A superconvergent LDG-hybridizable Galerkin method for second-order elliptic problems

By Bernardo Cockburn, Bo Dong, and Johnny Guzmán

Abstract

We identify and study an LDG-hybridizable Galerkin method, which is not an LDG method, for second-order elliptic problems in several space dimensions with remarkable convergence properties. Unlike all other known discontinuous Galerkin methods using polynomials of degree for both the potential as well as the flux, the order of convergence in of both unknowns is . Moreover, both the approximate potential as well as its numerical trace superconverge in -like norms, to suitably chosen projections of the potential, with order . This allows the application of element-by-element postprocessing of the approximate solution which provides an approximation of the potential converging with order in . The method can be thought to be in between the hybridized version of the Raviart-Thomas and that of the Brezzi-Douglas-Marini mixed methods.

1. Introduction

In this paper, we consider the LDG-hybridizable (LDG-H) Galerkin methods recently introduced in Reference 11 and show how to define their numerical traces in order to achieve the optimal order of convergence for the approximation to the flux, and to obtain superconvergence properties similar to those of the hybridized mixed methods of Raviart-Thomas (RT) Reference 16 and the Brezzi-Douglas-Marini (BDM) Reference 4 methods; see also Reference 10.

For the sake of simplicity of the exposition, we carry this out in the setting of the model second-order elliptic problem

where is a polyhedral domain (), , and is a symmetric matrix function that is uniformly positive definite on with components in . As usual, we assume that the -Lebesgue measure of is not zero, that and that .

To describe our results, we need to introduce what we will call hybridized Galerkin methods; they are one of the methods studied in Reference 11. To do that, let us introduce some notation. We denote by a triangulation of the domain of shape-regular simplexes and set . We associate to this triangulation the set of interior faces and the set of boundary faces . We say that if there are two simplexes and in such that , and we say that if there is a simplex in such that . We set .

The hybridized Galerkin methods Reference 11 are dual-mixed hybrid methods (see the definition in Reference 8 and an early example in Reference 15) which seek an approximation to the exact solution , , in a finite-dimensional space of the form

and determines it by requiring that

for all . Here, we have used the notation

for any functions in and in the space . The outward normal unit vector to is denoted by .

To complete the description of the hybridized Galerkin methods, the definition of numerical traces on the faces of the triangulation has to be provided. The choice which is relevant here is

where denotes an -projection defined as follows. Given any function and an arbitrary face , the restriction of to is defined as the element of that satisfies

Note that by suitably choosing the local spaces , , and , and the values of the local stabilization parameters , we can obtain the hybridized RT, the hybridized BDM and the LDG-H methods; see Tables 1 and 2. In Table 1 and in the remainder of this paper, we denote the space of polynomials of degree at most defined on by , and set . Since all these methods can be implemented in the same way and can be used in different elements while being automatically coupled, what is relevant, as argued in Reference 11, is to find out which method should be used in what element in order to optimize the computational effort. It is thus important from this perspective to develop DG methods as accurate and efficient as mixed methods so that they can be used in situations in which mixed method cannot. The LDG-H methods we uncover in this paper are the first example of those methods.

It is well known that the RT and BDM methods provide an approximation to the flux which converges in with order , that and superconverge in -like norms to suitably chosen projections of the potential with order , and that, as a consequence, it is possible to postprocess the approximate solution to obtain another approximation converging in with order ; see Reference 1 and Reference 4, and also Reference 10. In this paper, we use an extension of the postprocessing proposed in Reference 17Reference 18 and Reference 14. Given the similarities between these two mixed methods and the LDG-H method, it is natural to ask if it is possible to choose the local penalization parameters as to obtain similar convergence and superconvergence results. The main contribution of this paper is to show that this is actually possible.

Indeed, we show that this happens if we take, on each simplex ,

where is an arbitrary but fixed face of and is a strictly positive real number. Since the local penalization parameter is nonzero only on a single face of each simplex, we call this LDG-H method the single-face hybridizable DG method; for simplicity, we are going to refer to the method under consideration by the method. It is interesting to note two of the minimal dissipation DG methods considered in Reference 6, in the framework of a study of superconvergence properties of DG methods for one-dimensional steady-state convection-diffusion problems, happen to be an SF-H method. The first is called the md-DG method (see Table 1 in Reference 6) and is obtained, in our notation, by taking on each interior node ,

where is the size of the interval to the left of the node . The second is called the md-LDG method, and is obtained by taking the above choice of parameters formally letting go to infinity. The authors are not aware of any other instance of SF-H methods. In particular, let us emphasize that SF-H methods are not LDG methods whenever the stabilization parameters are finite; see the discussion about LDG-H methods in Reference 11.

In Table 3, we compare the orders of convergence for the flux of this method and the above-mentioned mixed methods. We have also included the order of convergence for the general LDG-H methods; it can be deduced from their characterization Reference 11 and the study of DG methods carried out in Reference 5. Finally, in Table 4, we display the orders of convergence of the postprocessed approximation to the potential.

We also uncover new relations between these three methods. One of the main features of the hybridizable Galerkin methods proposed in Reference 11 is that the only degrees of freedom that turn out to be globally coupled are those of the so-called Lagrange multiplier . This implies, in particular, that the LDG-H methods can be more efficiently implemented than the LDG methods introduced in Reference 12. In fact, see the discussion in Reference 11, they can be implemented as efficiently as the hybridized RT; see Reference 7 for the case in two dimensions, and BDM mixed methods, see Reference 10 for the case in multi-dimensions for both of these methods. Here we show that the stiffness matrix of the Lagrange multiplier for the RT, BDM and methods is actually identical and that, when for all , these methods provide the same approximation .

Next, let us briefly comment on the approach taken to carry out the a priori error analysis of the methods. We did not take the approach used in Reference 5 to analyze DG and LDG methods, or that used in the unified analysis of DG methods Reference 2. Instead, we exploited the unifying framework introduced in Reference 11 to render the analysis of the methods as close as possible to those of the hybridized RT and BDM methods. Since a key ingredient in those analyzes is the existence of a projection satisfying the so-called commutativity property

for all , the crucial step in the analysis was to find a similar projection. Unlike the above-mentioned mixed methods, the space of fluxes of the methods is not included in and, as a consequence, the above commutativity property can only be satisfied in a weak sense. We found a new projection satisfying the following weak version of the commutativity property:

for all such that . Just as the local spaces of the methods are, roughly speaking, “in between” the local spaces of the RT and BDM methods, this projection can also be considered to be “in between” the corresponding projections of those mixed methods. The construction of this projection, which is intimately linked to the definition of the numerical traces of the method and and to the choice of the local spaces, is certainly the most interesting aspect of the analysis of the methods. The first component of the projection, , was used in the error analysis of the minimal dissipation LDG method in Reference 9.

The paper is organized as follows. In Section 2, we state and discuss our main results and then prove them in Section 3. In Section 4, we display numerical experiments validating the theoretical results. Finally, in Section 5, we end with some concluding remarks.

2. The main results

2.1. The projection

In this subsection, we define the projection

and gather its main properties.

Given a function and an arbitrary simplex , the restriction of to is defined as the element of that satisfies

Similarly, given a function and an arbitrary simplex , the restriction of to is defined as the element of that satisfies

We gather the main properties of this projection in the following result. To state it, we need to recall the definition of some classical projections. Given a function and an arbitrary simplex , the restriction of to is defined as the element of that satisfies

Given a function and an arbitrary simplex , the restriction of to is defined as the element of that satisfies

To simplify the notation, we are going to write instead of . Note that is nothing but the projection for the RT method. We are now ready to state our result.

Proposition 2.1.

The projection given by Equation 2.1 and Equation 2.2 is well defined. Moreover, on each simplex , it satisfies the orthogonality properties

for all , and the weak commutativity property

for all . Finally, we have the following approximation estimates

where , is the diameter of , and depends only on and the shape-regularity parameters of the simplex , for any .

We are going to show that the three orthogonality properties imply all the others; they are thus the crucial properties for the analysis. Note also that, by simply adding the identity (iv) over all , we obtain the weak commutativity property discussed in the introduction.

2.2. Characterization of the approximate solution

Next we give a characterization of the approximate solution provided by the method. We begin by characterizing the difference between the numerical traces and the traces of the approximate solutions on each simplex.

Proposition 2.2.

For each simplex , we have that,

We see that the jump is independent of the value of whereas the jump is inversely proportional to . Moreover, by the estimate () of Proposition 2.1, we have that,

for any , and we see that the size of jump under consideration depends solely on the smoothness of . For example, if is a polynomial of degree , then on . This implies that on for every and, as a consequence, that . Now, if , for some , then we have that

by well-known approximation properties of the projection .

Next, we give a characterization of the approximate solution which follows from a similar result for more general methods obtained in Reference 11. To state it, we need to introduce the local solvers associated with the method. The first local solver is defined on the simplex as the mapping where

for all , where

The other local solver is defined on the simplex as the mapping where

for all , where

We can now state our characterization result.

Theorem 2.3.

The approximate solution given by the method is well defined. Moreover, we have that

where can be characterized as the function in satisfying

where

for all and .

This result allows us to shed light on the effect of local stabilization parameters on the approximate solution. It will also allow us to compare the RT, the BDM and the methods, see Reference 10 for a comparison of the hybridized version of the RT and the BDM methods. These results are gathered in the following theorem. To state it, we use the projection which is defined by Equation 2.4 for and which we take to be identically zero when . We keep this convention in the remainder of the paper.

Theorem 2.4.

We have that

(i)

The function , is independent of the values of the local stabilization parameters . Moreover, changes in the local stabilization parameters only affect the function .

(ii)

If for all simplexes , then is the same for the RT, the BDM (if ) and the methods. Moreover, for all .

(iii)

The bilinear form is always the same for the RT, the BDM (if ) and the methods.

2.3. A priori error estimates

In this subsection, we obtain a priori error estimates for the error of the approximation given by the and the numerical trace defined by Equation 1.4a. To state them, we need to introduce new notation.

For any real-valued function in , we set

For a vector-valued function we set

We can now state our results.

We begin by measuring the error in the approximation of the flux in the norm .

Theorem 2.5.

Suppose that the exact flux belongs to for some . Then

for some constant independent of and the exact solution .

Note that the upper bound of the error is independent of the local stabilization parameters , in complete agreement with the characterization of the approximate solution given by Theorem 2.3. It is interesting to realize that the first estimate also holds for the RT and BDM methods when the projection is suitably chosen; see Reference 13 and Reference 4. Such an estimate is obtained by using the commutativity property and the fact that the image of their projections is in . Since our projection only satisfies a weak version of the commutativity property, a much more delicate analysis has to be carried out to obtain it.

In Reference 5, it was shown that for general LDG methods with penalization parameters of order , the order of convergence of the approximations for flux using polynomials of degree is only ; this order is sharp because it is actually attained for some LDG methods. It was also shown that, for DG methods with both penalization parameters of order one, the order of convergence of the approximations for the flux using polynomials of degree is . Here, we obtain an order of convergence of . No other DG method has this property.

Next, we present several estimates for the error in the approximation of the potential . The first is a superconvergence result. To state it, we need to introduce the adjoint equations

We also need to assume that the following elliptic regularity result holds

for . Note that, since we are working with domains that can be triangulated by using straight-faced simplexes, the above result only holds if such a domain is convex and . However, we want to write this assumption in such a generality since the method will be extended to domains with smooth curved boundaries in a forthcoming paper.

Theorem 2.6.

Suppose that the exact flux belongs to for . Set . Then,

where

Moreover, for and general ,

where .

It is interesting to note that the above superconvergence result holds for any choice of local stabilization parameters such that is uniformly bounded, that is, such that is uniformly bounded with respect to . This shows that cannot be too small for superconvergence to take place.

A straightforward consequence of this theorem is the following result.

Corollary 2.7.

Suppose that the exact flux belongs to for . Then

where

Note that the above result shows that if is uniformly bounded for quasi-uniform triangulations, the convergence of is still optimal, provided is smoother than required, that is, provided instead of just . Of course, in this case, the superconvergence of to is lost.

The next result is a superconvergence result for the Lagrange multiplier . To state it, we use the following norm:

Theorem 2.8.

Suppose that the exact solution of Equation 1.1 belongs to for some . Then,

if , or if and .

There are no results of this type for any other DG method. However, the RT and the BDM methods have both similar results. Here we exploited the similarity of the methods with the RT and BDM methods to obtain these superconvergence results.

2.4. Postprocessing

We end this section by showing how to exploit the superconvergence results to postprocess , and to get a better approximation to defined as follows.

On the simplex , we define the new approximation of , , as the function of given by

whereand is the polynomial in satisfying

Here and is the collection of functions in with mean zero. The postprocessing technique just introduced is a slight modification of a postprocessing proposed in Reference 17Reference 18 and Reference 14; it consists of using the numerical trace instead of .

It is easy to see that this postprocessing is associated to a locally conservative method. Indeed, the scheme satisfied by on each simplex is

As a consequence, if we take to be the union of an arbitrary set of simplexes , we get that

which is nothing but the property of local conservativity.

Note that is well defined. Indeed, if we take in equation Equation 2.9c, the right-hand side is also equal to zero thanks to equation Equation 1.3b. The fact that it provides a better approximation to the potential than is contained in the following result.

Theorem 2.9.

Suppose that the exact solution belongs to for . Then, if ,

and if and ,

Note that when for all , by Theorem 2.4 we have that the function is the same for the RT, BDM () and methods. As a consequence, the postprocessed approximation is also the same for all these methods.

Note also that in Reference 3, a general postprocessing which is solely based on approximation results was obtained. When applied to the method for , it gives rise to an approximation of which converges with the same orders as ours. However, unlike such postprocessing, our postprocessed solution is associated to a locally conservative scheme; it is also easier to compute.

Let us end this section by noting that all the error estimates for hold if in the equation Equation 1.3b, we replace by any function such that for all and such that

Moreover, by statement (ii) of Theorem 2.4, the function provided by the RT, the BDM and the method is the same; in particular, we have that . The postprocessed approximation is also the same for those three methods.

3. Proofs

In this section, we present detailed proofs of all our results.

3.1. Proof of Proposition 2.1: The properties of

3.1.1. Two key auxiliary results about polynomials

To prove Proposition 2.1, we begin by stating and proving two lemmas whose use is crucial in our analysis.

Lemma 3.1.

Given the face of the simplex and a function , there is a unique function such that

Moreover,

where is the diameter of the simplex and depends solely on and the shape-regularity parameters of the simplex .

Lemma 3.2.

Given the face of the simplex and the function such that for all faces of different from , , there is a unique function such that

Moreover,

where is the diameter of the simplex and depends solely on and the shape-regularity parameters of the simplex .

We are only going to give a detailed proof of Lemma 3.2 since the proof of Lemma 3.1 is similar and simpler.

Proof of Lemma 3.2.

Let us begin by proving that the function satisfying (i) and (ii) exists and is unique. Since the linear system determined by equations (i) and (ii) is square, indeed,

and , we only need to show that if satisfies

then on .

Let be the affine mapping that transforms the element to the reference simplex . Moreover, let us denote by , , the faces of where . Assume that the mapping is such that is the face of lying on the plane , and set . Then the above equations become

since spaces of polynomials of a given total degree are invariant under affine transformations. Now, let be the basis of dual to , that is, Then we can write , where , and obtain that

The last equation implies that, for any , and hence that for some polynomial in . Taking , we get

and, since on , we conclude that . This implies that on . This proves the existence and uniqueness of satisfying the conditions (i) and (ii).

The estimate (iii) follows now from a simple scaling argument. This completes the proof.

3.1.2. Proof of the orthogonality properties

It is not difficult to see that the fact that is well defined is a direct corollary of Lemmas 3.1 and 3.2.

Now, let us prove the orthogonality properties. The property (i) follows from the property Equation 2.2a defining and the orthogonality property (ii) follows from the property Equation 2.1a defining . The orthogonality property (iii) follows from the properties Equation 2.2b and Equation 2.1b defining and , and from the definition of the projection , Equation 1.5. In fact, it follows from the fact that on each face of any simplex , we have that either or .

3.1.3. Proof of the weak commutativity property

The weak commutativity property (iv) is a direct consequence of the three orthogonality properties we just proved. Indeed, we have that

by the definition of the projection , Equation 1.5. This completes the proof of (iv).

3.1.4. Proof of the estimates (v) and (vi)

Note that, by the definition of the projections , Equation 2.1, and , Equation 2.3, we have that

for all , if , and for all and all faces of . By a well-known scaling argument, we immediately obtain that

It remains to estimate the above right-hand side. To do that, we note that, for any in , we have that

by the definition of the projections , Equation 2.1. Taking , where is given by Lemma 3.1 with , we get that, for any in ,

and, after a direct application of the Bramble-Hilbert lemma, we get

where . This completes the proof of the estimates (v) and (vi).

3.1.5. Proof of the estimate (vii)

Note that, by the definition of the projections , Equation 2.2, and , Equation 2.4, we have that

for all , if , and for all . This implies that Lemma 3.1 holds with and . As a consequence,

It remains to estimate the above right-hand side.

To do that, we note that, for any in , we have that

by the definition of the projection , Equation 2.4 and that of the projection , Equation 1.5. A well-known scaling argument states that given any function such that its restriction to each face of belongs to , there is a function in such that

where is the diameter of the simplex and depends solely on the shape-regularity constants of the simplex . Taking with in equation Equation 3.1, we obtain that

for any . Thus, after a direct application of the Bramble-Hilbert lemma, we get

where . This completes the proof of estimate (vii).

This completes the proof of Proposition 2.1.

3.2. Characterization of the approximate solution

To prove the results of the characterization of the approximate solution of the SF-H method, we begin by proving two auxiliary results concerning key properties of the local solvers.

3.2.1. Two auxiliary results about the local solvers

To state the first auxiliary result, we need to introduce the following decomposition of our local spaces:

where

and

Lemma 3.3.

Let be any simplex of the triangulation . Then the local mapping given by equations Equation 2.5 can be obtained as follows:

(i)

Set

(ii)

Compute by solving

(iii)

Compute by solving

Similarly, the local mapping given by equations Equation 2.6 can be obtained as follows:

()

Compute by solving

()

Compute by solving

()

Compute by solving

Proof.

Let us begin by proving the properties of the first local mapping. Thus, integrating by parts in the equation Equation 2.5b, we obtain

for all , by the definition of the numerical trace , Equation 1.4b and Equation 1.6. Taking , we see that

Using the fact that which follows by a simple application of Lemma 3.1, we have that (i) holds. As a consequence, we see that

and hence that . The property (ii) can now be obtained by restricting the test functions to the space in the equation Equation 2.5a. Now that we know , we obtain the formulation (iii) for by restricting the test functions to the space in the equation Equation 2.5a. It remains to show that is uniquely defined by those equations. But this follows from the fact that the system of equations is square and that , which in turn follows from the fact that and . This completes the proof of the properties of the first local lifting.

The proof the properties () and () of the second local mapping is similar to the proof of the properties (i) and (iii) of the first local mapping, respectively. Let us prove property (). If we take in the equation Equation 2.6a, we see that . Since the equation in () is obtained from Equation 2.6b by restricting the tests functions to , we only have to prove that given by () is well defined. But this follows from the fact that the system is a square system and . This completes the proof.

The second auxiliary result concerns the jumps of the local solvers.

Lemma 3.4.

For each simplex , we have that, on ,

Proof.

Let us begin by proving the second identity since its proof is more involved. Taking in the identity () of Lemma 3.3, where is given by Lemma 3.1 with , we obtain that

by the properties of the projection , Equation 2.1. As a consequence, we immediately obtain that, on ,

A similar argument gives that, on ,

This completes the proof of Lemma 3.4.

3.2.2. Proof of Theorem 2.3: Characterization of the approximate solution

The following result is a particular case of a general result proven in Reference 11.

Theorem 3.5.

The approximate solution given by the method is well defined. Moreover, we have that

where can be characterized as the function in satisfying

where

for all and .

Theorem 2.3 follows from this result if we show that on ,

for all . Since this is a straightforward consequence of Lemma 3.4, this completes the proof of Theorem 2.3.

3.2.3. Proof of Proposition 2.2: Characterization of the jumps

By the definition of the numerical traces Equation 1.4 and Equation 1.6, we have that

by the definition of the projection , Equation 2.1b, and that of the projection , Equation 1.5. So, we only have to prove that

But, by Theorem 2.3, we have that

on the face . This completes the proof of Proposition 2.2.

3.2.4. Proof of Theorem 2.4

The statement (i) of Theorem 2.4 follows directly from Theorem 2.3 and from Lemma 3.3.

To prove the remaining statements, we are going to use the fact that the RT, BDM and methods have exactly the same structure and satisfy the characterization Theorem 2.3; see Reference 11. The only difference between these methods is the choice of local spaces (see Table 1) and the choice of the local stabilization parameters ; see Table 2. Thus, to prove statement (ii) we only have to show that the functions and are the same for all these methods whenever for all . Similarly, to prove statement (iii), we only have to show that is the same for all these methods.

To do that, we begin by noting that we have, by Lemma 3.3, that the function is determined by

and that the function is determined by the equations

where , by () of Lemma 3.3, if . Since the four equations above also hold (the third whenever ) for the BDM method, we conclude that the statements (ii) and (iii) hold if we exclude the RT method.

To show that these statements also hold if we include it, we note that the above equations hold for the RT method if we modify the definition of the spaces and by

and if we replace the third equation by

Thus, the result follows from the fact that

and from the fact that, if and , then belongs to the space . This completes the proof of Theorem 2.4.

3.3. Proof of the error estimates

The proof of the error estimates is based on the error equations and the properties of the projection gathered in Proposition 2.1. The error equations are

for all .

A direct consequence of the weak commutativity identity (iv) of Proposition 2.1 that we find convenient to use in our analysis is contained in the following result.

Corollary 3.6.

For all , we have

where and depends only on and the shape-regularity parameters of the simplexes .

3.3.1. Proof of Theorem 2.5: The error in the flux

Theorem 2.5 follows immediately from the following auxiliary result.

Lemma 3.7.

We have

Indeed, this implies that

and hence, that

for some , by the estimate (vi) of Proposition 2.1. This proves Theorem 2.5.

Let us prove Lemma 3.7.

Proof.

By the error equation Equation 3.2a with , we have

By the identity () of Corollary 3.6 with , we get that

and by the error equation Equation 3.2b with ,

It we denote by the right-hand side of the above equations, it is not difficult to see that, after a few simple algebraic manipulations, we have that where

We are going to show that .

We begin by noting that,

by Proposition 2.2. By the definition of the projection , Equation 1.5,

by the definition of the projection , Equation 2.1b. Thus, .

Next, let us show that . By the fact that the numerical trace and the normal component of the numerical trace are single-valued on the interior faces, by definition of , Equation 1.4a, and the equation Equation 1.3c satisfied by , we have that

by the definition of the numerical traces at the boundary. By using the definition of the projection , Equation 1.5, we get

Finally, let us show that . By the definition of the numerical trace , Equation 1.4b,

by the definition of the projection , Equation 2.2b. This completes the proof.

3.3.2. Proof of Theorem 2.6: Superconvergence of

Since

we need to estimate the number . It is expressed in a suitable way in the following auxiliary result. Let us recall that is defined by Equation 2.4 for , and is for .

Lemma 3.8.

We have

Assume that . Then, applying the Cauchy-Schwarz inequality and using the estimate of in Theorem 2.5, and the approximation properties of the projections and , () in Proposition 2.1 and () in Corollary 3.6, we readily obtain

where . Since and using the elliptic regularity assumption Equation 2.8, we get

This completes the proof of Theorem 2.6 for .

In the case , we have that

and, after using the elliptic regularity assumption Equation 2.8, we get

Finally, let us consider the case and . By the identity (v) of Proposition 2.1 we have that , and by the identity (vi) of Proposition 2.1 we have that . This implies that

by the adjoint equation Equation 2.7c and the boundary condition Equation 1.1d. As a consequence, we get

and since

by the elliptic regularity assumption Equation 2.8, we get

It remains to prove Lemma 3.8.

Proof.

By the adjoint equation Equation 2.7b, we have that

by the identity () of Corollary 3.6 with . By the error equation Equation 3.2a with , we get

and, by the adjoint equation Equation 2.7a,

By the orthogonality property (ii) of Proposition 2.1, we get that

If we denote by the last three terms of the above right-hand side, we see that, after some simple algebraic manipulations, we can write , where

By the definition of the numerical trace , Equation 1.4a and Equation 1.6, we have that

by Proposition 2.2.

It remains to show that . By the definition of the projection , Equation 1.5,

By the definition of the numerical trace , Equation 1.4a,

Next, we show that . Integrating by parts, we obtain

by the identity () of Corollary 3.6 with . By the error equation Equation 3.2b with ,

by the definition of the projection , Equation 1.5, the definition of the projection , Equation 2.2b, and the definition of the numerical trace , Equation 1.4b. Hence

by the adjoint equation Equation 2.7c and the equation Equation 1.3c for .

This completes the proof.

3.3.3. Proof of Theorem 2.8: Superconvergence of

To prove this theorem, let us begin by estimating for each face of each simplex . For the face , we have that, by definition of the projection , Equation 2.2,

by Proposition 2.2 and the identity (v) of Proposition 2.1. By using a classical inverse inequality, we can conclude that

Now we consider the error in the faces of which are different from the face . By the error equation Equation 3.2a, we have that, for all ,

Taking given by Lemma 3.2 with , we obtain that

and using the estimate for the error in ,

As a consequence, we get

where . The result now follows from Theorems 2.6, 2.5 and 2.4 (i). This completes the proof of Theorem 2.8.

3.3.4. Proof of Theorem 2.9: The error estimate for

By the definition of , Equation 2.9a, we have that

where is defined in Equation 2.9b and . We estimate each of the two terms of the right-hand side separately.

We begin by estimating the second term. Since, by Poincaré’s inequality, we have

it is enough to estimate the error in the gradient. To do that, we note that, by the definition of , Equation 2.9c, we have

Then

Let us estimate the first term of the right-hand side. For any arbitrary , we have

where

By using Cauchy-Schwarz inequality, we get that

By using the definition of the Raviart-Thomas projection , Equation 2.3, and by using its commutativity property, we get that, for any ,

by Poincaré’s inequality. Finally, by the definition of the numerical trace , Equation 1.4b,

by Proposition 2.2 and identity (v) of Proposition 2.1. Applying a simple inverse inequality, we get

As a consequence,

This implies that

and so,

by Theorem 2.5 and the well-known approximation properties of .

Let us now estimate the error . We begin by considering the case . In this case, since , we get

by Theorem 2.6. Note that by Theorem 2.4, is independent of the value of the local stabilization parameters . This implies that the same is true for and so, we get that

It remains to consider the case and . We have that

Since, for any function , we have that

we readily obtain that

and so,

Since, by Theorem 2.4, is independent of the value of the local stabilization parameter , so is and so

This completes the proof of Theorem 2.9.

4. Numerical experiments

In this section, we carry out numerical experiments to validate the theoretical convergence properties of the method.

To do that, we use uniform meshes obtained by discretizing with squares of side which are then divided into two triangles as indicated in Figure 1; the resulting mesh is denoted by “mesh=”.

The test problem is obtained by taking and choosing and so that the exact solution is on the domain . The history of convergence of the SF-H method with

on the “mesh=”, is displayed in Table 5 for polynomials of degree , and . We observe optimal convergence rates of the quantities and for as predicted by Theorems 2.5 and 2.7. We also see that and superconverges with rate for just as predicted by Theorems 2.8 and 2.9. These results do not guarantee that these quantities are superconvergent if and . Since we do not observe superconvergence, we can conclude that the theoretical results for such a case are actually sharp.

Next we explore the effect of the size of on the quality of the approximation. In Table 6, we see that as diminishes the quality of the approximation to deteriorates. However, the effect of taking or is almost negligible especially when the grids are not coarse. We also see that the order of convergence is for and , but it is only for . This is in perfect agreement with Corollary 2.7.

We end with an example where the exact solution is harmonic, that is, , and display the convergence rates for in Table 7. We take and choose and so that is the solution. We see that the quantities and superconverge with the rate as our theoretical results predict.

5. Concluding remarks

The error analysis carried out here for the method also holds for the hybridized versions of the RT and the BDM methods. We simply have to replace the local space by the local space given by Table 1, use the definition of the local stabilization parameter given in Table 2, and suitably define the projection . Indeed, with such changes, the first four properties of Proposition 2.1, on which the whole analysis is based, hold. For this reason, we can consider this analysis to be a unifying analysis of these three methods.

A study of the optimal way to choose the local stabilization parameter falls beyond the scope of this paper and will be carried out elsewhere. Extensions of these results to more general second-order elliptic equations and other boundary conditions are straightforward. The extension of these results to the case of hanging nodes, variable-degree approximations and curved domains constitute the subject of ongoing work.

Figures

Table 1.

The local spaces

method
RT
LDG-H
BDM
Table 2.

The local stabilization parameters

method
RT
LDG-H
BDM
Table 3.

The orders of convergence in of the -errors

methodcondition
RT
LDG-H [5] and
LDG-H [5] and
BDM
Table 4.

The orders of convergence in of

methodordercondition
RT
,
BDM
BDM
Figure 1.

Example of a mesh with .

Graphic without alt text
Table 5.

History of convergence of the method.

mesh
errorordererrorordererrorordererrororder
1.11e+1-.17e+1-.28e-0-.22e-0-
2.36e-01.54.78e-01.12.92e-11.61.57e-11.96
03.12e-01.50.41e-00.94.35e-11.37.19e-11.57
4.53e-11.29.21e-00.97.14e-11.21.79e-21.27
5.24e-11.13.10e-00.98.69e-21.15.36e-21.14
6.12e-11.05.53e-10.99.32e-21.12.17e-21.10
1.21e-0-.23e-0-.31e-1-.21e-1-
2.43e-12.27.12e-00.94.75e-22.02.40e-22.38
13.78e-22.47.31e-11.94.10e-22.96.53e-32.91
4.17e-22.19.79e-21.99.12e-33.00.68e-42.98
5.42e-32.05.20e-22.00.15e-43.01.85e-52.99
6.10e-32.01.50e-32.00.19e-53.00.11e-62.99
1.68e-1-.89e-1-.72e-2-.81e-2-
2.38e-24.12.91e-23.29.41e-34.12.40e-34.35
23.32e-33.58.12e-22.96.27e-43.93.25e-43.97
4.32e-43.31.15e-32.98.18e-53.96.16e-53.99
5.37e-53.12.19e-42.99.11e-63.98.10e-64.00
6.45e-73.01.23e-53.00.70e-83.99.63e-84.00
Table 6.

Effect of on the convergence of .

mesh
errorordererrorordererrorordererrororder
1.61e+0-.11e+1-.21e+1-.40e+1-
2.21e+01.50.36e-01.54.11e+10.88.42e+1-0.06
03.95e-11.17.12e-01.50.57e-00.97.43e+1-0.02
4.46e-11.04.53e-11.29.28e-01.00.43e+1-0.00
5.23e-11.02.24e-11.13.14e-01.01.43e+10.00
6.11e-11.01.12e-11.05.70e-11.01.43e+10.00
1.15e+0-.21e-0-.35e-0-.67e-0-
2.27e-12.47.43e-12.27.14e-01.34.54e-00.29
13.66e-22.06.78e-22.47.35e-11.98.28e-00.97
4.16e-22.00.17e-22.19.88e-22.00.14e-00.99
5.41e-32.00.42e-32.05.22e-22.00.67e-11.00
6.10e-32.00.10e-32.01.60e-32.00.35e-11.00
1.33e-1-.68e-1-.14e-0-.28e-0-
2.19e-24.10.38e-24.12.14e-13.31.56e-12.33
23.23e-33.07.32e-33.58.18e-22.96.14e-11.98
4.28e-43.00.32e-43.31.23e-32.99.35e-22.00
5.36e-53.00.37e-53.12.28e-43.00.88e-32.00
6.45e-63.00.45e-73.01.35e-53.00.22e-32.00
Table 7.

History of convergence for a harmonic exact solution.

mesh
errorordererrorordererrorordererrororder
1.17e-0-.22e-0-.29e-1-.23e-1-
2.87e-10.94.11e-00.96.79e-21.84.62e-21.87
03.44e-10.99.57e-10.98.21e-21.90.16e-21.93
4.22e-10.99.29e-10.99.55e-31.96.41e-31.97
5.11e-11.00.14e-11.00.10e-31.98.10e-31.99
6.55e-21.00.72e-21.00.35e-42.00.26e-42.00

Mathematical Fragments

Equation (1.1)
Equation (1.3)
Equation (1.4)
Equation (1.5)
Equation (1.6)
Equation (2.1)
Equation (2.2)
Equation (2.3)
Equation (2.4)
Proposition 2.1.

The projection given by Equation 2.1 and Equation 2.2 is well defined. Moreover, on each simplex , it satisfies the orthogonality properties

for all , and the weak commutativity property

for all . Finally, we have the following approximation estimates

where , is the diameter of , and depends only on and the shape-regularity parameters of the simplex , for any .

Proposition 2.2.

For each simplex , we have that,

Equation (2.5)
for all , where
Equation (2.6)
for all , where
Theorem 2.3.

The approximate solution given by the method is well defined. Moreover, we have that

where can be characterized as the function in satisfying

where

for all and .

Theorem 2.4.

We have that

(i)

The function , is independent of the values of the local stabilization parameters . Moreover, changes in the local stabilization parameters only affect the function .

(ii)

If for all simplexes , then is the same for the RT, the BDM (if ) and the methods. Moreover, for all .

(iii)

The bilinear form is always the same for the RT, the BDM (if ) and the methods.

Theorem 2.5.

Suppose that the exact flux belongs to for some . Then

for some constant independent of and the exact solution .

Equation (2.7)
Equation (2.8)
Theorem 2.6.

Suppose that the exact flux belongs to for . Set . Then,

where

Moreover, for and general ,

where .

Corollary 2.7.

Suppose that the exact flux belongs to for . Then

where

Theorem 2.8.

Suppose that the exact solution of Equation 1.1 belongs to for some . Then,

if , or if and .

Equation (2.9)
whereand is the polynomial in satisfying
Theorem 2.9.

Suppose that the exact solution belongs to for . Then, if ,

and if and ,

Lemma 3.1.

Given the face of the simplex and a function , there is a unique function such that

Moreover,

where is the diameter of the simplex and depends solely on and the shape-regularity parameters of the simplex .

Lemma 3.2.

Given the face of the simplex and the function such that for all faces of different from , , there is a unique function such that

Moreover,

where is the diameter of the simplex and depends solely on and the shape-regularity parameters of the simplex .

Equation (3.1)
Lemma 3.3.

Let be any simplex of the triangulation . Then the local mapping given by equations Equation 2.5 can be obtained as follows:

(i)

Set

(ii)

Compute by solving

(iii)

Compute by solving

Similarly, the local mapping given by equations Equation 2.6 can be obtained as follows:

()

Compute by solving

()

Compute by solving

()

Compute by solving

Lemma 3.4.

For each simplex , we have that, on ,

Equation (3.2)
Corollary 3.6.

For all , we have

where and depends only on and the shape-regularity parameters of the simplexes .

Lemma 3.7.

We have

Lemma 3.8.

We have

References

Reference [1]
D. N. Arnold and F. Brezzi, Mixed and nonconforming finite element methods: implementation, postprocessing and error estimates, RAIRO Modél. Math. Anal. Numér. 19 (1985), 7–32. MR 813687 (87g:65126)
Reference [2]
D. N. Arnold, F. Brezzi, B. Cockburn, and L. D. Marini, Unified analysis of discontinuous Galerkin methods for elliptic problems, SIAM J. Numer. Anal. 39 (2002), 1749–1779. MR 1885715 (2002k:65183)
Reference [3]
J. H. Bramble and J. Xu, A local post-processing technique for improving the accuracy in mixed finite-element approximations, SIAM J. Numer. Anal. 26 (1989), no. 6, 1267–1275. MR 1025087 (90m:65193)
Reference [4]
F. Brezzi, J. Douglas, Jr., and L. D. Marini, Two families of mixed finite elements for second order elliptic problems, Numer. Math. 47 (1985), 217–235. MR 799685 (87g:65133)
Reference [5]
P. Castillo, B. Cockburn, I. Perugia, and D. Schötzau, An a priori error analysis of the local discontinuous Galerkin method for elliptic problems, SIAM J. Numer. Anal. 38 (2000), 1676–1706. MR 1813251 (2002k:65175)
Reference [6]
F. Celiker and B. Cockburn, Superconvergence of the numerical traces of discontinuous Galerkin and hybridized mixed methods for convection-diffusion problems in one space dimension, Math. Comp. 76 (2007), 67–96. MR 2261012
Reference [7]
Z. Chen, Equivalence between and multigrid algorithms for nonconforming and mixed methods for second-order elliptic problems, East-West J. Numer. Math. 4 (1996), 1–33. MR 1393063 (98c:65184)
Reference [8]
P. Ciarlet, The finite element method for elliptic problems, North-Holland, Armsterdam, 1978. MR 0520174 (58:25001)
Reference [9]
B. Cockburn and B. Dong, An analysis of the minimal dissipation local discontinuous Galerkin method for convection-diffusion problems, J. Sci. Comput. 32 (2007), 233–262. MR 2320571
Reference [10]
B. Cockburn and J. Gopalakrishnan, A characterization of hybridized mixed methods for second order elliptic problems, SIAM J. Numer. Anal. 42 (2004), 283–301. MR 2051067 (2005e:65183)
Reference [11]
B. Cockburn, J. Gopalakrishnan, and R. Lazarov, Unified hybridization of discontinuous Galerkin, mixed and continuous Galerkin methods for second order elliptic problems, Submitted.
Reference [12]
B. Cockburn and C.-W. Shu, The local discontinuous Galerkin method for time-dependent convection-diffusion systems, SIAM J. Numer. Anal. 35 (1998), 2440–2463. MR 1655854 (99j:65163)
Reference [13]
J. Douglas, Jr. and J. E. Roberts, Global estimates for mixed methods for second order elliptic equations, Math. Comp. 44 (1985), 39–52. MR 771029 (86b:65122)
Reference [14]
L. Gastaldi and R. H. Nochetto, Sharp maximum norm error estimates for general mixed finite element approximations to second order elliptic equations, RAIRO Modél. Math. Anal. Numér. 23 (1989), 103–128. MR 1015921 (91b:65125)
Reference [15]
J. T. Oden and J. K. Lee, Dual-mixed hybrid finite element method for second-order elliptic problems, Mathematical aspects of finite element methods (Proc. Conf., Consiglio Naz. delle Ricerche (C.N.R.), Rome, 1975), Springer, Berlin, 1977, pp. 275–291. Lecture Notes in Math., Vol. 606. MR 0520341 (58:25010)
Reference [16]
P. A. Raviart and J. M. Thomas, A mixed finite element method for second order elliptic problems, Mathematical Aspects of Finite Element Method (I. Galligani and E. Magenes, eds.), Lecture Notes in Math. 606, Springer-Verlag, New York, 1977, pp. 292–315. MR 0483555 (58:3547)
Reference [17]
Rolf Stenberg, A family of mixed finite elements for the elasticity problem, Numer. Math 53 (1988), 513–538. MR 954768 (89h:65192)
Reference [18]
Rolf Stenberg, Postprocessing schemes for some mixed finite elements, RAIRO Modél. Math. Anal. Numér. 25 (1991), 151–167. MR 1086845 (92a:65303)

Article Information

MSC 2000
Primary: 65M60 (Finite elements, Rayleigh-Ritz and Galerkin methods, finite methods), 65N30 (Finite elements, Rayleigh-Ritz and Galerkin methods, finite methods), 35L65 (Conservation laws)
Keywords
  • Discontinuous Galerkin methods
  • hybridization
  • superconvergence
  • second-order elliptic problems
Author Information
Bernardo Cockburn
School of Mathematics, 127 Vincent Hall, University of Minnesota, Minneapolis, Minnesota 55455
cockburn@math.umn.edu
Bo Dong
Division of Applied Mathematics, Brown University, Providence, Rhode Island 02912
bdong@dam.brown.edu
Johnny Guzmán
School of Mathematics, 127 Vincent Hall, University of Minnesota, Minneapolis, Minnesota 55455
guzma033@umn.edu
MathSciNet
Additional Notes

The first author was supported in part by the National Science Foundation (Grant DMS-0411254) and by the University of Minnesota Supercomputing Institute.

The third author was supported by an NSF Mathematical Science Postdoctoral Research Fellowship (DMS-0503050).

Journal Information
Mathematics of Computation, Volume 77, Issue 264, 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 2008 American Mathematical Society; reverts to public domain 28 years from publication
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/S0025-5718-08-02123-6
  • MathSciNet Review: 2429868
  • Show rawAMSref \bib{2429868}{article}{ author={Cockburn, Bernardo}, author={Dong, Bo}, author={Guzm\'an, Johnny}, title={A superconvergent LDG-hybridizable Galerkin method for second-order elliptic problems}, journal={Math. Comp.}, volume={77}, number={264}, date={2008-10}, pages={1887-1916}, issn={0025-5718}, review={2429868}, doi={10.1090/S0025-5718-08-02123-6}, }

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.