• Aucun résultat trouvé

As a byproduct, we show a k+ 2-th order superconvergence of the DG solution towards the Gauss-Radau projection of the exact solution

N/A
N/A
Protected

Academic year: 2022

Partager "As a byproduct, we show a k+ 2-th order superconvergence of the DG solution towards the Gauss-Radau projection of the exact solution"

Copied!
33
0
0

Texte intégral

(1)

SUPERCONVERGENCE OF DISCONTINUOUS GALERKIN METHOD FOR SCALAR NONLINEAR HYPERBOLIC EQUATIONS

WAIXIANG CAO, CHI-WANG SHU, YANG YANG, AND ZHIMIN ZHANG§ ¶

Abstract. In this paper, we study the superconvergence behavior of the semi-discrete discon- tinuous Galerkin (DG) method forscalarnonlinear hyperbolic equations in one spatial dimension.

Superconvergence results for problems with fixed and alternating wind directions are established.

On the one hand, we prove that, if the wind direction is fixed (i.e., the derivative of the flux function is bounded away from zero), both the cell average error and numerical flux error at cell interfaces converge at a rate of 2k+ 1 when upwind fluxes and piecewise polynomials of degreekare used.

Moreover, we also prove that the function value approximation of the DG solution is superconver- gent at interior right Radau points, and the derivative value approximation is superconvergent at interior left Radau points, with an order ofk+ 2 andk+ 1, respectively. As a byproduct, we show a k+ 2-th order superconvergence of the DG solution towards the Gauss-Radau projection of the exact solution. On the other hand, superconvergence results for problems with alternating wind directions (i.e., the derivative of the flux function either changes sign or otherwise achieves the value zero in the domain) are also established. To be more precise, we first prove that the DG flux function is superconvergent towards a particular flux function of the exact solution, with an order ofk+ 2, when Godunov fluxes are used. We then prove that the highest superconvergence rate of the DG solution itself isk+32 when sonic points (i.e., the derivative of the flux function achieves zero) appear in the computational domain. As byproducts, we obtain superconvergence properties for the DG solution and the DG flux function at special points and for cell average. Numerical experiments demonstrate that most of our results are optimal, i.e., the superconvergence rates are sharp.

Key words.Discontinuous Galerkin methods, superconvergence, nonlinear hyperbolic equations

AMS subject classifications. 65M15, 65M60, 65N30

1. Introduction. In this paper, we investigate the superconvergence behavior of the discontinuous Galerkin (DG) method for the following one-dimensional nonlinear hyperbolic equation

ut+f(u)x= 0, (x, t)[a, b]×[0, T], u(x,0) =u0(x), x∈Ω = [a, b],¯ (1)

where u0 is sufficiently smooth. We assume that the nonlinear flux functionf(u) is sufficiently smooth with respect touand the final timeT is not too large so that the exact solutions are smooth. In this paper, for simplicity, we only consider the periodic boundary condition.

The superconvergence behavior of the DG (see, e.g., [12, 13, 14, 15, 16, 17]) method has been studied for many years, and has been a hot research topic in recent

School of Mathematical Sciences, Beijing Normal University, Beijing 100875, China. Email:

caowx@bnu.edu.cn. Research was supported in part by NSFC grant No.11501026, and the China Postdoctoral Science Foundation grants 2016T90027 and 2015M570026.

Division of Applied Mathematics, Brown University, Providence, RI 02912, USA. E-mail:

shu@dam.brown.edu. Research was supported in part by DOE grant DE-FG02-08ER25863 and NSF grant DMS-1418750.

Department of Mathematical Sciences, Michigan Technological University, Houghton, MI 49931, USA. E-mail: yyang7@mtu.edu. Research was supported in part by NSFC grants No.11571367 and No.11601536 and Michigan Technological University Research Excellence Fund Scholarship and Creativity Grant 1605052

§Beijing Computational Science Research Center, Beijing 100193, China. E-mail:

zmzhang@csrc.ac.cn. Research was supported in part by NSFC grants No.11471031 and No.91430216.

Department of Mathematics, Wayne State University, Detroit, MI 48202, USA. E-mail:

zzhang@math.wayne.edu. Research was supported in part by NSF grant DMS-1419040.

1

(2)

years. We refer to [1,2,3,10,11,18,23,24,4,19] for an incomplete list of references on the superconvergence of DG methods for hyperbolic problems. However, all the studies in the literature are based on linear equations. Only recently, Meng et al.

studied the superconvergence of semi-discrete DG methods for (1) with fixed wind direction, i.e., |f(u)| possesses a uniform positive lower bound [20]. They proved that the order of the error between the DG solution and a particular projection of the exact solution can achievek+32 when upwind fluxes were used. Compared with the linear case, where the highest superconvergence rate can reach 2k+ 1 (see, e.g., [7, 6]), the superconvergence rate k+ 32 is far from optimal. Furthermore, to our best knowledge, no superconvergence results of DG methods applied to nonlinear hyperbolic equations with sonic points (i.e., points at whichf(u) = 0) are available in the literature.

The main purpose of this paper is to study superconvergence properties of DG methods for (1) with general flux functions. If the wind direction is fixed, our analy- sis indicates a 2k+ 1-th order convergence rate of the numerical flux at mesh nodes and for the cell average, and a k+ 2-th order of the error between the DG solution and the Gauss-Radau projection of the exact solution as well as the function value error at interior right Radau points, and a k+ 1-th order of the derivative error at interior left Radau points. As we may recall, these superconvergence results are the same as for non-degenerate linear hyperbolic problems (see, [7,6]). While if the wind direction is changing in the computational domain, we proved that the superconver- gence phenomenon still exists and the convergence rate may depend upon the specific property of the flux functions. Specifically, we first prove that the DG flux function is superconvergent towards a particular flux function of the exact solution, with an error boundO(hk+2), when Godunov fluxes are used. Then we establish the supercloseness result between the DG solution itself and a particular projection of the exact solution, and reveal that the highest superconvergence rate for nonlinear hyperbolic equations with sonic points isO(hk+32). As byproducts, we obtain superconvergence properties for the DG solution and the DG flux function at special points and for cell averages.

These superconvergence results are similar to linear problems with degenerate variable coefficients, see [5].

The contribution of this paper is to provide a symmetric method to study the superconvergence behavior of the DG methods for nonlinear problems. On the one hand, we establish the superconvergence results for the monotone flux, and prove that all the superconvergence results for the linear case in [7, 6] still hold true for nonlinear problems. Furthermore, the superconvergence results established in this paper improve those of [20] to the possible optimal superconvergence rates. On the other hand, we uncover the superconvergence phenomenon of the DG methods for (1) with the degenerate flux function, and extend the superconvergence result for linear problems to a more general nonlinear case. By doing so, we present a full picture for the superconvergence properties of DG methods applied to the (degenerate) nonlinear hyperbolic problems, and enrich the superconvergence theory of the DG method for linear hyperbolic equations in one dimension.

To end the introduction, we would like to emphasize that the main difficulty in the superconvergence analysis for nonlinear problems is how to deal with the nonlinear terms in the error equation. Our analysis is along the following lines: we first use a Taylor expansion to linearize the error equation; subsequently, we make an a priori error assumption to deal with the nonlinearity of the flux and other high-order terms in the linearization, then the superconvergence analysis for nonlinear problems is reduced to a linear one; finally, we introduce a correction function to deal with the

2

(3)

linear part to obtain higher-order accuracy.

The rest of the paper is organized as follows. In Section 2, we present DG schemes for the one-dimensional nonlinear hyperbolic conservation laws, and discuss the choice of numerical fluxes. Section 3 is dedicated to the superconvergence analysis of the DG methods for problems with fixed wind direction. Superconvergence results are established by using the idea of correction function, Taylor expansion, and an a priori error assumption. In Section 4, we study the superconvergence behavior of the DG approximation for problems with sonic points, and prove that the DG flux function is superconvergent with an order ofO(hk+2) to a particular flux function of the exact solution. Superconvergence behavior of the DG solution itself is also discussed in this case. We reveal a very important fact that the superconvergence phenomena are also valid for general fluxes, and the superconvergence rates may depend upon specific properties of the flux functions. In Section 5, we provide numerical examples to support our theoretical findings. Finally, concluding remarks and some possible future works are presented in Section 6.

Throughout this paper, we adopt standard notations for Sobolev spaces such as Wm,p(D) on sub-domainD Ω equipped with the norm∥ · ∥m,p,D and semi-norm

|·|m,p,D. WhenD= Ω, we omit the indexD; and ifp= 2, we setWm,p(D) =Hm(D),

∥ · ∥m,p,D=∥ · ∥m,D, and| · |m,p,D=| · |m,D. NotationAB implies thatA can be bounded byB multiplied by a constant independent of the mesh sizeh.

2. DG schemes and the energy inequality. Let Ω = [0,2π] and 0 =x1 2 <

x3

2 < . . . < xN+1

2 beN+ 1 distinct points on the interval Ω. For all positive integers r, we defineZr={1, . . . , r}, and denote by

τj= (xj1 2, xj+1

2), xj =1 2(xj1

2 +xj+1

2), jZN

the cells and cell centers, respectively. Let hj =xj+1

2 −xj1

2, hj = hj/2 and h= max

j hj. We assume that the mesh is regular, i.e., the ratio between the maximum and minimum mesh sizes shall stay bounded during mesh refinements.

Define the discontinuous finite element space

Vh={v: v|τj Pkj), jZN},

where Pk denotes the space of polynomials of degree at most k with coefficients as functions of t. The DG scheme for (1) reads as: Find uh Vh such that for any v∈Vh

(2) ((uh)t, v)j(f(uh), vx)j+ ˆf(uh)v

j+12 −fˆ(uh)v+

j12 = 0, jZN, where (u, v)j =∫

τjuvdx,v

j+12 andv+

j+12 separately denote the left and right limits of vat the pointxj+1

2, and ˆf(uh) is the numerical flux, which is a single-valued function defined at each cell interface and in general depends on the values of the numerical solutionuh from both sides of the interface, i.e.,

fˆ(uh)

j+12 = ˆf(uh(xj+1 2

), uh(x+j+1 2

)), ∀j∈ZN.

The choice of the numerical flux is of great significance in assuring the stability of the DG scheme. Iff(u)≥δ >0, for all x∈Ω, we define the numerical flux as

fˆ(uh) =f(uh).

3

(4)

Similarly, iff(u)≤ −δ <0, we choose

fˆ(uh) =f(u+h).

However, iff(u) changes its sign in the computational domain, we take the Godunov flux as our numerical flux, i.e.,

fˆ(uh) = ˆf(uh, u+h) =

{ minu

hµu+h f(µ), if uh < u+h, maxu+

hµuh f(µ), if uh ≥u+h. (3)

Note that the upwind flux for problems with alternating wind direction is defined as follows.

f(u˜ h) =

{ f(uh), if f(u)>0, f(u+h) if f(u)0.

(4)

The above flux is an essential part in our later energy estimate and error analysis for problems with alternating wind direction.

Throughout this paper, we will use the following notation (5) e=u−uh, ξ=uI−uh, η=u−uI. HereuI is some special interpolation function ofuto be determined.

To end with this section, we would like to present the following energy inequality, which will be frequently used in our later superconvergence analysis.

Theorem 1. Let ube the solution of (1)anduI ∈Vh be some special interpola- tion function of u. Assume that uh is the solution of (2) with the numerical fluxes chosen as the upwind flux (for fixed wind directions) or Godunov flux (for alternating wind directions) . Then for both numerical fluxes,

(6) 1 2

d

dt∥ξ∥20h2∥e∥20,∥η∥20+ (1 +h1∥e∥0,)∥ξ∥20+h1∥e∥20,

N j=1

ηj+1

2)2+|I|+|J|, whereη˜ is the numerical flux defined in (4)with f(uh)replaced byη, and

I=

N j=1

( ˜f(uh)−fˆ(uh))[ξ]

j+12, (7)

J =t, ξ) + (f(u)η, ξx) +

N j=1

(f(u)˜η[ξ])

j+12. (8)

Proof. Since the exact solution also satisfies (2), we have for allv∈Vh

(9) (et, v)j(f(u)−f(uh), vx)j+ (f(u)−fˆ(uh))v|j+12(f(u)−fˆ(uh))v+|j12 = 0.

Summing up overjfrom 1 toN and using the periodic boundary condition, we obtain (et, v) = (f(u)−f(uh), vx) +

N j=1

(

(f(u)−fˆ(uh))[v])

j+12,

4

(5)

where (u, v) =∑N

j=1(u, v)j and [v]j+1

2 =v+j+1

2 −vj+ 1 2

denotes the jump of v at the pointxj+1

2. Now we takev=ξin the above identity to get (10) (ξt, ξ) =−t, ξ) + (f(u)−f(uh), ξx) +

N j=1

(

(f(u)−f˜(uh))[ξ])

j+12 +I.

By the Taylor expansion off aboutu, we have

f(u)−f(uh) =f(u)ξ+f(u)η1 2

f¯u′′(ξ+η)2, f(u)−f(u˜ h) =f(u) ˜ξ+f(u)˜η−1

2

¯¯

f′′u( ˜ξ+ ˜η)2, (11)

where ¯fu′′=f′′1u+ (1−α1)uh) and ¯f¯′′u=f′′2u+ (1−α2)uh) with 0≤α1, α21.

Then (10) can be rewritten as

(12) (ξt, ξ) =I1−I2

2 +I+J, where

I1= (f(u)ξ, ξx) +

N j=1

(

f(u) ˜ξ[ξ])

j+12, I2= ( ¯fu′′(ξ+η)2, ξx) +

N j=1

(f¯¯u′′( ˜ξ+ ˜η)2[ξ])

j+12. Let

(13) Ω1={j∈ZN :fj+ 1

2 0},2={j∈ZN :fj+ 1 2

<0}, wherefj+ 1

2

=f(u(xj+1

2, t)). By a simple integration by parts, we have

I1=1

2(∂xf(u), ξ2)

N j=1

(f(u){ξ}[ξ])|j+12 + ∑

j1

(f(u)ξ[ξ])|j+12 + ∑

j2

(f(u)ξ+[ξ])|j+12

=1

2(∂xf(u), ξ2)

j1

(f(u)[ξ]2)|j+12 + ∑

j2

(f(u)[ξ]2)|j+12 ≤ −1

2(∂xf(u), ξ2)≤C∥ξ∥20. Here and in the following, C is a positive constant independent of the mesh sizeh,

and is not necessary to be the same at every occurrence. As for I2, we have from Cauchy-Schwarz inequality

|I2| ≤C∥e∥0,

∥e∥0∥ξx0+∥ξ∥2Γh+

N j=1

η˜j+1 2[ξ]j+1

2

h1∥e∥0,(

∥e∥0∥ξ∥0+∥ξ∥20

)+∥e∥0,

N j=1

˜ηj+1 2[ξ]j+1

2

h2∥e∥20,∥η∥20+ (1 +h1∥e∥0,)∥ξ∥20+h1∥e∥20,

N j=1

ηj+1 2)2,

5

(6)

where for any givenv,

∥v∥Γh =

∑N

j=1

(v

j+12)2+ (v+

j+12)2

1 2

,

and in the second and last steps, we have used the inverse inequality

∥vx0≤Ch1∥v∥0, ∥v∥Γh ≤Ch12∥v∥0, ∥v∥0,j ≤Ch12∥v∥0,τj, ∀v∈Vh. Then (6) follows by substituting the estimates of I1 and I2 into (12). The proof is complete.

3. Analysis for problems with fixed wind direction. In this section, we study the superconvergence of DG methods for problems with fixed wind direction.

Without loss of generality, we consider the case f(u)≥δ >0. The same arguments can be applied to the case withf(u)≤ −δ <0.

To study the superconvergence properties of the DG solution, our analysis is along this line: We first construct a special interpolation function uI Vh of the exact solution such that the DG solutionuh is super-close touI underL2 norm; and then we analyze the superconvergence behavior of the interpolation function uI; finally, the superconvergence for the DG solution is reduced to the superconvergence of the interpolation function uI due to the super-closeness betweenuh and uI. Therefore, our goal here is to design the special functionuI.

Note that

∥ξ(·, t)∥20=∥ξ(·,0)20+

t 0

d

dt∥ξ∥20dt, ∀t∈(0, T],

then the convergence rate of∥ξ∥0 depends upon the term dtd∥ξ∥20. In light of (6) and the fact that I in (6) vanishes for fixed wind directions, to achieve our superconver- gence goal, uI should be specially designed such that J in (6) is of high order. To this end, we begin with some preliminaries.

3.1. Preliminaries. First, we denote byLn andLj,n the traditional Legendre polynomials of degreenin the intervals [1,1] andτj, j∈ZN, respectively. For any x∈τj, let

(14) ω(x, t) =f(u(x, t)), w¯j(s, t) =ω(xj+ ¯hjs, t) =ω(x, t), s∈[1,1].

Note that for any fixed t and j, where j ZN, ¯ωj ≥δ > 0 L10), τ0 = [1,1], there exists a series of monic orthogonal polynomials¯j,n}n=1 with respect to the weight function ¯ωj, i.e.,

(15) ( ¯ϕj,n¯j,m)ω¯j :=

1

1

¯

ωjϕ¯j,nϕ¯j,mds=γnδmn,

where γn = ( ¯ϕj,n¯j,n)ω¯j and δmn is the Kronecker delta. Moreover, ¯ϕj,n can be constructed by the following three-term recurrence formula

(16) ϕ¯j,0= 1, ϕ¯j,1=s−α0, ϕ¯j,n+1= (s−αn) ¯ϕj,n−βnϕ¯j,n1, n≥1, where for alln≥0, m1,

(17) αn = (sϕ¯j,n¯j,n)ω¯j/( ¯ϕj,n¯j,n)ω¯j, βm= ( ¯ϕj,m¯j,m)ω¯j/( ¯ϕj,m1ϕ¯j,m1)ω¯j.

6

(7)

With the orthogonal polynomials¯j,n}on [1,1], we can construct a set of orthog- onal polynomialsϕj,n in each elementτj as follows.

ϕj,n(x, t) = ¯ϕj,n(s, t) = ¯ϕj,n(2(x−xj) hj , t)

, x∈τj.

Moreover, there holds the following orthogonal property (18)

τj

j,nϕj,mω)(x, t)dx= hj 2

1

1

( ¯ϕj,nϕ¯j,mω¯j)(s, t)ds= hj 2 γnδmn.

That is, j,n} are orthogonal polynomials with respect to the weight functionω = f(u) in τj.

Lemma 2. For any j ZN, let j,n} be a sequence of orthogonal polynomials with respect to the weight functionω=f(u)inτj, andf(u)be a sufficiently smooth function satisfying|∂tmf(u)|≲1, m≤k+ 1. Then

(19) ∥ϕj,n0,τjh12, ∥ϕj,n0,jh12∥ϕj,n0,τj ≲1.

Moreover, for all positivem, where1≤m≤k+ 1, there holds

(20) ∥∂tmϕj,n0,≲1.

Proof. First, (19) can be easily verified by a direct calculation. To show (20), it is sufficient to prove

(21) |∂mt ϕ¯j,n|≲1, 1≤m≤k+ 1.

We will prove (21) by induction. For any integer m≥1, we have, from (16) and the Newton-Leibniz formula of derivative,

tmϕ¯j,1=−∂mt α0=−∂mt (∫1

1¯jds

1

1ω¯jds )

m

r=0∥∂rtω¯j0,

1

1ω¯jdsm+1

≲1.

Then (21) holds for n= 1. Now we suppose (21) is valid for alln and prove it also holds forn+ 1. Actually, recall the three-term recurrence formula of ¯ϕj,n in (16), we easily get

|∂tmϕ¯j,n+1|≲(|∂tmαn|+|∂tmβn|) (|∂mt ϕ¯j,n|+|∂tmϕ¯j,n1|)≲(|∂mt αn|+|∂tmβn|). Again, we use the Newton-Leibniz formula of derivative to obtain

(|∂tmαn|+|∂mt βn|)≲

m

r=0|∂trϕ¯j,n| ( ¯ϕj,n¯j,n)m+1ω¯j +

m

r=0(|∂trϕ¯j,n|+|∂rtϕ¯j,n1|) ( ¯ϕj,n1¯j,n1)m+1ω¯j ≲1, and thus,

|∂mt ϕ¯j,n+1|≲1.

Then (21) is also valid forn+ 1, and this finishes our proof.

7

(8)

Let

Hh1={v: v|τj ∈H1j), jZN},

and for allψ∈Hh1, we denote byPhψ∈Vh the traditional Gauss-Radau projection ofψ, which is defined by

(22) (Phψ, v)j= (ψ, v)j, ∀v∈Pk1, Phψ(xj+1 2

) =ψj+ 1 2

, j∈ZN.

Note that the Gauss-Radau projectionPhis frequently used in the DG error analysis.

Now we extend the definition of Gauss-Radau projection to a more general one. Given a positive functionω, we define a projectionPh:Hh1→Vh by

(23)

τj

ωPhψvdx=

τj

ωψvdx, ∀v∈Pk1, Phψ(x

j+12) =ψ(x

j+12).

Note thatPh is reduced to the traditional Gauss-Radau projectionPh whenω = 1.

Furthermore, we have the following approximation properties for the projectionPh. Lemma 3. Let ω = f(u) δ > 0. Then the projection Ph in (23) is well- defined. Moreover, if ψ Wk+2,(Ω) and |∂xnf(u)| ≲ 1, n k, there holds the following results.

Phψ is super-close to the Gauss-Radau projectionPhψ, i.e., (24) ∥∂tm(Phψ−Phψ)∥0,jhk+2∥∂mt ψ∥k+1,, m= 0,1.

The function value approximation of Phψ is superconvergent at the interior right Radau pointsrj,m,1≤m≤k+ 1(zeros of the right Radau polynomial Lj,k+1−Lj,k except the pointxj+1

2), namely,

(25) (Phψ−ψ)(rj,m)≲hk+2∥ψ∥k+2,, m∈Zk.

The derivative value approximation ofPhψis superconvergent at the interior left Radau points lj,m, m∈Zk (zeros of left Radau polynomial Lj,k+1+Lj,k except the pointx=xj1

2 ), i.e.,

(26) x(Phψ−ψ)(lj,m)≲hk+1∥ψ∥k+2,, m∈Zk.

The cell average of Phψ in each element τj is superconvergent with an order of 2k+ 1, i.e.,

(27)

1 hj

τj

−Phψ)(x)dx

h2k+1∥ψ∥k+1,τj. Proof. SincePhψ∈Vh, we express in each elementτj

Phψ

τj =

k n=0

anϕj,n(x),

where {an}k0 are constants to be determined. Recall the definition of Ph and the orthogonality ofj,n} in (18), we easily obtain

an=

τjωψϕj,ndx

τjωϕj,nϕj,ndx = 2 hjγn

τj

ωψϕj,ndx, n≤k−1,

8

(9)

whereγn is the same as in (15). Moreover, noticing that the zeros of the orthogonal polynomials¯j,n}are all real, simple, and lies in the interval (1,1) (see [22], Theo- rem 3.2), we have ¯ϕj,n(1)̸= 0, or equivalently,ϕj,n(xj+1

2

)̸= 0, which yields, together with the second identity of (23),

ak= 1 ϕj,k(xj+1

2

) (

ψ(xj+1 2

)

k1

n=0

anϕj,n(xj+1 2

) )

= 1

ϕ¯j,k(1) (

ψ(xj+1 2

)

k1 n=0

anϕ¯j,n(1) )

.

Therefore, the projectionPhψfor any functionψ∈Hh1 is uniquely determined.

Now we assumePhψ−Phψ has the following expression inτj (Phψ−Phψ)(x, t) =

k n=0

bnϕj,n(x, t), x∈τj. Recalling the definitions ofPhandPh in (22) -(23), we have

bn =

τjω(Phψ−Phψ)ϕj,ndx

τjωϕj,nϕj,ndx = 2 hjγn

τj

ω(ψ−Phψ)ϕj,ndx

for alln≤k−1, and

bk =

k1

n=0

bn ϕ¯j,n(1) ϕ¯j,k(1).

Noticing that (ψ−Phψ)⊥Pk1, there holds for alln≤k−1,

|bn|= 2 hjγn

τj

−ω(xj))(ψ−Phψ)ϕj,ndx

∥ω−ω(xj)0,j∥ψ−Phψ∥0,j∥ϕj,n0,jh∥ω∥1,j∥ψ−Phψ∥0,j, where in the last step we have used (19). Consequently,

|bk|

k1

n=0

|bn|h∥ω∥1,j∥ψ−Phψ∥0,j,

and thus,

(28) ∥Phψ−Phψ∥0,j

k n=0

|bn|h∥ψ−Phψ∥0,j.

Following the same line, we can prove

(29) ∥∂tm(Phψ−Phψ)∥0,jh∥∂tm−Phψ)∥0,j, m≤k+ 1.

Then (24) follows from the standard approximation property of Ph (see, e.g., [21]).

Thanks to the super-closeness result (24), together with the superconvergence result forPh (see, [7])

−Phψ)(rj,i)≲hk+2∥ψ∥k+2,, x−Phψ)(lj,i)≲hk+1∥ψ∥k+2,,

9

(10)

the desired results (25)-(26) follow immediately.

Now we prove (27). On the one hand, using the orthogonality ofϕj,n, we get

τj

ϕj,ndx=

τj

ω1ωϕj,ndx=

τj

1−In1ω1)ωϕj,ndx,

where Inv Pn is some interpolation function of v. Since ω =f(u) is sufficiently smooth,i.e.,|∂xnω|≲1, we have

∥∂xnω10,j∥ω∥n,j

∥ω∥n+10,j

≲1.

Then

τj

ϕj,ndx

hn+1∥∂xnω10,jhn+1.

On the other hand, by the orthogonality ofPh, that is, (ψ−Phψ)⊥Pk1, we have

|bn|=

τj

ω(ψ−Phψ)ϕj,ndx =

τj

−Ik1nω)(ψ−Phψ)ϕj,ndx

h2k+1n∥ψ∥k+1,τj∥∂xknω∥0,τjh2k+1n∥ψ∥k+1,τj. Noticing that

τj

−Phψ)dx=

τj

(Phψ−Phψ)dx=

k n=0

bn

τj

ϕj,ndx,

then

τj

−Phψ)dx

h2k+2∥ψ∥k+1,τj, which yields (27) directly. The proof is complete.

To end this subsection, we introduce a special operatorL, which will be used in later analysis. For allψ∈Hh1, we define the operator L:Hh1→Vhby

(f(u)L(ψ), vx)j = (ψ, v)j, ∀v∈Pkj)\P0j), (30)

L(ψ)(xj+1 2

) = 0, ∀j ZN. (31)

Noticing that vx Pk1j) for all v∈Pkj)\P0j), we have from the definition ofPh

(f(u)Phψ, vx)j = (ωPhψ, vx)j = (ωψ, vx)j, v∈Pkj)\P0j).

Then the definition of L is similar to that of Ph except the term in the right-hand side (withωvxreplaced byv). Therefore, we can similarly prove that the operatorL is well-defined.

3.2. Construction of uI and its approximation properties. We first con- struct a series of functions starting withw0=u−Phuand define

(32) wi+1=L(∂twi), 1≤i≤k.

We have the following properties.

10

(11)

Lemma 4. Let ω=f(u)≥δ >0be a sufficiently smooth function satisfying (33) |∂xmf(u)|≲1 |∂tmf(u)|≲1, m≤k+ 1.

In each elementτj, we express wi’s in terms of orthogonal basis j,n} as

wi|τj =

k n=0

ci,nϕj,n.

Then the coefficientci,n satisfies

(34) |∂tmci,n|hmax(k+1+i,2k+1n)∥∂m+it u∥k+1,, n≤k, m= 0,1.

Consequently, there hold form= 0,1, (35)

∥∂tmwi0,jhk+i+1∥∂tm+iu∥k+1,,

1 hj

τj

tmwidx

h2k+1∥∂ti+mu∥k+1,. Proof. We only prove (34)-(35) for m = 0 since the similar argument can be applied to m = 1. We first prove (34) by induction. As v Pk \P0, we have vxPk1. Then we choosevx=ϕj,n, n= 0, . . . , k1 in (30) to obtain

ci+1,n=

τjtwiDx1ϕj,ndx

τjωϕj,nϕj,ndx = 2 hjγn

τj

twiDx1ϕj,ndx, n≤k−1, (36)

ci+1,k=

k1

n=0

ci+1,n

ϕ¯j,n(1) ϕ¯j,k(1), (37)

whereγn is given by (15), andDx1vis a function defined by

(38) Dx1v

τj =

x xj

1 2

v(x)dx.

SinceDx1ϕj,nPn+1j), we have, from the orthogonality ofPh, c1,n= 2

hjγn

τj

ω1ω∂t(u−Phu)Dx1ϕj,ndx

= 2

hjγn

τj

(ω1−Ik2nω1)

ω∂t(u−Phu)Dx1ϕj,ndx.

Here againImω∈Pmdenotes the interpolation function ofω. By (33) and standard approximation theory,

∥ω1−Ik2nω10,jhk1n∥∂xk1nω10,hk1n, which yields, together with (19) and (24),

|c1,n|hkn1∥∂t(u−Phu)∥0,j∥Dx1ϕj,n0,j

hkn(∥∂t(u−Phu)∥0,+|∂t(Phu−Phu)∥0,)∥ϕj,n0,h2k+1n∥∂tu∥k+1, 11

Références

Documents relatifs

Besides the Gevrey space analysis and the convergence results of Maday and Quarteroni already mentioned, our proof rests on previous work of Tadmor [25], who showed that for

Complexity of enumeration, first-order queries, structures of bounded degree, linear time, constant time, constant delay.. ∗ In Honour of Serge Grigorieff at the occasion of his

Consider the unsteady axisymmetric mixed convection boundary layer flow over a vertical permeable slender cylinder of radius R placed in a non-Newtonian power law fluid with power

In order to extend the energy conservation property of the the semi-discrete method to the fully discrete method, it is natural to employ time stepping methods conserving

and SKA~:S, D., Integer sequences having prescribed quadratic character,

PUBLISHED BY PACIFIC JOURNAL OF MATHEMATICS, A NON-PROFIT CORPORATION The Supporting Institutions listed above contribute to the cost of publication of this Journal, but they are

The cost analysis will look at what equipment (computers, control quipm nt, soft- ware), real estate a nd human resources ar required to implement Lhe

However, volumes of limit and market orders are equal in the two latter cases, and we have shown in section 3 that these different shapes can actually be obtained with different