Skip to content
Publicly Available Published by De Gruyter August 5, 2020

Quasi-Optimal and Pressure Robust Discretizations of the Stokes Equations by Moment- and Divergence-Preserving Operators

  • Christian Kreuzer ORCID logo EMAIL logo , Rüdiger Verfürth ORCID logo and Pietro Zanotti ORCID logo

Abstract

We approximate the solution of the Stokes equations by a new quasi-optimal and pressure robust discontinuous Galerkin discretization of arbitrary order. This means quasi-optimality of the velocity error independent of the pressure. Moreover, the discretization is well-defined for any load which is admissible for the continuous problem and it also provides classical quasi-optimal estimates for the sum of velocity and pressure errors. The key design principle is a careful discretization of the load involving a linear operator, which maps discontinuous Galerkin test functions onto conforming ones thereby preserving the discrete divergence and certain moment conditions on faces and elements.

MSC 2010: 65N30; 65N12; 65N15

1 Introduction

This paper is a new contribution to the research programme initiated in [17, 27], which aims at designing quasi-optimal and pressure robust discretizations of the Stokes equations

(1.1) { - μ Δ u + p = f in  Ω , div u = 0 in  Ω , u = 0 on  Ω

for the largest possible class of inf-sup stable pairs of finite element spaces.

To illustrate our results, let V / Q be an inf-sup stable pair and assume that a given discretization produces an approximation ( u ¯ , p ¯ ) V × Q to the solution ( u , p ) of (1.1). Moreover, let 1 be an H 1 -like norm. We say that the given discretization is quasi-optimal when there is a constant C qo 1 such that

(1.2) μ u - u ¯ 1 + p - p ¯ L 2 ( Ω ) C qo ( μ inf v V u - v 1 + inf q Q p - q L 2 ( Ω ) ) .

Analogously, we say that the given discretization is quasi-optimal and pressure robust when there is a constant C qopr 1 such that

(1.3) u - u ¯ 1 C qopr inf v V u - v 1 .

Any discretization fulfilling the above error estimates

  1. is defined for any admissible load f in the weak formulation of (1.1),

  2. inherits the approximation properties of the underlying spaces V and Q, irrespective of the regularity of ( u , p ) and f,

  3. is pressure robust, in the sense that (1.3) implies that large irrotational forces (or, equivalently, large pressure errors) do not affect the velocity error, cf. Remark 6 below.

Whereas the first two properties are desirable in the discretization of any equation, the third one is specific to the (Navier–)Stokes equations. Its importance has been pointed out in [19] and further investigated in various other references, see, e.g., [16].

Most Stokes discretizations based on nonconforming pairs fail to fulfill (1.2). Analogously, most discretizations with other pairs than divergence-free ones fail to fulfill (1.3). Both claims follow from the abstract results in [17, 25]. Indeed, the combination of (1.2) and (1.3) has been available for a long time only for discretizations based on conforming and divergence-free pairs, like the one of Scott and Vogelius [22]. The importance of pressure robustness was observed in [19], where pressure robust schemes are proposed using H ( div ) -conforming maps applied to the test functions. As a trade off, the quasi-optimality was weakened by involving additional consistency errors; compare also with the overview article [16]. Here and in [17, 27], we design quasi-optimal and pressure robust discretizations by devising, in particular, alternative H 1 -conforming maps applied to test functions.

The discretization proposed in [27] uses the first-order nonconforming Crouzeix-Raviart pair and can be written as follows: find u ¯ V and p ¯ Q such that

(1.4) { μ a ( u ¯ , v ) + b ( v , p ¯ ) = f , E v for all  v V , b ( u ¯ , q ) = 0 for all  q Q ,

where the forms a and b are as in the original discretization described in [9]. The operator E maps into continuous piecewise polynomials and preserves the discrete divergence and the averages on the mesh faces. This idea has been generalized in [17] to a wide class of pairs, under the same conditions on E. The only difference is that the form a needs to be augmented with additional terms. In this paper we propose a different approach, which does not require any augmentation of a, at the price of a more involved construction of E.

More precisely, we propose a class of discontinuous Galerkin discretizations of arbitrary order 1 , which differ from the ones in [14] only in the use of an operator E as in (1.4). Here E is required to preserve the discrete divergence and all moments up to the order - 1 on the mesh faces and up to the order - 2 in the mesh elements. The same approach applies also to H div -conforming pairs [7] and with higher-order Crouzeix–Raviart pairs [3, 8], but fails when dealing with pairs involving a reduced integration of the divergence.

The remaining part of this paper is organized as follows. In Section 2 we propose the new discretization and motivate the above-mentioned conditions on E. Section 3 is devoted to the construction of E and to the derivation of the error estimates. Finally, in Section 4 we investigate numerically the proposed discretization in the lowest-order case. We indicate Lebesgue and Sobolev spaces and their norms as usual, see, e.g., [5].

2 Discontinuous Galerkin Discretization of the Stokes Equations

2.1 Stokes Equations

Let Ω d , d { 2 , 3 } , be an open and bounded polyhedron with Lipschitz boundary. The variational formulation of the Stokes equations in Ω, with viscosity μ > 0 , load f H - 1 ( Ω ) := ( H 0 1 ( Ω ) d ) and homogeneous essential boundary conditions, reads as follows: find a velocity u H 0 1 ( Ω ) d and a pressure p L 0 2 ( Ω ) such that

(2.1) { μ Ω u : v - Ω p div v = f , v for all  v H 0 1 ( Ω ) d , Ω q div u = 0 for all  q L 0 2 ( Ω ) .

Here : denotes the euclidean scalar product of d × d tensors and , is the dual pairing of H - 1 ( Ω ) and H 0 1 ( Ω ) d . Note that we look for the pressure p in the space L 0 2 ( Ω ) := { q L 2 ( Ω ) Ω q = 0 } , according to the boundary condition u = 0 on Ω . This problem is well-posed and we have

(2.2) μ u L 2 ( Ω ) + p L 2 ( Ω ) c f H - 1 ( Ω ) ,

where c only depends on the geometry of Ω, see, e.g., [4, Theorem 8.2.1]. Moreover, introducing the kernel of the divergence operator

Z := { z H 0 1 ( Ω ) d div z = 0 } ,

we infer u Z and the a priori estimate

(2.3) μ u L 2 ( Ω ) f | Z Z := sup z Z f , z z L 2 ( Ω ) .

2.2 Meshes and Polynomials

Let be a face-to-face simplicial mesh of Ω. The shape constant γ of is given by

γ := max K h K ρ K ,

where h K is the diameter of a d-simplex K and ρ K is the diameter of the largest ball inscribed in K. We denote by and i the sets collecting all faces and all interior faces of , respectively. The skeleton of is Σ := F F . We let the meshsize h and the normal n be the piecewise constant functions on Σ given by

h | F := diam ( F ) and n | F := n F

for all F . Here n F is a unit normal vector of F, pointing outside Ω if F Ω .

We denote by 𝒟 the broken version of a differential operator 𝒟 , that is

( 𝒟 v ) | K := 𝒟 ( v | K )

for all K and for piecewise smooth v. We indicate by [ [ v ] ] and { { v } } , respectively, the jump and the average of v on the skeleton Σ of . More precisely, for an interior face F i and for x F , we have

[ [ v ] ] | F ( x ) = v | K 1 ( x ) - v | K 2 ( x ) and { { v } } | F ( x ) = v | K 1 ( x ) + v | K 2 ( x ) 2 ,

where K 1 , K 2 are such that F = K 1 K 2 and n points outside K 1 . Note that the sign of [ [ v ] ] depends on the orientation of n, which will however not be significant to our discussion. For boundary faces F i , it holds

[ [ v ] ] | F ( x ) = v | K ( x ) = { { v } } | F ( x ) ,

where K is such that F = K Ω . To alleviate the notation, we write [ [ ] ] and { { } } in place of [ [ ] ] and { { } } .

The spaces ( K ) and ( F ) , 0 , consist of all polynomials of total degree on a d-simplex K and a face F , respectively. For convenience, we set - 1 = { 0 } . The space of broken polynomials on with total degree reads

S 0 := { v : Ω v | K ( K )  for all  K } .

The approximation of the pressure space involved in the Stokes equations (2.1) motivates the use of the one-codimensional subspace

S ^ 0 := S 0 L 0 2 ( Ω ) .

We shall repeatedly make use of the following integration by parts formula:

(2.4) Ω ( div v ) q = - Ω v q + Σ [ [ v ] ] n { { q } } + Σ Ω { { v } } n [ [ q ] ] ,

where v ( H 0 1 ( Ω ) + S 0 ) d and q S - 1 0 , see, e.g., [1, equation (3.6)].

2.3 Discontinuous Galerkin Discretization

We consider a discontinuous Galerkin (dG) discretization of order of the Stokes equations (see, for instance, [14]) that builds on the bilinear forms a dG : ( S 0 ) d × ( S 0 ) d and b dG : ( S 0 ) d × S ^ - 1 0 given by

(2.5) a dG ( w , v ) := Ω w : v - Σ { { w } } n [ [ v ] ] - Σ [ [ w ] ] { { v } } n + Σ η h [ [ w ] ] [ [ v ] ]

and

b dG ( w , q ) := - Ω q div w + Σ [ [ w ] ] n { { q } } ,

where η > 0 is a penalty parameter.

Motivated by the abstract results in [25], we let E dG : ( S 0 ) d H 0 1 ( Ω ) d be a linear operator and consider the following dG discretization of the Stokes equations (2.1): find a discrete velocity u dG ( S 0 ) d and a discrete pressure p dG S ^ - 1 0 such that

(2.6) { μ a dG ( u dG , v ) + b dG ( v , p dG ) = f , E dG v for all  v ( S 0 ) d , b dG ( u dG , q ) = 0 for all  q S ^ - 1 0 .

Introducing the discrete divergence div dG : ( S 0 ) d S ^ - 1 0 through the problem

(2.7) Ω q div dG w = - b dG ( w , q ) , q S ^ - 1 0 ,

for all w ( S 0 ) d , we can equivalently rewrite (2.6) as follows:

{ μ a dG ( u dG , v ) - Ω p dG div dG v = f , E dG v for all  v ( S 0 ) d , Ω q div dG u dG = 0 for all  q S ^ - 1 0 .

This shows that

(2.8) u dG Z dG := { w ( S 0 ) d div dG w = 0 } ,

i.e. the discrete velocity u dG belongs to the kernel of the discrete divergence.

Remark 1 (Alternative Definition of div dG ).

Note that we could equivalently define the discrete divergence as the linear operator div dG : ( S 0 ) d S - 1 0 given by

Ω q div dG w = Ω q div w - Σ [ [ w ] ] n { { q } } for all  q S - 1 0 .

Indeed, testing with q = 1 and integrating by parts as in (2.4), we see that

Ω div dG w = Ω div w - Σ [ [ w ] ] n = 0

for all w ( S 0 ) d . This proves that div dG w S ^ - 1 0 . Then, testing with q S ^ 0 , we retrieve (2.7).

To assess the quality of the discretization (2.6), we introduce the scalar product

( w , v ) dG := Ω w : v + Σ η h [ [ w ] ] [ [ v ] ] , w , v H 0 1 ( Ω ) d + ( S 0 ) d ,

where the penalty parameter η is the same as in (2.5). We measure the velocity error u - u dG in the norm dG induced by ( , ) dG , that is an extension of the norm L 2 ( Ω ) to ( H 0 1 ( Ω ) + S 0 ) d . Since S ^ - 1 0 L 2 ( Ω ) , we measure the pressure error p - p dG in the L 2 -norm.

Remark 2 (Notation for dG Discretization).

The label “ dG ” identifies all objects and quantities that specifically depend on the discretization (2.6). In most (but not all) cases, such objects and quantities depend on the penalty parameter η.

Remark 3 (Comparison with Conforming and Nonconforming Discretizations).

One main advantage of dG discretizations of the Stokes equations is that they are inf-sup stable for all polynomial degrees 1 , irrespective of the underlying mesh, cf. (2.10c) below. Nonconforming discretizations based on the same pressure space, like, e.g., [9, 3, 8], have the disadvantage that the construction of the velocity space is not straight-forward for > 1 , in particular if d = 3 . Discretizations based on the conforming Scott–Vogelius pair also employ a discontinuous pressure space. Such discretizations are known to be inf-sup stable on the Alfeld refinement of any mesh for d , see [13]. Thus, for sufficiently large , the number of degrees of freedom involved in the discretization is approximately ( d + 1 ) times higher than in dG discretizations. For < d , the inf-sup stability of the Scott–Vogelius pair requires even more complicated mesh refinements and the elimination of local spurious pressure modes, see [12].

In what follows, we write C for a positive nondecreasing function of the shape constant γ of . Such function may depend also on other parameters (like Ω, d, or ) but is independent of the viscosity μ and the penalty parameter η. Furthermore, the value of C does not need to be the same at different occurrences. We sometimes abbreviate A C B as A B .

2.4 Stability

The so-called inverse trace inequality [10, Lemma 1.46] implies that there is a constant η ¯ > 0 , depending only on the shape parameter of and the polynomial degree , such that

(2.9) Σ h | { { v } } | 2 η ¯ v L 2 ( Ω ) 2 , v ( S 0 ) d × d .

Hence, simple algebraic manipulations reveal that the form a dG is bounded and coercive. More precisely, we have

(2.10)

(2.10a) a dG ( w , v ) α ¯ dG w dG v dG , α ¯ dG := 1 + η ¯ η ,
(2.10b) a dG ( w , w ) α ¯ dG w dG 2 , α ¯ dG := 1 - η ¯ η ,

for all w , v ( S 0 ) d . Furthermore, the form b dG is inf-sup stable, in that

(2.10c) β dG q L 2 ( Ω ) sup w ( S 0 ) d b dG ( w , q ) w dG , β dG - 1 max { 1 , η } ,

for all q S ^ - 1 0 , see, e.g., [16, Section 4.4]. Note that, without loss of generality, we can assume β dG 1 .

The following discrete counterpart of (2.2) follows from (2.10) and the theory of saddle point problems. The discrete stability constant involves, in particular, the operator norm of E dG

E dG := E dG ( ( S 0 ) d , H 0 1 ( Ω ) d ) .

Lemma 4 (Discrete Well-Posedness and Stability).

Let η ¯ > 0 be as in (2.9) and assume η > η ¯ . The discretization (2.6), with viscosity μ > 0 and load f H - 1 ( Ω ) , is uniquely solvable and its solution ( u dG , p dG ) satisfies the a priori estimates

μ u dG dG 1 α ¯ dG E dG f H - 1 ( Ω ) 𝑎𝑛𝑑 p dG L 2 ( Ω ) 2 α ¯ dG α ¯ dG β dG E dG f H - 1 ( Ω ) .

Proof.

Since ( S 0 ) d is finite-dimensional, the operator E dG is bounded. This implies that the adjoint operator E dG is well-defined and that the load in the first equation of (2.6) is E dG f . Then [4, Theorem 4.2.3] implies that (2.6) is uniquely solvable, as a consequence of (2.10), and yields the a priori estimates

μ u dG dG 1 α ¯ dG E dG f and p dG L 2 ( Ω ) 2 α ¯ dG α ¯ dG β dG E dG f

where E dG f is the norm of the functional E dG f in the dual space of ( S 0 ) d . We conclude by recalling that the operator norm E dG coincides with the one of E dG , see [6, Remark 2.16]. ∎

A discrete counterpart of (2.3) additionally holds, under the assumption that E dG maps discretely divergence-free functions into exactly divergence-free functions. To our best knowledge, the importance of this condition was first pointed out in [19].

Lemma 5 (Stability of the Discrete Velocity).

Under the assumptions of Lemma 4, for any load f H - 1 ( Ω ) the discrete velocity u dG ( S 0 ) d in (2.6) additionally enjoys the a priori estimate

(2.11) μ u dG dG E dG α ¯ dG f | Z Z

if and only if

(2.12) E dG ( Z dG ) Z .

Proof.

Assume first that (2.12) holds. Testing the first equation of (2.6) with the elements of Z dG , we see that u dG solves the reduced problem

μ a dG ( u dG , v ) = f , E dG v for all  v Z dG .

In view of (2.8), we are allowed to set v = u dG and exploit the coercivity (2.10b) of a dG

μ α ¯ dG u dG dG 2 f , E dG u dG .

Then the inclusion E dG u dG Z implies

f , E dG u dG f | Z Z E dG u dG .

We derive the claimed a priori estimate in view of the boundedness of E dG .

Conversely, assume (2.11) holds and there is v Z dG such that div E dG v 0 . Set f := ( div E dG v ) H - 1 ( Ω ) . On the one hand, we have f | Z = 0 , so that (2.11) implies u dG = 0 . On the other hand, the boundedness of a dG and the first equation of (2.6) reveal that

μ α ¯ dG u dG dG v dG f , E dG v = div E dG v L 2 ( Ω ) 2 0 .

This contradiction confirms that E dG maps Z dG into Z whenever (2.11) holds. ∎

Remark 6 (Pressure Robustness).

The a priori estimate (2.3) reveals that the velocity u in the Stokes equations (2.1) solely depends on f | Z . In particular, this entails that u is invariant with respect to irrotational perturbations of f, which only affect the pressure p, see Linke [19]. Whenever the estimate (2.11) holds, the discretization (2.6) reproduces such invariant property and we call it ‘pressure robust’. We refer to [16] and to the references therein for an extensive discussion on the importance of pressure robustness in the discretization of the (Navier–)Stokes equations.

2.5 Quasi-Optimality

We now look for conditions ensuring that the discretization (2.6) enjoys (1.2). To this end, we first investigate the approximation of the velocity field u in (2.1) by Z dG , i.e. by discretely divergence-free velocity fields. This is a standard question motivated by the inclusion (2.8) and several related results are available in the literature. We refer to [5, Theorem 12.5.17] for conforming discretizations and to [23, Lemma 8.1] for dG discretizations.

Lemma 7 (Approximation by Z dG ).

Let u H 0 1 ( Ω ) d be the velocity solving (2.1). Then there is a constant δ dG such that

inf z Z dG u - z dG δ dG inf w ( S 0 ) d u - w dG .

Moreover, it holds δ dG 1 + C β dG - 1 .

Proof.

Let w ( S 0 ) d be given and denote by Z dG the orthogonal complement of Z dG with respect to the scalar product ( , ) dG . Inequality (2.10c) implies that Z dG and S ^ - 1 0 have the same space dimension and

β dG q L 2 ( Ω ) sup w ~ Z dG b dG ( w ~ , q ) w ~ dG ,

cf. [5, Chapter 12.5]. Then the Banach–Nec̆as theorem (see, e.g., [11, Theorem 2.6]) ensures the existence of a unique solution w ~ Z dG to the problem

b dG ( w ~ , q ) = b dG ( w , q ) for all  q S ^ - 1 0 ,

together with the estimate

β dG w ~ dG b dG ( w , ) ( S ^ - 1 0 ) .

Recall that u is in H 0 1 ( Ω ) and divergence-free, in view of the second equation of (2.1). Hence, for all q S ^ - 1 0 , we have

b dG ( w , q ) = Ω div ( u - w ) q - Σ [ [ u - w ] ] n { { q } }
u - w dG q L 2 ( Ω ) ,

where we have used the inverse trace inequality (2.9) for the term involving { { q } } . This estimate and the previous one entail that β dG w ~ dG C u - w dG . Next, we set z := w - w ~ . By definition, we have b dG ( z , ) = 0 , showing that z Z dG . Moreover, it holds

u - z dG u - w dG + w ~ dG ( 1 + C β dG - 1 ) u - w dG .

We conclude taking the infimum over all w ( S 0 ) d . ∎

Remark 8 (Size of δ dG ).

The bound of δ dG in the above lemma is known to be potentially pessimistic if β dG is close to zero as, for instance, in channel-like stretched domains. A sharper bound of δ dG could be obtained in terms of the norm of a Fortin operator by arguing in the spirit of [16, Remark 4.1].

After this preparation, we observe that the dG discretization (2.6) fits into the abstract framework of [17, Section 2]. Hence, applying [17, Lemma 2.6], we derive that the following consistency conditions are necessary for quasi-optimality (1.2)

(2.13)

(2.13a) a dG ( w , v ) = Ω w : E dG v for all  v ( S 0 ) d ,
(2.13b) b dG ( w , q ) = - Ω q div E dG w for all  w ( S 0 ) d , q S ^ - 1 0 .

Differently from [17], we deal with these conditions assuming that E dG preserves sufficiently many moments of v on the d-simplices and on the faces of , in the vein of [26, Section 3.2].

Lemma 9 (Consistency by Moment-Preserving Operators).

Assume that the operator E dG : ( S 0 ) d H 0 1 ( Ω ) d is such that

(2.14)

(2.14a) F E dG v m F = F { { v } } m F for all  F i , m F - 1 ( F ) d ,
(2.14b) K E dG v m K = K v m K for all  K , m K - 2 ( K ) d ,

for all v ( S 0 ) d . Then E dG satisfies conditions (2.13a) and (2.13b).

Proof.

Let w ( S 0 ) d and q S ^ - 1 0 . The integration by parts formula (2.4) yields

b dG ( w , q ) = Ω w q - Σ Ω { { w } } n [ [ q ] ] .

In view of (2.14), we can replace w by E dG w in this identity. Then we integrate back by parts and note that [ [ E dG w ] ] = 0 on Σ, because of the inclusion E dG w H 0 1 ( Ω ) d ,

b dG ( w , q ) = Ω E dG w q - Σ Ω E dG w n [ [ q ] ] = - Ω q div E dG w .

This entails that (2.13b) holds. Arguing similarly, we infer that

(2.15) a dG ( w , v ) = Ω w : E dG v - Σ [ [ w ] ] { { v } } n + Σ η h [ [ w ] ] [ [ v ] ]

for all w , v ( S 0 ) d , cf. [26, Lemma 3.1]. Hence, we conclude that (2.13a) holds, because [ [ w ] ] = 0 on Σ in view of the inclusion w Z . ∎

Remark 10 (Alternative Approach to Consistency).

The implication (2.14)   (2.13) stated in the previous lemma relies on our choice of the forms a dG and b dG and fails to hold for different discretizations. Roughly speaking, this happens whenever some sort of reduced integration of the divergence is involved. In such cases the consistency conditions (2.13) need to be accommodated differently, for instance by the augmented Lagrangian formulation proposed in [17].

The next theorem states that (2.14) is indeed a sufficient condition for quasi-optimality. The essence of this result and a partial proof can be found also in [2, Section 6].

Theorem 11 (Quasi-Optimality).

Assume that the operator E dG satisfies (2.14) for all v ( S 0 ) d . Moreover, let η > η ¯ , where η ¯ is as in (2.9). Then, denoting by ( u , p ) and ( u dG , p dG ) the solutions of (2.1) and (2.6), respectively, with viscosity μ > 0 and load f H - 1 ( Ω ) , we have

μ u - u dG dG 1 + E dG α ¯ dG ( δ dG μ inf w ( S 0 ) d u - w dG + inf q S ^ - 1 0 p - q L 2 ( Ω ) )

and

p - p dG L 2 ( Ω ) ( 1 + E dG ) 2 α ¯ dG β dG ( δ dG μ inf w ( S 0 ) d u - w dG + inf q S ^ - 1 0 p - q L 2 ( Ω ) ) .

Proof.

We first estimate the velocity error. For this purpose, let Π u Z dG be the ( , ) dG -orthogonal projection of u onto Z dG , that is

Ω ( Π u - u ) : v + Σ η h [ [ Π u ] ] [ [ v ] ] = 0 for all  v Z dG .

Setting z := u dG - Π u , the coercivity (2.10b) and the first equation of problems (2.1) and (2.6) yield

(2.16) α ¯ dG u dG - Π u dG 2 ( Ω u : E dG z - a dG ( Π u , z ) ) - 1 μ ( Ω p div E dG z + b dG ( z , p dG ) ) = : 𝔗 1 - 𝔗 2 μ .

We bound 𝔗 1 according to the definition of Π u , identity (2.15) (whose validity is guaranteed by Lemma 9) and inequality (2.9)

𝔗 1 = Ω ( u - Π u ) : ( E dG z - z ) + Σ [ [ Π u ] ] { { z } } n ( 2 + E dG ) u - Π u dG z dG .

We bound 𝔗 2 in view of the inclusion z Z dG (which implies b dG ( z , p dG ) = 0 ), identity (2.13b) (whose validity is guaranteed by Lemma 9) and [20, Lemma 2.1]. Thus, we obtain

𝔗 2 = Ω ( p - q ) div E dG z E dG z dG p - q L 2 ( Ω )

for all q S ^ - 1 0 . We insert the estimates of 𝔗 1 and 𝔗 2 into (2.16) and apply the triangle inequality to obtain

u - u dG dG 3 + E dG α ¯ dG inf z Z dG u - z dG + E dG μ α ¯ dG inf q S ^ - 1 0 p - q L 2 ( Ω ) ,

where we have used α ¯ dG 1 . We derive the claimed estimate of the velocity error by invoking Lemma 7.

Next, in order to estimate the pressure error, let R p S ^ - 1 0 be the L 2 -orthogonal projection of p onto S ^ - 1 0 . The inf-sup stability (2.10c) and the triangle inequality yield

(2.17) p - p dG L 2 ( Ω ) p - R p L 2 ( Ω ) + 1 β dG sup v ( S 0 ) d b dG ( v , p dG - R p ) v dG .

Let v ( S 0 ) d . The first equations of problems (2.1) and (2.6) reveal

(2.18) b dG ( v , p dG - R p ) = μ ( Ω u : E dG v - a dG ( u dG , v ) ) - ( Ω p div E dG v + b dG ( v , R p ) ) = : μ 𝔘 1 + 𝔘 2 .

We bound 𝔘 1 according to identity (2.15) (which holds in view of Lemma 9) and inequality (2.9)

𝔘 1 = Ω ( u - u dG ) : E dG v + Σ [ [ u dG ] ] { { v } } n - Σ η h [ [ u dG ] ] [ [ v ] ] ( 2 + E dG ) u - u dG dG v dG .

We bound 𝔘 2 making use of identity (2.13b) (which holds in view of Lemma 9) and [20, Lemma 2.1]

𝔘 2 = Ω ( p - R p ) div E dG v E dG v dG p - R p L 2 ( Ω ) .

We insert the estimates of 𝔘 1 and 𝔘 2 into (2.17) and (2.18) to obtain

p - p dG L 2 ( Ω ) μ 2 + E dG β dG u - u dG dG + 1 + E dG β dG inf q S ^ - 1 0 p - q L 2 ( Ω ) .

We derive the claimed estimate of the pressure error by means of the previous estimate of the velocity error. ∎

2.6 Quasi-Optimality and Pressure Robustness

The assumptions in Theorem 11 do not guarantee that the discretization (2.6) is pressure robust in the sense of (1.3). We illustrate this by a numerical experiment in Section 4.2. Similarly as in [17], we achieve pressure robustness by the additional assumption that the operator E dG preserves the discrete divergence. Recalling the definition of div dG in (2.7), this corresponds to prescribing a reinforced version of (2.13b).

Theorem 12 (Quasi-Optimality and Pressure Robustness).

Assume that the operator E dG satisfies (2.14) and

(2.19) div E dG v = div dG v

for all v ( S 0 ) d . Moreover, let η > η ¯ , where η ¯ is as in (2.9). Then, denoting by ( u , p ) and ( u dG , p dG ) the solutions of (2.1) and (2.6), respectively, with viscosity μ > 0 and load f H - 1 ( Ω ) , we have

u - u dG dG C δ dG 1 + E dG α ¯ dG inf w ( S 0 ) d u - w dG

and

p - p dG L 2 ( Ω ) C δ dG ( 1 + E dG ) 2 α ¯ dG β dG μ inf w ( S 0 ) d u - w dG + inf q S ^ - 1 0 p - q L 2 ( Ω ) .

Proof.

The proof is the same as for Theorem 11, with the only difference that we have 𝔗 2 = 0 and 𝔘 2 = 0 in (2.16) and (2.18), respectively, as a consequence of (2.19). ∎

3 A Moment- and Divergence-Preserving Operator

Motivated by the error estimates in Theorem 12, we now aim at designing a linear operator

E dG : ( S 0 ) d H 0 1 ( Ω ) d

which satisfies the following conditions:

(3.1)

(3.1a) E dG  is stable, in that  E dG C ,
(3.1b) E dG  preserves  - 1 ( F ) d -moments on each  F i , see (2.14a) ,
(3.1c) E dG  preserves the discrete divergence  div dG , see (2.19) ,
(3.1d) E dG  preserves  - 2 ( K ) d -moments in each  K , see (2.14b) .

We restrict ourselves to the case d = 2 , in order to keep the discussion as easy as possible. In Section 3.7, we discuss the differences in the design for d = 3 .

3.1 Outline of the Construction

We first outline the strategy underlying our construction before we enter into the technical details. We shall obtain E dG from the combination of four operators, namely

(3.2) E dG := E 1 + E 2 + E 3 + E 4 .

Our construction has a recursive structure in the sense that the definition of E i , i { 2 , , 4 } , involves the one of E 1 , , E i - 1 . The role of each summand in (3.2) can be summarized as follows.

  1. The first operator E 1 maps ( S 0 ) 2 into ( S 0 H 0 1 ( Ω ) ) 2 by a simple averaging technique and is stable, in that (3.1a) holds.

  2. The second operator preserves the stability of E 1 , while correcting the moments on faces. This is obtained by mapping into a space of face-bubbles. As a result, the sum E 1 + E 2 enjoys both (3.1a) and (3.1b).

  3. The third operator E 3 additionally enforces (3.1c), while preserving the validity of the previous properties. This is achieved by mapping into a space of volume-bubbles.

  4. Finally, the fourth operator E 4 maps into a space of divergence-free volume-bubbles and is designed to guarantee that E dG enjoys also the last condition prescribed in (3.1d).

For v ( S 0 ) 2 , the definition of E 1 in a simplex K involves the values of v in the star around K, i.e. in the neighboring simplices. The operators E 2 , E 3 and E 4 are obtained solving local problems on the faces or on the simplices of . Each local problem is independent of the others and can efficiently be solved by resorting to a reference configuration. Thus, the resulting operator E dG is computationally feasible, in the sense that, for any nodal basis function Φ of ( S 0 ) 2 , the computation of E dG Φ requires only O ( 1 ) operations.

3.2 Preliminary Observations

The main difficulty in the construction of E dG is that conditions (3.1b), (3.1c) and (3.1d) are not linearly independent. In fact, prescribing sufficiently many moments of E dG v on the skeleton of as well as the divergence of E dG v can be expected to prescribe implicitly also the moments of E dG v times gradients on each simplex of . The next lemma states this observation more precisely, showing also that the above conditions are at least compatible.

Lemma 13 ( P - 1 ( K ) -Moments).

Let E : ( S 0 ) 2 H 0 1 ( Ω ) 2 be an operator fulfilling (3.1b) and (3.1c). Then, for all v ( S 0 ) 2 , K M and q P - 1 ( K ) , we have

K E v q = K v q .

Proof.

Let v ( S 0 ) 2 and q - 1 ( K ) be given. We extend q to Ω K by zero. The integration by parts formula (2.4) yields

K E v q = - Ω q div E v + Σ Ω E v n [ [ q ] ] = - Ω q div dG v + Σ Ω { { v } } n [ [ q ] ] ,

where the second identity follows from the assumption that E satisfies (3.1b) and (3.1c). The equivalent definition of the discrete divergence in Remark 1 entails that

K E v q = - Ω q div v + Σ [ [ v ] ] n { { q } } + Σ Ω { { v } } n [ [ q ] ] .

We conclude invoking once again the element-wise integration by parts formula and recalling that q vanishes in Ω K . ∎

The above lemma suggests that we should enforce (3.1d) only on some complement of - 1 ( K ) in - 2 ( K ) 2 . We proceed in analogy with [18, Section 4], where a similar issue was encountered. In particular, we shall identify one such complement and construct E dG with the help of the curl and rot operators, that are defined as

(3.3) curl ( w ) := ( 2 w , - 1 w ) and rot ( v ) := - 2 v 1 + 1 v 2

where w and v = ( v 1 , v 2 ) , respectively, are scalar- and vector-valued functions. Recall that assuming w H 0 1 ( K ) and v H 1 ( K ) , we have

(3.4) K curl ( w ) v = K w rot ( v )

for all K . Moreover, it holds

(3.5) div ( curl ( v ) ) = 0 .

Recall the convention - 1 = { 0 } . The next lemma provides the desired decomposition of - 2 ( K ) 2 .

Lemma 14 (Decomposition of Vector-Valued Polynomials).

For all k 0 and K M , define

x k - 1 ( K ) := { x r := ( - x 2 r , x 1 r ) r k - 1 ( K ) } .

The operator rot : P k ( K ) 2 P k - 1 ( K ) is injective on x P k - 1 ( K ) and its kernel coincides with P k + 1 ( K ) . As a consequence, we have

k ( K ) 2 = k + 1 ( K ) x k - 1 ( K ) .

Proof.

Assume k 1 , since the claim is clear for k = 0 . Let rot ( x r ) = 0 for some r = | α | k - 1 a α x α k - 1 ( K ) . We infer | α | k - 1 ( 2 + | α | ) a α x α = 0 , showing that r = 0 . This proves the injectivity of rot on x k - 1 ( K ) . Next, the fact that the kernel of rot on k ( K ) 2 coincides with k + 1 ( K ) is a standard result from vector calculus. This entails that k + 1 ( K ) x k - 1 ( K ) = { 0 } . Then the claimed decomposition of k ( K ) 2 follows from a dimensional argument. ∎

As mentioned before, the construction of the operators E 3 and E 4 in (3.2) involves the solution of local problems on each triangle in . For both theoretical and computational convenience, we shall formulate such problems on a reference triangle K ref , with the help of the Piola’s transformations, see, e.g., [4, Section 2.1.3]. Hence, for all K , we fix a one-to-one affine mapping F K : K ref K , with Jacobian matrix D F K . We set J K := | det D F K | . Note that D F K is a constant invertible matrix and that J K is a positive constant.

The contravariant and the covariant Piola’s transformations, respectively, map functions v ref , w ref from H 1 ( K ref ) 2 into H 1 ( K ) 2 and are given by

(3.6) 𝒫 K con v ref := J K - 1 D F K ( v ref F K - 1 ) and 𝒫 K cov w ref := D F K - T ( w ref F K - 1 ) .

Remarkably, we have

(3.7) K 𝒫 K con v ref 𝒫 K cov w ref = K ref v ref w ref .

Moreover, the contravariant Piola’s transformation is such that

(3.8) div ( 𝒫 K con v ref ) = J K - 1 ( div v ref ) F K - 1

and

(3.9) 𝒫 K con v ref L 2 ( K ) J K - 1 2 h K v ref L 2 ( K ref ) C 𝒫 K con v ref L 2 ( K ) .

3.3 Construction of E dG

We now construct an operator E dG : ( S 0 ) d H 0 1 ( Ω ) d , , which satisfies (3.1). As stated in (3.2), we set E dG := i = 1 4 E i , where each operator E i is defined as follows.

Definition of E 1 .

Each polynomial in ( K ) , K , is uniquely determined by its point values at the Lagrange nodes ( K ) of degree . Recall also that the nodal degrees of freedom of S 1 = S 0 H 0 1 ( Ω ) are given by the evaluations at the points := K ( K ) Ω . We denote by Φ z S 1 the nodal basis function associated with the evaluation at z , that is Φ z ( y ) = δ z y for all y , z . Then, for v ( S 0 ) d , we let E 1 v be defined by

(3.10) E 1 v := z 1 N z ( K , K z v | K ( z ) ) Φ z ,

where N z is the number of triangles in touching z. Averaging operators like E 1 or variants are common devices in the context of dG methods, see, e.g., [10, Section 5.5.2].

Definition of E 2 .

We define E 2 in the vein of [26, Section 3.2]. For every interior edge F i , let K 1 , K 2 be such that F = K 1 K 2 . Denote by ( F ) := ( K 1 ) ( K 2 ) the Lagrange nodes of degree on F and let b F := z 1 ( F ) Φ 1 z be the quadratic face bubble supported on K 1 K 2 . We introduce a linear operator E 2 , F : L 2 ( F ) 2 - 1 ( F ) 2 by solving the local problem

(3.11) F E 2 , F v m F b F = F v m F for all  m F - 1 ( F ) 2 .

Then, for v ( S 0 ) 2 , we set

(3.12) E 2 v := F i z - 1 ( F ) E 2 , F ( { { v } } - E 1 v ) ( z ) Φ - 1 z b F .

Note that each summand involves an extension from F to K 1 K 2 .

Definition of E 3 .

We define E 3 in the vein of [17, 27]. Let K ref be the reference triangle introduced in Section 3.2. We obtain a triangulation ref of K ref connecting each vertex with the barycenter, see Figure 1. The space S + 1 0 ( ref ) consists of all piecewise polynomials of degree ( + 1 ) on ref . We consider the subspaces

S + 1 1 ( ref ) := S + 1 0 ( ref ) H 0 1 ( K ref ) and S ^ 0 ( ref ) := S 0 ( ref ) L 0 2 ( K ref )

and introduce a linear operator E 3 , ref : S ^ 0 ( ref ) S + 1 1 ( ref ) 2 by imposing

(3.13) E 3 , ref ( q ref ) := argmin { v ref L 2 ( K ref ) 2 v ref S + 1 1 ( ref ) 2 , div v ref = q ref } .

This constrained quadratic minimization problem is uniquely solvable as a consequence of [13, Theorem 3.1]. Note that we can equivalently rewrite (3.13) as a discrete Stokes-like problem, with velocity space S + 1 1 ( ref ) 2 , pressure space S ^ 0 ( ref ) and right-hand side zero in the momentum equation and q ref in the continuity equation. Then, for v ( S 0 ) 2 , we define

(3.14) E 3 v := K 𝒫 K con E 3 , ref ( J K ( div dG v - i = 1 2 div E i v ) F K ) ,

where each summand vanishes on K and is extended by zero outside K. The discussion in the next section confirms that the argument of E 3 , ref is indeed an element of S ^ 0 ( ref ) .

Definition of E 4 .

Denote by b ref the cubic bubble function on K ref , that is obtained by taking the product of the Lagrange basis functions of 1 ( K ref ) associated with the evaluations at the vertices of K ref . For 3 , we introduce a linear operator E 4 , ref : L 2 ( K ref ) 2 x - 3 ( K ref ) by imposing

(3.15) K ref rot ( E 4 , ref q ref ) rot ( m ref ) b ref 2 = K ref q ref m ref for all  m ref x - 3 ( K ref ) .

Lemma 14 ensures that this problem is uniquely solvable. Then, for v ( S 0 ) 2 , we define E 4 v = 0 if { 1 , 2 } , otherwise

(3.16) E 4 v := K 𝒫 K con curl ( b ref 2 rot E 4 , ref ( 𝒫 K con ) - 1 ( v - i = 1 3 E i v ) ) ,

where each summand vanishes on K and is extended by zero outside K.

Figure 1

Reference triangle K ref (left) and triangulation ref (right).

3.4 Preservation Properties of E dG

In this subsection we prove that the operator E dG defined above satisfies conditions (3.1b), (3.1c) and (3.1d), i.e. it preserves the discrete divergence and all the prescribed moments on the faces and the triangles of . For this purpose, we make use of the following integration by parts formula, which generalizes Lemma 13.

Lemma 15.

Let E : ( S 0 ) 2 H 0 1 ( Ω ) 2 be a linear operator satisfying (3.1b). Then, for all v ( S 0 ) 2 , K M and q P - 1 ( K ) , we have

K ( div dG v - div E v ) q = - K ( v - E v ) q .

Proof.

Proceed as in the proof of Lemma 13, without assuming that E satisfies condition (3.1c). ∎

We are now prepared to prove the claimed properties of E dG .

Theorem 16 (Preservation Properties of E dG ).

The operator E dG defined in Section 3.3 satisfies conditions (3.1b), (3.1c) and (3.1d).

Proof.

Let v ( S 0 ) 2 . We check one by one the validity of the desired conditions.

Proof of (3.1b). By construction, each summand in the definitions (3.14) and (3.16) of E 3 and E 4 , respectively, is supported in one triangle K and vanishes on K . This entails that E dG v = E 1 v + E 2 v on the skeleton Σ. Moreover, for F i , we have ( E 2 v ) | F = E 2 , F ( { { v } } - E 1 v ) b F , as a consequence of (3.12). Hence, for all m F - 1 ( F ) 2 , the definition of E 2 , F in (3.11) implies

F E 2 v m F = F E 2 , F ( { { v } } - E 1 v ) m F b F = F ( { { v } } - E 1 v ) m F .

Rearranging terms, we infer that

F E dG v m F = F ( E 1 v + E 2 v ) m F = F { { v } } m F .

Proof of (3.1c). Each summand in the definition (3.16) of E 4 is divergence-free, as a consequence of (3.5) and (3.8). This entails that div E dG v = i = 1 3 div E i v in Ω. Moreover, for K , the identity (3.8) and the definitions (3.13) and (3.14) of E 3 , ref and E 3 , respectively, reveal that

( div E 3 v ) | K = ( div dG v - i = 1 2 div E i v ) | K .

Rearranging terms, we infer that

div E dG v = i = 1 3 div E i v = div dG v .

Proof of (3.1d). For all K , the covariant Piola’s transformation 𝒫 K cov from (3.6) maps - 2 ( K ref ) 2 into - 2 ( K ) 2 and is one-to-one. Then, according to the transformation rule (3.7), we see that the identity

(3.17) K ref ( 𝒫 K con ) - 1 E dG v m ref = K ref ( 𝒫 K con ) - 1 v m ref for all  m ref - 2 ( K ref ) 2 ,

is an equivalent formulation of (3.1d). Moreover, according to the decomposition stated in Lemma 14, we can split (3.17) into two independent conditions with test functions in - 1 ( K ref ) and x - 3 ( K ref ) , respectively. Let us first assume that m ref = q ref - 1 ( K ref ) . The definition of E 4 in (3.16), the integration by parts rule (3.4) and Lemma 14 imply that

K ref ( 𝒫 K con ) - 1 E 4 v m ref = K ref curl ( b ref 2 rot E 4 , ref ( 𝒫 K con ) - 1 ( v - i = 1 3 E i v ) ) q ref = K ref rot E 4 , ref ( 𝒫 K con ) - 1 ( v - i = 1 3 E i v ) rot ( q ref ) b ref 2 = 0 .

Next, recall the definitions of E 3 , ref and E 3 in (3.13) and (3.14), respectively. Integrating by parts, changing variables twice and invoking Lemma 15, we obtain

K ref ( 𝒫 K con ) - 1 E 3 v m ref = K ref E 3 , ref v q ref = - K ref ( div E 3 , ref v ) q ref = - K ref J K ( div dG v - i = 1 2 div E i v ) F K q ref = - K ( div dG v - i = 1 2 div E i v ) q ref F K - 1 = - K ( v - i = 1 2 E i v ) ( q ref F K - 1 ) = K ref ( 𝒫 K con ) - 1 ( v - i = 1 2 E i v ) m ref .

We combine this identity with the previous one and we rearrange terms according to the definition (3.2) of E dG . Then we infer that (3.17) holds for all m ref - 1 ( K ref ) . This concludes the proof for { 1 , 2 } . For 3 , assume further m ref x - 3 ( K ref ) . The definitions of E 4 , ref and E 4 in (3.15) and (3.16), respectively, and the integration by parts rule (3.4) reveal that

K ref ( 𝒫 K con ) - 1 E 4 v m ref = K ref rot ( E 4 , ref ( 𝒫 K con ) - 1 ( v - i = 1 3 E i v ) ) rot ( m ref ) b ref 2 = K ref ( 𝒫 K con ) - 1 ( v - i = 1 3 E i v ) m ref .

Rearranging terms, we infer that (3.17) holds for all m ref x - 3 ( K ref ) . Thus, Lemma 14 and the above discussion entail that E dG satisfies condition (3.1d). ∎

3.5 Stability of E dG

In this subsection we prove that the operator E dG defined in Section 3.3 satisfies condition (3.1a), i.e. it is stable in the norm dG and its stability constant E dG is bounded in terms of the shape constant γ of and of the polynomial degree . We begin by recalling a standard result concerning the operator E 1 defined in (3.10), see, e.g., [10, Section 5.5.2].

Lemma 17 (Local L 2 -Estimate of E 1 ).

For all v ( S 0 ) 2 and K M , we have

v - E 1 v L 2 ( K ) C F , F K h F 1 2 [ [ v ] ] L 2 ( F ) .

Next, we prove that E dG enjoys the same local estimate as E 1 , possibly up to a different constant.

Proposition 18 (Local L 2 -Estimate of E dG ).

The operator E dG defined in Section 3.3 is such that, for all v ( S 0 ) 2 and K M ,

v - E dG v L 2 ( K ) C F , F K h F 1 2 [ [ v ] ] L 2 ( F ) .

Proof.

First of all, we recall that E dG = i = 1 4 E i and apply the triangle inequality

v - E dG v L 2 ( K ) v - E 1 v L 2 ( K ) + i = 2 4 E i v L 2 ( K ) .

According to Lemma 17, we only need to bound the last three summands in the right-hand side. We estimate these terms one by one.

Estimate of E 2 . The definition of E 2 in (3.12) involves an extension from each interior face to the two neighboring triangles. A standard scaling argument reveals that such an extension is bounded, meaning that

E 2 v L 2 ( K ) C F i , F K h F 1 2 E 2 , F ( { { v } } - E 1 v ) b F L 2 ( F ) .

Recalling also the definition of E 2 , F in (3.11) and the bound b F 1 , we infer that

E 2 , F ( { { v } } - E 1 v ) b F L 2 ( F ) { { v } } - E 1 v L 2 ( F )

for all F i with F K . We insert this estimate into the previous one. Then we observe that

| { { v } } - v | K | = 1 2 | [ [ v ] ] |

on each face F involved in the above summation. This fact and an inverse trace inequality entail that

E 2 v L 2 ( K ) C ( v - E 1 v L 2 ( K ) + F i , F K h F 1 2 [ [ v ] ] L 2 ( F ) ) .

Then Lemma 17 yields

(3.18) E 2 v L 2 ( K ) C F , F K h F 1 2 [ [ v ] ] L 2 ( F ) .

Estimate of E 3 . The definition of E 3 in (3.14) and the transformation rule (3.9) imply that

E 3 v L 2 ( K ) J K - 1 2 h K E 3 , ref ( J K ( div dG v - i = 1 2 div E i v ) F K ) L 2 ( K ref ) .

Since E 3 , ref is a linear operator defined on a finite-dimensional space, it is bounded. We combine this observation with a change of variables

E 3 v L 2 ( K ) C h K div dG v - i = 1 2 div E i v L 2 ( K ) .

The inclusion E 1 v ( S 0 H 0 1 ( Ω ) ) 2 reveals that div E 1 v = div dG E 1 v . Recalling the equivalent definition of div dG in Remark 1, we obtain

div dG ( v - E 1 v ) L 2 ( K ) C ( div ( v - E 1 ) L 2 ( K ) + F , F K h F - 1 2 [ [ v ] ] L 2 ( F ) ) ,

where we have made use also of the identity [ [ E 1 v ] ] = 0 on Σ and of the inverse inequality (2.9). We combine this bound with the previous one and apply twice the inverse inequality div L 2 ( K ) C h K - 1 L 2 ( K ) . This entails that

E 3 v L 2 ( K ) C ( v - E 1 v L 2 ( K ) + E 2 v L 2 ( K ) + F , F K h F 1 2 [ [ v ] ] L 2 ( F ) ) .

Then Lemma 17 and inequality (3.18) yield

(3.19) E 3 v L 2 ( K ) C F , F K h F 1 2 [ [ v ] ] L 2 ( F ) .

Estimate of E 4 . Let K . The definition of E 4 in (3.16) and the transformation rule (3.9) imply that

E 4 v L 2 ( K ) J K - 1 2 h K curl ( b ref 2 rot E 4 , ref ( 𝒫 K con ) - 1 ( v - i = 1 3 E i v ) ) L 2 ( K ref ) .

Since E 4 , ref is a linear operator defined on a finite-dimensional space, it is bounded. We combine this observation with an inverse estimate, the transformation rule (3.9) and the triangle inequality

E 4 v L 2 ( K ) C J K - 1 2 h K ( 𝒫 K con ) - 1 ( v - i = 1 3 E i v ) L 2 ( K ref ) C ( v - E 1 v L 2 ( K ) + i = 2 3 E i v L 2 ( K ) ) .

Then Lemma 17 and inequalities (3.18) and (3.19) yield

E 3 v L 2 ( K ) C F , F K h F 1 2 [ [ v ] ] L 2 ( F ) .

The local estimate in Proposition 18 ensures that E dG satisfies condition (3.1a).

Theorem 19 (Stability of E dG ).

The operator E dG defined in Section 3.3 satisfies condition (3.1a) in that, for all v ( S 0 ) 2 , we have

E dG v L 2 ( Ω ) C max { 1 , 1 η } v dG .

Proof.

Let K . An inverse estimate and Proposition 18 imply that

( v - E dG v ) L 2 ( K ) C F , F K h F - 1 2 [ [ v ] ] L 2 ( F ) .

We square both sides in this inequality and sum over all K . Recalling that the number of triangles touching a given face is bounded in terms of the shape constant of , we obtain

( v - E dG v ) L 2 ( Ω ) C ( Σ h - 1 | [ [ v ] ] | 2 ) 1 2 .

We conclude by recalling the definition of the norm dG in Section 2.3. ∎

3.6 Main Results

We are now able to derive the main result of this paper. For this purpose, we invoke [24, Corollary 1] and derive the following upper bound of the velocity best error

inf w ( S 0 ) 2 u - w dG inf w ( S 0 H 0 1 ( Ω ) ) 2 ( u - w ) L 2 ( Ω ) C inf w ( S 0 ) 2 ( u - w ) L 2 ( Ω )

for all u H 0 1 ( Ω ) 2 . Notice that the right-hand side is independent of the penalty parameter η and bounds the left-hand side also from below. We combine this bound with Theorems 12, 16 and 19. Recall also the definition of α ¯ dG and the upper bounds of β dG - 1 and δ dG in (2.10) and Lemma 7, respectively.

Theorem 20 (Quasi-Optimality and Pressure Robustness by E dG ).

Let η > η ¯ , where η ¯ is as in (2.9). Denote by ( u , p ) and ( u dG , p dG ) the solutions of (2.1) and (2.6), respectively, in dimension d = 2 , with viscosity μ > 0 and load f H - 1 ( Ω ) . Moreover, let E dG be the operator defined in Section 3.3. Then we have

u - u dG dG C η inf w ( S 0 ) 2 ( u - w ) L 2 ( Ω )

and

p - p dG L 2 ( Ω ) C μ η inf w ( S 0 ) 2 ( u - w ) L 2 ( Ω ) + inf q S ^ - 1 0 p - q L 2 ( Ω ) .

The above design of E dG can be simplified when the sole quasi-optimality (without pressure robustness) is concerned. In this case, we can apply the operator E from [26, Proposition 3.4] component-wise. This gives rise to

(3.20) E ~ dG v := ( E v 1 , E v 2 ) , v = ( v 1 , v 2 ) ( S 0 ) 2 .

According to [26, Proposition 3.4], the resulting operator is moment-preserving and stable, in the sense that it satisfies conditions (3.1a), (3.1b) and (3.1d). Then the following weaker counterpart of Theorem 20 readily follows from Theorem 11.

Theorem 21 (Quasi-Optimality by E ~ dG ).

Let η > η ¯ , where η ¯ is as in (2.9). Denote by ( u , p ) and ( u dG , p dG ) the solutions of (2.1) and (2.6), respectively, in dimension d = 2 , with viscosity μ > 0 and load f H - 1 ( Ω ) . Moreover, let E dG be replaced by E ~ dG . Then we have

μ u - u dG dG C ( μ η inf w ( S 0 ) 2 ( u - w ) L 2 ( Ω ) + inf q S ^ - 1 0 p - q L 2 ( Ω ) )

and

p - p dG L 2 ( Ω ) C η ( μ η inf w ( S 0 ) 2 ( u - w ) L 2 ( Ω ) + inf q S ^ - 1 0 p - q L 2 ( Ω ) ) .

3.7 Construction of E dG for d = 3

We end this section with some comments concerning the extension of the previous results to the discretization of the Stokes equations in dimension d = 3 . First of all, the three-dimensional curl operator has to be used instead of the two-dimensional operators curl and rot from (3.3). The decomposition of vector-valued polynomials stated in Lemma 14 and used in the definition of E 4 reads

k ( K ) 3 = k + 1 ( K ) x k - 1 ( K ) 3

for all K and k 0 , where

x k - 1 ( K ) 3 := { x r := ( x 2 r 3 - x 3 r 2 x 3 r 1 - x 1 r 3 x 1 r 2 - x 2 r 1 ) | r = ( r 1 , r 2 , r 3 ) k - 1 ( K ) 3 } .

This can be verified by noticing that the operator curl is injective on x k - 1 ( K ) 3 .

The construction of E dG remains the same as in Section 3.3, up to the following minor modifications. The face bubble function b F involved in the definition of E 2 has degree three (and not two). Consequently, the operator E 3 , ref maps S ^ + 1 0 ( ref ) into S + 2 1 ( ref ) 3 , so as to guarantee that E 3 is well-defined. Finally, the volume bubble function b ref involved in the definition of E 4 has degree four (and not three).

With these ingredients, the statements and the proofs of the results in Sections 3.43.6 can easily be adapted to the case d = 3 .

4 Numerical Experiments

We now discuss the results obtained when approximating the solution of the Stokes equations (2.1) with

d = 2 , Ω = ( 0 , 1 ) × ( 0 , 1 ) , μ = 1 .

We test the discretization (2.6) of the Stokes equations with

  1. E dG = Id as in [14] (standard discretization),

  2. E dG defined by (3.20) as in Theorem 21 (quasi-optimal discretization),

  3. E dG defined by Section 3.3 as in Theorem 20 (quasi-optimal and pressure robust discretization).

The first option differs from the others, in that E dG does not map ( S 0 ) 2 into H 0 1 ( Ω ) 2 . Therefore, the duality in the right-hand side of (2.6) is not defined for a general load f H - 1 ( Ω ) . This observation clearly favors the second and the third discretizations when rough loads are concerned, cf. [27, Section 6.4].

We consider only the first-order discretization in our experiments, i.e. we set

= 1

in (2.6). The numerical results are obtained with the help of ALBERTA 3.0 [15, 21].

4.1 Smooth Exact Solution

We first consider a test case with smooth exact solution, namely

u ( x 1 , x 2 ) = curl ( x 1 2 ( 1 - x 1 ) 2 x 2 2 ( 1 - x 2 ) 2 ) , p ( x 1 , x 2 ) = ( x 1 - 0.5 ) ( x 2 - 0.5 ) .

We discretize the domain Ω by the following family of structured meshes. Given N 0 , we divide Ω into 2 N × 2 N identical squares with area 2 - 2 N . Then we obtain the “crisscross” mesh N C by drawing the two diagonals of each square, see Figure 2. We use the penalty parameter

η = 6 .

For N = 4 , , 8 , we report the velocity error u - u dG dG and the pressure error p - p dG L 2 ( Ω ) , for the three discretizations listed above, in Tables 1 and 2, respectively. For each sequence of errors ( e N ) , we compute the experimental order of convergence

EOC N := log ( e N e N - 1 ) log ( # N - 1 C # N C ) , N 1 ,

where # N C denotes the number of triangles in N C . Observing the numerical data, we see that the errors of the three discretizations behave quite similarly and converge to zero at the maximum decay rate ( # N C ) - 0.5 . In this case, the standard discretization should be preferred for the easier construction of the operator E dG .

Figure 2

The crisscross mesh N C with N = 2 .

(a)
(a)
Table 1

Velocity errors of the standard ( 𝚜𝚝𝚗𝚍 ), quasi-optimal ( 𝚚𝚘𝚙𝚝 ) and quasi-optimal and pressure robust ( 𝚙𝚛𝚘𝚋 ) discretizations with experimental orders of convergence.

N 𝚜𝚝𝚗𝚍 EOC 𝚚𝚘𝚙𝚝 EOC 𝚙𝚛𝚘𝚋 EOC
4 8.2516e - 03 8.3795e - 03 8.5337e - 03
5 3.8937e - 03 0.54 3.9344e - 03 0.55 4.1273e - 03 0.52
6 1.8797e - 03 0.53 1.8910e - 03 0.53 2.0231e - 03 0.51
7 9.2180e - 04 0.51 9.2477e - 04 0.52 1.0007e - 03 0.51
8 4.5621e - 04 0.51 4.5698e - 04 0.51 4.9756e - 04 0.50
Table 2

Pressure errors of the standard ( 𝚜𝚝𝚗𝚍 ), quasi-optimal ( 𝚚𝚘𝚙𝚝 ) and quasi-optimal and pressure robust ( 𝚙𝚛𝚘𝚋 ) discretizations with experimental orders of convergence.

N 𝚜𝚝𝚗𝚍 EOC 𝚚𝚘𝚙𝚝 EOC 𝚙𝚛𝚘𝚋 EOC
4 4.4477e - 03 4.4862e - 03 4.3843e - 03
5 2.2248e - 03 0.50 2.2377e - 03 0.50 2.2109e - 03 0.49
6 1.1142e - 03 0.50 1.1178e - 03 0.50 1.1109e - 03 0.50
7 5.5781e - 04 0.50 5.5878e - 04 0.50 5.5692e - 04 0.50
8 2.7912e - 04 0.50 2.7937e - 04 0.50 2.7884e - 04 0.50

4.2 Jumping Pressure

In order to investigate the pressure robustness of the three discretizations, we consider a test case with smooth exact velocity and rough exact pressure, namely

u ( x 1 , x 2 ) = curl ( x 1 2 ( x 1 - 1 ) 2 x 2 2 ( x 2 - 1 ) 2 ) , p ( x 1 , x 2 ) = { π π - 1 if  x 1 > π - 1 , - π if  x 1 < π - 1 .

We use unstructured meshes to make sure that the discontinuity of p along the line x 1 = π - 1 is never resolved. More precisely, we start from the initial mesh 0 U in Figure 3 (left) and we obtain the next meshes N U , N 1 , through uniform refinements by bisection. In this case, we set the penalty parameter to

η = 8 .

The data displayed in Figure 3 (right) show that the velocity error of the quasi-optimal and pressure robust discretization fully exploits the regularity of u and converges to zero at the maximum decay rate ( # N U ) - 0.5 , in accordance with Theorem 20. In contrast, the low regularity of p impairs the approximation of u in the standard discretization and in the quasi-optimal one. In fact, the corresponding velocity errors converge at the suboptimal decay rate ( # N U ) - 0.25 . This confirms, in particular, that the first estimate in Theorem 21 captures the correct behavior of the velocity error in the quasi-optimal discretization.

Figure 3

Initial mesh 0 U (left) and velocity error as a function of # N U (right) for the standard ( ), quasi-optimal ( * ) and quasi-optimal and pressure robust ( ) discretizations. Plain and dashed lines indicate the decay rates ( # N U ) - 0.5 and ( # N U ) - 0.25 , respectively.

(a)
(a)
(b)
(b)

Award Identifier / Grant number: KR 3984/5-1

Funding statement: Christian Kreuzer gratefully acknowledges partial support by the DFG research grant “Convergence Analysis for Adaptive Discontinuous Galerkin Methods” (KR 3984/5-1). The research of Pietro Zanotti was supported by the GNCS-INdAM through the program “Finanziamento giovani ricercatori 2019-2020”.

References

[1] D. N. Arnold, F. Brezzi, B. Cockburn and L. D. Marini, Unified analysis of discontinuous Galerkin methods for elliptic problems, SIAM J. Numer. Anal. 39 (2001/02), no. 5, 1749–1779. 10.1137/S0036142901384162Search in Google Scholar

[2] S. Badia, R. Codina, T. Gudi and J. Guzmán, Error analysis of discontinuous Galerkin methods for the Stokes problem under minimal regularity, IMA J. Numer. Anal. 34 (2014), no. 2, 800–819. 10.1093/imanum/drt022Search in Google Scholar

[3] A. Baran and G. Stoyan, Gauss–Legendre elements: A stable, higher order non-conforming finite element family, Computing 79 (2007), no. 1, 1–21. 10.1007/s00607-007-0219-1Search in Google Scholar

[4] D. Boffi, F. Brezzi and M. Fortin, Mixed Finite Element Methods and Applications, Springer Ser. Comput. Math. 44, Springer, Heidelberg, 2013. 10.1007/978-3-642-36519-5Search in Google Scholar

[5] S. C. Brenner and L. R. Scott, The Mathematical Theory of Finite Element Methods, 3rd ed., Texts Appl. Math. 15, Springer, New York, 2008. 10.1007/978-0-387-75934-0Search in Google Scholar

[6] H. Brezis, Functional Analysis, Sobolev Spaces and Partial Differential Equations, Universitext, Springer, New York, 2011. 10.1007/978-0-387-70914-7Search in Google Scholar

[7] B. Cockburn, G. Kanschat and D. Schötzau, A note on discontinuous Galerkin divergence-free solutions of the Navier–Stokes equations, J. Sci. Comput. 31 (2007), no. 1–2, 61–73. 10.1007/s10915-006-9107-7Search in Google Scholar

[8] M. Crouzeix and R. S. Falk, Nonconforming finite elements for the Stokes problem, Math. Comp. 52 (1989), no. 186, 437–456. 10.1090/S0025-5718-1989-0958870-8Search in Google Scholar

[9] M. Crouzeix and P.-A. Raviart, Conforming and nonconforming finite element methods for solving thestationary Stokes equations. I, Rev. Française Automat. Informat. Rech. Opér. Sér. Rouge 7 (1973), 33–75. 10.1051/m2an/197307R300331Search in Google Scholar

[10] D. A. Di Pietro and A. Ern, Mathematical Aspects of Discontinuous Galerkin Methods, Math. Appl. (Berlin) 69, Springer, Heidelberg, 2012. 10.1007/978-3-642-22980-0Search in Google Scholar

[11] A. Ern and J.-L. Guermond, Theory and Practice of Finite Elements, Appl. Math. Sci. 159, Springer, New York, 2004. 10.1007/978-1-4757-4355-5Search in Google Scholar

[12] J. Guzmán, A. Lischke and M. Neilan, Exact sequences on Powell-Sabin splits, Calcolo 57 (2020), no. 2, Paper No. 13. 10.1007/s10092-020-00361-xSearch in Google Scholar

[13] J. Guzmán and M. Neilan, 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. 10.1137/17M1153467Search in Google Scholar

[14] P. Hansbo and M. G. Larson, Discontinuous Galerkin methods for incompressible and nearly incompressible elasticity by Nitsche’s method, Comput. Methods Appl. Mech. Engrg. 191 (2002), no. 17–18, 1895–1908. 10.1016/S0045-7825(01)00358-9Search in Google Scholar

[15] C.-J. Heine, D. Köster, O. Kriessl, A. Schmidt and K. Siebert, ALBERTA: An adaptive hierarchical finite element toolbox, http://www.alberta-fem.de. Search in Google Scholar

[16] V. John, A. Linke, C. Merdon, M. Neilan and L. G. Rebholz, On the divergence constraint in mixed finite element methods for incompressible flows, SIAM Rev. 59 (2017), no. 3, 492–544. 10.1137/15M1047696Search in Google Scholar

[17] C. Kreuzer and P. Zanotti, Quasi-optimal and pressure robust discretizations of the Stokes equations by new augmented Lagrangian formulations, IMA J. Numer. Anal. (2019), 10.1093/imanum/drz044. 10.1093/imanum/drz044Search in Google Scholar

[18] P. L. Lederer, A. Linke, C. Merdon and J. 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. 10.1137/16M1089964Search in Google Scholar

[19] A. 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. 10.1016/j.cma.2013.10.011Search in Google Scholar

[20] R. H. Nochetto and J.-H. Pyo, Optimal relaxation parameter for the Uzawa method, Numer. Math. 98 (2004), no. 4, 695–702. 10.1007/s00211-004-0522-0Search in Google Scholar

[21] A. Schmidt and K. G. Siebert, Design of Adaptive Finite Element Software, Lect. Notes Comput. Sci. Eng. 42, Springer, Berlin, 2005. Search in Google Scholar

[22] 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), no. 1, 111–143. 10.1051/m2an/1985190101111Search in Google Scholar

[23] A. Toselli, hp discontinuous Galerkin approximations for the Stokes problem, Math. Models Methods Appl. Sci. 12 (2002), no. 11, 1565–1597. 10.1142/S0218202502002240Search in Google Scholar

[24] A. Veeser, Approximating gradients with continuous piecewise polynomial functions, Found. Comput. Math. 16 (2016), no. 3, 723–750. 10.1007/s10208-015-9262-zSearch in Google Scholar

[25] A. Veeser and P. Zanotti, Quasi-optimal nonconforming methods for symmetric elliptic problems. I—Abstract theory, SIAM J. Numer. Anal. 56 (2018), no. 3, 1621–1642. 10.1137/17M1116362Search in Google Scholar

[26] A. Veeser and P. Zanotti, Quasi-optimal nonconforming methods for symmetric elliptic problems. III—Discontinuous Galerkin and other interior penalty methods, SIAM J. Numer. Anal. 56 (2018), no. 5, 2871–2894. 10.1137/17M1151675Search in Google Scholar

[27] R. Verfürth and P. Zanotti, A quasi-optimal Crouzeix–Raviart discretization of the Stokes equations, SIAM J. Numer. Anal. 57 (2019), no. 3, 1082–1099. 10.1137/18M1177688Search in Google Scholar

Received: 2020-02-25
Revised: 2020-06-15
Accepted: 2020-06-29
Published Online: 2020-08-05
Published in Print: 2021-04-01

© 2021 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 7.5.2024 from https://www.degruyter.com/document/doi/10.1515/cmam-2020-0023/html
Scroll to top button