Optimal a priori error estimates for the -version of the local discontinuous Galerkin method for convection–diffusion problems

By Paul Castillo, Bernardo Cockburn, Dominik Schötzau, and Christoph Schwab

Abstract

We study the convergence properties of the -version of the local discontinuous Galerkin finite element method for convection-diffusion problems; we consider a model problem in a one-dimensional space domain. We allow arbitrary meshes and polynomial degree distributions and obtain upper bounds for the energy norm of the error which are explicit in the mesh-width , in the polynomial degree , and in the regularity of the exact solution. We identify a special numerical flux for which the estimates are optimal in both and . The theoretical results are confirmed in a series of numerical examples.

1. Introduction

This paper contains the first a priori error estimate of the -version of the so-called local discontinuous Galerkin (LDG) finite element method for convection-diffusion problems. Such an error analysis, which takes into account both the mesh-size of the element, , and the degree of the approximating polynomial in it, , is quite relevant for the LDG method since, being a locally conservative method that does not require any inter-element continuity, it is ideally suited for -adaptivity. In this paper, we consider a model convection-diffusion equation in one space dimension with Dirichlet boundary conditions and obtain, for a special choice of the numerical fluxes defining the LDG method, a priori error estimates that are optimal both in and , even for ; all other error estimates available in the current literature are suboptimal in both and and do not give a rate of convergence for .

The LDG method was introduced by Cockburn and Shu in Reference 11 as an extension to general convection-diffusion problems of the numerical scheme for the compressible Navier-Stokes proposed by Bassi and Rebay in Reference 1. This scheme was in turn an extension of the Runge-Kutta discontinuous Galerkin (RKDG) method developed by Cockburn and Shu Reference 10Reference 9Reference 8Reference 6Reference 12 for nonlinear hyperbolic systems. For a fairly complete set of references on RKDG and LDG methods, see the short monograph by Cockburn Reference 4; see also the review of the development of discontinuous Galerkin methods by Cockburn, Karniadakis, and Shu Reference 7.

To put our result in proper perspective, let us briefly describe the relevant results available in the current literature. There are only a few a priori error estimates for the LDG method and they are all for the -version of the method. The first a priori error estimate for the LDG method was obtained in 1998 by Cockburn and Shu Reference 11 who proved that, when polynomials of degree are used, the LDG method converges in the energy norm at a rate of order . This rate of convergence was obtained for the general form of the so-called numerical fluxes that appear in the definition of the LDG method and is sharp since for the numerical flux proposed by Bassi and Rebay Reference 1 this rate is actually achieved. Later, this analysis was extended by Cockburn and Dawson Reference 5 to the case in which the convective velocity and the diffusion tensor depend on and the domain is bounded; the rate of convergence of order was once again obtained.

Although the rate of convergence of order is sharp for general fluxes, Cockburn and Shu Reference 11 reported numerical experiments in the one-dimensional case indicating that, for a special numerical flux, a rate of convergence of order is achieved for very smooth solutions. This indication was later put on firm mathematical grounds by Castillo Reference 3 who showed, for the model problem of constant-coefficient, linear convection-diffusion in one space dimension, that the LDG method with a particular numerical flux converges with the optimal rate of convergence of order . Castillo’s result can be viewed as an extension to the convection-diffusion setting of the a priori error estimate for the discontinuous Galerkin (DG) method for the purely convective case obtained in 1974 by LeSaint and Raviart Reference 15 who prove that the rate of convergence is of order .

In this paper, we obtain an a priori error estimate for the -version of the LDG method for general numerical fluxes which is explicit in the mesh-width and the polynomial degree . Assuming that the -th derivative of the exact solution in the energy norm plus the -norm of the time derivative is finite, we show that, for general numerical fluxes, the energy norm of the error has a rate of convergence of order in the purely convective case and of in the convection-diffusion case. Moreover, by using the special numerical flux studied by Castillo Reference 3, we obtain the optimal rate of convergence of order for totally arbitrary meshes and polynomials of degree in all elements. This result holds in the purely convective case as well as in the purely parabolic case.

Let us give an idea of how the error estimate is obtained. First, using the technique employed by Cockburn and Shu Reference 11, we find an upper bound for the energy norm of a projection into the finite element space of the error. Then, following Castillo Reference 3, we eliminate as many as possible terms in the upper bound of the error by carefully defining the numerical flux of the LDG method and by suitably choosing such a projection. Indeed, instead of using the L-projection operator used by Cockburn and Shu Reference 11, the projections used by Houston, Schwab and Süli Reference 14Reference 30Reference 29, or the Lagrange interpolation of Gauss-Radau points used by Castillo Reference 3, we pick the more advantageous projection used in 1985 by Thomeé Reference 31 in his study of discontinuous Galerkin time-discretizations for parabolic problems and recently by Schötzau and Schwab Reference 23Reference 22 in their study of the -version of this method. Indeed, with this projection, many terms in the upper bound of the error become identically zero which allows us to obtain an optimal rate of convergence after a simple application of the sharp -approximation results for this projection.

Recent work on other discontinuous finite element methods for convection-diffusion (and for pure diffusion) problems has been reviewed by Cockburn, Karniadakis and Shu Reference 7. See, in particular, the numerical method of Baumann and Oden Reference 2, the optimal error estimates for the method as applied to nonlinear convection-diffusion equations by Rivière and Wheeler Reference 20, the analysis of several versions of the Baumann and Oden method for elliptic problems by Rivière, Wheeler and Girault Reference 21, and the -version analyses by Houston, Süli and Schwab Reference 14Reference 30Reference 29 of discontinuous Galerkin methods for second-order problems with nonnegative characteristic form. We also mention the recent work of Wihler and Schwab Reference 32 in which robust exponential rates of convergence of DG methods for (stationary) convection-diffusion problems in one space dimension are proven.

The organization of the paper is as follows. In section 2, we describe the LDG method. In section 3, we state and prove the a priori error estimate for the constant-coefficient convection-diffusion problem and the special numerical flux for which the estimates are optimal in both and . In section 4, we discuss several extensions and, in section 5, we perform numerical experiments that verify the theoretical results. We end our presentation with some concluding remarks in section 6.

2. The LDG method

In this section, we introduce and briefly discuss the various key elements of the LDG method for a simple model problem.

2.1. The model problem and its weak formulation

In this paper we consider the following model convection-diffusion equation in one space dimension:

with the initial condition

and the Dirichlet boundary conditions

The unknown function is a scalar, and we assume the velocity to be a positive and the diffusion coefficient to be a nonnegative number; we choose to work with a positive velocity simply to fix the location of the possible boundary layer at . Note that in the purely convective case , only the Dirichlet boundary condition at is taken into account.

The weak formulation we are going to use is obtained as follows. First, we introduce the new variable and the “flux” function

and rewrite Equation 2.1Equation 2.3 in the form

Next, given the nodes , we define the mesh and set ; furthermore, we define . To the mesh , we associate the so-called broken Sobolev space

For a function the one-sided limits at the nodes are denoted as follows:

Throughout, we assume the exact solution of (Equation 2.1)–(Equation 2.3) belongs to . Then, it satisfies the following equations:

for all test functions and for . Here, the time derivative is to be understood in the weak sense and .

2.2. The method

A discrete version of the above mixed formulation is obtained by restricting the trial and test functions to finite-dimensional subspaces and by replacing the flux function at the nodes by a numerical flux : Find such that for all and for the following equations hold:

Upon a choice of basis for the subspaces , and, more importantly, of the numerical fluxes, the semidiscrete problem Equation 2.5 becomes an ODE system of dimension on , where . In what follows we do not consider the impact of the time discretization and refer to Shu Reference 28, Shu and Osher Reference 26Reference 27 and Gottlieb and Shu Reference 13 for the analysis of certain TVD Runge-Kutta methods for the solution of ODE systems.

Our choice of the space is the space of discontinuous, piecewise polynomial functions

where denotes the set of all polynomials of degree less or equal than on . Notice that the polynomial degrees can vary from element to element.

To complete the definition of the LDG method, it remains to define the numerical flux .

2.3. The numerical flux

Crucial for the stability as well as for the accuracy of the LDG method is the choice of the numerical flux . To define it, we introduce with the notation in (Equation 2.4) the following quantities:

The numerical flux has the following general form:

which also holds at the boundary if we define

Let us stress several important points concerning this numerical flux:

In the purely hyperbolic case, i.e., in the case , if we take , we obtain the well-known “upwinding” flux of the original DG method; see, e.g., Reference 4.

Note that . This is so because we want to be able to solve for in terms of element by element. This local solvability, which gives the name to the LDG method, is not shared by most mixed methods and allows us to easily eliminate the unknown from the equations.

The main purpose of the coefficient is to enhance the stability of the method; that is why it must be a positive number. This results in an improvement of the accuracy of the method too.

The choice ensures the stability of the LDG method.

The main purpose of the coefficients is to enhance the accuracy of the method. Thus, if we take, following Bassi and Rebay Reference 1, , the rate of convergence of the energy norm is of order for smooth functions. If instead, following Castillo Reference 3, we take , we obtain the optimal rate of order .

Let us point out that, if we consider the general form of the numerical fluxes, it is possible to obtain exponential convergence for piecewise analytic exact solutions, but not optimality in both and . As we shall see, this optimality is guaranteed for completely arbitrary meshes if we take (an extension of) Castillo’s Reference 3 choice of the numerical flux , namely,

where .

This flux is obtained by setting and

Note again that in the purely convective case, , this numerical flux is nothing but the standard upwinding flux used by the original DG method. Note also that the fact that the coefficient has a special form is a reflection of the fact that at there might be a boundary layer which requires special treatment. The factor ensures the optimality in both and of the energy norm of the error.

The definition of the LDG method is now complete.

3. Error analysis

This section is devoted to our main a priori error estimates. First, we state and briefly discuss the results; the remainder of the section is devoted to their proof.

3.1. A priori estimates

Our main result follows naturally from an estimate of a suitably defined projection of the error and from the -approximation properties of the projection . Each of these results is contained in several lemmas that we state next. To do that, we need to introduce the projection and the norm in which we measure .

The projection is the operator from to that associates to where, for each interval , is defined by the following conditions:

Let us now introduce a norm that appears naturally in the analysis of the LDG method. In what follows, we denote by the L-norm in the subdomain and omit when . For the norm is defined as follows:

where the energy norm is given by

and

Note that information about the numerical flux is contained in the norm only through . We are now ready to state our results.

Lemma 3.1 (The basic estimate).

The error between the exact solution and the approximation given by the LDG method with numerical flux (Equation 2.6) satisfies the inequality

where

and

If we combine the above result with the estimates of , for and , we immediately obtain our desired error bound. To state the corresponding approximation results, we introduce on the reference interval and for the following weighted seminorm:

We can now state our approximation result.

Lemma 3.2 (The -approximation estimates for fixed ).

Let be the reference interval and a generic polynomial degree on . Assume for and . Then we have the following estimates:

where depends on but is independent of and .

For standard finite element methods, the introduction of the weighted norms enables one to show that for singular solutions the -version of the method, i.e., when the mesh is fixed and increases unboundedly, yields twice the convergence rate than the -version provided that the singularity lies at a mesh point ; see, e.g., Schwab Reference 24. The results in Lemma 3.2 are slightly suboptimal with regard to these aspects as will be shown in the numerical experiments in section 5 below. However, for smooth solutions we obtain optimal approximation properties for in and . This can be inferred immediately from Lemma 3.2 and standard scaling and interpolation arguments.

Lemma 3.3 (The -approximation estimates for fixed ).

Let for and . Then we have the following estimates:

where depends on but is independent of , and .

Now, assume that , where

Our main result is a simple consequence of the above lemmas. Indeed, since

from Lemmas 3.1 and 3.3, we obtain the following result.

Theorem 3.4 (The estimate of the energy norm).

Let be the error between the exact solution and the approximation given by the LDG method with numerical flux (Equation 2.6) and polynomial degree on each interval. Then, for totally arbitrary meshes, the energy norm of the error satisfies the inequality

Remark 3.5.

The error estimates in Theorem 3.4 are optimal in and for smooth solutions, even for the case in which piecewise constant approximations () are used. Note also that in the purely convective case, , the above error estimate is nothing but the extension of the super-convergence error estimate of LeSaint and Raviart Reference 15 for the -version of the DG method for purely convective problems.

Remark 3.6.

The proof of Lemma 3.3 actually gives us estimates which are completely explicit in the mesh-width , in the polynomial degree , and in the regularity of the exact solution (see Proposition 3.12 below). Hence, completely explicit error estimates in the energy norm can be obtained. In conjunction with geometric meshes and linearly increasing polynomial degrees, such estimates can be used in the -version to prove exponential rates of convergence in the presence of solution singularities; see, e.g., the recent monograph by Schwab Reference 24 and the references therein. However, since the corresponding analytic regularity in space-time still remains to be found, we do not further pursue these issues here.

Remark 3.7.

From Theorem 3.4, we conclude that in the -version of the LDG method, where the mesh is kept fixed and the polynomial degree is increased, we have . Hence, for smooth solutions convergence rates of arbitrarily high algebraic order in are possible. This is sometimes referred to as spectral convergence. Furthermore, for solutions which are analytic in , even exponential rates of convergence are obtained in the -version, i.e.,

with constants independent of . This result can immediately be derived from Lemma 3.1, properties of (see, e.g., (Equation 3.15) and (Equation 3.16) below) and standard approximation theory for analytic functions. We note that (Equation 3.4) holds true for general fluxes as well.

Remark 3.8.

For small diffusivities , i.e., for in (Equation 2.1), the solutions typically exhibit viscous boundary layers (or shock profiles) of length scale or . In principle, layer components in the solutions are still analytic (see the work of Melenk Reference 16 and Melenk and Schwab Reference 19Reference 18 for a complete characterization of boundary layers in stationary problems with analytic input data) and can thus be approximated at exponential rates of convergence, in agreement with (Equation 3.4). However, the estimate (Equation 3.4) is not robust with respect to the diffusivity and deteriorates as . A remedy is to employ a needle-element of the appropriate width or geometric mesh refinement near the boundary. It has been shown recently by Melenk Reference 16, Melenk and Schwab Reference 17Reference 19, Schwab and Suri Reference 25 and Wihler and Schwab Reference 32 that the use of these mesh-design principles yields exponential rates of convergence that are robust with respect to the diffusivity parameter . We demonstrate this robustness in our numerical examples in section 5.5 below.

3.2. Proof of the basic estimate

This section is devoted to the proof of Lemma 3.1. To do so, we follow the technique used by Cockburn and Shu Reference 11; see also Cockburn Reference 4, Castillo Reference 3 and Cockburn and Dawson Reference 5.

We start by rewriting the definition of the LDG method in compact form for general numerical fluxes. Integrating Equation 2.5 with respect to from to and summing over all elements, it turns out that the LDG solution is defined as follows:

Find such that

where the discrete bilinear form is given by

and the discrete linear form is given by

The basic error estimate now follows by standard manipulations. Indeed, since we have that

we obtain that

where , and this implies that

It only remains to obtain a suitable expression for and an upper bound for .

Lemma 3.9.

For any , it holds that

where

and is defined by (Equation 3.3).

Proof.

This is a direct consequence of the definition of the form .

Lemma 3.10.

For any and the numerical flux (Equation 2.6), we have

Proof.

Taking into account the definition of the form , (Equation 3.6), and the definition of the projection , (Equation 3.1) and (Equation 3.2), we easily get that

The result follows from simple applications of Cauchy-Schwarz’s and Young’s inequalities and from the definition of the functional defined in Lemma 3.9.

Now, inserting the results of Lemmas 3.9 and 3.10 into (Equation 3.8), we get the inequality

which is of the form

with

Since inequality (Equation 3.9) holds true for all , Lemma 3.1 now follows after a simple application of the following result.

Lemma 3.11.

Suppose that for all we have

where , , and are nonnegative functions. Then, for any ,

Proof.

Define and fix . Setting , the hypothesis implies that for

Integrating over yields

Hence,

Since , the assertion in Lemma 3.11 follows.

This completes the proof of Lemma 3.1.

3.3. Proof of the -approximation results

This section is devoted to the proof of Lemma 3.2 which follows from the subsequent finer approximation results after an application of Stirling’s formula.

Let and recall that

We have:

Proposition 3.12.

Let for . Let be defined by

Then we have

and

for any .

Proof.

The first estimate was obtained by Schötzau Reference 22 and Schötzau and Schwab Reference 23. Nevertheless, we present a detailed proof for the sake of completeness. We consider only since the proof for is similar. We proceed in several steps.

Step 1: First, we derive bounds on the difference in terms of the Legendre coefficients of . To do so, denote by , , the Legendre polynomial of degree on and expand the function into the series with . Since , it can be seen from (Equation 3.10) that is uniquely given by the series

Hence, the difference can be written as

Let be the -projection from onto . Since , and , we obtain from the definition of the following bounds:

Step 2: To estimate the sums in the above equalities, we start by expanding into the series . Integrating this expression yields

and employing for the identity

and rearranging terms, we obtain

Comparing coefficients in the Legendre expansions, we conclude that

Hence, after some simple algebraic manipulations,

and

As a consequence,

and since , we get

Step 3: Now, note that after inserting the estimates (Equation 3.13) and (Equation 3.14) into (Equation 3.11) and (Equation 3.12), respectively, we get

Replacing in these inequalities by , where is an arbitrary polynomial of degree , and taking into account that and give

Schwab Reference 24 proved that

for any , and the existence of a polynomial such that

for any . We now simply insert these estimates into (Equation 3.15) and (Equation 3.16) to conclude that

for any .

The corresponding estimates for are obtained by symmetry. Since

this proves Proposition 3.12.

From Proposition 3.12 we obtain by standard scaling and interpolation arguments the following -approximation properties of :

Corollary 3.13.

For each interval , , we have for , real, the estimates

and

for any . The constant is independent of , and .

4. Extensions

The a priori error estimate of Theorem 3.4 can be easily extended to the case of general boundary conditions and to general numerical fluxes.

4.1. Other boundary conditions

Theorem 3.4 holds unchanged for Neumann, Robin or mixed boundary conditions. To see this, let us consider, for example, the following Neumann boundary conditions:

First, we take

Then we redefine the numerical flux as follows:

Note that this flux is obtained by setting and

Now, we simply have to go through the proof of Lemma 3.1 in section 3.2 to verify that the basic error estimate of Lemma 3.1 still holds with as before and

This and the approximation result of Lemma 3.3 imply that the estimate of Theorem 3.4 holds in this case.

4.2. General numerical fluxes

In the case of general numerical fluxes, the optimality of the estimate of Theorem 3.4 is lost in both and . Theoretically, the main reason is that now there are new terms in which were equal to zero for the flux (Equation 2.6).

Indeed, in the purely convective case, , the new term is

which is estimated as follows:

The last two terms are absorbed by the term and the first two are bounded, using Lemma 3.3, by

where

Hence, the error estimate is

Note the loss of half a power in both and .

In the case in which , the following additional term appears:

By using the approximation results of Lemma 3.3, we see that we can bound as follows:

where

and

Next, we use the following inverse inequality:

for in order to get

Note that this produces an additional loss of half power in and a full power in . Thus, after a few simple manipulations, we obtain the following estimate for general numerical fluxes:

5. Numerical results

The purpose of this section is to numerically validate the a priori error estimates given in section 3. In all our experiments, we use a TVD Runge–Kutta time stepping method (see Shu and Osher Reference 26Reference 27), with sufficiently small time steps, such that the overall error is governed by the spatial error.

5.1. Exponential convergence

Our first example illustrates the exponential convergence in for analytic solutions. We solve (Equation 2.1) on the space-time domain , with exact solution . We use a fixed grid consisting of a uniform mesh with 4 elements and increase the polynomial degree . The corresponding errors in the energy norm at time are shown in Figure 1. The diffusion coefficient is and the convection coefficient is chosen as (left) and (right). The curves clearly show exponential rates of convergence as predicted in (Equation 3.4) of section 3. Since the quadrature points and weights used to determine the LDG solution are computed only with an accuracy of , the curves bottom out for .

5.2. Optimal order of convergence in

In these examples, we show that an optimal order of convergence of is achieved when using the numerical flux in (Equation 2.6). For this set of tests, we solve (Equation 2.1) on , again with exact solution . To determine numerically the convergence order we consider the two sequences of successively refined meshes shown in Figure 2. Since our analysis is valid for arbitrary meshes, we choose the second sequence to consist of nonuniform meshes whereas the first one contains uniform meshes. Note that in both cases the mesh-size parameter of is half of the one of .

If denotes the error on the th mesh in the energy norm, then the numerical rate of convergence is defined as

In Tables 1, 2 and 3, we present these numerical orders in the energy norm at for polynomials of degree to on the above two mesh-sequences. In all the experiments we use the same convection coefficient and increase the diffusion coefficient from to . The results show that our estimates are optimal in not only for convection dominated problems, but also for diffusion dominated problems. In all the cases the numerical orders agree with the theoretical orders of our error estimates in Theorem 3.4.

5.3. Nonsmooth solutions

In this subsection, we present some numerical results to illustrate the performance of the LDG method for a solution that is nonsmooth in space.

We consider first the -version and start by solving the purely convective () problem (Equation 2.1), on with and with data chosen in such a way that the exact solution is . The corresponding uniform and nonuniform spatial discretizations are similar to those used in section 5.2 (cf., Figure 2). In this purely convective problem an order of convergence of is expected from our error estimate in section 3. These orders can clearly be seen in Table 4. Again, they are calculated for the energy norm at .

Now, we consider the linear problem with diffusion, i.e., with and . Again, we choose the data in such a way that the exact solution is . The results at are shown in Table 5. From our a priori error estimate, an order of convergence of is expected. This is what we actually see for all values of except for . In this case, we observe an order of convergence of instead of the predicted . Since the order of convergence for is smaller than , we believe that an error cancellation might be taking place which occurs only for .

Since the solution is singular at the mesh point , we expect a doubling of the convergence rate in the -version where is increased on a fixed mesh ; see, e.g., Schwab Reference 24. This is shown in Figure 3 for the same model problems as above. We can see a convergence rate of in the purely hyperbolic case () and of in the convection-diffusion case, respectively, which corresponds to an exact doubling of the rates. However, in our theoretical results, if we insert the weighted bounds from Lemma 3.2 in the proof of Theorem 3.4, we obtain rates of in the hyperbolic case and of in the convection-diffusion case, resulting in a loss of one power of and indicating the suboptimality of Lemma 3.2 with respect to the weighted spaces .

5.4. Testing the optimality of the smoothness requirement

To test if the smoothness on the exact solution required by our Theorem 3.4 when is optimal, we consider problem (Equation 2.1) with , , homogeneous Dirichlet boundary conditions and initial data . The results at are given in Table 6. Theorem 3.4 predicts an order of convergence of but we actually see an order of convergence of . This gives a strong indication that, to obtain optimal orders of convergence at least in , less smoothness of the exact solution than required in Theorem 3.4 for is sufficient. However, obtaining optimality in the smoothness of the exact solution seems to ask for more sophisticated theoretical techniques than the ones available in the current literature and has to be addressed in future work.

5.5. Robust exponential convergence

Our last example shows that robust exponential convergence can be obtained in the presence of a boundary layer, when suitable meshes are used; see Remark 3.8. We solve (Equation 2.1) on with , and right-hand side such that the exact solution is . For small , this solution has an exponential boundary layer of strength at the outflow boundary . In Figure 4, we compare the -version of the LDG method when using uniform and geometric meshes. Both meshes are chosen in such a way that they have the same number of elements; however, the distribution of the grid points is different: In the geometric mesh the size of the first element near is on the order of the length of the boundary layer, , the size of the next element is twice the size of the previous and so forth. The errors in the energy norm at for and are depicted in Figure 4. All the curves show exponential rates of convergence. However, the robustness of the rates for the geometric boundary layer meshes can clearly be observed, whereas the uniform mesh performs orders of magnitude worse for .

6. Concluding remarks

In this paper, we have obtained optimal error estimates for the -version of the LDG method for the model problem of the initial boundary value problem for a one-dimensional convection-diffusion equation. We have shown that this is possible by a careful choice of the numerical fluxes and the associated projections and ; we have also shown how this optimality in and is lost, at least theoretically, when general numerical fluxes are used.

Our numerical results confirm the optimality in of our main result and the exponential convergence that follows when the solution is analytic. These results also indicate that the smoothness requirement on the exact solution is too stringent. The problem of obtaining optimality in the smoothness of the exact solution seems to ask for more sophisticated theoretical techniques than the ones available in the current literature and constitute the subject of ongoing work. Also, extensions of our main result to the more challenging cases of nonconstant coefficients and , and to the multidimensional case will be considered elsewhere.

Figures

Figure 1.

Exponential convergence in for an analytic exact solution. In both examples, the diffusion coefficient is . The convection coefficient is (left) and (right).

Graphic without alt text
Figure 2.

Sequence of uniform and nonuniform meshes in .

Graphic without alt text
Table 1.

Orders of convergence for the -version and a smooth exact solution with .

Nonuniform gridUniform grid
00.67330.69990.88010.49640.78170.8728
11.55272.02951.98461.81231.87391.9658
22.69722.86632.93842.55802.91902.9504
33.68914.18493.99484.03933.94723.9840
44.83924.93884.95624.70864.94894.9681
55.80426.20345.99376.12685.96605.9880
66.86736.97576.93656.78126.95666.9652
Table 2.

Orders of convergence for the -version and a smooth exact solution with .

Nonuniform gridUniform grid
00.64251.14820.93940.71240.92930.9405
11.65911.72622.01371.48811.95531.9914
23.05883.19972.96553.16562.94582.9738
33.82423.67913.99193.59693.97273.9888
44.96995.21524.97855.16624.95454.9840
55.87605.69785.99265.63915.97825.9915
66.96447.22176.97487.18136.96596.9860
Table 3.

Orders of convergence for the -version and a smooth exact solution with .

Nonuniform gridUniform grid
00.71981.20080.97730.90990.93300.9768
11.64891.60601.99931.24951.96961.9920
23.00703.22802.98583.22972.95822.9875
33.51193.57573.99393.12843.97283.9926
45.03405.22644.99055.27424.97044.9917
55.45125.59805.99495.08665.97905.9944
67.06327.22516.96897.30316.97746.9817
Table 4.

Orders of convergence for the -version and the non-smooth exact solution for the purely convective case .

Nonuniform gridUniform grid
00.09321.04171.00290.97510.99260.9976
10.27051.88391.91631.80821.89661.9524
20.44192.89612.96842.87952.97082.9776
31.38953.48323.56053.52433.60763.6162
42.88543.60593.62093.60673.62243.6310
53.53723.62503.63163.62303.62523.6332
63.61943.62883.62983.62523.62923.6343
Table 5.

Orders of convergence for the -version and the non-smooth exact solution for the convection-diffusion case .

Nonuniform gridUniform grid
00.75460.83270.96690.84050.96040.9957
11.82731.95421.99921.93831.98291.9956
22.98913.00663.00252.98532.98112.9704
33.00402.78692.68612.67502.64932.6430
42.64932.64262.64132.64242.64082.6409
52.63992.64032.64082.63972.64032.6408
62.63932.64032.64072.63932.64032.6408
Figure 3.

The -version for the nonsmooth exact solution . The convection coefficient is and the diffusion coefficient is (top curve) and (bottom curve).

Graphic without alt text
Table 6.

Orders of convergence for the -version and a non-smooth exact solution corresponding to the initial data with .

Nonuniform gridUniform grid
00.87650.90200.85550.92350.95460.9370
11.70591.84951.88871.84911.91581.9044
22.48442.48542.50472.48222.46722.5089
32.49842.49532.49082.48842.49792.3841
42.50012.49612.48082.48932.47162.0190
52.50012.50342.48752.49112.46662.4900
62.50222.50222.49322.49632.49492.4618
Figure 4.

Exponential rates of convergence in the presence of a boundary layer on uniform and geometric meshes for and .

Graphic without alt text

Mathematical Fragments

Equation (2.1)
Equation (2.3)
Equation (2.4)
Equation (2.5)
Equation (2.6)
Equations (3.1), (3.2)
Equation (3.3)
Lemma 3.1 (The basic estimate).

The error between the exact solution and the approximation given by the LDG method with numerical flux (Equation 2.6) satisfies the inequality

where

and

Lemma 3.2 (The -approximation estimates for fixed ).

Let be the reference interval and a generic polynomial degree on . Assume for and . Then we have the following estimates:

where depends on but is independent of and .

Lemma 3.3 (The -approximation estimates for fixed ).

Let for and . Then we have the following estimates:

where depends on but is independent of , and .

Theorem 3.4 (The estimate of the energy norm).

Let be the error between the exact solution and the approximation given by the LDG method with numerical flux (Equation 2.6) and polynomial degree on each interval. Then, for totally arbitrary meshes, the energy norm of the error satisfies the inequality

Remark 3.7.

From Theorem 3.4, we conclude that in the -version of the LDG method, where the mesh is kept fixed and the polynomial degree is increased, we have . Hence, for smooth solutions convergence rates of arbitrarily high algebraic order in are possible. This is sometimes referred to as spectral convergence. Furthermore, for solutions which are analytic in , even exponential rates of convergence are obtained in the -version, i.e.,

with constants independent of . This result can immediately be derived from Lemma 3.1, properties of (see, e.g., (3.15) and (3.16) below) and standard approximation theory for analytic functions. We note that (3.4) holds true for general fluxes as well.

Remark 3.8.

For small diffusivities , i.e., for in (Equation 2.1), the solutions typically exhibit viscous boundary layers (or shock profiles) of length scale or . In principle, layer components in the solutions are still analytic (see the work of Melenk Reference 16 and Melenk and Schwab Reference 19Reference 18 for a complete characterization of boundary layers in stationary problems with analytic input data) and can thus be approximated at exponential rates of convergence, in agreement with (Equation 3.4). However, the estimate (Equation 3.4) is not robust with respect to the diffusivity and deteriorates as . A remedy is to employ a needle-element of the appropriate width or geometric mesh refinement near the boundary. It has been shown recently by Melenk Reference 16, Melenk and Schwab Reference 17Reference 19, Schwab and Suri Reference 25 and Wihler and Schwab Reference 32 that the use of these mesh-design principles yields exponential rates of convergence that are robust with respect to the diffusivity parameter . We demonstrate this robustness in our numerical examples in section 5.5 below.

Equation (3.6)
Equation (3.8)
Lemma 3.9.

For any , it holds that

where

and is defined by (Equation 3.3).

Lemma 3.10.

For any and the numerical flux (Equation 2.6), we have

Equation (3.9)
Lemma 3.11.

Suppose that for all we have

where , , and are nonnegative functions. Then, for any ,

Proposition 3.12.

Let for . Let be defined by

Then we have

and

for any .

Equations (3.11), (3.12)
Equations (3.13), (3.14)
Equations (3.15), (3.16)

References

Reference [1]
F. Bassi and S. Rebay, A high-order accurate discontinuous finite element method for the numerical solution of the compressible Navier-Stokes equations, J. Comput. Phys. 131 (1997), 267–279. MR 97m:76078
Reference [2]
C.E. Baumann and J.T. Oden, A discontinuous finite element method for convection-diffusion problems, Comput. Methods Appl. Mech. Engrg. 175 (1999) 311–341. MR 2000d:65171
Reference [3]
P. Castillo, An optimal error estimate for the local discontinuous Galerkin method, Discontinuous Galerkin Methods: Theory, Computation and Applications (B. Cockburn, G.E. Karniadakis, and C.-W. Shu, eds.), Lectures Notes in Computational Science and Engineering, vol. 11, Springer-Verlag, 2000, pp. 285–290.
Reference [4]
B. Cockburn, Discontinuous Galerkin methods for convection-dominated problems, High-Order Methods for Computational Physics (T. Barth and H. Deconink, eds.), Lecture Notes in Computational Science and Engineering, vol. 9, Springer Verlag, 1999, pp. 69–224. MR 2000f:76095
Reference [5]
B. Cockburn and C. Dawson, Some extensions of the local discontinuous Galerkin method for convection-diffusion equations in multidimensions, in Proceedings of the Conference on the Mathematics of Finite Elements and Applications: MAFELAP X (J. Whiteman, ed.), Elsevier, 2000, pp. 225–238.
Reference [6]
B. Cockburn, S. Hou, and C.-W. Shu, TVB Runge-Kutta local projection discontinuous Galerkin finite element method for conservation laws IV: The multidimensional case, Math. Comp. 54 (1990), 545–581. MR 90k:65162
Reference [7]
B. Cockburn, G.E. Karniadakis, and C.-W. Shu, The development of discontinuous Galerkin methods, Discontinuous Galerkin Methods: Theory, Computation and Applications (B. Cockburn, G.E. Karniadakis, and C.-W. Shu, eds.), Lecture Notes in Computational Science and Engineering, vol. 11, Springer-Verlag, 2000, pp. 3-50.
Reference [8]
B. Cockburn, S.Y. Lin, and C.-W. Shu, TVB Runge-Kutta local projection discontinuous Galerkin finite element method for conservation laws III: One dimensional systems, J. Comput. Phys. 84 (1989), 90–113. MR 90k:65161
Reference [9]
B. Cockburn and C.-W. Shu, TVB Runge-Kutta local projection discontinuous Galerkin finite element method for scalar conservation laws II: General framework, Math. Comp. 52 (1989), 411–435. MR 90k:65160
Reference [10]
—, The Runge-Kutta local projection -discontinuous Galerkin method for scalar conservation laws, RAIRO Modél. Math. Anal.Numér. 25 (1991), 337–361. MR 92e:65128
Reference [11]
—, The local discontinuous Galerkin finite element method for convection-diffusion systems, SIAM J. Numer. Anal. 35 (1998), 2440–2463. MR 99j:65163
Reference [12]
—, The Runge-Kutta discontinuous Galerkin finite element method for conservation laws V: Multidimensional systems, J. Comput. Phys. 141 (1998), 199–224. MR 99c:65181
Reference [13]
S. Gottlieb and C.-W. Shu, Total variation diminishing Runge-Kutta schemes, Math. Comp. 67 (1998), 73–85. MR 98c:65122
Reference [14]
P. Houston, C. Schwab, and E. Süli, Discontinuous finite element methods for advection–diffusion problems, in preparation.
Reference [15]
P. LeSaint and P.A. Raviart, On a finite element method for solving the neutron transport equation, Mathematical aspects of finite elements in partial differential equations (C. de Boor, ed.), Academic Press, 1974, pp. 89–145. MR 58:31918
Reference [16]
J.M. Melenk, On the robust exponential convergence of -finite element methods for problems with boundary layers, IMA J. Num. Anal. 17 (1997), 577–601. MR 99b:65145
Reference [17]
J.M. Melenk and C. Schwab, -FEM for reaction-diffusion equations I: Robust exponential convergence, SIAM J. Numer. Anal. 35 (1998), 1520–1557. MR 99d:65333
Reference [18]
—, Analytic regularity for a singularly perturbed problem, SIAM J. Math. Anal. 30 (1999), 379–400. MR 2000b:35010
Reference [19]
—, An finite element method for convection-diffusion problems in one space dimension, IMA J. Num. Anal. 19 (1999), 425–453. MR 2000k:65176
Reference [20]
B. Rivière and M.F. Wheeler, A discontinuous Galerkin method applied to nonlinear parabolic equations, Discontinuous Galerkin Methods: Theory, Computation and Applications (B. Cockburn, G.E. Karniadakis, and C.-W. Shu, eds.), Lecture Notes in Computational Science and Engineering, vol. 11, Springer-Verlag, 2000, pp. 231–244.
Reference [21]
B. Rivière, M.F. Wheeler, and V. Girault, Improved energy estimates for interior penalty, constrained and discontinuous Galerkin methods for elliptic problems. Part I, Comput. Geosci. 3 (1999) 337–360. CMP 2000:10
Reference [22]
D. Schötzau, -DGFEM for parabolic evolution problems - applications to diffusion and viscous incompressible fluid flow, Ph.D. thesis No. 13041, Swiss Federal Institute of Technology, Zürich, 1999.
Reference [23]
D. Schötzau and C. Schwab, Time discretization of parabolic problems by the -version of the discontinuous Galerkin finite element method, SIAM J. Numer. Anal. 38 (2000), 837–875. CMP 2001:01
Reference [24]
C. Schwab, - and -FEM. Theory and Application to Solid and Fluid Mechanics, Oxford University Press, New York, 1998. MR 2000d:65003
Reference [25]
C. Schwab and M. Suri, The - and -version of the finite element method for problems with boundary layers, Math. Comp. 65 (1996), 1403–1429. MR 97a:65067
Reference [26]
C.-W. Shu and S. Osher, Efficient implementation of essentially non-oscillatory shock-capturing schemes, J. Comput. Phys. 77 (1988), 439–471. MR 89g:65113
Reference [27]
—, Efficient implementation of essentially non-oscillatory shock capturing schemes, II, J. Comput. Phys. 83 (1989), 32–78. MR 90i:65167
Reference [28]
C.-W. Shu, TVD time discretizations, SIAM J. Sci. Stat. Comput. 9 (1988), 1073–1084. MR 90a:65196
Reference [29]
E. Süli, P. Houston, and C. Schwab, -finite element methods for hyperbolic problems, The Proceedings of the Conference on the Mathematics of Finite Elements and Applications: MAFELAP X (J.R. Whiteman, ed.), Elsevier, 2000, pp. 143–162.
Reference [30]
E. Süli, C. Schwab, and P. Houston, -DGFEM for partial differential equations with non-negative characteristic form, Discontinuous Galerkin Methods: Theory, Computation and Applications (B. Cockburn, G. E. Karniadakis, and C.-W. Shu, eds.), Lecture Notes in Computational Science and Engineering, vol. 11, Springer-Verlag, 2000, pp. 221–230.
Reference [31]
V. Thomée, Galerkin Finite Element Methods for Parabolic Equations, Springer Verlag, 1997. MR 98m:65007
Reference [32]
T. Wihler and C. Schwab, Robust exponential convergence of the discontinuous Galerkin FEM for convection-diffusion problems in one space dimension, East West J. Num. Math. 8 (2000), 57–70. CMP 2000:12

Article Information

MSC 2000
Primary: 65N30 (Finite elements, Rayleigh-Ritz and Galerkin methods, finite methods)
Secondary: 65M60 (Finite elements, Rayleigh-Ritz and Galerkin methods, finite methods)
Keywords
  • Discontinuous Galerkin methods
  • -methods
  • convection-diffusion
Author Information
Paul Castillo
School of Mathematics, University of Minnesota, 206 Church Street S.E., Minneapolis, Minnesota 55455
castillo@math.umn.edu
Bernardo Cockburn
School of Mathematics, University of Minnesota, 206 Church Street S.E., Minneapolis, Minnesota 55455
cockburn@math.umn.edu
Dominik Schötzau
School of Mathematics, University of Minnesota, 206 Church Street S.E., Minneapolis, Minnesota 55455
schoetza@math.umn.edu
Christoph Schwab
Seminar of Applied Mathematics, ETHZ, 8092 Zürich, Switzerland
schwab@sam.math.ethz.ch
Additional Notes

The second author was partially supported by the National Science Foundation (Grant DMS-9807491) and by the University of Minnesota Supercomputer Institute. The third author was supported by the Swiss National Science Foundation (Schweizerischer Nationalfonds).

Journal Information
Mathematics of Computation, Volume 71, Issue 238, 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 2001 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/S0025-5718-01-01317-5
  • MathSciNet Review: 1885610
  • Show rawAMSref \bib{1885610}{article}{ author={Castillo, Paul}, author={Cockburn, Bernardo}, author={Sch\" otzau, Dominik}, author={Schwab, Christoph}, title={Optimal a priori error estimates for the $hp$-version of the local discontinuous Galerkin method for convection--diffusion problems}, journal={Math. Comp.}, volume={71}, number={238}, date={2002-04}, pages={455-478}, issn={0025-5718}, review={1885610}, doi={10.1090/S0025-5718-01-01317-5}, }

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.