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

A low order divergence-free H(div)-conforming finite element method for Stokes flows

Xu Li School of Mathematics, Shandong University, Jinan 250100, China xulisdu@126.com  and  Hongxing Rui School of Mathematics, Shandong University, Jinan 250100, China hxrui@sdu.edu.cn
Abstract.

In this paper, we propose a 𝑷1cRT0P0{\bm{P}_{1}^{c}}\oplus{RT0}-P0 discretization of the Stokes equations on general simplicial meshes in two/three dimensions (2D/3D), which yields an exactly divergence-free and pressure-independent velocity approximation with optimal order. Our method has the following features. Firstly, the global number of the degrees of freedom of our method is the same as the low order Bernardi and Raugel (BB-RR) finite element method (Bernardi and Raugel, 1985), while the number of the non-zero entries of the former is about half of the latter in the velocity-velocity region of the coefficient matrix. Secondly, the 𝑷1c{\bm{P}_{1}^{c}} component of the velocity, the RT0RT0 component of the velocity and the pressure seem to solve a popular 𝑷1cRT0P0{\bm{P}_{1}^{c}}-{RT0}-P0 discretization of a poroelastic-type system formally. Finally, our method can be easily transformed into a pressure-robust and stabilized 𝑷1cP0{\bm{P}_{1}^{c}}-P0 discretization for the Stokes problem via the static condensation of the RT0RT0 component, which has a much smaller number of global degrees of freedom. Numerical experiments illustrating the robustness of our method are also provided.

Key words and phrases:
Stokes problem; divergence-free velocity; pressure-robustness; optimal error estimates.
The second author is the corresponding author.

1. Introduction

The importance of strong mass conservation has been found in a variety of real-world applications (see, e.g., [36, 17, 34]). For the incompressible Stokes problem, pointwise mass conservation requires an exactly divergence-free velocity approximation. Another highly relevant concept is pressure-robustness (see, e.g., [25, 35, 18]), which describes a type of robustness that the velocity error is independent of the pressure. In this paper, we design a low order divergence-free and pressure-robust H(div)H(\operatorname{div})-conforming mixed method for the Stokes problem on general shape-regular simplicial meshes.

Due to the fascinating properties of divergence-free mixed finite element methods [16, 25], constructing efficient divergence-free methods has been a hot research topic during recent years and great progress has been made. However, this is still not an easy thing especially in three dimensions. For conforming divergence-free methods, we refer the readers to [46, 2, 50, 21, 23, 9, 15, 16], to name just a few. Among these methods, the lowest order cases on simplicial meshes might be the ones proposed in [9] and [23], where they have the same number of global degrees of freedom (DOFs) as the classical Bernardi and Raugel finite element method [4]. Like the Bernardi and Raugel element, they both use d+1d+1 face bubbles to enrich the linear polynomial space on each simplex in dd dimensions. The difference lies in the construction of the face bubbles: the bubbles in [9] are piecewise linear on the Powell-Sabin split (e.g., in 3D, a tetrahedron is split into 1212 small tetrahedrons) of a simplex, and the bubbles in [23] are typically piecewise polynomials of degree dd on the barycentric refinement of a simplex. Some applications of the former could be found in [8]. Another popular class of mixed methods is based on H(div)H(\operatorname{div})-conforming elements, which are a class of nonconforming elements for the Stokes equations. In general, there are two ways to stabilize the nonconformity (tangential discontinuity) of the elements. One is to modify the velocity-velocity bilinear form into a discontinuous Galerkin (DG) framework (see, e.g., [12, 48, 27, 26]), the other is to modify the discrete velocity space to impose tangential continuity in some weak sense such as [22, 39, 49, 47]. Among these methods the lowest order case on simplicial meshes might be the DG method based on BDM1P0BDM1-P0 pair [12, 48] (BDM1BDM1: the lowest order Brezzi-Douglas-Marini element [7], P0P0: the piecewise constant space with zero mean). Recently, there is also a class of pressure-robust discretizations developed on classical non-divergence-free mixed methods via divergence-free reconstruction of test functions (cf. [37, 38, 35, 30]), which can be seen as a class of Petrov-Galerkin methods.

One of the advantages of DG H(div)H(\operatorname{div})-conforming mixed methods lies in the simple construction of their velocity spaces. For instance, the lowest order velocity is piecewise linear in arbitrary dimensions. However, in a DG framework we should introduce some integral terms over element interfaces which may increase the cost and enlarge the stencil (sparsity). In addition, the stabilization parameters (the parameters of the jump-penalty term) need to be chosen carefully to guarantee the stability in some cases. The features of conforming and the second type of H(div)H(\operatorname{div})-conforming divergence-free mixed methods are exactly opposite to the DG methods mentioned above. They do not need to modify the bilinear form and are always stable. The constructions of their spaces are, however, a little complicated and high order polynomials are usually included. In a word, constructing divergence-free mixed methods is non-trivial.

The aim of this paper is to investigate a way on how to combine some advantages of the conforming and (DG-) H(div)H(\operatorname{div})-conforming divergence-free mixed methods. In our method, the velocity space is obtained by enriching the continuous vector-valued piecewise linear polynomial space (𝑷1c{\bm{P}_{1}^{c}}) with the lowest order Raviart-Thomas-Nédélec space (RT0RT0) [40, 42], and the pressure is approximated by piecewise constants with zero mean. In a sense, our method can be seen as replacing the Bernardi and Raugel bubbles [4] with the lowest order Raviart-Thomas-Nédélec elements (see Fig. 1). This replacement does bring some advantages. For example, the discrete velocity becomes pointwise divergence-free. However, this turns into a type of nonconforming method and some stabilization strategies are needed. To resolve this issue,

Refer to caption
Refer to caption
Refer to caption
Figure 1. DOFs of BB-RR (left), RT0RT0 (middle) and 𝑷1c{\bm{P}_{1}^{c}} (right) element in 2D. Solid circles indicate function evaluation and arrows indicate normal component evaluation.

we propose a conforming-like formulation which does not involve any integrals over element interfaces. The resulting method has several advantages. Firstly, the construction of the space is very simple in any dimensions. Secondly, we can use the lowest order quadrature rule to compute each entry of the coefficient matrix exactly in some cases. Thirdly, our method instead has much better sparsity than the Bernardi and Raugel method. Next, the discrete system is stable as long as the parameters introduced are positive. Finally, we can further obtain a stabilized pressure-robust 𝑷1cP0{\bm{P}_{1}^{c}}-P0 discretization for the Stokes equations via static condensation. Compared to a direct 𝑷1cP0{\bm{P}_{1}^{c}}-P0 discretization (unstable), this only changes the pressure region of the right-hand side and the pressure-pressure region of the coefficient matrix, which is different from the similar strategy for the Bernardi and Raugel element in [45]. The price to pay for these advantages is a consistency error bounded in an optimal order.

The motivation for studying low order divergence-free elements is threefold: (1) the low order element in this paper is easy to implement; (2) in some practical applications, the available mesh might be fixed due to the complex geometry [24], in which case lower order methods allow a smaller global system; (3) the divergence-free nature usually makes low order elements also give a satisfactory result for some non-trivial Stokes equations such as those with large pressure gradient, small viscosity or long time simulation (unsteady case), cf. [25, 1].

We also want to mention that hybridization is a very efficient way to reduce the computational cost of DG methods, cf. [11]. In [32] and [33], a class of H(div)H(\operatorname{div})-conforming hybrid DG (HDG) methods with discontinuous pressure elements was presented. A cheaper version of them with relaxed H(div)H(\operatorname{div})-conformity for velocity and pressure-robust reconstructions could be found in [28] and [29], where the lowest order pair (after static condensation) has one DOF per facet and dimension for velocity and one DOF per element for pressure, the same as CR1P0CR1-P0 pair (CR1CR1: the first order nonconforming Crouzeix-Raviart element, cf. [13]). A total hybridization (simultaneously for velocity and pressure) formulation for the Stokes equations was analyzed in [43], where the velocity is automatically H(div)H(\operatorname{div})-conforming if the order of the pressure space is one less than the order of the velocity space. To further lower the cost, an embedded DG (EDG) method with pressure-robust reconstruction and an embedded-hybridized DG (EDG-HDG, i.e., EDG for velocity and HDG for pressure) method were developed in [31] and [44], respectively. EDG methods use continuous facet unknowns which have the same number of DOFs as the corresponding conforming methods, while retaining the versatility of DG methods. In a sense, our method shares a similar topic with EDG methods, i.e., combining some advantages of conforming methods and DG methods. The number of the DOFs (in the reduced version) of the lowest order pair in [31] is equal to the one of 𝑷1cP1c{\bm{P}_{1}^{c}}-P_{1}^{c} pair, where P1cP_{1}^{c} is the scalar version of 𝑷1c{\bm{P}_{1}^{c}}. This is also equal to that of the lowest order MINI elements with static condensation of the cell bubbles, which can also be pressure-robust with reconstructions, cf. [30]. The two low order pairs have less DOFs than 𝑷1cP0{\bm{P}_{1}^{c}}-P0 due to a smaller pressure space and can be easily extended to higher order cases. Compared to them, the advantage of our method lies in its automatically divergence-free nature.

The rest of the paper is organized as follows. Section 2 is devoted to describing the model and its discretizations. In Section 3, we analyze the LBB stability, consistency and error estimates. Some numerical studies are shown in Section 4. Finally we do some conclusions in Section 5.

Throughout the paper we use C,C, with or without subscript, to denote a generic positive constant. The standard inner product and norm (seminorm) of the Sobolov space [Hm(X)]n[H^{m}(X)]^{n} or [Hm(X)]n×n[H^{m}(X)]^{n\times n} (n+n\in\mathbb{N}^{+}) are denoted by (,)m,X(\cdot,\cdot)_{m,X} and m,X\|\cdot\|_{m,X} (||m,X|\cdot|_{m,X}), respectively. When m=0m=0 (X=ΩX=\Omega), with the convention that the index mm (XX, respectively) is omitted. For any sub-dimensional face ee, we have ,e\langle\cdot,\cdot\rangle_{e} and ||||e||\cdot||_{e} similarly.

2. Model, notation and method

In this section we consider the incompressible Stokes equations in a bounded domain Ωd(d=2,3)\Omega\subset\mathbb{R}^{d}~(d=2,3)

(2.1a) νΔ𝒖+p\displaystyle-\nu\Delta\bm{u}+\nabla p =𝒇 in Ω,\displaystyle=\bm{f}\quad\text{ in }\Omega,
(2.1b) 𝒖\displaystyle\nabla\cdot\bm{u} =0 in Ω,\displaystyle=0\,\quad\text{ in }\Omega,
(2.1c) 𝒖\displaystyle\bm{u} =𝟎 on Ω,\displaystyle={\bm{0}}\quad\text{ on }\partial\Omega,

where the unknowns (𝒖,p)(\bm{u},p) represent the velocity and pressure, respectively; ν\nu denotes the constant fluid viscosity; 𝒇\bm{f} is the external body force; Eq. 2.1a is usually called the momentum equation and Eq. 2.1b is usually called the continuity equation. For simplicity, we assume that Ω\Omega has a Lipschitz-continuous polygonal/polyhedral boundary Γ\Gamma.

Let V=[H01(Ω)]dV=[H_{0}^{1}(\Omega)]^{d} and W=L02(Ω)W=L_{0}^{2}(\Omega). A commonly used variational form of Eq. 2.1 is:

find(𝒖,p)V×W\displaystyle{\rm find}~(\bm{u},p)\in V\times W suchthat\displaystyle~{\rm such~that}
(2.2a) νa(𝒖,𝒗)b(𝒗,p)\displaystyle\nu a(\bm{u},\bm{v})-b(\bm{v},p) =(𝒇,𝒗)𝒗V,\displaystyle=(\bm{f},\bm{v})\quad\forall~\bm{v}\in V,
(2.2b) b(𝒖,q)\displaystyle b(\bm{u},q) =0qW,\displaystyle=0\qquad\quad\forall~q\in W,

where

a(𝒖,𝒗)=(𝒖,𝒗),b(𝒗,q)=(𝒗,q).a(\bm{u},\bm{v})=(\nabla\bm{u},\nabla\bm{v}),\quad b(\bm{v},q)=(\nabla\cdot\bm{v},q).

And the spaces (V,W)(V,W) satisfy the following well-known inf-sup condition (See [5, §4.2])

(2.3) sup𝒗V{𝟎}b(𝒗,q)𝒗1βqqW,\sup_{\bm{v}\in V\setminus\{\bm{0}\}}\frac{b(\bm{v},q)}{||\bm{v}||_{1}}\geq\beta\|q\|\quad\forall~q\in W,

where β\beta is a positive constant.

2.1. Discretizations

Let 𝒯h\mathcal{T}_{h} be shape-regular triangular or tetrahedral partitions of Ω\Omega [10]. Let hTh_{T} and heh_{e} denote the diameters of elements TT and faces ee, respectively, and h=maxT𝒯hhTh=\max_{T\in\mathcal{T}_{h}}h_{T}. We denote the set of interior faces of 𝒯h\mathcal{T}_{h} by 0\mathcal{E}^{0}, the set of boundary faces by \mathcal{E}^{\partial} and =0\mathcal{E}=\mathcal{E}^{0}\cup\mathcal{E}^{\partial}. We define

H(div;Ω)={𝒗[L2(Ω)]d:𝒗L2(Ω)},H0(div;Ω)={𝒗H(div;Ω):𝒗𝒏|Γ=0}.H(\operatorname{div};\Omega)=\left\{\bm{v}\in[L^{2}(\Omega)]^{d}:\nabla\cdot\bm{v}\in L^{2}(\Omega)\right\},\quad H_{0}(\operatorname{div};\Omega)=\left\{\bm{v}\in H(\operatorname{div};\Omega):\left.\bm{v}\cdot\bm{n}\right|_{\Gamma}=0\right\}.

Let Pk(T)P_{k}(T) (k0k\geq 0) denote the space of polynomials of degree no more than kk on elements TT, and the finite element spaces are set as

Vh1={𝒗V:𝒗|T[P1(T)]dT𝒯h},V_{h}^{1}=\{\bm{v}\in V:\bm{v}|_{T}\in[P_{1}(T)]^{d}\quad\forall\ T\in\mathcal{T}_{h}\},
(2.4) RT0={𝒗H0(div;Ω):𝒗|T[P0(T)]d𝒙P0(T)T𝒯h},RT0=\{\bm{v}\in H_{0}(\operatorname{div};\Omega):\bm{v}|_{T}\in[P_{0}(T)]^{d}\oplus\bm{x}P_{0}(T)\quad\forall\ T\in\mathcal{T}_{h}\},
Wh={qL02(Ω):q|TP0(T)T𝒯h}.W_{h}=\left\{q\in L_{0}^{2}(\Omega):\left.q\right|_{T}\in P_{0}(T)\quad\forall~T\in\mathcal{T}_{h}\right\}.

We note that the functions in Vh1V_{h}^{1} are totally continuous, while the functions in RT0RT0 may have discontinuous tangential components across the interelement boundaries. The following lemma depicts a simple but important fact in our method.

Lemma 1.

Vh1RT0={𝟎}V_{h}^{1}\cap RT0=\{\bf 0\}.

Proof.

Consider a simplest case in two dimensions (see Fig. 2).

Refer to caption
Figure 2. Decompose Ω\Omega into two elements T1T_{1} and T2T_{2}, which share an edge \ell.

Two axes are respectively labeled as xx and yy. From ​​ 2.4, a function 𝒇rtRT0\bm{f}^{rt}\in RT0 can be written as follows,

𝒇rt|T1=(a1+c1x,b1+c1y),𝒇rt|T2=(a2+c2x,b2+c2y),ai,bi,ci,i=1,2.\bm{f}^{rt}|_{T_{1}}=\begin{pmatrix}a_{1}+c_{1}x,b_{1}+c_{1}y\end{pmatrix}^{\top},\quad\bm{f}^{rt}|_{T_{2}}=\begin{pmatrix}a_{2}+c_{2}x,b_{2}+c_{2}y\end{pmatrix}^{\top},\quad a_{i},b_{i},c_{i}\in\mathbb{R},\quad i=1,2.

It suffices to prove that 𝒇rtVh1𝒇rt=𝟎\bm{f}^{rt}\in V_{h}^{1}\Rightarrow\bm{f}^{rt}=\bm{0}. If 𝒇rtVh1\bm{f}^{rt}\in V_{h}^{1}, 𝒇rt\bm{f}^{rt} is totaly continuous across the element interface \ell. In this case, we have

(2.5a) a1+c1x=a2+c2xon,\displaystyle a_{1}+c_{1}x=a_{2}+c_{2}x\quad\text{on}~\ell,
(2.5b) b1+c1y=b2+c2yon.\displaystyle b_{1}+c_{1}y=b_{2}+c_{2}y\quad\text{on}~\ell.

We note that there is at least one of the variables (xx or yy) varying over \ell. For brevity, we assume that xx varies over \ell. From ​​ 2.5a we can obtain

a1=a2,c1=c2.a_{1}=a_{2},\quad c_{1}=c_{2}.

Substituting c1=c2c_{1}=c_{2} into ​​ 2.5b gives b1=b2.b_{1}=b_{2}. Then we have

𝒇rt|Ω=(a1+c1x,b1+c1y).\bm{f}^{rt}|_{\Omega}=\begin{pmatrix}a_{1}+c_{1}x,b_{1}+c_{1}y\end{pmatrix}^{\top}.

If we further demand that 𝒇rt|Γ=𝟎\bm{f}^{rt}|_{\Gamma}=\bm{0}, through the similar analysis, one can obtain a1=b1=c1=0.a_{1}=b_{1}=c_{1}=0. Thus we complete the proof of this simplest case. The proofs of general decompositions and three-dimensional case are a natural extension of this. ∎

Denote by Vh=Vh1RT0V_{h}=V_{h}^{1}\oplus RT0 the discrete velocity space. Let 𝒯:={𝚽e,e0}\mathcal{RT}:=\{\bm{\Phi}_{e},~e\in\mathcal{E}^{0}\} denote the set of the commonly used bases in RT0RT0, where 𝚽e\bm{\Phi}_{e} is only supported on two elements T1T_{1} and T2T_{2} which share the face ee and 𝚽e𝒏e\bm{\Phi}_{e}\cdot\bm{n}_{e^{\prime}} vanishes on any face eee^{\prime}\neq e. Here 𝒏e\bm{n}_{e^{\prime}} denotes the unit normal vector to ee^{\prime}.

For any 𝒗hVh\bm{v}_{h}\in V_{h}, we have the unique decompositions

𝒗h=𝒗h1+𝒗hR,𝒗hR=e0ve𝚽e with 𝒗h1Vh1,𝒗hRRT0.\bm{v}_{h}=\bm{v}_{h}^{1}+\bm{v}_{h}^{R},\quad\bm{v}_{h}^{R}=\sum_{e\in\mathcal{E}^{0}}v_{e}\bm{\Phi}_{e}\quad\text{ with }\bm{v}_{h}^{1}\in V_{h}^{1},\ \bm{v}_{h}^{R}\in RT0.

Our discretization is presented as:

Find(𝒖h,ph)Vh×Wh\displaystyle{\rm Find}~(\bm{u}_{h},p_{h})\in V_{h}\times W_{h}~ suchthat\displaystyle{\rm such~that}
(2.6a) νah(𝒖h,𝒗h)b(𝒗h,ph)\displaystyle\nu a_{h}(\bm{u}_{h},\bm{v}_{h})-b(\bm{v}_{h},p_{h}) =(𝒇,𝒗h)𝒗hVh,\displaystyle=(\bm{f},\bm{v}_{h})\quad\forall~\bm{v}_{h}\in V_{h},
(2.6b) b(𝒖h,qh)\displaystyle b(\bm{u}_{h},q_{h}) =0qhWh,\displaystyle=0\qquad\ \,\quad\forall~q_{h}\in W_{h},

where

(2.7) ah(𝒖h,𝒗h)=a(𝒖h1,𝒗h1)+aR(𝒖hR,𝒗hR).a_{h}(\bm{u}_{h},\bm{v}_{h})=a(\bm{u}_{h}^{1},\bm{v}_{h}^{1})+a^{R}(\bm{u}_{h}^{R},\bm{v}_{h}^{R}).

Here aR:RT0×RT0Ra^{R}:RT0\times RT0\rightarrow R has several choices as

(2.8) aR(𝒖hR,𝒗hR)=a0(𝒖hR,𝒗hR):=T𝒯hαThT2(𝒖hR,𝒗hR)T,a^{R}(\bm{u}_{h}^{R},\bm{v}_{h}^{R})=a^{0}(\bm{u}_{h}^{R},\bm{v}_{h}^{R}):=\sum_{T\in\mathcal{T}_{h}}\alpha_{T}h_{T}^{-2}(\bm{u}_{h}^{R},\bm{v}_{h}^{R})_{T},

or

(2.9) aR(𝒖hR,𝒗hR)=aD(𝒖hR,𝒗hR):=T𝒯heT0αThT2ueve(𝚽e,𝚽e)T,a^{R}(\bm{u}_{h}^{R},\bm{v}_{h}^{R})=a^{D}(\bm{u}_{h}^{R},\bm{v}_{h}^{R}):=\sum_{T\in\mathcal{T}_{h}}\sum_{e\in\partial T\cap\mathcal{E}^{0}}\alpha_{T}h_{T}^{-2}u_{e}v_{e}(\bm{\Phi}_{e},\bm{\Phi}_{e})_{T},

or

(2.10) aR(𝒖hR,𝒗hR)=adiv(𝒖hR,𝒗hR):=T𝒯heT0αTueve(𝚽e,𝚽e)T,a^{R}(\bm{u}_{h}^{R},\bm{v}_{h}^{R})=a^{\operatorname{div}}(\bm{u}_{h}^{R},\bm{v}_{h}^{R}):=\sum_{T\in\mathcal{T}_{h}}\sum_{e\in\partial T\cap\mathcal{E}^{0}}\alpha_{T}u_{e}v_{e}(\nabla\cdot\bm{\Phi}_{e},\nabla\cdot\bm{\Phi}_{e})_{T},

where αT\alpha_{T}, T𝒯hT\in\mathcal{T}_{h} are positive parameters.

Note that

a0(𝒖hR,𝒗hR)=T𝒯he0Te0TαThT2ueve(𝚽e,𝚽e)T,a^{0}(\bm{u}_{h}^{R},\bm{v}_{h}^{R})=\sum_{T\in\mathcal{T}_{h}}\sum_{e\in\mathcal{E}^{0}\cap\partial T}\sum_{e^{\prime}\in\mathcal{E}^{0}\cap\partial T}\alpha_{T}h_{T}^{-2}u_{e}v_{e^{\prime}}(\bm{\Phi}_{e},\bm{\Phi}_{e^{\prime}})_{T},

and

aD(𝒖hR,𝒗hR)=T𝒯he0Te0TδeeαThT2ueve(𝚽e,𝚽e)T,a^{D}(\bm{u}_{h}^{R},\bm{v}_{h}^{R})=\sum_{T\in\mathcal{T}_{h}}\sum_{e\in\mathcal{E}^{0}\cap\partial T}\sum_{e^{\prime}\in\mathcal{E}^{0}\cap\partial T}{\delta}_{ee^{\prime}}\alpha_{T}h_{T}^{-2}u_{e}v_{e^{\prime}}(\bm{\Phi}_{e},\bm{\Phi}_{e^{\prime}})_{T},

where δee=1{\delta}_{ee^{\prime}}=1 if e=ee=e^{\prime}, otherwise vanishes. Thus, aD(,)a^{D}(\cdot,\cdot) can be regarded as a diagonalization version of a0(,)a^{0}(\cdot,\cdot). Clearly the matrices from aD(,)a^{D}(\cdot,\cdot) and adiv(,)a^{\operatorname{div}}(\cdot,\cdot) are both diagonal. In the next section, we will prove that the three choices of aR(,)a^{R}(\cdot,\cdot) are equivalent and formula ​​ 2.6 is stable as long as αT>0\alpha_{T}>0 for all T𝒯hT\in\mathcal{T}_{h}.

Remark 1 (The motivation of aRa^{R}).

The three forms of aRa^{R} proposed here can also be regarded as some penalty-type stabilizations. From Lemma 1 we can further obtain RT0V={𝟎}RT0\cap V=\{\bm{0}\}. In other words, there is no conforming element in RT0RT0. Thus, we penalize the nonconformity in another way. That is, we penalize the whole RT0RT0 component, and enforce 𝒖hR=𝟎\bm{u}_{h}^{R}=\bm{0} when h0h\rightarrow 0.

The formulation ​​ 2.6 can be rewritten to the following equivalent three-field form:

Find(𝒖h1,𝒖hR,ph)Vh1×RT0×Wh\displaystyle{\rm Find}~(\bm{u}_{h}^{1},\bm{u}_{h}^{R},p_{h})\in V_{h}^{1}\times RT0\times W_{h}~ suchthat\displaystyle{\rm such~that}
(2.11a) νa(𝒖h1,𝒗h1)b(𝒗h1,ph)\displaystyle\nu a(\bm{u}_{h}^{1},\bm{v}_{h}^{1})-b(\bm{v}_{h}^{1},p_{h}) =(𝒇,𝒗h1)𝒗h1Vh1,\displaystyle=(\bm{f},\bm{v}_{h}^{1})~\quad\forall~\bm{v}^{1}_{h}\in V_{h}^{1},
(2.11b) νaR(𝒖hR,𝒗hR)b(𝒗hR,ph)\displaystyle\nu a^{R}(\bm{u}_{h}^{R},\bm{v}_{h}^{R})-b(\bm{v}_{h}^{R},p_{h}) =(𝒇,𝒗hR)𝒗hRRT0,\displaystyle=(\bm{f},\bm{v}_{h}^{R})\quad\forall~\bm{v}^{R}_{h}\in RT0,
(2.11c) b(𝒖h1,qh)+b(𝒖hR,qh)\displaystyle b(\bm{u}_{h}^{1},q_{h})+b(\bm{u}_{h}^{R},q_{h}) =0qhWh.\displaystyle=0\qquad\ \,\quad\forall~q_{h}\in W_{h}.
Remark 2.

Eq. 2.11a is a 𝑷1c{\bm{P}_{1}^{c}} discretization of the momentum equation ​​ 2.1a, while ​​ 2.11b is very similar to a low order discretization of the Darcy equation with the permeability κ=(ναThT2)1\kappa=(\nu\alpha_{T}h_{T}^{-2})^{-1} on the element TT, when a0a^{0} is used. Thus, our method is very similar to a 𝑷1cRT0P0{\bm{P}_{1}^{c}}-RT0-P0 discretization of a three-field (displacement - (Darcy) velocity - pressure) poroelastic-type system [41].

Since RT0=Wh\nabla\cdot RT0=W_{h} (see [5, Proposition 2.3.3]) and Vh1Wh\nabla\cdot V_{h}^{1}\subset W_{h}, we have Vh=Wh\nabla\cdot V_{h}=W_{h}. Thus, from ​​ 2.6b we know

Theorem 2.1 (Mass Conservation).

The finite element solution 𝐮h\bm{u}_{h} of ​​ 2.6 features a full satisfaction of the continuity equation, which means

(2.12) 𝒖h0.\nabla\cdot\bm{u}_{h}\equiv 0.

In other words, our method conserves mass pointwise.

Proof.

Taking qh=𝒖hq_{h}=\nabla\cdot\bm{u}_{h} in ​​ 2.6b gives 𝒖h2=0.||\nabla\cdot\bm{u}_{h}||^{2}=0. Then ​​ 2.12 follows. ∎

In the continuous case, there is a fundamental invariance property: Changing the external force by a gradient field changes only the pressure solution, and not the velocity; in symbols,

𝒇𝒇+ψ(𝒖,p)(𝒖,p+ψ),\bm{f}\rightarrow\bm{f}+\nabla\psi\quad\Longrightarrow\quad(\bm{u},p)\rightarrow(\bm{u},p+\psi),

where ψH1(Ω)/\psi\in H^{1}(\Omega)/\mathbb{R}. This phenomenon is highly relevant to pressure-robustness [25]. In our method, we have the following discrete invariance property.

Theorem 2.2.

The following invariance property holds in the discrete scheme ​​ 2.6:

(2.13) 𝒇𝒇+ψ(𝒖h,ph)(𝒖h,ph+Phψ),\bm{f}\rightarrow\bm{f}+\nabla\psi\quad\Longrightarrow\quad(\bm{u}_{h},p_{h})\rightarrow(\bm{u}_{h},p_{h}+P_{h}\psi),

where Ph:WWhP_{h}:W\rightarrow W_{h} is the L2L^{2} projection (see ​​ 3.1).

Proof.

Since Vh=Wh\nabla\cdot V_{h}=W_{h}, we have

(2.14) (ph+Phψ,𝒗h)=(ph+ψ,𝒗h)=(ph,𝒗h)(ψ,𝒗h).(p_{h}+P_{h}\psi,\nabla\cdot\bm{v}_{h})=(p_{h}+\psi,\nabla\cdot\bm{v}_{h})=(p_{h},\nabla\cdot\bm{v}_{h})-(\nabla\psi,\bm{v}_{h}).

Substituting ​​ 2.14 into ​​ 2.6 gives

(2.15a) ah(𝒖h,𝒗h)b(𝒗h,ph+Phψ)\displaystyle a_{h}(\bm{u}_{h},\bm{v}_{h})-b(\bm{v}_{h},p_{h}+P_{h}\psi) =(𝒇+ψ,𝒗h)𝒗hVh,\displaystyle=(\bm{f}+\nabla\psi,\bm{v}_{h})\quad\forall~\bm{v}_{h}\in V_{h},
(2.15b) b(𝒖h,qh)\displaystyle b(\bm{u}_{h},q_{h}) =0qhWh,\displaystyle=0\qquad\quad\quad\quad\,\quad\forall~q_{h}\in W_{h},

which implies that (𝒖h,ph+Phψ)(\bm{u}_{h},p_{h}+P_{h}\psi) is a solution corresponding to 𝒇+ψ\bm{f}+\nabla\psi. Together with the unique solvability of our methods (see next section), the invariance property ​​ 2.13 follows. ∎

2.2. A pressure-robust and stabilized P1cP0{P_{1}^{c}}-P0 scheme

A block form of the fully discrete problem ​​ 2.6 or ​​ 2.11 can be written as follows,

(2.16) 𝒜S(ULURP)=(FLFR𝟎), with 𝒜S=(ALL0GLT0ARRGRTGLGR0),\mathcal{A}_{S}\begin{pmatrix}U_{L}\\ U_{R}\\ P\end{pmatrix}=\begin{pmatrix}F_{L}\\ F_{R}\\ \bm{0}\end{pmatrix},\quad\text{ with }\mathcal{A}_{S}=\begin{pmatrix}A_{LL}&0&-G_{L}^{T}\\ 0&A_{RR}&-G_{R}^{T}\\ G_{L}&G_{R}&0\end{pmatrix},

where UL,UR,{U}_{L},{U}_{R}, and P{P} are the unknown vectors corresponding to the continuous linear component of the velocity, the RT0RT0 component of the velocity, and the pressure, respectively. The blocks ALLA_{LL}, ARRA_{RR}, GLG_{L}, GRG_{R}, FLF_{L} and FRF_{R} correspond to the bilinear forms a(𝒖h1,𝒗h1)a(\bm{u}_{h}^{1},\bm{v}_{h}^{1}), aR(𝒖hR,𝒗hR)a^{R}(\bm{u}_{h}^{R},\bm{v}_{h}^{R}), b(𝒖h1,qh)b(\bm{u}_{h}^{1},q_{h}), b(𝒖hR,qh)b(\bm{u}_{h}^{R},q_{h}), (𝒇,𝒗h1)(\bm{f},\bm{v}_{h}^{1}) and (𝒇,𝒗hR)(\bm{f},\bm{v}_{h}^{R}), respectively.

For a classical Bernardi and Raugel discretization, the cross region of the linear component and the bubble component of the velocity in the coefficient matrix is non-zero. In other words, the matrices from a(𝒖h1,𝒗hb)a(\bm{u}_{h}^{1},\bm{v}_{h}^{b}) and a(𝒖hb,𝒗h1)a(\bm{u}_{h}^{b},\bm{v}_{h}^{1}) are not zero matrices, where 𝒖hb\bm{u}_{h}^{b} and 𝒗hb\bm{v}_{h}^{b} denote the bubble parts. Thus, the number of the non-zero entries in the velocity-velocity region from our method is about half of the Bernardi and Raugel finite element method.

When aDa^{D} or adiva^{\operatorname{div}} is used, the block ARRA_{RR} is diagonal. In this situation, the component 𝒖hR\bm{u}_{h}^{R} can be locally expressed in terms of the discrete pressure php_{h}. In other words, the inverse of ARRA_{RR} is easy to obtain and also diagonal (sparse). It is then easy to eliminate the unknowns corresponding to RT0RT0 via static condensation, which leads to the following system:

𝒜^SD(ULP)=(FLGRARR1FR) with 𝒜^SD=(ALLGLTGLGRARR1GRT).\widehat{\mathcal{A}}_{S}^{D}\begin{pmatrix}U_{L}\\ P\end{pmatrix}=\begin{pmatrix}F_{L}\\ -G_{R}A_{RR}^{-1}F_{R}\end{pmatrix}\quad\text{ with }\widehat{\mathcal{A}}_{S}^{D}=\begin{pmatrix}A_{LL}&-G_{L}^{T}\\ G_{L}&G_{R}A_{RR}^{-1}G_{R}^{T}\end{pmatrix}.

Since 𝒖h\bm{u}_{h} is pressure-robust, 𝒖h1\bm{u}_{h}^{1} is also pressure-robust (see the error estimate section below). In the next section we will prove that although 𝒖h1\bm{u}_{h}^{1} is not divergence-free, it has optimal convergence properties in L2L^{2} and H1H^{1} norms. Thus, the above system can be regarded as a stabilized and pressure-robust 𝑷1cP0{\bm{P}_{1}^{c}}-P0 discretization of the Stokes equations.

Remark 3.

When aR=adiva^{R}=a^{\operatorname{div}}, all the terms in the left-hand side of ​​ 2.16 can be computed out exactly by the lowest order quadrature rule (barycentric quadrature rule). And the system can be reduced to a 𝑷1cP0{\bm{P}_{1}^{c}}-P0 system. Thus, we prefer the bilinear form adiva^{\operatorname{div}} in practice.

Remark 4.

To obtain a divergence-free solution from the above 𝑷1cP0{\bm{P}_{1}^{c}}-P0 system, one can get 𝒖h1\bm{u}_{h}^{1} and php_{h} first, then compute 𝒖hR\bm{u}_{h}^{R} by php_{h}. Since ARRA_{RR} is diagonal, the computation of 𝒖hR\bm{u}_{h}^{R} is very low-cost. Compared to a direct 𝑷1cP0{\bm{P}_{1}^{c}}-P0 discretization (unstable) of the Stokes equations, our method only changes the pressure-pressure region (lower right) of the coefficient matrix and the pressure region of the right-hand side.

3. Error estimates

3.1. Preliminaries

We define the standard Raviart-Thomas interpolation operator ΠhR:VRT0\Pi_{h}^{R}:V\rightarrow RT0 and the piecewise integral mean operator Ph:WWhP_{h}:W\rightarrow{W}_{h} by

((𝒗ΠhR𝒗)𝒏e,1)e=0e,\left(\left(\bm{v}-{\Pi}_{h}^{R}\bm{v}\right)\cdot\bm{n}_{e},1\right)_{e}=0\quad\forall~e\in\mathcal{E},

and

(3.1) (rPhr,q)=0qWh,\left(r-P_{h}r,q\right)=0\quad\forall~q\in{W}_{h},

respectively.

Lemma 2.

The operators ΠhR\Pi_{h}^{R} and PhP_{h} satisfy the following properties:

(3.2) 𝒗ΠhR𝒗TCRhT|𝒗|1,TT𝒯h,\|\bm{v}-\Pi_{h}^{R}\bm{v}\|_{T}\leq C_{R}h_{T}|\bm{v}|_{1,T}\quad\forall~T\in\mathcal{T}_{h},
(3.3) rPhrTCPhTm|r|m,TT𝒯h,m=0,1,\left\|r-P_{h}r\right\|_{T}\leq C_{P}h_{T}^{m}|r|_{m,T}\quad\forall~T\in\mathcal{T}_{h},\quad m=0,1,

and

(3.4) ΠhR𝒗=Ph𝒗,\nabla\cdot\Pi_{h}^{R}\bm{v}=P_{h}\nabla\cdot\bm{v},

for any 𝐯V\bm{v}\in V and rWHm(Ω)r\in W\cap H^{m}(\Omega), where CRC_{R} and CPC_{P} depend on the shape regularity of 𝒯h\mathcal{T}_{h}.

Proof.

See [5, §2.5]. ∎

Next, let V(h)=VRT0.V(h)=V\oplus RT0. be a larger space containing VV and VhV_{h}. Recall that V=[H01(Ω)]dV=\left[H_{0}^{1}(\Omega)\right]^{d}. For any 𝒗V(h)\bm{v}\in V(h), we have the unique decomposition 𝒗=𝒗1+𝒗R\bm{v}=\bm{v}^{1}+\bm{v}^{R} with 𝒗1V,𝒗RRT0.\bm{v}^{1}\in{V},\bm{v}^{R}\in RT0. An extension of the definition of ah(,)a_{h}(\cdot,\cdot), ​​ 2.7, to the whole V(h)×V(h)V(h)\times V(h) is needed prior to analysis. We define that, for any 𝒘,𝒗V(h)\bm{w},\bm{v}\in V(h),

(3.5) ah(𝒘,𝒗):=a(𝒘1,𝒗1)+aR(𝒘R,𝒗R).a_{h}(\bm{w},\bm{v}):=a(\bm{w}^{1},\bm{v}^{1})+a^{R}(\bm{w}^{R},\bm{v}^{R}).

Restricted to the finite element space Vh×VhV_{h}\times V_{h}, the definition ​​ 3.5 coincides with the definition ​​ 2.7. And on V×VV\times V we have ah(,)|V×V=a(,).a_{h}(\cdot,\cdot)|_{V\times V}=a(\cdot,\cdot). Thus the definition of ah(,)a_{h}(\cdot,\cdot) is logical and natural.

To investigate the properties of the discretization formula ​​ 2.6, we introduce a norm |||||||||\cdot||| for the set V(h)V(h) as follows,

(3.6) |𝒗|2=ah(𝒗,𝒗)𝒗V(h).|||\bm{v}|||^{2}=a_{h}(\bm{v},\bm{v})\quad\forall~\bm{v}\in V(h).

In this definition, for any 𝒗1V\bm{v}^{1}\in{V}, we have |𝒗1|=|𝒗1|1.|||\bm{v}^{1}|||=|\bm{v}^{1}|_{1}.

For the case that aR=a0a^{R}=a^{0}, it is not difficult to verify that 𝒗𝒗1+𝒗R|𝒗1|1+𝒗RC|𝒗|\|\bm{v}\|\leq\|\bm{v}^{1}\|+\|\bm{v}^{R}\|\leq|\bm{v}^{1}|_{1}+\|\bm{v}^{R}\|\leq C|||\bm{v}||| for any 𝒗V(h)\bm{v}\in V(h). Thus, ​​ 3.6 really defines a norm for aR=a0a^{R}=a^{0}. For the other two choices of aRa^{R}, that is, aDa^{D} and adiva^{\operatorname{div}}, we have the following results.

Lemma 3.

For any 𝐯RRT0\bm{v}^{R}\in RT0, we have

(3.7) Ca0(𝒗R,𝒗R)aD(𝒗R,𝒗R)Ca0(𝒗R,𝒗R),C_{*}a^{0}(\bm{v}^{R},\bm{v}^{R})\leq a^{D}(\bm{v}^{R},\bm{v}^{R})\leq C^{*}a^{0}(\bm{v}^{R},\bm{v}^{R}),

and

(3.8) γadiv(𝒗R,𝒗R)aD(𝒗R,𝒗R)γadiv(𝒗R,𝒗R),\gamma_{*}a^{\operatorname{div}}(\bm{v}^{R},\bm{v}^{R})\leq a^{D}(\bm{v}^{R},\bm{v}^{R})\leq\gamma^{*}a^{\operatorname{div}}(\bm{v}^{R},\bm{v}^{R}),

where the constant CC_{*} only depends on the parameters αT,T𝒯h\alpha_{T},T\in\mathcal{T}_{h}, and CC^{*}, γ\gamma_{*}, γ\gamma^{*} only depend on the shape regularity of 𝒯h\mathcal{T}_{h} and the parameters.

Proof.

For simplicity, we will only prove the lemma in the case that αTα\alpha_{T}\equiv\alpha for all T𝒯hT\in\mathcal{T}_{h}, where α\alpha is an arbitrary but fixed positive constant. Let us prove the first inequality in ​​ 3.7. Applying the same technique in [45, Lemma 4.2], by the Cauchy-Schwarz inequality we can similarly obtain

(3.9) (𝒗R,𝒗R)T(d+1)e0Tve2(𝚽e,𝚽e)TT𝒯h,\left(\bm{v}^{R},\bm{v}^{R}\right)_{T}\leq(d+1)\sum_{e\in\mathcal{E}^{0}\cap\partial T}v_{e}^{2}(\bm{\Phi}_{e},\bm{\Phi}_{e})_{T}\quad\forall~T\in\mathcal{T}_{h},

where vev_{e} is the coefficient of 𝚽e\bm{\Phi}_{e} such that 𝒗R=e0ve𝚽e\bm{v}^{R}=\sum_{e\in\mathcal{E}^{0}}v_{e}\bm{\Phi}_{e}. Summation of the above inequality over all the elements T𝒯hT\in\mathcal{T}_{h} gives the first inequality in ​​ 3.7 with C=1/(d+1)C_{*}=1/(d+1).

Let us prove the second inequality of ​​ 3.7. For any basis 𝚽e𝒯\bm{\Phi}_{e}\in\mathcal{RT}, eTe\subset\partial T, since 𝚽e𝒏e\bm{\Phi}_{e}\cdot\bm{n}_{e^{\prime}} vanishes on any face eee^{\prime}\neq e, we have 𝚽e|e𝒏e=(|T|/|e|)𝚽e|T,\bm{\Phi}_{e}|_{e}\cdot\bm{n}_{e}=(|T|/|e|)\nabla\cdot\bm{\Phi}_{e}|_{T}, where |||\cdot| denotes the measure and we use the divergence theorem here. By a series of simple calculations we arrive at

(3.10) 𝚽e𝒏ee2=(|T|/|e|)𝚽eT2.\|\bm{\Phi}_{e}\cdot\bm{n}_{e}\|_{e}^{2}=(|T|/|e|)\|\nabla\cdot\bm{\Phi}_{e}\|_{T}^{2}.

From [5, Eqs. 2.1.75&2.1.78] and scaling arguments one can obtain

(3.11) 𝚽eTChT𝚽eT.||\bm{\Phi}_{e}||_{T}\leq Ch_{T}||\nabla\cdot\bm{\Phi}_{e}||_{T}.

Since we assume the partition is shape-regular, the combination of ​​ 3.10 and ​​ 3.11 gives that 𝚽eTChe1/2𝚽e𝒏ee||\bm{\Phi}_{e}||_{T}\leq Ch_{e}^{1/2}||\bm{\Phi}_{e}\cdot\bm{n}_{e}||_{e}. Then for the elements T1T_{1} and T2T_{2} which share the face ee, we have

ve2𝚽eTi2Cheve2𝚽e𝒏ee2=Cheve𝚽e𝒏ee2=Che𝒗R𝒏ee2C𝒗RTi,v_{e}^{2}||\bm{\Phi}_{e}||_{T_{i}}^{2}\leq Ch_{e}v_{e}^{2}||\bm{\Phi}_{e}\cdot\bm{n}_{e}||_{e}^{2}=Ch_{e}||v_{e}\bm{\Phi}_{e}\cdot\bm{n}_{e}||_{e}^{2}=Ch_{e}||\bm{v}^{R}\cdot\bm{n}_{e}||_{e}^{2}\leq C||\bm{v}^{R}||_{T_{i}},

for i=1,2i=1,2, where in the last inequality we use the trace inequality. Sum over all the elements and apply the fact that the mesh is shape-regular, then the second inequality in ​​ 3.7 can be proven.

Finally, the two inequalities in ​​ 3.8 follow immediately from an inverse inequality and the summation of ​​ 3.11 over all the elements, respectively. Thus we complete the proof. ∎

3.2. Approximation

In this subsection, we will construct an interpolation operator Πh:V[C0(Ω¯)]dVh\Pi_{h}:V\cap[C^{0}(\bar{\Omega})]^{d}\rightarrow V_{h} which satisfies the following properties

(3.12) Πh𝒗=Ph𝒗\nabla\cdot\Pi_{h}\bm{v}=P_{h}\nabla\cdot\bm{v}
(3.13) 𝒗Πh𝒗T+hT|𝒗Πh𝒗|TCIhT2|𝒗|2,T,||\bm{v}-\Pi_{h}\bm{v}||_{T}+h_{T}|||\bm{v}-\Pi_{h}\bm{v}|||_{T}\leq{C_{I}h_{T}^{2}|\bm{v}|_{2,T},}

for all T𝒯h,𝒗[H2(Ω)]dT\in\mathcal{T}_{h},\bm{v}\in[H^{2}({\Omega})]^{d}. Here ||||||T|||\cdot|||_{T} is defined as

|𝒗|T2=ahT(𝒗,𝒗),|||\bm{v}|||_{T}^{2}=a_{h}^{T}(\bm{v},\bm{v}),

where ahT(,)a_{h}^{T}(\cdot,\cdot) is defined in the natural sense such that ah(,)=T𝒯hahT(,)a_{h}(\cdot,\cdot)=\sum_{T\in\mathcal{T}_{h}}a_{h}^{T}(\cdot,\cdot).

To this end, denote by Πh1:V[C0(Ω¯)]dVh1\Pi_{h}^{1}:V\cap[C^{0}(\bar{\Omega})]^{d}\rightarrow V_{h}^{1} the usual nodal interpolation and we define Πh\Pi_{h} as

(3.14) Πh𝒗=Πh1𝒗+ΠhR(𝒗Πh1𝒗)𝒗V[C0(Ω¯)]d.\Pi_{h}\bm{v}=\Pi_{h}^{1}\bm{v}+\Pi_{h}^{R}(\bm{v}-\Pi_{h}^{1}\bm{v})\quad\forall~\bm{v}\in V\cap[C^{0}(\bar{\Omega})]^{d}.

For the properties of ΠhR\Pi_{h}^{R}, see Lemma 2. We have the following lemma:

Lemma 4.

The interpolation operator Πh:V[C0(Ω¯)]dVh\Pi_{h}:V\cap[C^{0}(\bar{\Omega})]^{d}\rightarrow V_{h} defined in ​​ 3.14 satisfies the properties ​​ 3.12 and ​​ 3.13.

Proof.

Since Vh1Wh,\nabla\cdot V_{h}^{1}\subset W_{h}, we have

(3.15) Ph𝒗h1=𝒗h1𝒗h1Vh1.P_{h}\nabla\cdot\bm{v}_{h}^{1}=\nabla\cdot\bm{v}_{h}^{1}\quad\forall~\bm{v}_{h}^{1}\in V_{h}^{1}.

Eqs. 3.4 and 3.15 imply

Πh𝒗=Πh1𝒗+ΠhR(𝒗Πh1𝒗)=Πh1𝒗+Ph(𝒗Πh1𝒗)=Ph𝒗,\nabla\cdot\Pi_{h}\bm{v}=\nabla\cdot\Pi_{h}^{1}\bm{v}+\nabla\cdot\Pi_{h}^{R}(\bm{v}-\Pi_{h}^{1}\bm{v})=\nabla\cdot\Pi_{h}^{1}\bm{v}+P_{h}\nabla\cdot(\bm{v}-\Pi_{h}^{1}\bm{v})=P_{h}\nabla\cdot\bm{v},

Thus ​​ 3.12 has been proved.

Let us prove ​​ 3.13. For simplicity, we will only prove it for a0(,)a^{0}(\cdot,\cdot). From the approximation theory we know

(3.16) 𝒗Πh1𝒗T+hT|𝒗Πh1𝒗|1,TCI1hT2|𝒗|2,TT𝒯h,\|\bm{v}-\Pi_{h}^{1}\bm{v}\|_{T}+h_{T}|\bm{v}-\Pi_{h}^{1}\bm{v}|_{1,T}\leq{C_{I_{1}}h_{T}^{2}|\bm{v}|_{2,T}}\quad\forall~T\in\mathcal{T}_{h},

which, together with ​​ 3.2, gives

(3.17) ΠhR(𝒗Πh1𝒗)TCRCI1hT2|𝒗|2,T.\|\Pi_{h}^{R}(\bm{v}-\Pi_{h}^{1}\bm{v})\|_{T}\leq{C_{R}C_{I_{1}}h_{T}^{2}|\bm{v}|_{2,T}.}

Then ​​ 3.13 follows immediately from ​​ 3.16 and ​​ 3.17, together with the triangle inequality. ∎

3.3. Stability and boundedness

From ​​ 3.6, the coercivity and the boundedness of ah(,)a_{h}(\cdot,\cdot) are trivial. We just state the following lemmas without proof.

Lemma 5 (Coercivity).

For any 𝐯V(h)\bm{v}\in V(h) we have

(3.18) ah(𝒗,𝒗)=|𝒗|2.a_{h}(\bm{v},\bm{v})=|||\bm{v}|||^{2}.
Lemma 6 (Boundedness).

For any 𝐰,𝐯V(h)\bm{w},\bm{v}\in V(h) we have

(3.19) ah(𝒗,𝒘)|𝒗||𝒘|.a_{h}(\bm{v},\bm{w})\leq|||\bm{v}||||||\bm{w}|||.

Next, let us prove the discrete inf-sup condition here. Similarly to the Bernardi and Raugel element, here we can not use the interpolation operator Πh\Pi_{h} to prove the inf-sup conditions since it is not well-defined on [H1(Ω)]d[H^{1}(\Omega)]^{d}. However, using the similar technique in [4] and Lemma 4, we can construct an interpolation πh:VVh\mathcal{\pi}_{h}:V\rightarrow V_{h} which satisfies

(3.20) πh𝒗=Ph𝒗𝒗V,\nabla\cdot\mathcal{\pi}_{h}\bm{v}=P_{h}\nabla\cdot\bm{v}\quad\forall~\bm{v}\in V,

and

(3.21) |πh𝒗|CF𝒗1𝒗V.|||\mathcal{\pi}_{h}\bm{v}|||\leq C_{F}||\bm{v}||_{1}\quad\forall~\bm{v}\in V.

For more details, we refer the readers to [4] and [20, pp. 136-138]. Based on this interpolation, we will prove the following inf-sup conditions.

Lemma 7 (LBB Stability).

There exists a positive constant β1,\beta_{1}, independent of h,h, such that

(3.22) sup𝒗hVh{𝟎}b(𝒗h,qh)|𝒗h|β1qhqhWh.\sup_{\bm{v}_{h}\in V_{h}\setminus\{\bm{0}\}}\frac{b(\bm{v}_{h},q_{h})}{|||\bm{v}_{h}|||}\geq\beta_{1}\|q_{h}\|\quad\forall~q_{h}\in W_{h}.
Proof.

We first use the operator πh\mathcal{\pi}_{h} to obtain

(3.23) sup𝒗hVh{𝟎}b(𝒗h,qh)|𝒗h|sup𝒗V{𝟎}b(πh𝒗,qh)|πh𝒗|=sup𝒗V{𝟎}b(𝒗,qh)|πh𝒗|.\sup_{\bm{v}_{h}\in V_{h}\setminus\{\bm{0}\}}\frac{b(\bm{v}_{h},q_{h})}{|||\bm{v}_{h}|||}\geq\sup_{\bm{v}\in V\setminus\{\bm{0}\}}\frac{b\left(\mathcal{\pi}_{h}\bm{v},q_{h}\right)}{|||\mathcal{\pi}_{h}\bm{v}|||}=\sup_{\bm{v}\in V\setminus\{\bm{0}\}}\frac{b(\bm{v},q_{h})}{|||\mathcal{\pi}_{h}\bm{v}|||}.

Thus, substituting ​​ 3.21 into the inequality ​​ 3.23 gives

sup𝒗hVh{𝟎}b(𝒗h,qh)|𝒗h|CF1sup𝒗V{𝟎}b(𝒗,qh)𝒗1β1qh,\sup_{\bm{v}_{h}\in V_{h}\setminus\{\bm{0}\}}\frac{b(\bm{v}_{h},q_{h})}{|||\bm{v}_{h}|||}\geq C_{F}^{-1}\sup_{\bm{v}\in V\setminus\{\bm{0}\}}\frac{b(\bm{v},q_{h})}{\|\bm{v}\|_{1}}\geq\beta_{1}\|q_{h}\|,

where we have used the inf-sup condition for the continuous case ​​ 2.3. ∎

3.4. Consistency Error

The continuous solutions (𝒖,p)(\bm{u},p) satisfy

(3.24) ν(Δ𝒖,𝒗)b(𝒗,p)=(𝒇,𝒗)𝒗V(h),b(𝒖,q)=0qW,\begin{split}\nu\left(-\Delta\bm{u},\bm{v}\right)-b\left(\bm{v},p\right)&=\left(\bm{f},\bm{v}\right)\quad\forall~\bm{v}\in V(h),\\ b\left(\bm{u},q\right)&=0\quad\forall~q\in W,\end{split}

which can be rewritten as

(3.25) νah(𝒖,𝒗)b(𝒗,p)=(𝒇,𝒗)ν((Δ𝒖,𝒗)ah(𝒖,𝒗))𝒗V(h),b(𝒖,q)=0qW.\begin{split}\nu a_{h}(\bm{u},\bm{v})-b(\bm{v},p)&=(\bm{f},\bm{v})-\nu\left((-\Delta\bm{u},\bm{v})-a_{h}(\bm{u},\bm{v})\right)\quad\forall~\bm{v}\in V(h),\\ b(\bm{u},q)&=0\quad\forall~q\in W.\end{split}

Then (Δ𝒖,𝒗)ah(𝒖,𝒗)\left(-\Delta\bm{u},\bm{v})-a_{h}(\bm{u},\bm{v}\right) is the consistency error of our method. From the arguments in [48, Sect. 3], one can see, in general, |(Δ𝒖,𝒗)ah(𝒖,𝒗)|0.\left|(-\Delta\bm{u},\bm{v})-a_{h}\left(\bm{u},\bm{v}\right)\right|\neq 0.

Lemma 8 (Consistency).

The following inequality holds:

(3.26) |(Δ𝒖,𝒗)ah(𝒖,𝒗)|(T𝒯hhT2Δ𝒖T2)1/2(T𝒯hhT2𝒗RT2)1/2𝒗V(h).\left|(-\Delta\bm{u},\bm{v})-a_{h}\left(\bm{u},\bm{v}\right)\right|\leq(\sum_{T\in\mathcal{T}_{h}}h_{T}^{2}\|\Delta\bm{u}\|_{T}^{2})^{1/2}(\sum_{T\in\mathcal{T}_{h}}h_{T}^{-2}\left\|\bm{v}^{R}\right\|_{T}^{2})^{1/2}\quad\forall~\bm{v}\in{{V}(h)}.
Proof.

Note that ah(𝒖,𝒗h)=a(𝒖,𝒗1)=(Δ𝒖,𝒗1)a_{h}(\bm{u},\bm{v}_{h})=a(\bm{u},\bm{v}^{1})=(-\Delta\bm{u},\bm{v}^{1}). Thus, we have

(3.27a) |(Δ𝒖,𝒗)ah(𝒖,𝒗)|\displaystyle\left|(-\Delta\bm{u},\bm{v})-a_{h}\left(\bm{u},\bm{v}\right)\right| =|(Δ𝒖,𝒗R)|T𝒯hhTΔ𝒖ThT1𝒗RT\displaystyle=\left|(-\Delta\bm{u},\bm{v}^{R})\right|\leq\sum_{T\in\mathcal{T}_{h}}h_{T}\|\Delta\bm{u}\|_{T}h_{T}^{-1}\left\|\bm{v}^{R}\right\|_{T}
(3.27b) (T𝒯hhT2Δ𝒖T2)1/2(T𝒯hhT2𝒗RT2)1/2.\displaystyle\leq(\sum_{T\in\mathcal{T}_{h}}h_{T}^{2}\|\Delta\bm{u}\|_{T}^{2})^{1/2}(\sum_{T\in\mathcal{T}_{h}}h_{T}^{-2}\left\|\bm{v}^{R}\right\|_{T}^{2})^{1/2}.

3.5. Error estimates

Based on the above results, the error analysis can be done now. In addition, the following error equations are needed:

(3.28a) νah(𝒖𝒖h,𝒗h)b(𝒗h,pph)\displaystyle\nu a_{h}\left(\bm{u}-\bm{u}_{h},\bm{v}_{h}\right)-b\left(\bm{v}_{h},p-p_{h}\right) =νδh(𝒖,𝒗h)𝒗hVh,\displaystyle=-\nu\delta_{h}(\bm{u},\bm{v}_{h})\quad\forall~\bm{v}_{h}\in V_{h},
(3.28b) b(𝒖𝒖h,qh)\displaystyle b\left(\bm{u}-\bm{u}_{h},q_{h}\right) =0qhWh,\displaystyle=0\quad\forall~q_{h}\in W_{h},

where δh:V(h)×V(h)\delta_{h}:V(h)\times V(h)\rightarrow\mathbb{R} is the consistency error defined as

δh(𝒖,𝒗)=(Δ𝒖,𝒗)ah(𝒖,𝒗).\delta_{h}(\bm{u},\bm{v})=(-\Delta\bm{u},\bm{v})-a_{h}\left(\bm{u},\bm{v}\right).

These error equations can be obtained by subtracting ​​ 2.6 from ​​ 3.25. In what follows we shall use CαC_{\alpha} to denote the constant in the upper bound of the consistency error such that

|δh(𝒖,𝒗)|CαhΔ𝒖{aR(𝒗R,𝒗R)}1/2.|\delta_{h}(\bm{u},\bm{v})|\leq C_{\alpha}h||\Delta\bm{u}||\left\{a^{R}(\bm{v}^{R},\bm{v}^{R})\right\}^{1/2}.

It follows from ​​ 3.26 that CαC_{\alpha} only depends on the parameters αT\alpha_{T} when aR=a0a^{R}=a^{0}. Now we are in a position to present an error estimate for the finite element scheme ​​ 2.6. Our attention is mainly paid to the velocity estimates, which are independent of the pressure and the viscosity.

Theorem 3.1.

Let (𝐮,p)(\bm{u},p) be the solution of ​​ 2.1 and (𝐮h,ph)Vh×Wh\left(\bm{u}_{h},p_{h}\right)\in V_{h}\times W_{h} be the solution of ​​ 2.6. Assume that (𝐮,p)[H2(Ω)]d×H1(Ω)(\bm{u},p)\in[H^{2}(\Omega)]^{d}\times H^{1}(\Omega). Then there exists a constant CC independent of hh, ν\nu and β1\beta_{1} such that

(3.29) |𝒖𝒖h|Ch|𝒖|2,phPhpνβ1Ch|𝒖|2,|||\bm{u}-\bm{u}_{h}|||\leq Ch|\bm{u}|_{2},\quad\|p_{h}-P_{h}p\|\leq\frac{\nu}{\beta_{1}}Ch|\bm{u}|_{2},

and

(3.30) pphCh(νβ1|𝒖|2+|p|1).\left\|p-p_{h}\right\|\leq Ch\left(\frac{\nu}{\beta_{1}}|\bm{u}|_{2}+\left|p\right|_{1}\right).

Here β1\beta_{1} is the coefficient of the discrete inf-sup condition.

Proof.

Introduce the notation:

ξh=𝒖hΠh𝒖,ηh=phPhp,ξ=𝒖Πh𝒖,η=pPhp.\xi_{h}=\bm{u}_{h}-\Pi_{h}\bm{u},\quad\eta_{h}=p_{h}-P_{h}p,\quad\xi=\bm{u}-\Pi_{h}\bm{u},\quad\eta=p-P_{h}p.

We note that ​​ 3.12 implies Πh𝒖=0\nabla\cdot\Pi_{h}\bm{u}=0, which, together with ​​ 2.12 gives that

(3.31) ξh=0.\nabla\cdot\xi_{h}=0.

It follows from the error equations ​​ 3.28 that

(3.32a) νah(ξh,𝒗h)b(𝒗h,ηh)\displaystyle\nu a_{h}\left(\xi_{h},\bm{v}_{h}\right)-b\left(\bm{v}_{h},\eta_{h}\right) =νah(ξ,𝒗h)b(𝒗h,η)+νδh(𝒖,𝒗h)𝒗hVh,\displaystyle=\nu a_{h}(\xi,\bm{v}_{h})-b(\bm{v}_{h},\eta)+\nu\delta_{h}(\bm{u},\bm{v}_{h})\quad\forall~\bm{v}_{h}\in V_{h},
(3.32b) b(ξh,qh)\displaystyle b\left(\xi_{h},q_{h}\right) =b(ξ,qh)=0qhWh.\displaystyle=b(\xi,q_{h})=0\quad\forall~q_{h}\in W_{h}.

By taking 𝒗h=ξh\bm{v}_{h}=\xi_{h} in ​​ 3.32a and qh=ηhq_{h}=\eta_{h} in ​​ 3.32b, the sum of ​​ 3.32a and ​​ 3.32b together with ​​ 3.31 gives

(3.33) ah(ξh,ξh)=ah(ξ,ξh)+δh(𝒖,ξh),a_{h}\left(\xi_{h},\xi_{h}\right)=a_{h}\left(\xi,\xi_{h}\right)+\delta_{h}(\bm{u},\xi_{h}),

where ν\nu has been eliminated. Thus, it follows from the coercivity ​​ 3.18, the boundedness ​​ 3.19 and the consistency error ​​ 3.26 that

|ξh|2|ξ|×|ξh|+Cαh|𝒖|2|ξh|,|||\xi_{h}|||^{2}\leq|||\xi|||\times|||\xi_{h}|||+C_{\alpha}h|\bm{u}|_{2}|||\xi_{h}|||,

which implies

|ξh||ξ|+Cαh|𝒖|2.|||\xi_{h}|||\leq|||\xi|||+C_{\alpha}h|\bm{u}|_{2}.

The above estimate is equivalent to

|𝒖hΠh𝒖||𝒖Πh𝒖|+Cαh|𝒖|2.|||\bm{u}_{h}-\Pi_{h}\bm{u}|||\leq|||\bm{u}-\Pi_{h}\bm{u}|||+C_{\alpha}h|\bm{u}|_{2}.

By taking s=2s=2 in ​​ 3.13, a combination of the triangle inequality and the above error estimate gives

(3.34) |𝒖𝒖h|2|𝒖Πh𝒖|+Cαh|𝒖|2(2CI+Cα)h|𝒖|2.|||\bm{u}-\bm{u}_{h}|||\leq 2|||\bm{u}-\Pi_{h}\bm{u}|||+C_{\alpha}h|\bm{u}|_{2}\leq\left(2C_{I}+C_{\alpha}\right)h|\bm{u}|_{2}.

Let us estimate the pressure error. Since Vh=Wh\nabla\cdot V_{h}=W_{h}, it follows from ​​ 3.1 that b(𝒗h,pPhp)=0b(\bm{v}_{h},p-P_{h}p)=0 for all 𝒗hVh,\bm{v}_{h}\in V_{h}, which, together with the discrete inf-sup condition ​​ 3.22 gives

phPhp\displaystyle\left\|p_{h}-P_{h}p\right\| 1β1sup𝒗hVh{𝟎}b(𝒗h,phPhp)|𝒗h|=1β1sup𝒗hVh{𝟎}b(𝒗h,php)|𝒗h|\displaystyle\leq\frac{1}{\beta_{1}}\sup_{\bm{v}_{h}\in V_{h}\setminus\{\bm{0}\}}\frac{b\left(\bm{v}_{h},p_{h}-P_{h}p\right)}{|||\bm{v}_{h}|||}=\frac{1}{\beta_{1}}\sup_{\bm{v}_{h}\in V_{h}\setminus\{\bm{0}\}}\frac{b\left(\bm{v}_{h},p_{h}-p\right)}{|||\bm{v}_{h}|||}
=1β1sup𝒗hVh{𝟎}νah(𝒖𝒖h,𝒗h)νδh(𝒖,𝒗h)|𝒗h|\displaystyle=\frac{1}{\beta_{1}}\sup_{\bm{v}_{h}\in V_{h}\setminus\{\bm{0}\}}\frac{-\nu a_{h}\left(\bm{u}-\bm{u}_{h},\bm{v}_{h}\right)-\nu\delta_{h}(\bm{u},\bm{v}_{h})}{|||\bm{v}_{h}|||}
νβ1sup𝒗hVh{𝟎}1|𝒗h||𝒗h|(|𝒖𝒖h|+Cαh|𝒖|2)νβ1(|𝒖𝒖h|+Cαh|𝒖|2).\displaystyle\leq\frac{\nu}{\beta_{1}}\sup_{\bm{v}_{h}\in V_{h}\setminus\{\bm{0}\}}\frac{1}{|||\bm{v}_{h}|||}|||\bm{v}_{h}|||\left(|||\bm{u}-\bm{u}_{h}|||+C_{\alpha}h|\bm{u}|_{2}\right)\leq\frac{\nu}{\beta_{1}}\left(|||\bm{u}-\bm{u}_{h}|||+C_{\alpha}h|\bm{u}|_{2}\right).

Now using the above estimate and the estimate ​​ 3.34 we get

phPhp2νβ1(CI+Cα)h|𝒖|2.\left\|p_{h}-P_{h}p\right\|\leq 2\frac{\nu}{\beta_{1}}\left(C_{I}+C_{\alpha}\right)h|\bm{u}|_{2}.

Thus we complete the proof of ​​ 3.29. Then the error estimate ​​ 3.30 follows from the standard triangle inequality, the above estimate and taking m=1m=1 in ​​ 3.3. ∎

Remark 5.

Classical mixed methods which do not satisfy divergence-free (or pressure-robust) property can only obtain an abstract estimate as

(𝒖𝒖hcla)L2(Ω)C(inf𝒗hVhcla(𝒖𝒗h)L2(Ω)+ν1infqhWhclapqhL2(Ω)).\left\|\nabla\left(\bm{u}-\bm{u}_{h}^{cla}\right)\right\|_{L^{2}(\Omega)}\leq C\left(\inf_{{\bm{v}}_{h}\in V_{h}^{cla}}\left\|\nabla\left(\bm{u}-{\bm{v}}_{h}\right)\right\|_{L^{2}(\Omega)}+{\nu}^{-1}\inf_{q_{h}\in W_{h}^{cla}}\left\|p-q_{h}\right\|_{L^{2}(\Omega)}\right).

See [25, eq. 3.5]. In classical mixed methods, when ν\nu is small, or when the best approximation error of the pressure is large, the error bound of the velocity may be large.

To derive an optimal order L2L^{2}-error estimate for the velocity approximation, we seek (𝒛,λ)(\bm{z},\lambda) satisfying

Δ𝒛+λ=𝒖𝒖h,𝒛=0 in Ω,𝒛=𝟎 on Ω,-\Delta\bm{z}+\nabla\lambda=\bm{u}-\bm{u}_{h},\ \nabla\cdot\bm{z}=0\text{ in }\Omega,\quad\bm{z}=\bm{0}\text{ on }\partial\Omega,

which is also a Stokes system with the viscosity equal to 1.

Here we assume that Ω\Omega is convex, and since 𝒖𝒖h[L2(Ω)]d\bm{u}-\bm{u}_{h}\in\left[L^{2}\left(\Omega\right)\right]^{d}, we have (𝒛,λ)[H2(Ω)]d×H1(Ω)(\bm{z},\lambda)\in[H^{2}(\Omega)]^{d}\times H^{1}(\Omega) and the following regularity estimate holds true (see [14, 19]):

(3.35) 𝒛2+λ1C𝒖𝒖h.\|\bm{z}\|_{2}+\|\lambda\|_{1}\leq C\left\|\bm{u}-\bm{u}_{h}\right\|.

Define δh:V(h)×V(h)\delta_{h}^{*}:V(h)\times V(h)\rightarrow\mathbb{R} as δh(𝒗,𝒛)=δh(𝒛,𝒗),\delta_{h}^{*}(\bm{v},\bm{z})=\delta_{h}(\bm{z},\bm{v}), which is the adjoint consistency error [3]. In fact, since Δ-\Delta is a self-adjoint operator and ah(,)a_{h}(\cdot,\cdot) is symmetric, the properties of consistency and adjoint consistency are equivalent.

Similarly to ​​ 3.25, note that ah(,)a_{h}(\cdot,\cdot) is symmetric, then we have

(3.36) ah(𝒗,𝒛)b(𝒗,λ)=(𝒖𝒖h,𝒗)δh(𝒗,𝒛)𝒗V(h).a_{h}(\bm{v},\bm{z})-b(\bm{v},\lambda)=\left(\bm{u}-\bm{u}_{h},\bm{v}\right)-\delta_{h}^{*}(\bm{v},\bm{z})\quad\forall~\bm{v}\in V(h).
Theorem 3.2.

Let (𝐮,p)(\bm{u},p) be the solution of ​​ 2.1 and (𝐮h,ph)Vh×Wh\left(\bm{u}_{h},p_{h}\right)\in V_{h}\times W_{h} be the solution of ​​ 2.6. Assume that (𝐮,p)[H2(Ω)]d×H1(Ω)(\bm{u},p)\in[H^{2}(\Omega)]^{d}\times H^{1}(\Omega). Then there exists a constant CC independent of hh, ν\nu and β1\beta_{1} such that

(3.37) 𝒖𝒖hCh2|𝒖|2.\left\|\bm{u}-\bm{u}_{h}\right\|\leq Ch^{2}|\bm{u}|_{2}.
Proof.

By taking 𝒗=𝒖𝒖h\bm{v}=\bm{u}-\bm{u}_{h} in ​​ 3.36 we get

(3.38) ah(𝒖𝒖h,𝒛)b(𝒖𝒖h,λ)=𝒖𝒖h2δh(𝒖𝒖h,𝒛).a_{h}\left(\bm{u}-\bm{u}_{h},\bm{z}\right)-b\left(\bm{u}-\bm{u}_{h},\lambda\right)=\left\|\bm{u}-\bm{u}_{h}\right\|^{2}-\delta_{h}^{*}(\bm{u}-\bm{u}_{h},\bm{z}).

Eq. 3.28b implies that

(3.39) b(𝒖𝒖h,λ)=b(𝒖𝒖h,λPhλ).b\left(\bm{u}-\bm{u}_{h},\lambda\right)=b\left(\bm{u}-\bm{u}_{h},\lambda-P_{h}\lambda\right).

Meanwhile, Eq. 3.28a and the fact Πh𝒛=0\nabla\cdot\Pi_{h}\bm{z}=0 give

(3.40) ah(𝒖𝒖h,𝒛)\displaystyle a_{h}\left(\bm{u}-\bm{u}_{h},\bm{z}\right) =ah(𝒖𝒖h,𝒛Πh𝒛)+ah(𝒖𝒖h,Πh𝒛)\displaystyle=a_{h}\left(\bm{u}-\bm{u}_{h},\bm{z}-\Pi_{h}\bm{z}\right)+a_{h}\left(\bm{u}-\bm{u}_{h},\Pi_{h}\bm{z}\right)
=ah(𝒖𝒖h,𝒛Πh𝒛)+1νb(Πh𝒛,pph)δh(𝒖,Πh𝒛)\displaystyle=a_{h}\left(\bm{u}-\bm{u}_{h},\bm{z}-\Pi_{h}\bm{z}\right)+\frac{1}{\nu}b\left(\Pi_{h}\bm{z},p-p_{h}\right)-\delta_{h}(\bm{u},{\Pi_{h}\bm{z}})
=ah(𝒖𝒖h,𝒛Πh𝒛)δh(𝒖,Πh𝒛).\displaystyle=a_{h}\left(\bm{u}-\bm{u}_{h},\bm{z}-\Pi_{h}\bm{z}\right)-\delta_{h}(\bm{u},{\Pi_{h}\bm{z}}).

Substituting ​​ 3.39 and ​​ 3.40 into ​​ 3.38 one can obtain

𝒖𝒖h2=ah(𝒖𝒖h,𝒛Πh𝒛)b(𝒖𝒖h,λPhλ)+δh(𝒖𝒖h,𝒛)δh(𝒖,Πh𝒛).\left\|\bm{u}-\bm{u}_{h}\right\|^{2}=a_{h}\left(\bm{u}-\bm{u}_{h},\bm{z}-\Pi_{h}\bm{z}\right)-b\left(\bm{u}-\bm{u}_{h},\lambda-P_{h}\lambda\right)+\delta_{h}^{*}(\bm{u}-\bm{u}_{h},\bm{z})-\delta_{h}(\bm{u},{\Pi_{h}\bm{z}}).

From ​​ 3.14 and ​​ 3.26 we have

|δh(𝒖,Πh𝒛)|(T𝒯hhT2Δ𝒖T2)1/2(T𝒯hhT2ΠhR(𝒛Πh1𝒛)T2)1/2Ch2|𝒖|2𝒛2(see​​ 3.17),|\delta_{h}(\bm{u},\Pi_{h}\bm{z})|\leq(\sum_{T\in\mathcal{T}_{h}}h_{T}^{2}\|\Delta\bm{u}\|_{T}^{2})^{1/2}(\sum_{T\in\mathcal{T}_{h}}h_{T}^{-2}\left\|\Pi_{h}^{R}(\bm{z}-\Pi_{h}^{1}\bm{z})\right\|_{T}^{2})^{1/2}\leq Ch^{2}|\bm{u}|_{2}||\bm{z}||_{2}\quad({\rm see~\lx@cref{creftype~refnum}{rt0part}}),

and |δh(𝒖𝒖h,𝒛)|Cαh𝒛2|𝒖𝒖h|,|\delta_{h}^{*}(\bm{u}-\bm{u}_{h},\bm{z})|\leq C_{\alpha}h||\bm{z}||_{2}|||\bm{u}-\bm{u}_{h}|||, which imply

𝒖𝒖h2C|𝒖𝒖h|(|𝒛Πh𝒛|+λPhλ)+Ch(h|𝒖|2+|𝒖𝒖h|)𝒛2.\left\|\bm{u}-\bm{u}_{h}\right\|^{2}\leq C|||\bm{u}-\bm{u}_{h}|||\left(|||\bm{z}-\Pi_{h}\bm{z}|||+\left\|\lambda-P_{h}\lambda\right\|\right)+Ch(h|\bm{u}|_{2}+|||\bm{u}-\bm{u}_{h}|||)||\bm{z}||_{2}.

It follows from the above estimate, the interpolation error estimates ​​ 3.3 and ​​ 3.13, and the regularity estimate ​​ 3.35 that

𝒖𝒖h2Ch(|𝒖𝒖h|+h|𝒖|2)𝒖𝒖h,\left\|\bm{u}-\bm{u}_{h}\right\|^{2}\leq Ch\left(|||\bm{u}-\bm{u}_{h}|||+h|\bm{u}|_{2}\right)\left\|\bm{u}-\bm{u}_{h}\right\|,

which implies that

𝒖𝒖hCh(|𝒖𝒖h|+h|𝒖|2).\left\|\bm{u}-\bm{u}_{h}\right\|\leq Ch\left(|||\bm{u}-\bm{u}_{h}|||+h|\bm{u}|_{2}\right).

Then the estimate ​​ 3.37 follows immediately from the above inequality and the estimate ​​ 3.29. Thus we complete the proof. ∎

Remark 6.

From Theorem 3.1 and ​​ 3.6 we have (𝒖𝒖h1)Ch|u|2\|\nabla(\bm{u}-\bm{u}_{h}^{1})\|\leq Ch|u|_{2}. Thus, 𝒖h1\bm{u}_{h}^{1} is a conforming approximation of the continuous solution 𝒖\bm{u} with optimal order in H1H^{1} norm. In addition, Theorem 3.1 also imply that 𝒖hRCh2|𝒖|2||\bm{u}_{h}^{R}||\leq Ch^{2}|\bm{u}|_{2}, which, together with Theorem 3.2 and a triangle inequality, demonstrates that 𝒖h1\bm{u}_{h}^{1} is also convergent to 𝒖\bm{u} with the optimal order in L2L^{2} norm.

We summarize the above remark in the following corollary.

Corollary 1.

Let (𝐮,p)(\bm{u},p) be the solution of ​​ 2.1 and (𝐮h,ph)Vh×Wh\left(\bm{u}_{h},p_{h}\right)\in V_{h}\times W_{h} be the solution of ​​ 2.6. Assume that (𝐮,p)[H2(Ω)]d×H1(Ω)(\bm{u},p)\in[H^{2}(\Omega)]^{d}\times H^{1}(\Omega). Then there exists a constant CC independent of hh, ν\nu and β1\beta_{1} such that

𝒖𝒖h1+h(𝒖𝒖h1)Ch2|𝒖|2.\left\|\bm{u}-\bm{u}_{h}^{1}\right\|+h\left\|\nabla(\bm{u}-\bm{u}_{h}^{1})\right\|\leq Ch^{2}|\bm{u}|_{2}.
Remark 7.

Although our method is stable and optimally convergent as long as the parameters αT,T𝒯h\alpha_{T},T\in\mathcal{T}_{h} are positive, here we can not give a set of optimal choices in the sense of the accuracy. This is still an open and interesting topic. Consider the simplest case. Assuming that αTα\alpha_{T}\equiv\alpha and taking aR=a0a^{R}=a^{0} may give some rough answers for this topic. Note that ​​ 3.33, ​​ 3.26 and the definition of |||||||||\cdot||| imply

|ξh|{(𝒖Πh1𝒖)2+αT𝒯hhT2ΠhR(𝒖Πh1𝒖)T2}1/2+{α1T𝒯hhT2||Δ𝒖||T2}1/2.|||\xi_{h}|||\leq\left\{||\nabla(\bm{u}-\Pi_{h}^{1}\bm{u})||^{2}+\alpha\sum_{T\in\mathcal{T}_{h}}h_{T}^{-2}||\Pi_{h}^{R}(\bm{u}-\Pi_{h}^{1}\bm{u})||_{T}^{2}\right\}^{1/2}+\left\{\alpha^{-1}\sum_{T\in\mathcal{T}_{h}}h_{T}^{2}||\Delta\bm{u}||_{T}^{2}\right\}^{1/2}.

Since ΠhR(𝒖Πh1𝒖)TCRCI1hT2|𝒖|2,T||\Pi_{h}^{R}(\bm{u}-\Pi_{h}^{1}\bm{u})||_{T}\leq C_{R}C_{I_{1}}h_{T}^{2}|\bm{u}|_{2,T} (see ​​ 3.17) and Δ𝒖T2d|𝒖|2,T2||\Delta\bm{u}||_{T}^{2}\leq d|\bm{u}|_{2,T}^{2}, by a series of simple calculations we arrive at

|ξh|22(𝒖Πh1𝒖)2+2(αCR2CI12+α1d)T𝒯hhT2|𝒖|2,T2.|||\xi_{h}|||^{2}\leq 2||\nabla(\bm{u}-\Pi_{h}^{1}\bm{u})||^{2}+2(\alpha C_{R}^{2}C_{I_{1}}^{2}+\alpha^{-1}d)\sum_{T\in\mathcal{T}_{h}}h_{T}^{2}|\bm{u}|_{2,T}^{2}.

The above inequality implies that taking α=d/(CRCI1)\alpha=\sqrt{d}/(C_{R}C_{I_{1}}) might be a good choice. This requires a severe estimate of CRC_{R} and CI1C_{I_{1}}, which depends on the shape regularity of the mesh. In the numerical experiments we will study the performance of our method with different parameters on various meshes.

Remark 8.

In Remark 3 we point out that in practice we prefer the bilinear form adiv(,)a^{\operatorname{div}}(\cdot,\cdot). However, in the numerical experiments we find that this bilinear form is a little sensitive to the change of the parameters. To reduce the sensitivity, some strategies could be applied. For example, we can modify the bilinear form ah(,)a_{h}(\cdot,\cdot) to

(3.41) ahm(𝒖,𝒗)=ah(𝒖,𝒗)T𝒯hγT(𝒖1,𝒗1)T,a_{h}^{m}(\bm{u},\bm{v})=a_{h}(\bm{u},\bm{v})-\sum_{T\in\mathcal{T}_{h}}\gamma_{T}(\nabla\cdot\bm{u}^{1},\nabla\cdot\bm{v}^{1})_{T},

where 0γT<αT/(d+1),T𝒯h0\leq\gamma_{T}<\alpha_{T}/(d+1),T\in\mathcal{T}_{h} are the parameters. The analysis of ahm(,)a_{h}^{m}(\cdot,\cdot) is very similar to ah(,)a_{h}(\cdot,\cdot). The main difference is that we can not prove the coercivity of ahm(,)a_{h}^{m}(\cdot,\cdot) on the whole V(h)V(h). However, we can prove that ahm(,)a_{h}^{m}(\cdot,\cdot) is coercive on the subspace of V(h)V(h)

Z(h)={𝒗V(h):b(𝒗,q)=0qW}={𝒗V(h):𝒗=0}.Z(h)=\{\bm{v}\in V(h):b(\bm{v},q)=0\quad\forall\ q\in W\}=\{\bm{v}\in V(h):\nabla\cdot\bm{v}=0\}.

Indeed, for any 𝒗Z(h)\bm{v}\in Z(h), we have 𝒗1=𝒗R\nabla\cdot\bm{v}^{1}=-\nabla\cdot\bm{v}^{R}, which implies that

(3.42a) γT(𝒗1,𝒗1)T\displaystyle\gamma_{T}(\nabla\cdot\bm{v}^{1},\nabla\cdot\bm{v}^{1})_{T} =γT(𝒗R,𝒗R)T(d+1)γTeT0ve2(𝚽e,𝚽e)T\displaystyle=\gamma_{T}(\nabla\cdot\bm{v}^{R},\nabla\cdot\bm{v}^{R})_{T}\leq(d+1)\gamma_{T}\sum_{e\in\partial T\cap\mathcal{E}^{0}}v_{e}^{2}(\nabla\cdot\bm{\Phi}_{e},\nabla\cdot\bm{\Phi}_{e})_{T}
(3.42b) <αTeT0ve2(𝚽e,𝚽e)T.\displaystyle<\alpha_{T}\sum_{e\in\partial T\cap\mathcal{E}^{0}}v_{e}^{2}(\nabla\cdot\bm{\Phi}_{e},\nabla\cdot\bm{\Phi}_{e})_{T}.

Then the coercivity of ahm(,)a_{h}^{m}(\cdot,\cdot) on Z(h)Z(h) follows immediately from the above inequality. Note that Z(h)Z(h) include the finite element space Zh={𝒗hVh:b(𝒗h,qh)=0qhWh}={𝒗hVh:𝒗h=0}.Z_{h}=\{\bm{v}_{h}\in V_{h}:b(\bm{v}_{h},q_{h})=0\quad\forall\ q_{h}\in W_{h}\}=\{\bm{v}_{h}\in V_{h}:\nabla\cdot\bm{v}_{h}=0\}. In addition, the boundedness of the bilinear forms and the inf-sup conditions could be similarly obtained. Then from the classical theory on saddle point systems [6], the existence and uniqueness of the solutions are guaranteed. The error estimate also follows similarly. The numerical experiments below (see section 4.2) demonstrate that this modified scheme does reduce the sensitivity.

4. Numerical experiments

We perform some numerical experiments on four examples. The examples in section 4.1 and section 4.3 are from [25]. In some examples, we also give some results computed by the classical low order Bernardi and Raugel pair [4] to make a comparison.

4.1. Convergence test

The exact solutions are prescribed as follows,

𝒖=200(x2(1x)2y(1y)(12y),x(1x)(12x)y2(1y)2),\bm{u}=200\left(x^{2}(1-x)^{2}y(1-y)(1-2y),-x(1-x)(1-2x)y^{2}(1-y)^{2}\right)^{\top},

and

p=10((x1/2)3y2+(1x)3(y1/2)3).p=10\left(\left(x-1/2\right)^{3}y^{2}+(1-x)^{3}\left(y-1/2\right)^{3}\right).

The velocity field has the form of a large vortex [25]. We set Ω=(0,1)×(0,1)\Omega=(0,1)\times(0,1) and consider a small viscosity case, ν=106\nu=10^{-6}, to show the robustness of our methods. The corresponding results are shown in Table 1, where we also give some results from the Bernardi and Raugel element. For the choice of parameters, we take αT20,T𝒯h\alpha_{T}\equiv 20,T\in\mathcal{T}_{h} for aR=a0a^{R}=a^{0} and aR=aDa^{R}=a^{D}. In the case that aR=adiva^{R}=a^{\operatorname{div}}, we take αT1.5\alpha_{T}\equiv 1.5. One could see, the velocity errors of our methods (“compact H(div) method”) are roughly 10410^{4} times smaller than the Bernardi and Raugel method. Moreover, comparing with the bottom line of Table 1, one can find that the discrete pressures obtained by our methods are almost equal to the best approximation of the continuous pressure in L2(Ω)L^{2}(\Omega).

Table 1. section 4.1. The L2L^{2} and H1H^{1} (seminorm) velocity errors and L2L^{2} pressure errors. CHdiva0: the method ​​ 2.6 with a0a^{0} ​​ 2.8; CHdivaD: the method ​​ 2.6 with aDa^{D} ​​ 2.9; CHdivadiv: the method ​​ 2.6 with adiva^{\operatorname{div}} ​​ 2.10; B-R: the Bernardi and Raugel finite element method [4]. The L2L^{2} projection errors of the pressure are listed at the bottom.
h=0.1h=0.1 0.050.05 0.0250.025 0.01250.0125 0.006250.00625
𝒖𝒖h\left\|\bm{u}-\bm{u}_{h}\right\| CHdiva0{\rm CHdiva0} 1.51E-2 3.87E-3 9.73E-4 2.43E-4 6.09E-5
CHdivaD{\rm CHdivaD} 1.60E-2 4.11E-3 1.03E-3 2.60E-4 6.52E-5
CHdivadiv{\rm CHdivadiv} 1.97E-2 5.05E-3 1.27E-3 3.19E-4 8.16E-5
BR{\rm B-R} 130.45 33.80 8.55 2.14 5.37E-1
(𝒖𝒖h1)\left\|\nabla(\bm{u}-\bm{u}_{h}^{1})\right\| CHdiva0{\rm CHdiva0} 1.12 5.67E-1 2.84E-1 1.42E-1 7.11E-2
CHdivaD{\rm CHdivaD} 1.13 5.68E-1 2.84E-1 1.42E-1 7.11E-2
CHdivadiv{\rm CHdivadiv} 1.13 5.69E-1 2.84E-1 1.42E-1 7.11E-2
BR{\rm B-R} 7257.64 3895.73 2010.36 1020.13 513.69
pph\left\|p-p_{h}\right\| CHdiva0{\rm CHdiva0} 2.52E-2 1.26E-2 6.32E-3 3.16E-3 1.58E-3
CHdivaD{\rm CHdivaD} 2.52E-2 1.26E-2 6.32E-3 3.16E-3 1.58E-3
CHdivadiv{\rm CHdivadiv} 2.52E-2 1.26E-2 6.32E-3 3.16E-3 1.58E-3
BR{\rm B-R} 2.61E-2 1.31E-2 6.55E-3 3.27E-3 1.63E-3
pPhp\left\|p-P_{h}p\right\| 2.52E-2 1.26E-2 6.32E-3 3.16E-3 1.58E-3

4.2. Effects of different values of the parameters αT\alpha_{T}

In this example we focus on the case that aR=adiva^{R}=a^{\operatorname{div}} and investigate the effects of different choices of the parameters on four meshes (see Fig. 3). The exact solutions are the same as section 4.1. We set αTα\alpha_{T}\equiv\alpha with α=0.1:0.1:10\alpha=0.1:0.1:10, that is, we test 100 different values of the parameters which are from 0.1 to 10. The results from the original scheme ​​ 2.6 and the modified scheme ​​ 3.41 are shown in Fig. 4 and Fig. 5, respectively. It is demonstrated that the most stable quantity is 𝒖𝒖h1||\bm{u}-\bm{u}_{h}^{1}|| (L2u1L^{2}u^{1}), which is very close to the nodal interpolation error (L2Π1uL^{2}\Pi^{1}u) in the whole region. The quantities 𝒖𝒖h||\bm{u}-\bm{u}_{h}|| (L2uL^{2}u) and (𝒖𝒖h1)||\nabla(\bm{u}-\bm{u}_{h}^{1})|| (H1u1H^{1}u^{1}) reduce violently before α<1\alpha<1. All the minimum points of 𝒖𝒖h\|\bm{u}-\bm{u}_{h}\| are in (1,5)(1,5). The comparison between the results in Fig. 4 and Fig. 5 shows that the modified scheme ​​ 3.41 does reduce the sensitivity to the change of α\alpha when α>1\alpha>1.

Based on the above observations, we recommend that one choose α>1\alpha>1 in practice.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3. section 4.2. Shown above are four test meshes.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 4. section 4.2. Shown above are plots of the parameter α\alpha versus the natural logarithms of 𝒖𝒖h||\bm{u}-\bm{u}_{h}|| (L2uL^{2}u), (𝒖𝒖h1)||\nabla(\bm{u}-\bm{u}_{h}^{1})|| (H1u1H^{1}u^{1}), 𝒖𝒖h1||\bm{u}-\bm{u}_{h}^{1}|| (L2u1L^{2}u^{1}), 𝒖Πh1𝒖||\bm{u}-\Pi_{h}^{1}\bm{u}|| (L2Π1uL^{2}\Pi^{1}u) and (𝒖Πh1𝒖)||\nabla(\bm{u}-\Pi_{h}^{1}\bm{u})|| (H1Π1uH^{1}\Pi^{1}u) on four meshes with the original scheme ​​ 2.6.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5. section 4.2. Shown above are plots of the parameter α\alpha versus the natural logarithms of 𝒖𝒖h||\bm{u}-\bm{u}_{h}|| (L2uL^{2}u), (𝒖𝒖h1)||\nabla(\bm{u}-\bm{u}_{h}^{1})|| (H1u1H^{1}u^{1}), 𝒖𝒖h1||\bm{u}-\bm{u}_{h}^{1}|| (L2u1L^{2}u^{1}), 𝒖Πh1𝒖||\bm{u}-\Pi_{h}^{1}\bm{u}|| (L2Π1uL^{2}\Pi^{1}u) and (𝒖Πh1𝒖)||\nabla(\bm{u}-\Pi_{h}^{1}\bm{u})|| (H1Π1uH^{1}\Pi^{1}u) on four meshes with the modified scheme ​​ 3.41.

4.3. Stokes flow with Coriolis force

In this example, we consider the flows with strong Coriolis forces which appear in several real-world applications such as meteorology. The simplest model is described as follows (see [34, 25]),

(4.1) νΔ𝒖+p+2𝝎×𝒖=𝒇,𝒖=0inΩ,-\nu\Delta\bm{u}+\nabla p+2\bm{\omega}\times\bm{u}=\bm{f},\quad\nabla\cdot\bm{u}=0\quad\text{in}\ \Omega,

where 𝝎\bm{\omega} is a constant angular velocity vector.

Here 𝒖=(u1,u2)T\bm{u}=(u_{1},u_{2})^{T}. We embed it into three-dimensional space as 𝒖=(u1,u2,0)T\bm{u}=(u_{1},u_{2},0)^{T}, and similarly for other variables. We take 𝝎=(0,0,ω)T\bm{\omega}=(0,0,\omega)^{T} then we get a two-dimensional model since 𝝎×𝒖=ω(u2,u1,0)T.\bm{\omega}\times\bm{u}=\omega(-u_{2},u_{1},0)^{T}. The computational domain Ω\Omega is taken as a L-shaped domain: (0,4)×(0,2)[2,4)×(0,1](0,4)\times(0,2)\setminus[2,4)\times(0,1]. We choose ν=1\nu=1. And the value of ω\omega changes from 11, 100100 to 1000010000. We note that changing the magnitude ω\omega of the Coriolis force will change only the pressure, and not the velocity (see [25, Example 6.5]). The inlet is set at x=0x=0 and outlet is set at x=4x=4. Pure Dirichlet boundary condition is considered, where the volume preserving parabolic inflow and outflow profiles are given by

𝒖in=(y(2y)/2,0),𝒖out=(4(2y)(y1),0),\bm{u}_{in}=\left(y(2-y)/2,0\right)^{\top},\quad\bm{u}_{out}=\left(4(2-y)(y-1),0\right)^{\top},

and homogeneous conditions are used in other parts of the boundary. Here the Coriolis term is discretized by (2𝝎×𝒖h,𝒗h)(2\bm{\omega}\times\bm{u}_{h},\bm{v}_{h}), and the nonhomogeneous boundary condition is imposed by enforcing 𝒖h|Γ=Πh𝒖|Γ\bm{u}_{h}|_{\Gamma}=\Pi_{h}\bm{u}|_{\Gamma}. Of course, we do not know the exact solution 𝒖\bm{u} but the boundary value of Πh𝒖\Pi_{h}\bm{u} can be obtained by the known boundary value of 𝒖\bm{u}. The computational results are shown in Fig. 6.

Refer to captionRefer to caption
Refer to captionRefer to caption
Refer to captionRefer to caption
Refer to captionRefer to caption
Figure 6. section 4.3. Absolute value of the velocity (speed) obtained by the Bernardi and Raugel element [4] (top) and the method ​​ 2.6 with aR=adiva^{R}=a^{\operatorname{div}} (bottom).

One could see, our method is robust with respect to the magnitude ω\omega of the Coriolis force, while undesired instabilities occur for the Bernardi and Raugel element when a strong Coriolis force exists (ω=100,10000\omega=100,10000).

4.4. The robustness of the pressure-robust P1cP0{P_{1}^{c}-P0} scheme

In this example we want to show that the construction method of the 𝑷1cP0{\bm{P}_{1}^{c}-P0} discretization on the Stokes equations could be extended to general incompressible fluid flows. We will use the Brinkman equation and the Coriolis force problem (like section 4.3 but the exact velocity is known) as two examples. The (scaled) Brinkman equation is described as

(4.2) νΔ𝒖+𝒖+p=𝒇,𝒖=0inΩ,-\nu\Delta\bm{u}+\bm{u}+\nabla p=\bm{f},\quad\nabla\cdot\bm{u}=0\quad\text{in}~\Omega,

and the Coriolis force problem is described as ​​ 4.1. In two problems the exact velocity are the same as section 4.1 and we apply the homogenous velocity boundary conditions. In ​​ 4.2 we choose ν=106\nu=10^{-6} and the pressure is also the same as section 4.1. In the Coriolis problem we choose ν=1\nu=1 and ω=104\omega=10^{4} with 𝒇0\bm{f}\equiv 0 (resulting in a very large pressure gradient that balances the Coriolis force). The corresponding discretizations respectively read: (in ah(,)a_{h}(\cdot,\cdot) we choose aR=adiva^{R}=a^{\operatorname{div}} with αT1.5\alpha_{T}\equiv 1.5)

Find(𝒖h,ph)Vh×Whsuchthat\displaystyle{\rm Find}~(\bm{u}_{h},p_{h})\in V_{h}\times W_{h}~{\rm such~that}~~~~~~~~
(4.3a) νah(𝒖h,𝒗h)+aB(𝒖h,𝒗h)b(𝒗h,ph)\displaystyle\nu a_{h}(\bm{u}_{h},\bm{v}_{h})+a^{B}(\bm{u}_{h},\bm{v}_{h})-b(\bm{v}_{h},p_{h}) +b(𝒖h,qh)=(𝒇,𝒗h)𝒗hVh,qhWh,\displaystyle+b(\bm{u}_{h},q_{h})=(\bm{f},\bm{v}_{h})\quad\forall~\bm{v}_{h}\in V_{h},~q_{h}\in W_{h},

and

Find(𝒖h,ph)Vh×Whsuchthat\displaystyle{\rm Find}~(\bm{u}_{h},p_{h})\in V_{h}\times W_{h}~{\rm such~that}~~~~~~~~
(4.4a) νah(𝒖h,𝒗h)+aC(𝒖h,𝒗h)b(𝒗h,ph)\displaystyle\nu a_{h}(\bm{u}_{h},\bm{v}_{h})+a^{C}(\bm{u}_{h},\bm{v}_{h})-b(\bm{v}_{h},p_{h}) +b(𝒖h,qh)=(𝒇,𝒗h)𝒗hVh,qhWh,\displaystyle+b(\bm{u}_{h},q_{h})=(\bm{f},\bm{v}_{h})\quad\forall~\bm{v}_{h}\in V_{h},~q_{h}\in W_{h},

where

aB(𝒖h,𝒗h)=(𝒖h1,𝒗h1)+(𝒖hR,𝒗h1)+(𝒖h1,𝒗hR)+(d+1)T𝒯heT0ueve(𝚽e,𝚽e)T,a^{B}(\bm{u}_{h},\bm{v}_{h})=(\bm{u}_{h}^{1},\bm{v}_{h}^{1})+(\bm{u}_{h}^{R},\bm{v}_{h}^{1})+(\bm{u}_{h}^{1},\bm{v}_{h}^{R})+(d+1)\sum_{T\in\mathcal{T}_{h}}\sum_{e\in\partial T\cap\mathcal{E}^{0}}u_{e}v_{e}(\bm{\Phi}_{e},\bm{\Phi}_{e})_{T},

and

aC(𝒖h,𝒗h)=(2𝝎×𝒖h1,𝒗h1)+(2𝝎×𝒖hR,𝒗h1)+(2𝝎×𝒖h1,𝒗hR).a^{C}(\bm{u}_{h},\bm{v}_{h})=(2\bm{\omega}\times\bm{u}_{h}^{1},\bm{v}_{h}^{1})+(2\bm{\omega}\times\bm{u}_{h}^{R},\bm{v}_{h}^{1})+(2\bm{\omega}\times\bm{u}_{h}^{1},\bm{v}_{h}^{R}).

The final term in aBa^{B} is to preserve the stability since the second term and the third term in aBa^{B} are two indefinite terms. A similar term does not exist in aCa^{C} since one could verify that aCa^{C} itself has a skew-symmetric structure, that is, aC(𝒗h,𝒗h)=0a^{C}(\bm{v}_{h},\bm{v}_{h})=0 for all 𝒗hVh\bm{v}_{h}\in V_{h}. This time the RT0RT0-RT0RT0 regions of the coefficient matrices are both diagonal for the Brinkman and Coriolis force equations. Thus, we can easily obtain the corresponding 𝑷1cP0{\bm{P}_{1}^{c}-P0} schemes via static condensation. Unlike the scheme for the Stokes equations, in these two cases, the 𝒖hR\bm{u}_{h}^{R} parts are expressed in terms of 𝒖h1\bm{u}_{h}^{1} and php_{h} together, not only php_{h}. Some convergence results are shown in Table 2 and Table 3.

Table 2. section 4.4. Errors for the Brinkman equation with ν=106\nu=10^{-6}.
h 𝒖𝒖h||\bm{u}-\bm{u}_{h}|| order (𝒖𝒖h1)||\nabla(\bm{u}-\bm{u}_{h}^{1})|| order pph||p-p_{h}|| order
1E-1 1.23E-2 1.15 2.51E-2
5E-2 2.93E-3 2.07 5.73E-1 1.01 1.26E-2 0.99
2.5E-2 7.10E-4 2.04 2.85E-1 1.00 6.32E-3 0.99
1.25E-2 1.74E-4 2.02 1.42E-1 1.00 3.16E-3 0.99
6.25E-3 4.32E-5 2.01 7.11E-2 1.00 1.58E-3 0.99
Table 3. section 4.4. Errors for the Stokes equation with strong Coriolis force (ω=104\omega=10^{4}).
h 𝒖𝒖h||\bm{u}-\bm{u}_{h}|| order (𝒖𝒖h1)||\nabla(\bm{u}-\bm{u}_{h}^{1})|| order
1E-1 4.27E-2 1.18
5E-2 8.41E-3 2.34 5.71E-1 1.04
2.5E-2 1.37E-3 2.61 2.84E-1 1.00
1.25E-2 3.19E-4 2.10 1.42E-1 1.00
6.25E-3 8.11E-5 1.97 7.11E-2 1.00

5. Conclusions

We have developed a class of divergence-free mixed methods for the Stokes problem based on 𝑷1cRT0P0{\bm{P}_{1}^{c}}\oplus RT0-P0 pair. By putting this pair into a modified conforming-like formulation, our methods combine some advantages of conforming methods and H(div)H(\operatorname{div})-conforming methods. In addition, a stabilized and pressure-robust 𝑷1cP0{\bm{P}_{1}^{c}}-P0 discretization is proposed in this paper which is computationally cheap. Future work includes:

  • Extending our methods to the mixed velocity and stress boundary value problems which are also an important class of problems in real-world applications. We note that, in this extension, a minor modification to our methods is required to preserve the convergence speed.

  • Extending our methods to the Navier-Stokes equations (NSEs). The key issue is the discretization of the nonlinear term in NSE. In this regard, we will propose a modified convective formulation which conserves the energy, linear momentum and angular momentum in some cases. This is based on the construction of our velocity space.

  • We are planning to propose a general framework on constructing stable and pressure-robust 𝑷1cP0{\bm{P}_{1}^{c}}-P0 discretizations for incompressible fluid flows, including the Brinkman equations, the Oseen equations, the Navier-Stokes equations etc. Some related numerical results have been shown in this paper (see section 4.4).

Funding

This work was supported by the National Natural Science Foundation of China grant 12131014.

References

  • [1] Naveed Ahmed, Alexander Linke, and Christian Merdon, On really locking-free mixed finite element methods for the transient incompressible Stokes equations, SIAM J. Numer. Anal. 56 (2018), no. 1, 185–209.
  • [2] D. N. Arnold and J. Qin, Quadratic velocity/linear pressure Stokes elements, Advances in Computer Methods for Partial Differential Equations-VII, R. Vichnevetsky, D. Knight & G. Richter, eds., IMACS, New Brunswick, NJ (1992), 28–34.
  • [3] Douglas N. Arnold, Franco Brezzi, Bernardo Cockburn, and L. Donatella Marini, Unified analysis of discontinuous Galerkin methods for elliptic problems, SIAM J. Numer. Anal. 39 (2002), no. 5, 1749–1779 (en).
  • [4] Christine Bernardi and Genevieve Raugel, Analysis of some finite elements for the Stokes problem, Math. Comp. 44 (1985), no. 169, 71–79.
  • [5] Daniele Boffi, Franco Brezzi, and Michel Fortin, Mixed finite element methods and applications, Springer Series in Computational Mathematics, vol. 44, Springer Berlin Heidelberg, 2013.
  • [6] F. Brezzi, On the existence, uniqueness and approximation of saddle-point problems arising from lagrangian multipliers, ESAIM: Math. Model. Numer. Anal. 8 (1974), no. R2, 129–151.
  • [7] Franco Brezzi, Jim Douglas, and L. D. Marini, Two families of mixed finite elements for second order elliptic problems, Numer. Math. 47 (1985), no. 2, 217–235 (en).
  • [8] Erik Burman, Snorre H. Christiansen, and Peter Hansbo, Application of a minimal compatible element to incompressible and nearly incompressible continuum mechanics, Comput. Methods Appl. Mech. Engrg. 369 (2020), 113224.
  • [9] Snorre H. Christiansen and Kaibo Hu, Generalized finite element systems for smooth differential forms and Stokes’ problem, Numer. Math. 140 (2018), no. 2, 327–371.
  • [10] Philippe G. Ciarlet, The finite element method for elliptic problems, North-Holland, New York, 1978.
  • [11] Bernardo Cockburn, Jayadeep Gopalakrishnan, and Raytcho Lazarov, Unified hybridization of discontinuous galerkin, mixed, and continuous galerkin methods for second order elliptic problems, SIAM J. Numer. Anal. 47 (2009), no. 2, 1319–1365.
  • [12] Bernardo Cockburn, Guido Kanschat, and Dominik Scho¨{\rm\ddot{o}}tzau, A note on discontinuous Galerkin divergence-free solutions of the Navier-Stokes equations, J. Sci. Comput. 31 (2007), no. 1-2, 61–73.
  • [13] Crouzeix, M. and Raviart, P.-A., Conforming and nonconforming finite element methods for solving the stationary Stokes equations I, R.A.I.R.O. 7 (1973), 33–75.
  • [14] Monique Dauge, Stationary Stokes and Navier-Stokes systems on two- or three-dimensional domains with corners. Part I. linearized equations, SIAM J. Math. Anal. 20 (1989), no. 1, 74–97.
  • [15] John A. Evans and Thomas J. R. Hughes, Isogeometric divergence-conforming B-splines for the steady Navier-Stokes equations, Math. Models Methods Appl. Sci. 23 (2013), no. 08, 1421–1478.
  • [16] by same author, Isogeometric divergence-conforming B-splines for the unsteady Navier-Stokes equations, J. Comput. Phys. 241 (2013), 141–167.
  • [17] Keith J. Galvin, Alexander Linke, Leo G. Rebholz, and Nicholas E. Wilson, Stabilizing poor mass conservation in incompressible flow problems with large irrotational forcing and application to thermal convection, Comput. Methods Appl. Mech. Engrg. 237-240 (2012), 166–176.
  • [18] Nicolas R. Gauger, Alexander Linke, and Phillip W. Schroeder, On high-order pressure-robust space discretisations, their advantages for incompressible high Reynolds number generalised Beltrami flows and beyond, SMAI J. Comput. Math. 5 (2019), 89–129.
  • [19] Vivette Girault and Jacques-Louis Lions, Two-grid finite-element schemes for the transient Navier-Stokes problem, ESAIM Math. Model. Numer. Anal. 35 (2001), no. 5, 945–980.
  • [20] Vivette Girault and Pierre-Arnaud Raviart, Finite element methods for Navier-Stokes equations, Springer Series in Computational Mathematics, vol. 5, Springer Berlin Heidelberg, Berlin, Heidelberg, 1986.
  • [21] J. Guzmán and M. Neilan, Conforming and divergence-free Stokes elements in three dimensions, IMA J. Numer. Anal. 34 (2014), 1489–1508.
  • [22] Johnny Guzmán and Michael Neilan, A family of nonconforming elements for the Brinkman problem, IMA J. Numer. Anal. 32 (2012), no. 4, 1484–1508.
  • [23] by same author, Inf-sup stable finite elements on barycentric refinements producing divergence–free approximations in arbitrary dimensions, SIAM J. Numer. Anal. 56 (2018), no. 5, 2826–2844.
  • [24] Volker John, Petr Knobloch, and Julia Novo, Finite elements for scalar convection-dominated equations and incompressible flow problems: a never ending story?, Comput. Visual Sci. 19 (2018), no. 5-6, 47–63.
  • [25] Volker John, Alexander Linke, Christian Merdon, Michael Neilan, and Leo G. Rebholz, On the divergence constraint in mixed finite element methods for incompressible flows, SIAM Rev. 59 (2017), no. 3, 492–544.
  • [26] Guido Kanschat and Natasha Sharma, Divergence-conforming discontinuous Galerkin methods and c0c^{0} interior penalty methods, SIAM J. Numer. Anal. 52 (2014), no. 4, 1822–1842.
  • [27] Juho Könnö and Rolf Stenberg, H(div)H(\operatorname{div})-conforming finite elements for the Binkman problem, Math. Models Methods Appl. Sci. 21 (2011), no. 11, 2227–2248.
  • [28] Philip L. Lederer, Christoph Lehrenfeld, and Joachim Schöberl, Hybrid discontinuous Galerkin methods with relaxed H(div)-conformity for incompressible flows. Part I, SIAM J. Numer. Anal. 56 (2018), no. 4, 2070–2094.
  • [29] by same author, Hybrid discontinuous Galerkin methods with relaxed H(div)-conformity for incompressible flows. Part II, ESAIM: Math. Model. Numer. Anal. 53 (2019), no. 2, 503–522.
  • [30] Philip L. Lederer, Alexander Linke, Christian Merdon, and Joachim Schöberl, Divergence-free reconstruction operators for pressure-robust Stokes discretizations with continuous pressure finite elements, SIAM J. Numer. Anal. 55 (2017), no. 3, 1291–1314.
  • [31] Philip L. Lederer and Sander Rhebergen, A pressure-robust embedded discontinuous Galerkin method for the Stokes problem by reconstruction operators, SIAM J. Numer. Anal. 58 (2020), no. 5, 2915–2933.
  • [32] Christoph Lehrenfeld, Hybrid discontinuous galerkin methods for solving incompressible flow problems, Master’s thesis, RWTH Aachen University, 2010.
  • [33] Christoph Lehrenfeld and Joachim Schöberl, High order exactly divergence-free hybrid discontinuous Galerkin methods for unsteady incompressible flows, Comput. Methods Appl. Mech. Engrg. 307 (2016), 339–361.
  • [34] A. Linke and C. Merdon, On velocity errors due to irrotational forces in the Navier-Stokes momentum balance, J. Comput. Phys. 313 (2016), 654–661.
  • [35] by same author, Pressure-robustness and discrete Helmholtz projectors in mixed finite element methods for the incompressible Navier-Stokes equations, Comput. Methods Appl. Mech. Engrg. 311 (2016), 304–326.
  • [36] Alexander Linke, Collision in a cross-shaped domain-A steady 2d Navier-Stokes example demonstrating the importance of mass conservation in CFD, Comput. Methods Appl. Mech. Engrg. 198 (2009), no. 41, 3278–3286.
  • [37] Alexander Linke, On the role of the Helmholtz decomposition in mixed methods for incompressible flows and a new variational crime, Comput. Methods Appl. Mech. Engrg. 268 (2014), 782–800.
  • [38] Alexander Linke, Gunar Matthies, and Lutz Tobiska, Robust arbitrary order mixed finite element methods for the incompressible Stokes equations with pressure independent velocity errors, ESAIM: M2AN 50 (2016), no. 1, 289–309.
  • [39] Kent Andre Mardal, Xue-Cheng Tai, and Ragnar Winther, A robust finite element method for Darcy–Stokes flow, SIAM J. Numer. Anal. 40 (2002), no. 5, 1605–1631.
  • [40] J. C. Nédélec, Mixed finite elements in R3, Numer. Math. 35 (1980), 315–341.
  • [41] Phillip Joseph Phillips and Mary F. Wheeler, A coupling of mixed and continuous Galerkin finite element methods for poroelasticity I: the continuous in time case, Comput. Geosci. 11 (2007), no. 2, 131–144.
  • [42] P. A. Raviart and J. M. Thomas, A mixed finite element method for 2nd order elliptic problems, in Mathematical Aspects of Finite Element Methods, I. Galligani, E. Magenes, eds., Lectures Notes in Math. 606 (1977), 292–315.
  • [43] Sander Rhebergen and Garth N. Wells, Analysis of a hybridized/interface stabilized finite element method for the Stokes equations, SIAM J. Numer. Anal. 55 (2017), no. 4, 1982–2003.
  • [44] Sander Rhebergen and Garth N. Wells, An embedded-hybridized discontinuous Galerkin finite element method for the Stokes equations, Comput. Methods Appl. Mech. Engrg. 358 (2020), 112619.
  • [45] C. Rodrigo, X. Hu, P. Ohm, J.H. Adler, F.J. Gaspar, and L.T. Zikatanov, New stabilized discretizations for poroelasticity and the Stokes’ equations, Comput. Methods Appl. Mech. Engrg. 341 (2018), 467–484 (en).
  • [46] L. R. Scott and M. Vogelius, Norm estimates for a maximal right inverse of the divergence operator in spaces of piecewise polynomials, RAIRO Modél. Math. Anal. Numér. 19 (1985), 111–143.
  • [47] X.-C. Tai and R. Winther, A discrete de Rham complex with enhanced smoothness, Calcolo 43 (2006), no. 4, 287–306.
  • [48] Junping Wang and Xiu Ye, New finite element methods in computational fluid dynamics by H(div){H}(\operatorname{div}) elements, SIAM J. Numer. Anal. 45 (2007), no. 3, 1269–1286 (en).
  • [49] Xiaoping Xie, Jinchao Xu, and Guangri Xue, Uniformly-stable finite element methods for Darcy-Stokes-Brinkman models, J. Comput. Math. 26 (2008), no. 003, 437–455.
  • [50] Shangyou Zhang, A new family of stable mixed finite elements for the 3D Stokes equations, Math. Comp. 74 (2005), no. 250, 543–554.