• Aucun résultat trouvé

On the time growth of the error of the DG method for advective problems

N/A
N/A
Protected

Academic year: 2022

Partager "On the time growth of the error of the DG method for advective problems"

Copied!
26
0
0

Texte intégral

(1)

doi:10.1093/imanum/dry013

On the time growth of the error of the DG method for advective problems

Václav Kuˇcera

Faculty of Mathematics and Physics, Charles University, Sokolovská 83, Praha 8, 18675, Czech Republic

Corresponding author: vaclav.kucera@email.cz and

Chi-Wang Shu

Division of Applied Mathematics, Brown University, Providence, RI 02912, USA shu@dam.brown.edu

[Received on 16 March 2017; revised on 15 January 2018]

In this paper we derive a priori L(L2) and L2(L2) error estimates for a linear advection–reaction equation with inlet and outlet boundary conditions. The goal is to derive error estimates for the discontinuous Galerkin method that do not blow up exponentially with respect to time, unlike the usual case when Gronwall’s inequality is used. While this is possible in special cases, such as divergence-free advection fields, we take a more general approach using exponential scaling of the exact and discrete solutions. Here we use a special scaling function, which corresponds to time taken along individual pathlines of the flow. For advection fields, where the time that massless particles carried by the flow spend inside the spatial domain is uniformly bounded from above by someT, we deriveO(hp+1/2) error estimates where the constant factor depends only onT, but not on the final time T. This can be interpreted as applying Gronwall’s inequality in the error analysis along individual pathlines (Lagrangian setting), instead of physical time (Eulerian setting).

Keywords: discontinuous Galerkin method; advection–reaction problem; a priori error estimates; time growth of error; exponential scaling.

1. Introduction

The discontinuous Galerkin (DG) finite element method introduced in the study byReed & Hill (1973) is an increasingly popular method for the numerical solution of partial differential equations. The DG method was first formulated for a neutron transport equation and such problems remain the major focus of the DG community. It is for problems of an advective or convective nature that the DG method is suited best and shows its strengths compared to other numerical methods for such problems.

In this paper, we shall consider a scalar time-dependent linear advection–reaction equation of the form

∂u

∂t +a· ∇u+cu=0. (1.1)

We will discretize the problem in space using the DG method with the upwind numerical flux on unstructured simplicial meshes inRd which may contain hanging nodes. The first analysis of the DG

© The authors 2018. Published by Oxford University Press on behalf of the Institute of Mathematics and its Applications. All rights reserved.

(2)

method for a stationary advection–reaction problem with constant coefficients was made in the study by Lesaint & Raviart (1974), later improved in the study byJohnson & Pitkäranta (1986).

In this paper we derive a priori estimates of the error eh=uuh, where uhis the DG solution.

The goal is to derive estimates of ehin the L L2

- and L2 L2

-norms of the order Chp+1/2, where the constant C does not depend exponentially on the final time T→ +∞. Such results already exist in the literature; however, they are derived under the ellipticity condition

c12div aγ0>0 (1.2)

for some constantγ0. In the paper byFeistauer & Švadlenka (2004),γ0 =0 is also admissible, which corresponds to the interesting case of a divergence-free advection field a. Without assuming (1.2), when one proceeds straightforwardly, at some point Gronwall’s inequality must be used in the proofs. This, however, results in exponential growth of the constant factor C in the error estimate with respect to time.

Here, we will circumvent condition (1.2) by considering an exponential scaling transformation u(x, t)=eμ(x,t)u(x, t),˜ uh(x, t)=eμ(x,t)u˜h(x, t) (1.3) of the exact and discrete solutions. Substituting (1.3) into (1.1) and its discretization results in a new equation foru with a modified reaction term˜ c˜=∂μ∂t +a· ∇μ+c. The functionμcan then be chosen in various ways in order to satisfy the ellipticity condition (1.2) for the new equation. After performing the error analysis, the resulting error estimates fore˜h= ˜u− ˜uhcan be transformed back via (1.3) to obtain estimates for eh.

The use of transformations similar to (1.3) is not new. The trick known as exponential scaling, corresponding to takingμ(x, t)=αt for some sufficiently large constantα, is very well known from the partial differential equation and numerical analysis communities. However, the application of such a scaling transformation inherently leads to exponential dependence of the error on t after transforming frome˜hback to eh. Other choices ofμare possible, e.g.μ(x, t)=μ0·x for some constant vectorμ0. This choice was used in the stationary case in the studies byNävert (1982)andJohnson & Pitkäranta (1986). Takingμ(x) independent of t corresponds to the analysis of the stationary case from the study byAyuso & Marini (2009). In the nonstationary case the use of this stationary transform would lead to too restrictive assumptions on the advection field.

In this paper we construct the function μ from (1.3) using characteristics of the advection field.

Namely,μwill be proportional to time along individual pathlines of the flow. If pathlines exist only for a finite time bounded by someT for each particle entering and leaving the spatial domainΩ, we obtain error estimates exponential inT and not the final physical time T. This is a result we would expect if we applied Gronwall’s lemma in the Lagrangian framework along individual pathlines (which exist only for a finite time uniformly bounded byT) instead of the usual application of Gronwall with respect to physical time in the Eulerian framework. The analysis can be carried out under mild assumptions that a(·,·) satisfies the assumptions of the Picard–Lindelöf theorem and that there are no characteristic boundary points on the inlet (i.e. a·n is uniformly bounded away from zero on the inlet).

The paper is organized as follows. After introducing the continuous problem in Section2, we discuss the exponential scaling transform, its variants and application in the weak formulation in Section3. In Section4 we construct the scaling functionμand prove results on its regularity and other properties needed in the analysis. In Section5 we introduce the DG formulation and its basic properties. In Sections6and7we analyse the DG advection and reaction forms and estimate the error of the method.

(3)

In order to focus on the main ideas of the analysis, we postpone some of the technical estimates of the DG forms to the Appendix.

We use (·,·) to denote the L2(Ω) scalar product and·for the L2(Ω)-norm. To simplify the notation, we shall drop the argumentΩin Sobolev norms, e.g. · Hp+1 denotes the Hp+1(Ω)-norm. We will also denote the Bochner norms over the whole considered interval (0, T) in concise form, e.g.uL2(Hp+1) denotes the L2(0, T; Hp+1(Ω))-norm of u.

2. Continuous problem

LetΩ ⊂ Rd, d ∈ Nbe a bounded polygonal (polyhedral) domain with Lipschitz boundary∂Ω. Let 0 < T +∞and let QT =Ω ×(0, T) be the space–time domain. We note that we admit the time interval to be infinite in the special case of T= +∞.

We seek u : QT →Rsuch that

∂u

∂t +a· ∇u+cu=0 in QT, (2.1)

u=uD on∂Ω×(0, T), u(x, 0)=u0(x), xΩ.

Here a : QT →Rdand c : QT →Rare the given advective field and reaction coefficient, respectively.

By∂Ωwe denote the inflow part of the boundary, i.e.

∂Ω= {x∈∂Ω; a(x, t)·n(x) <0 ∀t(0, T)},

where n(x) is the unit outer normal to∂Ω at x. For simplicity we assume that the inflow boundary is independent of t, in other words we assume that a(x, t)·n(x)<0 for all t and all x∂Ω, and a(x, t) · n(x) is non-negative on∂Ω \ ∂Ω for all t. In general ∂Ω can depend on t due to the nonstationary nature of a. However, we do not consider this case since it would only make the notation more complicated without changing the key points of the analysis itself.

We assume that the reaction coefficient satisfies cC([0, T); L(Ω))L(QT). Furthermore, let aC([0, T); W1,(Ω))with a,∇a uniformly bounded a.e. in QT. In other words, a (·,·) satisfies the assumptions of the Picard–Lindelöf theorem: continuity with respect to x, t and uniform Lipschitz continuity with respect to x. The spatial Lipschitz constant of a will be denoted by La.

The DG method for the stationary version of problem (2.1) was analysed in the studies byNävert (1982),Johnson & Pitkäranta (1986)andAyuso & Marini (2009). In this case the well-posedness of the problem is guaranteed by the assumption that the flow field possesses neither closed curves nor stationary points, cf. the study by Devinatz et al. (1974). For the nonstationary problem (2.1), the situation is simpler. Here, the problem can be solved using the method of characteristics, similarly to the method we use in Section4, cf.Evans (2010). Along each characteristic curve (pathline), problem (2.1) reduces to a simple, ordinary differential equation, which can be solved in closed form. The result is exponential growth or decay of u along the pathline with the rate given by the reaction coefficient c. As the flow field a satisfies the assumptions of the Picard–Lindelöf theorem, we have existence and uniqueness of the pathlines. Since the Dirichlet boundary condition is prescribed only at the inflow boundary, this gives the well-posedness of problem (2.1). We note that also a ‘parabolic’ approach to

(4)

(2.1) is possible to obtain ‘energy’ estimates of the time growth of the L2(Ω)-norm of the solution. This is done by Gronwall’s lemma after testing the weak form of (2.1) with the solution itself and applying Green’s theorem in the advection terms; cf. e.g.Mizohata (1973)andLarsson & Thomée (2003). The result is an upper bound on the time growth of the solution that is exponential in the quantity c12diva if this quantity is non-negative; compare with the ellipticity condition (3.1). We note that from the theoretical and practical points of view, the case of systems of equations of the considered type is much more interesting; cf. e.g.Benzoni-Gavage & Serre (2007).

3. Exponential scaling

The standard assumption on the coefficients a, c found throughout the numerical literature is

c12div aγ0>0 on QT (3.1)

for some constantγ0. This assumption comes from the requirement that the weak formulations of the advection and reaction terms give an elliptic bilinear form on the corresponding function space. In the study byFeistauer & Švadlenka (2004), this assumption is avoided by using exponential scaling in time, i.e. the transformation u(x, t)=eαtw(x, t) forα∈ R. Substitution into (2.1) and division of the whole equation by the common positive factor eαt leads to a modified reaction coefficient˜c= c+αin the new equation for w. By choosing the constantα sufficiently large, (3.1) will be satisfied for the new equation. In the study byFeistauer & Švadlenka (2004), error estimates that grow linearly in time are then derived for the DG scheme. However, forα >0, if one transforms the resulting estimates back to the original problem using the exponential scaling transformation, the result is an estimate that depends exponentially on T.

Another possibility to avoid assumption (3.1) is a transformation similar to exponential scaling, but with respect to the spatial variable; cf. the studies byNävert (1982),Johnson & Pitkäranta (1986)and Roos et al. (2008): letμ0∈Rd, then we write

u(x, t)=eμ0·xu(x, t).˜ (3.2)

If we assume that u is sufficiently regular, as we shall in this paper, by substituting (3.2) into (2.1) and dividing the whole equation by the strictly positive function eμ0·x, we get the new problem

∂u˜

∂t +a· ∇ ˜u+(a·μ0+c)u˜=0. (3.3) The condition corresponding to (3.1) now reads as follows: there existsμ0∈Rdsuch that

a·μ0+c12div aγ0>0 on QT. (3.4) A possible generalization of the exponential scaling transformation (3.2) is taking a sufficiently smooth functionμ:Ω→Rand setting

u(x, t)=eμ(x)u(x, t).˜ (3.5)

(5)

Again, substituting into (2.1) and dividing by eμ(x), we get the new problem

∂u˜

∂t +a· ∇ ˜u+(a· ∇μ+c)u˜=0.

The condition corresponding to (3.1) and (3.4) now reads as follows: there existsμ:Ω →Rsuch that a· ∇μ+c12div aγ0>0 on QT. (3.6) This is essentially the approach used in the study byAyuso & Marini (2009)for a stationary advection–

diffusion–reaction problem. As shown in the study byDevinatz et al. (1974), the existence of a function μ:Ω →Rsuch that a· ∇μγ0is equivalent to the property that the advective field a possesses neither closed curves nor stationary points. The uniformly positive term a· ∇μcan then be used to dominate the possibly negative term c12diva in order to satisfy condition (3.6). If we used the choice (3.5) in our analysis, we would need to assume the nonexistence of closed curves or stationary points of the flow field for all t. This assumption is too restrictive. Furthermore, in the study byAyuso & Marini (2009), rather high regularity ofμis needed in the analysis, namelyμWp+1,(Ω), where p is the polynomial degree of the DG scheme.

In this paper, we shall generalize the transformation (3.5) using a functionμ: QT →Rand set

u(x, t)=eμ(x,t)u(x, t).˜ (3.7)

In our analysis we will need only minimum regularity of the scaling functionμ, in contrast to the study byAyuso & Marini (2009). Namely, we assume thatμis Lipschitz continuous both in space and time.

This implies that the derivativesμtand∇μexist for almost all x and t. Hence, the following steps are valid for almost all x and t and particularly, they will hold in the integral (i.e. weak) sense. We note that the Lipschitz continuity in space and time will hold for the specific construction ofμfrom Section4, as shown in Section4.1.

As in the previous cases, substituting (3.7) into (2.1) and dividing by eμ(x,t)gives the new problem

∂u˜

∂t +a· ∇˜u+ ∂μ

∂t +a· ∇μ+c

˜

u=0. (3.8)

The condition corresponding to (3.1), (3.4) and (3.6) now reads as follows: there existsμ : QT → R such that

∂μ

∂t +a· ∇μ+c12div aγ0>0 a.e. in QT. (3.9) This is the condition we will assume throughout the paper. Using the transformation (3.7) instead of (3.5) gives sub-exponential growth or even uniform boundedness of the constant factor in the error estimate with respect to time.

If the original problem (2.1) does not satisfy (3.1), one is tempted to numerically solve the trans- formed equation (3.8) instead, using DG or any other method. This is done e.g. inRoos et al. (2008) for the case of linearμ, i.e. (3.2). However, then we are numerically solving a different problem and obtain different results, as the DG solutions of (2.1) and (3.8) are not related by the simple relation (3.7), unlike the exact solutions. However, as we will show in this paper, one can analyse the DG method

(6)

for the original problem (2.1), while taking advantage of the weaker ellipticity condition (3.9) for the transformed problem (3.8).

Since the DG scheme is based on a suitable weak formulation, the first step is to reformulate the transformation (3.7) within the weak, rather than strong formulation of (2.1). The key step in deriving (3.8) from (2.1) is dividing the whole equation by the common factor eμ(x,t). However, if we substitute (3.7) into the weak formulation

Ω

∂u

∂tv+a· ∇uv+cuv dx=0, (3.10)

it is not easy to divide the equation by eμ(x,t), since this is inside the integral. The solution is to test (3.10) by the new test functionˆv(x, t)=eμ(x,t)v(x, t). Due to the opposite signs in the exponents, the exponential factors cancel each other and we obtain the weak formulation of (3.8):

Ω

∂u˜

∂tv+a· ∇ ˜uv+ ∂μ

∂t +a· ∇μ+c

˜

uv dx=0.

We note that the transformations u→ ˜u and v→ ˆv are bijections. Furthermore, since we assume thatμ is Lipschitz continuous in space, then the factor eμ(·,t)lies in W1,(Ω)and the transformation (3.7) is a bijection from the function space V into itself, where

V =

uH1(Ω); u|∂Ω =0

is the space for weak solutions and test functions for (2.1) we consider. This is the case uD =0, the general case can be treated by the standard Dirichlet lifting procedure. We note that in the general case, the appropriate function spaces for the solution of (2.1) would be e.g. L1(Ω) or BV(Ω) with respect to space. However, since we assume smooth solutions and the linear case, it is possible to consider the Hilbert setting of H1(Ω) which makes the analysis easier in the finite element or DG framework.

In the DG method, the above procedure using transform (3.7) cannot be directly applied, since if vSh—the discrete space of discontinuous piecewise polynomial functions—thenvˆ = eμv will no longer lie in Shand therefore cannot be used as a test function in the formulation of the method. The solution is to test with a suitable projection ofv onto Sˆ hand estimate the difference, as we shall do in Section6.

4. Construction of the functionμ

In this section we show a construction of the functionμsatisfying (3.9). Many different constructions ofμare possible depending on the assumptions on the vector field a. For example, we can always take μ(x, t)=αt for someα0. This corresponds to standard exponential scaling and, as we have seen, this choice leads to exponential growth in time of the error estimate. Another possibility was the approach ofAyuso & Marini (2009)mentioned in Section2, where a suitable functionμ(x) exists if a, which is stationary, possesses no closed curves or stationary points. Here we show another possibility with an interesting interpretation.

(7)

If c−12diva is negative or changes sign frequently, we can use the expressionμt+ a·∇μto dominate this term everywhere. If we chooseμ1such that

∂μ1

∂t +a· ∇μ1=1 on QT, (4.1)

then by multiplyingμ1by a sufficiently large constant, we can satisfy the ellipticity condition (3.9) for a chosenγ0>0.

Equation (4.1) can be explicitly solved using characteristics. We define pathlines of the flow, i.e. the family of curves S(t; x0, t0)by

S(t0; x0, t0)=x0Ω, dS(t; x0, t0)

dt =a(S(t; x0, t0), t). (4.2) This means that S(·; t0, x0) is the trajectory of a massless particle in the nonstationary flow field a passing through point x0at time t0. It is convenient to choose the parameter t0minimal for each pathline—then the pair (x0, t0) is the ‘origin’ of the pathline. In other words, for each x0Ω, there is a pathline S(·; x0, 0) corresponding to trajectories of particles present inΩat the initial time t0=0. Then there are trajectories of particles entering through the inlet part of∂Ω: for each x0∂Ωand all t0there exists a pathline S(·; x0, t0) originating at (x0, t0)∈∂Ω×[0, T).

Equation (4.1) can then be rewritten along the pathlines as dμ1(S(t; x0, t0), t)

dt =

∂μ1

∂t +a· ∇μ1

(S(t; x0, t0), t)=1, therefore

μ1(S(t; x0, t0), t)=tt0. (4.3) Here any constant can be chosen instead of t0; however, this choice is the most convenient. With this choice, we have at the origin of a pathline

μ1(S(t0; x0, t0), t0)=0

and the value ofμ1along this pathline is simply the time elapsed since t0. In other words,μ1(x, t) is the time a particle carried by the flow passing through xΩat time t has spent inΩup to the time t. In this paper we assume this quantity to be uniformly bounded in order forμ1andμto be uniformly bounded;

cf. assumption (4.7). It is important to note that t0depends on the specific trajectory considered and that t0will in general be different for each (x, t). Therefore, even though t and t0may be arbitrarily large if we consider an infinite time interval, the difference tt0=μ1, which measures the time elapsed since the entry of the corresponding particle intoΩ, will remain bounded under this assumption.

Now that we have constructed a function satisfying (4.1), we can chooseγ0and define e.g.

μ(x, t)=μ1(x, t) inf

QT

c12div a

+γ0

, (4.4)

(8)

where f=min{0, f}is the negative part of f . Then

∂μ

∂t +a· ∇μ+c12div a= inf

QT

c12div a

+γ0+c12div aγ0.

In the main result, of this paper, Theorem7.1and Corollary7.2, this choice ofμleads to estimates of the following form for the DG error eh(cf. (7.1) and (7.10)):

ehL(L2)+√

γ0ehL2(L2)CeT(|infQT(c12div a)|+γ0)hp+1/2, (4.5) whereT=supQTμ1, i.e.T is the maximal particle ‘lifetime’, i.e. the maximal time any particle carried by the flow field a spends inΩ, since by (4.3),μ1(x, t) is the time the particle at (x, t) carried by the flow has spent inΩ. If we compare this to estimates obtained by a straightforward analysis using Gronwall’s lemma without any ellipticity assumption, we would expect

ehL(L2)CeT(supQT|c12div a|)hp+1/2. (4.6) Comparing (4.6) to (4.5), we see that the estimates are essentially of similar form, only the exponential dependence on global physical time T has been replaced by dependence on timeT along pathlines, which is bounded. Effectively, our analysis replaces the application of Gronwall’s lemma in the Eulerian framework with its application in the Lagrangian framework—along individual pathlines which have bounded length.

Throughout the paper, we assume the particle lifetimeT to be bounded. In general, we could consider dependencies of the form

T(t)= sup

(x,ϑ)Ω×(0,t)

μ1(x,ϑ),

i.e.T(t)is the maximal time any particle carried by the flow a spends inΩ up to time t. From (4.5), we can expect exponential dependence of the L(0, t; L2)and L2(0, t; L2)norms onT(t). The result of the standard analysis (4.6) corresponds to the ‘worst case’T(t)=t, i.e. that there is a particle that stays insideΩ for all t[0, T). However, considering more general dependencies on time is possible, e.g.

T=√

t, leading to growth of the error that is exponential int.

4.1 Regularity of the functionμ

In our analysis of the DG scheme we will assume thatμsatisfies 0μ(x, t)μmax,

|μ(x, t)μ(y, t)|Lμ|x−y|, (4.7) for all x, yΩ and t(0, T). In other words,μis non-negative, uniformly bounded and Lipschitz continuous in space, where the Lipschitz constant is uniformly bounded for all t. SinceΩis a Lipschitz domain, hence quasiconvex, this means thatμ(t)W1,(Ω)for all t, with its W1,(Ω)seminorm uniformly bounded by Lμfor all t.

(9)

Fig. 1. Left: proof of Lemma4.1. Right: vortex touching∂Ωwith t0 ˜t0=0, i.e. t0− ˜t0is large, henceμ1is not Lipschitz continuous at x.

Now we show whenμ1defined by (4.3) satisfies conditions (4.7), especially Lipschitz continuity in space which is necessary in the analysis. We note that sinceμis obtained fromμ1by multiplication by a constant, properties derived forμ1imply those forμ.

An obvious example of whenμ1defined by (4.3) will not be Lipschitz continuous is when a vortex

‘touches’∂Ωas in Fig.1. Then if (x, t) lies on a pathline originating at (x0, t0) such that a(x0, t0) is tangent to∂Ω, we can findx arbitrarily close to x such that the corresponding pathline is much longer,˜ perhaps winding several times around the vortex. Thenμ1(x, t)μ1(x, t)˜ = ˜t0t0can be very large, whilex− ˜x is arbitrarily small, hence μ1 is not Lipschitz continuous. In factμ1 is discontinuous at (x, t). In the following lemma, we show that inlet points where a is tangent to∂Ω are the only troublemakers.

We note that sinceΩis a bounded Lipschitz domain (hence quasiconvex), when proving Lipschitz continuity ofμ1in space it is sufficient to prove local Lipschitz continuity ofμ1in some neighborhood of each xΩwith a Lipschitz constant independent of x.

Lemma 4.1 Let aL(QT)be continuous with respect to time and Lipschitz continuous with respect to space. Let there exist a constant amin>0 such that

−a(x, t)·namin (4.8)

for all x∂Ω, t[0, T). Letμ1be defined by (4.3) onΩ×[0, T). Let the time any particle carried by the flow field a(·,·) spends inΩbe uniformly bounded byT. Thenμ1satisfies assumption (4.7).

Proof . By definition, we haveμ1(x, t)0 for all x, t. By the above considerations,μ1is bounded by the maximal time particles spend inΩ, which is uniformly bounded. This impliesμ1(x, t)T =μmax

for someμmax. Now we will prove Lipschitz continuity.

Let t(0, T) be fixed and let x,x˜ ∈ Ω such that|x− ˜x| ε, whereεwill be chosen sufficiently small in the following. Due to the assumptions on a, the pathlines passing through x andx are uniquely˜ determined and originate at some (x0, t0) and(x˜0t0), respectively. In other words,

x=S(t; x0, t0), x˜=S t;˜x0t0

.

Without loss of generality, let t0≥ ˜t0 >0, hence x0,x˜0∂Ω. The case when˜t0=0 can be treated similarly. Furthermore, we assume that x0is not a vertex of∂Ω—this case will be treated at the end of the proof.

(10)

Since a is uniformly bounded and Lipschitz continuous in space, there exists δ, such that if t[0, T) and dist{x,∂Ω}δthen

−a(x, t)·namin/2, (4.9) due to (4.8). Ifεis sufficiently small, then for the distance of the two considered pathlines at time t0we have|x0S(t0;x˜0t0)| = |S(t0; x0, t0)S(t0;x˜0t0)|δ. Since x0∂Ωthis means that S(t0;x˜0t0) is in theδ-neighborhood of∂Ωand by (4.9), dist{S(ϑ;x˜0t0),∂Ω}decreases asϑgoes from t0to˜t0

with a rate of at least amin/2 due to the uniformity of the bound (4.9). Therefore, S(ϑ;x˜0t0)stays in the δ-neighborhood of∂Ωfor allϑ

˜ t0, t0

and

−a S

ϑ;x˜0t0

,ϑ

·namin/2 for allϑ

˜ t0, t0

. Moreover, since x0lies in the interior of an edge of∂Ω, by choosingεsufficiently small, we can ensure thatx˜0also lies on this edge (face).

Now we estimate|μ1(x, t)μ1(˜x, t)| =t0− ˜t0. We have xx0=

t

t0

dS

dt(ϑ; x0, t0)dϑ = t

t0

a(S(ϑ; x0, t0),ϑ) dϑ,

˜ x− ˜x0=

t

˜ t0

dS dt

ϑ;x˜0t0

dϑ = t

˜ t0

a S

ϑ;˜x0t0

,ϑ dϑ.

Subtracting these two identities gives us

x0− ˜x0=x− ˜x+ t0

˜ t0

a S

ϑ;x˜0t0

,ϑ dϑ+

t t0

a S

ϑ;x˜0t0

,ϑ

a(S(ϑ; x0, t0),ϑ)dϑ.

(4.10) If we consider n, the normal to∂Ωat x0, we see that(x0− ˜x0)·n=0, as both x0andx˜0lie on the same edge on∂Ω. Therefore, if we multiply (4.10) by−n, we get

0=(x− ˜x)·(−n)+ t0

˜ t0

a S

ϑ;x˜0t0

,ϑ

·(−n)dϑ +

t

t0

a S

ϑ;x˜0t0

,ϑ

a(S(ϑ; x0, t0),ϑ)

·(−n)dϑ. (4.11)

Due to (4.9), we can estimate the first integral as t0

˜ t0

a S

ϑ;x˜0t0

,ϑ

·(−n)dϑ amin

2

t0− ˜t0

.

(11)

As for the second integral in (4.11), we have t

t0

a S

ϑ;x˜0t0

,ϑ

a(S(ϑ; x0, t0),ϑ)

·(−n)dϑ TLa sup

ϑ(t0,t)

S ϑ;x˜0t0

S(ϑ; x0, t0)TLa

eTLa −1

x− ˜x, (4.12)

where Lais the Lipschitz constant of a with respect to x. The last inequality in (4.12) follows from standard results on ordinary differential equations, namely continuous dependence of the solution on the initial condition—here we consider the ordinary differential equations defining S(·;·,·) backward in time on the interval t0, t with ‘initial’ conditions x and˜x at time t.

Finally,|(x− ˜x)·(n)|x− ˜x. Therefore, we get from (4.11), amin

2 |μ1(x, t)μ1(x, t)˜ | = amin

2 t0− ˜t0=amin

2

t0− ˜t0

1+TLa

eTLa−1

x− ˜x.

(4.13) Dividing by amin/2>0 gives Lipschitz continuity ofμ1, t).

Now we return to the case when x0 is a vertex of ∂Ω. Reasoning as above, by choosing ε sufficiently small, we can ensure thatx˜0∂Ω is sufficiently close to x0, i.e ˜x0 lies on one of the edges adjoining x0. Then we can again multiply (4.11) by−n, the normal to∂Ωatx˜0. Hence, (x˜− ˜x0)·(n)=0 will also be satisfied and we can proceed as in the previous case.

Remark 4.2 We note that the Lipschitz constant Lμ1 ofμ1can be estimated from (4.13) by Lμ1 2

amin

1+TLa

eTLa −1

.

The Lipschitz constant Lμofμcan then be obtained by multiplying by the constant factor from (4.4).

Having established uniform Lipschitz continuity ofμ1in space, we can prove its Lipschitz continuity with respect to time. This implies the differentiability ofμ1with respect to x and t a.e. in QT; therefore the left-hand side of (4.1) is well defined a.e. in QTand this expression is equal to the derivative ofμ1 along pathlines. This is important, as all the following considerations are thus justified.

Lemma 4.3 Let a satisfy the assumptions of Lemma4.1. Then μ1is uniformly Lipschitz continuous with respect to time t: there exists Lt0 such that

μ(x, t)μ

x,˜tLtt− ˜t for all xΩand t,˜t[0, T).

Proof . Let e.g.˜t>t and let (x0, t0) and

˜ x0t0

denote the origin of the pathlines passing through (x, t) and

x,˜t

, respectively. Therefore x=S(t; x0, t0)=S

˜ t;x˜0t0

. Denotex˜=S t;x˜0t0

. Then

|x− ˜x| =

˜t t

a S

ϑ;x˜0t0

,ϑ

aL(QT)|t− ˜t|. (4.14)

(12)

Therefore, by (4.3), μ(x, t)μ

x,˜t|t− ˜t| + |˜t0t0| =t− ˜t+ |μ(x, t)μ(x, t)˜ |t− ˜t+Lμ|x− ˜x|

1+LμaL(QT) t− ˜t

due to (4.14) and Lemma4.1. This completes the proof.

Example 4.4 In simple cases, the functionμ1 can be explicitly written down. As a trivial example, we take the one-dimensional stationary flow field a(x, t)=x + 1 onΩ =(0, 1) and the time interval (0,+∞). Then (4.2) can be easily solved to obtain

S(t; x0, t0)=(x0+1)ett0−1.

The (x,t)–plane is then separated into two regions separated by the pathline S(t;0, 0), which is the curve x=et1, i.e. t=ln(x+1). For points (x, t) beneath this curve, i.e. tln(x+1), we have t0=0, henceμ1(x, t)=tt0=t. For points (x, t) above the separation curve, we have x0=0, hence

x=S(t; 0, t0)=ett0 −1 ⇒ μ1(x, t)=tt0=ln(x+1).

Altogether, we have

μ1(x, t)=

t if tln(x+1), ln(x+1) otherwise.

This function is globally bounded and globally Lipschitz continuous. We note that the standard exponential scaling trick which gives exponential growth of the error corresponds to takingμ1(x, t)=t for all x, t, which is an unbounded function.

5. DG method

Having introduced the necessary tools for the analysis using the general exponential scaling argument, we can finally proceed to the DG method itself.

LetThbe a triangulation ofΩ, i.e. a partition ofΩ into a finite number of closed simplices with mutually disjoint interiors. As usual with DG, Th need not be conforming, i.e. hanging nodes are allowed. For KThwe set hK=diam(K), h=maxK∈ThhK.

For each KThwe define its inflow and outflow boundaries by

∂K(t)= {x∂K; a(x, t)·n(x) <0},

∂K+(t)= {x∈∂K; a(x, t)·n(x)≥0},

where n(x) is the unit outer normal to∂K. To simplify the notation, we shall usually omit the argument t in the following and write simply∂K±. Here we remark thatThneed not conform to∂Ω. In other words, the intersection of∂Ωand∂K need not be an entire face (edge) of K.

(13)

The triangulationThdefines the broken Sobolev space H1(Th)=

vL2(Ω); v|KH1(K)KTh

.

The approximate solution will be sought in the space of discontinuous piecewise polynomial functions Sh=

vh; vh|KPp(K)KTh , where Pp(K) is the space of all polynomials on K of degree at most p∈N.

Given an element KTh, for vhShwe define vh as the trace of vhon∂K from the side of the element adjacent to K. Furthermore, on∂K\∂Ω we define the jump of vhas [vh]=vhvh, where vh

is the trace from inside K.

The DG formulation of (2.1) then reads as follows: we seek uhC1([0, T); Sh)such that uh(0)= u0h, an Sh-approximation of the initial condition u0, and for all t(0, T),

∂uh

∂t , vh

+bh(uh, vh)+ch(uh, vh)=lh(vh)vhSh. (5.1)

Here bh, chand lhare bilinear and linear forms, respectively, defined for u, vH1(Th)as follows. The bilinear advection form bhis

bh(u, v)=

K∈Th

K

(a· ∇u)v dx

K∈Th

∂K\∂Ω

(a·n)[u]v dS

K∈Th

∂K∂Ω

(a·n)uv dS,

the reaction form is defined by

ch(u, v)=

Ω

cuv dx

and lhis the right-hand side form

lh(v)= −

K∈Th

∂K∂Ω

(a·n)uDv dx.

The definition of bh corresponds to the concept of upwinding which has already been applied in the context of DG in the original paper byReed & Hill (1973). The right-hand side form lhis based on the weak enforcement of the Dirichlet boundary condition on∂Ωby penalization. For a detailed derivation of the specific forms of bhand lhas used in the paper, we refer e.g. to the study byFeistauer & Švadlenka (2004).

6. Analysis of the advection and reaction terms

In this section, we prove the estimates of the bilinear forms bhand chnecessary for the error analysis.

First, we need some standard results from approximation theory.

(14)

6.1 Auxiliary results

In the following analysis we assume that the exact solution u is sufficiently regular, namely u, ut:=∂u

∂tL2

Hp+1

.

We consider a system{Th}h(0, h0), h0>0 of nonconforming triangulations ofΩthat are shape regular and satisfy the inverse assumption, cf.Ciarlet (1980). Under these assumptions we have the following standard results:

Lemma 6.1 (Inverse inequality). There exists a constant CI>0 independent of h, K such that for all KTh, and all vPp(K),

|v|H1(K)CIhK1vL2(K).

For vL2(Ω) we denote byΠhvShthe L2)-projection of v onto Sh:

hvv,ϕh)=0 ∀ϕhSh. (6.1)

Letηh(t)=u(t)Πhu(t) andξh(t)=Πhu(t)uh(t)Sh. Then we can write the error of the method as eh(t) :=u(t)uh(t)=ηh(t) +ξh(t). For simplicity, we shall usually drop the subscript h. We have the following standard approximation result; cf.Ciarlet (1980).

Lemma 6.2 There exists a constant CΠ >0 independent of h, K such that for all h(0, h0), η(t)L2(K)CΠhp+1|u(t)|Hp+1(K),

|η(t)|H1(K)CΠhp|u(t)|Hp+1(K), η(t)L2(∂K)CΠhp+1/2|u(t)|Hp+1(K), ∂η(t)

∂t

L2(K)

CΠhp+1|ut(t)|Hp+1(K).

Throughout the paper, C will be a generic constant independent of h, t, T. In order to track the dependence of the estimates on the functionμ, we will also assume that the generic constant C is independent ofμ, particularlyμmaxand Lμfrom (4.7), and we will explicitly track these dependencies.

The dependence of the resulting estimates of Theorem7.1onμwill then be clarified in Section7for the specific construction ofμfrom Section4.

6.2 Estimates of the advection and reaction terms

In our analysis we will assume there exists a constantγ0 and a functionμ : QT → Rsuch that the ellipticity assumption (3.9) holds. Furthermore, we assume thatμsatisfies (4.7), i.e.μis non-negative, bounded and Lipschitz continuous in space. Such a functionμwas constructed in Section4.

Similarly to Section2, we wish to write ξ(x, t) = eμ(x,t)ξ (x, t)˜ and test the error equation with φ (x, t)=eμ(x,t)ξ (x, t)˜ =e2μ(x,t)ξ(x, t)to obtain estimates forξ˜. However, sinceφ(t)/Shthis is not possible. One possibility is to test byΠhφ(t)Shand estimate the resulting differenceΠhφ(t)φ(t).

(15)

This is done in the stationary case in the study byAyuso & Marini (2009)and the analysis is carried out under the assumptionμWp+1,(Ω). Such high regularity can be achieved by mollification ofμ.

However, this would be somewhat technical in the evolutionary case, as space–time smoothing would be required in which case the dependence of all constants on T must be carefully considered. Also QT

is potentially an unbounded domain (for T= +∞) which leads to technical difficulties. Here we carry out the analysis under the weaker assumption (4.7), i.e.μ(t)W1,(Ω).

Lemma 6.3 Let μ satisfy assumptions (4.7). Let φ (x, t) = eμ(x,t)ξ (x, t)˜ = e2μ(x,t)ξ(x, t), where ξ(t)Sh. Then there exists C independent of h, t,ξ,ξ˜andμsuch that

Πhφ (t)φ (t)L2(K)CLμehKLμhKmax

xKeμ(x,t)˜ξ (t)L2(K), Πhφ (t)φ (t)L2(∂K)CLμehKLμh1/2K max

xKeμ(x,t)˜ξ (t)L2(K). (6.2) Proof . Let xKbe the centroid of K. On element K we introduce the constantμK(t)=μ(xK, t); then the function eK(t)ξ(·, t)lies in Pp(K), hence is fixed by the projectionΠh. Therefore,

Πhφ (t)φ (t)=Πh

e2μ(t)ξ(t)eK(t)ξ(t)

e2μ(t)ξ(t)eK(t)ξ(t)

=Πhw(t)w(t),

where w(x, t)=(e2μ(x,t)e2μ(xK,t))ξ(x, t). Standard estimates of the interpolation error ofΠhgive

Πhφ (t)φ (t)2L2(K) = Πhw(t)w(t)2L2(K)Ch2K|w(t)|2H1(K). (6.3)

For the right-hand side seminorm we have

|w(t)|2H1(K)=

K

e2μ(x,t)e2μ(xK,t)

ξ(x, t)2dx

2

K

∇e2μ(x,t)ξ(x, t)dx+2

K

e2μ(x,t)e2μ(xK,t)

ξ(x, t)2dx.

Ifμ(t)C1(Ω), by the mean value theorem

e2μ(xK,t)e2μ(x,t)hK∇e2μ(ζ,t)=hKe2μ(ζ,t)2|∇μ(ζ, t)|

(16)

for some pointζ on the line between x and xK. Therefore, by the inverse inequality,

|w(t)|2H1(K)2

K

4e4μ(x,t)|∇μ(x, t)|2|ξ(x, t)|2dx+2

K

h2Ke4μ(ζ,t)4|∇μ(ζ, t)|2|∇ξ(x, t)|2dx 8|μ(t)|2W1,∞

K

e4μ(x,t)e2μ(x,t)ξ (x, t)|2dx+8C2Imax

xKe4μ(x,t)|μ(t)|2W1,∞ξ(t)2 8L2μmax

xKe2μ(x,t)e2hKLμ˜ξ (t)2+8C2Imax

xKe2μ(x,t)e2hKLμL2μ˜ξ (t)2.

Substituting into (6.3) gives us the first inequality in (6.2) for μ(t)C1(Ω). The case μ(t)W1,(Ω) follows by a density argument. The second inequality in (6.2) can be obtained similarly, only intermediately applying the trace inequalityξ(t)L2(∂K) ChK1/2ξ(t)L2(K). Remark 6.4 The estimates of Lemma6.3 are effectively O(hK)˜ξ and O(h1/2K )˜ξ estimates, respectively. We note that the dependence of the estimates on Lμis rather mild, effectively linear, due to the small factor hKin the exponent.

Now we shall estimate individual terms in the DG formulation. Due to the consistency of the DG scheme, the exact solution u also satisfies (5.1). We subtract the formulations for u and uhto obtain the error equation

∂ξ

∂t, vh

+

∂η

∂t, vh

+bh(ξ, vh)+bh(η, vh)+ch(ξ, vh)+ch(η, vh)=0 (6.4)

for all vhSh.

As stated earlier we want to test (6.4) byφ (x, t)=eμ(x,t)ξ (x, t); however,˜ φ(t)/Sh. We therefore set vh=Πhφ(t) and estimate the difference using Lemma6.3. We write (6.4) as

∂ξ

∂t,Πhφ

+bh,φ)+bh,Πhφφ)+bh(η,φ)+bh(η,Πhφφ) + ch,φ)+ch,Πhφφ)+ch(η,Πhφ)+

∂η

∂t,Πhφ

=0. (6.5)

We will estimate the individual terms of (6.5) in a series of lemmas. For this purpose, we introduce the following norm on a subsetωof∂K or∂Ω:

fa,ω =

|a·n|f

L2(ω),

where n is the outer normal to∂K or∂Ω. We will usually omit the argument t to simplify the notation.

In the following lemma we show that selected terms from (6.5) indeed give us the desired ellipticity similarly to (3.8) and (3.9) plus additional elliptic terms due to the use of the upwind flux.

Références

Documents relatifs

The main purpose of this paper is to analyze the stability and error estimates of the local discontinuous Galerkin (LDG) methods coupled with implicit-explicit (IMEX)

Quantum chemistry, electronic Schr¨ odinger equation, coupled Cluster method, numerical analysis, non- linear operator equation, quasi-optimality, error estimators.. ∗ This work

The almost optimal error estimâtes for the discontinuous Galerkin method may be taken as a basis for rational methods for automatic time step control Further, the discontinuous

In the numencal solution of elhptic boundary value problems, ît is well- known that the présence of corners m the domain can cause a loss of accuracy in the solution Many methods

R.A.I.R.O. Analyse numérique/Numerical Analysis, vol.. In fact, we shall adapt to the situation described above Osborn's method [ 5 ] ; note that, independently of the fact that A h

In order to estimate the distance between the solutions u of (2) and u h of (7) we shall follow the procedure described in [9] for another unilatéral problem; of course our

In order to suppress the round-off error and to avoid blow-up, a regularized logarithmic Schr¨ odinger equation (RLogSE) is proposed with a small regularization parameter 0 &lt; ε ≪

An analysis in the energy norms for the model unsteady Stokes problem shows that this scheme enjoys the time convergence properties of both underlying methods: for low value of