The inf-sup constant for -Crouzeix-Raviart triangular elements
Abstract
In this paper, we consider the discretization of the two-dimensional stationary Stokes equation by Crouzeix-Raviart elements for the velocity of polynomial order on conforming triangulations and discontinuous pressure approximations of order . We will bound the inf-sup constant from below independent of the mesh size and show that it depends only logarithmically on . Our assumptions on the mesh are very mild: for odd we require that the triangulations contain at least one inner vertex while for even we assume that the triangulations consist of more than a single triangle.
AMS Subject Classification: 65N30, 65N12, 76D07, 33C45,
Key Words:Non-conforming finite elements, Crouzeix-Raviart elements, high order finite elements, Stokes equation.
1 Introduction
In this paper we consider the numerical discretization of the two-dimensional stationary Stokes problem by Crouzeix-Raviart elements. They were introduced in the seminal paper [16] in 1973 by Crouzeix and Raviart with the goal to obtain a stable and economic discretization of the Stokes equation. They can be considered as an non-conforming enrichment of conforming finite elements of polynomial degree for the velocity and discontinuous pressures of degree . It is well known that the conforming pair of finite elements can be unstable; for two-dimensions the proof of the inf-sup stability of Crouzeix-Raviart discretizations of general order has been evolved over the last 50 years, the inf-sup stability for has been proved in [16] and only recently the last open case , has been proved in [11]. We mention the papers [21], [32], [15], [7], [27], [10] which contain essential milestones in this development. There is a vast of literature on various further aspects of Crouzeix-Raviart elements; we omit to present a comprehensive review here but refer to the overview article [9] instead.
Since higher order methods are becoming increasingly popular a natural question arises how the inf-sup constant depends on the polynomial degree . It is the goal of this paper to investigate this dependency.
The paper is organized as follows. In Section 1.1 we introduce the Stokes problem and the Crouzeix-Raviart discretization of polynomial order . We state our main theorem that the discrete inf-sup constant can be estimated from below by , where the positive constant depends only on the shape-regularity of the mesh and on the maximal outer angle of the domain . The explicit value of depends on the mesh topology. The simplest case is that each triangle in triangulation contains at least one inner vertex and then holds. We will give the value of also for more general triangulations.
The proof is given in Section 2. The key ingredient is to show that for any discrete pressure , there exists a velocity field from the Crouzeix-Raviart space such that belongs to the Scott-Vogelius pressure space [37, R.1, R.2], [32, R.1, R.2] (see (2.25)) and depends continuously on . These properties allow us to construct a conforming velocity field of order with . For this step, the construction in [32, Thm. 5.1], [27, Thm. 1] is modified such that the norm of the right-inverse does not deteriorate if a triangle vertex is a “nearly-critical” point – a notion which will be introduced in Definition 2.11. This key result is proved for odd polynomial degree in Section 2.2 and for even polynomial degree in Section 2.3.
In the conclusions (Sec. 3) we summarize our main findings and compare our results with existing results in the literature on some other pairs of finite elements for the Stokes equation.
In the appendices, we prove a technical result for a Gram matrix related to the bilinear form applied to the Crouzeix-Raviart element and discontinuous pressure space (see §A), an estimate of traces of non-conforming Crouzeix-Raviart basis functions (see §B), give some explicit formulae for integrals related to orthogonal polynomials (see §C), and prove a discrete Friedrichs inequality for Crouzeix-Raviart spaces (see §D).
1.1 The Stokes problem and its numerical discretization
Let denote a bounded polygonal Lipschitz domain with boundary . For a vertex in , we denote by the exterior angle between the two segments in with joint . The minimal outer angle at the boundary vertices is given by
(1.1) |
and satisfies since is Lipschitz. We consider the Stokes equation
with Dirichlet boundary conditions for the velocity and a usual normalization condition for the pressure
To formulate this equation in a variational form we first introduce the relevant function spaces. Throughout the paper we restrict to vector spaces over the field of real numbers.
For , , denote the classical Sobolev spaces of functions with norm . As usual we write instead of and for . For , we denote by the closure of the space of infinitely smooth functions with compact support in with respect to the norm. Its dual space is denoted by . For the pressure , the space will be relevant.
The scalar product and norm in are denoted respectively by
Vector-valued and tensor-valued analogues of the function spaces are denoted by bold and blackboard bold letters, e.g., and .
The scalar product and norm for vector valued functions are given by
where denotes the Euclidean scalar product in . In a similar fashion, we define for the scalar product and norm by
where . We also need fractional order Sobolev norms on boundary of triangles and introduce the relevant notation; for details see, e.g., [29]. For a bounded Lipschitz domain with boundary , let and denote the usual Lebesgue and Sobolev space on with norm and . For the fractional Sobolev space on of order is denoted by and equipped with the norm
and seminorm
We introduce the bilinear form by
(1.2) |
where and denote the derivatives of and . The variational form of the Stokes problem is given by: For given
(1.3) |
It is well-known (see, e.g., [22]) that (1.3) is well posed. Since we consider non-conforming discretizations we restrict the space for the right-hand side to a smaller space and assume from now on for simplicity that for some ; for a more general setting we refer to [35], [36].
In the following a discretization for problem (1.3) is introduced. Let denote a triangulation of consisting of closed triangles which are conforming: the intersection of two different triangles is either empty, a common edge, or a common point. We also assume , where
(1.4) |
and denotes the interior of a set .
Piecewise versions of differential operators such as and are defined for functions and vector fields which are sufficiently smooth in the interior of the triangles by
The values on are arbitrary since has measure zero.
An important measure for the quality of a finite element triangulation is the shape-regularity constant given by
(1.5) |
with the local mesh width and denoting the diameter of the largest inscribed ball in . The global mesh width is .
Remark 1.1
It is well known that the shape-regularity implies that there exists some minimal angle depending only on such that every triangle angle in is bounded from below by . In turn, every triangle angle in is bounded from above by .
The set of edges in is denoted by , while the subset of boundary edges is ; the subset of inner edges is given by . For each edge we fix a unit vector orthogonal to with the convention that is the outer normal vector for boundary edges .
The set of triangle vertices in is denoted by , while the subset of inner vertices is and . For , the set of its vertices is denoted by . For , we define the edge patch by
For , the nodal patch is defined by
(1.6) |
with local mesh width . For , we set
(1.7) |
For a subset , we denote by its convex hull; in this way an edge with endpoints can be written as .
Let and . For , we employ the usual multiindex notation for and points
Let denote the space of -variate polynomials of maximal degree , consisting of functions of the form
for real coefficients . Formally, we set . To indicate the domain explicitly in notation we write sometimes for and skip the index since it is then clear from the argument .
We introduce the following finite element spaces
(1.8) |
Furthermore, let
The vector-valued versions are denoted by and . Finally, we define the Crouzeix-Raviart space by
(1.9a) | ||||
(1.9b) |
Here, denotes the jump of across an edge
and is the space of polynomials of maximal degree with respect to the local variable in .
We have collected all ingredients for defining the Crouzeix-Raviart discretization for the Stokes equation. For , let the discrete velocity space and pressure space be defined by
Then, the discretization is given by: find such that
(1.10) |
where the bilinear forms and are given by
It is well known that problem (1.10) is well-posed if (i): the bilinear form is coercive and (ii): satisfies the inf-sup condition.
To verify the condition (i) we introduce, for a conforming triangulation of the domain , the broken Sobolev space
and define, for , the broken –seminorm by
In [16, Lem. 2]) it is proved that defines a norm in which is equivalent to the norm with equivalence constants independent of the polynomial degree and the mesh width (see Theorem D.1). This directly implies the coercivity of :
Hence, well-posedness of (1.10) follows from the inf-sup condition for .
Definition 1.2
Let denote a conforming triangulation for . The pair is inf-sup stable if there exists a constant such that
(1.11) |
We are now in the position to formulate our main theorem.
Theorem 1.3
Let be a bounded polygonal Lipschitz domain and let denote a conforming triangulation of consisting of more than a single triangle. Let . If is odd we assume that contains at least one inner vertex. Then, the inf-sup condition (1.11) holds:
(1.12) |
for a constant depending only on the shape-regularity of the mesh and on the maximal outer angle . In particular is independent of the mesh width and the polynomial degree . The value of is given by
where depends only on the mesh topology via the number of steps involved in the step-by-step construction introduced in (2.13).
Proof. The estimate follows, for from [16], for from [12, Thm. 3.1], for even from [7], for odd from [10], and for from [11]. We set
Estimate (1.12) for some for odd is proved in Section 2.2, Lem. 2.22, while the estimate for some for even is proved in Section 2.3. Both constants , depend only on the shape-regularity of the mesh and . Hence, .
2 Proof of Theorem 1.3
In this section, we will analyse the -dependence of the inf-sup constant in the form (1.12), first for odd polynomial degree and then for even degree .
2.1 Barycentric coordinates and basis functions for the velocity
In this section, we introduce basis functions for the finite element spaces in Section 1.1. We begin with introducing some general notation.
Notation 2.1
For vectors , , we write for the matrix with column vectors . For we set . Let be the -th canonical unit vector in .
For , is the Euclidean vector norm while the induced matrix norm is given for by .
Vertices in a triangle are numbered counterclockwise. In a triangle with vertices , , the angle at is called . If a triangle is numbered by an index (e.g., ), the angle at is called . For quantities in a triangle as, e.g., angles , , we use the cyclic numbering convention and .
For a -dimensional measurable set we write for its measure; for a discrete set, say , we denote by its cardinality.
In the proofs, we consider frequently nodal patches for inner vertices . The number denotes the number of triangles in . Various quantities in this patch such as, e.g., the triangles in , have an index which runs from to . Here, we use the cyclic numbering convention and and apply this analogously for other quantities in the nodal patch.
Let the closed reference triangle be the triangle with vertices , , . The nodal points on the reference element of order are given by
For a triangle , we denote by an affine bijection. The mapped nodal points of order on are given by
Nodal points of order on are defined by
We introduce the Lagrange basis for the space , which is indexed by the nodal points and characterized by
(2.1) |
where is the Kronecker delta. A basis for the space is given by , .
Let denote a triangle with vertices , , and let be the barycentric coordinate for the node defined by
(2.2) |
If the numbering of the vertices in is fixed, we write short for . For the barycentric coordinate on the reference element for the vertex we write , . Elementary calculation yield (see, e.g., [10, Appendix A])
(2.3) |
where is the edge of opposite to , the outward unit normal at , and the angle in at .
Definition 2.2
Let denote the usual univariate Legendre polynomial of degree (see [17, Table 18.3.1]). Let be even and . Then, the non-conforming triangle bubble is given by
For odd and , the non-conforming edge bubble is given by
(2.4) |
where denotes the vertex in opposite to .
Different representations of the functions , exist in the literature, see [34], [5], [12, for ], [13] while the formula for has been introduced in [7] and the one for in [10].
Proposition 2.3
A basis for the space is given
-
1.
for even by
-
2.
for odd by
The proof of this proposition and the following corollary can be found, e.g., in [34, Rem. 3], [13, Thm. 22], [10, Cor. 3.4].
Corollary 2.4
A basis for the space is given
-
1.
for even by
(2.5) -
2.
for odd by
(2.6)
Remark 2.5
The original definition of Crouzeix-Raviart spaces by [16] is implicit and given for conforming simplicial finite element meshes in , . For their practical implementation, a basis is needed and Corollary 2.4 provides a simple definition. A basis for Crouzeix-Raviart finite elements in is introduced in [20] for , a general construction is given in [14], and a basis for a minimal Crouzeix-Raviart spaces in general dimension is presented in [30].
2.2 The case of odd
In this section, we assume for the following
(2.7) |
This section is structured as follows. In §2.2.1 we generalize the concept of critical points (see [37], [32]) to -critical points which turn out to be essential for estimates with constants depending on the mesh only via the shape-regularity constant and . We split these -critical points into a set of “obtuse” critical points and “acute” critical points. In §2.2.2, we provide the proof of Theorem 1.3 for a maximal partial triangulation that does not contain acute critical points and satisfies (2.7). Finally, in §2.2.3 we present the argument to allow for acute critical points.
2.2.1 Geometric preliminaries
For the analysis of the inf-sup constant we start with the definition of critical points (see [37], [32]).
Definition 2.6
Let denote a triangulation as in §1.1. For , let
The point is a critical point for if there exist two straight infinite lines , in such that all edges satisfy . The set of all critical points in is .
Remark 2.7
Geometric configurations where critical points occur are well studied in the literature (see, e.g., [32]). Any critical point belongs to one of the following cases (see Fig. 1):
-
1.
and consists of four triangles and is the intersections of the two diagonals in the quadrilateral .
-
2.
and , i.e., both edges are boundary edges with joint .
-
3.
and and two edges are boundary edges which lie on a straight boundary piece.
-
4.
and and each of the two boundary edges is aligned with one edge of .
Definition 2.8
Remark 2.9
It is easy to see that if and only if .
Lemma 2.10
Proof. Let and consider an edge . Then, there are two triangles which are adjacent to . The angle in resp. at is denoted by resp. .
1st case. Let or . Then, we conclude from Remark 1.1 that
For the left inequality to hold, the minimal angle must satisfy while for the right inequality, it must hold . For , the 1st case is empty. For we get
Since this case cannot appear.
2nd case. Let . The condition implies that with
(2.8) |
Consequently the two angles in and in at satisfy
where (resp. ) denotes the third angle in (resp. ). Hence, in this case
Definition 2.11
Let be as in Lemma 2.10. For , the set of -critical points is given by
A point is called a nearly critical point. An -critical point is isolated if all edge satisfy: is not an -critical point.
By perturbing the geometric configurations in Remark 2.7 we obtain the following subcases (see Fig. 2).
Definition 2.12
Let be as in Lemma 2.10 and . If satisfies
-
1.
and and . Then is an inner -critical point. Let
-
2.
and . Then is an acute critical point. Let111Note that the set of acute critical points is independent of .
-
3.
and and . Then is flat -critical point. Let
-
4.
and and . Then is a (locally) concave -critical point. Let
The acute critical points require some special treatment and we denote the union of the others by
The following lemma states that for a possibly adjusted , still depending only on the shape-regularity of the mesh and the maximal outer angle , the -critical points belong to one of the four categories described in Definition 2.12.
Lemma 2.13
Let be a conforming triangulation such that is a Lipschitz domain.
Then, there exists some depending only on the shape-regularity of the mesh and the minimal outer angle such that for any -critical point belongs to one of the four categories described in Definition 2.12.
Proof. Let be an -critical point. We set and choose a counterclockwise numbering for the triangles in , i.e., , . The shape-regularity of the mesh implies that there is some depending only on such that . Denote by the angle in at . Let which will be fixed later and assume . Since is an -critical it holds
The shape-regularity implies and, for , we get
(2.9) |
Since is monotonously increasing with and , we can select such . In turn, the first and last interval in (2.9) are empty and
(2.10) |
Case 1: .
In this case we obtain
(2.11) |
By adjusting such that we conclude that and . Hence, is an inner -critical point according to Definition 2.12(1).
Case 2a: and .
These cases correspond to acute critical/flat -critical/concave -critical points according to Definition 2.12(2-4).
Case 2b: and .
We argue as in Case 1 but take into account that the patch is not “closed” since is a boundary point. Let be the “outer angle” of the domain at . Then
By the definition of , we obtain
where if is even and if is odd. By rearranging the terms we get
(2.12) |
We adjust such that satisfies and, in turn, . Then, it is easy to verify that
holds for all . Hence, (2.12) cannot hold and there exists no -critical boundary point for .
Next, we collect the critical points in pairwise disjoint, edge-connected sets which we will define in the following.
We say two points are edge-connected if there is an edge with endpoints . A subset is edge-connected if there is a numbering of the points in such that , are edge-connected for all . A point is edge-connected to if or there is such that , are edge-connected.
From Lemma 2.10 we know that two edge-connected points can be both critical only if the connecting edge belongs to ; in this case it holds . Next, we will group the points in into subsets called fans.
From Lemma 2.10 it follows that the points in are isolated (see Def. 2.11). All other -critical points lie on the boundary. Next, we define mappings , , and . The construction is illustrated in Figure 3.
For , Definition 2.12 implies that and hence . We fix one edge and set . Note that the choice of is unique for . For the choice is arbitrary. For , the set consists of two edges, say , . We fix one of them and set . Let be such that . Then . Lemma 2.10 implies that is not an -critical point. A unit vector orthogonal to is defined by the condition that and form a right-handed system.
Definition 2.14
We decompose into disjoint fans , , such that the following conditions are satisfies
-
1.
-
2.
for any , the set is edge-connected,
-
3.
for any , there is such that for all it holds and, vice versa:
-
4.
any which is edge-connected to some and satisfies belongs to .
The following lemma will allow us to construct a right-inverse for the divergence operator separately for each fan.
Lemma 2.15
Let be as in in Lemma 2.10 and let be fixed.
-
a.
Then, the mapping is injective.
-
b.
For , let . The domains have pairwise disjoint interior.
Proof. Part a. The injectivity of the mapping follows from Lemma 2.10: if and is such that then .
Part b. The following construction is illustrated in Figure 4.
For fixed , we number the points in by , , such that and are edge-connected for all and and are the endpoints in the polygonal line through these points. Let be the triangle with vertices , , , and let . Note that this set is empty if . Let be two different triangles such that and . Let be the third vertex in and observe that it does not belong to . Since is an inner edge and is an -critical point, Lemma 2.10 implies that is not an -critical point. Next, we show that does not belong to ; from this the assertion follows. We assume and derive a contradiction. Since , this assumption implies that . If then Definition 2.14(4) implies that and this is a contradiction. If , then .
Since the acute critical points need some special treatment we define a sequence of triangulations , , with the properties
-
1.
(2.13) -
2.
is a maximal subset of such that ,
-
3.
for ,
By this step-by-step procedure, triangles are attached to a previous triangulation which have an edge in common with the set of edges in . The proof of Theorem 1.3 under assumption (2.7) then consists of first proving the inf-sup stability for and then to investigate the effect of attaching a triangle to an inf-sup stable triangulation. A sufficient condition for is that every triangle in has an interior point.
2.2.2 The case
In this section, we prove the inf-sup stability for the triangulation in (2.13) where . For simplicity we skip the index and write , , etc.
Next, we define some fundamental non-conforming Crouzeix-Raviart vector fields which will be used to eliminate the critical pressures in the Stokes element .
Essential properties of the Crouzeix-Raviart function are: it is a polynomial of degree on each and a Legendre polynomial on each edge so that the jump relations in (1.9b) are satisfied. Furthermore, and .
In the first step, we modify the function by adding a conforming edge bubble in such that the norm of the modified function has an improved behaviour with respect to .
Let with endpoints , . Set and consider a function with and
(2.14a) | ||||
(2.14b) |
Then, also belongs to the space and
(2.15) |
The last relation can be derived from the following reasoning. For and , set for all . Let be arbitrary. Clearly, for some . The conditions in (2.14) imply that for it holds
Hence,
(2.16) | ||||
Since was arbitrary, (2.15) follows.
Lemma 2.16
Let be odd and for , let be as in (2.4). Then, there exists a function with
-
i.
-
ii.
for all , for all :
(2.17) -
iii.
for all , for all :
-
iv.
for all
(2.18) -
v.
The piecewise gradient is bounded:
(2.19)
Proof. We employ the reference triangle as in [6] in order to apply the polynomial extension theorem therein. Let be the equilateral triangle with vertices , , and let denote the edge in opposite to , .
Let with endpoints , , and let . Choose an affine pullback such that . We employ the function given by
with the Jacobi polynomials (see, e.g., [17, §18.3]) and the normalisation factor . These functions have been analysed in the proof of Lemma A.1 in [3, denoted by ] and we recall relevant properties. It holds and (cf. [17, §18.3]). Their norms can be estimated by
We set
(2.20) |
Clearly, it holds
By using Lemma C.1, we get
From [6, Thm. 7.4] we conclude that there is with
which satisfies
for a constant independent of . Lemma B.2 implies the following estimate of the norm of :
(2.21) |
In turn, we get
(2.22) |
By using the affine lifting to the triangle we define the function by
This function is continuous across (with value ) and, on , it is a lifted Legendre polynomial. This implies property (iii) for . The function vanishes outside so that (i) holds. Since the construction implies that the derivative of in the direction of , evaluated at the endpoints , of , is zero, we may apply the reasoning in (2.16) to obtain property (ii) for .
From (2.22) we obtain by the transformation rule for integrals and the chain rule for differentiation
Next, we modify such that property (iv) holds without affecting the other properties. Let with and
(2.23) |
The modified function finally is defined by
(2.24) |
Since property (iv) follows by construction. The gradient vanishes in the vertices of so that for all and (ii) is inherited from . Properties (i), (iii) are obvious. Next, we verify (v). Let denote the vertex in opposite to . We first compute
In this way, and an inverse inequality for quartic polynomials gives us
Hence, property (iv) follows.
Next we recall a result which goes back to Vogelius [37] and Scott-Vogelius [32], see also [27, Proof of Thm. 1].
Definition 2.17
Let be as in Lemma 2.10. For , the subspace of the pressure space is given by
(2.25) |
where, for , the functional is as follows: fix the counterclockwise numbering , , of the triangles in the patch by the condition and set
(2.26) |
Note that is the pressure space introduced by Vogelius [37] and Scott-Vogelius [32] and the following inclusions hold: for
with the pressure space in [27, p. 517].
For the Scott-Vogelius pressure space , the existence of a continuous right-inverse of the divergence operator into was proved in [37] and [32].
Proposition 2.18 (Scott-Vogelius)
For any there exists some such that
for a constant which only depends on the shape-regularity of the mesh, the polynomial degree , and on , where
(2.27) |
In particular, the constant is independent of .
In Lemma 2.20, we will show that, by subtracting the divergence of a suitable Crouzeix-Raviart velocity from a given pressure in , the resulting modified pressure belongs to the reduced pressure space . As a preliminary, we need a bound of the functional in (2.26) which is explicit with respect to the local mesh size and polynomial degree.
Lemma 2.19
There exists a constant which only depends on the shape-regularity of the mesh such that
for any .
Proof. Let and . The affine pullback to the reference triangle is denoted by . For , let and . Then
(2.28) |
A summation over all leads to
where only depends on the shape-regularity of the mesh.
The following lemma shows that the non-conforming Crouzeix-Raviart elements allow us to modify a general pressure in such a way that the result belongs to provided .
Lemma 2.20
Proof. Let . Let the fans , , be as in Definition 2.14. For each fan we employ an ansatz
(2.32) |
for as in (2.24), where the coefficients are defined next. The global function is then given by
Property (2.29) follows from this ansatz by using Lemma 2.16. Next, we define the coefficients in (2.32) such that (2.30) holds and prove the norm estimates for . Our construction of the fans implies that open interiors of the supports of are pairwise disjoint (see Lem. 2.15); as a consequence the definition of and the estimate of can be performed for each fan separately.
For , let , . Let , , denote the triangles in with the convention that points into . The vertex in opposite to is denoted by . We use
and compute the divergence of
(2.33) |
Well-known properties of Legendre polynomials applied to (2.33) imply that for any vertex of and odd polynomial degree
(2.34) |
Hence the condition for all is equivalent to the system of linear equation
(2.35) |
with
(2.36) |
The matrix is explicitly given by
We use (cf. Fig. 5)
In Lemma A.1, we will prove that the matrix is invertible and the inverse is bounded by a constant independent of and . Hence,
(2.44) |
Let and . We estimate the function in (2.32) by
The constant only depends on the shape-regularity of the mesh. We use Lemma 2.19 and conclude that
Since the interiors of the supports have pairwise empty intersection the estimate
follows.
Definition 2.21
For the proof of the following lemma we recall the definitions of some linear maps from the literature which are related to the right inverse of the divergence operator acting on some polynomial spaces.
Bernardi and Raugel introduced in [8, Lem. II.4] a linear mapping with the property: for any , the function satisfies
(2.45) |
and
(2.46) |
for a constant which is independent of the mesh width and the polynomial degree.
Next we consider some right inverse of the divergence operator on the space
Let
There exists a linear operator such that for all
(2.47) |
where the constant is independent of the mesh width and the polynomial degree. Note that in the original paper [37, Lem. 2.5] by M. Vogelius, the right-hand side in the estimate (2.47) contains an additional factor for some positive (independent of the mesh width). In [3, Thm. 3.4], the operator in [37, Lem. 2.5] is modified and the estimate in the form (2.47) is proved for the modified operator.
Finally, we reconsider the linear operator introduced by Guzmán and Scott in [27, Proof of Lem. 6 and Lem. 7] with the property that, for any , it holds
(2.48a) | ||||
(2.48b) |
We emphasize that in [27, Lemma 7] the constant (cf. (2.27)) instead of appears in (2.48b) so that the estimate of for deteriorates in cases where the is a nearly critical point, i.e., very close to the geometric situations described in Remark 2.7. The proof of [27, Lemma 7] is split into an estimate related to points with (cf. (2.26)) and an estimate for the remaining points with . Only in this second part, the constant is involved. The result has been improved in [24, Lem. 4.5] and it was shown that there is an operator such that the properties in (2.48) hold for : for any it holds
From Lemma 2.20, we conclude that and the second part of the proof in [27, Lemma 7] is applied only to points with
Hence, (2.48b) follows for depending only on the shape-regularity of the mesh.
Lemma 2.22
Let assumption (2.7) be satisfied and let . There exists a constant which only depends on the shape-regularity of the mesh and as in (1.1) such that for any fixed and any , there exists some such that
and
The constant only depends on the shape-regularity of the mesh and but is independent of the mesh width and the polynomial degree .
Proof. For the construction of we follow and modify the lines of proof in [27, Thm. 1] by a) involving the operator and b) employing the concept of -critical points.
For given , we employ the operators , , , in the definition of the function
(2.49) | ||||
By construction we have
The first two summands in (2.49) satisfy
(2.50) | ||||
(2.51) | ||||
For the third term in (2.49) we get
(2.52) |
The combination with (2.46), (2.31) leads to
(2.53) |
For the fourth term we get in a similar way
(2.54) |
The combination of (2.50), (2.51), (2.53), (2.54) with (2.49) leads to the assertion.
2.2.3 The case
In this section, we remove the condition and construct a bounded right-inverse of the piecewise divergence operator for odd and conforming triangulations which contain at least one inner point. The construction is based on the step-by-step procedure (cf. (2.13)) from the triangulation to . Inductively, we assume that there is a triangulation along a bounded right-inverse of the piecewise divergence operator. A single extension step is analysed by the following lemma.
Lemma 2.23
Let denote a conforming triangulation for the domain and let be a subset such that every triangle has one edge, say , which belongs to . We assume that has at least one inner vertex and set . Assume that there exists a bounded linear operator with on and
Then, there exists a linear operator with on and
for a constant which depends only on the shape-regularity of the mesh and on .
Proof. Let . We set where the operator is as in (2.45) and satisfies
Hence, belongs to and has trianglewise integral mean zero.
The following construction is illustrated in Figure 6.
Let be such that there exists a non-empty subset having the property that any shares an edge with . We have ; indeed, if , then all three edges of are boundary edges which implies and violates the condition that must contain an inner vertex.
For , let denote the vertex in opposite to and set . The endpoints of are denoted by , . We employ the ansatz (cf. (2.24))
(2.55) |
with the convention that is the unit vector orthogonal to and directed into . By construction it holds and . We determine in (2.55) such that and employ (2.34) to get
Hence and we conclude as in the proof of Lemma 2.20 that
We set
and note that
(2.56) |
The construction implies , has trianglewise integral mean zero, and . Next, we employ the vector field defined in [24, Lem. 4.9] which is a modification of the cubic vector field defined in [27, (3.5)] but allows for better -explicit estimates. We recall the relevant lemma from [24] for the existence of such vector fields and collect important properties.
Lemma 2.24 ([24, Lem. 4.9])
Let be a conforming triangulation of and let . Let with endpoints , . Then there exist vector fields , , with the following properties
(2.57) |
We employ this vector field to the edge , set
and define
(2.58) |
The function has trianglewise integral mean zero and vanishes in all vertices of . The norm can be estimated in the same way as the function in (2.53); however the factor does not appear as in (2.53) since the last estimate in (2.57) does not depend on these quantities. In this way, we get
(2.59) | ||||
(2.60) |
Hence, from [3, Thm. 3.4] we deduce that there exists with such that on and
In this way we have constructed the function by
such that on , and
where only depends on the shape-regularity of the mesh and through the constants , , . Let and note that by construction
and
Finally, the linear map is defined by
and satisfies on and
for some which only depends on the shape-regularity of the mesh and .
Theorem 2.25
Let be a conforming triangulation which contains at least one interior vertex. Let be odd and let the number of steps in the construction (2.13). Then, the inf-sup constant for the corresponding Crouzeix-Raviart discretization satisfies
for some constant which depends only on the shape-regularity of the mesh and . If every triangle in has an interior vertex, then .
2.3 The case of even
This case is slightly simpler that the case of odd since the non-conforming Crouzeix-Raviart functions for even have smaller support (i.e., one triangle) compared to two triangles (which share an edge) for odd .
In this section, we assume
(2.61) |
Remark 2.26
It is easy to verify that implies that there exists a mapping with and not all vertices of are -critical.
For an -critical point , let and fix a counterclockwise numbering of the triangles in
(2.62) |
such that and share an edge for all . With this notation at hand, the functional is given by
Lemma 2.27
Let assumption (2.61) be satisfied. There exists a constant which only depends on the shape-regularity of the mesh and such that for any fixed and any , there exists some such that
(2.63) | ||||
(2.64) |
and
(2.65) |
The constant depends only on the shape-regularity of the mesh and .
Proof. For a triangle , we set and . Their cardinalities are denoted by and . Note that (cf. Rem. 2.26). We number the vertices in with the convention and . The angle in at is denoted by . Let be the edge in opposite to and let denote the outward unit normal vector at .
In a similar way as for the construction of in Lemma 2.16 (and employing Lemma B.1 instead of Lemma B.2 for (2.21)) there exists a function with
-
1.
-
2.
for all , for all
(2.66) -
3.
for all
(2.67) -
4.
for all and any
(2.68) -
5.
The piecewise gradient is bounded by
(2.69)
For , we define
(2.70) |
i.e., we fix and in (2.5). The divergence of evaluated at a vertex , , is given by
Let . We choose by the conditions for
(2.71) |
For , let be defined by (cf. (2.62)). Then,
for
We define by
so that is the solution of
Observe that
and due to the shape-regularity of the mesh. For the coefficient we get explicitly
(2.72) |
with an estimate
where only depends on the shape-regularity of the mesh. Note that this is the analogue for even to (2.44). If , it holds .
Next, we verify (2.64). Let and recall the notation and convention as in (2.62). Let . Then (2.64) follows from
The estimate
for a constant which only depends on the shape-regularity of the mesh and follows directly from (2.69) and the final estimate (2.65) is derived by repeating the arguments as in the proof of Lemma 2.20.
This lemma allows us to extend Definition 2.21 to the case of even by defining the coefficients by (2.72) and the functions by (2.70) and set (cf. (2.73)) . Since we may apply the further steps in the proof of Lemma 2.22 to obtain the inf-sup stability for even .
Theorem 2.28
Let be a conforming triangulation satisfies (2.61). Then, the inf-sup constant for the corresponding Crouzeix-Raviart discretization satisfies
for some constant which depends only on the shape-regularity of the mesh and .
3 Conclusion
In this paper, we have derived lower bounds for the inf-sup constant for Crouzeix-Raviart elements for the Stokes equation which are explicit with respect to the polynomial degree and are independent of the mesh size.
-
1.
The inf-sup constant can be bounded from below by if
-
(a)
for odd ,
-
i.
has at least one interior point and
-
ii.
for , has no acute critical point,
-
i.
-
(b)
for even , contains more than one triangle,
-
(a)
-
2.
If for odd , condition 1.a.ii. is not satisfied but a step-by-step construction (2.13) for some is possible, then, .
Finally, we compare these findings with some other stable pairs of Stokes elements on triangulations in the literature. The element has a discrete inf-sup constant which can be estimated from below by (see [31], [33]). The discrete inf-sup constant for the Scott-Vogelius element for can be estimated from below by for some integer sufficiently large (see [32], [37]). The pressure-wired Stokes element in LABEL:Sauter_eta_wired (again for ) is a mesh-robust generalization of the Scott-Vogelius element with a lower bound of the inf-sup constant of the form . In [3], a conforming stable pair of Stokes elements on triangulations is introduced and it is proved that the discrete inf-sup constant can be estimated from below by for a constant independent of and and . However, the implementation requires finite elements for the velocity with continuity at the triangle vertices and pressures which are continuous in the triangle vertices.
Acknowledgement 3.1
Thanks are due to Benedikt Gräßle, HU Berlin, for fruitful discussions on Lemma 2.13.
I am grateful to my colleagues from TU Vienna, Profs. Joachim Schöberl and Markus Melenk. Joachim showed by numerical experiments that the lower bound of the inf-sup constant in the first arxiv version of the paper, namely , might be too pessimistic and “it should be at least ” and Markus raised the suspicion that the interpolation argument might be too pessimistic for Legendre polynomials.
Appendix A The inverse of the matrix in (2.37)
Lemma A.1
Note that the matrix in (2.43) is the same as the matrix which has been analysed in [10, (3.36)]. In particular, the formula
was proved.
Next we show that the sum is bounded away from and . The bound follows from Remark 1.1. Since , are critical points the sum of both angles adjacent to at satisfy
We write . From the proof of Lemma 2.10, in particular from the estimate (2.8) we conclude that .
Since all points are edge-connected to the same point , the number is bounded from above by a constant which only depends on the shape-regularity of the mesh. Hence,
By adjusting the constant in Lemma 2.10 to it follows that
By using the trivial estimate , we may conclude that
(A.1) |
Note that the entries in the matrix (cf. 2.43) satisfy
(A.2) |
and hence the Frobenius norm can be estimated by
It is well known that and hence the bound on follows.
We combine (A.1), (A.2), and to obtain by Cramer’s rule that there exists a constant which only depends on the shape-regularity such that
By the same arguments as before we conclude that for a constant which only depends on the shape-regularity of the mesh.
Next we estimate . Since is diagonal it suffices to estimate the diagonal entries
We write and obtain
Next, we adjust the upper bound by setting to obtain with implies the invertibility of with bound
Appendix B Estimate of the norm of traces of non-conforming Crouzeix-Raviart functions
In this appendix, we prove the norm estimate (2.69) for and (2.19) for . We first introduce some norms and semi-norms on the unit interval in a formal way:
(B.1) |
Lemma B.1
Let be even and . Let be defined by (2.67). Then there exists an absolute constant depending only on the shape regularity of such that
Proof. We first prove the estimate for the reference element . By construction (see (2.67)) the function coincides with on . Let the vertices of be numbered counterclockwise and denoted by , . The edge opposite to is (with cyclic numbering convention and ). We choose the pullbacks to by
(B.2) |
and observe . From [25] (see also [4, p 1870]) we deduce that the norm is equivalent to
(B.3) |
where for :
(B.4) |
In our application (and even ) we have so that . We use to get
(B.5) |
This integral can be evaluated analytically for (see Lem. C.2) and we obtain
(B.6) |
for a generic constant . This leads to the final estimate on the reference element:
For a triangle , let be an affine pullback and set . Then, the transformation rule for integrals yields
(B.7) |
where only depends on the shape regularity of the mesh. The leads to the claim.
The case of odd is considered in the following lemma.
Lemma B.2
Let be odd. For , let the function be as in the proof of Lemma 2.16. Then there exists a generic constant depending only on the shape regularity of such that for any , it holds
Proof. Let and . Similarly as in (B.7) we have
with and affine pullback . Number the vertices in counterclockwise , , such that is opposite to . The edge in opposite to is denoted by and this implies . The edgewise pullbacks are defined as in (B.2). We employ the equivalence of the norm with (see (B.3)) and obtain
with
Note that
(B.8) |
with and as in (2.20). The antisymmetry of the Legendre polynomial for odd implies and
(B.9) |
The estimates for follow from (B.6). For the last term, we employ
The last term can be estimated by taking into account (obtained, e.g., by evaluating and differentiating [17, 18.5.8 for the choice and 18.5.10 for the choice ] at ):
For the third term in (B.9), we simply employ so that
The last integral can be evaluated analytically: By using the recurrence relation (cf. [17, 18.9.1])
we obtain
(B.10) |
This and the orthogonality properties of the Legendre polynomials lead to
(B.11) |
The integral in (B.11) will be evaluated in (C.1). We obtain
and the explicit recursion formula
with starting value
It is easy to verify that solves the recursion and hence,
From this, the assertion follows.
Appendix C Analytic evaluation of some integrals involving Legendre polynomials
In the proof of Lemma 2.20 some integrals over Legendre polynomials appear and we present here their explicit evaluation.
Lemma C.1
For , it holds
(C.1) |
and
(C.2) |
Proof. The relation follows from [23, 7.221(1)].
For , these relations follows from . Let . The recurrence relation in [17, 18.9.1, Table 18.9.1] imply
(C.3) |
Substituting under the integral in (C.1) by this and taking into account the orthogonality relations of the Legendre polynomials leads to
For the second integral (C.2) we employ integration by parts:
(C.4) |
Since the orthogonality properties of Legendre polynomials imply that the integral in the right-hand side of (C.4) is zero. By using (cf. [17, Table 18.6.1]) and (cf. [17, combine 18.9.15 with Table 18.6.1]) we get
Finally
The last integral is zero due the orthogonality of the Legendre polynomials. The endpoint values of and lead to the assertion.
In the final part of this section, we will compute the value of explicitly. We set
Lemma C.2
It holds
Proof. We write
with
Hence,
with
We perform the integration with respect to explicitly and get
with
From now on we restrict to the case . Since is a polynomial of maximal degree the second integral vanishes: . We apply recursively integration by parts and use the orthogonality of the Legendre polynomials to obtain
We interchange the ordering of the summation, introduce the new variable , and obtain
By using a telescoping sum argument, it is easy to verify that the inner sum equals
Hence,
(C.5) |
for
For and we get
(C.6) |
We use the endpoint formula for and for
This leads to and the Leibniz rule for differentiation yields
The last sum in (C.6) is the Taylor expansion of about , evaluated at , i.e.,
Hence,
(C.7) |
For the quantity we obtain by similar arguments
The combination of this with (C.5), (C.7) leads to
Since is a polynomial of maximal degree the orthogonality of leads to
(C.8) |
We employ the recursion formula in [17, Table 18.9.1] for
so that
(C.9) |
From (C.8) and , , we get
(C.10) |
Now, it is easy to verify that satisfies the recursion (C.9) and the initial value conditions (C.10).
Appendix D Norm equivalence for Crouzeix-Raviart spaces
It is well known that for , the norms and are equivalent. In this section, we state estimates for the constants in these equivalencies – the proof is a repetition of the well-known arguments for the case (see, e.g., [19, Lem. 36.6]).
Theorem D.1
There exists a constant depending only on the shape-regularity of the mesh and the domain such that
In particular is independent of the polynomial degree and the mesh size .
Proof. We prove this result only under the regularity assumption that the Poisson problem:
is regular. For less regularity we refer to [19, Lem. 36.6]. For , we have
(D.1) |
For , there exists some such that and for a constant which only depends on . Hence,
where is the unit normal vector pointing to the exterior of . Next, we rewrite the sum over the triangle boundaries as a sum over the edges. For we fix the direction of a unit vector which is orthogonal to and for let denote the unit vector, orthogonal to , pointing to the exterior of . Then
Let be fixed and let be the function with constant value . The orthogonality conditions of the Crouzeix-Raviart elements across edges (see (1.9b)) imply
(D.2) | ||||
We employ first a weighted trace inequality (see, e.g., [18, Lem. 12.15]) and then a Poincaré-Steklov estimate (see, e.g., [19, (12.17) for and .]) to get for
(D.3) |
where only depends on the shape-regularity of the mesh.
Next we estimate the jump of across . For , we define as the function with constant value on and observe . Hence,
for a constant which only depends on the shape-regularity of the mesh. For the estimate for follows in a similar fashion. The combination of (D.2) with (D.3) and the two trace estimates for leads to
Using this estimate in (D.1) finishes the proof.
References
- [1] M. Abramowitz and I. A. Stegun. Handbook of Mathematical Functions. Applied Mathematics Series 55. National Bureau of Standards, U.S. Department of Commerce, 1972.
- [2] M. Ainsworth and S. Jiang. Preconditioning the mass matrix for high order finite element approximation on triangles. SIAM J. Numer. Anal., 57(1):355–377, 2019.
- [3] M. Ainsworth and C. Parker. Mass conserving mixed -FEM approximations to Stokes flow. Part I: uniform stability. SIAM J. Numer. Anal., 59(3):1218–1244, 2021.
- [4] M. Ainsworth and C. Parker. Mass conserving mixed -FEM approximations to Stokes flow. Part II: optimal convergence. SIAM J. Numer. Anal., 59(3):1245–1272, 2021.
- [5] M. Ainsworth and R. Rankin. Fully computable bounds for the error in nonconforming finite element approximations of arbitrary order on triangular elements. SIAM J. Numer. Anal., 46(6):3207–3232, 2008.
- [6] I. Babuška, A. Craig, J. Mandel, and J. Pitkäranta. Efficient preconditioning for the -version finite element method in two dimensions. SIAM J. Numer. Anal., 28(3):624–661, 1991.
- [7] Á. Baran and G. Stoyan. Gauss-Legendre elements: a stable, higher order non-conforming finite element family. Computing, 79(1):1–21, 2007.
- [8] C. Bernardi and G. Raugel. Analysis of some finite elements for the Stokes problem. Math. Comp., 44(169):71–79, 1985.
- [9] S. C. Brenner. Forty years of the Crouzeix-Raviart element. Numer. Methods Partial Differential Equations, 31(2):367–396, 2015.
- [10] C. Carstensen and S. Sauter. Critical functions and inf-sup stability of Crouzeix-Raviart elements. Comput. Math. Appl., 108:12–23, 2022.
- [11] C. Carstensen and S. A. Sauter. Crouzeix-Raviart triangular elements are inf-sup stable. Math. Comp., 91(337):2041–2057, 2022.
- [12] Y. Cha, M. Lee, and S. Lee. Stable nonconforming methods for the Stokes problem. Appl. Math. Comput., 114(2-3):155–174, 2000.
- [13] P. G. Ciarlet, P. Ciarlet, Jr., S. A. Sauter, and C. Simian. Intrinsic finite element methods for the computation of fluxes for Poisson’s equation. Numer. Math., 132(3):433–462, 2016.
- [14] P. Ciarlet, Jr., C. F. Dunkl, and S. A. Sauter. A family of Crouzeix-Raviart finite elements in 3D. Anal. Appl. (Singap.), 16(5):649–691, 2018.
- [15] M. Crouzeix and R. Falk. Nonconforming finite elements for Stokes problems. Math. Comp., 186:437–456, 1989.
- [16] M. Crouzeix and P.-A. Raviart. Conforming and nonconforming finite element methods for solving the stationary Stokes equations. I. Rev. Française Automat. Informat. Recherche Opérationnelle Sér. Rouge, 7(R-3):33–75, 1973.
- [17] NIST Digital Library of Mathematical Functions. http://dlmf.nist.gov/, Release 1.0.13 of 2016-09-16. F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller and B. V. Saunders, eds.
- [18] A. Ern and J.-L. Guermond. Finite elements I—Approximation and interpolation. Springer, Cham, 2021.
- [19] A. Ern and J.-L. Guermond. Finite elements II—Galerkin approximation, elliptic and mixed PDEs. Springer, Cham, 2021.
- [20] M. Fortin. A three-dimensional quadratic nonconforming element. Numer. Math., 46(2):269–279, 1985.
- [21] M. Fortin and M. Soulie. A nonconforming quadratic finite element on triangles. International Journal for Numerical Methods in Engineering, 19:505–520, 1983.
- [22] V. Girault and P. Raviart. Finite Element Methods for Navier-Stokes Equations. Springer, Berlin, 1986.
- [23] I. S. Gradshteyn and I. Ryzhik. Table of Integrals, Series, and Products. Academic Press, New York, London, 1965.
- [24] B. Gräßle, N.-E. Bohne, and S. A. Sauter. The pressure-wired stokes element: a mesh-robust version of the Scott-Vogelius element, (extended edition), arXiv: 2212.09673, 2022.
- [25] P. Grisvard. Elliptic Problems in Nonsmooth Domains. Pitman, Boston, 1985.
- [26] J. Guzmán and M. Neilan. Conforming and divergence-free Stokes elements on general triangular meshes. Math. Comp., 83(285):15–36, 2014.
- [27] J. Guzmán and L. R. Scott. The Scott-Vogelius finite elements revisited. Math. Comp., 88(316):515–529, 2019.
- [28] G. Matthies and L. Tobiska. Inf-sup stable non-conforming finite elements of arbitrary order on triangles. Numer. Math., 102(2):293–309, 2005.
- [29] W. McLean. Strongly Elliptic Systems and Boundary Integral Equations. Cambridge, Univ. Press, 2000.
- [30] S. Sauter and C. Torres. On the Inf-Sup Stabillity of Crouzeix-Raviart Stokes Elements in 3D. Math. Comp., (https://doi.org/10.1090/mcom/3793), 2022.
- [31] C. Schwab and M. Suri. Mixed finite element methods for Stokes and non-Newtonian flow. Comput. Methods Appl. Mech. Engrg., 175(3-4):217–241, 1999.
- [32] 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(1):111–143, 1985.
- [33] R. Stenberg and M. Suri. Mixed finite element methods for problems in elasticity and Stokes flow. Numer. Math., 72(3):367–389, 1996.
- [34] G. Stoyan and Á. Baran. Crouzeix-Velte decompositions for higher-order finite elements. Comput. Math. Appl., 51(6-7):967–986, 2006.
- [35] A. Veeser and P. Zanotti. Quasi-optimal nonconforming methods for symmetric elliptic problems. I—Abstract theory. SIAM J. Numer. Anal., 56(3):1621–1642, 2018.
- [36] A. Veeser and P. Zanotti. Quasi-optimal nonconforming methods for symmetric elliptic problems. II—Overconsistency and classical nonconforming elements. SIAM J. Numer. Anal., 57(1):266–292, 2019.
- [37] M. Vogelius. A right-inverse for the divergence operator in spaces of piecewise polynomials. Numerische Mathematik, 41(1):19–37, 1983.
- [38] T. Warburton and J. S. Hesthaven. On the constants in -finite element trace inverse inequalities. Comput. Methods Appl. Mech. Engrg., 192(25):2765–2773, 2003.