Mixed finite element methods for linear elasticity with weakly imposed symmetry

By Douglas N. Arnold, Richard S. Falk, and Ragnar Winther

Abstract

In this paper, we construct new finite element methods for the approximation of the equations of linear elasticity in three space dimensions that produce direct approximations to both stresses and displacements. The methods are based on a modified form of the Hellinger–Reissner variational principle that only weakly imposes the symmetry condition on the stresses. Although this approach has been previously used by a number of authors, a key new ingredient here is a constructive derivation of the elasticity complex starting from the de Rham complex. By mimicking this construction in the discrete case, we derive new mixed finite elements for elasticity in a systematic manner from known discretizations of the de Rham complex. These elements appear to be simpler than the ones previously derived. For example, we construct stable discretizations which use only piecewise linear elements to approximate the stress field and piecewise constant functions to approximate the displacement field.

1. Introduction

The equations of linear elasticity can be written as a system of equations of the form

Here the unknowns and denote the stress and displacement fields engendered by a body force acting on a linearly elastic body which occupies a region . Then takes values in the space of symmetric matrices and takes values in . The differential operator is the symmetric part of the gradient, the div operator is applied row-wise to a matrix, and the compliance tensor is a bounded and symmetric, uniformly positive definite operator reflecting the properties of the body. If the body is clamped on the boundary of , then the proper boundary condition for the system Equation 1.1 is on . For simplicity, this boundary condition will be assumed here. The issues that arise when other boundary conditions are assumed (e.g., the case of pure traction boundary conditions ) are discussed in Reference 9.

The pair can alternatively be characterized as the unique critical point of the Hellinger–Reissner functional

The critical point is sought among all , the space of square-integrable symmetric matrix fields with square-integrable divergence, and all , the space of square-integrable vector fields. Equivalently, is the unique solution to the following weak formulation of the system Equation 1.1:

A mixed finite element method determines an approximate stress field and an approximate displacement field as the critical point of over where and are suitable piecewise polynomial subspaces. Equivalently, the pair is determined by the weak formulation Equation 1.3, with the test space restricted to . As is well known, the subspaces and cannot be chosen arbitrarily. To ensure that a unique critical point exists and that it provides a good approximation of the true solution, they must satisfy the stability conditions from Brezzi’s theory of mixed methods Reference 12Reference 13.

Despite four decades of effort, no stable simple mixed finite element spaces for elasticity have been constructed. For the corresponding problem in two space dimensions, stable finite elements were presented in Reference 10. For the lowest order element, the space is composed of piecewise cubic functions, with 24 degrees of freedom per triangle, while the space consists of piecewise linear functions. Another approach which has been discussed in the two-dimensional case is the use of composite elements, in which consists of piecewise polynomials with respect to one triangulation of the domain, while consists of piecewise polynomials with respect to a different, more refined, triangulation Reference 5Reference 21Reference 23Reference 31. In three dimensions, a partial analogue of the element in Reference 10 has been proposed and shown to be stable in Reference 1. This element uses piecewise quartic stresses with 162 degrees of freedom per tetrahedron, and piecewise linear displacements.

Because of the lack of suitable mixed elasticity elements, several authors have resorted to the use of Lagrangian functionals which are modifications of the Hellinger–Reissner functional given above Reference 2Reference 4Reference 6Reference 27Reference 28Reference 29Reference 30, in which the symmetry of the stress tensor is enforced only weakly or abandoned altogether. In order to discuss these methods, we consider the compliance tensor as a symmetric and positive definite operator mapping into , where is the space of matrices. In the isotropic case, for example, the mapping has the form

where are positive scalar coefficients, the Lamé coefficients. A modification of the variational principle discussed above is obtained if we consider the extended Hellinger–Reissner functional

over the space , where denotes the space of skew symmetric matrices. We note that the symmetry condition for the space of matrix fields is now enforced through the introduction of the Lagrange multiplier, . A critical point of the functional is characterized as the unique solution of the system

Clearly, if is a solution of this system, then is symmetric, i.e., , and therefore the pair solves the corresponding system Equation 1.3. On the other hand, if solves Equation 1.3, then and, if we set to the skew-symmetric part of , then solves Equation 1.5. In this respect, the two systems Equation 1.3 and Equation 1.5 are equivalent. However, the extended system Equation 1.5 leads to new possibilities for discretization. Assume that we choose finite element spaces and consider a discrete system corresponding to Equation 1.5. If is a discrete solution, then will not necessarily inherit the symmetry property of . Instead, will satisfy the weak symmetry condition

Therefore, these solutions in general will not correspond to solutions of the discrete system obtained from Equation 1.3.

Discretizations based on the system Equation 1.5 will be referred to as mixed finite element methods with weakly imposed symmetry. Such discretizations were already introduced by Fraejis de Veubeke in Reference 21 and further developed in Reference 4. In particular, the so-called PEERS element proposed in Reference 4 for the corresponding problem in two space dimensions used a combination of piecewise linear functions and cubic bubble functions, with respect to a triangulation of the domain, to approximate the stress , piecewise constants to approximate the displacements, and continuous piecewise linear functions to approximate the Lagrange multiplier . Prior to the PEERS paper, Amara and Thomas Reference 2 developed methods with weakly imposed symmetry using a dual hybrid approach. The lowest order method they discussed approximates the stresses with quadratic polynomials plus bubble functions and the multiplier by discontinuous constant or linear polynomials. The displacements are approximated on boundary edges by linear functions. Generalizations of the idea of weakly imposed symmetry to other triangular elements, rectangular elements, and three space dimensions were developed in Reference 28, Reference 29, Reference 30 and Reference 24. In Reference 29, a family of elements is developed in both two and three dimensions. The lowest order element in the family uses quadratics plus the curls of quartic bubble functions in two dimensions or quintic bubble functions in three dimensions to approximate the stresses, discontinuous linears to approximate the displacements, and discontinuous quadratics to approximate the multiplier. In addition, a lower order method is introduced that approximates the stress by piecewise linear functions augmented by the curls of cubic bubble functions plus a cubic bubble times the gradient of local rigid motions. The multiplier is approximated by discontinuous piecewise linear functions and the displacement by local rigid motions. Morley Reference 24 extends PEERS to a family of triangular elements, to rectangular elements, and to three dimensions. In addition, the multiplier is approximated by nonconforming rather than continuous piecewise polynomials.

There is a close connection between mixed finite elements for linear elasticity and discretization of an associated differential complex, the elasticity complex, which will be introduced in §3 below. In fact, the importance of this complex was already recognized in Reference 10, where mixed methods for elasticity in two space dimensions were discussed. The new ingredient here is that we utilize a constructive derivation of the elasticity complex starting from the de Rham complex. This construction is described in Eastwood Reference 18 and is based on the the Bernstein–Gelfand–Gelfand resolution; cf. Reference 11 and also Reference 14. By mimicking the construction in the discrete case, we are able to derive new mixed finite elements for elasticity in a systematic manner from known discretizations of the de Rham complex. As a result, we can construct new elements in both two and three space dimensions which are significantly simpler than those derived previously. For example, we will construct stable discretizations of the system Equation 1.5 which only use piecewise linear and piecewise constant functions, as illustrated in the figure below. For simplicity, the entire discussion of the present paper will be given in the three-dimensional case. A detailed discussion in two space dimensions can be found in Reference 8. Besides the methods discussed here, we note that by slightly generalizing the approach of this paper, one can also analyze some of the previously known methods mentioned above that are also based on the weak symmetry formulation (see Reference 19 for details).

Figure 1.

Elements for the stress, displacement, and multiplier in the lowest order case in two dimensions and three dimensions.

Graphic without alt text

An alternative approach to construct finite element methods for linear elasticity is to consider a pure displacement formulation. Since the coefficient in Equation 1.1 is invertible, the stress can be eliminated using the first equation in Equation 1.1, the stress-strain relation. This leads to the second order equation

for the displacement . A weak solution of this equation can be characterized as the global minimizer of the energy functional

over the Sobolev space . Here denotes the space of all square integrable vector fields on , with square integrable derivatives, and which vanish on the boundary . A finite element approach based on this formulation, where we seek a minimum over a finite element subspace of is standard and discussed in textbooks, (e.g., Reference 16). However, for more general models, arising, for example, in viscoelasticity and plasticity (cf. Reference 15), the stress–strain relation is not local and an elimination of the stress is impossible. For such models, a pure displacement model is excluded, and a mixed approach seems to be an obvious alternative. The construction of stable mixed elements for linear elasticity is an important step in the construction of mixed methods for these more complicated models. Another advantage of the mixed approach is that we automatically obtain schemes which are uniformly stable in the incompressible limit, i.e., as the Lamé parameter tends to infinity. Since this behavior of mixed methods is well known, we will not focus further on this property here. A more detailed discussion in this direction can, for example, be found in Reference 5.

An outline of the paper is as follows. In §2, we describe the notation to be used, state our main result, and provide some preliminary discussion on the relation between stability of mixed finite element methods and discrete exact complexes. In §3, we present two complexes related to the two mixed formulations of elasticity given by Equation 1.3 and Equation 1.5. In §4, we introduce the framework of differential forms and show how the elasticity complex can be derived from the de Rham complex. In §5, we derive discrete analogues of the elasticity complex beginning from discrete analogues of the de Rham complex and identify the required properties of the discrete spaces necessary for this construction. This procedure is our basic design principle. In §6, we apply the construction of the preceding section to specific discrete analogues of the de Rham complex to obtain a family of discrete elasticity complexes. In §7 we use this family to construct stable finite element schemes for the approximation of the mixed formulation of the equations of elasticity with weakly imposed symmetry. Finally, in §8, we show how a slightly more complicated procedure leads to a simplified elasticity element.

2. Notation, statement of main results, and preliminaries

We begin with some basic notation and hypotheses. We continue to denote by the space of -vectors, by the space of real matrices, and by and the subspaces of symmetric and skew symmetric matrices, respectively. The operators and denote the symmetric and skew symmetric parts, respectively. Note that an element of the space can be identified with its axial vector in given by the map :

i.e., for any vectors and .

We assume that is a domain in with boundary . We shall use the standard function spaces, like the Lebesgue space and the Sobolev space . For vector-valued functions, we include the range space in the notation following a semicolon, so denotes the space of square integrable functions mapping into a normed vector space . The space denotes the subspace of (vector-valued) functions in whose divergence belongs to . Similarly, denotes the subspace of (matrix-valued) functions in whose divergence (by rows) belongs to .

Assuming that is an inner product space, then has a natural norm and inner product, which will be denoted by and , respectively. For a Sobolev space , we denote the norm by and for , the norm is denoted by . The space denotes the space of polynomial functions on of total degree . Usually we abbreviate this to just .

In this paper we shall consider mixed finite element approximations derived from Equation 1.5. These schemes take the form:

Find such that

where now , and .

Following the general theory of mixed finite element methods (cf. Reference 12Reference 13) the stability of the saddle–point system Equation 2.1 is ensured by the following conditions:

(A)

whenever satisfies , and ,

(A)

for all nonzero , there exists nonzero with ,

where and are positive constants independent of .

The main result of this paper, given in Theorem 7.1, is to construct a new family of stable finite element spaces , , that satisfy the stability conditions (A) and (A). We shall show that for , the choices of the Nédélec second family of H(div) elements of degree for (cf. Reference 26) and of discontinuous piecewise polynomials of degree for and provide a stable finite element approximation. In contrast to the previous work described in the introduction, no stabilizing bubble functions are needed; nor is interelement continuity imposed on the multiplier. In §8 we also discuss a somewhat simpler lowest order element () in which the local stress space is a strict subspace of the full space of linear matrix fields.

Our approach to the construction of stable mixed elements for elasticity is motivated by the success in developing stable mixed elements for steady heat conduction (i.e., the Poisson problem) based on discretizations of the de Rham complex. We recall (see, e.g., Reference 7) that there is a close connection between the construction of such elements and discretizations of the de Rham complex

More specifically, a key to the construction and analysis of stable mixed elements is a commuting diagram of the form

Here, the spaces and are the finite element spaces used to discretize the flux and temperature fields, respectively. The spaces and are additional finite element spaces, which can be found for all well-known stable element choices. The bottom row of the diagram is a discrete de Rham complex, which is exact when the de Rham complex is (i.e., when the domain is contractible). The vertical operators are projections determined by the natural degrees of freedom of the finite element spaces. As pointed out in Reference 7, there are many such discretizations of the de Rham complex.

A diagram analogous to Equation 2.3, but with the de Rham complex replaced by the elasticity complex defined just below, will be crucial to our construction of stable mixed elements for elasticity. Discretization of the elasticity complex also gives insight into the difficulties of constructing finite element approximations of the mixed formulation of elasticity with strongly imposed symmetry; cf. Reference 8.

3. The elasticity complex

We now proceed to a description of two elasticity complexes, corresponding to strongly or weakly imposed symmetry of the stress tensor. In the case of strongly imposed symmetry, relevant to the mixed elasticity system Equation 1.3, the characterization of the divergence-free symmetric matrix fields will be needed. In order to give such a characterization, define to be the differential operator defined by taking of each row of the matrix. Then define a second order differential operator by

It is easy to check that and that . In other words,

is a complex. Here the dependence of the domain is suppressed, i.e., , and denotes the six-dimensional space of infinitesimal rigid motions on , i.e., functions of the form with and . In fact, when is contractible, then Equation 3.2 is an exact sequence, a fact which will follow from the discussion below. The complex Equation 3.2 will be referred to as the elasticity complex.

A natural approach to the construction of stable mixed finite elements for elasticity would be to extend the complex Equation 3.2 to a complete commuting diagram of the form Equation 2.3, where Equation 3.2 is the top row and the bottom row is a discrete analogue. However, due to the pointwise symmetry requirement on the discrete stresses, this construction requires piecewise polynomials of high order. For the corresponding problem in two space dimensions, such a complex was proposed in Reference 10 with a piecewise cubic stress space; cf. also Reference 8. An analogous complex was derived in the three-dimensional case in Reference 3. It uses a piecewise quartic space, with 162 degrees of freedom on each tetrahedron for the stresses.

We consider the formulation based on weakly imposed symmetry of the stress tensor, i.e., the mixed system Equation 1.5. Then the relevant complex is, instead of Equation 3.2,

Here,

and denotes the extension of the operator defined on by Equation 3.1 such that on . We remark that may be written

where is the algebraic operator

with the identity matrix. Indeed, if is symmetric, then is trace free, and therefore the definition Equation 3.4 reduces to Equation 3.1 on . On the other hand, if is skew with axial vector , then , and so .

Observe that there is a close connection between Equation 3.2 and Equation 3.3. In fact, Equation 3.2 can be derived from Equation 3.3 by performing a projection step. To see this, consider the diagram

where the projection operators are defined by

We may identify with a subspace of , namely,

Under this identification, corresponds to . We identify the on the right with a different subspace of , namely,

With these identifications, the bottom row is a subcomplex of the top row, and the operators are all projections. Furthermore, the diagram commutes. It follows easily that the exactness of the upper row implies exactness of the bottom row.

In the next section, we shall discuss these complexes further. In particular, we show the elasticity complex with weakly imposed symmetry, i.e., Equation 3.3 follows from the de Rham complex Equation 2.2. Hence, as a consequence of the discussion above, both Equation 3.2 and Equation 3.3 will follow from Equation 2.2.

4. From the de Rham to the elasticity complex

The main purpose of this section is to demonstrate the connection between the de Rham complex Equation 2.2 and the elasticity complex Equation 3.3. In particular, we show that whenever Equation 2.2 is exact, Equation 3.3 is exact. This section serves as an introduction to a corresponding construction of a discrete elasticity complex, to be given in the next section. In the following section, the discrete complex will be used to construct stable finite elements for the system Equation 1.5.

The de Rham complex Equation 2.2 is most clearly stated in terms of differential forms. Here we briefly recall the definitions and properties we will need. We use a completely coordinate-free approach. For a slightly more expanded discussion and the expressions in coordinates see, e.g., Reference 7, §4. We let denote the space of smooth differential -forms on , i.e. , where denotes the vector space of alternating -linear maps on . If we let denote evaluated at , i.e., we use subscripts to indicate the spatial dependence.

Using the inner product on inherited from the inner product on (see equation (4.1) of Reference 7, §4), we may also define the Hilbert space of square integrable differential forms with norm denoted by , and also the th order Sobolev space , consisting of square integrable -forms for which the norm

is finite (where the sum is over multi-indices of degree at most ).

Thus, -forms are scalar functions and -forms are covector fields. We will not emphasize the distinction between vectors and covectors, since, given the inner product in , we may identify a -form with the vector field for which , . In the three-dimensional case, we can identify a -form with a vector field and a -form with a scalar field by

The exterior derivative is defined by

where the hat is used to indicate a suppressed argument and denotes the directional derivative in the direction of the vector . It is useful to define

with norm given by . Using the identifications given above, the correspond to , , and for , respectively, and the correspond to , , , and, for , .

The de Rham complex Equation 2.2 can then be written

It is a complex since .

A differential -form on may be restricted to a differential -form on any submanifold ; at each point of the restriction of is an alternating linear form on tangent vectors. Moreover, if , the integral is defined.

If is a vector space, then refers to the -forms with values in , i.e., , where are alternating -linear forms on with values in . Given an inner product on , we obtain an inner product on . Obviously the corresponding complex

is exact whenever the de Rham complex is.

We now construct the elasticity complex as a subcomplex of a complex isomorphic to the de Rham complex with values in the six-dimensional vector space . First, for any we define by . We then define an operator by

Next, we define an isomorphism by

with inverse given by

Next, define the operator by . Inserting the isomorphisms in the -valued de Rham sequence, we obtain a complex

which is exact whenever the de Rham complex is.

The operator has a simple form. Using the definition of , we obtain for ,

where , is given by . Using the definition Equation 4.1 of the exterior derivative, the definition Equation 4.4 of , and the Leibniz rule

we obtain

Note that the operator is purely algebraic, and independent of .

Since , we have

or

Noting that

we find, using the identity

that is invertible with

We now define the desired subcomplex. Define

with projections and given by

Using Equation 4.8, it is straightforward to check that maps into and into , and that the diagram

commutes, and therefore the subcomplex in the bottom row is exact when the de Rham complex is. This subcomplex is, essentially, the elasticity complex. Indeed, by identifying elements with , and elements with , the subcomplex becomes

This complex may be identified with Equation 3.3. As an initial step of this identification we observe that the algebraic operator appearing in Equation 3.3 via Equation 3.4 and the operator are connected by the identity

where and are given by and for . In fact, using Equation 4.9, we have for any ,

On the other hand,

and hence Equation 4.12 follows from the algebraic identity

which holds for any .

We may further identify the four spaces of fields in Equation 3.3 with the corresponding spaces of forms in Equation 4.11 in a natural way:

.

given by .

given by .

given by , .

Under these identifications, we find that

corresponds to the row-wise gradient .

corresponds to the inclusion of .

corresponds to .

corresponds to the row-wise divergence .

corresponds to the operator .

Thus, modulo these identifications and the (unimportant) constant factor in the last identification, Equation 3.3 and Equation 4.11 are identical. Hence we have established the following result.

Theorem 4.1.

When the de Rham complex Equation 2.2 is exact, (i.e., the domain is contractible), then so is the elasticity complex Equation 3.3.

To end this section, we return to the operator defined by . Let be the adjoint of (with respect to the Euclidean inner product on and the Frobenius inner product on ), which is given by . Define by . Recall that the wedge product is given by

where the sum is over the set of all permutations of , for which and . This extends as well to differential forms with values in an inner product space, using the inner product to multiply the terms inside the summation. Using the Leibniz rule Equation 4.6, we have

We thus have

and

Subtracting these two expressions gives Equation 4.13.

For later reference, we note that, analogously to Equation 4.7, we have

5. The discrete construction

In this section we derive a discrete version of the elasticity sequence by adapting the construction of the previous section. To carry out the construction, we will use two discretizations of the de Rham sequence. For , let denote a finite-dimensional space of for which , and for which there exist projections which make the following diagram commute:

This is simply the diagram Equation 2.3 written in the language of differential forms. We do not make a specific choice of the discretization yet, but, as recalled in §2, there exist many such discrete de Rham complexes based on piecewise polynomials. In fact, as explained in Reference 7, for each polynomial degree we may choose to be the space of all piecewise polynomial -forms with respect to some simplicial decomposition of , and construct four such diagrams. We make the assumption that , which is true in all the cases mentioned.

Let be a second set of finite dimensional spaces with corresponding projection operators enjoying the same properties, giving us a second discretization of the de Rham sequence. Supposing a compatibility condition between these two discretizations, which we describe below, we shall construct a discrete elasticity complex.

We start with the complex

where denotes the -valued analogue of and similarly for . For brevity, we henceforth write for . As a discrete analogue of the operator , we define by where is the interpolation operator onto .

Next define by , for . Observe that the discrete version of Equation 4.8,

follows exactly as in the continuous case. From the commutative diagram Equation 5.1, we see that

Continuing to mimic the continuous case, we define the automorphism on by

and the operator by , which leads to

Thanks to the assumption that , we have . Hence, inserting the isomorphisms into Equation 5.2, we obtain

In analogy to the continuous case, we define

As in the continuous case, we can identify with , but, unlike in the continuous case, we cannot identify with , since we do not require that be invertible (it is not in the applications). Hence, in order to derive the analogue of the diagram Equation 4.10 we require a surjectivity assumption:

Under this assumption, the operator has a right inverse mapping into . This allows us to define discrete counterparts of the projection operators and by

and obtain the discrete analogue of Equation 4.10:

It is straightforward to check that this diagram commutes. For example, if , then

where the last equality follows from Equation 5.3. Thus the bottom row of Equation 5.6 is a subcomplex of the top row, and the vertical maps are commuting projections. In particular, when the top row is exact, so is the bottom. Thus we have established the following result.

Theorem 5.1.

For , let be a finite dimensional subspace of for which and for which there exist projections that make the diagram Equation 5.1 commute. Let be a second set of finite dimensional spaces with corresponding projection operators with the same properties. If maps onto , and the bottom row of Equation 5.1 is exact for both sequences and , then the discrete elasticity sequence given by the bottom row of Equation 5.6 is also exact.

The exactness of the bottom row of Equation 5.6 suggests that the following choice of finite element spaces will lead to a stable discretization of Equation 2.1:

In the next section we will make specific choices for the discrete de Rham complexes, and then verify the stability in the following section.

For use in the next section, we state the following result, giving a sufficient condition for the key requirement that be surjective.

Proposition 5.2.

If the diagram

commutes, then is surjective.

6. A family of discrete elasticity complexes

In this section, we present a family of examples of the general discrete construction presented in the previous section by choosing specific discrete de Rham complexes. These furnish a family of discrete elasticity complexes, indexed by an integer degree . In the next section we use these complexes to derive finite elements for elasticity. In the lowest order case, the method will require only piecewise linear functions to approximate stresses and piecewise constants to approximate the displacements and multipliers.

We begin by recalling the two principal families of piecewise polynomial spaces of differential forms, following the presentation in Reference 7. We henceforth assume that the domain is a contractible polyhedron. Let be a triangulation of . Let be a triangulation of by tetrahedra, and set

Here where is the Koszul differential defined by

The spaces correspond to the usual degree Lagrange piecewise polynomial subspaces of , and the spaces correspond to the usual degree subspace of discontinuous piecewise polynomials in . For and , the spaces correspond to the discretizations of and , respectively, presented by Nédélec in Reference 25, and the spaces are the ones presented by Nédélec in Reference 26. An element is uniquely determined by the following quantities:

Here is the set of vertices, edges, faces, or tetrahedra in the mesh , for , respectively, and for , we interpret . Note that for , naturally restricts on the face to an element of . Therefore, for , the wedge product belongs to and hence the integral of on the -dimensional face of is a well-defined and natural quantity. Using the quantities in Equation 6.1, we obtain a projection operator from to .

Similarly, an element is uniquely determined by

and so these quantities determine a projection.

If is a vector space, we use the notation and to denote the corresponding spaces of piecewise polynomial differential forms with values in . Furthermore, if is an inner product space, the corresponding degrees of freedom are given by Equation 6.1 and Equation 6.2, but where the test spaces are replaced by the corresponding valued spaces.

To carry out the construction described in the previous section we need to choose the two sets of spaces and for . We fix and set , , and , , , and . As explained in Reference 7, both these choices give a discrete de Rham sequence with commuting projections, i.e., a diagram like Equation 5.1 makes sense and is commutative.

We establish the key surjectivity assumption for our choice of spaces by verifying the commutativity of Equation 5.7.

Lemma 6.1.

Let and with projections , defined via the corresponding vector-valued moments of the form Equation 6.1 and Equation 6.2. If then

and so is surjective.

Proof.

We must show that for . Defining , the required condition becomes . Since , we have

(in fact Equation 6.4 holds for as well, but this is not used here). We must show that Equation 6.4 implies

From Equation 4.13, we have , where for , as is evident from Equation 4.14. Hence Equation 6.5 follows from Equation 6.4.

7. Stable mixed finite elements for elasticity

Based on the discrete elasticity complexes just constructed, we obtain mixed finite element spaces for the formulation Equation 2.1 of the elasticity equations by choosing , , and as the spaces of matrix and vector fields corresponding to appropriate spaces of forms in the - and -valued de Rham sequences used in the construction. Specifically, these are

In other terminology, may be thought of as the product of three copies of the Nédélec space of the second kind of degree , and and are spaces of all piecewise polynomials of degree at most with values in and , respectively, with no imposed interelement continuity. In this section, we establish stability and convergence for this finite element method.

The stability of the method requires the two conditions (A) and (A) stated in §2. The first condition is obvious since, by construction, , i.e., . The condition (A) is more subtle. We will prove a stronger version, namely,

(A)

for all nonzero , there exists nonzero with , and

where is the projection into and is a constant.

Recalling that and , and that the operator corresponds on the matrix level to , we restate condition (A) in the language of differential forms.

Theorem 7.1.

Given that , there exists such that and

where the constant is independent of and .

Before proceeding to the proof, we need to establish some bounds on projection operators. We do this for the corresponding scalar-valued spaces. The extensions to vector-valued spaces are straightforward. First we claim that

Here the constant may depend on the shape regularity of the mesh, but not on the meshsize. The second bound is obvious (with ), since is just the projection. The first bound follows by a standard scaling argument. Namely, let denote the reference simplex. For any , we have

where ranges over the faces of , over a basis for , and over a basis for . This is true because the integrals on the right hand side of Equation 7.4 form a set of degrees of freedom for (see Equation 6.1), and so we may use the equivalence of all norms on this finite dimensional space. We apply this result with , where is the projection defined to preserve the integrals on the right hand side of Equation 7.4. It follows that

where we have used a standard trace inequality in the last step. Next, if is an arbitrary simplex and , we map the reference simplex onto by an affine map , and define by

for any and any vectors . It is easy to check that , and that

Squaring and adding this over all the simplices in the mesh gives the first bound in Equation 7.3.

We also need a bound on the projection of a form in into . However, the projection operator is not bounded on , because its definition involves integrals over edges. A similar problem has arisen before (see, e.g., Reference 10), and we use the same remedy. Namely we start by defining an operator by the conditions

Note that, in contrast to , in the definition of , we have set the troublesome edge degrees of freedom to zero. Let be defined analogously on the reference element.

Now for , , so

where again ranges over the faces of , over a basis of , and over a basis of . But

where we have used Stokes’ theorem and the fact that the vanishing of the edge quantities in the definition of to obtain the first equality, and the face degrees of freedom entering the definition of to obtain the second. Similarly,

It follows that

When we scale this result to an arbitrary simplex and add over the mesh, we obtain

To remove the problematic in the last estimate, we introduce the Clement interpolant mapping into continuous piecewise linear -forms (still following Reference 10). Then

Defining by

we obtain

Thus we have shown that

Having modified to obtain the bounded operator , we now verify that the key property Equation 6.3 in Lemma 6.1 carries over to

where we now use the vector-valued forms of the projection operators. It follows easily from Equation 7.8, Equation 7.5, and Equation 7.6 that Equation 6.4 holds with , so that the proof of Equation 7.10 is the same as for Equation 6.3.

We can now give the proof of Theorem 7.1.

Proof of Theorem 7.1.

Given there exists such that with the bound (since maps onto ). Similarly, given there exists with with the bound . Let (recall that is an isomorphism) and set

We will now show that . From the definition of , we have

Then, using Equation 5.3, Equation 7.10, and the commutativity , we see

Combining, we get as desired. Furthermore, from the commutativity and the definition of , we get

and so we have established that .

It remains to prove the bound Equation 7.2. Using Equation 7.3, we have

Thus . Using Equation 7.9, we then get , and, using Equation 7.3, that . Therefore , while , and thus we have the desired bound Equation 7.2.

We have thus verified the stability conditions (A) and (A), and so may apply the standard theory of mixed methods (cf. Reference 12, Reference 13, Reference 17, Reference 20) and standard results about approximation by finite element spaces to obtain convergence and error estimates.

Theorem 7.2.

Suppose is the solution of the elasticity system Equation 1.5 and is the solution of discrete system Equation 2.1, where the finite element spaces , , and are given by Equation 7.1 for some integer . Then there is a constant , independent of , such that

If and are sufficiently smooth, then

Remark.

Note that the errors and depend on the approximation of both and . For the choices made here, the approximation of is one order less than the approximation of , and thus we do not obtain improved estimates, as one does in the case of the approximation of Poisson’s equation, where the extra variable does not enter.

8. A simplified element

Recall that the lowest order element in the stable family described above, for a discretization based on Equation 1.5, is of the form

The purpose of this section is to present a stable element which is slightly simpler than this one. More precisely, the spaces and are unchanged, but will be simplified from full linears to matrix fields whose tangential–normal components on each two-dimensional face of a tetrahedron are only a reduced space of linears.

We will still adopt the notation of differential forms. By examining the proof of Theorem 7.1, we realize that we do not use the complete sequence Equation 5.2 for the given spaces. We only use the sequences

The purpose here is to show that it is possible to choose subspaces of some of the spaces in Equation 8.1 such that the desired properties still hold. More precisely, compared to Equation 8.1, the spaces and are simplified, while the three others remain unchanged. If we denote by and the simplifications of the spaces and , respectively, then the properties we need are that

is a complex and that the surjectivity assumption Equation 5.5 holds, i.e., maps the space onto . Note that if , then maps onto .

We first show how to construct as a subspace of . Since the construction is done locally on each tetrahedron, we will show how to construct a space as a subspace of . We begin by recalling that the face degrees of freedom of have the form

We then observe that this six-dimensional space can be decomposed into

i.e., into forms with values in the tangent space to , or the normal space . This is a -dimensional decomposition. Furthermore,

where is in if and only if for orthonormal tangent vectors and . Finally, we obtain a -dimensional decomposition

where

In more explicit terms, if has the form

where and are orthonormal tangent vectors on , is the unit normal to , and is a tangent vector on , then we can write , with and , where

The reason for this particular decomposition of the degrees of freedom is that if we examine the proof of Lemma 6.1, where equation Equation 6.3 is established, we see that the only degrees of freedom that are used for an element are the subset of the face degrees of freedom given by

However, for , is given by . Since the general element can be written in the form , for a tangent vector, and thus . Hence, we have split the degrees of freedom into three on each face that we need to retain for the proof of Lemma 6.1 and three on each face that are not needed. The reduced space that we now construct has two properties. The first is that it still contains the space and the second is that the unused face degrees of freedom are eliminated (by setting them equal to zero). We can achieve these conditions by first writing an element as , where denotes the usual projection operator into defined by the edge degrees of freedom. Then the elements in will satisfy

i.e., their traces on the edges will be zero. Thus, they are completely defined by the face degrees of freedom

Since this is the case, we henceforth denote by .

We then define our reduced space

where denotes the set of forms satisfying

i.e., we have set the unused degrees of freedom to be zero.

Then

The degrees of freedom for this space are then given by

It is clear from this definition that the space will have 48 degrees of freedom (36 edge degrees of freedom and 12 face degrees of freedom). The unisolvency of this space follows immediately from the unisolvency of the spaces and .

The motivation for this choice of the space is that it easily leads to a definition of the space that satisfies the property that Equation 8.2 is a complex. We begin by defining

It is easy to see that this space will have 24 face degrees of freedom. Note this is a reduction of the space , since

We then define

It is clear that . The fact that the complex Equation 8.2 is exact now follows directly from the fact that the complex

is exact and the definition

We will define appropriate degrees of freedom for the space by using a subset of the 36 degrees of freedom for , i.e., of , . In particular, we take as degrees of freedom for ,

where denotes the set of that satisfy . It is easy to check that such will have the form

where .

Since is a six-dimensional space on each face, the above quantities specify 24 degrees of freedom for the space . To see that these are a unisolvent set of degrees of freedom for , we let , where and and set all degrees of freedom equal to zero. Then for , since

we see that by the unisolvency of the standard degrees of freedom for . In addition, for all and ,

Since , by the unisolvency of the degrees of freedom of the space .

Using an argument completely parallel to that used previously, it is straightforward to show that the simplified spaces also satisfy assumption Equation 5.5, i.e., that is onto.

To translate the degrees of freedom of the space to more standard finite element degrees of freedom, we use the identification of an element with a matrix given by . Then and . Since and hence is of the form Equation 8.5, we get on each face the six degrees of freedom

Finally, we note that the analogue of Theorem 7.2 holds with for the simplified spaces.

Acknowledgments

The authors are grateful to Geir Ellingsrud and Snorre Christiansen at CMA and to Joachim Schöberl, Johannes Kepler University Linz, for many useful discussions.

Mathematical Fragments

Equation (1.1)
Equation (1.3)
Equation (1.5)
Equation (2.1)
Equation (2.2)
Equation (2.3)
Equation (3.1)
Equation (3.2)
Equation (3.3)
Equation (3.4)
Equation (4.1)
Equation (4.4)
Equation (4.6)
Equation (4.7)
Equation (4.8)
Equation (4.9)
Equation (4.10)
Equation (4.11)
Equation (4.12)
Equation (4.13)
Equation (4.14)
Equation (5.1)
Equation (5.2)
Equation (5.3)
Equation (5.5)
Equation (5.6)
Proposition 5.2.

If the diagram

commutes, then is surjective.

Equation (6.1)
Equation (6.2)
Lemma 6.1.

Let and with projections , defined via the corresponding vector-valued moments of the form Equation 6.1 and Equation 6.2. If then

and so is surjective.

Equation (6.4)
Equation (6.5)
Equation (7.1)
Theorem 7.1.

Given that , there exists such that and

where the constant is independent of and .

Equation (7.3)
Equation (7.4)
Equations (7.5), (7.6), (7.7)
Equation (7.8)
Equation (7.9)
Equation (7.10)
Theorem 7.2.

Suppose is the solution of the elasticity system Equation 1.5 and is the solution of discrete system Equation 2.1, where the finite element spaces , , and are given by Equation 7.1 for some integer . Then there is a constant , independent of , such that

If and are sufficiently smooth, then

Equation (8.1)
Equation (8.2)
Equation (8.5)

References

Reference [1]
Scot Adams and Bernardo Cockburn, A mixed finite element method for elasticity in three dimensions, Journal of Scientific Computing 25 (2005), 515–521. MR 2221175 (2006m:65251)
Reference [2]
Mohamed Amara and Jean-Marie Thomas, Equilibrium finite elements for the linear elastic problem, Numer. Math. 33 (1979), no. 4, 367–383. MR 553347 (81b:65096)
Reference [3]
Douglas N. Arnold, Gerard Awanou, and Ragnar Winther, Finite elements for symmetric tensors in three dimensions, Submitted to Math. Comp.
Reference [4]
Douglas N. Arnold, Franco Brezzi, and Jim Douglas, Jr., PEERS: a new mixed finite element for plane elasticity, Japan J. Appl. Math. 1 (1984), no. 2, 347–367. MR 840802 (87h:65189)
Reference [5]
Douglas N. Arnold, Jim Douglas, Jr., and Chaitan P. Gupta, A family of higher order mixed finite element methods for plane elasticity, Numer. Math. 45 (1984), no. 1, 1–22. MR 761879 (86a:65112)
Reference [6]
Douglas N. Arnold and Richard S. Falk, A new mixed formulation for elasticity, Numer. Math. 53 (1988), no. 1-2, 13–30. MR 946367 (89f:73020)
Reference [7]
Douglas N. Arnold, Richard S. Falk, and Ragnar Winther, Differential complexes and stability of finite element methods. I: The de Rham complex, in Compatible Spatial Discretizations, D. Arnold, P. Bochev, R. Lehoucq, R. Nicolaides, and M. Shashkov, eds., IMA Volumes in Mathematics and its Applications 142, Springer-Verlag 2005, 23-46. MR 2249344
Reference [8]
Douglas N. Arnold, Richard S. Falk, and Ragnar Winther, Differential complexes and stability of finite element methods. II: The elasticity complex, in Compatible Spatial Discretizations, D. Arnold, P. Bochev, R. Lehoucq, R. Nicolaides, and M. Shashkov, eds., IMA Volumes in Mathematics and its Applications 142, Springer-Verlag 2005, 47-67. MR 2249345
Reference [9]
Douglas N. Arnold, Richard S. Falk, and Ragnar Winther, Finite element exterior calculus, homological techniques, and application, Acta Numerica (2006), 1-155. MR 2269741
Reference [10]
Douglas N. Arnold and Ragnar Winther, Mixed finite elements for elasticity, Numer. Math. 92 (2002), no. 3, 401–419. MR 1930384 (2003i:65103)
Reference [11]
I.N. Bernstein, I.M. Gelfand, and S.I. Gelfand, Differential operators on the base affine space and a study of –modules, Lie groups and their representation, I.M. Gelfand, ed., (1975), 21–64. MR 0578996 (58:28285)
Reference [12]
Franco Brezzi, On the existence, uniqueness and approximation of saddle-point problems arising from Lagrangian multipliers, Rev. Française Automat. Informat. Recherche Opérationnelle Sér. Rouge 8 (1974), no. R-2, 129–151. MR 0365287 (51:1540)
Reference [13]
Franco Brezzi and Michel Fortin, Mixed and Hybrid Finite Element Methods, Springer-Verlag, New York, 1991. MR 1115205 (92d:65187)
Reference [14]
Andreas Čap, Jan Slovák, and Vladimír Souček, Bernstein–Gelfand–Gelfand sequences, Ann. Math. (2) 154 (2001), 97–113. MR 1847589 (2002h:58034)
Reference [15]
Richard M. Christensen, Theory of Viscoelasticity, Dover Publications, 1982.
Reference [16]
Philippe G. Ciarlet, The finite element method for elliptic problems, North-Holland, Amsterdam, 1978. MR 0520174 (58:25001)
Reference [17]
Jim Douglas, Jr. and Jean E. Roberts, Global estimates for mixed methods for second order elliptic equations, Math. Comp. 44 (1985), no. 169, 39–52. MR 771029 (86b:65122)
Reference [18]
Michael Eastwood, A complex from linear elasticity, Rend. Circ. Mat. Palermo (2) Suppl. (2000), no. 63, 23–29. MR 1758075 (2001j:58033)
Reference [19]
Richard S. Falk, Finite element methods for linear elasticity, to appear in Mixed Finite Elements, Compatibility Conditions, and Applications, Lectures given at the C.I.M.E. Summer School held in Cetraro, Italy, June 26-July 1, 2006, Lecture Notes in Mathematics, Springer-Verlag.
Reference [20]
Richard S. Falk and John E. Osborn, Error estimates for mixed methods, R.A.I.R.O. Analyse numérique/Numerical Analysis, 14 (1980), no. 3, 249–277. MR 592753 (82j:65076)
Reference [21]
Baudoiun M. Fraejis de Veubeke, Stress function approach, Proc. of the World Congress on Finite Element Methods in Structural Mechanics, Vol. 1, Bournemouth, Dorset, England (Oct. 12-17, 1975), J.1–J.51.
[22]
V. Girault and P.-A. Raviart, Finite element methods for Navier-Stokes equations. Theory and algorithms, Springer Series in Computational Mathematics, 5, Springer-Verlag, Berlin, 1986. MR 851383 (88b:65129)
Reference [23]
Claes Johnson and Bertrand Mercier, Some equilibrium finite element methods for two-dimensional elasticity problems, Numer. Math. 30 (1978), no. 1, 103–116. MR 0483904 (58:3856)
Reference [24]
Mary E. Morley, A family of mixed finite elements for linear elasticity, Numer. Math. 55 (1989), no. 6, 633–666. MR 1005064 (90f:73006)
Reference [25]
Jean-Claude Nédélec, Mixed finite elements in , Numer. Math. 35 (1980), no. 3, 315–341. MR 592160 (81k:65125)
Reference [26]
Jean-Claude Nédélec, A new family of mixed finite elements in , Numer. Math. 50 (1986), no. 1, 57–81. MR 864305 (88e:65145)
Reference [27]
Erwin Stein and Raimund Rolfes, Mechanical conditions for stability and optimal convergence of mixed finite elements for linear plane elasticity, Comput. Methods Appl. Mech. Engrg. 84 (1990), no. 1, 77–95. MR 1082821 (91i:73045)
Reference [28]
Rolf Stenberg, On the construction of optimal mixed finite element methods for the linear elasticity problem, Numer. Math. 48 (1986), no. 4, 447–462. MR 834332 (87i:73062)
Reference [29]
—, A family of mixed finite elements for the elasticity problem, Numer. Math. 53 (1988), no. 5, 513–538. MR 954768 (89h:65192)
Reference [30]
—, Two low-order mixed methods for the elasticity problem, The mathematics of finite elements and applications, VI (Uxbridge, 1987), Academic Press, London, 1988, pp. 271–280. MR 956898 (89j:73074)
Reference [31]
Vernon B. Watwood Jr. and B. J. Hartz, An equilibrium stress field model for finite element solution of two–dimensional elastostatic problems, Internat. Jour. Solids and Structures 4 (1968), 857–873.

Article Information

MSC 2000
Primary: 65N30 (Finite elements, Rayleigh-Ritz and Galerkin methods, finite methods)
Secondary: 74S05 (Finite element methods)
Keywords
  • Mixed method
  • finite element
  • elasticity
Author Information
Douglas N. Arnold
Institute for Mathematics and its Applications, University of Minnesota, Minneapolis, Minnesota 55455
arnold@ima.umn.edu
MathSciNet
Richard S. Falk
Department of Mathematics, Rutgers University, Piscataway, New Jersey 08854-8019
falk@math.rutgers.edu
Ragnar Winther
Centre of Mathematics for Applications and Department of Informatics, University of Oslo, P.O. Box 1053, Blindern, 0316 Oslo, Norway
ragnar.winther@cma.uio.no
MathSciNet
Additional Notes

The work of the first author was supported in part by NSF grant DMS-0411388.

The work of the second author was supported in part by NSF grant DMS03-08347.

The work of the third author was supported by the Norwegian Research Council.

Journal Information
Mathematics of Computation, Volume 76, Issue 260, 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 2007 American Mathematical Society; reverts to public domain 28 years from publication
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/S0025-5718-07-01998-9
  • MathSciNet Review: 2336264
  • Show rawAMSref \bib{2336264}{article}{ author={Arnold, Douglas}, author={Falk, Richard}, author={Winther, Ragnar}, title={Mixed finite element methods for linear elasticity with weakly imposed symmetry}, journal={Math. Comp.}, volume={76}, number={260}, date={2007-10}, pages={1699-1723}, issn={0025-5718}, review={2336264}, doi={10.1090/S0025-5718-07-01998-9}, }

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.