Korn’s inequalities for piecewise vector fields

By Susanne C. Brenner

Abstract

Korn’s inequalities for piecewise vector fields are established. They can be applied to classical nonconforming finite element methods, mortar methods and discontinuous Galerkin methods.

1. Introduction

In this paper we use a boldface italic lower-case Roman letter such as to denote a vector (or vector function) with components () and a boldface lower-case Greek letter such as to denote a matrix (or matrix function) with components (). The Euclidean norm of the vector (resp. the Frobenius norm of the matrix ) will be denoted by (resp. ).

Let be a bounded connected open polyhedral domain in ( or ). The classical Korn inequality (cf. Reference 8, Reference 14, Reference 5 and the references therein) states that there exists a (generic) positive constant such that

where the strain tensor is the matrix with components

and the (semi)norms are defined by

Let be the space of (infinitesimal) rigid motions on defined by

where is the position vector function on and is the Lie algebra of anti-symmetric matrices. The space is precisely the kernel of the strain tensor; i.e., for ,

Let be a seminorm on with the following properties:

where is a generic positive constant depending on , and

Note that such a seminorm is invariant under the addition of a constant vector ; i.e.,

Then Equation 1.1, Equation 1.4, Equation 1.5 and the compactness of the embedding of into imply that

In particular, the inequality Equation 1.6 implies

where

is the orthogonal projection operator from onto the orthogonal complement of the constant vector functions;

for all , where is the infinitesimal -dimensional volume and is a measurable subset of with a positive -dimensional volume; and

where is the vector function (the curl of ) defined by

when , and the scalar function (the rotation of ) defined by

when .

Remark 1.1.

The inequality Equation 1.7 is of course equivalent to Korn’s inequality Equation 1.1. Inequalities Equation 1.8 and Equation 1.9 imply Korn’s first inequality

and Korn’s second inequality

Henceforth we will also refer to Equation 1.9 as Korn’s second inequality.

In this paper we establish analogs of Equation 1.7, Equation 1.8 and Equation 1.9 for piecewise vector fields (functions) with respect to a partition of consisting of nonoverlapping polyhedral subdomains, which is not necessarily a triangulation of . In other words, we only assume that

Typical two- and three-dimensional examples of partitions are depicted in Figure 1, where the square is partitioned into 7 subdomains and the cube is partitioned into 5 subdomains.

The space of piecewise vector fields (functions) is defined by

and the seminorm is given by

We also use the notation to denote the matrix function defined by

Let be the set of all the (open) sides (i.e., edges () or faces ()) common to two subdomains in . For example, there are 10 such edges in the two-dimensional example in Figure 1 and 8 such faces in the three-dimensional example. (Precise definitions of will be given in Section 4 and Section 5.) For , we denote by the orthogonal projection operator from onto , the space of vector polynomial functions on of degree .

The following are analogs of the classical Korn inequalities for :

where is the jump of across the side and the positive constant depends only on the shape regularity of the partition . In particular these inequalities are valid for partitions that are not quasi-uniform. (More details on the shape regularity assumptions are given in Section 4 and Section 5.)

Inequalities Equation 1.11Equation 1.13 imply

provided for all , or equivalently

Thus we immediately obtain Equation 1.14Equation 1.16 for certain classical nonconforming finite elements (cf. Reference 11, Reference 10, Reference 9, Reference 13). With some modifications, these estimates can also be applied to certain mortar elements (cf. Reference 18). Details will be carried out elsewhere.

The inequalities Equation 1.11Equation 1.13 also immediately imply

for all . These estimates are useful for the analysis of discontinuous Galerkin methods for elasticity problems (cf. Reference 15, Reference 6, Reference 12 and the references therein).

Remark 1.2.

Note that classical Korn’s inequalities can also be expressed in terms of the full norm (cf. Reference 8, Reference 14, Reference 5). In view of Equation 1.11Equation 1.13 and the following Poincaré-Friedrichs inequalities (cf. Reference 2):

for all , where is the orthogonal projection operator from onto , the space of constant vector functions on , we also have the following “full norm” versions of Korn’s inequalities for piecewise vector fields:

for all . The “full norm” versions of Equation 1.14Equation 1.16 and Equation 1.18Equation 1.20 can be readily derived from Equation 1.21Equation 1.23.

Remark 1.3.

Let be the restriction of to and let be the orthogonal projection operator from onto . Then Korn’s first inequalities Equation 1.8 and Equation 1.15 (resp. Equation 1.12 and Equation 1.22) remain valid if the terms involving the integral over in these inequalities are replaced by (resp. ).

The rest of the paper is organized as follows. First we derive Korn’s inequalities for piecewise linear and piecewise vector fields with respect to simplicial triangulations of . These are carried out in Section 2 and Section 3. Korn’s inequalities for piecewise vector fields with respect to general partitions are then established in Section 4 for two-dimensional domains and in Section 5 for three-dimensional domains. A generalization of the result in Section 2 to piecewise polynomial vector fields is given in Section 6, which can be used to derive Equation 1.14Equation 1.16 for some nonconforming finite elements that violate Equation 1.17. The appendix contains a discussion of the dependence of the constant in Korn’s second inequality Equation 1.9 on the underlying domain, which is used in Section 3 and Section 5.

Throughout this paper we use to denote the -dimensional volume of a -dimensional geometric object in a Euclidean space.

2. A generalized Korn’s inequality for piecewise linear vector fields with respect to simplicial triangulations

In this and the next two sections we restrict our attention to the case where the partition is actually a triangulation of by simplexes (i.e., triangles for and tetrahedra for ). The intersection of the closures of any two simplexes in is therefore either empty, a vertex, a closed edge or a closed face. In this case coincides with the set of interior open edges () or open faces (). The minimum angle of the triangles or tetrahedra in will be denoted by .

To avoid the proliferation of constants, we henceforth use the notation to represent the statement , where the (generic) function is continuous and independent of . The notation is equivalent to and .

Let be the space of piecewise linear vector fields and be the space of continuous piecewise linear vector fields. We define a linear map as follows. Let be the set of all the vertices of . Then is defined by

where

is the set of simplexes sharing as a common vertex and is the number of simplexes in . Note that

The following lemma contains the basic estimate for the operator .

Lemma 2.1.

It holds that

where is the set of the vertices of the simplex ,

is the set of interior sides sharing as a common vertex, and is the jump of across .

Proof.

Let , and . We have, by Equation 2.1,

Let be a simplex in . There exists a chain of simplexes such that (i) and , and (ii) and share a common side . (A two-dimensional example is depicted in Figure 2.)

Note that Equation 2.2 implies and hence

The estimate Equation 2.3 follows from Equation 2.4 and Equation 2.5.

We can now prove a generalized Korn’s inequality for functions in .

Lemma 2.2.

Let be a seminorm such that

where is the set of the vertices of , and

Then the following estimate holds

for all , where for all .

Proof.

Observe first that Equation 2.2, Equation 2.3 and a standard finite element estimate for (cf. Reference 4, Reference 3) imply

where we have also used the relation

From Equation 1.6, Equation 2.7 and Equation 2.10 we then find, for arbitrary ,

The following are examples of that satisfy conditions Equation 2.6Equation 2.8. The validity of Equation 2.6 and Equation 2.8 is obvious in all three examples.

Example 2.3.

Let be defined by

where is the orthogonal projection from onto the orthogonal complement of the constant vector fields. Condition Equation 2.7 can be verified as follows:

where we have used Equation 2.2, Equation 2.3, Equation 2.11 and a standard finite element estimate for the -norm.

Example 2.4.

Let be defined by

where is a measurable subset of with a positive -dimensional volume. Using Equation 2.2, Equation 2.3, Equation 2.11 and a standard finite element estimate for the -norm, condition Equation 2.7 can be verified as follows:

Example 2.5.

Let be defined by

Using Equation 2.10, we can easily verify condition Equation 2.7:

Remark 2.6.

The definitions of the seminorms , and can be extended to for a general partition of . In fact, if we denote by the set of all the partitions of , then is a well-defined function on for .

3. Korn’s inequalities for piecewise vector fields with respect to simplicial triangulations

Let be a simplicial triangulation of . First we define on each an interpolation operator from onto (the space of the rigid motions restricted to ) by the following conditions:

These conditions determine because

Note that Equation 1.4, Equation 3.2 and Corollary A.3 in the appendix imply

for all , and Equation 3.1 together with the classical Poincaré-Friedrichs inequality (with scaling) yields

Let , the space of piecewise linear vector fields with respect to , be defined by

We can now prove a generalized Korn’s inequality for functions in .

Theorem 3.1.

Let be a seminorm satisfying conditions Equation 2.6Equation 2.8 and, in addition, the condition that

where is defined by Equation 3.5. Then the following estimate holds

where is a continuous function independent of .

Proof.

Let be arbitrary. From Equation 1.4, Equation 2.9 and Equation 3.3 we have

Using condition Equation 3.6, we immediately find

Let be arbitrary and . We have, by a standard inverse estimate (cf. Reference 4, Reference 3),

where we have also used the fact that since .

Let be the set of the two simplexes in sharing as a common side. It follows from Equation 3.3, Equation 3.4 and the trace theorem (with scaling) that

Combining Equation 2.11, Equation 3.10 and Equation 3.11, we find

The estimate Equation 3.7 follows from Equation 3.8, Equation 3.9 and Equation 3.12.

Finally we observe that the seminorms in Examples 2.32.5 satisfy the condition Equation 3.6. In view of Equation 3.2, this is trivial for . Using Equation 3.3 and Equation 3.4, the case of can be established as follows:

For the case of , we apply Equation 3.3, Equation 3.4 and the trace theorem to obtain

Remark 3.2.

From here on we assume that is a seminorm for every (cf. Remark 2.6) and that it satisfies the conditions Equation 2.6Equation 2.8 and Equation 3.6 for every .

Remark 3.3.

By choosing to be , or , we immediately obtain Korn’s inequalities Equation 1.11Equation 1.13 in the case where is a simplicial triangulation. Similar remarks apply in the next three sections.

4. Korn’s inequalities for on a two-dimensional

First we need a precise definition of the set of interior (open) edges for a general partition , which in turn requires the concept of a vertex of . We define a vertex of to be a vertex of any of the subdomains in . (For example, the partition of the square in Figure 1 has 14 vertices.) We then define an open edge of to be an open line segment on the boundary of a subdomain in bounded between two of the vertices of . The set consists of the open edges of that are common to the boundaries of two subdomains in .

Remark 4.1.

The concept of an edge of a polygon and the concept of an edge of on are different. For example, a square always has 4 edges while there are 5 edges of the two-dimensional partition in Figure 1 on the boundary of the square at the lower right corner.

In order to apply Theorem 3.1 we introduce the set

By definition Equation 4.1, is a subspace of for every . Since functions in are continuous on the edges of that are not in , the following result is an immediate consequence of Theorem 3.1.

Theorem 4.2.

Let be as in Remark 3.2. Then we have

for all , where is a continuous function independent of .

The set provides an abstract measure of the shape regularity of the partition and the number can be viewed as a constant depending on the shape regularity of . However, in applications one may want to relate the abstract estimate Equation 4.2 to a concrete description of the shape regularity of given in terms of (i) the shape regularity of individual subdomains and (ii) the relative positions of subdomains that share a common edge of .

We can measure the shape regularity of a polygon (or a polyhedron in 3D) by using an affine homeomorphism between and a reference domain and by using the aspect ratio of defined by (diameter of )/(radius of the largest disc (or ball) in the closure of ).

The relative positions between subdomains sharing a common edge of can be measured in terms of the quantity

The following corollary gives an application of Theorem 4.2 to a fairly general class of two-dimensional partitions.

Corollary 4.3.

Let be as in Remark 3.2 and let be a family of partitions of such that

(i)

the polygons appearing in all the partitions are affine homeomorphic to a fixed finite set of reference polygons and the aspect ratios of the polygons in all the ’s are uniformly bounded,

(ii)

the set is bounded.

Then there exists a positive constant , independent of , such that

for any and .

Proof.

It suffices to show that under the assumptions on the family of partitions we can construct one partition for each such that . Then the estimate Equation 4.4 follows from Equation 4.2 if we take to be an upper bound of the bounded set .

First we construct a simplicial triangulation on each reference polygon so that each edge of the reference polygon is also an edge of the triangulation and each triangle in the triangulation can have at most one edge on the boundary of the reference polygon.

Let . We can induce a triangulation on using the triangulation on a reference polygon and the corresponding affine homeomorphism. Let be a vertex of which is not a vertex of . Then belongs to an edge of which is an edge of a triangle , and we connect to the vertex of not on by a straight line. In this way we have created a triangulation . (This construction is carried out in Figure 3 for the two-dimensional partition in Figure 1, where the reference square is triangulated by its two diagonals.)

Let be the reference polygon affine homeomorphic to and let be the corresponding affine map from to . The uniform boundedness of the aspect ratios implies (cf. Theorem 3.1.3 in Reference 4) the existence of a positive constant , independent of , such that

where is the matrix 2-norm induced by the Euclidean vector norm. Hence we have

where , and are any four points such that and .

We conclude from Equation 4.6 and the boundedness of the set that is bounded away from zero.

Remark 4.4.

If the family of partitions in Corollary 4.3 is actually a family of triangulations (simplicial or otherwise), then the condition on the boundedness of is redundant.

An example of a family of partitions satisfying the assumptions of Corollary 4.3 is depicted in Figure 4, where a square is being refined successively towards the upper right corner.

5. Korn’s inequalities for on a three-dimensional

In order to give a precise definition of , we first introduce the concept of an edge of , which is just an edge of any of the subdomains in . We then define an open face of to be an open subset of the boundary of a subdomain in enclosed by edges of . The set consists of open faces of common to the boundaries of two subdomains in .

Remark 5.1.

Again the concept of a face of a polyhedron and the concept of a face of on are different. For example, there are always 6 faces on a parallelepiped while there are 9 faces of the three-dimensional partition in Figure 1 on the boundary of the subdomain in the back.

As in the two-dimensional case, we would like to derive a generalized Korn’s inequality for partitions from Theorem 3.1. But here the situation is more complicated since the faces in may not be triangles. Accordingly we introduce the following family of triangulations:

Since a face in may not be a face in for , we cannot immediately derive an analog of Theorem 4.2. We need to introduce two more parameters related to the shape regularity of in addition to the parameter already defined in Equation 4.3.

Let . For we will denote by the triangulation of by faces of , i.e., , and define the parameter

Note the following obvious bound for (the number of elements in ):

Moreover Equation 4.3 and Equation 5.2 imply that

The other parameter is the smallest number with the property that

where is any subdomain in , is any function in and (the space of rigid motions restricted to ) is defined by the conditions

The existence of is a consequence of Equation 1.4, Equation 5.6, the trace theorem, the Poincaré-Friedrichs inequality and Korn’s second inequality Equation 1.9.

We can now state and prove a generalized Korn’s inequality.

Theorem 5.2.

Let be as in Remark 3.2. Then we have

for all , where is a continuous function independent of .

Proof.

Let and . We have, from Equation 5.3 and the Cauchy-Schwarz inequality,

where is a continuous function independent of . Therefore it suffices to show that

for all , where is a continuous function independent of .

Since definition Equation 5.1 implies that is a subspace of for every , we immediately obtain from Equation 3.7 the estimate

for any , where is a continuous function independent of .

Let be arbitrary, let and let be the set of the two polyhedra in that share as a common face. It follows from the Cauchy-Schwarz inequality that, for any ,

since on .

From Equation 5.5 we have

Note also that Equation 5.2 and Equation 5.3 imply

Combining Equation 4.3 and Equation 5.10Equation 5.12, we find

for any and .

Finally we observe that the number of faces in that appear on the boundary of any subdomain in is less than or equal to , and hence

The inequality Equation 5.8 follows from Equation 5.9, Equation 5.13 and Equation 5.14, with the function given by, for example,

The set provides an abstract measure of the shape regularity of the partition and we can think of

as a constant depending on the shape regularity of . Under appropriate concrete shape regularity assumptions one can also obtain from Theorem 5.2 Korn’s inequalities for a family of partitions with a uniform constant. For simplicity we only give an analog of Corollary 4.3 for partitions by convex polyhedra.

Since a face of a partition consisting of convex polyhedra is a convex polygon, it can be triangulated by connecting its center to the vertices of on its boundary by straight lines. Such a triangulation will be referred to as the canonical triangulation of the face.

Corollary 5.3.

Let be as in Remark 3.2 and be a family of partitions of with the following properties

(i)

The polyhedra appearing in all the partitions are affine homeomorphic to a fixed finite set of convex reference polyhedra and the aspect ratios of the polyhedra in all the ’s are uniformly bounded.

(ii)

The set is bounded.

(iii)

The angles of the triangles in the canonical triangulations of the faces of all the partitions are bounded below by a positive constant.

Then there exists a positive constant , independent of , such that

for any and .

Proof.

Let be affine homeomorphic to the reference polyhedron and let be the corresponding affine map from to . Then the estimates Equation 4.5 and Equation 4.6 again follow from condition (i).

From Equation 1.4, Equation 5.6, the trace theorem (with scaling), the classical Poincaré-Friedrichs inequality (with scaling), condition (i) and Lemma A.2 in the appendix, we have

where is a positive constant depending only on . Since there are only finitely many reference polyhedra for the partitions , we conclude from Equation 5.16 that the set

For each we can construct a triangulation by first imposing the canonical triangulation on each face of and then triangulating each subdomain using its center and the triangles on its faces.

Let be a face of . Condition (iii) implies that the number of triangles in the canonical triangulation of is uniformly bounded (since these triangles share the center of as a common vertex) and the areas of any two triangles in the canonical triangulation are also comparable. It follows that

Condition (ii) implies that the number of faces of on the face of a subdomain is uniformly bounded, which together with the observation in the previous paragraph implies that the number of triangles of on is also uniformly bounded. It then follows from condition (iii) that the triangulation of by the triangular faces from is quasi-uniform. Moreover condition (i) implies that the number of faces of is uniformly bounded and that the sizes of any two faces of are comparable. Therefore the triangulation of by the faces from is also quasi-uniform, which together with Equation 4.6 implies

Combining condition (ii) and Equation 5.17Equation 5.19, we see that is a precompact subset of . The estimate Equation 5.15 then follows from Equation 5.7 if we take to be an upper bound of the bounded set .

Remark 5.4.

If the family of partitions in Corollary 5.3 is actually a family of triangulations (simplicial or otherwise), then conditions (ii) and (iii) are redundant. Moreover, it is not necessary to assume that the subdomains are convex, since there are only finitely many different reference polygons for the faces of the ’s.

An example of a family of partitions satisfying the assumptions of Corollary 5.3 is depicted in Figure 5, where a cube is being refined successively towards the upper left front corner.

6. Korn’s inequalities for piecewise polynomial vector fields with respect to triangulations by polyhedral subdomains

Attentive readers may have already noticed that the inequality Equation 2.9 for piecewise linear vector fields is different from the inequalities Equation 3.7, Equation 4.2 and Equation 5.7 for piecewise vector fields. Since pointwise evaluation is not well defined for functions in and , the formulation of Korn’s inequalities in Equation 3.7, Equation 4.2 and Equation 5.7 is the appropriate one for piecewise vector fields. However, for piecewise polynomial vector fields associated with a triangulation (simplicial or otherwise), pointwise evaluation of the jump across is possible. The following theorem generalizes Lemma 2.2 to such vector fields.

Theorem 6.1.

Let be as in Remark 3.2 and let be a family of triangulations of by polygons or polyhedra . Assume that the subdomains appearing in all the triangulations are affine homeomorphic to a fixed finite set of reference domains and that the aspect ratios of the subdomains in all the ’s are uniformly bounded. Let for , where is a positive integer. Then there exists a positive constant , independent of , such that

for any and .

Proof.

We will use to denote a generic positive constant independent of .

Let and be arbitrary. Recall that is the set of the two simplexes sharing as a common face. For , we have

where (respectively ) is the linear nodal interpolant of (respectively ).

The trace theorem (with scaling) and the Bramble-Hilbert lemma (cf. Reference 1) imply that

Moreover, from the well-known relation (cf. Reference 8)

and a standard inverse estimate we have

Combining Equation 2.11, Equation 6.3 and Equation 6.4, we find

On the other hand we obtain, from a standard finite element estimate for the -norm,

The inequality Equation 6.1 follows from Equation 6.2, Equation 6.5, Equation 6.6, Corollary 4.3, Corollary 5.3, Remark 4.4 and Remark 5.4.

Using Theorem 6.1, we can immediately obtain Korn’s inequalities Equation 1.14Equation 1.16 for Wilson’s brick/rectangle (cf. Reference 17, Reference 4, Reference 16, Reference 19) and other nonconforming quadrilateral elements in Reference 20 which are continuous at the vertices of the triangulation. Note that these elements do not satisfy the weak continuity condition Equation 1.17.

Appendix A. Dependence of the constant in Korn’s second inequalityon the underlying domain

Let be a bounded connected open polyhedral domain in and let be the smallest positive number such that

for all . In this appendix we briefly discuss the behavior of under affine homeomorphisms. More precisely, we assume that is homeomorphic to a reference domain under the affine transformation defined by , and we consider the dependence of on , the Lie group of nonsingular matrices.

Remark A.1.

The estimates Equation A.3 and Equation A.4 below are crucial for Lemma 2.2 and Corollary 5.3. These estimates, though elementary, do not seem to be in the literature.

Without loss of generality, we may assume and write the constant in Equation A.1 as . We have,

Observe that, since , the quotients on the right-hand side of Equation A.2 form a family of equicontinuous functions on . Therefore , defined as the supremum of this equicontinuous family, is continuous on (cf. Reference 7).

Lemma A.2.

Let be a family of domains affine homeomorphic to the reference domain . Assume also that the aspect ratios of the ’s are uniformly bounded. Then there exists a positive constant , independent of , such that

for any and .

Proof.

We may assume without loss of generality that diam for all . It follows from the uniform boundedness of the aspect ratios of the ’s that the norm of the nonsingular matrix in the affine transformation and the norm of its inverse are uniformly bounded for all (cf. Equation 4.5). Hence is a precompact subset of and the boundedness of follows from the continuity of on .

For a simplex we can also control the constant in terms of the minimum angle of .

Corollary A.3.

There exists a continuous decreasing function such that, for any simplex ,

for all .

Proof.

It follows from Lemma A.2 that, for any , the set

is bounded. Then defines a nonnegative decreasing function on . In view of Equation A.1 and Equation A.5 we have

for all .

The estimate Equation A.4 follows from Equation A.6 if we choose to be any continuous decreasing function satisfying the condition . (There are infinitely many such functions.)

Mathematical Fragments

Equation (1.1)
Equation (1.4)
Equation (1.5)
Equation (1.6)
Equation (1.7)
Equation (1.8)
Equation (1.9)
Equations (1.11), (1.12), (1.13)
Equations (1.14), (1.15), (1.16)
Equation (1.17)
Equations (1.18), (1.19), (1.20)
Remark 1.2.

Note that classical Korn’s inequalities can also be expressed in terms of the full norm (cf. Reference 8, Reference 14, Reference 5). In view of Equation 1.11Equation 1.13 and the following Poincaré-Friedrichs inequalities (cf. Reference 2):

for all , where is the orthogonal projection operator from onto , the space of constant vector functions on , we also have the following “full norm” versions of Korn’s inequalities for piecewise vector fields:

for all . The “full norm” versions of Equation 1.14Equation 1.16 and Equation 1.18Equation 1.20 can be readily derived from 1.211.23.

Equation (2.1)
Equation (2.2)
Lemma 2.1.

It holds that

where is the set of the vertices of the simplex ,

is the set of interior sides sharing as a common vertex, and is the jump of across .

Equation (2.4)
Equation (2.5)
Lemma 2.2.

Let be a seminorm such that

where is the set of the vertices of , and

Then the following estimate holds

for all , where for all .

Equation (2.10)
Equation (2.11)
Example 2.3.

Let be defined by

where is the orthogonal projection from onto the orthogonal complement of the constant vector fields. Condition Equation 2.7 can be verified as follows:

where we have used Equation 2.2, Equation 2.3, Equation 2.11 and a standard finite element estimate for the -norm.

Example 2.5.

Let be defined by

Using Equation 2.10, we can easily verify condition Equation 2.7:

Remark 2.6.

The definitions of the seminorms , and can be extended to for a general partition of . In fact, if we denote by the set of all the partitions of , then is a well-defined function on for .

Equations (3.1), (3.2)
Equation (3.3)
Equation (3.4)
Equation (3.5)
Theorem 3.1.

Let be a seminorm satisfying conditions Equation 2.6Equation 2.8 and, in addition, the condition that

where is defined by Equation 3.5. Then the following estimate holds

where is a continuous function independent of .

Equation (3.8)
Equation (3.9)
Equation (3.10)
Equation (3.11)
Equation (3.12)
Remark 3.2.

From here on we assume that is a seminorm for every (cf. Remark 2.6) and that it satisfies the conditions Equation 2.6Equation 2.8 and Equation 3.6 for every .

Equation (4.1)
Theorem 4.2.

Let be as in Remark 3.2. Then we have

for all , where is a continuous function independent of .

Equation (4.3)
Corollary 4.3.

Let be as in Remark 3.2 and let be a family of partitions of such that

(i)

the polygons appearing in all the partitions are affine homeomorphic to a fixed finite set of reference polygons and the aspect ratios of the polygons in all the ’s are uniformly bounded,

(ii)

the set is bounded.

Then there exists a positive constant , independent of , such that

for any and .

Equation (4.5)
Equation (4.6)
Remark 4.4.

If the family of partitions in Corollary 4.3 is actually a family of triangulations (simplicial or otherwise), then the condition on the boundedness of is redundant.

Equation (5.1)
Equation (5.2)
Equation (5.3)
Equation (5.5)
Equation (5.6)
Theorem 5.2.

Let be as in Remark 3.2. Then we have

for all , where is a continuous function independent of .

Equation (5.8)
Equation (5.9)
Equation (5.10)
Equation (5.12)
Equation (5.13)
Equation (5.14)
Corollary 5.3.

Let be as in Remark 3.2 and be a family of partitions of with the following properties

(i)

The polyhedra appearing in all the partitions are affine homeomorphic to a fixed finite set of convex reference polyhedra and the aspect ratios of the polyhedra in all the ’s are uniformly bounded.

(ii)

The set is bounded.

(iii)

The angles of the triangles in the canonical triangulations of the faces of all the partitions are bounded below by a positive constant.

Then there exists a positive constant , independent of , such that

for any and .

Equation (5.16)
Equation (5.17)
Equation (5.19)
Remark 5.4.

If the family of partitions in Corollary 5.3 is actually a family of triangulations (simplicial or otherwise), then conditions (ii) and (iii) are redundant. Moreover, it is not necessary to assume that the subdomains are convex, since there are only finitely many different reference polygons for the faces of the ’s.

Theorem 6.1.

Let be as in Remark 3.2 and let be a family of triangulations of by polygons or polyhedra . Assume that the subdomains appearing in all the triangulations are affine homeomorphic to a fixed finite set of reference domains and that the aspect ratios of the subdomains in all the ’s are uniformly bounded. Let for , where is a positive integer. Then there exists a positive constant , independent of , such that

for any and .

Equation (6.2)
Equation (6.3)
Equation (6.4)
Equation (6.5)
Equation (6.6)
Equation (A.1)
Equation (A.2)
Lemma A.2.

Let be a family of domains affine homeomorphic to the reference domain . Assume also that the aspect ratios of the ’s are uniformly bounded. Then there exists a positive constant , independent of , such that

for any and .

Corollary A.3.

There exists a continuous decreasing function such that, for any simplex ,

for all .

Equation (A.5)
Equation (A.6)

References

Reference [1]
J.H. Bramble and S.R. Hilbert. Estimation of linear functionals on Sobolev spaces with applications to Fourier transforms and spline interpolation. SIAM J. Numer. Anal., 7:113–124, 1970. MR 41:7819
Reference [2]
S.C. Brenner. Poincaré-Friedrichs inequalities for piecewise functions. SIAM J. Numer. Anal. 41: 306–324, 2003 (electronic).
Reference [3]
S.C. Brenner and L.R. Scott. The Mathematical Theory of Finite Element Methods (Second Edition). Springer-Verlag, New York-Berlin-Heidelberg, 2002. MR 2003a:65103
Reference [4]
P.G. Ciarlet. The Finite Element Method for Elliptic Problems. North-Holland, Amsterdam, 1978. MR 58:25001
Reference [5]
P.G. Ciarlet. Mathematical Elasticity Volume I: Three-Dimensional Elasticity. North-Holland, Amsterdam, 1988. MR 89e:73001
Reference [6]
B. Cockburn, G.E. Karniadakis, and C.-W. Shu, editors. Discontinuous Galerkin Methods. Springer-Verlag, Berlin-Heidelberg, 2000. MR 2002b:65004
Reference [7]
J. Dieudonné. Foundations of Modern Analysis. Pure and Applied Mathematics, Vol. X. Academic Press, New York, 1960.MR 22:11074
Reference [8]
G. Duvaut and J.L. Lions. Inequalities in Mechanics and Physics. Springer-Verlag, Berlin, 1976. MR 58:25191
Reference [9]
R.S. Falk. Nonconforming finite element methods for the equations of linear elasticity. Math. Comp., 57:529–550, 1991. MR 92a:65290
Reference [10]
M. Fortin. A three-dimensional quadratic nonconforming element. Numer. Math., 46:269–279, 1985. MR 86f:65192
Reference [11]
M. Fortin and M. Soulie. A non-conforming piecewise quadratic finite element on triangles. Internat. J. Numer. Methods Engrg., 19:505–520, 1983. MR 84g:76004
Reference [12]
P. Hansbo and M.G. Larson. Discontinuous Galerkin methods for incompressible and nearly incompressible elasticity by Nitsche’s method. Comput. Methods Appl. Mech. Engrg., 191:1895–1908, 2002.
Reference [13]
P. Knobloch. On Korn’s inequality for nonconforming finite elements. Technische Mechanik, 20:205–214 and 375 (Errata), 2000.
Reference [14]
J.A. Nitsche. On Korn’s second inequality. RAIRO Anal. Numér., 15:237–248, 1981. MR 83a:35012
Reference [15]
Ch. Schwab. -FEM for fluid flow simulation. In T.J. Barth and H. Deconinck, editors, High-Order Methods for Computational Physics, pages 325–438. Springer-Verlag, Berlin-Heidelberg, 1999.MR 2000k:76093
Reference [16]
M. Wang. The generalized Korn inequality on nonconforming finite element spaces. Chinese J. Numer. Math. Appl., 16:91–96, 1994. MR 98a:65166
Reference [17]
E.L. Wilson, R.L. Taylor, W. Doherty, and J. Ghaboussi. Incompatible displacement models. In S.J. Fenves, N. Perrone, A.R. Robinson, and W.C. Schnobrich, editors, Numerical and Computer Methods in Structural Mechanics, pages 43–57. Academic Press, New York, 1973.
Reference [18]
B.I. Wohlmuth. Discretization Methods and Iterative Solvers Based on Domain Decomposition. Springer-Verlag, Heidelberg, 2001.MR 2002c:65231
Reference [19]
X. Xu. A discrete Korn’s inequality in two and three dimensions. Appl. Math. Letters, 13:99–102, 2000.MR 2000m:74087
Reference [20]
Z. Zhang. Analysis of some quadrilateral nonconforming elements for incompressible elasticity. SIAM J. Numer. Anal., 34:640–663, 1997.MR 98b:73041

Article Information

MSC 2000
Primary: 65N30 (Finite elements, Rayleigh-Ritz and Galerkin methods, finite methods), 74S05 (Finite element methods)
Keywords
  • Korn’s inequalities
  • piecewise vector fields
  • nonconforming finite elements
  • mortar methods
  • discontinuous Galerkin methods
Author Information
Susanne C. Brenner
Department of Mathematics, University of South Carolina, Columbia, South Carolina 29208
brenner@math.sc.edu
Additional Notes

This work was supported in part by the National Science Foundation under Grant No. DMS-00-74246.

Journal Information
Mathematics of Computation, Volume 73, Issue 247, ISSN 1088-6842, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2003 American Mathematical Society
Article References

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.