A locally conservative LDG method for the incompressible Navier-Stokes equations

By Bernardo Cockburn, Guido Kanschat, and Dominik Schötzau

Abstract

In this paper a new local discontinuous Galerkin method for the incompressible stationary Navier-Stokes equations is proposed and analyzed. Four important features render this method unique: its stability, its local conservativity, its high-order accuracy, and the exact satisfaction of the incompressibility constraint. Although the method uses completely discontinuous approximations, a globally divergence-free approximate velocity in is obtained by simple, element-by-element post-processing. Optimal error estimates are proven and an iterative procedure used to compute the approximate solution is shown to converge. This procedure is nothing but a discrete version of the classical fixed point iteration used to obtain existence and uniqueness of solutions to the incompressible Navier-Stokes equations by solving a sequence of Oseen problems. Numerical results are shown which verify the theoretical rates of convergence. They also confirm the independence of the number of fixed point iterations with respect to the discretization parameters. Finally, they show that the method works well for a wide range of Reynolds numbers.

1. Introduction

In this paper we propose and analyze a local discontinuous Galerkin (LDG) method for the stationary incompressible Navier-Stokes equations

Here is the kinematic viscosity, the velocity, the pressure, and the external body force. For the sake of simplicity, we take to be a polygonal domain of .

This paper is the fourth in a series (Reference 9, Reference 8, and Reference 7) devoted to the study of the LDG method as applied to incompressible fluid flow problems. In Reference 9, we considered the Stokes equations

and focused on the problem of how to deal with the incompressibility condition. Later, in Reference 8, we considered the Oseen equations

where the convective velocity was taken to be a smooth function, and focused on the problem of how to incorporate the linear convective term. The resulting method was shown to be optimally convergent and robust for a wide range of Reynolds numbers. A succinct review of this work can be found in Reference 7.

In this paper, we continue our study of LDG methods for incompressible flows and consider their application to the Navier-Stokes equations Equation 1.1. Our main concern is to devise LDG methods that are locally conservative and can be proven to be stable. The local conservativity, a property highly valued by practitioners of computational fluid dynamics, is a discrete version of the identity

where is an arbitrary subdomain of with outward normal unit vector . This property can be easily enforced by LDG methods as soon as the equations are written in divergence form. However, to devise an LDG method that can be proven to be stable (that is, that satisfies a discrete version of the stability estimate for the continuous case,

where the constant is the Poincaré constant) is extremely difficult.

The reason for this is that, in order to obtain the stability estimate Equation 1.5, the incompressibility condition must be used. It is well known that, for many numerical methods for the Stokes and the Oseen equations, a weakly enforced incompressibility is enough to guarantee stability. However, this is not so for the Navier-Stokes equations because of the presence of nonlinear convection. Moreover, the now standard solution to this problem, Reference 21, Reference 22, which is based on a suitable modification of the nonlinearity of the Navier-Stokes equations, cannot be used. This happens because such a modification does not have divergence form and hence prevents LDG methods from being locally conservative.

The main contribution of this paper is to show how to overcome this difficulty. In fact, we show that this can be done in two ways. The first one focuses on the convective nonlinearity and is based on a new modification in divergence form of the nonlinearity; it will be explored elsewhere in detail. The second one, which constitutes the main subject of this paper, focuses on the incompressibility constraint and is based on discretizing the Oseen equations Equation 1.3 where the convective velocity is taken to be a projection of the approximate velocity ,

into the space of globally divergence-free functions. This projection is a slight modification of well-known projections with the property

where is an -projection; see Reference 5. Its implementation is very efficient as it can be computed in an element-by-element fashion.

Thus, given the convective velocity , the resulting scheme is nothing but the LDG method Reference 8 already studied for the Oseen equations Equation 1.3. Since that method is stable, high-order accurate and locally conservative, so is the LDG method under consideration. Moreover, the approximation to the velocity given by  has continuous normal components across elements and is globally divergence free in . To the knowledge of the authors, no other numerical scheme for the incompressible Navier-Stokes equations has all these properties.

Of course, the convective velocity depends on the discrete velocity field  through Equation 1.6 and, hence, we need to use an iterative method to compute it. To do so, we note that if is the LDG approximate velocity of the Oseen problem with convective velocity , then the approximate velocity  of the LDG method under consideration is a fixed point of . If is proven to be a contraction, to compute the approximate solution , we can use the fixed point iteration

This is nothing but a discrete version of the argument used to prove the existence and uniqueness of the exact solution of Equation 1.1. It ensures the existence and uniqueness of the exact solution of Equation 1.1 under a smallness condition of the type

where only depends on ; see Reference 17, Theorem 10.1.1 and the references therein. We mimic this argument to show that the approximate solution of the LDG method exists and is unique under a similar condition.

Let us point out that exact incompressibility can be achieved trivially if the LDG method has a velocity space that is div-conforming, i.e., it is included in . In this particular case, weak incompressibility implies exact incompressibility, provided that the discrete spaces are matched correctly. As will be discussed below, this approach can be viewed as a particular LDG method for which the operator  is chosen to be the identity. Consequently, all the results of this paper hold true verbatim for methods that are based on div-conforming velocity spaces. Furthermore, we note that, although we have used the LDG method to discretize the terms associated with the viscosity effects, any other DG discretization whose primal form is both coercive and continuous could have been used to that effect; see the discussions in Reference 2 and Reference 19.

The organization of the paper is as follows. In Section 2 we discuss the ideas that motivate the devising of the LDG method we propose in this paper. In Section 3 we present the LDG discretization in detail and verify its local conservativity. In Section 4 we state and discuss the main results, namely, the stability of the method, the convergence of the fixed point iteration, and the a priori error estimates, and in Section 5 we present their proofs. In Section 6 we present numerical experiments verifying the theoretical results. We end in Section 7 with some concluding remarks.

2. Devising the LDG method

In this section we discuss the ideas that led us to the devising of an LDG method that is both stable and locally conservative. To keep the discussion as simple and clear as possible, we do not work with the numerical method. Instead, we work directly with the equations Equation 1.1 and infer, from their structure, the properties of the corresponding LDG method.

2.1. A locally conservative LDG method

Since the incompressible Navier-Stokes equations Equation 1.1 are written in divergence form, a locally conservative LDG method can be easily constructed. However, it is very difficult to prove its stability. Let us illustrate this difficulty by using the equations for the exact solution.

If we multiply the first equation of Equation 1.1 by , integrate by parts and use the boundary conditions, we get

We see that we must use the incompressibility condition to obtain the equation

from which the stability estimate Equation 1.5 immediately follows.

In general, since exact incompressibility is very difficult to achieve after discretization, it is usually only enforced weakly. This weak incompressibility is enough, in a wide variety of cases, to guarantee that the discrete version of the term

is exactly zero, as for most mixed methods for the Stokes and Navier-Stokes equations, or nonnegative, as for the LDG methods considered for the Stokes Reference 9 and Oseen Reference 8 problems. Unfortunately, this is not true for the discrete version of the term

because the square of the modulus of the approximate velocity does not necessarily belong to the space of the approximate pressure.

2.2. The classical modification of the nonlinearity

A solution to this impasse can be obtained by using a now classical technique proposed back in the 1960s; see Reference 21 and Reference 22. From our perspective, it consists in modifying the nonlinearity of the equations as follows:

To see that this solves the problem, multiply the first equation by , integrate by parts, and use the boundary conditions to get

As a consequence, stability can follow from weak incompressibility. This is the case for most mixed methods for the Navier-Stokes equations. It is also the case for the first discontinuous Galerkin method for the incompressible Navier-Stokes equations Reference 13, a method which uses locally divergence-free polynomial approximations of the velocity, and for the more recent discontinuous Galerkin method developed in Reference 10.

The only problem with this approach is that local conservativity cannot be achieved because the first equation is not written in divergence form.

2.3. A new modification of the nonlinearity

To overcome this difficulty, it is enough to take another glance at the first equation in this section to realize that instead of working with the kinematic pressure , we should work with the new variable

If we incorporate this unorthodox pressure into the Navier-Stokes equations, we get

We now see that, since the above modification is in divergence form, locally conservative LDG methods can easily be constructed. Moreover, since we have

we also see that the stability of the LDG method can follow from weak incompressibility. The LDG method obtained with this approach can indeed be proven to have those properties; it is going to be studied thoroughly in a forthcoming paper.

2.4. Enforcing exact incompressibility

As we have seen, suitable modifications of the nonlinearity can be introduced which allow us to obtain stability by using only weakly incompressible approximations to the velocity. The approach we consider in this paper does not rely on a modification of that type. Instead, it is based on enforcing exact incompressibility in the space

The idea that allows this to happen is based on two observations. The first is that we can rewrite the Navier-Stokes equations as the Oseen problem

where, of course, . If we multiply the first equation by , integrate by parts, and use the boundary conditions, we get

This suggests that we consider an LDG method with two different (but strongly related) approximations to the velocity: one approximation for and another for . The stability for the LDG method would then be achieved if the approximation to  is weakly incompressible and if the approximation to is exactly incompressible. Further, local conservativity can be readily achieved for such an LDG method. Indeed, the fact that the equations are not written in conservative form can be compensated for by the fact that the approximation to is globally divergence free.

The second observation is that, if the approximation to given by an LDG method is weakly incompressible, it is possible to compute, in an element-by-element fashion, another approximation, , which is exactly incompressible. As we shall see, the post-processing operator is a slight modification of the well-known Brezzi-Douglas-Marini (BDM) interpolation operator; see Reference 4.

We note that the post-processing procedure can be omitted in the particular case where the velocity space is div-conforming. Indeed, in this case the fact that the approximation to the velocity is weakly incompressible does imply that it is exactly incompressible, provided the pressure space is chosen suitably. Hence, we can take .

We are now ready to describe the LDG method in detail.

3. The LDG method

In this section, we introduce a locally conservative LDG discretization for the Navier-Stokes equations Equation 1.1.

3.1. Meshes and trace operators

We begin by introducing some notation. We denote by a regular and shape-regular triangulation of mesh-size of the domain  into triangles . We further denote by  the set of all interior edges of  and by  the set of all boundary edges. We set .

Next we introduce notation associated with traces. Let and be two adjacent elements of . Let be an arbitrary point of the interior edge . Let be a piecewise smooth scalar-, vector-, or matrix-valued function and let us denote by the traces of on taken from within the interior of . Then, we define the mean value at as

Further, for a generic multiplication operator , we define the jump at as

Here, denotes an outward unit normal vector on the boundary of element . On boundary edges, we set accordingly , and , with denoting the outward unit normal vector on .

3.2. The LDG method for the Oseen equations

We now recall the LDG method for the Oseen equations Equation 1.3. We assume that the convective velocity field  is in the space

We begin by introducing the auxiliary variable and rewriting the Oseen equations as

Next, we introduce the space where

for an approximation order . Here denotes the space of polynomials of total degree at most  on . For simplicity we consider here only so-called mixed-order elements where the approximation degree in the pressure is of one order lower than the one in the velocity.

Finally, we define the approximate solution by requesting that for each ,

for all test functions . Each of the above equations is enforced locally, that is, element by element, due to the appearance of the so-called numerical fluxes , , , , and .

Thanks to this structure of the LDG method we immediately get that

and since is globally divergence free, we obtain a discrete version of the property of local conservativity Equation 1.4, namely,

In other words, the LDG method is locally conservative.

To ensure that the method is also stable (and high-order accurate), the numerical fluxes, which are nothing but discrete approximations to the traces on the boundary of the elements, must be defined carefully. As we shall prove, the numerical fluxes that define the LDG method for the Oseen equations Reference 8 do ensure stability. For the sake of clarity we consider the fluxes in their simplest form.

The convective numerical flux

For the convective flux in Equation 3.3, we take the standard upwind flux introduced in Reference 15Reference 18. For an element , we set

where is the inflow part of given by

The diffusive numerical fluxes

If a face lies inside the domain , we take

and, if lies on the boundary, we take

As will be shown later, the role of the parameter is to ensure the stability of the method; see also Reference 6.

The numerical fluxes related to the incompressibility constraint

The numerical fluxes associated with the incompressibility constraint, and , are defined by using an analogous recipe. If the face lies on the interior of , we take

On the boundary, we set

This completes the definition of the LDG method for the Oseen problem in Equation 3.1.

Remark 3.1.

Notice that one can take the div-conforming velocity space

while keeping the other spaces and definitions above unchanged.

3.3. The post-processing operator

To complete the definition of the LDG method for the Navier-Stokes equations Equation 1.1, it only remains to introduce what we refer to as the post-processing operator and to set in the approximation Equation 3.2Equation 3.4.

For a piecewise smooth velocity field , we define the operator by

where is the numerical flux Equation 3.8Equation 3.9 related to the incompressibility constraint. For each element, the local operator is given via the moments

where

Here, denotes the elemental mapping and its Jacobian. On the reference triangle , the space is defined by

The post-processing operator  is well defined and can be computed in an element-by-element fashion. Moreover, if satisfies the equations Equation 3.4, that is, if it is weakly incompressible, then is exactly incompressible. These results are gathered in the next result and are given in terms of the Piola transformation, which maps any vector field on the reference triangle into

and of the BDM projection on , ; see Reference 4.

Proposition 3.2.

We have the following results.

(1)

is well defined and is in the space

(2)

If and , then .

(3)

If satisfies Equation 3.4, then in and .

Proof.

The proof of the first assertion is straightforward, and that of the second easily follows from the definitions of the projections and from the fact that if , then .

To prove the third assertion, we first observe that . This is due to the fact that for all and

in view of the definitions of and in Equation 3.9.

Now, let satisfy Equation 3.4. For , we obtain

Here, we have used integration by parts, the properties of and Equation 3.4. Thus, we have in . It follows that .

Remark 3.3.

For the LDG method using the div-conforming space in Equation 3.10, it can be readily seen that a field satisfying Equation 3.4 already is exactly incompressible and belongs to . Hence, for this particular LDG method, we can take as the identity.

3.4. The mixed setting of the LDG method

Next we recast the LDG method under consideration in a classical mixed setting, not only to facilitate its analysis but to be able to state our main results in a more precise way. Thus, we eliminate the auxiliary variable  and show that the approximation given by the LDG method satisfies

for all where

Here the forms , , and are associated to the discretization of the Laplacian, the convective term, and the incompressibility constraint, respectively. We proceed in several steps.

Step 1: Solving for in terms of

To be able to eliminate the auxiliary variable , we introduce the lifting operator by

It is now easy to see that the equation defining in terms of , equation Equation 3.2, can be rewritten as

with denoting the elementwise gradient. Note that, can be computed from  in an element-by-element fashion. Using this identity, it is possible to eliminate from the equations, as we show next.

Step 2: Eliminating

To eliminate from equation Equation 3.3, we make use of a second lifting operator given by

If we insert the expression of into equation Equation 3.3 and use the definitions of the numerical fluxes and the lifting operators, we readily get (see Reference 2, Reference 8, and Reference 19)

where

This completes the elimination of the auxiliary variable from the equations defining the LDG method. Note that exactly the same form is also used in the mixed DG approaches of Reference 11, Reference 23, Reference 19, Reference 10.

Step 3: Rewriting the incompressibility constraint

Finally, it is a simple exercise to see that equation Equation 3.4 can be rewritten as

This shows that the LDG method in Equation 3.2Equation 3.4 can be cast in the form given in Equation 3.11Equation 3.12.

4. The main results

In this section, we state and discuss the main results of this paper.

4.1. Preliminaries

We consider LDG methods with a very specific stabilization function . To define it, we introduce on the edges the local mesh-size function by for all , with denoting the length of the edge . We then set

with independent of the mesh-size and the viscosity .

The results will be stated in terms of norms we introduce next. We consider the space

endowed with the norm

For this norm, we have the Poincaré inequality

for a constant independent of the mesh-size; see, e.g., Reference 1Reference 3. Finally, the space is equipped with the -norm .

4.2. Stability properties of the bilinear forms

Here we collect crucial stability properties of the forms that are used to define the LDG method. Our main results will be stated in terms of the corresponding stability constants.

Continuity

First we study the continuity of the forms involved in the LDG method. We observe that the lifting operators  and  can be extended to operators  and , respectively, using the same definitions. It is then well known (from, e.g., Reference 16, Section 3 and Reference 19, Lemma 7.2) that

for a constant only depending on , the shape-regularity of the mesh, and the polynomial degree . As a consequence, the forms and are well defined and continuous on and .

Proposition 4.1.

We have that

for continuity constants and that are independent of the mesh-size.

Proof.

The continuity properties of and follow from Equation 4.4, Equation 4.5, and the Cauchy-Schwarz inequality; see Reference 16, Proposition 3.1 and Reference 19, Lemma 7.5.

Proposition 4.2.

Let , , and . Then we have

for a continuity constant that is independent of the mesh-size.

The proof is given in subsection 5.1. It is obtained by using the embedding and trace theorems of Reference 13 and Reference 10.

Coercivity and inf-sup condition

Next, we discuss the coercivity properties of the forms and . We have the following result.

Proposition 4.3.

Let be given by Equation 4.1. Then, for any , there exists a constant independent of the mesh-size such that

Furthermore, for and , it holds that

In the integrals over edges in , we denote by any unit normal to the edge under consideration.

For a proof of the coercivity of the form , we refer to Reference 2 or Reference 16. A similar coercivity result involving also the discrete velocity gradient was used in Reference 9 and Reference 8. We further note that, for the similar symmetric interior penalty forms used in the DG approach of Reference 11, the parameter has to be chosen large enough. The proof of the second assertion is standard; see, e.g., Reference 8.

Finally, we have the following inf-sup condition for the form .

Proposition 4.4 (Reference 11).

There exists a constant independent of the mesh-size such that

Extensions of this result to the -version of the finite element method and to quadrilateral meshes can be found in Reference 23, Reference 19.

Remark 4.5.

A careful inspection of the proof in Reference 11 reveals that the discrete inf-sup condition in Proposition 4.4 also holds for the smaller space in Equation 3.10. Consequently, all the stability results of this section hold for the particular LDG method in Remark 3.1.

Stability of the post-processing operator

The next result states that the operator is a bounded linear operator from to with respect to the norm . It is one of the main technical results needed to analyze the locally conservative LDG method Equation 3.11Equation 3.13.

Proposition 4.6.

Let . Then we have

with a stability constant that is independent of the mesh-size.

The proof of this proposition is given in subsection 5.2 below. We remark that, for the LDG method in Remark 3.1, we have since is chosen as the identity.

We are now ready to state our main results.

4.3. The results

Our first result states that, under a smallness condition similar to the one for the exact solution Equation 1.7, the LDG method Equation 3.11Equation 3.13 defines a unique discrete approximation. Moreover, it actually gives an efficient way to compute it.

Theorem 4.7 (Existence and uniqueness of discrete solutions).

Assume that

Then the LDG method Equation 3.11Equation 3.13 defines a unique solution . It satisfies the bounds

as well as

Moreover, if is the approximate solution given by the LDG method in Equation 3.11Equation 3.12 for the Oseen equations with , , then

for any initial guess .

This result, whose proof is given in subsection 5.3, states that the LDG method in Equation 3.11Equation 3.13 is well defined and that we can compute its approximate solution by solving a sequence of Oseen problems. Since the parameter is independent of the mesh-size, the convergence of that sequence is always exponential and so computationally efficient.

Note that if we set

then both the smallness assumptions in Equation 1.7 and Equation 4.6 are satisfied if we have that

Hence, both the Navier-Stokes equations and their LDG approximations are uniquely solvable. Under a smallness condition that is slightly more restrictive, we obtain the following estimates.

Theorem 4.8 (Error estimates).

Assume that

and that the exact solution of the Navier-Stokes equations Equation 1.1 satisfies

Then

where only depends on the regularity of the mesh and the polynomial degree , and

This result, whose proof is given in subsection 5.4, states that the LDG method under consideration converges with optimal order. Note also that, since the function is exactly divergence free, it provides an optimally convergent globally solenoidal approximation to the velocity!

Let us briefly discuss some extensions of these results.

• First, we point out that all the results of this section are valid verbatim for the LDG method in Remark 3.1 where is replaced by the div-conforming space  in Equation 3.10 and is chosen to be the identity.

• Although here we only considered the case of triangular meshes, the results of this paper can be straightforwardly extended to simplicial meshes in three dimensions. Furthermore, the LDG approach we propose here can be easily extended to elements on quadrilateral or hexahedral affine meshes by using a post-processing operator that is given by a slight modification of the BDM projection on quadrilaterals or hexahedra; see Reference 5. The results in this section then hold true for this LDG method as well. This fact is actually verified in our numerical experiments for which we have used square meshes and elements. However, the extension of our results to elements on quadrilateral meshes is not straightforward. Although, by using a post-processing operator that is a slight modification of the standard Raviart-Thomas projection, it is easy to define a solenoidal velocity field that belongs to the anisotropic polynomial space . The approximation properties in this space give rise to only suboptimal convergence rates. If, on the other hand, the polynomial degree of the post-processed velocity is increased, the field cannot be shown to be solenoidal, as .

• Finally, let us remark that here we have used the LDG approach to discretize the viscous terms. However, our results remain valid for any other DG discretization of these terms whose primal bilinear form is both coercive and continuous, such as, e.g., the interior penalty form. For detail, we refer the reader to the discussions in Reference 2 and Reference 19.

5. Proofs

In this section, we provide the proofs of our main results.

5.1. Proof of Proposition 4.2

We begin by proving Proposition 4.2. To do this, note that, if we insert the definition of the upwinding numerical flux into the form , we have

Here, denotes the exterior trace of taken over the edge under consideration and set to zero on the boundary. If we perform now a simple integration by parts, we get

This implies that

where

To bound the term , we recall the following embedding result proved in Reference 13, Proposition 4.5 for smooth and convex domains and in Reference 10, Lemma 6.2 for general polygons:

with a constant independent of the mesh-size. (We point out that the broken -norm used in Reference 13 is slightly different than the one we use here. However, a careful inspection of the proof of Proposition 4.5 therein shows that, in fact the result holds for our definition of .) It is then clear that we can use Hölder’s inequality to obtain

It remains to estimate the term . Using the Lipschitz continuity of the function , we get

and, proceeding as in the proof of Reference 13, Proposition 4.5, we obtain

with a constant independent of the mesh-size. This completes the proof of Proposition 4.2.

5.2. Proof of Proposition 4.6

We prove Proposition 4.6 by first establishing local stability bounds over patches of elements and then by summing up these local results.

Let be fixed. We proceed in several steps.

Step 1: Local bounds in the interior. Let be an interior edge shared by two elements and . We wish to establish a local stability bound over the patch formed by and . Namely, by defining the local seminorm

we claim that

with a constant independent of the mesh-size.

To prove Equation 5.1, it is enough to consider the case where and form a reference patch of unit size. The general case then follows from a scaling argument and the shape-regularity assumption on the mesh, by mapping onto the reference patch using elementwise Piola transforms.

By the triangle inequality, we have

It remains to bound . To do so, we note that, for any element , the restriction belongs to and is uniquely defined by the moments

Hence, the restriction of to the patch belongs to the space

Furthermore, it can be easily seen that the mappings and , given by

define norms on . By the equivalence of all norms on a finite dimensional space, there holds

with a constant only depending on the the polynomial degree. Thus, we obtain that

On the other hand, since, for ,

we conclude that

This estimate and the inequality in Equation 5.2 prove the local stability bound in Equation 5.1.

Step 2: Local bounds on the boundary. The analogous stability result holds on the boundary. Let be an element on the boundary and a boundary edge. By setting

there exists a constant independent of the mesh-size such that

Step 3: Summing up the local bounds. We complete the proof of Proposition 4.6 by summing up the local stability estimates established in Equation 5.1 and Equation 5.4. To this end, we first note that on interior edges and on boundary edges. Here, we write to denote the exterior trace of  over the edge under consideration. Therefore, we have for any edge

Using the local bounds in Equation 5.1 and Equation 5.4 and the above estimate for , we obtain

with constants independent of the mesh-size. This completes the proof of Proposition 4.6.

5.3. Proof of Theorem 4.7

To prove Theorem 4.7, we proceed as follows. First, we eliminate the pressure from the problem by restricting ourselves to the weakly divergence-free subspace of ,

The approximate velocity is thus characterized as the only function such that

Then, we construct a contractive mapping defined on a ball of whose only fixed point is precisely the above velocity. The properties for the corresponding pressure then follow from its characterization

and from the - condition for the incompressibility form .

We proceed in several steps.

Step 1: The operator . We begin by introducing the operator . For , denotes the solution of the following problem. Find such that

Note that since we have, by Proposition 3.2, that . As a consequence, this problem is uniquely solvable.

Furthermore, by the coercivity of the form and in Proposition 4.3,

By the Poincaré inequality in Equation 4.3, we obtain

Hence, the solution to the above problem satisfies

This implies that maps into , where

Step 2: The operator is a contraction. Next, we show that is a contraction on under the smallness condition Equation 4.6. To do so, let be in , and set , . Then

Since

for any , taking , we get

By the coercivity property of the form in Proposition 4.3,

Moreover, by the continuity property of in Proposition 4.2, the bound Equation 5.7 and the continuity of the post-processing operator in Proposition 4.6,

This implies that

and so, if , that is, if the smallness condition Equation 4.6 is satisfied, the mapping is a contraction. Hence, has a unique fixed point , which is the solution to the problem Equation 5.6.

Step 3: Recovering the pressure. Now that the velocity has been computed, the pressure is the solution of

Due to Proposition 4.1, Proposition 4.2, and the Poincaré inequality in Equation 4.3, the right-hand side defines a continuous linear functional on . The inf-sup condition in Proposition 4.4 then guarantees the existence of a unique solution to the above problem. It can then be easily seen that the tuple is the unique solution to the LDG method in Equation 3.11 and Equation 3.12 with .

Step 4: The stability bounds. Next, let us show the stability bounds for in Theorem 4.7. The bound for in Equation 4.7 follows in a straightforward way since . To obtain the bound for the upwind term in Equation 4.9, note that

Similarly to the previous arguments, here we have used the coercivity of , equation Equation 5.6 with test function , and the Poincaré inequality Equation 4.3. Bringing the term to the left-hand side and observing the coercivity of give the stability bound in Equation 4.9.

Moreover, using the inf-sup condition in Proposition 4.4, the Poincaré inequality in Equation 4.3, the continuity properties in Proposition 4.1 and Proposition 4.2, and the stability of in Proposition 4.6, we have from Equation 5.8

Taking into account the stability bound for and assumption Equation 4.6 gives

This gives the desired bound Equation 4.8 for .

Step 5: The convergence estimates. It remains to prove the error estimates for the sequence . Let us begin with that of the velocity. From Step 3, we have that . As a consequence, since is a contraction with Lipschitz constant , we immediately get

The result now follows from the fact that, by the stability bound Equation 5.7,

for .

To obtain the estimate for the pressure, we proceed as follows. First, we note that, from Step 4, we have

This implies that

We insert this expression in the inf-sup condition for , use the stability properties of , , and , take into account the bounds for , , and the contraction property of to obtain

The desired bound for then follows from the bound for .

This completes the proof of Theorem 4.7.

5.4. Proof of Theorem 4.8

Here, we derive the error estimates in Theorem 4.8. To do that, we modify the approach used in Reference 8 to get error estimates for the LDG method for the Oseen problem in two ways. First, we use the nonconforming approach introduced in Reference 16 and later used in Reference 19, and consider the expression

where

The second modification is, of course, due to the presence of the convective nonlinearity.

Proof.

We proceed in several steps.

Step 1: The abstract estimate for . Let us begin with the estimate for the error . We claim that we have

where is given by Equation 4.13.

To prove this result, we proceed as in the error analysis of standard mixed methods (see, e.g., Reference 5) and consider first an element where is the kernel in Equation 5.5. We then have the trivial inequality

Next, we obtain an estimate of . By the coercivity property of the form in Proposition 4.3, we have

and since

for any , for , we get that

Hence,

The terms , , and can be easily estimated as follows. First, by the continuity of the form ,

Then, by the definition of and ,

Finally, since , we have

for any . From Proposition 4.1, it follows that

It remains to estimate the term . To do that, consider the identity

Note that due to Propositions 4.2 and 4.6, the stability bound for in Equation 1.5, and the definitions of the parameters in Equation 4.10, we have

by the smallness condition Equation 4.11.

Next, by the coercivity property of in Proposition 4.3,

and, by Propositions 4.2 and 4.6, and the bound for in Theorem 4.7,

by the smallness condition Equation 4.11.

Finally, by the Lipschitz property of the form in Proposition 4.2, we have

by the bound for in Equation 1.5 and the smallness condition Equation 4.11. Now, take an arbitrary function in . Since reproduces functions in , we have , and so

by Proposition 4.6. This implies that

Thus, gathering all the estimates above and inserting them in the right-hand side of inequality Equation 5.11, and bringing to the left-hand side, we obtain

Inserting this estimate in the right-hand side of inequality Equation 5.10, we get

for any , , and .

It remains to replace by an arbitrary function in . To this end, fix and consider the problem: Find such that

The inf-sup condition in Proposition 4.4 guarantees that such a solution exists. Furthermore, it can be easily seen that we have

in view of the inf-sup condition for and the continuity of the form . By construction and since for any , we further have that . Inserting in Equation 5.12, employing the triangle inequality, and taking into account the above bound for yield the abstract error estimate Equation 5.9 for the velocity.

Step 2: The abstract estimate for . As a consequence of the approximation result Equation 5.9, we obtain the estimate of the error between and its globally solenoidal approximation 

To see this, note that

where is any element of . Here we have used the stability bound in Proposition 4.6 and the fact that reproduces polynomials in . This shows the inequality Equation 5.13.

Step 3: The abstract estimate for the pressure. Now, let us obtain the estimate for the pressure. We claim that the error in the pressure satisfies

where is given by Equation 4.15.

To see this, we proceed in a way similar to the one used to deal with the velocity. Thus, we begin by noting that for

where we have used the inf-sup condition in Proposition 4.4. Therefore,

by the continuity of the form .

To bound the first term on the right-hand side of Equation 5.15, we note that

for any , and proceed as in the previous step to obtain

and since

from Step 2, we get

Inserting this inequality in Equation 5.15 and using the bound Equation 5.9 for from Step 1, we immediately obtain the abstract estimate Equation 5.14 for the pressure.

Step 4: The abstract estimate for the velocity gradient. To derive the error estimate for , we note that from Equation 3.14 we have

Hence,

Using the stability bound Equation 4.4 for the lifting operator yields

The last inequality follows as the jumps of the exact solution vanish. This shows that

Step 5: Approximation estimates. Under the regularity assumption Equation 4.12, the standard approximation property

holds. Moreover, from the results in Reference 5 (see also Reference 11, Section 3) we have

Finally, we have that

with a constant independent of the mesh-size.

To see the estimate of the residual, we proceed as follows. For , it is easy to see that is given by

with and denoting the -projections onto and , respectively. The desired estimate then follows by proceeding as the proof of Reference 19, Proposition 8.1 and using standard approximation results for and .

Step 6: Conclusion. It is now a simple matter to see that the error estimates of Theorem 4.8 follow by inserting the approximation estimates obtained in the previous step into the abstract bounds for the velocity Equation 5.9, for its globally solenoidal post-processing Equation 5.13, the pressure Equation 5.14, and the velocity gradient Equation 5.16. This completes the proof of Theorem 4.8.

6. Numerical results

In this section, we present numerical experiments that show that the theoretical rates of convergence are sharp. We also display the behavior of the iterative method as a function of the Reynolds number.

As a reference solution in our tests, we take the analytical solution  of the incompressible Navier-Stokes equations that was obtained by Kovasznay in Reference 14. For a given viscosity , this solution is given by

where

It solves Equation 1.1 with a suitably chosen right-hand side . Here, the constant is such that . Further we use the values of the exact solution to prescribe inhomogeneous Dirichlet boundary data for the velocity on the whole boundary of the computational domain, which we take to be . Note that in this case, the numerical fluxes (on edges lying on the boundary) must be modified as

We consider square meshes that are generated by refining the single grid cell uniformly. Therefore, a mesh on “level” consists of by cells. All computations are performed with bilinear shape functions for and  and piecewise constants for , according to the remarks in subsection 4.3. The stabilization function is chosen as in Equation 4.1 with . Our analysis then predicts first-order convergence for in the norm and for the pressure in the -norm.

In Table 1 we show the errors and convergence rates in , and obtained for . The errors in and are measured in the -norm while and are evaluated in the norm . We observe the predicted first-order convergence for all the error components, in full agreement with the results of Theorem 4.8. Note that we have scaled the -error in by so that this error can be directly compared to the -errors in and . These three errors are all of the same magnitude, with a slight advantage for the post-processed solution.

In Table 2 we show the seminorm of the errors and which measures their jumps. Note that it superconverges with order . This means that the relative contribution of this seminorm to the -norm diminishes as decreases. An analysis of this phenomenon remains to be carried out.

In Table 3 we show the -errors in the velocities and their corresponding convergence orders. In the first column we observe that the velocities converge with second-order accuracy. In the second column, we notice that by post-processing the error is reduced by a factor of roughly 3/2. Therefore, the post-processed solution should be used as the best approximation obtained by our scheme. Furthermore, we show the -norms of the divergence of (evaluated at the points of a 4-by-4 Gauss formula on each cell). These are of the order of the residual of the nonlinear iteration, confirming that the post-processed solution is indeed divergence free.

The convergence of the nonlinear iteration under consideration is illustrated in Table 4. The number of steps required to reduce the start residual by a factor of is displayed. The initial guess for the iterations is the vector .

The linear system in each step is solved by a preconditioned GMRES method up to a relative accuracy of , using the preconditioners described in Reference 12. Therefore, the error of the linear iterations is small enough to be neglected. Table 4 shows that the number of iteration steps is not only bounded independently of the mesh-size, but in fact decreasing. This is in perfect agreement with our theoretical results in Theorem 4.7. If we decrease the viscosity, the increase of the number of iteration steps for convergence is quite moderate on fine grids. Of course, this only holds as long as there is convergence. With , the nonlinear iteration does not converge anymore, probably because the stationary solution is not stable in this case.

7. Concluding remarks

In this paper, we described and analyzed a new LDG method for the approximation of the two-dimensional incompressible Navier-Stokes equations on triangular meshes. The approximation is based on discontinuous elements for the approximation of the velocity, the velocity gradient, and the pressure, respectively, and on a post-processing procedure. Alternatively, the approximation of the velocity field can be based on div-conforming BDM elements for which this procedure can be omitted. These are the only DG methods that are locally conservative and stable for the incompressible Navier-Stokes equations. They are also the only methods that provide a systematic and simple way to obtain an approximate velocity that is exactly divergence free for polynomials of degree greater than or equal to 1; see Reference 20.

As discussed in subsection 4.3, the results of this paper can be extended in a straightforward way to simplicial meshes in three dimensions and to elements on quadrilateral or hexahedral affine meshes. Future work will be devoted to extending the methods to the case in which the numerical flux depends also on the pressure. This happens, for example, when all the unknowns are approximated with the same polynomial space; see Reference 9, Reference 8. Indeed, in this case, the operator depends not only on the velocity but also on the pressure. As a consequence, the analysis of the convergence of the fixed point iteration is much more delicate. It will be carried out in a forthcoming paper.

Figures

Table 1.

Errors and orders of convergence for .

32.2e+01.2e+18.1e+07.0e-0
41.0e+01.125.4e+01.113.2e+01.333.4e-01.05
54.8e-11.102.4e+01.161.4e+01.181.6e-01.07
62.3e-11.041.1e+01.186.8e-11.067.8e-11.04
71.2e-11.014.7e-11.173.4e-11.023.9e-11.02
85.8e-21.002.2e-11.131.7e-11.011.9e-11.02
Table 2.

Errors and orders of convergence for in the seminorm .

39.1e+04.8e+0
44.2e+01.111.5e+01.72
51.8e+01.234.7e-11.64
67.2e-11.321.6e-11.54
72.8e-11.395.6e-21.52
81.0e-11.442.0e-21.51
Table 3.

-errors and orders of convergence in the velocity and -norm of the divergence of the post-processed solution  for .

36.4e-14.9e-11.4e-12
41.6e-12.031.1e-12.221.4e-12
53.3e-22.222.0e-22.373.2e-12
67.1e-32.244.2e-32.271.5e-11
71.6e-32.199.8e-42.121.8e-12
83.5e-42.132.4e-42.042.9e-11
Table 4.

Number of iterations for convergence of the nonlinear iteration.

31433865
41021106
581654
661229
751018
861013

Mathematical Fragments

Equation (1.1)
Equation (1.3)
Equation (1.4)
Equation (1.5)
Equation (1.6)
Equation (1.7)
Equation (3.1)
Equations (3.2), (3.3), (3.4)
Equation (3.8)
Equation (3.9)
Remark 3.1.

Notice that one can take the div-conforming velocity space

while keeping the other spaces and definitions above unchanged.

Proposition 3.2.

We have the following results.

(1)

is well defined and is in the space

(2)

If and , then .

(3)

If satisfies Equation 3.4, then in and .

Equations (3.11), (3.12)
Equation (3.13)
Equation (3.14)
Equation (4.1)
Equation (4.3)
Equations (4.4), (4.5)
Proposition 4.1.

We have that

for continuity constants and that are independent of the mesh-size.

Proposition 4.2.

Let , , and . Then we have

for a continuity constant that is independent of the mesh-size.

Proposition 4.3.

Let be given by Equation 4.1. Then, for any , there exists a constant independent of the mesh-size such that

Furthermore, for and , it holds that

In the integrals over edges in , we denote by any unit normal to the edge under consideration.

Proposition 4.4 (Reference 11).

There exists a constant independent of the mesh-size such that

Proposition 4.6.

Let . Then we have

with a stability constant that is independent of the mesh-size.

Theorem 4.7 (Existence and uniqueness of discrete solutions).

Assume that

Then the LDG method Equation 3.11Equation 3.13 defines a unique solution . It satisfies the bounds

as well as

Moreover, if is the approximate solution given by the LDG method in Equation 3.11Equation 3.12 for the Oseen equations with , , then

for any initial guess .

Equation (4.10)
Theorem 4.8 (Error estimates).

Assume that

and that the exact solution of the Navier-Stokes equations Equation 1.1 satisfies

Then

where only depends on the regularity of the mesh and the polynomial degree , and

Equation (5.1)
Equation (5.2)
Equation (5.4)
Equation (5.5)
Equation (5.6)
Equation (5.7)
Equation (5.8)
Equation (5.9)
Equation (5.10)
Equation (5.11)
Equation (5.12)
Equation (5.13)
Equation (5.14)
Equation (5.15)
Equation (5.16)

References

Reference [1]
D. N. Arnold, An interior penalty finite element method with discontinuous elements, SIAM J. Numer. Anal. 19 (1982), 742–760. MR 0664882 (83f:65173)
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]
S. Brenner, Poincaré-Friedrichs inequalities for piecewise H functions, SIAM J. Numer. Anal. 41 (2003), 306–324. MR 1974504 (2004d:65140)
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 0799685 (87g:65133)
Reference [5]
F. Brezzi and M. Fortin, Mixed and hybrid finite element methods, Springer Verlag, 1991. MR 1115205 (92d:65187)
Reference [6]
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 [7]
B. Cockburn, G. Kanschat, and D. Schötzau, The local discontinuous Galerkin methods for linear incompressible flow: A review, Computers and Fluids (Special Issue: Residual based methods and discontinuous Galerkin schemes), to appear.
Reference [8]
—, Local discontinuous Galerkin methods for the Oseen equations, Math. Comp. 73 (2004), 569–593. MR 2031395
Reference [9]
B. Cockburn, G. Kanschat, D. Schötzau, and C. Schwab, Local discontinuous Galerkin methods for the Stokes system, SIAM J. Numer. Anal. 40 (2002), 319–343. MR 2003g:65141
Reference [10]
V. Girault, B. Rivière, and M. F. Wheeler, A discontinuous Galerkin method with non-overlapping domain decomposition for the Stokes and Navier-Stokes problems, Math. Comp., 74 (2005), 53–84.
Reference [11]
P. Hansbo and M.G. Larson, Discontinuous finite element methods for incompressible and nearly incompressible elasticity by use of Nitsche’s method, Comput. Methods Appl. Mech. Engrg., 191 (2002), 1895–1908. MR 1886000 (2003j:74057)
Reference [12]
G. Kanschat, Block preconditioners for LDG discretizations of linear incompressible flow problems, J. Sci. Comput., 25 (2003), 815–831.
Reference [13]
O. A. Karakashian and W.N. Jureidini, A nonconforming finite element method for the stationary Navier-Stokes equations, SIAM J. Numer. Anal., 35 (1998), 93–120. MR 99d:65320
Reference [14]
L. I. G. Kovasznay, Laminar flow behind a two-dimensional grid, Proc. Camb. Philos. Soc. 44 (1948), 58–62. MR 0024282 (9:476d)
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 0658142 (58:31918)
Reference [16]
I. Perugia and D. Schötzau, An -analysis of the local discontinuous Galerkin method for diffusion problems, J. Sci. Comput. (Special Issue: Proceedings of the Fifth International Conference on Spectral and High Order Methods (ICOSAHOM-01), Uppsala, Sweden) 17 (2002), 561–571. MR 1910752
Reference [17]
A. Quarteroni and A. Valli, Numerical approximation of partial differential equations, Springer, New York, 1994. MR 1299729 (95i:65005)
Reference [18]
W.H. Reed and T.R. Hill, Triangular mesh methods for the neutron transport equation, Tech. Report LA-UR-73-479, Los Alamos Scientific Laboratory, 1973.
Reference [19]
D. Schötzau, C. Schwab, and A. Toselli, hp-DGFEM for incompressible flows, SIAM J. Numer. Anal. 40 (2003), 2171–2194. MR 1974180
Reference [20]
L. R. Scott and M. Vogelius, Norm estimates for a maximal right inverse of the divergence operator in spaces of piecewise polynomials, RAIRO Modél. Math. Anal. Numér. 19 (1985), 111–143. MR 0813691 (87i:65190)
Reference [21]
R. Témam, Sur l’approximation des solutions des équations de Navier-Stokes, C. R. Acad. Sci. Paris Sér. A 216 (1966), 219–221. MR 0211059 (35:1941)
Reference [22]
—, Une méthode d’approximation de la solutions des équations de Navier-Stokes, Bull. Soc. Math. France 98 (1968), 115–152. MR 0237972 (38:6249)
Reference [23]
A. Toselli, hp-discontinuous Galerkin approximations for the Stokes problem, Math. Models Methods Appl. Sci. 12 (2002), 1565–1616. MR 1938957 (2003m:65211)

Article Information

MSC 2000
Primary: 65N30 (Finite elements, Rayleigh-Ritz and Galerkin methods, finite methods)
Keywords
  • Finite element methods
  • discontinuous Galerkin methods
  • incompressible Navier-Stokes equations
Author Information
Bernardo Cockburn
School of Mathematics, University of Minnesota, Vincent Hall, Minneapolis, Minnesota 55455
cockburn@math.umn.edu
Guido Kanschat
Institut für Angewandte Mathematik, Universität Heidelberg, Im Neuenheimer Feld 293/294, 69120 Heidelberg, Germany
kanschat@dgfem.org
MathSciNet
Dominik Schötzau
Mathematics Department, University of British Columbia, 1984 Mathematics Road, Vancouver, British Columbia V6T 1Z2, Canada
schoetzau@math.ubc.ca
Additional Notes

The first author was supported in part by the National Science Foundation (Grant DMS-0107609) and by the University of Minnesota Supercomputing Institute.

This work was carried out in part while the authors were at the Mathematisches Forschungsinstitut Oberwolfach for the meeting on Discontinuous Galerkin Methods in April 21–27, 2002 and while the second and third authors visited the School of Mathematics, University of Minnesota, in September 2002.

Journal Information
Mathematics of Computation, Volume 74, Issue 251, 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 2004 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/S0025-5718-04-01718-1
  • MathSciNet Review: 2136994
  • Show rawAMSref \bib{2136994}{article}{ author={Cockburn, Bernardo}, author={Kanschat, Guido}, author={Sch\"otzau, Dominik}, title={A locally conservative LDG method for the incompressible Navier-Stokes equations}, journal={Math. Comp.}, volume={74}, number={251}, date={2005-07}, pages={1067-1095}, issn={0025-5718}, review={2136994}, doi={10.1090/S0025-5718-04-01718-1}, }

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.