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

Consistency enforcement for the iterative solution of weak Galerkin finite element approximation of Stokes flow

Weizhang Huang Department of Mathematics, the University of Kansas, 1460 Jayhawk Blvd, Lawrence, KS 66045, USA (). [email protected]    Zhuoran Wang Department of Mathematics, the University of Kansas, 1460 Jayhawk Blvd, Lawrence, KS 66045, USA (). [email protected]

Abstract. Finite element discretization of Stokes problems can result in singular, inconsistent saddle point linear algebraic systems. This inconsistency can cause many iterative methods to fail to converge. In this work, we consider the lowest-order weak Galerkin finite element method to discretize Stokes flow problems and study a consistency enforcement by modifying the right-hand side of the resulting linear system. It is shown that the modification of the scheme does not affect the optimal-order convergence of the numerical solution. Moreover, inexact block diagonal and triangular Schur complement preconditioners and the minimal residual method (MINRES) and the generalized minimal residual method (GMRES) are studied for the iterative solution of the modified scheme. Bounds for the eigenvalues and the residual of MINRES/GMRES are established. Those bounds show that the convergence of MINRES and GMRES is independent of the viscosity parameter and mesh size. The convergence of the modified scheme and effectiveness of the preconditioners are verified using numerical examples in two and three dimensions.

Keywords: Stokes flow, GMRES, MINRES, Weak Galerkin, Compatibility condition.

Mathematics Subject Classification (2020): 65N30, 65F08, 65F10, 76D07

1 Introduction

We consider the lowest-order weak Galerkin (WG) finite element approximation of Stokes flow problems in the form

{μΔ𝐮+p=𝐟,in Ω,𝐮=0, in Ω,𝐮=𝐠,on Ω,\begin{cases}\displaystyle-\mu\Delta\mathbf{u}+\nabla p=\mathbf{f},\quad\mbox{in }\;\Omega,\\ \displaystyle\nabla\cdot\mathbf{u}=0,\quad\text{ in }\Omega,\\ \displaystyle\mathbf{u}=\mathbf{g},\quad\mbox{on }\;\partial\Omega,\end{cases} (1)

where Ωd\Omega\subset\mathbb{R}^{d} (d=2,3)(d=2,3) is a bounded polygonal/polyhedral domain, μ>0\mu>0 is the fluid kinematic viscosity, 𝐮\mathbf{u} is the fluid velocity, pp is the fluid pressure, 𝐟\mathbf{f} is a body force, 𝐠\mathbf{g} is a non-zero boundary datum of the velocity satisfying the compatibility condition Ω𝐠𝐧=0\int_{\partial\Omega}\mathbf{g}\cdot\mathbf{n}=0, and 𝐧\mathbf{n} is the unit outward normal to the boundary of the domain. The resulting linear algebraic system is a singular saddle point system where the singularity reflects the nonuniqueness of the pressure solution. While it is well known (e.g., see [4, Remark 6.12] and [16, Section 10.2]) that Krylov subspace methods, such as the minimal residual method (MINRES) and the generalized minimal residual method (GMRES), work well for consistent singular systems, the underlying system is nonconsistent in general when gg is not identically zero (cf. Section 2). This inconsistency can cause MINRES and GMRES to fail to converge.

The objective of this work is to study a simple consistency enforcement strategy by modifying the right-hand side of the linear system that results from the WG discretization of (1). We shall prove that the optimal-order convergence is not affected by the modification. Moreover, we consider inexact block diagonal and triangular Schur complement preconditioners for the efficient iterative solution of the singular but consistent saddle point system resulting from the modified scheme. Bounds for the eigenvalues and the residual of MINRES and GMRES for the corresponding preconditioned systems are established. These bounds show that the convergence of MINRES and GMRES with the inexact block diagonal and triangular Schur complement preconditioning is independent of the fluid kinematic viscosity μ\mu and mesh size. It is worth mentioning that the preconditioned system associated a block diagonal Schur complement preconditioner is diagonalizable and this MINRES can be used for system solving and spectral analysis can be used to analyze the convergence of MINRES. On the other hand, the preconditioned system with a block triangular Schur complement preconditioner is not diagonalizable and we need to use GMRES for the iterative solution of the system. Moreover, the spectral analysis cannot be used directly to analyze the convergence of GMRES. Instead, it needs to be combined with Lemmas A.1 and A.2 of [7] that provide a bound for the residual of GMRES in terms of the norm of the off-diagonal blocks in the preconditioned system and the performance of GMRES to the preconditioned Schur complement.

Numerical solution of Stokes flow problems has continuously gained attention from researchers. Particularly, a variety of finite element methods have been studied for those problems; e.g., see [1] (mixed finite element methods), [3, 17] (virtual element methods), [8, 15] (hybrid discontinuous Galerkin methods), and [11, 14, 18] (weak Galerkin (WG) finite element methods). We use here the lowest-order WG method for the discretization of Stokes flow problems. It is known (cf. Lemma 2.1 or [19]) that the lowest-order WG method, without using stabilization terms, satisfies the inf-sup condition (for stability) and has the optimal-order convergence. Moreover, the error in the velocity is independent of the error in the pressure (pressure-robustness) and the error in the pressure is independent of the viscosity μ\mu (μ\mu-semi-robustness). On the other hand, little work has been done so far for the efficient iterative solution of the saddle point system arising from the WG approximation of Stokes problems. Recently, the authors considered in [7] the iterative solution of the lowest-order WG approximation of Stokes problems through regularization and provided a convergence analysis for MINRES and GMRES with inexact block diagonal and triangular Schur complement preconditioners, respectively.

The rest of this paper is organized as follows. In Section 2, the weak formulation for Stokes flow and its discretization by the lowest-order WG method are described. Consistency enforcement and related error analysis are discussed in Section 3. In Section 4, the block diagonal and block triangular Schur complement preconditioning and convergence of MINRES and GMRES for the modified system are studied. Section 5 presents two- and three dimensional numerical experiments to verify the theoretical findings and showcase the effectiveness of the preconditioners. Finally, the conclusions are drawn in Section 6.

2 Weak Galerkin discretization for Stokes flow

In this section we describe the lowest-order weak Galerkin approximation of the Stokes flow problem (1).

The weak formulation of (1) is to find 𝐮H1(Ω)d\mathbf{u}\in H^{1}(\Omega)^{d} and pL2(Ω)p\in L^{2}(\Omega) such that 𝐮|Ω=𝐠\mathbf{u}|_{\partial\Omega}=\mathbf{g} (in the weak sense) and

{μ(𝐮,𝐯)(p,𝐯)=(𝐟,𝐯),𝐯H01(Ω)d,(𝐮,q)=0,qL2(Ω).\begin{cases}\mu(\nabla\mathbf{u},\nabla\mathbf{v})-(p,\nabla\cdot\mathbf{v})=(\mathbf{f},\mathbf{v}),\quad\forall\mathbf{v}\in H^{1}_{0}(\Omega)^{d},\\ -(\nabla\cdot\mathbf{u},q)=0,\quad\forall q\in L^{2}(\Omega).\end{cases} (2)

Consider a connected quasi-uniform simplicial mesh 𝒯h={K}\mathcal{T}_{h}=\{K\} on Ω\Omega. A mesh is called to be connected if any pair of its elements is linked by a chain of elements that share an interior facet with each other. We introduce the following discrete weak function spaces on Ω\Omega:

𝐕h\displaystyle\displaystyle\mathbf{V}_{h} ={𝐮h={𝐮h,𝐮h}:𝐮h|KP0(K)d,𝐮h|eP0(e)d,K𝒯h,eK},\displaystyle=\{\mathbf{u}_{h}=\{\mathbf{u}_{h}^{\circ},\mathbf{u}_{h}^{\partial}\}:\;\mathbf{u}_{h}^{\circ}|_{K}\in P_{0}(K)^{d},\;\mathbf{u}_{h}^{\partial}|_{e}\in P_{0}(e)^{d},\;\forall K\in\mathcal{T}_{h},\;e\in\partial K\}, (3)
Wh\displaystyle\displaystyle W_{h} ={phL2(Ω):ph|KP0(K),K𝒯h},\displaystyle=\{p_{h}\in L^{2}(\Omega):\;p_{h}|_{K}\in P_{0}(K),\;\forall K\in\mathcal{T}_{h}\}, (4)

where P0(K)P_{0}(K) and P0(e)P_{0}(e) denote the set of constant polynomials defined on element KK and facet ee, respectively. Note that 𝐮h𝐕h\mathbf{u}_{h}\in\mathbf{V}_{h} is approximated on both interiors and facets of mesh elements while phWhp_{h}\in W_{h} is approximated only on element interiors. Then, we can define the lowest-order WG approximation of the Stokes problem (1) as: finding 𝐮h𝐕h\mathbf{u}_{h}\in\mathbf{V}_{h} and phWhp_{h}\in W_{h} such that 𝐮h|Ω=Qh𝐠\mathbf{u}_{h}^{\partial}|_{\partial\Omega}=Q_{h}^{\partial}\mathbf{g} and

{μK𝒯h(w𝐮h,w𝐯)KK𝒯h(ph,w𝐯)K=K𝒯h(𝐟,𝚲h𝐯)K,𝐯𝐕h0,K𝒯h(w𝐮h,q)K=0,qWh,\begin{cases}\displaystyle\mu\sum_{K\in\mathcal{T}_{h}}(\nabla_{w}\mathbf{u}_{h},\nabla_{w}\mathbf{v})_{K}-\sum_{K\in\mathcal{T}_{h}}(p_{h}^{\circ},\nabla_{w}\cdot\mathbf{v})_{K}=\sum_{K\in\mathcal{T}_{h}}(\mathbf{f},\mathbf{\Lambda}_{h}\mathbf{v})_{K},\quad\forall\mathbf{v}\in\mathbf{V}_{h}^{0},\\ \displaystyle-\sum_{K\in\mathcal{T}_{h}}(\nabla_{w}\cdot\mathbf{u}_{h},q^{\circ})_{K}=0,\quad\forall q\in W_{h},\end{cases} (5)

where QhQ_{h}^{\partial} is a L2L^{2}-projection operator onto 𝐕h\mathbf{V}_{h} restricted on each facet and the lifting operator 𝚲h:𝐕hRT0(𝒯h)\mathbf{\Lambda}_{h}:\mathbf{V}_{h}\to RT_{0}(\mathcal{T}_{h}) is defined [10, 19] as

((𝚲h𝐯)𝐧,w)e=(𝐯𝐧,w)e,wP0(e),𝐯𝐕h,eK.\displaystyle((\mathbf{\Lambda}_{h}\mathbf{v})\cdot\mathbf{n},w)_{e}=(\mathbf{v}^{\partial}\cdot\mathbf{n},w)_{e},\quad\forall w\in P_{0}(e),\;\forall\mathbf{v}\in\mathbf{V}_{h},\;\forall e\subset\partial K. (6)

Here, RT0(K)RT_{0}(K) is the lowest-order Raviart-Thomas space, viz.,

RT0(K)=(P0(K))d+𝐱P0(K).\displaystyle RT_{0}(K)=(P_{0}(K))^{d}+\mathbf{x}\,P_{0}(K).

Notice that 𝚲h𝐯\mathbf{\Lambda}_{h}\mathbf{v} depends on 𝐯\mathbf{v}^{\partial} but not on 𝐯\mathbf{v}^{\circ}.

The discrete weak gradient and divergence operators in (5) are defined as follows. For a scalar function or a component of a vector-valued function, uh=(uh,uh)u_{h}=(u_{h}^{\circ},u_{h}^{\partial}), the discrete weak gradient operator w:WhRT0(𝒯h)\nabla_{w}:W_{h}\rightarrow RT_{0}(\mathcal{T}_{h}) is defined as follows

(wuh,𝐰)K=(uh,𝐰𝐧)K(uh,𝐰)K,𝐰RT0(K),K𝒯h,(\nabla_{w}u_{h},\mathbf{w})_{K}=(u^{\partial}_{h},\mathbf{w}\cdot\mathbf{n})_{\partial K}-(u^{\circ}_{h},\nabla\cdot\mathbf{w})_{K},\quad\forall\mathbf{w}\in RT_{0}(K),\quad\forall K\in\mathcal{T}_{h}, (7)

where 𝐧\mathbf{n} is the unit outward normal to K\partial K and (,)K(\cdot,\cdot)_{K} and (,)K(\cdot,\cdot)_{\partial K} are the L2L^{2} inner product on KK and K\partial K, respectively. For a vector-valued function 𝐮h\mathbf{u}_{h}, w𝐮h\nabla_{w}\mathbf{u}_{h} is viewed as a matrix with each row representing the weak gradient of a component. By choosing 𝐰\mathbf{w} properly in (7) and using the fact that wuhRT0(K)\nabla_{w}u_{h}\in RT_{0}(K), we can obtain (e.g., see [6])

wφK=CK(𝐱𝐱K),\displaystyle\nabla_{w}\varphi_{K}^{\circ}=-C_{K}(\mathbf{x}-\mathbf{x}_{K}), (8)
wφK,i=CKd+1(𝐱𝐱K)+|eK,i||K|𝐧K,i,i=1,,d+1,\displaystyle\nabla_{w}\varphi_{K,i}^{\partial}=\frac{C_{K}}{d+1}(\mathbf{x}-\mathbf{x}_{K})+\frac{|e_{K,i}|}{|K|}\mathbf{n}_{K,i},\;i=1,...,d+1, (9)

where φK\varphi_{K}^{\circ} and φK,i\varphi_{K,i}^{\partial} denote the basis functions of P0(K)P_{0}(K) and P0(eK,i)P_{0}(e_{K,i}), respectively, eK,ie_{K,i} denotes the ii-th facet of KK, 𝐧K,i\mathbf{n}_{K,i} is the unit outward normal to eK,ie_{K,i},

CK=d|K|𝐱𝐱KK2,𝐱K=1d+1i=1d+1𝐱K,i,\displaystyle C_{K}=\frac{d\;|K|}{\|\mathbf{x}-\mathbf{x}_{K}\|_{K}^{2}},\quad\mathbf{x}_{K}=\frac{1}{d+1}\sum_{i=1}^{d+1}\mathbf{x}_{K,i},

and 𝐱K,i\mathbf{x}_{K,i}, i=1,,d+1i=1,...,d+1 denote the vertices of KK.

The discrete weak divergence operator w:𝐕h𝒫0(𝒯h)\nabla_{w}\cdot:\mathbf{V}_{h}\to\mathcal{P}_{0}(\mathcal{T}_{h}) is defined independently through

(w𝐮,w)K=(𝐮,w𝐧)e(𝐮,w)K,wP0(K).(\nabla_{w}\cdot\mathbf{u},w)_{K}=(\mathbf{u}^{\partial},w\mathbf{n})_{e}-(\mathbf{u}^{\circ},\nabla w)_{K},\quad\forall w\in P_{0}(K). (10)

Note that w𝐮|KP0(K)\nabla_{w}\cdot\mathbf{u}|_{K}\in P_{0}(K). Moreover, by taking w=1w=1 we have

(w𝐮,1)K=i=1d+1|eK,i|𝐮eK,iT𝐧K,i,(\nabla_{w}\cdot\mathbf{u},1)_{K}=\sum_{i=1}^{d+1}|e_{K,i}|\langle\mathbf{u}\rangle_{e_{K,i}}^{T}\mathbf{n}_{K,i}, (11)

where 𝐮eK,i\langle\mathbf{u}\rangle_{e_{K,i}} denotes the average of 𝐮\mathbf{u} on facet eK,ie_{K,i} and |eK,i||e_{K,i}| is the (d1)(d-1)-dimensional measure of eK,ie_{K,i}.

The scheme (5) achieves the optimal-order convergence as shown in the following lemma.

      Lemma 2.1.

Let 𝐮H2(Ω)d\mathbf{u}\in H^{2}(\Omega)^{d} and pH1(Ω)p\in H^{1}(\Omega) be the exact solutions for the Stokes problem (2) and let 𝐮h𝐕h\mathbf{u}_{h}\in\mathbf{V}_{h} and phWhp_{h}\in W_{h} be numerical solutions for (5). Assume that 𝐟L2(Ω)d\mathbf{f}\in L^{2}(\Omega)^{d}. Then, the followings hold true:

pphCh𝐟,\displaystyle\|p-p_{h}\|\leq Ch\|\mathbf{f}\|, (12)
𝐮w𝐮hCh𝐮2,\displaystyle\|\nabla\mathbf{u}-\nabla_{w}\mathbf{u}_{h}\|\leq Ch\|\mathbf{u}\|_{2}, (13)
𝐮𝐮h=𝐮𝐮hCh𝐮2,\displaystyle\|\mathbf{u}-\mathbf{u}_{h}\|=\|\mathbf{u}-\mathbf{u}_{h}^{\circ}\|\leq Ch\|\mathbf{u}\|_{2}, (14)
Qh𝐮𝐮hCh2𝐮2,\displaystyle\|Q_{h}^{\circ}\mathbf{u}-\mathbf{u}_{h}^{\circ}\|\leq Ch^{2}\|\mathbf{u}\|_{2}, (15)

where =L2(Ω)\|\cdot\|=\|\cdot\|_{L^{2}(\Omega)}, 2=H2(Ω)\|\cdot\|_{2}=\|\cdot\|_{H^{2}(\Omega)}, CC is a constant independent of hh and μ\mu, and QhQ_{h}^{\circ} is a L2L^{2}-projection operator for element interiors satisfying Qh𝐮|K=𝐮K,K𝒯hQ_{h}^{\circ}\mathbf{u}|_{K}=\langle\mathbf{u}\rangle_{K},\;\forall K\in\mathcal{T}_{h}.

Proof.

The proof of these error estimates can be found in [12, Theorem 4.5] and [19, Theorem 3]. ∎

To cast (5) into a matrix-vector form, hereafter we use 𝐯h\mathbf{v}_{h} interchangeably for the WG approximation of 𝐯\mathbf{v} in 𝐕h\mathbf{V}_{h} and the vector formed by its values (𝐯h,K,𝐯h,K,i)(\mathbf{v}_{h,K}^{\circ},\mathbf{v}_{h,K,i}^{\partial}) for i=1,,d+1i=1,...,d+1 and K𝒯hK\in\mathcal{T}_{h}, excluding those on Ω\partial\Omega. Similarly, 𝐪h\mathbf{q}_{h} is used to denote both the WG approximation of qq and the vector formed by qh,Kq_{h,K} for all K𝒯hK\in\mathcal{T}_{h}. Then, we can write (5) into

[μA(B)TB𝟎][𝐮h𝐩h]=[𝐛1𝐛2],\begin{bmatrix}\mu A&-(B^{\circ})^{T}\\ -B^{\circ}&\mathbf{0}\end{bmatrix}\begin{bmatrix}\mathbf{u}_{h}\\ \mathbf{p}_{h}\end{bmatrix}=\begin{bmatrix}\mathbf{b}_{1}\\ \mathbf{b}_{2}\end{bmatrix}, (16)

where the matrices AA and BB^{\circ} and vectors 𝐛1\mathbf{b}_{1} and 𝐛2\mathbf{b}_{2} are defined as

𝐯TA𝐮h\displaystyle\mathbf{v}^{T}A\mathbf{u}_{h} =K𝒯h(w𝐮h,w𝐯)K=K𝒯h(𝐮h,KwφK,w𝐯)K\displaystyle=\sum_{K\in\mathcal{T}_{h}}(\nabla_{w}\mathbf{u}_{h},\nabla_{w}\mathbf{v})_{K}=\sum_{K\in\mathcal{T}_{h}}(\mathbf{u}_{h,K}^{\circ}\nabla_{w}\varphi_{K}^{\circ},\nabla_{w}\mathbf{v})_{K} (17)
+K𝒯hi=1eK,iΩd+1(𝐮h,K,iwφK,i,w𝐯)K,𝐮h,𝐯𝐕h0,\displaystyle\displaystyle+\sum_{K\in\mathcal{T}_{h}}\sum^{d+1}_{\begin{subarray}{c}i=1\\ e_{K,i}\notin\partial\Omega\end{subarray}}(\mathbf{u}_{h,K,i}^{\partial}\nabla_{w}\varphi_{K,i}^{\partial},\nabla_{w}\mathbf{v})_{K},\quad\forall\mathbf{u}_{h},\mathbf{v}\in\mathbf{V}_{h}^{0},
𝐪TB𝐮h\displaystyle\mathbf{q}^{T}B^{\circ}\mathbf{u}_{h} =K𝒯h(w𝐮h,q)K=K𝒯hi=1eK,iΩd+1|eK,i|qK(𝐮h,K,i)T𝐧K,i,𝐮h𝐕h0,qWh,\displaystyle=\sum_{K\in\mathcal{T}_{h}}(\nabla_{w}\cdot\mathbf{u}_{h},q^{\circ})_{K}=\sum_{K\in\mathcal{T}_{h}}\sum^{d+1}_{\begin{subarray}{c}i=1\\ e_{K,i}\notin\partial\Omega\end{subarray}}|e_{K,i}|q_{K}^{\circ}(\mathbf{u}_{h,K,i}^{\partial})^{T}\mathbf{n}_{K,i},\quad\forall\mathbf{u}_{h}\in\mathbf{V}_{h}^{0},\quad\forall q\in W_{h}, (18)
𝐯T𝐛1\displaystyle\mathbf{v}^{T}\mathbf{b}_{1} =K𝒯h(𝐟,𝚲h𝐯)KμK𝒯hi=1eK,iΩd+1((Qh𝐠)wφK,i,w𝐯)K,𝐯𝐕h0,\displaystyle=\sum_{K\in\mathcal{T}_{h}}(\mathbf{f},\mathbf{\Lambda}_{h}\mathbf{v})_{K}-\mu\sum_{K\in\mathcal{T}_{h}}\sum_{\begin{subarray}{c}i=1\\ e_{K,i}\in\partial\Omega\end{subarray}}^{d+1}((Q_{h}^{\partial}\mathbf{g})\nabla_{w}\varphi_{K,i}^{\partial},\nabla_{w}\mathbf{v})_{K},\;\forall\mathbf{v}\in\mathbf{V}_{h}^{0}, (19)
𝐪T𝐛2\displaystyle\mathbf{q}^{T}\mathbf{b}_{2} =K𝒯hi=1eK,iΩd+1|eK,i|qK(Qh𝐠|eK,i)T𝐧K,i,qWh.\displaystyle=\sum_{K\in\mathcal{T}_{h}}\sum^{d+1}_{\begin{subarray}{c}i=1\\ e_{K,i}\in\partial\Omega\end{subarray}}|e_{K,i}|q_{K}^{\circ}(Q_{h}^{\partial}\mathbf{g}|_{e_{K,i}})^{T}\mathbf{n}_{K,i},\quad\forall q\in W_{h}. (20)

Notice that (16) is a singular saddle point system where the pressure solution is unique up to a constant. Moreover, the system is inconsistent in general when 𝐠\mathbf{g} is not identically zero. To explain this, from (20) we have

𝐛2(K)\displaystyle\mathbf{b}_{2}(K) =i=1eK,iΩd+1|eK,i|(Qh𝐠|eK,i)T𝐧K,i,K𝒯h.\displaystyle=\sum^{d+1}_{\begin{subarray}{c}i=1\\ e_{K,i}\in\partial\Omega\end{subarray}}|e_{K,i}|(Q_{h}^{\partial}\mathbf{g}|_{e_{K,i}})^{T}\mathbf{n}_{K,i},\quad\forall K\in\mathcal{T}_{h}. (21)

This gives rise to

K𝒯h𝐛2(K)\displaystyle\sum_{K\in\mathcal{T}_{h}}\mathbf{b}_{2}(K) =K𝒯hi=1eK,iΩd+1|eK,i|(Qh𝐠|eK,i)T𝐧K,i\displaystyle=\sum_{K\in\mathcal{T}_{h}}\sum^{d+1}_{\begin{subarray}{c}i=1\\ e_{K,i}\in\partial\Omega\end{subarray}}|e_{K,i}|(Q_{h}^{\partial}\mathbf{g}|_{e_{K,i}})^{T}\mathbf{n}_{K,i} (22)
=eΩ|e|(Qh𝐠|e)T𝐧e=eΩe(Qh𝐠|e)T𝐧𝑑S,\displaystyle=\sum_{e\in\partial\Omega}|e|(Q_{h}^{\partial}\mathbf{g}|_{e})^{T}\mathbf{n}_{e}=\sum_{e\in\partial\Omega}\int_{e}(Q_{h}^{\partial}\mathbf{g}|_{e})^{T}\mathbf{n}dS,

where we have used the fact that Ω\partial\Omega consists of element facets because Ω\Omega is assumed to be a polygon or a polyhedron. If we choose Qh𝐠|e=𝐠eQ_{h}^{\partial}\mathbf{g}|_{e}=\langle\mathbf{g}\rangle_{e}, eΩ\forall e\in\partial\Omega, from the compatibility condition we have

K𝒯h𝐛2(K)=eΩe𝐠eT𝐧𝑑S=eΩe𝐠T𝐧𝑑S=0.\sum_{K\in\mathcal{T}_{h}}\mathbf{b}_{2}(K)=\sum_{e\in\partial\Omega}\int_{e}\langle\mathbf{g}\rangle_{e}^{T}\mathbf{n}dS=\sum_{e\in\partial\Omega}\int_{e}\mathbf{g}^{T}\mathbf{n}dS=0.

Unfortunately, in general the averages 𝐠e\langle\mathbf{g}\rangle_{e} need to be approximated numerically. The approximation error can cause a non-zero K𝒯h𝐛2(K)\sum_{K\in\mathcal{T}_{h}}\mathbf{b}_{2}(K) and thus the inconsistency of the system. This can also happen if QhQ_{h}^{\partial} is defined differently, for instance, Qh𝐠|eQ_{h}^{\partial}\mathbf{g}|_{e} is defined as the value of 𝐠\mathbf{g} at the barycenter of ee. Krylov subspace methods, like MINRES and GMRES, may fail to converge when applied to this inconsistent singular system although they are known to work well for consistent singular systems, at least when the initial guess is taken as zero; e.g., see [4, Remark 6.12] and [16, Section 10.2].

3 Consistency enforcement and its effects on the optimal-order convergence

In this section we consider an approach of enforcing the consistency/compatibility condition by modifying the right-hand term 𝐛2\mathbf{b}_{2} and study the effects of this modification on the optimal-order convergence of the numerical solution.

We propose to modify 𝐛2\mathbf{b}_{2} to enforce the consistency. Specifically, we define

𝐛~2(K)=𝐛2(K)αhN,K𝒯h,\displaystyle\tilde{\mathbf{b}}_{2}(K)=\mathbf{b}_{2}(K)-\frac{\alpha_{h}}{N},\quad\forall K\in\mathcal{T}_{h}, (23)

where NN is the number of elements in 𝒯h\mathcal{T}_{h} and

αh=K𝒯h𝐛2(K)=eΩe(Qh𝐠|e)T𝐧𝑑S.\displaystyle\alpha_{h}=\sum_{K\in\mathcal{T}_{h}}\mathbf{b}_{2}(K)=\sum_{e\in\partial\Omega}\int_{e}(Q_{h}^{\partial}\mathbf{g}|_{e})^{T}\mathbf{n}dS.

By definition, we have K𝐛~2(K)=0\sum_{K}\tilde{\mathbf{b}}_{2}(K)=0. Moreover, from the compatibility condition Ω𝐠𝐧𝑑S=0\int_{\partial\Omega}\mathbf{g}\cdot\mathbf{n}dS=0, we can rewrite αh\alpha_{h} as

αh=eΩe(Qh𝐠|e𝐠)T𝐧𝑑S=eΩ|e|(Qh𝐠|e𝐠e)T𝐧e.\displaystyle\alpha_{h}=\sum_{e\in\partial\Omega}\int_{e}(Q_{h}^{\partial}\mathbf{g}|_{e}-\mathbf{g})^{T}\mathbf{n}dS=\sum_{e\in\partial\Omega}|e|\left(Q_{h}^{\partial}\mathbf{g}|_{e}-\langle\mathbf{g}\rangle_{e}\right)^{T}\mathbf{n}_{e}. (24)

The modified scheme reads as

[μA(B)TB𝟎][𝐮h𝐩h]=[𝐛1𝐛~2],\begin{bmatrix}\mu A&-(B^{\circ})^{T}\\ -B^{\circ}&\mathbf{0}\end{bmatrix}\begin{bmatrix}\mathbf{u}_{h}\\ \mathbf{p}_{h}\end{bmatrix}=\begin{bmatrix}\mathbf{b}_{1}\\[3.61371pt] \tilde{\mathbf{b}}_{2}\end{bmatrix}, (25)

where AA, BB^{\circ}, and 𝐛1\mathbf{b}_{1} are given in (17), (18), and (19), respectively. Note that (25) is still singular but consistent.

An immediate question about the modified scheme is how much the accuracy of the numerical solution is affected by the modification. To answer this, we provide an error analysis for the modified scheme (25) in the following.

Let 𝐮H2(Ω)d\mathbf{u}\in H^{2}(\Omega)^{d} and pH1(Ω)p\in H^{1}(\Omega) be the solutions of the Stokes problem (1) and 𝐮h𝐕h\mathbf{u}_{h}\in\mathbf{V}_{h} and phWhp_{h}\in W_{h} be the numerical solutions of (25). We split the error into the projection and discrete error as

{𝐮𝐮h=(𝐮Qh𝐮)+𝐞h,𝐞h=Qh𝐮𝐮h,pph=(pQhp)+ehp,ehp=Qhpph,\begin{cases}\displaystyle\mathbf{u}-\mathbf{u}_{h}=(\mathbf{u}-Q_{h}\mathbf{u})+\mathbf{e}_{h},\quad\mathbf{e}_{h}=Q_{h}\mathbf{u}-\mathbf{u}_{h},\\ \displaystyle p-p_{h}=(p-Q_{h}p)+e_{h}^{p},\qquad e_{h}^{p}=Q_{h}p-p_{h},\end{cases} (26)

where the L2L^{2} projection operator is defined as Qh=(Qh,Qh)Q_{h}=(Q_{h}^{\circ},Q_{h}^{\partial}). The operator QhQ_{h} is considered to be componentwise when applied to vector-valued functions. The projection error 𝐮Qh𝐮\mathbf{u}-Q_{h}\mathbf{u} and pQhpp-Q_{h}p are determined by the approximation capacity of the finite element spaces. Our primary focus is on 𝐞h\mathbf{e}_{h} and ehpe_{h}^{p}.

      Lemma 3.1.

There hold

we2C(h1wK2+hwK2),wH1(K),eK,K𝒯h,\displaystyle\|w\|_{e}^{2}\leq C(h^{-1}\|w\|_{K}^{2}+h\|\nabla w\|_{K}^{2}),\quad\forall w\in H^{1}(K),\quad\forall e\in\partial K,\quad\forall K\in\mathcal{T}_{h}, (27)
WK2ChW𝐧K2,WRT0d,K𝒯h,\displaystyle\|W\|_{K}^{2}\leq Ch\|W\mathbf{n}\|^{2}_{\partial K},\quad\forall W\in RT_{0}^{d},\quad\quad\forall K\in\mathcal{T}_{h}, (28)
K𝒯hh1𝐯𝐯K2C|𝐯|2,𝐯𝐕h0,\displaystyle\displaystyle\sum_{K\in\mathcal{T}_{h}}h^{-1}\|\mathbf{v}^{\partial}-\mathbf{v}^{\circ}\|_{\partial K}^{2}\leq C|||\mathbf{v}|||^{2},\quad\forall\mathbf{v}\in\mathbf{V}_{h}^{0}, (29)
K𝒯h𝚲h𝐯𝐯KC(K𝒯heKh𝐯𝐯e2)1/2,𝐯𝐕h,\displaystyle\sum_{K\in\mathcal{T}_{h}}\|\mathbf{\Lambda}_{h}\mathbf{v}-\mathbf{v}^{\circ}\|_{K}\leq C(\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}h\|\mathbf{v}^{\partial}-\mathbf{v}^{\circ}\|_{e}^{2})^{1/2},\quad\forall\mathbf{v}\in\mathbf{V}_{h}, (30)

where |𝐯|=w𝐯|||\mathbf{v}|||=\|\nabla_{w}\mathbf{v}\|.

Proof.

These results can be found in [19, Lemma 2] for (27), [9] for (28), [19, Lemma 3] for (29), [19, Lemma 4] for (30). ∎

      Lemma 3.2.

There exists a constant β>0\beta>0 independent of μ,h\mu,h such that

sup𝐯𝐕h0,|𝐯|0𝐯T(B)𝐩|𝐯|β𝐩,pWh.\displaystyle\sup_{\mathbf{v}\in\mathbf{V}_{h}^{0},|||\mathbf{v}|||\neq 0}\frac{\mathbf{v}^{T}(B^{\circ})\mathbf{p}}{|||\mathbf{v}|||}\geq\beta\|\mathbf{p}\|,\qquad\forall p\in W_{h}. (31)
Proof.

Proof can be found in [19, Lemma 7]. ∎

      Lemma 3.3.

The error equations for the modified scheme (25) read as

{𝐯TA𝐞h𝐯T(B)T𝐞hp=μ1(𝐮,𝐯)+μ2(𝐮,𝐯),𝐯𝐕h0,𝐪TB𝐞h=αhNK𝒯hq|K,qWh,\begin{cases}\displaystyle\mathbf{v}^{T}A\mathbf{e}_{h}-\mathbf{v}^{T}(B^{\circ})^{T}\mathbf{e}_{h}^{p}=\;\mu\,\mathcal{R}_{1}(\mathbf{u},\mathbf{v})+\mu\,\mathcal{R}_{2}(\mathbf{u},\mathbf{v}),\quad\forall\mathbf{v}\in\mathbf{V}_{h}^{0},\\[5.78172pt] \displaystyle-\mathbf{q}^{T}B^{\circ}\mathbf{e}_{h}=\;\frac{\alpha_{h}}{N}\sum_{K\in\mathcal{T}_{h}}q|_{K},\quad\forall q\in W_{h},\end{cases} (32)

where

{1(𝐮,𝐯)=K𝒯heK(𝐯𝐯,(𝒬h𝐮𝐮)𝐧e)e,2(𝐮,𝐯)=K𝒯h(Δ𝐮,𝚲h𝐯𝐯)K,\left\{\begin{array}[]{l}\displaystyle\mathcal{R}_{1}(\mathbf{u},\mathbf{v})=\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}(\mathbf{v}^{\partial}-\mathbf{v}^{\circ},\,(\mathcal{Q}_{h}\nabla\mathbf{u}-\nabla\mathbf{u})\mathbf{n}_{e})_{e},\\[14.45377pt] \displaystyle\mathcal{R}_{2}(\mathbf{u},\mathbf{v})=\sum_{K\in\mathcal{T}_{h}}(\Delta\mathbf{u},\,\mathbf{\Lambda}_{h}\mathbf{v}-\mathbf{v}^{\circ})_{K},\end{array}\right. (33)

and 𝒬h\mathcal{Q}_{h} is L2L^{2} projection from L2(Ω)d×dL^{2}(\Omega)^{d\times d} onto RT0(𝒯h)dRT_{0}(\mathcal{T}_{h})^{d} space.

Proof.

Using the discrete errors defined in (26), we substitute

𝐮h=Qh𝐮𝐞h,ph=Qhpehp\mathbf{u}_{h}=Q_{h}\mathbf{u}-\mathbf{e}_{h},\quad p_{h}=Q_{h}p-e_{h}^{p}

into the modified scheme (25) and obtain

{μK𝒯h(w𝐞h,w𝐯)KK𝒯h(ehp,w𝐯)K=K𝒯h(𝐟,𝚲h𝐯)K+μK𝒯h(wQh𝐮,w𝐯)KK𝒯h(Qhp,w𝐯)K,𝐯𝐕h0,K𝒯h(w𝐞h,q)K=αhNK𝒯hq|KK𝒯h(w(Qh𝐮),q)K,qWh.\displaystyle\begin{cases}\displaystyle&\mu\sum\limits_{K\in\mathcal{T}_{h}}(\nabla_{w}\mathbf{e}_{h},\nabla_{w}\mathbf{v})_{K}-\sum\limits_{K\in\mathcal{T}_{h}}(e_{h}^{p},\nabla_{w}\cdot\mathbf{v})_{K}\displaystyle=-\sum\limits_{K\in\mathcal{T}_{h}}(\mathbf{f},\mathbf{\Lambda}_{h}\mathbf{v})_{K}\\ &\qquad\qquad\qquad\displaystyle+\;\mu\sum_{K\in\mathcal{T}_{h}}(\nabla_{w}{Q}_{h}\mathbf{u},\nabla_{w}\mathbf{v})_{K}-\sum\limits_{K\in\mathcal{T}_{h}}(Q_{h}p,\nabla_{w}\cdot\mathbf{v})_{K},\quad\forall\mathbf{v}\in\mathbf{V}_{h}^{0},\\[21.68121pt] \displaystyle&-\sum\limits_{K\in\mathcal{T}_{h}}(\nabla_{w}\cdot\mathbf{e}_{h},q^{\circ})_{K}=-\frac{\alpha_{h}}{N}\sum\limits_{K\in\mathcal{T}_{h}}q|_{K}-\sum\limits_{K\in\mathcal{T}_{h}}(\nabla_{w}\cdot({Q}_{h}\mathbf{u}),q)_{K},\quad\forall q\in W_{h}.\end{cases} (34)

Next, we estimate the terms on the right-hand sides. For the first term on the right-hand side of the first equation of (34), by testing the first equation of (1) by 𝚲h𝐯\mathbf{\Lambda}_{h}\mathbf{v} we obtain

K𝒯h(𝐟,𝚲h𝐯)K\displaystyle\sum_{K\in\mathcal{T}_{h}}(\mathbf{f},\mathbf{\Lambda}_{h}\mathbf{v})_{K} =K𝒯hμ(Δ𝐮,𝚲h𝐯)K+K𝒯h(p,𝚲h𝐯)K\displaystyle=\sum_{K\in\mathcal{T}_{h}}-\mu(\Delta\mathbf{u},\mathbf{\Lambda}_{h}\mathbf{v})_{K}+\sum_{K\in\mathcal{T}_{h}}(\nabla p,\mathbf{\Lambda}_{h}\mathbf{v})_{K}
=K𝒯hμ(Δ𝐮,𝚲h𝐯𝐯)KK𝒯hμ(Δ𝐮,𝐯)K+K𝒯h(p,𝚲h𝐯)K.\displaystyle=-\sum_{K\in\mathcal{T}_{h}}\mu(\Delta\mathbf{u},\mathbf{\Lambda}_{h}\mathbf{v}-\mathbf{v}^{\circ})_{K}-\sum_{K\in\mathcal{T}_{h}}\mu(\Delta\mathbf{u},\mathbf{v}^{\circ})_{K}+\sum_{K\in\mathcal{T}_{h}}(\nabla p,\mathbf{\Lambda}_{h}\mathbf{v})_{K}. (35)

For the second term on the right-hand side of the above equation, using the divergence theorem and 𝐯=0\nabla\mathbf{v}^{\circ}=0 we have

K𝒯hμ(Δ𝐮,𝐯)K\displaystyle-\sum_{K\in\mathcal{T}_{h}}\mu(\Delta\mathbf{u},\mathbf{v}^{\circ})_{K} =K𝒯hμ(𝐮,𝐯)KeKμ(𝐮𝐧,𝐯)e=eKμ(𝐮𝐧,𝐯𝐯)e.\displaystyle=\sum_{K\in\mathcal{T}_{h}}\mu(\nabla\mathbf{u},\nabla\mathbf{v}^{\circ})_{K}-\sum_{e\in\partial K}\mu(\nabla\mathbf{u}\cdot\mathbf{n},\mathbf{v}^{\circ})_{e}=\sum_{e\in\partial K}\mu(\nabla\mathbf{u}\cdot\mathbf{n},\mathbf{v}^{\partial}-\mathbf{v}^{\circ})_{e}. (36)

For the last term on the right-hand side of (35), using the divergence theorem, the definitions of the lifting and discrete divergence operators, and the fact w=0\nabla w=0 for wP0(K)w\in P_{0}(K), we have

(Λh(𝐯),w)K\displaystyle(\nabla\cdot\Lambda_{h}(\mathbf{v}),w)_{K} =eK(Λh(𝐯)𝐧,w)e(Λh(𝐯),w)K\displaystyle=\sum_{e\in\partial K}(\Lambda_{h}(\mathbf{v})\cdot\mathbf{n},w)_{e}-(\Lambda_{h}(\mathbf{v}),\nabla w)_{K}
=eK(Λh(𝐯)𝐧,w)e=eK(𝐯𝐧,w)e\displaystyle=\sum_{e\in\partial K}(\Lambda_{h}(\mathbf{v})\cdot\mathbf{n},w)_{e}=\sum_{e\in\partial K}(\mathbf{v}\cdot\mathbf{n},w)_{e}
=(w𝐯,w)K+(𝐯,w)K=(w𝐯,w)K,\displaystyle=(\nabla_{w}\cdot\mathbf{v},w)_{K}+(\mathbf{v},\nabla w)_{K}=(\nabla_{w}\cdot\mathbf{v},w)_{K},

or

(Λh(𝐯),w)K=(w𝐯,w)K,wP0(K),𝐯𝐕h.\displaystyle(\nabla\cdot\Lambda_{h}(\mathbf{v}),w)_{K}=(\nabla_{w}\cdot\mathbf{v},w)_{K},\quad\forall w\in P_{0}(K),\;\forall\mathbf{v}\in\mathbf{V}_{h}. (37)

From this and the divergence theorem, definition of projection QhQ_{h}, and normal continuity of 𝚲h𝐯\mathbf{\Lambda}_{h}\mathbf{v}, we obtain

K𝒯h(p,𝚲h𝐯)K\displaystyle\sum_{K\in\mathcal{T}_{h}}(\nabla p,\mathbf{\Lambda}_{h}\mathbf{v})_{K} =eK(p,𝚲h𝐯𝐧)eK𝒯h(p,𝚲h𝐯)K\displaystyle=\sum_{e\in\partial K}(p,\mathbf{\Lambda}_{h}\mathbf{v}\cdot\mathbf{n})_{e}-\sum_{K\in\mathcal{T}_{h}}(p,\nabla\cdot\mathbf{\Lambda}_{h}\mathbf{v})_{K}
=eK(p,𝚲h𝐯𝐧)eK𝒯h(Qhp,𝚲h𝐯)K\displaystyle=\sum_{e\in\partial K}(p,\mathbf{\Lambda}_{h}\mathbf{v}\cdot\mathbf{n})_{e}-\sum_{K\in\mathcal{T}_{h}}(Q_{h}p,\nabla\cdot\mathbf{\Lambda}_{h}\mathbf{v})_{K}
=eK(p,𝚲h𝐯𝐧)eK𝒯h(Qhp,w𝐯)K\displaystyle=\sum_{e\in\partial K}(p,\mathbf{\Lambda}_{h}\mathbf{v}\cdot\mathbf{n})_{e}-\sum_{K\in\mathcal{T}_{h}}(Q_{h}p,\nabla_{w}\cdot\mathbf{v})_{K}
=K𝒯h(Qhp,w𝐯)K.\displaystyle=-\sum_{K\in\mathcal{T}_{h}}(Q_{h}p,\nabla_{w}\cdot\mathbf{v})_{K}. (38)

Inserting (36) and (38) into (35), we obtain

K𝒯h(𝐟,𝚲h𝐯)K\displaystyle\sum_{K\in\mathcal{T}_{h}}(\mathbf{f},\mathbf{\Lambda}_{h}\mathbf{v})_{K} =K𝒯hμ(Δ𝐮,𝚲h𝐯𝐯)K+eKμ(𝐮𝐧,𝐯𝐯)eK𝒯h(Qhp,w𝐯)K.\displaystyle=-\sum_{K\in\mathcal{T}_{h}}\mu(\Delta\mathbf{u},\mathbf{\Lambda}_{h}\mathbf{v}-\mathbf{v}^{\circ})_{K}+\sum_{e\in\partial K}\mu(\nabla\mathbf{u}\cdot\mathbf{n},\mathbf{v}^{\partial}-\mathbf{v}^{\circ})_{e}-\sum_{K\in\mathcal{T}_{h}}(Q_{h}p,\nabla_{w}\cdot\mathbf{v})_{K}. (39)

For the second term on the right-hand side of the first equation of (32) using the WG commuting identity

(wQh𝐮,𝐰)K=(𝒬h𝐮,𝐰)K,𝐰RT0(K)d,(\nabla_{w}{Q}_{h}\mathbf{u},\mathbf{w})_{K}=(\mathcal{Q}_{h}\nabla\mathbf{u},\mathbf{w})_{K},\quad\forall\mathbf{w}\in RT_{0}(K)^{d},

the definition of the discrete weak gradient operator, the divergence theorem, and the fact 𝐯=0\nabla\mathbf{v}^{\circ}=0, we have

μK𝒯h(wQh𝐮,w𝐯)K\displaystyle\mu\sum_{K\in\mathcal{T}_{h}}(\nabla_{w}{Q}_{h}\mathbf{u},\nabla_{w}\mathbf{v})_{K} =μK𝒯h(𝒬h𝐮,w𝐯)K\displaystyle=\mu\sum_{K\in\mathcal{T}_{h}}(\mathcal{Q}_{h}\nabla\mathbf{u},\nabla_{w}\mathbf{v})_{K}
=eKμ(𝒬h𝐮𝐧,𝐯)eK𝒯hμ((𝒬h𝐮),𝐯)K\displaystyle=\sum_{e\in\partial K}\mu(\mathcal{Q}_{h}\nabla\mathbf{u}\cdot\mathbf{n},\mathbf{v}^{\partial})_{e}-\sum_{K\in\mathcal{T}_{h}}\mu(\nabla\cdot(\mathcal{Q}_{h}\nabla\mathbf{u}),\mathbf{v}^{\circ})_{K}
=eKμ(𝒬h𝐮𝐧,𝐯)e+K𝒯hμ(𝒬h𝐮,𝐯)K\displaystyle=\sum_{e\in\partial K}\mu(\mathcal{Q}_{h}\nabla\mathbf{u}\cdot\mathbf{n},\mathbf{v}^{\partial})_{e}+\sum_{K\in\mathcal{T}_{h}}\mu(\mathcal{Q}_{h}\nabla\mathbf{u},\nabla\mathbf{v}^{\circ})_{K}
eK(𝒬h𝐮𝐧,𝐯)e\displaystyle\qquad-\sum_{e\in\partial K}(\mathcal{Q}_{h}\nabla\mathbf{u}\cdot\mathbf{n},\mathbf{v}^{\circ})_{e}
=eKμ(𝒬h𝐮𝐧,𝐯𝐯)e.\displaystyle=\sum_{e\in\partial K}\mu(\mathcal{Q}_{h}\nabla\mathbf{u}\cdot\mathbf{n},\mathbf{v}^{\partial}-\mathbf{v}^{\circ})_{e}. (40)

Substituting (39) and (40) into (34), we obtain (32).

Now we estimate the second term on the right-hand side of the second equation of (34). Using the fact 𝐮=0\nabla\cdot\mathbf{u}=0 and the commuting identity of WG

(w(Qh𝐮),w)K=(Qh(𝐮),w)K,wP0(K),(\nabla_{w}\cdot({Q}_{h}\mathbf{u}),w)_{K}=({Q}_{h}(\nabla\cdot\mathbf{u}),w)_{K},\quad\forall w\in P_{0}(K),

we have

(w(Qh𝐮),q)K=(Qh(𝐮),q)K=(𝐮,q)K=0.\displaystyle(\nabla_{w}\cdot(Q_{h}\mathbf{u}),q)_{K}=({Q}_{h}(\nabla\cdot\mathbf{u}),q)_{K}=(\nabla\cdot\mathbf{u},q)_{K}=0.

Inserting this into (34) gives the second error equation in (32). ∎

      Lemma 3.4.

Let 𝐮H2(Ω)d\mathbf{u}\in H^{2}(\Omega)^{d} and pH1(Ω)p\in H^{1}(\Omega) be the solutions of the Stokes problem and 𝐮h𝐕h\mathbf{u}_{h}\in\mathbf{V}_{h} and phWhp_{h}\in W_{h} be the numerical solutions of the modified scheme (25). Then,

|Qh𝐮𝐮h|C(h𝐮2+|αh|),\displaystyle|||Q_{h}\mathbf{u}-\mathbf{u}_{h}|||\leq C\Big{(}h\|\mathbf{u}\|_{2}+|\alpha_{h}|\Big{)}, (41)
QhpphCμ(h𝐮2+|αh|),\displaystyle\|Q_{h}p-p_{h}\|\leq C\mu\Big{(}h\|\mathbf{u}\|_{2}+|\alpha_{h}|\Big{)}, (42)

where C>0C>0 is a constant independent of hh and μ\mu.

Proof.

By the Cauchy-Schwarz inequality and trace inequalities ((27), (29), and (30) in Lemma 3.1), we have

|1(𝐮,𝐞h)|\displaystyle|\mathcal{R}_{1}(\mathbf{u},\mathbf{e}_{h})| =|K𝒯heK(𝐞h𝐞h,(𝒬h𝐮𝐮)𝐧e)e|\displaystyle=|\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}(\mathbf{e}_{h}^{\partial}-\mathbf{e}_{h}^{\circ},\,(\mathcal{Q}_{h}\nabla\mathbf{u}-\nabla\mathbf{u})\mathbf{n}_{e})_{e}|
(K𝒯heKh1𝐞h𝐞he2)1/2(K𝒯heKh(𝒬h𝐮𝐮)𝐧ee2)1/2\displaystyle\leq(\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}h^{-1}\|\mathbf{e}_{h}^{\partial}-\mathbf{e}_{h}^{\circ}\|_{e}^{2})^{1/2}(\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}h\|(\mathcal{Q}_{h}\nabla\mathbf{u}-\nabla\mathbf{u})\mathbf{n}_{e}\|_{e}^{2})^{1/2}
Ch𝐮2|𝐞h|,\displaystyle\leq Ch\|\mathbf{u}\|_{2}|||\mathbf{e}_{h}|||, (43)

and

|2(𝐮,𝐞h)|\displaystyle|\mathcal{R}_{2}(\mathbf{u},\mathbf{e}_{h})| =|K𝒯h(Δ𝐮,𝚲h𝐞h𝐞h)|C𝐮2(K𝒯h𝚲h𝐞h𝐞h2)1/2\displaystyle=|\sum_{K\in\mathcal{T}_{h}}(\Delta\mathbf{u},\mathbf{\Lambda}_{h}\mathbf{e}_{h}-\mathbf{e}_{h}^{\circ})|\leq C\|\mathbf{u}\|_{2}(\sum_{K\in\mathcal{T}_{h}}\|\mathbf{\Lambda}_{h}\mathbf{e}_{h}-\mathbf{e}_{h}^{\circ}\|^{2})^{1/2}
Ch𝐮2|𝐞h|.\displaystyle\leq Ch\|\mathbf{u}\|_{2}|||\mathbf{e}_{h}|||. (44)

Taking 𝐯=𝐞h\mathbf{v}=\mathbf{e}_{h} and 𝐪=𝐞hp\mathbf{q}=\mathbf{e}_{h}^{p} in the error equations (32), we have

μ|𝐞h|2\displaystyle\displaystyle\mu|||\mathbf{e}_{h}|||^{2} =𝐞hTA𝐞h=μ1(𝐮,𝐞h)+μ2(𝐮,𝐞h)+𝐞hT(B)T𝐞hp\displaystyle=\mathbf{e}_{h}^{T}A\mathbf{e}_{h}=\mu\mathcal{R}_{1}(\mathbf{u},\mathbf{e}_{h})+\mu\mathcal{R}_{2}(\mathbf{u},\mathbf{e}_{h})+\mathbf{e}_{h}^{T}(B^{\circ})^{T}\mathbf{e}_{h}^{p}
Chμ𝐮2|||𝐞h|||+|αh|NK𝒯h|ehp|K|\displaystyle\leq Ch\mu\|\mathbf{{u}}\|_{2}|||\mathbf{e}_{h}|||+\frac{|\alpha_{h}|}{N}\sum_{K\in\mathcal{T}_{h}}|e_{h}^{p}|_{K}|
Chμ𝐮2|||𝐞h|||+C|αh|K𝒯h|K||ehp|K|\displaystyle\leq Ch\mu\|\mathbf{{u}}\|_{2}|||\mathbf{e}_{h}|||+C|\alpha_{h}|\sum_{K\in\mathcal{T}_{h}}|K|\,|e_{h}^{p}|_{K}|
Chμ𝐮2|𝐞h|+C|αh|ehp,\displaystyle\leq Ch\mu\|\mathbf{{u}}\|_{2}|||\mathbf{e}_{h}|||+C|\alpha_{h}|\,\|e_{h}^{p}\|, (45)

where we have used the assumption that the mesh is quasi-uniform. From the inf-sup condition Lemma 3.2, we get

β𝐞hp\displaystyle\beta\|\mathbf{e}_{h}^{p}\| sup𝐯𝐕h,|𝐯|0𝐯T(B)T𝐞hp|𝐯|\displaystyle\leq\sup_{\mathbf{v}\in\mathbf{V}_{h}^{\circ},\;|||\mathbf{v}|||\neq 0}\frac{\mathbf{v}^{T}(B^{\circ})^{T}\mathbf{e}_{h}^{p}}{|||\mathbf{v}|||}
=sup𝐯𝐕h,|𝐯|0𝐯TA𝐞hμ1(𝐮,𝐯)μ2(𝐮,𝐯)|𝐯|\displaystyle=\sup_{\mathbf{v}\in\mathbf{V}_{h}^{\circ},\;|||\mathbf{v}|||\neq 0}\frac{\mathbf{v}^{T}A\mathbf{e}_{h}-\mu\mathcal{R}_{1}(\mathbf{u},\mathbf{v})-\mu\mathcal{R}_{2}(\mathbf{u},\mathbf{v})}{|||\mathbf{v}|||}
=μ|𝐞h|+Chμ𝐮2.\displaystyle=\mu|||\mathbf{e}_{h}|||+Ch\mu\|\mathbf{u}\|_{2}. (46)

Combining (45) and (46), we have

μ|𝐞h|2\displaystyle\mu|||\mathbf{e}_{h}|||^{2} Cμ|𝐞h|(h𝐮2+|αh|)+C|αh|h𝐮2.\displaystyle\leq C\mu|||\mathbf{e}_{h}|||(h\|\mathbf{u}\|_{2}+|\alpha_{h}|)+C|\alpha_{h}|h\|\mathbf{u}\|_{2}.

Applying the quadratic formula to the above inequality we obtain

|𝐞h|C(h𝐮2+|αh|)+Ch|αh|𝐮2C(h𝐮2+|αh|),\displaystyle|||\mathbf{e}_{h}|||\leq C(h\|\mathbf{u}\|_{2}+|\alpha_{h}|)+C\sqrt{h|\alpha_{h}|\,\|\mathbf{u}\|_{2}}\leq C(h\|\mathbf{u}\|_{2}+|\alpha_{h}|),

which gives (41). The inequality (42) follows from (41) and (46). ∎

      Theorem 3.1.

Let 𝐮H2(Ω)d\mathbf{u}\in H^{2}(\Omega)^{d} and pH1(Ω)p\in H^{1}(\Omega) be the solutions of Stokes problem and 𝐮h𝐕h\mathbf{u}_{h}\in\mathbf{V}_{h} and phWhp_{h}\in W_{h} be the numerical solutions of the modified scheme (25). Assume 𝐟L2(Ω)d\mathbf{f}\in L^{2}(\Omega)^{d}. Then,

pphCh𝐟+Cμ|αh|,\displaystyle\|p-p_{h}\|\leq Ch\|\mathbf{f}\|+C\mu|\alpha_{h}|, (47)
𝐮w𝐮hCh𝐮2+C|αh|,\displaystyle\|\nabla\mathbf{u}-\nabla_{w}\mathbf{u}_{h}\|\leq Ch\|\mathbf{u}\|_{2}+C|\alpha_{h}|, (48)
𝐮𝐮h=𝐮𝐮hCh𝐮2+C|αh|,\displaystyle\|\mathbf{u}-\mathbf{u}_{h}\|=\|\mathbf{u}-\mathbf{u}_{h}^{\circ}\|\leq Ch\|\mathbf{u}\|_{2}+C|\alpha_{h}|, (49)
Qh𝐮𝐮hCh2𝐮2+C|αh|,\displaystyle\displaystyle\|Q^{\circ}_{h}\mathbf{u}-\mathbf{u}^{\circ}_{h}\|\leq Ch^{2}\|\mathbf{u}\|_{2}+C|\alpha_{h}|, (50)

where CC is a constant independent of hh and μ\mu.

Proof.

It is known that the solutions of the Stokes problem have the regularity

μ𝐮2+p1C𝐟.\mu\|\mathbf{u}\|_{2}+\|p\|_{1}\leq C\|\mathbf{f}\|.

Using this, Lemma 3.4, and the triangle inequality, we obtain (47).

Let 𝒬h\mathcal{Q}_{h} be a L2L^{2}-projection operator from L2(Ω)d×dL^{2}(\Omega)^{d\times d} onto broken RT0(𝒯h)dRT_{0}(\mathcal{T}_{h})^{d} space. From the commuting identity of WG, Lemma 3.4, and projection properties, we have

w𝐮h𝐮\displaystyle\|\nabla_{w}\mathbf{u}_{h}-\nabla\mathbf{u}\| w𝐮h𝒬h𝐮+𝒬h𝐮𝐮C(h𝐮2+|αh|),\displaystyle\leq\|\nabla_{w}\mathbf{u}_{h}-\mathcal{Q}_{h}\nabla\mathbf{u}\|+\|\mathcal{Q}_{h}\nabla\mathbf{u}-\nabla\mathbf{u}\|\leq C(h\|\mathbf{u}\|_{2}+|\alpha_{h}|),

which gives (48).

Consider the dual problem

{μΔ𝚿+ψ=𝐞h,inΩ,𝚿=0,inΩ,𝚿=𝟎,onΩ,\begin{cases}\displaystyle-\mu\Delta\mathbf{\Psi}+\nabla\psi=\mathbf{e}_{h}^{\circ},\quad\mbox{in}\;\;\Omega,\\[3.61371pt] \displaystyle\nabla\cdot\mathbf{\Psi}=0,\quad\mbox{in}\;\;\Omega,\\[3.61371pt] \displaystyle\mathbf{\Psi}=\mathbf{0},\quad\mbox{on}\;\;\partial\Omega,\end{cases} (51)

where 𝐞h=Qh𝐮𝐮h\mathbf{e}_{h}^{\circ}=Q^{\circ}_{h}\mathbf{u}-\mathbf{u}^{\circ}_{h} is the velocity discrete error. It is known 𝚿H2(Ω)d\mathbf{\Psi}\in H^{2}(\Omega)^{d} and ψH1(Ω)\psi\in H^{1}(\Omega). Taking 𝐞h\mathbf{e}_{h}^{\circ} as the test function in the first equation of the dual problem, we have

𝐞h2=μK𝒯h(Δ𝚿,𝐞h)K+K𝒯h(ψ,𝐞h)K.\displaystyle\|\mathbf{e}_{h}^{\circ}\|^{2}=-\mu\sum_{K\in\mathcal{T}_{h}}(\Delta\mathbf{\Psi},\,\mathbf{e}_{h}^{\circ})_{K}+\sum_{{K\in\mathcal{T}_{h}}}(\nabla\psi,\,\mathbf{e}_{h}^{\circ})_{K}. (52)

We estimate the terms on the right-hand side separately.

For the second term in (52), using the divergence theorem and the definition of QhQ_{h} and discrete weak divergence, we have

(ψ,𝐞h)K\displaystyle\displaystyle(\nabla\psi,\mathbf{e}_{h}^{\circ})_{K} =(ψ𝐧,𝐞h)e(ψ,𝐞h)K\displaystyle=(\psi\mathbf{n},\,\mathbf{e}_{h}^{\circ})_{e}-(\psi,\nabla\cdot\mathbf{e}_{h}^{\circ})_{K}
=(ψ𝐧,𝐞h)e(Qhψ,𝐞h)K\displaystyle=(\psi\mathbf{n},\,\mathbf{e}_{h}^{\circ})_{e}-(Q_{h}\psi,\nabla\cdot\mathbf{e}_{h}^{\circ})_{K}
=(ψ𝐧,𝐞h)e((Qhψ)𝐧,𝐞h)e+((Qhψ),𝐞h)K\displaystyle=(\psi\mathbf{n},\mathbf{e}_{h}^{\circ})_{e}-((Q_{h}\psi)\mathbf{n},\mathbf{e}_{h}^{\circ})_{e}+(\nabla(Q_{h}\psi),\mathbf{e}_{h}^{\circ})_{K}
=(ψ𝐧,𝐞h)e((Qhψ)𝐧,𝐞h)e+((Qhψ)𝐧,𝐞h)e(Qhψ,w𝐞h)K.\displaystyle=(\psi\mathbf{n},\mathbf{e}_{h}^{\circ})_{e}-((Q_{h}\psi)\mathbf{n},\mathbf{e}_{h}^{\circ})_{e}+((Q_{h}\psi)\mathbf{n},\mathbf{e}_{h}^{\partial})_{e}-(Q_{h}\psi,\nabla_{w}\cdot\mathbf{e}_{h})_{K}.

Summing over 𝒯h\mathcal{T}_{h} and using K𝒯heK(ψ𝐧,𝐞h)e=0\displaystyle\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}(\psi\mathbf{n},\,\mathbf{e}_{h}^{\partial})_{e}=0, we get

K𝒯h(ψ,𝐞h)K=K𝒯heK((Qhψψ)𝐧,𝐞h𝐞h)eK𝒯h(Qhψ,w𝐞h)K.\displaystyle\displaystyle\sum_{K\in\mathcal{T}_{h}}(\nabla\psi,\mathbf{e}_{h}^{\circ})_{K}=\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}((Q_{h}\psi-\psi)\mathbf{n},\,\mathbf{e}_{h}^{\partial}-\mathbf{e}_{h}^{\circ})_{e}-\sum_{K\in\mathcal{T}_{h}}(Q_{h}\psi,\nabla_{w}\cdot\mathbf{e}_{h})_{K}.

From the Cauchy-Schwarz inequality, (27) and (29) of Lemma 3.1, the projection properties, Lemma 3.4, and

K𝒯h(Qhψ,w𝐞h)K=B(𝐞h,Qhψ)=αhNK𝒯hQhψ|KCψ1|αh|,\sum_{K\in\mathcal{T}_{h}}(Q_{h}\psi,\nabla_{w}\cdot\mathbf{e}_{h})_{K}=B^{\circ}(\mathbf{e}_{h},Q_{h}\psi)=\frac{\alpha_{h}}{N}\sum_{K\in\mathcal{T}_{h}}Q_{h}\psi|_{K}\leq C\|\psi\|_{1}|\alpha_{h}|,

we have

K𝒯h(ψ,𝐞h)K\displaystyle\displaystyle\sum_{K\in\mathcal{T}_{h}}(\nabla\psi,\mathbf{e}_{h}^{\circ})_{K} =K𝒯heK(Qhψψ)𝐧,𝐞h𝐞heK𝒯h(Qhψ,w𝐞h)K\displaystyle=\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}\langle(Q_{h}\psi-\psi)\mathbf{n},\,\mathbf{e}_{h}^{\partial}-\mathbf{e}_{h}^{\circ}\rangle_{e}-\sum_{K\in\mathcal{T}_{h}}(Q_{h}\psi,\nabla_{w}\cdot\mathbf{e}_{h})_{K}
(K𝒯heKhQhψψe2)12(K𝒯heKh1𝐞h𝐞he2)12\displaystyle\leq\left(\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}h\|Q_{h}\psi-\psi\|_{e}^{2}\right)^{\frac{1}{2}}\left(\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}h^{-1}\|\mathbf{e}_{h}^{\partial}-\mathbf{e}_{h}^{\circ}\|_{e}^{2}\right)^{\frac{1}{2}}
+Cψ1|αh|\displaystyle+C\|\psi\|_{1}|\alpha_{h}|
(K𝒯h(QhψψK2+h2(Qhψψ)K2))12|𝐞h|+Cψ1|αh|\displaystyle\leq\left(\sum_{K\in\mathcal{T}_{h}}(\|Q_{h}\psi-\psi\|^{2}_{K}+h^{2}\|\nabla(Q_{h}\psi-\psi)\|_{K}^{2})\right)^{\frac{1}{2}}|||\mathbf{e}_{h}|||+C\|\psi\|_{1}|\alpha_{h}|
Chψ1|𝐞h|+Cψ1|αh|\displaystyle\displaystyle\leq Ch\|\psi\|_{1}|||\mathbf{e}_{h}|||+C\|\psi\|_{1}|\alpha_{h}|
Cψ1(h2𝐮2+|αh|).\displaystyle\leq C\|\psi\|_{1}(h^{2}\|\mathbf{u}\|_{2}+|\alpha_{h}|). (53)

Next, we focus on the first term of (52), μK𝒯h(Δ𝚿,𝐞h)K-\mu\sum_{K\in\mathcal{T}_{h}}(\Delta\mathbf{\Psi},\,\mathbf{e}_{h}^{\circ})_{K}. By the divergence theorem and K𝒯heK(𝚿𝐧,𝐞h)e=0\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}(\nabla\mathbf{\Psi}\cdot\mathbf{n},\mathbf{e}_{h}^{\partial})_{e}=0 (from the continuity of 𝚿\mathbf{\Psi}), there holds

(Δ𝚿,𝐞h)\displaystyle-(\Delta\mathbf{\Psi},\mathbf{e}_{h}^{\circ}) =(𝚿,𝐞h)K𝒯heK(𝚿𝐧,𝐞h)e\displaystyle=(\nabla\mathbf{\Psi},\nabla\mathbf{e}_{h}^{\circ})-\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}(\nabla\mathbf{\Psi}\cdot\mathbf{n},\mathbf{e}_{h}^{\circ})_{e}
=(𝚿,𝐞h)K𝒯heK(𝚿𝐧,𝐞h𝐞h)e.\displaystyle=(\nabla\mathbf{\Psi},\nabla\mathbf{e}_{h}^{\circ})-\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}(\nabla\mathbf{\Psi}\cdot\mathbf{n},\mathbf{e}_{h}^{\circ}-\mathbf{e}_{h}^{\partial})_{e}. (54)

For the first term in (54), by the property of 𝒬h\mathcal{Q}_{h}, divergence theorem, definition of (w)(\nabla_{w}), and property of QhQ_{h}, we have

(𝚿,𝐞h)\displaystyle(\nabla\mathbf{\Psi},\nabla\mathbf{e}_{h}^{\circ}) =(𝒬h(𝚿),𝐞h)\displaystyle=(\mathcal{Q}_{h}(\nabla\mathbf{\Psi}),\nabla\mathbf{e}_{h}^{\circ})
=(𝐞h,((𝒬h(𝚿)))+K𝒯heK(𝐞h,(𝒬h(𝚿))𝐧)e\displaystyle=-(\mathbf{e}_{h}^{\circ},\nabla\cdot((\mathcal{Q}_{h}(\nabla\mathbf{\Psi})))+\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}(\mathbf{e}_{h}^{\circ},(\mathcal{Q}_{h}(\nabla\mathbf{\Psi}))\cdot\mathbf{n})_{e}
=(𝒬h(𝚿),w𝐞h)+K𝒯heK(𝒬h(𝚿)𝐧,𝐞h𝐞h)e\displaystyle=(\mathcal{Q}_{h}(\nabla\mathbf{\Psi}),\nabla_{w}\mathbf{e}_{h})+\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}(\mathcal{Q}_{h}(\nabla\mathbf{\Psi})\cdot\mathbf{n},\mathbf{e}_{h}^{\circ}-\mathbf{e}_{h}^{\partial})_{e}
=(wQh𝚿,w𝐞h)+K𝒯heK(𝒬h(𝚿)𝐧,𝐞h𝐞h)e.\displaystyle=(\nabla_{w}Q_{h}\mathbf{\Psi},\nabla_{w}\mathbf{e}_{h})+\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}(\mathcal{Q}_{h}(\nabla\mathbf{\Psi})\cdot\mathbf{n},\mathbf{e}_{h}^{\circ}-\mathbf{e}_{h}^{\partial})_{e}. (55)

Plugging (55) into (54), we get

μ(Δ𝚿,𝐞h)\displaystyle-\mu(\Delta\mathbf{\Psi},\mathbf{e}_{h}^{\circ}) =μ(wQh𝚿,w𝐞h)\displaystyle=\mu(\nabla_{w}Q_{h}\mathbf{\Psi},\nabla_{w}\mathbf{e}_{h})
K𝒯heKμ((𝚿𝒬h(𝚿))𝐧,𝐞h𝐞h)e.\displaystyle\quad-\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}\mu((\nabla\mathbf{\Psi}-\mathcal{Q}_{h}(\nabla\mathbf{\Psi}))\cdot\mathbf{n},\mathbf{e}_{h}^{\circ}-\mathbf{e}_{h}^{\partial})_{e}. (56)

Combining (52), (53), and (56) gives

𝐞h2\displaystyle\|\mathbf{e}_{h}^{\circ}\|^{2} Cψ1(h2𝐮2+|αh|)+μ(wQh𝚿,w𝐞h)\displaystyle\leq C\|\psi\|_{1}(h^{2}\|\mathbf{u}\|_{2}+|\alpha_{h}|)+\mu(\nabla_{w}Q_{h}\mathbf{\Psi},\nabla_{w}\mathbf{e}_{h})
K𝒯heKμ((𝚿𝒬h(𝚿))𝐧,𝐞h𝐞h)e.\displaystyle\quad-\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}\mu((\nabla\mathbf{\Psi}-\mathcal{Q}_{h}(\nabla\mathbf{\Psi}))\cdot\mathbf{n},\mathbf{e}_{h}^{\circ}-\mathbf{e}_{h}^{\partial})_{e}. (57)

Taking 𝐯=Qh𝚿\mathbf{v}=Q_{h}\mathbf{\Psi} in the first equation of error equation (32) yields

μ(w𝐞h,wQh𝚿)=μ1(𝐮,Qh𝚿)+μ2(𝐮,Qh𝚿)+(ehp,wQh𝚿).\displaystyle\mu(\nabla_{w}\mathbf{e}_{h},\nabla_{w}Q_{h}\mathbf{\Psi})=\mu\mathcal{R}_{1}(\mathbf{u},Q_{h}\mathbf{\Psi})+\mu\mathcal{R}_{2}(\mathbf{u},Q_{h}\mathbf{\Psi})+(e_{h}^{p},\nabla_{w}\cdot Q_{h}\mathbf{\Psi}).

By the property of projection and 𝚿=0\nabla\cdot\mathbf{\Psi}=0, we have

(ehp,wQh𝚿)=(ehp,Qh(𝚿))=(ehp,𝚿)=0,\displaystyle(e_{h}^{p},\nabla_{w}\cdot Q_{h}\mathbf{\Psi})=(e_{h}^{p},Q_{h}(\nabla\cdot\mathbf{\Psi}))=(e_{h}^{p},\nabla\cdot\mathbf{\Psi})=0,

which gives

μ(w𝐞h,wQh𝚿)=μ1(𝐮,Qh𝚿)+μ2(𝐮,Qh𝚿).\displaystyle\mu(\nabla_{w}\mathbf{e}_{h},\nabla_{w}Q_{h}\mathbf{\Psi})=\mu\mathcal{R}_{1}(\mathbf{u},Q_{h}\mathbf{\Psi})+\mu\mathcal{R}_{2}(\mathbf{u},Q_{h}\mathbf{\Psi}).

Using this, from (3) we obtain

𝐞h2\displaystyle\|\mathbf{e}_{h}^{\circ}\|^{2} Cψ1(h2𝐮2+|αh|)+μ1(𝐮,Qh𝚿)+μ2(𝐮,Qh𝚿)\displaystyle\leq C\|\psi\|_{1}(h^{2}\|\mathbf{u}\|_{2}+|\alpha_{h}|)+\mu\mathcal{R}_{1}(\mathbf{u},Q_{h}\mathbf{\Psi})+\mu\mathcal{R}_{2}(\mathbf{u},Q_{h}\mathbf{\Psi})
K𝒯heKμ((𝚿𝒬h(𝚿))𝐧,𝐞h𝐞h)e.\displaystyle\quad-\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}\mu((\nabla\mathbf{\Psi}-\mathcal{Q}_{h}(\nabla\mathbf{\Psi}))\cdot\mathbf{n},\mathbf{e}_{h}^{\circ}-\mathbf{e}_{h}^{\partial})_{e}. (58)

Next, we estimate the last three terms in the above inequality.

By the trace inequalities (27) and (29) and Lemma 3.4, the third term is bounded as

K𝒯heKμ((𝚿Qh𝚿)𝐧,𝐞h𝐞h)e\displaystyle\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}\mu((\nabla\mathbf{\Psi}-Q_{h}\mathbf{\Psi})\cdot\mathbf{n},\mathbf{e}_{h}^{\circ}-\mathbf{e}_{h}^{\partial})_{e}
μ(K𝒯heKh𝚿𝒬h(𝚿)e2)1/2(K𝒯heKh1𝐞h𝐞he2)1/2\displaystyle\leq\mu(\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}h\|\nabla\mathbf{\Psi}-\mathcal{Q}_{h}(\nabla\mathbf{\Psi})\|^{2}_{e})^{1/2}(\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}h^{-1}\|\mathbf{e}_{h}^{\circ}-\mathbf{e}_{h}^{\partial}\|^{2}_{e})^{1/2}
Cμh𝚿2|𝐞h|\displaystyle\leq C\mu h\|\mathbf{\Psi}\|_{2}|||\mathbf{e}_{h}|||
Cμh2𝚿2𝐮2+Cμ𝚿2|αh|.\displaystyle\leq C\mu h^{2}\|\mathbf{\Psi}\|_{2}\|\mathbf{u}\|_{2}+C\mu\|\mathbf{\Psi}\|_{2}|\alpha_{h}|. (59)

By the Cauchy-Schwarz inequality and the trace inequality (27), the second term is bounded as

μ1(𝐮,Qh𝚿)\displaystyle\mu\mathcal{R}_{1}(\mathbf{u},Q_{h}\mathbf{\mathbf{\Psi}}) =μK𝒯heK(Qh𝚿Qh𝚿,(𝒬h𝐮𝐮)𝐧e)e\displaystyle=\mu\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}(Q_{h}^{\partial}\mathbf{\Psi}-Q_{h}^{\circ}\mathbf{\Psi},\,(\mathcal{Q}_{h}\nabla\mathbf{u}-\nabla\mathbf{u})\mathbf{n}_{e})_{e}
μ(K𝒯heKQh𝚿𝚿e2)1/2(K𝒯heK𝒬h𝐮𝐮e2)1/2\displaystyle\leq\mu(\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}\|Q_{h}^{\circ}\mathbf{\Psi}-\mathbf{\Psi}\|_{e}^{2})^{1/2}(\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}\|\mathcal{Q}_{h}\nabla\mathbf{u}-\nabla\mathbf{u}\|_{e}^{2})^{1/2}
Cμh2𝐮2𝚿2.\displaystyle\leq C\mu h^{2}\|\mathbf{u}\|_{2}\|\mathbf{\Psi}\|_{2}. (60)

By the trace inequalities (27) and (30), for the third term we have

μ2(𝐮,Qh𝚿)\displaystyle\mu\mathcal{R}_{2}(\mathbf{u},Q_{h}\mathbf{\mathbf{\Psi}}) =K𝒯hμ(Δ𝐮,𝚲hQh𝚿Qh𝚿)\displaystyle=\sum_{K\in\mathcal{T}_{h}}\mu(\Delta\mathbf{u},\mathbf{\Lambda}_{h}Q_{h}\mathbf{\Psi}-Q_{h}^{\circ}\mathbf{\Psi})
=K𝒯hμ(Δ𝐮,𝚲hQh𝚿Qh𝚿)\displaystyle=\sum_{K\in\mathcal{T}_{h}}\mu(\Delta\mathbf{u},\mathbf{\Lambda}_{h}Q_{h}\mathbf{\Psi}-Q_{h}^{\circ}\mathbf{\Psi})
Cμ𝐮2(K𝒯heKhQh𝚿Qh𝚿e2)1/2\displaystyle\leq C\mu\|\mathbf{u}\|_{2}(\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}h\|Q_{h}^{\partial}\mathbf{\Psi}-Q_{h}^{\circ}\mathbf{\Psi}\|_{e}^{2})^{1/2}
Cμ𝐮2(K𝒯heKhQh𝚿𝚿e2)1/2\displaystyle\leq C\mu\|\mathbf{u}\|_{2}(\sum_{K\in\mathcal{T}_{h}}\sum_{e\in\partial K}h\|Q_{h}^{\circ}\mathbf{\Psi}-\mathbf{\Psi}\|_{e}^{2})^{1/2}
Cμh2𝐮2𝚿2.\displaystyle\leq C\mu h^{2}\|\mathbf{u}\|_{2}\|\mathbf{\Psi}\|_{2}. (61)

Combining (58), (59), (60), and (61), and using the regularity

μ𝚿2+ψ1C𝐞h,\mu\|\mathbf{\Psi}\|_{2}+\|\psi\|_{1}\leq C\|\mathbf{e}_{h}^{\circ}\|,

we obtain (50). Inequality (49) follows from (50), the triangle inequality, and approximation properties. ∎

Comparing the above theorem with Lemma 2.1, we can see that the effects of the consistency enforcement (23) are reflected by the factor |αh||\alpha_{h}| in the error bounds (47)–(50). In particular, if αh=0\alpha_{h}=0, the bounds in Theorem 3.1 are the same as those in Lemma 2.1. Moreover, from (24) we can see that the magnitude of αh\alpha_{h} depends on how closely 𝐠e\langle\mathbf{g}\rangle_{e} is approximated by Qh𝐠|eQ_{h}^{\partial}\mathbf{g}|_{e}. Such approximation can be made in 𝒪(h2)\mathcal{O}(h^{2}), for instance, by choosing Qh𝐠|eQ_{h}^{\partial}\mathbf{g}|_{e} as the value of 𝐠\mathbf{g} at the barycenter of ee. It can also be made at higher order using a Gaussian quadrature rule for computing 𝐠e\langle\mathbf{g}\rangle_{e}. When the approximation is at second order or higher, Theorem 3.1 shows that the optimal convergence order of the WG scheme is not affected by the consistency enforcement (23).

4 Convergence analysis of MINRES/GMRES with block Schur complement preconditioning

In this section we study the MINRES/GMRES iterative solution for the modified scheme (25) (that is singular but consistent) with block diagonal/triangular Schur complement preconditioning. We establish bounds for the eigenvalues of the preconditioned systems and for the residual of MINRES/GMRES.

We start with rescaling the unknown variables and rewriting (25) into

[A(B)TB0][μ𝐮h𝐩h]=[𝐛1μ𝐛~2],𝒜=[A(B)TB0].\begin{bmatrix}A&-(B^{\circ})^{T}\\ -B^{\circ}&0\end{bmatrix}\begin{bmatrix}\mu\mathbf{u}_{h}\\ \mathbf{p}_{h}\end{bmatrix}=\begin{bmatrix}\mathbf{b}_{1}\\ \mu\tilde{\mathbf{b}}_{2}\end{bmatrix},\quad\mathcal{A}=\begin{bmatrix}A&-(B^{\circ})^{T}\\ -B^{\circ}&0\end{bmatrix}. (62)

The following lemma provide a bound for the Schur complement of the above system.

      Lemma 4.1.

The Schur complement S=BA1(B)TS=B^{\circ}A^{-1}(B^{\circ})^{T} for (62) satisfies

0SdMp,\displaystyle 0\leq S\leq d\;M_{p}^{\circ}, (63)

where MpM_{p}^{\circ} is the mass matrix of interior pressure, the sign “\leq” is in the sense of negative semi-definite.

Proof.

From the definitions of operator BB^{\circ} in (18) and the mass matrix and the fact w𝐮P0(𝒯h)\nabla_{w}\cdot\mathbf{u}\in P_{0}(\mathcal{T}_{h}), we have

𝐮T(B)T(Mp)1B𝐮=K𝒯h(w𝐮,w𝐮)K.\displaystyle\mathbf{u}^{T}(B^{\circ})^{T}(M_{p}^{\circ})^{-1}B^{\circ}\mathbf{u}=\sum_{K\in\mathcal{T}_{h}}(\nabla_{w}\cdot\mathbf{u},\nabla_{w}\cdot\mathbf{u})_{K}.

Moreover, it can be verified directly that

K𝒯h(w𝐮𝐡,w𝐮𝐡)KdK𝒯h(w𝐮h,w𝐮h)K.\displaystyle\sum_{K\in\mathcal{T}_{h}}(\nabla_{w}\cdot\mathbf{u_{h}},\nabla_{w}\cdot\mathbf{u_{h}})_{K}\leq d\sum_{K\in\mathcal{T}_{h}}(\nabla_{w}\mathbf{u}_{h},\nabla_{w}\mathbf{u}_{h})_{K}.

Then, since both AA (16) and MpM_{p}^{\circ} are symmetric and positive definite, we have

sup𝐩0𝐩TBA1(B)T𝐩𝐩TMp𝐩\displaystyle\sup_{\mathbf{p}\neq 0}\frac{\mathbf{p}^{T}B^{\circ}{A}^{-1}(B^{\circ})^{T}\mathbf{p}}{\mathbf{p}^{T}M_{p}^{\circ}\mathbf{p}} =sup𝐩0𝐩T(Mp)12BA1(B)T(Mp)12𝐩𝐩T𝐩\displaystyle=\sup_{\mathbf{p}\neq 0}\frac{\mathbf{p}^{T}(M_{p}^{\circ})^{-\frac{1}{2}}B^{\circ}{A}^{-1}(B^{\circ})^{T}(M_{p}^{\circ})^{-\frac{1}{2}}\mathbf{p}}{\mathbf{p}^{T}\mathbf{p}}
=sup𝐮0𝐮T(B)T(Mp)1B𝐮𝐮TA𝐮d,\displaystyle=\sup_{\mathbf{u}\neq 0}\frac{\mathbf{u}^{T}(B^{\circ})^{T}(M_{p}^{\circ})^{-1}B^{\circ}\mathbf{u}}{\mathbf{u}^{T}A\mathbf{u}}\leq d, (64)

which implies SdMpS\leq d\;M_{p}^{\circ}. ∎

      Lemma 4.2.

The eigenvalues of S^1S\hat{S}^{-1}S, where S^=Mp\hat{S}=M_{p}^{\circ} and S=BA1(B)TS=B^{\circ}A^{-1}(B^{\circ})^{T}, are bounded by γ1=0\gamma_{1}=0 and γi[β2,d]\gamma_{i}\in[\beta^{2},d] for i=2,,Ni=2,...,N.

Proof.

Denoting the eigenvalues of S^1S=(Mp)1BA1(B)T\hat{S}^{-1}S=(M_{p}^{\circ})^{-1}B^{\circ}A^{-1}(B^{\circ})^{T} by γ1=0<γ2γN\gamma_{1}=0<\gamma_{2}\leq\cdots\leq\gamma_{N}. From (64), we have γNd\gamma_{N}\leq d. Moreover, γ2\gamma_{2} is the smallest positive eigenvalue of S^1S\hat{S}^{-1}S, which is equal to the smallest positive eigenvalue of A1/2(B)T(Mp)1BA1/2A^{-1/2}(B^{\circ})^{T}(M_{p}^{\circ})^{-1}B^{\circ}A^{-1/2}, which is equal to the square of the inf-sup constant β\beta, i.e., γ2=β2\gamma_{2}=\beta^{2} (cf. Lemma 3.2). Thus, γ1=0\gamma_{1}=0 and γi[β2,d]\gamma_{i}\in[\beta^{2},d] for i=2,,Ni=2,...,N. ∎

4.1 Block diagonal Schur complement preconditioning

We first consider a block diagonal Schur complement preconditioner. Based on Lemma 4.1, we take S^=Mp\hat{S}=M_{p}^{\circ} as an approximation to the Schur complement SS and choose the block diagonal preconditioner as

𝒫d=[A00Mp].\displaystyle\mathcal{P}_{d}=\begin{bmatrix}A&0\\ 0&M_{p}^{\circ}\end{bmatrix}. (65)

Since 𝒫d\mathcal{P}_{d} is SPD, the preconditioned system 𝒫d1𝒜\mathcal{P}_{d}^{-1}\mathcal{A} is similar to 𝒫d12𝒜𝒫d12\mathcal{P}_{d}^{-\frac{1}{2}}\mathcal{A}\mathcal{P}_{d}^{-\frac{1}{2}} which can be expressed as

𝒫d12𝒜𝒫d12\displaystyle\mathcal{P}_{d}^{-\frac{1}{2}}\mathcal{A}\mathcal{P}_{d}^{-\frac{1}{2}} =[A1200(Mp)12]1[A(B)TB0][A1200(Mp)12]1\displaystyle=\begin{bmatrix}A^{\frac{1}{2}}&0\\ 0&(M_{p}^{\circ})^{\frac{1}{2}}\end{bmatrix}^{-1}\begin{bmatrix}A&-(B^{\circ})^{T}\\ -B^{\circ}&0\end{bmatrix}\begin{bmatrix}A^{\frac{1}{2}}&0\\ 0&(M_{p}^{\circ})^{\frac{1}{2}}\end{bmatrix}^{-1}
=[A12(B)T(Mp)12(Mp)12BA120].\displaystyle=\begin{bmatrix}\mathcal{I}&-A^{-\frac{1}{2}}(B^{\circ})^{T}(M_{p}^{\circ})^{-\frac{1}{2}}\\ -(M_{p}^{\circ})^{-\frac{1}{2}}B^{\circ}A^{-\frac{1}{2}}&0\end{bmatrix}. (66)

The symmetry of the preconditioned system indicates that MINRES can be used for its iterative solution, with convergence analyzed through spectral analysis and residual estimates.

      Lemma 4.3.

The eigenvalues of 𝒫d1𝒜\mathcal{P}_{d}^{-1}\mathcal{A} are bounded by

[11+4d2,\displaystyle\Bigg{[}\frac{1-\sqrt{1+4d}}{2},\, 11+4β22]{0}[1+1+4β22,1+1+4d2].\displaystyle\frac{1-\sqrt{1+4\beta^{2}}}{2}\Bigg{]}\cup\Bigg{\{}0\Bigg{\}}\cup\Bigg{[}\frac{1+\sqrt{1+4\beta^{2}}}{2},\,\frac{1+\sqrt{1+4d}}{2}\Bigg{]}. (67)
Proof.

The eigenvalue problem of the preconditioned system 𝒫d1𝒜\mathcal{P}_{d}^{-1}\mathcal{A} can be expressed as

[A(B)TB0][𝐮𝐩]=λ[A00Mp][𝐮𝐩]\displaystyle\begin{bmatrix}A&-(B^{\circ})^{T}\\ -B^{\circ}&0\end{bmatrix}\begin{bmatrix}\mathbf{u}\\ \mathbf{p}\end{bmatrix}=\lambda\begin{bmatrix}A&0\\ 0&M_{p}^{\circ}\end{bmatrix}\begin{bmatrix}\mathbf{u}\\ \mathbf{p}\end{bmatrix} (68)

It is clear that λ=1\lambda=1 is not an eigenvalue. By eliminating 𝐮\mathbf{u}, we obtain

λ(λ1)S^𝐩=S𝐩.\displaystyle\lambda(\lambda-1)\hat{S}\mathbf{p}=S\mathbf{p}.

Denote the eigenvalues of S^1S\hat{S}^{-1}S by γ1=0<γ2γN\gamma_{1}=0<\gamma_{2}\leq...\leq\gamma_{N}. Then, we have

λ2λγi=0, or λ=1±1+4γi2,i=2,,N.\displaystyle\lambda^{2}-\lambda-\gamma_{i}=0,\quad\text{ or }\quad\lambda=\frac{1\pm\sqrt{1+4\gamma_{i}}}{2},\qquad i=2,...,N.

Combining this with the bound of γi\gamma_{i} (cf. Lemma 4.2), we obtain the bound (67) for the eigenvalues of 𝒫d1𝒜\mathcal{P}_{d}^{-1}\mathcal{A}. ∎

      Proposition 4.1.

The residual of MINRES applied to the preconditioned system 𝒫d1/2𝒜𝒫d1/2\mathcal{P}_{d}^{-1/2}\mathcal{A}\mathcal{P}_{d}^{-1/2} is bounded by

𝐫2k+1𝐫02(dβd+β)k.\displaystyle\frac{\|\mathbf{r}_{2k+1}\|}{\|\mathbf{r}_{0}\|}\leq 2\Bigg{(}\frac{\sqrt{d}-\beta}{\sqrt{d}+\beta}\Bigg{)}^{k}. (69)
Proof.

It is known [13] that the residual of MINRES is given by

𝐫2k+1=minp2k+1p(0)=1p(𝒫d1𝒜)𝐫0minp2k+1p(0)=1p(𝒫d1𝒜)𝐫0,\|\mathbf{r}_{2k+1}\|=\min\limits_{\begin{subarray}{c}p\in\mathbb{P}_{2k+1}\\ p(0)=1\end{subarray}}\|p(\mathcal{P}_{d}^{-1}\mathcal{A})\mathbf{r}_{0}\|\leq\min\limits_{\begin{subarray}{c}p\in\mathbb{P}_{2k+1}\\ p(0)=1\end{subarray}}\|p(\mathcal{P}_{d}^{-1}\mathcal{A})\|\;\|\mathbf{r}_{0}\|,

where 2k+1\mathbb{P}_{2k+1} is the set of polynomials of degree up to 2k+12k+1. Denote the eigenvalues of 𝒫d1/2𝒜𝒫d1/2\mathcal{P}_{d}^{-1/2}\mathcal{A}\mathcal{P}_{d}^{-1/2} by λi\lambda_{i}, i=2,,2N1i=2,...,2N-1. It is known (e.g., see [4, Section 8.3.4] or [16, Chapter 10]) that the convergence of MINRES with zero initial guess is not affected by the singular eigenvalues when applied to consistent systems. Based on the bounds for the eigenvalues of 𝒫d1𝒜\mathcal{P}_{d}^{-1}\mathcal{A} (cf. Lemma 4.3) and [4, Theorem 6.13] (for convergence of MINRES), we have

𝐫2k+1𝐫0\displaystyle\frac{\|\mathbf{r}_{2k+1}\|}{\|\mathbf{r}_{0}\|} minp2kp(0)=1maxi=1,2,..,2N1|p(λi)|\displaystyle\leq\min\limits_{\begin{subarray}{c}p\in\mathbb{P}_{2k}\\ p(0)=1\end{subarray}}\max\limits_{i=1,2,..,2N-1}|p(\lambda_{i})|
minp2kp(0)=1maxλ[11+4d2,11+4β22][1+1+4β22,1+1+4d2],|p(λi)|\displaystyle\leq\min\limits_{\begin{subarray}{c}p\in\mathbb{P}_{2k}\\ p(0)=1\end{subarray}}\max_{\lambda\in[\frac{1-\sqrt{1+4d}}{2},\,\frac{1-\sqrt{1+4\beta^{2}}}{2}]\cup[\frac{1+\sqrt{1+4\beta^{2}}}{2},\,\frac{1+\sqrt{1+4d}}{2}],}|p(\lambda_{i})|
2(dβd+β)k.\displaystyle\leq 2\Bigg{(}\frac{\sqrt{d}-\beta}{\sqrt{d}+\beta}\Bigg{)}^{k}.

Note that 𝐫2k+2𝐫2k+1\|\mathbf{r}_{2k+2}\|\leq\|\mathbf{r}_{2k+1}\| holds due to the minimization property of MINRES [4]. An important consequence of Proposition 4.1 is that the number of MINRES required to reach a prescribed level of the residual, when preconditioned by 𝒫d\mathcal{P}_{d}, is independent of size of the discrete problem and the parameter μ\mu. Hence, 𝒫d\mathcal{P}_{d} is an effective preconditioner for (62).

4.2 Block triangular Schur complement preconditioning

We now consider block triangular Schur complement preconditioners and the convergence of GMRES.

We start by pointing out that several inexact block triangular Schur complement preconditioners have been studied in [7, Appendix A] for general saddle point systems and are shown to perform very similarly. As such, we focus here only on the following block lower triangular preconditioner,

𝒫t=[A0BMp].\mathcal{P}_{t}=\begin{bmatrix}A&0\\ -B^{\circ}&-M_{p}^{\circ}\end{bmatrix}. (70)

This preconditioner is nonsingular. Moreover, it corresponds to the choice S^=Mp\hat{S}=M_{p}^{\circ} (the pressure mass matrix) for the approximation of the Schur complement S=BA1(B)TS=B^{\circ}A^{-1}(B^{\circ})^{T}.

      Proposition 4.2.

The residual of GMRES applied to the preconditioned system 𝒫t1𝒜\mathcal{P}_{t}^{-1}\mathcal{A} is bounded by

𝐫k𝐫02(1+d+(dλmax(Mp)λmin(A))12)(dβd+β)k2.\displaystyle\frac{\|\mathbf{r}_{k}\|}{\|\mathbf{r}_{0}\|}\leq 2\left(1+d+\Big{(}\frac{d\,\lambda_{\max}(M_{p}^{\circ})}{\lambda_{\min}(A)}\Big{)}^{\frac{1}{2}}\right)\Big{(}\frac{\sqrt{d}-\beta}{\sqrt{d}+\beta}\Big{)}^{k-2}. (71)
Proof.

Applying [7, Lemma A.1] to (62) with preconditioner (70), we obtain the bound for the GMRES residual as

𝐫k𝐫0(1+S^1S+A1(B)T)minpk1p(0)=1p(S^1S).\displaystyle\frac{\|\mathbf{r}_{k}\|}{\|\mathbf{r}_{0}\|}\leq\left(1+\|\hat{S}^{-1}S\|+\|A^{-1}(B^{\circ})^{T}\|\right)\min\limits_{\begin{subarray}{c}p\in\mathbb{P}_{k-1}\\ p(0)=1\end{subarray}}\|p(\hat{S}^{-1}S)\|. (72)

From Lemma 4.1, we have S^1Sd\|\hat{S}^{-1}S\|\leq d. In the following, we estimate A1(B)T\|A^{-1}(B^{\circ})^{T}\| and minpp(S^1S)\min\limits_{p}\|p(\hat{S}^{-1}S)\|.

For A1(B)T\|A^{-1}(B^{\circ})^{T}\|, using (64) and the fact that AA and MpM_{p}^{\circ} are SPD, we have

A1(B)T2\displaystyle\|A^{-1}(B^{\circ})^{T}\|^{2} =sup𝐩0𝐩BA1A1(B)T𝐩𝐩T𝐩\displaystyle=\sup_{\mathbf{p}\neq 0}\frac{\mathbf{p}B^{\circ}A^{-1}A^{-1}(B^{\circ})^{T}\mathbf{p}}{\mathbf{p}^{T}\mathbf{p}}
=sup𝐩0𝐩(Mp)12(Mp)12BA1A1(B)T(Mp)12(Mp)12𝐩𝐩T𝐩\displaystyle=\sup_{\mathbf{p}\neq 0}\frac{\mathbf{p}(M_{p}^{\circ})^{\frac{1}{2}}(M_{p}^{\circ})^{-\frac{1}{2}}B^{\circ}A^{-1}A^{-1}(B^{\circ})^{T}(M_{p}^{\circ})^{-\frac{1}{2}}(M_{p}^{\circ})^{\frac{1}{2}}\mathbf{p}}{\mathbf{p}^{T}\mathbf{p}}
λmax(A1)λmax(Mp)sup𝐮0𝐮TB(Mp)1(B)T𝐮𝐮TA𝐮\displaystyle\leq\lambda_{\max}(A^{-1})\lambda_{\max}(M_{p}^{\circ})\sup_{\mathbf{u}\neq 0}\frac{\mathbf{u}^{T}B^{\circ}(M_{p}^{\circ})^{-1}(B^{\circ})^{T}\mathbf{u}}{\mathbf{u}^{T}A\mathbf{u}}
dλmax(Mp)λmin(A).\displaystyle\leq\frac{d\,\lambda_{\max}(M_{p}^{\circ})}{\lambda_{\min}(A)}. (73)

Next, we look into the factor minpp(S^1S)\min_{p}\|p(\hat{S}^{-1}S)\|. Notice that both SS and S^\hat{S} are symmetric for the current situation. Moreover, SS is singular with its null space given by Null((B)T)\text{Null}((B^{\circ})^{T}). Since the modified scheme is consistent, the corresponding system with S^1S\hat{S}^{-1}S is consistent as well. It is known (e.g., see [4, Section 8.3.4] or [16, Chapter 10]) that the convergence of GMRES with zero initial guess is not affected by the singular eigenvalues when applied to consistent systems. Thus, denoting the eigenvalues of S^1S\hat{S}^{-1}S by γ1=0<γ2γN\gamma_{1}=0<\gamma_{2}\leq\cdots\leq\gamma_{N}, we have

minpk1p(0)=1p(S^11S1)=minpk2p(0)=1maxi=2,,N|p(γi)|minpk2p(0)=1maxγ[γ2,γN]|p(γ)|.\min\limits_{\begin{subarray}{c}p\in\mathbb{P}_{k-1}\\ p(0)=1\end{subarray}}\|p(\hat{S}_{1}^{-1}S_{1})\|=\min\limits_{\begin{subarray}{c}p\in\mathbb{P}_{k-2}\\ p(0)=1\end{subarray}}\max_{i=2,...,N}|p(\gamma_{i})|\leq\min\limits_{\begin{subarray}{c}p\in\mathbb{P}_{k-2}\\ p(0)=1\end{subarray}}\max_{\gamma\in[\gamma_{2},\gamma_{N}]}|p(\gamma)|.

From Lemma 4.2 and using shifted Chebyshev polynomials for the above minmax problem (e.g., see [5, Pages 50-52]), we get

minpk2p(0)=1maxγ[β2,d]|p(γ)|2(dβd+β)k2.\min\limits_{\begin{subarray}{c}p\in\mathbb{P}_{k-2}\\ p(0)=1\end{subarray}}\max_{\gamma\in[\beta^{2},d]}|p(\gamma)|\leq 2\Big{(}\frac{\sqrt{d}-\beta}{\sqrt{d}+\beta}\Big{)}^{k-2}.

Combining the above results, we obtain (71). ∎

Recall that AA is the stiffness matrix of the WG approximation of the Laplacian operator. It is known that for quasi-uniform meshes, both λmax(Mp)\lambda_{\max}(M_{p}^{\circ}) and λmin(A)\lambda_{\min}(A) are in the order of 1/N1/N and λmax/λmin(A)\lambda_{\max}/\lambda_{\min}(A) is bounded above by a constant independent of hh and μ\mu. Then, (71) implies that the convergence of GMRES, applied to the modified scheme (62), with the block triangular Schur complement conditioner (70), is independent of hh and μ\mu.

5 Numerical experiments

In this section we present some two- and three-dimensional numerical results to demonstrate the performance of MINRES/GMRES with the block Schur complement preconditioners for the modified system (62). We use MATLAB’s function minres with tol=109tol=10^{-9} for 2D examples and tol=108tol=10^{-8} for 3D example, with a maximum of 1000 iterations and the zero vector as the initial guess. Similarly, we use MATLAB’s function gmres with tol=109tol=10^{-9} for 2D examples and tol=108tol=10^{-8} for 3D examples, restart=30restart=30, and the zero vector as the initial guess. The implementation of the block preconditioners requires the exact inversion of the diagonal blocks. The inversion of the diagonal mass matrix MpM_{p}^{\circ} is straightforward. The leading block AA represents the WG approximation of the Laplacian operator, and linear systems associated with AA are solved using the conjugate gradient method preconditioned with incomplete Cholesky decomposition. The incomplete Cholesky decomposition is carried out using MATLAB’s function ichol with threshold dropping and the drop tolerance is 10310^{-3}. Numerical experiments are conducted on triangular and tetrahedral meshes as shown in Fig. 1.

Refer to caption
(a) A triangular mesh
Refer to caption
(b) A tetrahedral mesh
Figure 1: Examples of meshes used for the computation in two and three dimensions.

5.1 The two-dimensional example

This two-dimensional (2D) example is adopted from [10] where Ω=(0,1)2\Omega=(0,1)^{2},

𝐮=[ex(ycos(y)+sin(y))exysin(y)],p=2exsin(y),𝐟=[2(1μ)exsin(y)2(1μ)excos(y)].\displaystyle\mathbf{u}=\begin{bmatrix}-e^{x}(y\cos(y)+\sin(y))\\ e^{x}y\sin(y)\end{bmatrix},\quad p=2e^{x}\sin(y),\quad\mathbf{f}=\begin{bmatrix}2(1-\mu)e^{x}\sin(y)\\ 2(1-\mu)e^{x}\cos(y)\end{bmatrix}.

We take μ=1\mu=1 and 10410^{-4}.

To examine the convergence of the modified scheme (25), we show the L2L^{2} error of the velocity in Table 1. Theoretically, this error has the optimal first-order convergence as stated in Theorem 3.1. In Table 1, the L2L^{2} error of velocity for both μ=1\mu=1 and 10410^{-4} shows the first-order convergence rate. Interestingly, the error is almost the same for μ=1\mu=1 and 10410^{-4}.

Table 1: The 2D Example. The L2L^{2} error of the velocity for the modified scheme (62) (consistency enforcement) with μ=1\mu=1 and μ=104\mu=10^{-4}.
μ=1\mu=1 μ=104\mu=10^{-4}
NN 𝐮𝐮h\|\mathbf{u}-\mathbf{u}_{h}\| conv. rate 𝐮𝐮h\|\mathbf{u}-\mathbf{u}_{h}\| conv. rate
232 9.428879e-02 9.428878e-02
918 4.719342e-02 1.006 4.719342e-02 1.006
3680 2.352019e-02 1.003 2.352019e-02 1.003
14728 1.174930e-02 1.000 1.174930e-02 1.000
58608 5.893651e-03 0.999 5.893651e-03 0.999

Recall from Proposition 4.1 that the residual of MINRES with the preconditioner 𝒫d\mathcal{P}_{d} is independent of μ\mu and hh. Table 2 shows a consistent number of MINRES iterations as the mesh is refined. Furthermore, the iteration remains stable for small values of μ\mu. Similar observations can be made for GMRES from Table 3.

The good performance of the block Schur complement preconditioners 𝒫d\mathcal{P}_{d} and 𝒫t\mathcal{P}_{t} can be explained from the number of MINRES/GMRES iterations shown in Tables 2 and 3. The number of iterations required to reach convergence remains relatively small and consistent with mesh refinement and various values of μ\mu. If without preconditioning, on the other hand, MINRES and GMRES would take more than 30,000 iterations to reach convergence, which is a significant difference compared to the solvers with preconditioning. We also observe that the number of MINRES iterations is almost double that of GMRES, as shown by comparing Tables 2 and 3. This observation is consistent with the estimates in Proposition 4.1 and Proposition 4.2.

Table 2: The 2D Example. The number of MINRES iterations required to reach convergence for the preconditioned system 𝒫d1𝒜\mathcal{P}_{d}^{-1}\mathcal{A} with block diagonal Schur complement preconditioning.
μ\mu NN 232 918 3680 14728 58608
11 43 47 49 49 47
10410^{-4} 42 48 54 58 60
Table 3: The 2D Example. The number of GMRES iterations required to reach convergence for the preconditioned system 𝒫t1𝒜\mathcal{P}_{t}^{-1}\mathcal{A} with block triangular Schur complement preconditioning.
μ\mu NN 232 918 3680 14728 58608
11 21 23 24 25 25
10410^{-4} 23 25 27 27 27

5.2 The three-dimensional example

This three-dimensional (3D) example is adopted from deal.II [2] step-56. Here, Ω=(0,1)3\Omega=(0,1)^{3} and

𝐮=[2sin(πx)πycos(πx)πzcos(πx)],p=sin(πx)cos(πy)sin(πz),\displaystyle\mathbf{u}=\begin{bmatrix}2\sin(\pi x)\\ -\pi y\cos(\pi x)\\ -\pi z\cos(\pi x)\end{bmatrix},\quad p=\sin(\pi x)\cos(\pi y)\sin(\pi z),
𝐟=[2μπ2sin(πx)+πcos(πx)cos(πy)sin(πz)μπ3ycos(πx)πsin(πy)sin(πx)sin(πz)μπ3zcos(πx)+πsin(πx)cos(πy)cos(πz)].\displaystyle\mathbf{f}=\begin{bmatrix}2\mu\pi^{2}\sin(\pi x)+\pi\cos(\pi x)\cos(\pi y)\sin(\pi z)\\ -\mu\pi^{3}y\cos(\pi x)-\pi\sin(\pi y)\sin(\pi x)\sin(\pi z)\\ -\mu\pi^{3}z\cos(\pi x)+\pi\sin(\pi x)\cos(\pi y)\cos(\pi z)\end{bmatrix}.

The number of MINRES/GMRES iterations for the preconditioned systems for μ=1\mu=1 and 10410^{-4} is listed in Tables 4 and 5, respectively. The number of iterations remains relatively small and consistent with various in μ\mu and hh, which is consistent with the fact that the bounds of the residual in Propositions 4.1 and 4.2 are independent of μ\mu and hh. Once again, MINRES/GMRES without preconditioning would require more than 30,000 iterations to reach convergence.

Table 4: The 3D Example. The number of MINRES iterations required to reach convergence for the preconditioned system 𝒫d1𝒜\mathcal{P}_{d}^{-1}\mathcal{A} with block diagonal Schur complement preconditioning.
μ\mu NN 4046 7915 32724 112078 266555
11 59 59 63 67 67
10410^{-4} 62 62 70 76 78
Table 5: The 3D Example. The number of GMRES iterations required to reach convergence for the preconditioned system 𝒫t1𝒜\mathcal{P}_{t}^{-1}\mathcal{A} with block triangular Schur complement preconditioning.
μ\mu NN 4046 7915 32724 112078 266555
11 30 30 31 32 33
10410^{-4} 34 35 37 38 38

6 Conclusions

In the previous sections we have studied the iterative solution of the singular saddle point system arising from the lowest-order weak Galerkin discretization of Stokes flow problems. In general the system is inconsistent when the boundary datum is not identically zero. This inconsistency can cause iterative methods such as MINRES and GMRES to fail to converge. We have proposed a simple strategy to enforce the consistency by modifying the right-hand side of the second equation (cf. (23)). We have shown in Theorem 3.1 that the optimal-order convergence of the numerical solutions is not affected by the modification.

We have considered block diagonal and triangular Schur complement preconditioning for the iterative solution of the modified system. In Section 4.1, we have studied the block diagonal Schur complement preconditioner (65) and established bounds for the eigenvalues of the preconditioned system (see Lemma 4.3) as well as for the residual of MINRES applied to the preconditioned system (cf. Proposition 4.1). In Section 4.2, we have studied the block triangular Schur complement preconditioner (70) and derived the asymptotic error bound in Proposition 4.2 for GMRES applied to the preconditioned system. These bounds show that both the convergence of MINRES and GMRES for the corresponding preconditioned systems is independent of hh and μ\mu. The optimal-order convergence of the modified scheme (62) and effectiveness of the block diagonal and triangular Schur complement preconditioners have been verified by two- and three-dimensional numerical examples in Section 5.

Acknowledgments

W. Huang was supported in part by the Air Force Office of Scientific Research (AFOSR) grant FA9550-23-1-0571 and the Simons Foundation grant MPS-TSM-00002397.

References

  • [1] M. Ainsworth and C. Parker. A mass conserving mixed hphp-FEM scheme for Stokes flow. Part III: Implementation and preconditioning. SIAM J. Numer. Anal., 60:1574–1606, 2022.
  • [2] D. Arndt, W. Bangerth, B. Blais, T. C. Clevenger, M. Fehling, A. Grayver, T. Heister, L. Heltai, M. Kronbichler, M. Maier, P. Munch, J.-P. Pelteret, R. Rastak, I. Tomas, B. Turcksin, Z. Wang, and D. Wells. The deal.ii library, version 9.2. J. Numer. Math., 28(3):131–146, 2020.
  • [3] T. Bevilacqua, F. Dassi, S. Zampini, and S. Scacchi. BDDC preconditioners for virtual element approximations of the three-dimensional Stokes equations. SIAM J. Sci. Comput., 46:A156–A178, 2024.
  • [4] H. C. Elman, D. J. Silvester, and A. J. Wathen. Finite elements and fast iterative solvers: with applications in incompressible fluid dynamics. Numerical Mathematics and Scientific Computation. Oxford University Press, Oxford, second edition, 2014.
  • [5] A. Greenbaum. Iterative methods for solving linear systems, volume 17 of Frontiers in Applied Mathematics. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 1997.
  • [6] W. Huang and Y. Wang. Discrete maximum principle for the weak Galerkin method for anisotropic diffusion problems. Comm. Comput. Phys., 18:65–90, 2015.
  • [7] W. Huang and Z. Wang. Convergence analysis of iterative solution with inexact block preconditioning for weak Galerkin finite element approximation of Stokes flow. arXiv:2409.16477, 2024.
  • [8] P. L. Lederer and C. Merdon. Gradient-robust hybrid DG discretizations for the compressible Stokes equations. J. Sci. Comput., page 54, 2024.
  • [9] J. Liu, S. Tavener, and Z. Wang. Lowest-order weak Galerkin finite element method for Darcy flow on convex polygonal meshes. SIAM J. Sci. Comput., 40:B1229–B1252, 2018.
  • [10] L. Mu. Pressure robust weak Galerkin finite element methods for Stokes problems. SIAM J. Sci. Comput., 42:B608–B629, 2020.
  • [11] L. Mu, X. Wang, and X. Ye. A modified weak Galerkin finite element method for the Stokes equations. J. Comput. Appl. Math., 275:79–90, 2015.
  • [12] L. Mu, X. Ye, and S. Zhang. A stabilizer-free, pressure-robust, and superconvergence weak Galerkin finite element method for the Stokes equations on polytopal mesh. SIAM J. Sci. Comput., 43:A2614–A2637, 2021.
  • [13] C. C. Paige and M. A. Saunders. Solutions of sparse indefinite systems of linear equations. SIAM J. Numer. Anal., 12:617–629, 1975.
  • [14] X. Tu and B. Wang. A BDDC algorithm for the Stokes problem with weak Galerkin discretizations. Comput. Math. Appl., 76:377–392, 2018.
  • [15] X. Tu, B. Wang, and J. Zhang. Analysis of BDDC algorithms for Stokes problems with hybridizable discontinuous Galerkin discretizations. Electron. Trans. Numer. Anal., 52:553–570, 2020.
  • [16] H. A. van der Vorst. Iterative Krylov Methods for Large Linear Systems. Cambridge Monographs on Applied and Computational Mathematics. Cambridge University Press, 2003.
  • [17] G. Wang, L. Mu, Y. Wang, and Y. He. A pressure-robust virtual element method for the Stokes problem. Comput. Meth. Appl. Mech. Engrg., 382:113879, 2021.
  • [18] J. Wang and X. Ye. A weak Galerkin finite element method for the Stokes equations. Adv. Comput. Math., 42:155–174, 2016.
  • [19] Z. Wang, R. Wang, and J. Liu. Robust weak Galerkin finite element solvers for Stokes flow based on a lifting operator. Comput. Math. Appl., 125:90–100, 2022.