Each averaging technique yields reliable a posteriori error control in FEM on unstructured grids. Part I: Low order conforming, nonconforming, and mixed FEM

By Carsten Carstensen and Sören Bartels

Abstract

Averaging techniques are popular tools in adaptive finite element methods for the numerical treatment of second order partial differential equations since they provide efficient a posteriori error estimates by a simple postprocessing. In this paper, their reliablility is shown for conforming, nonconforming, and mixed low order finite element methods in a model situation: the Laplace equation with mixed boundary conditions. Emphasis is on possibly unstructured grids, nonsmoothness of exact solutions, and a wide class of averaging techniques. Theoretical and numerical evidence supports that the reliability is up to the smoothness of given right-hand sides.

1. Introduction

Error control and efficient mesh-design in finite element simulations of computational engineering and scientific computing finite element simulations is frequently based on a posteriori error estimates. One of the more popular techniques is local or global averaging, e.g., in form of the ZZ-error indicator Reference ZZ. Efficiency and reliability of this estimator were known only for very structured grids and for solutions of higher regularity and then we have even asymptotic exactness Reference V. Numerical experiments in Reference Baetal showed that averaging techniques were quite more reliable on irregular meshes than expected. For homogeneous Dirichlet conditions and conforming finite element methods, the reliability and efficiency of the ZZ-estimator is proven on unstructured, merely shape-regular grids Reference R2.

This work is devoted to give theoretical and numerical support for the robust reliability of all averaging techniques, robust with respect to violated (local) symmetry of meshes and superconvergence and robust with respect to other boundary conditions or other finite element methods.

For a more precise description of averaging techniques, let us discuss a discretisation of a conservation equation

with a given right-hand side and a known approximation to the unknown exact flux in the bounded Lipschitz domain with piecewise flat boundary. The test function finite element space should include the continuous piecewise affines (with homogeneous Dirichlet boundary conditions) based on a regular (in the sense of Reference Ci, cf. Section 2) triangulation of . Suppose a Galerkin property for with , i.e.,

What can be said about the error when we regard as an unknown and as a known variable?

In averaging techniques, the error estimator is based on a smoother approximation, e.g., in , the continuous -piecewise linears, to the (components of the) discrete solution . For instance,

may serve as a computable error estimator and the elementwise contributions as local error indicators in an adaptive mesh-refining algorithm.

The triangle inequality shows that is efficient with constant up to higher order terms of the exact solution , indeed,

The last term converges as (provided is smooth enough and denotes the maximal mesh–size in ) and so, generically, is of higher order than the error in the lowest order finite element method. If the second term in the right-hand side of Equation 1.4 fails to be of higher order, one can still prove efficiency of using equivalence of global and local averaging (cf. Theorem 3.2) and that local averaging is equivalent to weighted jumps across interelement boundaries. An efficiency estimate with higher order terms that depend on local smoothness of right-hand sides but with unknown constants then follows as in Reference V.

In practise, we may apply an averaging operator to and compute the upper bound of . Then, efficiency depends strongly on the approximation properties of and deserves further investigation.

In this paper, the focus is on the reliability of , i.e., we investigate under which conditions an estimate

holds, we study what the constant depends on, what affects the higher-order contributions ”, how to modify the definition of in the presence of mixed boundary conditions, and how to modify the general setting presented for nonconforming and mixed lowest order finite element methods.

Recall from Equation 1.5 that any averaging technique, described by , then is reliable up to higher order terms. We also prove equivalence to local modifications of where the minimisation is over smaller domains, e.g., patches of nodes or edges.

The outline of the paper is as follows. Preliminaries and notation are introduced in Section 2 where we state and prove stability and first order estimates for a certain approximation operator essentially designed to yield further local orthogonality properties as in Reference CVReference C2. Basic estimates are provided in Section 3 for a local and global averaging technique and their equivalence. The subsequent Sections 4, 5, and 6 display the consequences to averaging techniques in a posteriori error control for first order conforming, nonconforming and mixed finite element schemes. Numerical evidence, reported in Section 7, supports the theoretical results for adaptively refined and evenly perturbed meshes. Although asymptotic exactness is not claimed in this paper, our numerical experiments illustrate that is a very good approximate to even on perturbed grids.

The proofs are given for a simple elliptic model example with mixed boundary conditions for conforming, nonconforming, and mixed finite elements in two dimensions for notational simplicity. More interesting examples such as higher-order schemes, the application to the Stokes problem or the Navier-Lamé equations without incompressibility-locking will appear elsewhere Reference BCReference CF2Reference CF3Reference CF4Reference CF5.

2. Approximation in finite element spaces

The Lipschitz boundary of the bounded domain is split into a closed Dirichlet part with positive surface measure and a remaining, relatively open and possibly empty, Neumann part . Suppose be a regular triangulation of the domain , , in the sense of Ciarlet Reference BSReference Ci (no hanging node, domain is matched exactly) with piecewise affine Lipschitz boundary , i.e., consists of a finite number of closed subsets of , that cover . Each element is either an interval if , a triangle , or a parallelogram if . The extremal points are called vertices, the faces , e.g., , are called edges. The set of all vertices and all edges appearing for some in are denoted as and . Two distinct and intersecting and share either an entire edge or a vertex. Each edge on the boundary belongs either to , written , or to , written . Therefore the set of edges is partitioned into , , and . We stress that , the union of all egdes, denotes the skeleton of egdes in , i.e., the set of all points that belong to some boundary of some element . Finally, denotes the set of free nodes.

We do not explicitly distinguish between nodes and vertices when we consider conforming finite elements (and avoid these concepts for nonconforming or mixed finite element schemes).

For , let if is a triangle or if is a parallelogram. Here, , resp. , denotes the set of algebraic polynomials in variables on of total, resp. partial, degree . The space of (possibly discontinuous) -piecewise polynomials of degree is the set of all with for all in . Set

Let denote the nodal basis of , i.e., satisfies if and . Note that is a partition of unity and the open patches

form an open cover of with finite overlap.

In order to define a weak interpolation operator , we modify to a partition of unity . For each fixed node , we choose a node and let if . In this way, we define a partition of into classes , . For each set

and notice that is a partition of unity. It is required that

is connected and that implies that has a positive surface measure.

For and let be

and then define

The local mesh-sizes are denoted by and , where denotes the element-size, for , and the edge-size is defined on the union or skeleton of all edges in by . The patch-size is defined for each node separately.

Theorem 2.1.

There exist -independent constants such that for all and there holds

The constants only depend on , , and the shape of the elements and patches (not on their sizes).

Remark 2.1.

The assertion of the theorem holds verbatim for three space dimensions where consists of tetrahedra or parallelepipeds with the same proof.

Proof.

In this proof and at similar occasions, abbreviates an inequality up to a constant -independent factor. Also, abbreviates and we neglect if is meant, i.e., . Hence, e.g., Equation 2.6 could be phrased as .

The key estimate for the stability and the approximation property of will be

For the proof of Equation 2.10, let denote the integral mean of on . Then, using the definition Equation 2.4 for the coefficients , Cauchy’s and Young’s inequality, we infer, with ,

Absorbing , we deduce

A Poincaré inequality yields

Note that is nonzero only if has positive surface measure. Since vanishes there, Friedrichs’ inequality shows

Therefore, (Equation 2.12) results in

To prove (Equation 2.10), we use the triangle inequality, (Equation 2.13), and again Cauchy’s and Friedrichs’ inequality to verify

To prove (Equation 2.7), we use that is a partition of unity and obtain with (Equation 2.10), (Equation 2.4) for any that

In the last step we used that has a finite overlap that depends on the shape of the elements only. This concludes the proof of (Equation 2.7).

The remaining part of the proof uses standard arguments and is therefore sketched for brevity. To prove (Equation 2.8) we let and , , in (Equation 2.7). To verify (Equation 2.6) we use and and repeat the triangle inequality several times for

With Friedrichs’ and Poincaré’s inequality we infer

It remains to estimate with Equation 2.10, Friedrichs’ inequality, and the above arguments. A trace inequality Reference BSReference ClReference CF1 of the form

for and with together with Equation 2.6 and Equation 2.8 implies Equation 2.9.

3. Basic estimates

In this section we first derive with the approximation operator a global error estimate for a posteriori error control by averaging processes in an abstract setting. We then show the equivalence of local and global averaging techniques. The estimates of this section are then specified, and thereby proved to be substantial, in the subsequent sections to conforming, nonconforming, and mixed finite element methods.

Theorem 3.1.

Suppose and with and

Then there holds

Proof.

According to Equation 3.1, Equation 2.6, Cauchy’s inequality, and an integration by parts we have, for each with , that

since and vanish on . Owing to Equation 2.7 and Equation 2.9 in Theorem 2.1, we conclude Equation 3.2 from Equation 3.3 and Cauchy’s inequality.

The second result justifies local averaging. For each edge , let and for the two distinct elements with and for each edge , let and for the element with . Let denote the (possibly discontinuous) -piecewise polynomials of degree on and let .

Theorem 3.2.

There exists an -independent constant which depends on the shape of the elements in and on the polynomial degree , , such that, for all , we have

Proof.

The upper estimate follows from for all and a rearrangement of the sums over edges and elements. To verify the lower estimate in Equation 3.4 we consider a subspace of ,

where denotes the restriction of the triangulation to . Since is a closed convex subset of , the best-approximation problem

defines an orthogonal relation, namely, for all ,

where , , denotes the minimiser in Equation 3.5. From , (Equation 3.6), and Cauchy’s inequality we deduce, for arbitrary ,

For each , we consider the semi-norms on a finite dimensional subspace of

Then, Equation 3.7 and yield

We claim . For a proof, suppose . Then, for each that is an inner edge of , we have on the open set for some . Since , we find that . The set of all such is a cover of and there is a sequence of inner edges such that , so that we deduce . Moreover, on each edge with , while for edges with we have . Altogether, we deduce . A compactness and scaling argument then shows our claim

Utilizing Equation 3.9 in Equation 3.8, we conclude

Remark 3.1.

The assertions of Theorems 3.1 and 3.2 hold verbatim for three space dimensions where consists of tetrahedra or parallelepipeds with the same proofs.

4. Applications to conforming finite element schemes

Given right-hand sides , , and , let denote the unique weak solution to

Suppose a finite element scheme, based on a regular triangulation , provided a discrete flux to the exact flux such that , for all and

Theorem 4.1.

There exists an -independent constant (that depends on and the shape of the elements and patches) such that

In the infimum, stands for all with on .

Proof.

Abbreviate and let . Assume that satisfies on and . Recall and . Then Equation 4.1-Equation 4.4 imply Equation 3.1. Hence, we may choose and in Theorem 3.1 to obtain with Cauchy’s inequality for the second term that

Since , we can divide Equation 4.6 by to verify

Let denote the -piecewise action of the -operator. The triangle inequality in the last summand in Equation 4.7 and for and and a summation over elements show

Note that for our choices of . A -elementwise inverse estimate shows (with a constant that depends on the shape of the finite elements only). Utilising this in Equation 4.7Equation 4.8, we deduce Equation 4.5.

Remark 4.1.

In the proof of Theorem 4.1 we used the assumption that is of lowest order, i.e., , for the purpose of estimating by . We refer to Reference BC for related error estimates for higher order methods.

The subsequent lemma shows that is a higher order term.

Lemma 4.1.

Suppose that for all . Then there exists an -independent constant (that depends on the shapes of the elements only) such that

If , we have

Proof.

Let belong to some and denote . We determine by extending the boundary values and . Note that is continuous on since interpolates at each node on . An harmonic extension of to yields

where we applied an interpolation estimate. A one-dimensional integration argument shows . Consequently,

A scaling argument guarantees that the constant in Equation 4.12 is -independent. Defining by on elements on and by zero on other elements then shows the lemma. The second estimate follows from .

Lemma 4.2.

Suppose and, for each node where the outer unit normal on is continuous (hence constant in a neighbourhood of as is a polygon), let be continuous. Then, the set

is nonvoid and, for each ,

Proof.

Elementary estimates on each edge on verify Equation 4.14; the proof of follows from an explicit construction in Example 4.1.

Example 4.1.

We define an operator by

where for while we incorporate for . In case for two distinct edges with distinct outer unit normals , on , at a corner we choose to be the unique solution of the linear system

In the remaining cases for or with two parallel edges with the unit tangent vector let solve

The following corollary is Equation 1.5 with a constant as in Theorem 4.1 and with specified higher order terms from Lemma 4.1 and 4.2 and a Poincaré inequality.

Corollary 4.1.

Under the conditions of Theorem 4.1 and Lemmas 4.1 and 4.2 we have for that

The -independent constant depends on the shape of the elements and patches only.

Remark 4.2.

Let us emphasise that the derivatives along are required only -piecewisely while needs to be patch-wise (not only elementwise) in and so . For a nonsmooth right-hand side , may be replaced by a patch-wise –best approximation error in the approximation through constants of (cf. Equation 2.7).

The global averaging process might be too expensive or its approximation may be inefficient and hence a local averaging process of interest. Recall that is the (interior of the) union of all elements in that share the edge .

Corollary 4.2.

Under the conditions of Theorem 4.1 and Lemmas 4.1 and 4.2 we have for that

The -independent constant depends on the shape of the elements and patches only.

Proof.

Theorem 4.1, Lemma 4.1, an approximation of as in Lemma 4.2, and a Poincaré inequality show

This and the first inequality of Theorem 3.2 imply the assertion.

Remark 4.3.

The results of this section hold also in three dimensions where consists of tetrahedra or parallelepipeds. The proofs of some details as Lemma 4.1 or Lemma 4.2 require much more technical preparations and so are omitted in this overview.

Remark 4.4.

It is shown in Reference CVReference C2 that the edge-contributions (jump differences in the normal fluxes components across edges) dominate in standard residual a posteriori error estimates Reference BaRReference BReference BSReference CF1Reference EEHJReference V. Arguing as in Reference R1Reference R2Reference DMR, one can hence derive alternative proofs of Equation 4.18 and then of Equation 4.17.

Remark 4.5.

In an -estimate of Reference HSWW it is suggested to average over a domain of size instead of merely over patches or the entire domain to obtain asymptotic exact results.

5. Applications to nonconforming finite element schemes

In the Laplace problem with mixed boundary conditions Equation 4.1Equation 4.3, we suppose that the discrete flux , where denotes the -piecewise application of the gradient, satisfies

The usual conformity conditions read for all ,

where denotes the jump of across and denotes on . Those conditions are satisfied by construction for Crouzeix–Raviart finite elements of lowest order.

Remark 5.1.

It is stressed that is a conforming test function space which is included in the nonconforming finite element spaces for triangles or tetrahedra. For parallelograms, Equation 5.1 means that the polynomial degrees are at least of second order to include the conforming term . This technical detail could actually be dropped since the contribution from an enhanced finite element space leads to a higher order term Reference KS. We restrict our analysis to triangles or tetrahedra for simplicity.

Theorem 5.1.

Suppose that is connected and that belongs to only one connectivity component of . Then, there exists an -independent constant (that depends on and the shape of the elements and patches) such that

Here, denotes the unit tangent vector on .

Remark 5.2.

The following lemma is based on the Helmholtz decomposition of a vector field. The decomposition is available in three dimensions as well (cf., e.g., [GR]) but the notation is more involved so we restrict the discussion to the two-dimensional setting for brevity.

Lemma 5.1.

For all , there exist that satisfy boundary conditions and is constant such that

Proof.

The lemma follows from the Helmholtz decomposition where solves and with proper boundary conditions (cf., e.g., Reference GR).

Proof of Theorem 5.1.

For and , Lemma 5.1 yields

Since is connected, we may and will assume without loss of generality that on . According to Equation 4.1Equation 4.3 and Equation 5.1, we infer Equation 3.1 and hence may choose and, in case , in Theorem 3.1 to obtain

The estimate of the last term in Equation 5.5 will follow from Theorem 3.1 as well once we establish an analogy to Equation 3.1, namely

where . It is essential to notice that is constant and has a vanishing integral on any edge. An elementwise integration by parts on the left-hand side of Equation 5.7 yields volume terms and edge terms whose integral vanishes on any (the case is indicated and the assertion is true for as well; on shows it for ). In this way we establish Equation 5.7.

To employ Theorem 3.1, we interchange components, writing in this proof for vectors, and we interchange the role of the boundaries and adopt Theorem 2.1 and 3.1 where acts as the Dirichlet boundary and acts as the Neumann boundary. Writing and , Equation 5.7 reads for all and this is Equation 3.1. Reading Theorem 3.1 in the present notation, we obtain

with . In the second last term, denotes the unit tangent vector and in the last term we used that .

The remaining arguments are similar to those in the proof of Theorem 4.1 and hence are omitted.

In contrast to the conforming situation, Theorem 5.1 demands averaging functions to satisfy some conditions on the Dirichlet boundary.

Lemma 5.2.

Suppose and, for each node where the outer unit normal on is continuous, let be continuous. Then, the set

is nonvoid and, for each ,

Proof.

Similar to Equation 4.10 in Lemma 4.1 or 4.2.

Example 5.1.

Assume the conditions of Lemmas 4.2 and 5.2 on the data and . We define an operator by Equation 4.15 and for . In case we preceed as in Equation 4.16a, resp. Equation 4.16b. In case we consider the analogous systems

(cf. Equation 4.16a and notation from Example 4.1), resp., as an analog to Equation 4.16b,

For we require a compatibility condition if , namely . Then, we define as in Equation 5.11b when (the case is analogous). For , we need no further compatibility of the data and solve the linear system

(Here ; the case is analogous.)

The modification of Equation 1.5 in the nonconforming setting is a direct consequence of Theorem 5.1, Lemma 5.2 and Example 5.1. Note that Corollary 4.1 is a special case apart from the different treatment of the Dirichlet boundary conditions.

Corollary 5.1.

Under the conditions of Theorem 5.1 and Lemmas 4.2 and 5.2, we have for that

The analog to Corollary 4.2 concludes this section on lowest order Raviart–Crouzeix finite elements.

Corollary 5.2.

Under the conditions of Theorem 5.1 and Lemmas 4.2 and 5.2, there exists a constant such that we have for

Here, denotes an approximation of as in Lemma 5.2, i.e., on for some .

Remark 5.3.

The results of this section can be generalised to three space dimensions as all the required tools such as a Helmholtz decomposition are available then as well. Details on the three-dimensional case are omitted for notational simplicity.

Remark 5.4.

Arguing as in Reference CVReference C2, one can prove that edge contributions (jumps in the fluxes across edges) dominate the residual based error estimates from Reference DDPVReference C2Reference KS. Arguing in the spirit of Reference R1Reference R2Reference DMR, one can hence derive alternative proofs of Equation 5.14 and then of Equation 5.13.

6. Applications to mixed finite element schemes

In the Laplace problem with mixed boundary conditions Equation 4.1Equation 4.3, we suppose that the discrete flux and the displacement approximation satisfy, for all with on , and for all and that

Remark 6.1.

Standard mixed finite element methods of any order, such as Raviart–Thomas (RT), Brezzi–Douglas–Marini (BDM), or Brezzi–Douglas–Fortin–Marini (BDFM) elements (cf. Reference BF for details), provide Equation 6.1Equation 6.3 Reference C1.

Theorem 6.1.

Suppose that is connected and that belongs to only one connectivity component of and let , i.e., for all . Then, there exists an -independent constant (that depends on and the shape of the elements and patches) such that

Proof.

Lemma 5.1 provides Equation 5.5, and we may and will assume without loss of generality that on . An integration by parts and Equation 6.2Equation 6.3 show for the -piecewise integral mean of that

The second last term is estimated with an elementwise Poincaré inequality while the last term in Equation 6.5 involves a trace theorem Reference BSReference CF1Reference Cl, namely

for on the triangle and the edge , . With a second application of Poincaré’s inequality, Equation 6.6, and Cauchy’s inequality we show that

The second contribution on the right-hand side of Equation 5.5 is analysed with Theorem 3.1, where, as in the proof of Theorem 5.1, we interchange components and the role of the boundary conditions. As already employed in Reference C1Reference C2, for all . Moreover, on . Hence, Equation 6.1 and an integration by parts for yield Equation 5.7 because of Equation 4.2. Arguing as in the proof of Theorem 5.1, we deduce for arbitrary that

The remaining details are analogous to the proof of Theorem 5.1 and hence are omitted.

The precise version of Equation 1.5 for lowest order mixed finite element methods is summarised as follows.

Corollary 6.1.

Suppose that the discrete flux satisfies , and . Then,

Proof

Combine Theorem 6.1 and Lemmas 4.1 and 5.2, and use an inverse estimate to prove

Remark 6.2.

The assumptions in Corollary 6.1 are satisfied for lowest order Raviart–Thomas and Brezzi–Douglas–Fortin–Marini finite elements.

Example 6.1.

Assume the conditions of Lemma 5.2 on the data . We define an operator by Equation 4.15 and for . In case we consider systems

(cf. Equation 4.16a and notation from Example 4.1), resp., as an analog to Equation 4.16b,

A local version follows from Theorem 3.2 and concludes this section on mixed finite element methods.

Corollary 6.2.

Under the conditions of Theorem 6.1 and Corollary 6.1 we have

Remark 6.3.

The results of this section could be generalised to three space dimensions. Details are omitted for brevity.

Remark 6.4.

For related residual based a posteriori error estimates we refer to Reference AReference BVReference C1Reference C2Reference HW.

7. Numerical experiments

The theoretical results of this paper are supported by numerical experiments. In this section, we report on two examples of the problem Equation 4.1Equation 4.3 on uniform, adapted, and perturbed meshes for conforming, nonconforming, and mixed finite element methods.

Example 7.1.

Let on the L-shaped domain , on the Dirichlet boundary , and on the Neumann boundary ,

using polar coordinates . The exact solution of Equation 4.1Equation 4.3 has a typical corner singularity at the origin. In this example, the right-hand sides are smooth, but the solution is not. The coarsest triangulation consists of three squares halved by diagonals parallel to the vector (cf. Figure 1).

Example 7.2.

Let for the function

on the unit square and set on the entire boundary (). The solution to Equation 4.1Equation 4.3 is -regular but (although theoretically smooth) has huge gradients on the circle with radius around . The coarsest triangulation consists of four squares halved by diagonals parallel to the vector (cf. Figure 3).

The following adaptive algorithm generates all the sequences of meshes in this paper which are uniform for or adapted for in Equation 7.2. Since the resulting meshes might show local symmetries, we considered meshes that are either unperturbed (relative to ) for and randomly perturbed for in step (e). The implementation was performed in Matlab in the spirit of Reference ACF with a direct solution of linear systems of equations. For details on the red-blue-green refinements we refer to Reference V.

Algorithm

(a)

Start with a coarse mesh , .

(b)

Compute the discrete solution on the actual mesh .

(c)

Compute error indicators

for all and plot energy error and its estimator versus the degree of freedom of the triangulation .

(d)

Mark the element for red-refinement provided

(e)

Mark further elements (red-blue-green refinement) to avoid hanging nodes. Generate a new triangulation using edge-midpoints if and points on the edges at a random distance at most from the edge-midpoints if . Perturbe the nodes of the mesh at random with values taken uniformly from a ball around of radius . Correct boundary nodes by orthogonal projection onto that boundary piece they are expected such that are matched by the resulting mesh exactly. Update and go to (b).

7.1. Results for conforming finite element methods

In the conforming finite element scheme, we use operator from Example 4.1 in Equation 7.1 of Algorithm and report on results obtained for (uniform), (adaptive), and (adaptive, perturbed).

Some meshes obtained for Example 7.1 are shown in Figure 1 and illustrate a high automatic mesh-refinement of the adapted meshes towards the origin, which is expected to improve the convergence rate of possibly to the optimal value . The result of the perturbation in step (e) of Algorithm is seen in the right half of Figure 1. We believe that the meshes generated by Algorithm have less local symmetry than that by . According to local extrapolation, symmetry could cause superconvergence phenomena. To check the practical convergence behaviour, we plotted in Figure 2 for each mesh an entry and . A log-scaling on both axes allows a slope of a straight line in the plot that connects two subsequent entries for a series of meshes generated by Algorithm , to be interpreted as an experimental convergence rate (owing to in two dimensions). We observe experimental convergence rates , resp. , for uniform, resp. adapted, meshes (generated by Algorithm for , resp. ). Furthermore, even for coarse meshes, appears to be a very good approximation to ; corresponding entries almost coincide for . If these meshes are perturbed (cf. Figure 1), the quotient is almost a constant very close to . Numerical checks with different numerical quadrature rules (used to evaluate ) convinced us that, in general, behaves not asymptotically exact in practise but is very accurate.

In Example 7.2 we obtained meshes and experimental convergence rates displayed in Figures 3 and 4. Although belongs to and we expect linear convergence, has huge second order derivatives along a circular arc where is steep. We observe high refinements in the adapted meshes towards this arc. In this example the preasymptotic range is very large, an experimental convergence rate can be observed only for for all refinement strategies. In this regime the estimator appears as a good approximate for and the entries and almost conincide for . This is not the case for , but the quotient is still close to .

7.2. Results for nonconforming finite element methods

The operator from Example 5.1 serves in Equation 7.1 to define for first order Crouzeix–Raviart finite elements (cf., e.g., Reference BSReference Ci) in Algorithm . The generated meshes look similar to those shown in Figure 1, resp. Figure 3, and therefore are not displayed in this paper. The experimental convergence rates for Example 7.1, resp. 7.2, are illustrated in Figure 5, resp. Figure 6. The overall picture appears similarly to the above discussions and we draw the same conclusions. For uniform meshes, the quotient is nearly constant but significantly larger than in Figure 2.

7.3. Results for mixed finite element methods

For the Raviart–Thomas finite element method (cf. Reference BReference BFReference BS) we use the operator from Example 6.1 to define . The adapted meshes look similar to those shown in Figures 1 and 3 and therefore are not displayed in this paper. Figures 7 and 8 display the error and the estimator for the mixed finite element scheme in Examples 7.1 and 7.2 obtained from Algorithm and . We obtain the same experimental convergence rates as in the previous methods.

7.4. Remarks

(i)

Our overall experience with Algorithm and other (e.g., residual-based) adaptive algorithms supports that all such adaptive algorithms yield a considerable convergence improvement.

(ii)

Although asymptotic exactness of is not observed, the reliability constant in Equation 1.5 and the efficiency constant are experimentally very close to since is a very good approximation to for very fine meshes (i.e., when h.o.t. is neglegible, say, for ).

(iii)

Note that the efficiency constant is not known to be one as (based on the averaging operator ) is different from

For conforming linear triangular finite elements, the efficiency of with an averaging operator follows from Reference R1Reference R2.

(iv)

Instead of the averaging operator , we tested the error estimator from Equation 7.3 and found that sometimes the performance is poorer than that of : In Figure 4, for instance, the results of are much smaller than those of . The averaging technique suggested in Reference HSWW; average over a domain of size , is expected to give values between and .

(v)

The error estimation in Example 7.1 is very accurate even for very coarse meshes. Hence the higher order terms do not seem to be important here although is nonsmooth. This agrees with our theoretical prediction in Equation 1.5 since and are zero and is piecewise analytic. For a generic corner singularity of , we expect

and so the h.o.t. in Equation 1.4 are not expected to be dominant even for coarse meshes.

(vi)

For coarse meshes in Example 7.2, higher order terms may cause the overall observation that is much smaller than . Assuming (which is seen for fine meshes, whence for neglegible h.o.t.) nonsmooth data ( is large) indicate that cannot be improved to .

Acknowledgments

The second author (S.B.) thankfully acknowledges the partial support by the German research foundation at the Graduiertenkolleg 357 “Effiziente Algorithmen und Mehrskalenmethoden”.

Figures

Figure 1.

Adaptively refined meshes (left upper) to (right lower) (left) and perturbed triangulation with free nodes (right) in Example 7.1 for the conforming finite element scheme.

Graphic without alt text
Figure 2.

Error and error estimator for uniform, adaptive, and perturbed adaptive mesh-refinement in the conforming finite element scheme in Example 7.1.

Graphic without alt text
Figure 3.

Adaptively refined meshes (left upper) to (right lower) (left) and perturbed triangulation with free nodes (right) for the conforming finite element scheme in Example 7.2.

Graphic without alt text
Figure 4.

Error and error estimator for uniform, adaptive, and perturbed adaptive mesh-refinement in the conforming finite element scheme in Example 7.2.

Graphic without alt text
Figure 5.

Error and error estimator for uniform, adaptive, and perturbed adaptive mesh-refinement in the nonconforming finite element scheme in Example 7.1.

Graphic without alt text
Figure 6.

Error and error estimator for uniform, adaptive, and perturbed adaptive mesh-refinement in the nonconforming finite element scheme in Example 7.2.

Graphic without alt text
Figure 7.

Error and error estimator for uniform, adaptive, and perturbed adaptive mesh-refinement in the mixed finite element scheme in Example 7.1.

Graphic without alt text
Figure 8.

Error and error estimator for uniform, adaptive, and perturbed adaptive mesh-refinement in the mixed finite element scheme in Example 7.2.

Graphic without alt text

Mathematical Fragments

Equation (1.4)
Equation (1.5)
Equation (2.4)
Theorem 2.1.

There exist -independent constants such that for all and there holds

The constants only depend on , , and the shape of the elements and patches (not on their sizes).

Equation (2.10)
Equation (2.12)
Equation (2.13)
Theorem 3.1.

Suppose and with and

Then there holds

Equation (3.3)
Theorem 3.2.

There exists an -independent constant which depends on the shape of the elements in and on the polynomial degree , , such that, for all , we have

Equation (3.5)
Equation (3.6)
Equation (3.7)
Equation (3.8)
Equation (3.9)
Equations (4.1), (4.2), (4.3)
Equation (4.4)
Theorem 4.1.

There exists an -independent constant (that depends on and the shape of the elements and patches) such that

In the infimum, stands for all with on .

Equation (4.6)
Equation (4.7)
Equation (4.8)
Lemma 4.1.

Suppose that for all . Then there exists an -independent constant (that depends on the shapes of the elements only) such that

If , we have

Equation (4.12)
Lemma 4.2.

Suppose and, for each node where the outer unit normal on is continuous (hence constant in a neighbourhood of as is a polygon), let be continuous. Then, the set

is nonvoid and, for each ,

Example 4.1.

We define an operator by

where for while we incorporate for . In case for two distinct edges with distinct outer unit normals , on , at a corner we choose to be the unique solution of the linear system

In the remaining cases for or with two parallel edges with the unit tangent vector let solve
Corollary 4.1.

Under the conditions of Theorem 4.1 and Lemmas 4.1 and 4.2 we have for that

The -independent constant depends on the shape of the elements and patches only.

Corollary 4.2.

Under the conditions of Theorem 4.1 and Lemmas 4.1 and 4.2 we have for that

The -independent constant depends on the shape of the elements and patches only.

Equation (5.1)
Theorem 5.1.

Suppose that is connected and that belongs to only one connectivity component of . Then, there exists an -independent constant (that depends on and the shape of the elements and patches) such that

Here, denotes the unit tangent vector on .

Lemma 5.1.

For all , there exist that satisfy boundary conditions and is constant such that

Equation (5.5)
Equation (5.7)
Lemma 5.2.

Suppose and, for each node where the outer unit normal on is continuous, let be continuous. Then, the set

is nonvoid and, for each ,

Example 5.1.

Assume the conditions of Lemmas 4.2 and 5.2 on the data and . We define an operator by Equation 4.15 and for . In case we preceed as in Equation 4.16a, resp. Equation 4.16b. In case we consider the analogous systems

(cf. Equation 4.16a and notation from Example 4.1), resp., as an analog to Equation 4.16b,

For we require a compatibility condition if , namely . Then, we define as in 5.11b when (the case is analogous). For , we need no further compatibility of the data and solve the linear system

(Here ; the case is analogous.)

Corollary 5.1.

Under the conditions of Theorem 5.1 and Lemmas 4.2 and 5.2, we have for that

Corollary 5.2.

Under the conditions of Theorem 5.1 and Lemmas 4.2 and 5.2, there exists a constant such that we have for

Here, denotes an approximation of as in Lemma 5.2, i.e., on for some .

Equations (6.1), (6.2), (6.3)
Theorem 6.1.

Suppose that is connected and that belongs to only one connectivity component of and let , i.e., for all . Then, there exists an -independent constant (that depends on and the shape of the elements and patches) such that

Equation (6.5)
Equation (6.6)
Corollary 6.1.

Suppose that the discrete flux satisfies , and . Then,

Example 6.1.

Assume the conditions of Lemma 5.2 on the data . We define an operator by Equation 4.15 and for . In case we consider systems

(cf. Equation 4.16a and notation from Example 4.1), resp., as an analog to Equation 4.16b,
Example 7.1.

Let on the L-shaped domain , on the Dirichlet boundary , and on the Neumann boundary ,

using polar coordinates . The exact solution of Equation 4.1Equation 4.3 has a typical corner singularity at the origin. In this example, the right-hand sides are smooth, but the solution is not. The coarsest triangulation consists of three squares halved by diagonals parallel to the vector (cf. Figure 1).

Example 7.2.

Let for the function

on the unit square and set on the entire boundary (). The solution to Equation 4.1Equation 4.3 is -regular but (although theoretically smooth) has huge gradients on the circle with radius around . The coarsest triangulation consists of four squares halved by diagonals parallel to the vector (cf. Figure 3).

Equation (7.1)
Equation (7.2)
Equation (7.3)

References

Reference [ACF]
J. Alberty, C. Carstensen, S.A. Funken: Remarks around lines of Matlab: short finite element implementation. Numer. Algorithms 20 (1999) 117-137. CMP 2001:01
Reference [A]
A. Alonso: Error estimators for a mixed method. Numer. Math. 74 (1996), 385–395. MR 97g:65212
Reference [BaR]
I. Babuška, W.C. Rheinboldt: Error estimates for adaptive finite element computations. SIAM J. Numer. Anal. 15 (1978) 736–754. MR 58:3400
Reference [Baetal]
I. Babuška, T. Strouboulis, C.S. Upadhyay, S.K. Gangaraj, K. Copps: Validation of a posteriori error estimators by numerical approach. Int. J. Numer. Meth. Engrg. 37 (1994) 1073–1123. MR 95e:65096
Reference [BC]
S. Bartels, C. Carstensen: Each averaging technique yields reliable a posteriori error control in FEM on unstructured grids. Part II: Higher order FEM, Math. Comp., posted on February 4, 2002, PII S 0025-5718(02)01412-6 (to appear in print).
[BeR]
R. Becker, R. Rannacher: A feed-back approach to error control in finite element methods: basic analysis and examples. East-West J. Numer. Math., 4, No. 4 (1996) 237–264. MR 98m:65185
Reference [B]
D. Braess: Finite Elements. Cambridge University Press (1997). MR 98f:65002
Reference [BV]
D. Braess, R. Verfürth: A posteriori error estimators for the Raviart-Thomas element. SIAM J. Numer. Anal. 33 (1996) 2431–2444. MR 97m:65201
Reference [BS]
S.C. Brenner, L.R. Scott: The mathematical theory of finite element methods. Texts Appl. Math. 15, Springer, New-York (1994). MR 95f:65001
Reference [BF]
F. Brezzi, M. Fortin: Mixed and hybrid finite element methods. Springer-Verlag (1991). MR 92d:65187
Reference [C1]
C. Carstensen: A posteriori error estimate for the mixed finite element method. Math. Comp. 66 (1997) 465–476. MR 98a:65162
Reference [C2]
C. Carstensen: Quasi interpolation and a posteriori error analysis in finite element method. M2AN Math. Model Numer. Anal. 33 (1999) 1187-1202. MR 2001a:65135
Reference [CF1]
C. Carstensen, S.A. Funken: Constants in Clément-interpolation error and residual based a posteriori error estimates in Finite Element Methods. East-West J. Numer. Math. 8 (3) (2000) 153-175. CMP 2001:07
Reference [CF2]
C. Carstensen, S.A. Funken: A posteriori error control in low-order finite element discretisations of incompressible stationary flow problems, Math. Comp., 70 (2001), 1353–1381.
Reference [CF3]
C. Carstensen, S.A. Funken: Averaging technique for FE—a posteriori error control in elasticity. Part I: conforming FEM, Comp. Meth. Appl. Mech. Engrg. 190 (2001) 2483-2498. MR 2002a:74114
Reference [CF4]
C. Carstensen, S.A. Funken: Averaging technique for FE—a posteriori error control in elasticity. Part II: -independent estimates, Comput. Methods Appl. Mech. Engrg., 190 (2001), 4663–4675.
Reference [CF5]
C. Carstensen, S.A. Funken: Averaging technique for FE—a posteriori error control in elasticity. Part III: Locking-free nonconforming FEM, Comput. Methods Appl. Mech. Engrg. (to appear).
Reference [CV]
C. Carstensen, R. Verfürth: Edge residuals dominate a posteriori error estimates for low order finite element methods. SIAM J. Numer. Anal. 36 No. 5 (1999) 1571-1587. MR 2000g:65115
Reference [Cl]
P. Clément: Approximation by finite element functions using local regularization. RAIRO Sér. Rouge Anal. Numér. R-2 (1975) 77–84. MR 53:4569
Reference [Ci]
P.G. Ciarlet: The finite element method for elliptic problems. North-Holland, Amsterdam (1978). MR 58:25001
Reference [DDPV]
E. Dari, R. Duran, C. Padra, and V. Vampa: A posteriori error estimators for nonconforming finite element methods. RAIRO Model. Math. Anal. Numer. 30 (1996) 385–400. MR 97f:65066
Reference [DMR]
R. Duran, M.A. Muschietti, R. Rodriguez: On the asymptotic exactness of error estimators for linear triangular elements. Numer. Math. 59 (1991) 107–127. MR 92b:65086
Reference [EEHJ]
K. Eriksson, D. Estep, P. Hansbo, C. Johnson: Introduction to adaptive methods for differential equations. Acta Numer. (1995) 105–158. MR 96k:65057
Reference [GR]
V. Girault, P.A. Raviart: Finite Element Methods for Navier-Stokes Equations. Springer, Berlin (1986). MR 88b:65129
Reference [HSWW]
W. Hoffmann, A.H. Schatz, L.B. Wahlbin, G. Wittum: Asymptotically exact a posteriori error estimators for the pointwise gradient error on each element in irregular meshes. Part 1: A smooth problem and globally quasi–uniform meshes. Math. Comp. 70 (2001) 897–909.
Reference [HW]
R.H.W. Hoppe, B. Wohlmuth: Element-oriented and edge-oriented local error estimators for nonconforming finite element methods. RAIRO Model. Math. Anal. Numer. 30 (1996) 237–263. MR 97e:65124
Reference [KS]
G. Kanschat, F.-T. Suttmeier: A posteriori error estimates for nonconforming finite element schemes. Calcolo 36, No.3 (1999) 129-141. MR 2000k:65208
Reference [R1]
R. Rodriguez: Some remarks on Zienkiewicz–Zhu estimator. Numer. Methods Partial Differential Equations 10 (1994) 625–635. MR 95de:65103
Reference [R2]
R. Rodriguez: A posteriori error analysis in the finite element method. Finite element methods. 50 years of the Courant element. Conference held at the University of Jyvaeskylae, Finland, 1993. Inc. Lect. Notes Pure Appl. Math. 164, 389-397 (1994). MR 95g:65158
Reference [V]
R. Verfürth: A review of a posteriori error estimation and adaptive mesh-refinement techniques. Wiley-Teubner (1996).
Reference [ZZ]
O.C. Zienkiewicz, J.Z. Zhu: A simple error estimator and adaptive procedure for practical engineering analysis. Int. J. Numer. Methods Engrg. 24 (1987) 337–357. MR 87m:73055

Article Information

MSC 2000
Primary: 65N30 (Finite elements, Rayleigh-Ritz and Galerkin methods, finite methods), 65R20 (Integral equations), 74B20 (Nonlinear elasticity), 74G99 (None of the above, but in this section), 74H99 (None of the above, but in this section)
Keywords
  • A posteriori error estimates
  • residual based error estimate
  • adaptive algorithm
  • reliability
  • finite element method
  • mixed finite element method
  • nonconforming finite element method
Author Information
Carsten Carstensen
Institute for Applied Mathematics and Numerical Analysis, Vienna University of Technology, Wiedner Hauptstrasse 8-10, A-1040 Vienna, Austria
Carsten.Carstensen@tuwien.ac.at
Sören Bartels
Mathematisches Seminar, Christian-Albrechts-Universität zu Kiel Ludewig-Meyn-Str. 4, D-24098 Kiel, FRG.
sba@numerik.uni-kiel.de
Journal Information
Mathematics of Computation, Volume 71, Issue 239, ISSN 1088-6842, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on and published on .
Copyright Information
Copyright 2002 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/S0025-5718-02-01402-3
  • MathSciNet Review: 1898741
  • Show rawAMSref \bib{1898741}{article}{ author={Carstensen, Carsten}, author={Bartels, S\"{o}ren}, title={Each averaging technique yields reliable a posteriori error control in FEM on unstructured grids. Part I: Low order conforming, nonconforming, and mixed FEM}, journal={Math. Comp.}, volume={71}, number={239}, date={2002-07}, pages={945-969}, issn={0025-5718}, review={1898741}, doi={10.1090/S0025-5718-02-01402-3}, }

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.