution. The convergence history against the number of degrees of freedom for various values of c0c_{0} is shown in Figure 9 on square meshes and anisotropic meshes. We can observe that optimal convergence rates matching the theoretical results can be obtained, which indicates that our method can work well on anisotropic meshes and the method works for c0=0c_{0}=0. Further, the errors on anisotropic meshes are much smaller than that of rectangular meshes, which illustrates the superior performances of anisotropic meshes in particular for layered problem.

Refer to caption
Refer to caption
Figure 7: Example 6.3. Exact velocity: u1u_{1} (left) and u2u_{2} (right).
Refer to caption
Refer to caption
Figure 8: Example 6.3. Exact pressure (left) and example mesh (right) for ϑ=102,N=8\vartheta=10^{-2},N=8 used in the simulations.
Refer to caption
Refer to caption
Refer to caption
Figure 9: Example 6.3. Convergence history with uniform rectangular meshes (solid line) and anisotropic meshes (dashed line). c0=1c_{0}=1 (left), c0=103c_{0}=10^{-3} (middle), and c0=0c_{0}=0 (right).

6.4 A model problem of vertical compaction

In this example, we consider a more practical case motivated by the example given in [25], where three sedimentary layers overlying an impermeable bedrock in a basin. The sediment stack totals 420m at the deepest point of the basin (x=0x=0m) but thins to 120m above the step (x>4000x>4000m). The top two layers of the sequence are 20m thick each, see Figure 10 for an illustration of the domain geometry. The first and third layers are aquifers; the middle layer is relatively impermeable to flow. The flow field initially is at steady state, but pumping from the lower aquifer reduces hydraulic head by 6m per year at the basin center. The head drop moves fluid away from the step. The fluid supply in the upper reservoir is limitless. The initial condition is set to zero and the boundary conditions for region A to E are described in Table 1, in addition, we let f=q=0f=q=0. The values for the parameters are given in Table 2. The final simulation time is 10 years, where we take the simulation time to be (0:360:360×10)(0:360:360\times 10). Figure 11 shows the plot for pressure and displacement at 2 years and 10 years.

Refer to caption
Refer to caption
Figure 10: Example 6.4. Model geometry showing boundary segments (left) and profile of the mesh used (right). Vertical exaggeration ×5\times 5
A zn=0z\cdot n=0 u=(0,0)Tu=(0,0)^{T}
B zn=0z\cdot n=0 u1=0u_{1}=0
C p=0p=0 u1=0u_{1}=0
D p=0p=0 σn=(0,0)T\sigma n=(0,0)^{T}
E p=H(t)p=H(t) u1=0u_{1}=0
Table 1: Example 6.4. Description of boundary conditions. u=(u1,u2)Tu=(u_{1},u_{2})^{T}.
Variable Description Value
c0c_{0} storage coefficient, aquifer layers 106m110^{-6}\mbox{m}^{-1}
c0c_{0} storage coefficient, confining layers 105m110^{-5}\mbox{m}^{-1}
KK permeability, aquifer layers 25 m/day
KK permeability, confining layers 0.01 m/day
α\alpha Biot-Willis coefficient 1
E Young’s modulus, aquifer layers 81088\cdot 10^{8} Pa
E Young’s modulus, confining layers 81078\cdot 10^{7} Pa
ν\nu Poisson ratio, all regions 0.25
H(t)H(t) Declining head boundary (6 m/year)t\cdot t
Table 2: Example 6.4. Values of parameters.
Refer to caption
Refer to caption
Figure 11: Example 6.4. Numerical results at 2 years (left) and 10 years (right). Surface: pressure; Contour: displacement (|uh||u_{h}|).

7 Conclusion

In this paper we analyzed a staggered DG method for a five-field formulation of the Biot system of poroelasticity on general polygonal meshes. The method is proved to converge in optimal rates for all the variables in their natural norms. In addition, the error estimates are independent of the storativity coefficient c0c_{0} and are valid even for c0=0c_{0}=0, which is also confirmed by our numerical simulation. Several numerical experiments including cantilever bracket benchmark problem are presented, which further confirms the locking-free property of our method. In addition, the accuracy of our method is slightly influenced by the shape of the grid. Numerical results illustrate that our method is a good option for solving problems on anisotropic meshes. Moreover, our method can handle nonmatching meshes straightforwardly, which is advantageous in solving problem with local features. In the future we will extend this method to solve nonlinear poroelasticity.

Acknowledgment

The research of Eric Chung is partially supported by the Hong Kong RGC General Research Fund (Project numbers 14304719 and 14302018) and CUHK Faculty of Science Direct Grant 2019-20 and NSFC/RGC Joint Research Scheme (Project number HKUST620/15). The research of Eun-Jae Park was supported by the National Research Foundation of Korea (NRF) grant funded by the Ministry of Science and ICT (NRF-2015R1A5A1009350 and NRF-2019R1A2C2090021).

References

  • [1] I. Ambartsumyan, E. Khattatov, and I. Yotov. A coupled multipoint stress – multipoint flux mixed finite element method for the Biot system of poroelasticity. arXiv:2001.04582, 2020.
  • [2] T. Apel, A. Linke, and C. Merdon. A nonconforming pressure-robust finite element method for the Stokes equations on anisotropic meshes. arXiv:2002.12127, 2020.
  • [3] M. A. Biot. General theory of three‐dimensional consolidation. J. Appl. Phys., 12(2):155–164, 1941.
  • [4] D. Boffi, M. Botti, and D. A. Di Pietro. A nonconforming high-order method for the Biot problem on general meshes. SIAM J. Sci. Comput., 38(3):A1508–A1537, 2016.
  • [5] D. Braess. Finite Elements: Theory, Fast Solvers, and Applications in Solid Mechanics. Cambridge University Press, 2007.
  • [6] S. C. Brenner. Poincaré-Friedrichs inequalities for piecewise H1H^{1} functions. SIAM J. Numer. Anal., 41(1):306–324, 2003.
  • [7] F. Brezzi and R. S. Falk. Stability of higher-order Hood-Taylor methods. SIAM J. Numer. Anal., 28(3):581–590, 1991.
  • [8] S. Cheung, E. T. Chung, H. H. Kim, and Y. Qian. Staggered discontinuous Galerkin methods for the incompressible Navier–Stokes equations. J. Comput. Phys., 302(1):251–266, 2015.
  • [9] E. T. Chung, B. Cockburn, and G. Fu. The staggered DG method is the limit of a hybridizable DG method. SIAM J. Numer. Anal., 52(2):915–932, 2014.
  • [10] E. T. Chung, B. Cockburn, and G. Fu. The staggered DG method is the limit of a hybridizable DG method. Part II: The Stokes flow. J. Sci. Comput., 66(2):870–887, 2016.
  • [11] E. T. Chung, J. Du, and C. Lam. Discontinuous Galerkin methods with staggered hybridization for linear elastodynamics. Comput. Math. Appl., 74(6):1198–1214, 2017.
  • [12] E. T. Chung and B. Engquist. Optimal discontinuous Galerkin methods for wave propagation. SIAM J. Numer. Anal., 44(5):2131–2158, 2006.
  • [13] E. T. Chung and B. Engquist. Optimal discontinuous Galerkin methods for the acoustic wave equation in higher dimensions. SIAM J. Numer. Anal., 47(5):3820–3848, 2009.
  • [14] E. T. Chung, H. H. Kim, and O. B. Widlund. Two-level overlapping Schwarz algorithms for a staggered discontinuous Galerkin method. SIAM J. Numer. Anal., 51(1):47–67, 2013.
  • [15] E. T. Chung and C. Lee. A staggered discontinuous Galerkin method for the curl–curl operator. IMA J. Numer. Anal., 32(3):1241–1265, 2012.
  • [16] E. T. Chung and W. Qiu. Analysis of an SDG method for the incompressible Navier–Stokes equations. SIAM J. Numer. Anal., 55(2):543–569, 2017.
  • [17] P. G. Ciarlet. The Finite Element Method for Elliptic Problems. Society for Industrial and Applied Mathematics, 2002.
  • [18] J. Du and E. T. Chung. An adaptive staggered discontinuous Galerkin method for the steady state convection–diffusion equation. J. Sci. Comput., 77(3):1490–1518, 2018.
  • [19] G. Fu. A high-order HDG method for the Biot’s consolidation model. Comput. Math. Appl., 77(1):237–252, 2019.
  • [20] F. J. Gaspar, F. J. Lisbona, and P. N. Vabishchevich. A finite difference analysis of Biot’s consolidation model. Appl. Numer. Math., 44(4):487–506, 2003.
  • [21] J. Guzmán and L. R. Scott. The Scott-Vogelius finite elements revisited. Math. Comp., 88(316):515–529, 2019.
  • [22] X. Hu, L. Mu, and X. Ye. Weak Galerkin method for the Biot’s consolidation model. Comput. Math. Appl., 75(6):2017–2030, 2018.
  • [23] H. H. Kim, E. T. Chung, and C. Lee. A staggered discontinuous Galerkin method for the Stokes system. SIAM J. Numer. Anal., 51(6):3327–3350, 2013.
  • [24] J. Korsawe and G. Starke. A least-squares mixed finite element method for Biot’s consolidation problem in porous media. SIAM J. Numer. Anal., 43(1):318–339, 2005.
  • [25] S. A. Leake and P. A. Hsieh. Simulation of deformation of sediments from decline of ground-water levels in an aquifer underlain by a bedrock step. U.S. Geological Survey Open File Report: 97-47, 1997.
  • [26] J. J. Lee. Robust error analysis of coupled mixed methods for Biot’s consolidation model. J. Sci. Comput., 69(2):610–632, 2016.
  • [27] J. J. Lee. Robust three-field finite element methods for Biot’s consolidation model in poroelasticity. BIT Numer. Math, 58(2):347–372, 2018.
  • [28] J. J. Lee and H. H. Kim. Analysis of a staggered discontinuous Galerkin method for linear elasticity. J. Sci. Comput., 66(2):625–649, 2016.
  • [29] J. J. Lee, K.-A. Mardal, and R. Winther. Parameter-robust discretization and preconditioning of Biot’s consolidation model. SIAM J. Sci. Comput., 39(1):A1–A24, 2017.
  • [30] A. Mikelić and M. F. Wheeler. Convergence of iterative coupling for coupled flow and geomechanics. Comput. Geosci., 17(3):455–461, 2013.
  • [31] M. A. Murad and A. F. D. Loula. Improved accuracy in finite element analysis of Biot’s consolidation problem. Comput. Methods Appl. Mech. Engrg., 95(3):359–382, 1992.
  • [32] M. A. Murad and A. F. D. Loula. On stability and convergence of finite element approximations of Biot’s consolidation problem. Internat. J. Numer. Methods Engrg., 37(4):645–667, 1994.
  • [33] M. A. Murad, V. Thomée, and A. F. D. Loula. Asymptotic behavior of semidiscrete finite-element approximations of Biot’s consolidation problem. SIAM J. Numer. Anal., 33(3):1065–1083, 1996.
  • [34] C. Niu, H. Rui, and X. Hu. A stabilized hybrid mixed finite element method for poroelasticity. Comput. Geosci., 2020.
  • [35] J. M. Nordbotten. Stable cell-centered finite volume discretization for Biot equations. SIAM J. Numer. Anal., 54(2):942–968, 2016.
  • [36] R. Oyarzúa and R. Ruiz-Baier. Locking-free finite element methods for poroelasticity. SIAM J. Numer. Anal., 54(5):2951–2973, 2016.
  • [37] P. J. Phillips and M. F. Wheeler. A coupling of mixed and continuous Galerkin finite element methods for poroelasticity I: the continuous in time case. Comput. Geosci., 11(2):131–144, 2007.
  • [38] P. J. Phillips and M. F. Wheeler. A coupling of mixed and continuous Galerkin finite element methods for poroelasticity II: the discrete-in-time case. Comput. Geosci., 11(2):145–158, 2007.
  • [39] P. J. Phillips and M. F. Wheeler. Overcoming the problem of locking in linear elasticity and poroelasticity: an heuristic approach. Comput. Geosci., 13(1):5–12, 2009.
  • [40] C. Rodrigo, F. J. Gaspar, X. Hu, and L.T. Zikatanov. Stability and monotonicity for some discretizations of the Biot’s consolidation model. Comput. Methods Appl. Mech. Engrg., 298(1):183–204, 2016.
  • [41] R. E. Showalter. Diffusion in poro-elastic media. J. Math. Anal. Appl., 251(1):310–340, 2000.
  • [42] R. E. Showalter. Nonlinear degenerate evolution equations in mixed formulation. SIAM J. Numer. Anal., 42(5):2114–2131, 2010.
  • [43] R. E. Showalter. Monotone operators in Banach space and nonlinear partial differential equations. American Mathematical Society, 2013.
  • [44] M. Wheeler, G. Xue, and I. Yotov. Coupling multipoint flux mixed finite element methods with continuous Galerkin methods for poroelasticity. Comput. Geosci., 18(1):57–75, 2014.
  • [45] S.-Y. Yi. A coupling of nonconforming and mixed finite element methods for Biot’s consolidation model. Numer. Methods Partial Differential Equations, 29(5):1749–1777, 2013.
  • [46] S.-Y. Yi. Convergence analysis of a new mixed finite element method for Biot’s consolidation model. Numer. Methods Partial Differential Equations, 30(4):1189–1210, 2014.
  • [47] S.-Y. Yi. A study of two modes of locking in poroelasticity. SIAM J. Numer. Anal., 55(4):1915–1936, 2017.
  • [48] L. Zhao, E. T. Chung, and M. Lam. A new staggered DG method for the Brinkman problem robust in the Darcy and Stokes limits. Comput. Methods Appl. Mech. Engrg., 364(1), 2020.
  • [49] L. Zhao, E. T. Chung, E.-J. Park, and G. Zhou. Staggered DG method for coupling of the Stokes and Darcy-Forchheimer problems. SIAM J. Numer. Anal., 2020, to appear.
  • [50] L. Zhao, D. Kim, E.-J. Park, and E. Chung. Staggered DG method with small edges for Darcy flows in fractured porous media. arXiv:2005.10955, 2020.
  • [51] L. Zhao and E.-J. Park. A staggered discontinuous Galerkin method of minimal dimension on quadrilateral and polygonal meshes. SIAM J. Sci. Comput., 40(4):A2543–A2567, 2018.
  • [52] L. Zhao and E.-J. Park. A lowest-order staggered DG method for the coupled Stokes–Darcy problem. IMA J. Numer. Anal., 40(4):2871–2897, 2020.
  • [53] L. Zhao and E.-J. Park. A staggered cell-centered DG method for linear elasticity on polygonal meshes. SIAM J. Sci. Comput., 42(4):A2158–A2181, 2020.
  • [54] L. Zhao, E.-J. Park, and D. w. Shin. A staggered DG method of minimal dimension for the Stokes equations on general meshes. Comput. Methods Appl. Mech. Engrg., 345(1):854–875, 2019.