• Aucun résultat trouvé

A finite element approximation for grazing collisions

N/A
N/A
Protected

Academic year: 2021

Partager "A finite element approximation for grazing collisions"

Copied!
29
0
0

Texte intégral

(1)

HAL Id: hal-00076838

https://hal.archives-ouvertes.fr/hal-00076838

Submitted on 24 Oct 2012

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de

To cite this version:

Brigitte Lucquin-Desreux, Simona Mancini. A finite element approximation for grazing collisions.

Transport Theory and Statistical Physics, Taylor & Francis, 2003, 32, 3& 4, pp.293-319. �hal-00076838�

(2)

Vol. 32, Nos. 3 & 4, pp. 293–319, 2003

A Finite Element Approximation of Grazing Collisions

B. Lucquin-Desreux * and S. Mancini Laboratoire Jacques-Louis Lions, Universite´ Pierre et Marie Curie, Paris, France

ABSTRACT

In this article, we propose a finite element discretization of the Boltzmann-Lorentz operator for which it is possible to define a grazing collision limit. We illustrate this discretization by considering the evolution of a system of particles in a slab, subject to collisions both with the boundaries and with themselves. A comparison is made with isotropic collisions or with the Laplace-Beltrami operator.

Moreover, in the case of multiplying boundary conditions, it is proven by numerical simulations the existence of a critical value for the absorption coefficient, which is independent on the grazing collision parameter. Finally, the focalization of a beam is studied and the numerical results are compared with previous simulations.

AMS Classification: 35B40; 65N30; 82C40.

*Correspondence: B. Lucquin-Desreux, Laboratoire Jacques-Louis Lions, Universite´ Pierre et Marie Curie, Boite Courrier 187, 75252 Paris, Cedex 05, France; E-mail: lucquin@ann.jussieu.fr.

293

DOI: 10.1081/TT-120024765 0041-1450 (Print); 1532-2424 (Online) Copyright&2003 by Marcel Dekker, Inc. www.dekker.com

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

(3)

1. INTRODUCTION

Kinetic equations, as Boltzmann (or Fokker-Planck) equations, model the evolution of a system of particles subject to the action of a force field and to collisions. They are applied for example in semiconductor physics (see Markowich et al. (1994)), in the modeling of rarefied gases or plasmas (see Cercignani (1988)), in astrophysics (see Belleni-Morante and Moro (1996) for problems relative to interstellar clouds) or in bio- mathematics (see Arlotti et al. (2000) for the study of dynamics of popula- tions). In the more general frame, the operator describing the scattering event is nonlinear and due to its complexity, many other models have been derived. Notably, the Lorentz operators, derived from the Boltzmann one or from the Fokker-Planck one, model the elastic collisions of particles against heaviest target ones (see Degond and Lucquin-Desreux (1996) and Lucquin-Desreux (2000) for the disparate masses particles asymptotics).

In plasma physics for example, the Boltzmann-Lorentz operator repre- sents (in first approximation with respect to the mass ratio) the effect of collisions due to neutral particles on the electron distribution function.

On the other hand, the Fokker-Planck-Lorentz operator is the leading order term (still with respect to the mass ratio) describing the collisions of electrons against ions.

The Fokker-Planck operator is usually considered as an approxi- mation of the Boltzmann collision operator when collisions become grazing, i.e., when the angle of deviation of a particle during a collision is very small. This essentially occurs for potentials of long range inter- action, such as the Coulombian potential in the case of a plasma.

This grazing collision limit of the Boltzmann operator towards the Fokker-Planck one has been proven in Degond and Lucquin-Desreux (1992) and Desvillettes (1992). In Desvillettes (1992), the scattering cross section is smooth; it depends on a small parameter " and it is localized around the value ¼ 0 of the scattering angle when " goes to zero. On the other hand, the Coulombcase is studied in Degond (1992) by means of a quite different asymptotics: the scattering cross section has a nonintegrable singularity when the relative velocity of the colliding particles tends to zero but also when the scattering angle tends to zero. There is thus a necessary truncation of this angle around the value zero ( ") and the small parameter " has a precise physical meaning: it is clearly identified as the plasma parameter. We also refer the reader to the works Goudon (1997a, 1997b) concerning the grazing collision limit of the Boltzmann operator when the scattering cross section corresponds to a inverse power force (i.e., 1=r

s

, with s >1) and without angular cut-off. More recently, the convergence of Boltzmann-Lorentz operator towards the

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

81

82

83

84

(4)

Fokker-Planck-Lorentz (or Laplace-Beltrami) operator in the so-called grazing collision limit has been proven in Buet et al. (2001), both for the asymptotics developed in Degond and Lucquin-Desreux (1992) and Desvillettes (1992).

In this article, we are concerned with the numerical approximation of the Boltzmann-Lorentz operator and with its grazing collision limit. Up to our knowledge, only few recent articless have been devoted to the study of the grazing collision limit at the discrete level: we can refer for example to Gue´rin and Me´le´ard (2002) and Pareschi et al. (2002) concerning the full nonlinear Boltzmann operator discretized either by spectral methods (see Pareschi et al. (2002)) or by particle methods (see Gue´rin and Me´le´ard (2002)). In our article, the considered scattering cross section is a simplified version of the one defined in Buet et al.

(2001): it is assumed to be constant on a very small interval depending on the small parameter " and zero outside this interval. Our goal is to give a numerical approximation of the Boltzmann-Lorentz operator for graz- ing collisions for which it is possible to pass to the limit " ! 0. We first decided to apply a finite element method (FEM) because this method is particularly well adapted to the limit problem (we refer for example to Cordier et al. (2000) for 3D computations). Secondly, in the FEM described below, we decided to do exact computations of the collision terms (instead of doing quadrature formulas): denoting by h the velocity step, we can then choose " h and thus study the limit " ! 0 for a fixed discretization in velocity variable, i.e., for a fixed h (which would not be possible when using quadrature formulas).

In order to show the efficiency of the proposed FEM approximation, we will consider a transport equation in which the collision operator is the Boltzmann-Lorentz operator for grazing collisions defined on a bounded region. Thus, we equip the kinetic equation with some kind of boundary conditions: in a first approach, we only take into account specular reflec- tion boundary conditions. The numerical results obtained with the pro- posed FEM approximation are compared with those obtained considering the Laplace-Beltrami operator (i.e., its limit when " goes to zero), with the isotropic collisions (i.e., the scattering cross section is constant), and also with the collisionless case. Successively, we consider multiplying reflection boundary conditions (see for instance Belleni-Morante and Totaro (1996) and Mancini and Totaro (1998)): each particle colliding on the boundary of the considered region is reflected and multiplied by a factor bigger than one. In this optic, as the region where the evolution of particles takes place is bounded, it seems physically necessary to consider also an absorption term in our transport equation. The existence and uniqueness of the solution of such a problem, with the absorption cross section bigger

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

104

105

106

107

108

109

110

111

112

113

114

115

116

117

118

119

120

121

122

123

124

125

126

(5)

than a theoretical value

t

, has been proven in Belleni-Morante (1996) for multiplying Maxwell boundary conditions, and in Mancini and Totaro (1998) in the general case. We will numerically prove the existence of a critical absorption cross section

n

<

t

for which the total number of particles does not sharply increase nor vanishes with time. Moreover, we also remark that both

n

and the equilibrium profile of the particle density do not depend on the initial data.

The article is organized as follows. In Sec. 2, we present the finite element approximation for the Boltzmann-Lorentz operator for grazing collisions and also for isotropic collisions. In Sec. 3, we present the test problems used to validate our numerical approximation. Section 4 con- cerns the results of the numerical simulations. Finally, in Sec. 5, we give some concluding remarks.

2. NUMERICAL SCHEMES

The goal of this article is the numerical approximation of the Boltzmann-Lorentz operator for grazing collisions. We recall that the Boltzmann-Lorentz operator models for example the elastic collisions of a heavy particle against a light one (see Degond and Lucquin-Desreux, (1996) and Lucquin-Desreux (2000)). Thus, it acts only on the velocity direction of the particles, leaving unchanged the velocity modulus jvj (which we will consider normalized to one). In the bi-dimensional case, the Boltzmann-Lorentz operator is given by:

f ÞðÞ ¼ Z

S1

0

Þ ð f ð

0

Þ f ðÞÞ d

0

, ð1Þ where

0

Þ is the scattering cross section which represents the probability for a particle entering a collision with velocity direction

0

to out-go the scattering event with a velocity direction . For instance,

0

Þ may be determined from the interaction force between the particles (see Degond and Lucquin-Desreux (1992), Desvillettes (1992), and Goudon (1997b)); in particular, we recall that, in the case of isotropic collisions, it does not depend on nor on

0

.

We are interested in grazing collisions, i.e., we want to take into account also the very small deviations in the trajectories of the particles.

This fact is emphasized by introducing the dependence of the scattering cross section B on a small parameter ". Recently, Buet, Cordier, and Lucquin-Desreux (see Buet et al. (2001)) have proven the convergence, when the small parameter " tends to zero, of the Boltzmann-Lorentz

127

128

129

130

131

132

133

134

135

136

137

138

139

140

141

142

143

144

145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162

163

164

165

166

167

168

(6)

operator for grazing collisions towards the Fokker-Planck-Lorentz operator, which is simply the Laplace-Beltrami operator.

Our goal here is to give a numerical approximation of the Boltzmann-Lorentz operator for which it is possible to pass to the limit

" ! 0 for a fixed velocity discretization. As shown in the sequel, this is possible when applying for example a finite element approach, where the integral terms are all exactly computed. The numerical results we obtain are then compared with the collisionless case, the case of isotropic collisions, and those obtained when considering the Laplace-Beltrami operator.

2.1. An Implicit Finite Element Approximation

We want to approximate by means of a finite element method the following Cauchy problem:

@f

@t ¼ f Þ, f ðt ¼ 0Þ ¼ f

0

, 8

<

:

ð2Þ

where f Þ is defined by Eq. (1). We note that the dependence of the density f on the space variable is not underlined in this paragraph because the position of the particles does not play any role during a collision. The angle variable belongs to the unit circle S

1

, i.e., 2 ½, and the distribution function f satisfies periodic boundary conditions.

We first consider the following piecewise affine approximation f

N

of the function f :

f ðt, Þ ’ f

N

ðt, Þ ¼ X

N1

j¼0

j

ðÞ f

j

ðtÞ,

where ’

j

, j 2 f0, . . . , N

1g, are the classical ‘‘hat functions’’ of the P

1

finite element approximation. More precisely, denoting by

j

¼ þ hj, with h ¼ 2=N

, we have ’

i

ð

j

Þ ¼

ij

(=1 if i=j and 0 otherwise), so that f

j

ðtÞ ¼ f

N

ðt,

j

Þ.

The weak discretized form of Eq. (2) then reads:

X

N1

j¼0

@f

j

@t Z

S1

j

ðÞ’

i

ðÞ d

¼

N

X

1

j¼0

f

j

Z

S1

Z

S1

0

Þ½’

j

ð

0

Þ’

i

ðÞ ’

j

ðÞ’

i

ðÞ d

0

d :

169

170

171

172

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

191

192

193

194

195

196

197

198

199

200

201

202

203

204

205

206

207

208

209

210

(7)

Setting:

A

ij

¼ Z

S1

i

ðÞ’

j

ðÞ d, Q

ij

¼

Z

S1

Z

S1

0

Þ ½’

i

ðÞ ’

j

ð

0

Þ ’

i

ðÞ ’

j

ðÞ d d

0

, 8

> >

<

> >

:

ð3Þ

we write the discretized problem more compactly as follows:

X

N1

j¼0

@f

j

@t A

ij

¼

N

X

1

j¼0

f

j

Q

ij

: ð4Þ

The matrix A with entries A

ij

is the so called ‘‘mass matrix,’’ while the matrix Q with entries Q

ij

represents the collision matrix. In the next paragraph, we give the expression of these two matrices in the context of grazing collisions; the detailed computations are performed in the Appendix A1. Before this, let us precise the time discretization of Eq. (4). In order to avoid a limitation of the time step t in terms of the angular step ¼ h, we will use a fully implicit scheme. We introduce some notations. Let us denote by F

n

the N dimensional vector with entries F

jn

¼ f ðnt,

j

Þ, for j 2 f0, . . . , N 1g. The implicit finite element scheme we consider reads:

F

nþ1

¼ ðA tQÞ

1

AF

n

, n 0 F

0

¼ ðF

i0

Þ

0jN1

,

(

ð5Þ

with F

j0

¼ f

0

ð

j

Þ.

2.2. Grazing Collisions

In the case of grazing collisions, the scattering cross section B is localized and depends on a small parameter ". In the nonCoulombian case, B has the following expression (see Buet et al. (2001) and Desvillettes (1992)):

B

"

ðzÞ ¼ 1

"

3

Bðz="Þ

S1

ðz="Þ

where denotes the characteristic function (i.e.,

S1

ðzÞ ¼ 1 if z 2 S

1

and 0 otherwise). We will here consider a simplified model for which the scattering cross section is given by:

B

"

ðzÞ ¼ B

0

if jzj < ", 0 else

n

211

212

213

214

215

216

217

218

219

220

221

222

223

224

225

226

227

228

229

230

231

232

233

234

235

236

237

238

239

240

241

242

243

244

245

246

247

248

249

250

251

252

(8)

with B

0

¼ 1="

3

. We then have:

Proposition 2.1. Setting ¼ "=h and assuming that 1, the grazing collision matrix Q is given by:

Q ¼

3

3h

2 þ

34

1

12

18

0 . . . 0

18

1

12

1

12

. . . . . .

. . . . . .

. . . . . .

1 8

1

8

. . . . . . . . . . . . . . . . . .

0

0 . . .

. . . . . .

. . . . . .

. . .

.. . .. .

. . . . . .

. . . . . .

. . . . . .

0 0 . . . . . . . . . . . . . . . . . .

1 8

1

8

. . .

. . . . . .

. . . . . .

. . .

1

12

1

12

18

0 . . . 0

18

1

12

2 þ

34

0 B B B B B B B B B B B B B B B B

@

1 C C C C C C C C C C C C C C C C A :

ð6Þ Proof. The proof is detailed in the Appendix A1.

By means of the same arguments applied in order to compute the matrix Q (see Appendix A1), we have that the mass matrix A reads:

A ¼ h 3

2

12

0 . . . 0

12

1 2

. . .

. . . . . .

. . . 0 0 . . .

. . . . . .

. . . .. . .. .

. . . . . .

. . . . . .

0 0 . . .

. . . . . .

. . .

1 1 2

2

0 . . . 0

12

2

0 B B B B B B B B B B

@

1 C C C C C C C C C C A

: ð7Þ

We remark that if we let " ! 0 then Q

ij

! ðh

3

Þ=ð3D

ij

Þ, where D

ij

is the classical finite difference discretization of the Laplace-Beltrami operator f ¼ @

f :

D ¼ 1 h

2

2 1 0 . . . 0 1

1 . . . . . .

. . . . . .

0 0 . . .

. . . . . .

. . . .. . .. .

. . . . . .

. . . . . .

0 0 . . .

. . . . . .

. . . 1

1 0 . . . 0 1 2

0 B B B B B B B B

@

1 C C C C C C C C A :

253

254

255

256

257

258

259

260

261

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

277

278

279

280

281

282

283

284

285

286

287

288

289

290

291

292

293

294

(9)

Moreover, the coefficient

3

=3 is strictly related to the second order moment of the scattering cross section, which in our case is equal to

3

=3 (since B ¼ 1).

2.3. Isotropic Collisions

In the following simulations, in order to compare the evolution of the system of particles with grazing collisions and with isotropic collisions, we will not compute the exact solution for the isotropic Boltzmann-Lorentz operator (although it is possible), but we will approx- imate the isotropic operator following the same discretization as before.

Recalling that for isotropic collisions the scattering cross section B is a constant (which we will assume equal to one), we get the following finite element discretization of the isotropic Lorentz operator. The gain matrix G is given by: G ¼ ðG

ij

Þ

0i,jN1

, with G

ij

¼ h

2

, whereas, the loss matrix L is such that: L ¼ 2A. Then we can apply the implicit finite element scheme (5), where now the matrix Q ¼ G L represents the P

1

finite element approximation of the isotropic Lorentz operator.

3. THE TEST PROBLEM

We illustrate the finite element approximation for the Boltzmann-Lorentz operator with grazing collisions on the following transport equation:

@

t

f þ cos @

x

f þ f ¼ 1

f Þ: ð8Þ

This transport type equation describes the evolution of a system of particles moving between two parallel plates (slab), subject to absorption by the host medium and to elastic collisions with other particles present in the slab. We recall that f ¼ f ðx, , represents the number of particles which at time t 0, are in the position x 2 ½L, þ L with velocity v ¼ ðcos , sin Þ, 2 ½, . Moreover, f Þ is the Lorentz collision operator defined by Eq. (1), 0 is the absorption cross section, and is the relaxation time; this one is assumed to be equal to one, except in the last numerical test concerning the defocusing of a beam (see Sec. 4.4 below and Cordier et al. (2000)).

The interactions of the particles with the two plates of the slabmay be described by means of a linear, bounded, and positive operator

295

296

297

298

299

300

301

302

303

304

305

306

307

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

325

326

327

328

329

330

331

332

333

334

335

336

(10)

:

out

!

in

which relates the incoming and outgoing flux of particles as follows:

f

in

¼ f

out

: ð9Þ

In Eq. (9), f

in

and f

out

are the traces of the distribution function respectively to the incoming and outgoing boundary sets:

in

¼ fðL, Þ, cos > 0g fðL, Þ, cos < 0g,

out

¼ fðL, Þ, cos > 0g fðL, Þ, cos < 0g:

We define the norms on the incoming and outgoing sets as usual:

k f k

out

¼ Z

ðx,Þ2out

f ðx, Þj cos j d , k f k

in

¼ Z

ðx,Þ2in

f ðx, Þj cos j d , and we consider the norm for the operator to be given by:

kk ¼ max

kfoutkout¼1

kf

out

k

in

:

We will consider multiplying boundary conditions (see Mancini and Totaro (1998)), i.e., the operator is such that kf

out

k

in

ckf

out

k

out

, with c > 1. In other words, a multiplication of particles takes place on the boundaries. For instance, the boundary operator may represent reflec- tion, or diffusion boundary conditions, as well as a linear combination of them. In the following simulations, we will consider multiplying reflective boundary conditions:

f ðL, Þ ¼

L

f ðL,

0

Þ , for cos > 0, f ðL, Þ ¼

R

f ðL,

0

Þ , for cos < 0,

ð10Þ with

0

¼ , where

L

and

R

respectively denote the reflection coeffi- cients on the left and right boundaries of the slab (i.e., on x ¼ L and x ¼ L respectively). As the region where the evolution takes place is bounded, it is natural to consider also an absorption term f in order to avoid the growth of the number of particles (see the results of Fig. 7 below). Summarizing, the problem which we approximate numerically reads as follows:

@

t

f þ cos @

x

f þ f ¼

1

f Þ, f

in

¼ f

out

,

f ðx, , 0Þ ¼ f

0

ðx, Þ:

8

> <

> :

ð11Þ

It is clear that the efficiency of the absorption coefficient will depend on the ‘‘power’’ of the multiplication coefficient (i.e., the norm of the operator ), on the modulus of the particle velocities jvj (here

337

338

339

340

341

342

343

344

345

346

347

348

349

350

351

352

353

354

355

356

357

358

359

360

361

362

363

364

365

366

367

368

369

370

371

372

373

374

375

376

377

378

(11)

equal to 1) and on the dimension of the considered region (here the distance 2L between the two plates). It has been proven in Mancini and Totaro (1998), by means of the semi-groups theory and for a fixed

", that if kk 1, then problem (11) admits a unique solution whatever is 0; whereas, if kk > 1, then problem (11) admits a unique solution provided that:

> jvj ln kk

2L ¼

t

; ð12Þ

t

is called the theoretical critical absorption cross section. In the particular case of multiplying reflective boundary conditions (10), the theoretical value for the critical absorption cross section is given by (see Belleni-Morante and Totaro (1996)):

t

¼ jvj lnðmaxf

L

,

R

2L : ð13Þ

We will numerically compute a critical value

n

for the absorption cross section and we will precise the behavior of the solution of problem (11) in terms of the absorption coefficient : we will see that, if <

n

the particle density significantly grows in a finite time, for >

n

the particle density rapidly goes to zero in a finite time, and for ¼

n

it remains bounded without vanishing (see Sec. 4.3 below). This numerical critical value

n

does not depend on "; moreover, it is smaller than the theoretical one

t

. This fact does not contradict at all the results proven in Belleni-Morante and Totaro (1996) or Mancini and Totaro (1998); in fact, it just shows the existence of a solution also for absorption cross sections 2

n

,

t

½ (since

n

<

t

), which was not possible to prove by means of the semigroup theory.

4. NUMERICAL SIMULATIONS

In order to simulate the evolution problem (11), we apply a first order splitting in time method. First, we approximate by means of an upwind explicit finite difference method the following transport problem (see paragraph 4.1 below):

@f

@t þ cos @f

@x þ f ¼ 0, f

in

¼ f

out

,

f ðx, , 0Þ ¼ f

0

ðx, Þ:

8

> <

> :

ð14Þ

379

380

381

382

383

384

385

386

387

388

389

390

391

392

393

394

395

396

397

398

399

400

401

402

403

404

405

406

407

408

409

410

411

412

413

414

415

416

417

418

419

420

(12)

Second, we apply the implicit finite element scheme of Sec. 2.1 to the pure collision problem:

@f

@t ¼

1

Qf f ðx, , 0Þ ¼ f

0

ðx, Þ:

(

ð15Þ

At each iteration, the program solves separately the transport and collision parts. Our whole numerical scheme is first order in all variables.

We have done this choice in order to simplify the presentation of the method, but it would be naturally possible to improve the accuracy of the method by using appropriate second order schemes for each step (convection, collision) coupled with a second order (of Strang type) splitting algorithm.

We will here perform essentially three different numerical tests. In the first two ones, developed in paragraphs 4.2 and 4.3 below, the relaxation time is equal to one. In paragraph 4.2, we essentially focalize on the discrete collision operator itself: the boundary conditions are classical specular reflection, and there is no absorption; the grazing collision operator is then compared to the Laplace-Beltrami operator (i.e., to its limit when " goes to zero), to the isotropic operator and also to the collisionless case. The dependence of the solution on the grazing parameter " is also numerically investigated, first for " h=, but also for bigger values of ". The influence of multiplicative boundary condi- tions is then studied in paragraph 4.3: in particular, the existence of a numerical critical absorption cross section, which does not depend on "

(for sufficiently small values of ") nor on the initial data is shown.

Moreover, we underline the existence of a unique profile of equilibrium (independent on " and on the initial data) for the particle density. Finally, in paragraph 4.4, we study the influence of the time relaxation for a photonic type problem already considered in a previous work (see Cordier et al. (2000)). This last test is further investigated in Cordier et al. (2002).

Let us firstly precise the numerical scheme used to approximate the convective part of the equation, i.e., problem (14).

4.1. Transport Problem Approximation

We consider N

x

(=50 in simulations, unless otherwise specified) discretization points x

i

, i 2 f1, . . . , N

x

g, for the space variable x 2 ½L, L (with L =1) and N

(=50 in simulations, unless otherwise specified) discretization points for the angle variable 2 S

1

. We introduce the following symmetric discretization of the circle S

1

:

j

¼ þ hj with

421

422

423

424

425

426

427

428

429

430

431

432

433

434

435

436

437

438

439

440

441

442

443

444

445

446

447

448

449

450

451

452

453

454

455

456

457

458

459

460

461

462

(13)

h ¼ 2=N

and j ¼ 0, . . . , N

1. We have that if cos

j

is a discretization point, then also cos

j

is one. Moreover, if j and j

0

are such that j

0

¼ jÞ, where jÞ ¼ ðN

=2 jÞmodðNÞ (see Fig. 1), then: cos

j0

¼ cos

j

.

Considering the above discretization, we approximate the transport problem (14) by means of an upwind explicit finite difference scheme as follows:

.

If cos

j

> 0 then, f

1,nþ1j

¼

L

f

1,nþ1

and for i ¼ 2 . . . N

x

f

ijnþ1

¼ ð1 ðt=xÞ cos

j

tÞf

ijn

þ ðt=xÞ cos

j

f

i1,n j

: ð16Þ

.

If cos

j

< 0 then f

Nnþ1x,j

¼

R

f

Nnþ1x,

and for i ¼ 1 . . . N

x

1

f

ijnþ1

¼ ð1 þ ðt=xÞ cos

j

tÞf

ijn

ðt=xÞ cos

j

f

iþ1,n j

: ð17Þ We recall that f

ijn

¼ f ðx

i

, cos

j

, t

n

Þ is the value of the distribution function f at the position x

i

¼ L þ xði 1Þ, i ¼ 1 . . . N

x

, and x ¼ 2L=ðN

x

1Þ, with velocity cos

j

, j ¼ 0 . . . N 1, and at time t

n

¼ n t, n ¼ 1 . . . N

iter

.

Setting ¼ t=x, the CFL condition writes (for any 0): 1.

In the simulations below, is taken equal to 0.8 (but ¼ 1 would give exactly the same results). Moreover, we choose

L

¼ 2

R

¼ 4 and

" ¼ 0:001, unless otherwise specified.

j=0

j s(j)

Figure 1. Discretization on the sphere.

F1

463

464

465

466

467

468

469

470

471

472

473

474

475

476

477

478

479

480

481

482

483

484

485

486

487

488

489

490

491

492

493

494

495

496

497

498

499

500

501

502

503

504

(14)

4.2. Grazing Collisions and Comparison with Other Collision Terms

In this section we present numerical results concerning the grazing collision limit at the discrete level and also a comparison with different types of collisions. We take into account only specular reflection bound- ary conditions (i.e.,

L

¼

R

¼ 1), the absorption cross section is given by ¼ 0 (i.e., the number of particles is constant along all the evolution), and the relaxation time is ¼ 1. We introduce the density function:

nðx, tÞ ¼ Z

f ðx, , d

and the number of particles inside the slab at time t:

NðtÞ ¼ Z

þL

L

nðx, dx: ð18Þ

In Fig. 2, we show how the behavior of the density nðx, changes when changing the value of the grazing parameter ". We have first considered values of " such that " h=, but also larger values of ": in this last case,

ε ε x

π π n(x,T)

ε π

-1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 0.01

0.05 0.10 0.14 0.19 0.23 0.28 0.32 0.37 0.41 0.46

+ + + + + + ++ + ++ + ++ + ++ + + ++ + + ++ + ++ + ++ ++++ +

+ +

+ +

+ +

+ +++ ++ + +

grazing 0.01

+

Laplace-Beltrami =2h/

=h/ =3h/

Figure 2. The density nðx; TÞ, with ¼ 0,

R

¼

L

¼ 1, for different values of ":

" ¼ 10

2

; h=; 2h=;3h= and for the Laplace-Beltrami operator, with initial data given by a Dirac, t ¼ tN

iter

, with N

iter

¼ 100, N

x

¼ 50, N

¼ 10:

F2

505

506

507

508

509

510

511

512

513

514

515

516

517

518

519

520

521

522

523

524

525

526

527

528

529

530

531

532

533

534

535

536

537

538

539

540

541

542

543

544

545

546

(15)

the collision terms have been computed by use of standard first order quadrature formulas. In particular, we note that for " < 10

2

the difference between the grazing collision case and the Laplace-Beltrami operator is too small to be observed. In fact, the difference is of order "h, so that it is rapidly very small for small values of " (and for small h as well). In order to precise this behavior, we have computed in Fig. 3 the L

1

norm of the relative error between the density n

"

of the grazing collision problem and that of the limiting one, denoted by n

LB

, i.e.,

max

i

n

"

ði, T Þ n

LB

ði, n

LB

ði, T Þ

,

in terms of " and for different values of h (with T ¼ 50t). We first remark that this error is very small for " ¼ 10

3

. But as expected, we observe a difference between the case " ¼ h= and " ¼ 10

2

for large values of h (h ¼ =2), while this difference decreases for smaller values of h. Figure 2 is related to the ‘‘medium’’ case h ¼ =5 (i.e., N

¼ 10).

In Fig. 4, we compare four types of collisions: the collisionless case (Q ¼ 0), the case of isotropic collisions (with B ¼ 1=4 in order to fit the first nonzero eigenvalues), the grazing collisions case (with " ¼ 10

3

) and its limit when " ! 0, i.e., the Laplace-Beltrami operator. We trace the evolution of the density nð, for four different times of computation T ¼ tN

iter

, with N

iter

¼ 50, 100, 150, 200. We observe that the grazing collisions case and the Laplace-Beltrami operator give the same results (up to a difference of order 10

4

which is not noticeable on the figure).

Naturally, in all cases, we have checked numerically that the total density NðtÞ defined by Eq. (18) was constant with respect to time.

4.3. Absorption Cross Section and Multiplying Boundary Conditions

Still assuming ¼ 1, we now take into account multiplying boundary conditions, in particular

L

¼ 2,

R

¼ 4. We consider two initial

Figure 3. Influence of parameters " and h.

F3

F4

547

548

549

550

551

552

553

554

555

556

557

558

559

560

561

562

563

564

565

566

567

568

569

570

571

572

573

574

575

576

577

578

579

580

581

582

583

584

585

586

587

588

(16)

x n(x,T)

1.0 0.8 0.6 0.4 0.2 0.0 0.2 0.4 0.6 0.8 1.0

+ + + + + + + + ++ + + + ++ + + + ++ + + + + ++ + + + ++ +++++ +

+ +

+ +++ ++ + + + + +

0.12 0.10 0.08 0.06 0.04 0.02 0.00

LaplaceBeltrami50

+ grazing50

isotrope50 no collisions50

x n(x,T)

1.0 0.8 0.6 0.4 0.2 0.0 0.2 0.4 0.6 0.8 1.0

+ + + ++ + ++ + ++ + + + ++ + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + +

no collisions100 isotrope100 LaplaceBeltrami100

grazing100

+

0.12 0.10 0.08 0.06 0.04 0.02 0.00

x n(x,T)

1.0 0.8 0.6 0.4 0.2 0.0 0.2 0.4 0.6 0.8 1.0

+ + + + + + + + + + + + + + + + + + + + + + ++ + + + + + + + + + + + + + + + + + + + + + + + + + +

no collisions150 isotrope150 grazing150

+

LaplaceBeltrami150 0.20

0.18 0.16 0.14 0.12 0.10 0.08 0.06 0.04 0.02 0.00

x n(x,T)

1.0 0.8 0.6 0.4 0.2 0.0 0.2 0.4 0.6 0.8 1.0

+ + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + +

isotrope200 LaplaceBeltrami200

0.20 0.18 0.16 0.14 0.12 0.10 0.08 0.06 0.04 0.02 0.00

+ grazing200 no collisions200

Figure 4. The density nð; TÞ, with ¼ 0,

R

¼

L

¼ 1, for the grazing collisions with " ¼ 0:001, the Laplace-Beltrami operator, the isotropic collisions case, and the collisionless case, with initial data given by a Dirac, T ¼ tN

iter

, with N

iter

¼ 50 (top on the left), 100 (top on the right), 150 (bottom on the left), 200 (bottom on the right).

1 2002 47 6 5_ TT32 _ 03 -0 4_ R 1_ 0 717 03

[PageNo.307]F:/MDI/Tt/32(3&4)/120024765_TT_032_003-004_R1.3dTransportTheoryandStatisticalPhysics(TT)

Approximation of Grazing Collisions 307

(17)

data: first, a constant function both in space and velocity; second, a Dirac mass both in space and velocity.

In Fig. 5, we show the convergence of the sequence of critical absorp- tion cross sections towards a finite value

n

for " ¼0:001. The critical absorption value we compute is independent (for large times) on the initial data: in fact we observe that both the constant initial data and the Dirac mass give the same value. Moreover, this value is smaller than the theoretical one

t

given by Eq. (13). The way the computation is performed is the following one:

.

We compute the total densities Nðt

i

, Þ for t

i

¼ ði þ 1Þ50t, with i ¼ 0, . . . , 9, and for different values of .

.

We determine the intersection point of two successive curves Nðt

i

, Þ and Nðt

iþ1

, Þ describing the evolution of the density in terms of ; the respective value of for the intersection points are called critical and denoted by

c

ðiÞ. In fact, for each

c

ðiÞ, the total density is constant for the two successive times of iteration t

i

and t

iþ1

.

1 2 3 4 5 6 7 8 9

0.40 0.41 0.42 0.43 0.44 0.45 0.46 0.47 0.48 0.49 0.50

constant dirac

Figure 5. The interpolation curve of

c

ðiÞ. Each point is given by the intersection of the total density curves for two successive iteration times, i.e. 1 corresponds to the intersection point of 50 and 100 iteration curves. We increase of 50 iterations at every ordinate point, finishing by 500 iterations.

F5

589

590

591

592

593

594

595

596

597

598

599

600

601

602

603

604

605

606

607

608

609

610

611

612

613

614

615

616

617

618

619

620

621

622

623

624

625

626

627

628

629

630

(18)

We observe in Fig. 5 that these values of critical absorption cross sections numerically converge to a finite value

n

, which is the searched value. Moreover, additional numerical simulations show that this value

n

does not depend on " (for sufficiently small values of ").

In Fig. 6, we show the limit profile of the curve nð, for the com- puted critical absorption ¼

n

¼ 0:466 and t ¼ 500t. Again the initial condition is given by a constant function or a Dirac mass. We underline the fact that for both initial data, we end up with the same (up to a normalization) profile which does not depend on ", for " small enough.

In Fig. 7, we show the influence of multiplying reflection boundary conditions with respect to different values of the absorption cross section.

When time evolves, we clearly see the significant growth of the number of particles NðtÞ when <

n

, and its fast shoot to zero for >

n

. For ¼

n

, this number remains bounded without ‘‘exploding’’ both in the case of the initial condition given by a constant function or a Dirac mass. Moreover, it seems to reach an almost constant value. Hence, this numerical approximation

n

of the critical cross-section

c

done in Fig. 5 seems to be satisfactory.

-1.0 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1.0

0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

+ + + + + ++ + + + + + + + + + + + + + + + + + + + ++ + + ++ + ++ + ++ ++ ++++++ +

+ +

+

constant dirac

+

Figure 6. The final profile of nð; with initial data a Dirac mass or a constant function, both in position and velocity; ¼ 0:466, " ¼ 0:001,

R

¼ 2

L

¼ 4, t ¼ 500t:

F6

F7

631

632

633

634

635

636

637

638

639

640

641

642

643

644

645

646

647

648

649

650

651

652

653

654

655

656

657

658

659

660

661

662

663

664

665

666

667

668

669

670

671

672

(19)

Finally, in Fig. 8, we show the behavior of the density function nð, for three different iteration times. The initial data which we consider is a Dirac mass both in space and velocity. The absorption cross section is the numerical critical coefficient

n

¼ 0:466. It is clearly seen that the Dirac mass is convected and diffused respectively by means of the trans- port equation and of the Boltzmann-Lorentz operator for grazing colli- sions (with " ¼ 0:001). Finally, the density is also multiplied and reflected at the boundaries, and we can see the beginning of boundary layers.

4.4. Defocusing of a Beam

Following the test performed in Cordier et al. (2000) for a photonic type model, we now assume that between the two plates there is vacuum.

Moreover, a focussed beam enters in the slab on the left side (i.e., at x ¼ L) and scatters. The absorption cross section is fixed to ¼ 0, the grazing coefficient to " ¼ 0:001 and the discretization points in x and respectively to N

x

¼ 50 and N

¼ 50. We now consider the influ- ence of the relaxation time . The boundary condition corresponding to

0 100 200 300 400 500

0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0

sigma=0.466 sigma=0.1 sigma=1

Figure 7. The total density N(t) for different values of the absorption ( >

n

, <

n

, and ¼

n

¼ 0:466) with initial data a constant function, " ¼ 0:001 and t ¼ 500t:

F8

673

674

675

676

677

678

679

680

681

682

683

684

685

686

687

688

689

690

691

692

693

694

695

696

697

698

699

700

701

702

703

704

705

706

707

708

709

710

711

712

713

714

(20)

the injection of a mono-kinetic beam in the direction perpendicular to the planes of the slabis given by:

f ðL, cos > 0, tÞ ¼ ,

where denotes the delta measure with respect to the angle variable ; on the right boundary we have:

f ðL, cos < 0, tÞ ¼ 0,

which means that the particles leaving the slabon the right side never return.

In Fig. 9, we draw the focalization coefficient, defined as the ratio between the particles leaving the slab on the right side with a velocity perpendicular to the plane, and the injected particles:

F ð, tÞ ¼ f ðL, ¼ 0, tÞ:

The computations are done for t sufficiently large (t ¼ 150t) so that the density has reached a stationary state.

Finally, in Fig. 10 we show for three values of the relaxation time, ¼ 2

5

, 1, 2

5

, and for four iteration times N

iter

¼ 25, 50, 100, 200 the particle density nð, evolution (with t ¼ tN

iter

).

Let us note that the numerical results we have obtained for this model are in perfect agreement with those performed in Cordier et al. (2000)

-1.0 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1.0

0 1 2 3 4 5

1 iter 20 iter 100 iter

Figure 8. The evolution of nð; with initial data a Dirac mass (normalized) in position and velocity.

F9

F10

715

716

717

718

719

720

721

722

723

724

725

726

727

728

729

730

731

732

733

734

735

736

737

738

739

740

741

742

743

744

745

746

747

748

749

750

751

752

753

754

755

756

(21)

for a 1D in space and 2D in velocity (i.e., R S

2

) analogous problem, where the collision operator involved was the Laplace-Beltrami operator on the unit sphere.

5. CONCLUDING REMARKS

We have described an implicit finite element approximation for the grazing collision Boltzmann-Lorentz operator. This approximation allows to pass to the limit " ! 0 for a fixed discrete scheme (with respect to the velocity variable). This fact is due essentially to the possibility of carrying out the exact computations of both the mass matrix and the collision matrix with no need of quadrature formulas. Other methods would have been possible: for example, a finite volume method based on a piecewise parabolic operator or a spectral method (see Buet et al.

(2001), Pareschi et al. (2002)). Our choice of a FEM method has been essentially motivated by the nature of the limiting problem and also by previous computations done for the Laplace-Beltrami operator in 3D.

The generalization to the three-dimensional case seems to be possi- ble: we can in fact compute the matrices A and Q, but the coefficients we find are much more complicated. According to Buet et al. (2001),

F

0 k 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

8 6 4 2 0 2 4 6 8 10

10

Figure 9. The focalization F ð; t ¼ 150tÞ. ¼ 2

k

; k ¼ 10; . . . ; 10.

757

758

759

760

761

762

763

764

765

766

767

768

769

770

771

772

773

774

775

776

777

778

779

780

781

782

783

784

785

786

787

788

789

790

791

792

793

794

795

796

797

798

(22)

n(x,T)

x

1.0 0.0 0.2 0.4 0.6 0.8 1.0

1.40 1.26 1.12 0.98 0.84 0.70 0.56 0.42 0.28 0.14

0.00 0.8 0.6 0.4 0.2

25 iterations 50 iterations

100 iterations 200 iterations

n(x,T)

x

0.0 0.2 0.4 0.6 0.8 1.0

1.26 1.40

1.12 0.98 0.84 0.70 0.56 0.42 0.28 0.14 0.00

0.8 0.6 0.4 0.2

1.0

25 iterations 50 iterations

100 iterations 200 iterations n(x,T)

x

0.2 0.4 0.6 0.8 1.0

1.40 1.26 1.12 0.98 0.84 0.70 0.56 0.42 0.28 0.14 0.00

1.0 0.8 0.6 0.4 0.2 0.0

25 iterations 50 iterations

100 iterations 200 iterations

Figure 10. The density nð; TÞ, ¼ 2

5

(left),1 (center) , 2

5

(right), T ¼ N

iter

t with N

iter

¼ 25; 50; 100; 200.

799

800

801

802

803

804

805

806

807

808

809

810

811

812

813

814

815

816

817

818

819

820

821

822

823

824

825

826

827

828

829

830

831

832

833

834

835

836

837

838

839

840

(23)

a spectral method should be a good idea for the numerical approximation of the 3D grazing collision Lorentz operator.

It should be also interesting to generalize the approximation presented in this article to nonconstant grazing collisions scattering cross sections and also to the Coulombian case (we refer to Cordier et al. (2002) for a numerical comparison between the Boltzmann-Lorentz operator and the Fokker-Planck one in the Coulombian case).

A1. APPENDIX

Here we perform the computation of the coefficients Q

ij

of the graz- ing collision matrix Q. We first express the hat functions

i

as follows:

i

ðÞ ¼ ’

h

ð

i

Þ, where:

h

ðzÞ ¼ h

1

ðh jzjÞ, if jzj < h,

0 else:

Replacing in Eq. (3) ¼

i

and

0

¼

0

j

yields:

Q

ij

¼ Z

S1

Z

S1

B

"

ð

0

þ Þ ½’

h

ðÞ ’

h

ð

0

Þ ’

h

ðÞ ’

h

ð þ Þ d d

0

, where we have set ¼

i

j

¼ ði jÞh. Introducing the new hat function

’ðzÞ ¼ ð1 jzjÞ, if jzj < 1,

0 else

n

we have ’

h

ðzÞ ¼ ’ðz=hÞ. Now using the new variables ¼ =h and

0

¼

0

=h, we can write:

Q

ij

¼ h

2

Z

þ1

1

Z

þ1 1

B

"

ðh h

0

þ Þ ½’ðÞ ’ð

0

Þ ’ðÞ ’ð þ ði jÞÞ d d

0

¼ h

2

Z

þ1

1

Z

þ1 1

B

"

ðhð

0

þ ði jÞÞÞ ½’ðÞ ’ð

0

Þ ’ðÞ ’ð þ ði jÞÞ d d

0

:

Let us set B

h

ðzÞ ¼ BðhzÞ, i.e., (with B

0

¼ 1="

3

):

B

h

ðzÞ ¼ B

0

if jzj < ¼ "=h, 0 else;

841

842

843

844

845

846

847

848

849

850

851

852

853

854

855

856

857

858

859

860

861

862

863

864

865

866

867

868

869

870

871

872

873

874

875

876

877

878

879

880

881

882

(24)

then the collision coefficient Q

ij

reads:

Q

ij

¼ h

2

Z

þ1

1

Z

þ1 1

B

h

ð

0

þ ði jÞÞ ½’ðÞ ’ð

0

Þ ’ðÞ ’ð þ ði j ÞÞ d d

0

¼ h

2

B

0

Z Z

j0þðijÞj<

½’ðÞ ’ð

0

Þ ’ðÞ ’ð þ ði jÞÞ d d

0

with ,

0

2 ½1, þ 1. We split Q

ij

into a gain part G

ij

and a loss part L

ij

, i.e., we set Q

ij

¼ G

ij

L

ij

, where:

G

ij

¼ h

2

B

0

Z Z

j0þðijÞj<

’ðÞ ’ð

0

Þ d d

0

, L

ij

¼ h

2

B

0

Z Z

j0þðijÞj<

’ðÞ ’ð þ ði jÞÞ d d

0

:

We now have to compute separately these gain and loss terms. For this, we will assume that 1, so that " h= ¼ 2=N : it is thus possible to pass to the limit first for " ! 0 and then for h ! 0.

Let us first compute the gain coefficients G

ij

. As the ’ function is null outside a domain of length 2, it is clear that for ji jj 3, G

ij

¼ 0 (see also Fig. 11). Thus, we have to compute G

ij

for i ¼ j, ji jj ¼ 1, and ji jj ¼ 2. Considering also the symmetry of the collision operator, we only have to compute the integrals for i j ¼ 2, i j ¼ 1, and i ¼ j.

θ θ

θ θ θ θ

θ

’= − 1

’ = − 2

’ =

−β β

θ’

Figure 11. Set of definition of B

h

ð

0

þ i jÞ:

F11

883

884

885

886

887

888

889

890

891

892

893

894

895

896

897

898

899

900

901

902

903

904

905

906

907

908

909

910

911

912

913

914

915

916

917

918

919

920

921

922

923

924

(25)

Looking at Fig. 12, we define the following integrals for a given > 0:

IðÞ ¼ h

2

B

0

Z

0

ð1 þ Þ

Z

þ1 1

ð1 þ

0

Þ d

0

d þ h

2

B

0

Z

1 0

ð1 Þ

Z

þ1 1

ð1 þ

0

Þ d

0

d þ h

2

B

0

Z

1 1

ð1 Þ Z

0

1

ð1 þ

0

Þ d

0

d þ h

2

B

0

Z

1 1

ð1 Þ

Z

þ1 0

ð1

0

Þ d

0

d

¼ h

2

B

0

1

12

3

ð 4Þ þ 1

24 ð1 Þð1 þ Þð

2

þ 4 þ 1Þ þ 1 4

2

¼ h

2

B

0

1 8

4

þ 1

6

3

þ 1 4

2

þ 1

6 þ 1 24

and

J ð Þ ¼ h

2

B

0

Z

1

ð1 Þ

Z

1 1

ð1 þ

0

Þ d

0

d

¼ h

2

B

0

1

24 ð 1Þ

4

θ

−β

I(−β)

β β

θ’

J( )

Figure 12. The integrals I ðÞ and JðÞ.

F12

925

926

927

928

929

930

931

932

933

934

935

936

937

938

939

940

941

942

943

944

945

946

947

948

949

950

951

952

953

954

955

956

957

958

959

960

961

962

963

964

965

966

Références

Documents relatifs

Besides the dimension and the volume of the sub- manifold and the order of the eigenvalue, these bounds depend on either the maximal number of intersection points of M with a p-

Druet [1] has proved that for small volumes (i.e. τ &gt; 0 small), the solutions of the isoperimetric problem are close (in a sense to be made precise) to geodesic spheres of

Barrett and S¨ uli have introduced, and proved the existence of global-in-time weak solutions for d = 2 and 3 to, microscopic-macroscopic dumbbell models of dilute polymers

It is known that the discrete and semi-discrete models obtained by dis- cretizing the wave equation with the classical finite difference or finite element method do not

In all of these cases, the corresponding classical system has a set of initial conditions with non-zero measure for which the orbits are bounded but to which no

Elliott, On the finite element approximation of an elliptic variational inequality arising from an implicit time discretization of the Stefan problem.. Elliott, Error analysis of

In this paper we considered the edge finite element approximation of the Maxwell resolvent operator and we proved the uniform convergence in the L 2 (Ω) 3 -norm.. The proof

Also, this section strives to critically review techniques designed to boost the L 2 -convergence rate without excessive additional computational effort An a priori and an a