• Aucun résultat trouvé

KANIEL-SHINBROT ITERATION AND GLOBAL SOLUTIONS OF THE CAUCHY PROBLEM FOR THE BOLTZMANN EQUATION

N/A
N/A
Protected

Academic year: 2022

Partager "KANIEL-SHINBROT ITERATION AND GLOBAL SOLUTIONS OF THE CAUCHY PROBLEM FOR THE BOLTZMANN EQUATION"

Copied!
25
0
0

Texte intégral

(1)

OF THE CAUCHY PROBLEM FOR THE BOLTZMANN EQUATION

CLAUDE BARDOS, IRENE GAMBA, AND C. DAVID LEVERMORE

Abstract. Kaniel-Shinbrot iteration can be used to establish the existence of solutions to the Boltzmann equation over the spatial domain RD for some initial data that is pointwise bounded above and below by a class of global Maxwellians larger than previously considered.

When these solutions are global in time, we obain results on their large-time asymptotics.

1. Introduction

We consider a kinetic density F(v, x, t) over the velocity-position domain RD×RD that is governed by the Cauchy problem for the Boltzmann equation:

(1.1) ∂tF +v·∇xF =B(F, F), F!

!t=0 =Fin.

We assume that the initial data Fin(v, x) is nonnegative and satisfies the bounds

(1.2) 0<

" "

RD×RD

#1 +|v|2+|x|2$

Findvdx <∞. The collision operatorB(F, F) has the form

(1.3) B(F, F) =

" "

SD−1×RD

(F"F"−FF) bdωdv,

where ω ∈ SD−1, b(ω, v −v) is the collision kernel, while F, F", and F" denote F(·, x, t) evaluated at v, v" =v−ω ω·(v−v), andv" =v+ω ω·(v−v) respectively.

1.1. Collision Kernels. We will assume that the collision kernel has the separable form (1.4) b(ω, v−v) = ˆb(ω·n)|v−v|β, where n= v−v

|v−v|,

for someβ ∈(−D,2) while ˆb(ω·n) is positive almost everywhere and satisfies the weak small- deflection cutoff condition

(1.5) %

%ˆb%

%L1(dω) =

"

SD−1

ˆb(ω·n) dω <∞.

The conditions β > −D and (1.5) are required for b(ω, v −v) given by (1.4) to be locally integrable with respect to dωdv. They allow us to decompose the collision operator (1.3) as (1.6) B(F, F) =G(F, F)−A(F)F ,

where the attenuation operator A(F) and the gain operatorG(F, F) are defined by (1.7) A(F) =

" "

SD−1×RD

Fbdωdv, G(F, F) =

" "

SD−1×RD

F"F"bdωdv.

Notes on our join work started at ICERM, draft December 4, 2011.

1

(2)

The form (1.4) arises from the classical scattering cross section calculation for identical hard spheres of massm and diameter do, which yields

(1.8) b(ω, v−v) = doD−1

2m |ω·(v−v)|. This corresponds to the case β = 1 and ˆb(ω·n) = do2mD−1 |ω·n| in (1.4).

The form (1.4) also arises from the classical scattering cross-section calculation for an inverse- power interparticle potential after a small-deflection cutoff is imposed. For potentials propor- tional to r−k, where r is the distance between the center of masses, one has

(1.9) β= 1−2D−1

k for k > 2D−1 D + 1.

The cases β ∈ (−D,0), β = 0, β ∈ (0,1], and β ∈ (1,2) are called respectively the “soft”,

“Maxwell”, “hard”, and “super-hard” cases. The super-hard cases do not arise from an inverse- power interparticle potential.

1.2. Conservation Laws. Any solution F of the Boltzmann equation (1.1) formally satisfies the local conservation law

t

"

RD

ξF dv+∇x·

"

RD

vξF dv = 0, when ξ(v, x, t) is any quantity that satisfies

(1.10) ξ(·, x, t)∈Span{1, v1, v2,· · · , vD,|v|2}, ∂tξ+v·∇xξ = 0.

It has been known essentially since Boltzmann [4], who worked out the case D = 3, that the only such quantities ξ are linear combinations of the 4 + 2D + D(D−1)2 quantities

(1.11) 1, v , x−vt , 12|v|2, v∧x v·(x−vt), 12|x−vt|2,

where v ∧x = v xT −x vT is the skew tensor product. By integrating the coresponding local conservation laws over space and time, we formally obtain the global conservation laws

(1.12)

" "

RD×RD

 1 v x−vt

1 2|v|2 v∧x v·(x−vt)

1

2|x−vt|2

F(v, x, t) dvdx=

" "

RD×RD

 1 v x

1 2|v|2 v∧x v·x

1 2|x|2

Fin(v, x) dvdx ,

where the right-hand sides above exist by the bounds (1.2). These conserved quantities are associated respectively with the conservation laws of mass, momentum, initial center of mass, energy, angular momentum, scalar momentum moment, and scalar inertial moment. The last two are not general physical laws, but are shared by the solutions of other kinetic equations.

1.3. Global Maxwellians. One consequence of Boltzmann’s celebratedH-theorem [4] is that if f(v) is any nonnegative integrable function that has appropriate behavior as |v| → ∞then B(f, f) = 0 if and only if

(1.13) f = ρ

(2πθ)D2 exp ,

− |v−u|2

-

for some (ρ, u,θ)∈R+×RD×R+.

(3)

Functions of the form (1.13) where ρ, u, and θ functions of (x, t) are called local Maxwellians.

Any local Maxwellian that also satisfies the kinetic equation (1.1) is called aglobal Maxwellian.

The family of all global Maxwellians over the spatial domain RD with positive mass, zero net momentum, and center of mass at the origin has the form

(1.14a) M= m

(2π)D

.det(Q) exp#

−q(v, x, t)$

, Q= (ac−b2)I+B2,

(1.14b) q(v, x, t) = 1 2

, v

x−vt -T ,

cI bI +B bI−B aI

- , v

x−vt -

,

with m >0 and (a, b, c, B)∈Ω whereΩ is the open cone in R×R×R×RD∧D defined by

(1.15) Ω=/

(a, b, c, B)∈R+×R×R+×RD∧D : (ac−b2)I +B2 >00 .

HereR+ denotes the positive real numbers and RD∧D denotes the skew-symmetric D×D real matrices. The form (1.14) can be derived from the fact that log(M) must satisfy

log(M)∈Span{1, v1, v2,· · · , vD,|v|2}, #

t+v·∇x$

log(M) = 0,

whereby log(M) must be a linear combination of the quantities (1.11). The form (1.14) then comes from the requirements thatMhave finite mass, zero net momentum, and center of mass at the origin. The larger family of global Maxwellians with positive mass is obtained from (1.14) by introducing translations inv and x; whereby it has a total of 4 + 2D + D(D−1)2 parameters.

Remark. The positive definiteness condition (1.15) can be thought of as a bound on b and B in terms of ac by expressing it as

(1.16) b2+(B(2 < ac ,

where (B( is the(2-norm, which equals the spectral radius for the skew-symmetric matrix B.

We can bring Minto the local Maxwellian form (1.13). By expanding (1.14b) we find that q(v, x, t) = 12#

at2−2bt+c$

|v|2−#

axt−bx−Bx$

·v+12a|x|2.

Because a, c >0 and ac > b2 by the remark above, we know that at2−2bt+c > 0 for everyt.

We can therefore complete the square in v to obtain q(v, x, t) = 1

2θ(t)|v−u(x, t)|2+ θ(t)

2 xTQx , where θ(t) and u(x, t) are given by

(1.17) θ(t) = 1

at2−2bt+c, u(x, t) =θ(t)#

axt−bx−Bx$ . We thereby see that Mhas the local Maxwellian form

(1.18) M= ρ(x, t)

#2πθ(t)$D2 exp ,

−|v−u(x, t)|2 2θ(t)

- ,

where the temperatureθ(t) and bulk velocityu(x, t) are given by (1.17), while the mass density ρ(x, t) is given by

(1.19) ρ(x, t) =m

,θ(t) 2π

-D2

.det(Q) exp ,

−θ(t) 2 xTQx

- . Because Q >0 by (1.15), we see that ρ(x, t) is integrable over RD.

(4)

2. Properties of Global Maxwellians

2.1. Moments of Global Maxwellians. We can evaluate the zeroth, first, second, and any other moment of the global Maxwellian familyMbecause they are Gaussian densities over the velocity-position space RD×RD. We will use some basic facts about matrices in the form

(2.1) A=

, cI bI+B

bI−B aI -

, where (a, b, c, B)∈R×R×R×RD∧D.

Specifically, if c)= 0 then the matrices A and Q= (ac−b2)I+B2 are related by the facts

(2.2)

(i) det(A) = det(Q),

(ii) A is invertible ⇐⇒ Qis invertible, in which case A−1 =

,Q−1 0 0 Q−1

- , aI −bI −B

−bI +B cI -

, (iii) A ≥0 ⇐⇒ Q≥0 and c >0,

(iv) A >0 ⇐⇒ Q >0 and c >0 ⇐⇒ (a, b, c, B)∈Ω. These facts can all be seen from the block LDU factorization

A=

, I 0

1

c(bI −B) I

- ,cI 0 0 1cQ

- ,I 1c(bI +B)

0 I

- .

This factorization holds for anyAof the form (2.1) withc)= 0 andB ∈RD×D. Fact (ii) follows from fact (i). The formula for A−1 can be checked by direct computation. Facts (iii) and (iv) require that B ∈ RD∧D, which insures that A and Qare symmetric. Fact (iv) shows that Ω is a cone. Because it is defined by strict inequalites, Ω⊂R+×R×R+×RD∧D is open.

By evaluating the Gaussian integral and using facts (i) and (iv) of (2.2) we see that

" "

RD×RDMdvdx=

"

RD

ρ(x, t) dx=m . Hence, m is the total mass ofM. Because Mis an even function of #

v x−vt$T

, we see that

" "

RD×RD

, v

x−vt -

Mdvdx= ,0

0 -

.

This reflects the facts thatMhas no net momentum and that its center of mass is at the origin.

In fact, every odd-order moment will vanish because M has even symmetry.

The second moments of M may be computed by evaluating the Gaussian integrals and by using the formula forA−1 given in fact (ii) of (2.2) as

(2.3)

" "

RD×RD

, v vT v(x−vt)T

(x−vt)vT (x−vt) (x−vt)T -

Mdvdx

=

" "

RD×RD

, v

x−vt

- , v

x−vt -T

Mdvdx

=mA−1 =m

, aQ−1 −bQ−1−Q−1B

−bQ−1+Q−1B cQ−1

- .

(5)

In particular, the values of the quadratic conserved quantities from (1.12) are given by

(2.4)

" "

RD×RD|v|2Mdvdx=mtr# Q−1$

a ,

" "

RD×RD

v·(x−vt)Mdvdx=−mtr# Q−1$

b ,

" "

RD×RD|x−vt|2Mdvdx=mtr# Q−1$

c ,

" "

RD×RD

v∧xMdvdx=−2mQ−1B .

Every even-order moment of Mmay also be computed, but we will not do so here.

2.2. Conserved Quantities and Global Maxwellians. Every Cauchy problem (1.1) with initial data Fin(v, x) that satisfies the bounds (1.2) can be associated with a unique global Maxwellian determined by the values of the conserved quantities computed from Fin. By choosing an appropriate rescaling and Galilean frame, we may assume without loss of generality that

(2.5)

" "

RD×RD

Findvdx= 1,

" "

RD×RD

v Findvdx=

" "

RD×RD

x Findvdx= 0. The following theorem was proved in [8].

Theorem 2.1. Let Fin(v, x) be a nonnegative function that satisfies the bounds (1.2) and the normalizations (2.5). Let a, b, c, and B be given by

(2.6)

" "

RD×RD|v|2Findvdx=a,

" "

RD×RD

v·x Findvdx=b,

" "

RD×RD|x|2Findvdx=c,

" "

RD×RD

v∧x Findvdx=B. Then(a, b, c, B)∈Ω, where Ω is the open cone in R×R×R×RD∧D defined by (2.7) Ω =1

(a, b, c, B)∈R+×R×R+×RD∧D : 12tr(|B|)<.

ac−b22 , with |B|=.

BTB =.

−B2.

Conversely, if (a, b, c, B) ∈ Ω then there exists a unique (up to a time shift) global MaxwellianMgiven by (1.14) withm= 1 and(a, b, c, B)∈Ωsuch that the quadratic converved quantities associated with M through (2.4) have values (a, b, c, B) — i.e. such that

(2.8) a = tr# Q−1$

a , b =−tr# Q−1$

b , c = tr# Q−1$

c , B =−2Q−1B , where Q= (ac−b2)I+B2 and the set Ω was defined by (1.15).

Remark. The first part of this theorem states that Ω contains the set of values that can be realized by the conserved quantities given by (2.6). The second part asserts that every point inΩ can be so realized. Therefore Ω characterizes all such values.

Remark. Because B is skew-symmetric, its nonzero eigenvalues will come in conjugate pairs.

When B has at most two nonzero eigenvalues then (B( = 12tr(|B|), where (B( is the (2 matrix norm, which equals the spectral radius of B. If D = 2 or D = 3 then B can have at most two nonzero eigenvalues and we see by (1.16) that Ω = Ω. If D > 3 then Ω will be a proper subset ofΩbecause (B(< 12tr(|B|) when B has more than two nonzero eigenvalues.

(6)

2.3. Ordering Global Maxwellians. The following lemma characterizes when one global Maxwellian with finite mass bounds another that has the same net momentum and center of mass. By choosing an appropriate Galilean frame, we can assume without loss of generality that the net momentum is zero and the center of mass is the origin. In that case the global Maxwellians will have the form (1.14).

Lemma 2.1. Let M1 and M2 be global Maxwellians of the form (1.14) with parameters given by (m1, a1, b1, c1, B1)∈R+×Ω and (m2, a2, b2, c2, B2)∈R+×Ω respectively. Then M1 ≤M2 for every (v, x, t) if and only if

, c2I b2I+B2

b2I−B2 a2I -

, c1I b1I +B1

b1I −B1 a1I -

. (2.9)

m1

3 det##

a1c1−b12$

I+B12$

≤m2

3 det##

a2c2−b22$

I+B22$ , (2.10)

Similarly, M1 <M2 for every (v, x, t) if and only if (2.9) holds and

(2.11) m1

3 det##

a1c1−b12$

I+B12$

< m2

3 det##

a2c2−b22$

I+B22$ , Proof. It follows from (1.14) that

log ,M2

M1 -

= log

 m23

det##

a2c2 −b22$

I +B22$ m13

det##

a1c1 −b12$

I +B12$

+ 1 2

, v

x−vt -T ,

(c1 −c2)I (b1 −b2)I+ (B1−B2) (b1−b2)I−(B1−B2) (a1 −a2)I

- , v

x−vt -

.

The assertions of the lemma can be read off from this. !

Remark. The matrix inequality (2.9) is equivalent to a2 ≤a1,c2 ≤c1, and

(2.12) #

(a1−a2)(c1 −c2)−(b1−b2)2$

I+ (B1−B2)2 ≥0.

This positive definiteness condition can be thought of as a bound on b1−b2 and B1 −B2 in terms of (a1−a2)(c1−c2) by expressing it as

(b1 −b2)2+(B1−B2(2 <(a1−a2)(c1−c2),

where (B1−B2( is the matrix (2-norm, which equals the spectral radius for B1−B2. Remark. The matrix inequality (2.9) implies that

3 det##

a2c2−b22$

I+B22$

≤ 3

det##

a1c1−b12$

I+B12$ , with equality if and only if

, c2I b2I +B2

b2I −B2 a2I -

=

, c1I b1I +B1

b1I−B1 a1I -

. Therefore ifM1 ≤M2 then m1 < m2 unless M1 =M2.

Remark. In this section we assumed that M1 and M2 have the same net momentum and the same center of mass. It is clear that if M1 ≤ M2 then M1 and M2 must have the same net momentum, but they need not have the same same center of mass. Allowing for different centers of mass does not change condition (2.9), but does modify condition (2.10).

(7)

2.4. Bounds by Global Maxwellians. Let M1 and M2 be global Maxwellians of the form (1.14) with parameters given by (m1, a1, b1, c1, B1) ∈R+×Ω and (m2, a2, b2, c2, B2)∈R+×Ω respectively such thatM1 ≤M2 for every (v, x, t). Let M1(t) and M2(t) denote their values at timet. Let Fin(v, x) satisfy the pointwise bounds

(2.13) M1(s)< Fin<M2(s) over RD×RD for some s∈R. and the normalizations

(2.14)

" "

RD×RD

Findvdx= 1,

" "

RD×RD

v Findvdx=

" "

RD×RD

x Findvdx= 0. We would like to show that under certain conditions there exists a unique solutionF(v, x, t) of the kinetic equation posed over the spatial domain RD with initial data Fin(v, x).

In addition, we hope to show that there exists a global Maxwellian M of the form (1.14) with m= 1 and (a, b, c, B)∈Ω such that

F(v, x, t)∼M(v, x, t) as t→ ∞ in some topology. If we let (a, b, c, B)∈Ω be given by

(2.15)

" "

RD×RD|v|2Findvdx=a,

" "

RD×RD

v·x Findvdx =b,

" "

RD×RD

|x|2Findvdx=c,

" "

RD×RD

v∧x Findvdx =B, then by Theorem 2.1 there exists (a, b, c, B)∈Ωsuch that

(2.16) a = tr# Q−1$

a , b =−tr# Q−1$

b , c = tr# Q−1$

c , B =−2Q−1B , where Q= (ac−b2)I+B2.

The global Maxwellian bounds (2.13) implies that (a, b, c, B) can be bounded by the corresponding quantities associated withmin1 Min1 and min2Min2, which are

(2.17)

a1∗ =m1tr# Q−11 $

a1, a2∗ =m2tr# Q−12 $

a2, b1∗ =−m1tr#

Q−11 $

b1, b2∗ =−m2tr# Q−12 $

b2, c1∗ =m1tr#

Q−11 $

c1, c2∗ =m2tr# Q−12 $

c2, B1∗ =−2m1Q−11 B1, B2∗ =−2m2Q−12 B2,

where Q1 = (a1c1−b12)I+B12 and Q2 = (a2c2−b22)I +B22. Theorem 2.1 implies that

(2.18)

#a −a1∗, b −b1∗, c−c1∗, B−B1∗

$∈Ω,

#a2∗−a, b2∗−b, c2∗−c, B2∗ −B

$∈Ω,

#a2∗ −a1∗, b2∗−b1∗, c2∗ −c1∗, B2∗−B1∗$

∈Ω, which yields the bounds

(2.19)

a1∗ < a < a2∗, c1∗ < c < c2∗,

1

2tr(|B−B1∗|)<.

(a −a1∗)(c−c1∗)−(b−b1∗)2,

1

2tr(|B2∗−B|)<.

(a2∗−a)(c2∗−c)−(b2∗−b)2,

1

2tr(|B2∗−B1∗|)<.

(a2∗−a1∗)(c2∗−c1∗)−(b2∗ −b1∗)2.

(8)

3. Kaniel-Shinbrot Iteration

3.1. Kaniel-Shinbrot Theorem. Kaniel-Shinbrot iteration can be used to prove the existence of solutions F to the Boltzmann initial-value problem posed over the spatial domain RD with initial data Fin(v, x) that satisfies certain pointwise bounds. More specifically, it is used to construct two sequences of approximate solutions, /

FjL0

j∈N and / FjU0

j∈N, the first of which converges monotonically toF from below, while the second converges monotonically toF from above. In other words, for some T ∈(0,∞] these opposing monotone sequences should satisfy the order relationships

(3.1) FjL≤Fj+1L ≤Fj+1U ≤FjU over RD×RD×[0, T) for every j ∈N.

The expectation is that these sequences will converge toF(v, x, t) over RD×RD×[0, T). When this construction can be carried out forT =∞then the solution F will be global in time.

Kaniel-Shinbrot iteration constructs sequences as follows. GivenFj−1L andFj−1U for anyj ∈Z+ we define the Kaniel-Shinbrot iteratesFjL and FjU to be the solution of the linear system

tFjU+v·∇xFjU+A# Fj−1L $

FjU =G#

Fj−1U , Fj−1U $ ,

tFjL+v·∇xFjL+A# Fj−1U $

FjL=G#

Fj−1L , Fj−1L $ (3.2a) ,

FjU!

!t=0 =FjL!

!t=0 =Fin. (3.2b)

The existence of FjL and FjU can be inferred from the mild formulation of system (3.2).

The form of the Kaniel-Shinbrot iteration (3.2) is motivated by the fact that A and G have the monotonicity properties

(3.3) F ≤G =⇒ A(F)≤A(G) and G(F, F)≤G(G, G).

The following lemma shows that this form insures that Kaniel-Shinbrot iterates preserve the order relationships that appear in (3.1).

Lemma 3.1. (Order Preservation Lemma) If for some T ∈ (0,∞] the Kaniel-Shinbrot iterates Fj−1L , Fj−1U , FjL, andFjU satisfy

(3.4) Fj−1L ≤FjL≤FjU ≤Fj−1U over RD×RD×[0, T), then the Kaniel-Shinbrot iterates Fj+1L and Fj+1U satisfy

(3.5) FjL≤Fj+1L ≤Fj+1U ≤FjU over RD×RD×[0, T). Proof. By (3.2a) of the Kaniel-Shinbrot construction we see that

#∂t+v·∇x +A#

FjU$$ #

Fj+1L −FjL$

=G#

FjL, FjL$

−G#

Fj+1L , Fj+1L $ +A#

Fj−1U −FjU$ FjL,

#∂t+v·∇x +A#

FjL$$ #

FjU −Fj+1U $

=G#

Fj−1U , Fj−1U $

−G#

FjU, FjU$ +A#

FjL−Fj−1L $ FjU. By the monotonicity (3.3) of A and G and hypothesis (3.4) the right-hand sides above are nonnegative over RD×RD× [0, T), while (3.2b) of the Kaniel-Shinbrot construction implies that

#Fj+1L −FjL$ !

!t=0 =#

FjU−Fj+1U $ !

!t=0 = 0, over RD×RD. The maximum principle then implies that

(3.6) #

Fj+1L −FjL$

≥0, #

FjU−Fj+1U $

≥0, over RD×RD×[0, T). This yields two of the three inequalities asserted by (3.5).

(9)

Similarly, by (3.2a) of the Kaniel-Shinbrot construction we have

#∂t+v·∇x+A#

FjU$$ #

Fj+1U −Fj+1L $

=G#

FjU, FjU$

−G#

FjL, FjL$ +A#

FjU −FjL$ Fj+1U . By the monotonicity (3.3) of A and G and hypothesis (3.4) the right-hand side above is non- negative over RD×RD × [0, T), while by (3.2b) of the Kaniel-Shinbrot construction we see that

#Fj+1U −Fj+1L $ !

!t=0 = 0, over RD×RD. The maximum principle then implies that

#Fj+1U −Fj+1L $

≥0, over RD×RD×[0, T).

When this inequality is combimed with the two in (3.6), we see that assertion (3.5) holds. ! Induction, Lemma 3.1, the Lebesgue Monotone Convergence Theorem, and a stability bound can be used [7, 6] to prove the following.

Theorem 3.1. (Kaniel-Shinbrot) If for some T ∈ (0,∞] the Kaniel-Shinbrot iterates F0L, F0U, F1L, and F1U satisfy the so-called beginning condition

(3.7) F0L≤F1L≤F1U ≤F0U over RD×RD×[0, T), then the Kaniel-Shinbrot iteration yields opposing monotone sequences/

FjL0

j∈N and / FjU0

j∈N

overRD×RD×[0, T)— i.e. sequences that satisfy the order relationship (3.1). These sequences converge to a unique mild solution of the initial-value problem (1.1) for the Boltzmann equation.

Remark. The hard part of applying the Kaniel-Shinbrot Theorem is being sure that the first two iterates satisfy the beginning condition (3.7).

3.2. Beginning with Local Maxwellians. When the initial Kaniel-Shinbrot iterates are local Maxwellians then there is a simple criterion that insures the beginning condition (3.7) is satisfied for some T ∈(0,∞].

Proposition 3.1. (Local Maxwellian Beginning Lemma) Suppose ML and MU are local Maxwellians that satisfy

(3.8a) ML!

!t=0 ≤MU!

!t=0 over RD×RD, and for some T ∈(0,∞] satisfy

(3.8b) ∂tMU +v·∇xMU ≥A(MU −ML)MU over RD×RD×[0, T),

−∂tML−v·∇xML≥A(MU −ML)ML over RD×RD×[0, T), Then for every initial data Fin such that

ML!

!t=0 ≤Fin≤MU!

!t=0 over RD×RD,

the Kaniel-Shinbrot iterates obtained by setting F0L=ML and F0U =MU satisfy the beginning condition (3.7) over[0, T). Moreover, the Kaniel-Shinbrot Theorem (3.1) yields the unique mild solution F(v, x, t) of the initial-value problem (1.1) for the Boltzmann equation that satisfies the bounds

(3.9) ML(v, x, t)≤F(v, x, t)≤MU(v, x, t) over RD×RD×[0, T).

(10)

Proof. Because F0L=ML and F0U =MU, the Kaniel-Shinbrot construction (3.2) yields

tF1U +v·∇xF1U+A(ML)F1U =G(MU, MU),

tF1L+v·∇xF1L+A(MU)F1L=G(ML, ML), F1U!

!t=0 =F1L!

!t=0 =Fin.

The above right-hand sides satisfy G(ML, ML) = A(ML)ML and G(MU, MU) = A(MU)MU because ML and MU are local Maxwellians. Therefore

#∂t+v·∇x+A(ML)$ #

MU−F1U$

=∂tMU +v·∇xMU−A(MU −ML)MU,

#∂t+v·∇x+A(MU)$ #

F1L−ML$

=−∂tML−v·∇xML−A(MU −ML)ML.

By hypothesis (3.8b) the right-hand sides above are nonnegative over RD×RD×[0, T), while by hypothesis (3.8a)

#MU −F1U$ !

!t=0 ≥0, #

F1L−ML$ !

!t=0 ≥0, over RD×RD. The maximum principle then implies that

(3.10) #

MU −F1U$

≥0, #

F1L−ML$

≥0, over RD×RD×[0, T).

This yields two of the three inequalities that are required by the beginning condition (3.7).

Similarly, we have

#∂t+v·∇x+A(MU)$ #

F1U−F1L$

=G(MU, MU)−G(ML, ML) +A(MU −ML)F1U. The right-hand side above is nonnegative by the monotonicity (3.3) ofA and G, while we see by (3.2b) of the Kaniel-Shinbrot construction that

#F1U−F1L$ !

!t=0 = 0 overRD×RD. The maximum principle then implies that

#F1U −F1L$

≥0 over RD×RD×[0, T).

When this inequality is combined with the two in (3.10), we see that the beginning condition (3.7) is satisfied. The Kaniel-Shinbrot Theorem (3.1) thereby yields a unique mild solution of the initial-value problem (1.1) for the Boltzmann equation that satisfies (3.9). ! Proposition 3.1 requires us to find local Maxwellians ML and MU that meet criterion (3.8) for some T ∈(0,∞]. When ML >0 this criterion may be recast as

#∂t+v·∇x$ log#

MU$

≥A#

MU−ML$

over RD×RD×[0, T),

−#

t+v·∇x$ log#

ML$

≥A#

MU−ML$

over RD×RD×[0, T), (3.11a)

ML!

!t=0 ≤MU!

!t=0 over RD×RD. (3.11b)

When ML= 0 criterion (3.8) reduces to simply

(3.12) #

t+v·∇x$

log(MU)≥ A# MU$

over RD×RD×[0, T).

We will construct such local Maxwellians from the family of global Maxwellians given by (1.14).

When we can do this with T =∞, the Kaniel-Shinbrot theorem will yield global solutions.

(11)

3.3. Constructing Local Maxwellians from Global Ones. Let M1 and M2 be global Maxwellians in the form (1.14) that are given by parameters (m1, a1, b1, c1, B1)∈ R+×Ω and (m2, a2, b2, c2, B2) ∈ R+ ×Ω respectively, and that satisfy M1 ≤ M2 for every (v, x, t). Let M1(t) andM2(t) denote their values at time t. Set

(3.13)

Q1 = (a1c1−b12)I +B12, Q2 = (a2c2−b22)I+B22, q1(v, x, t) = 1

2

, v

x−vt -T,

c1I b1I+B1

b1I−B1 a1I

- , v

x−vt -

.

q2(v, x, t) = 1 2

, v

x−vt -T,

c2I b2I+B2 b2I−B2 a2I

- , v

x−vt -

, so that

M1 = m1

(2π)D

.det(Q1) exp#

−q1(v, x, t)$

, M2 = m2

(2π)D

.det(Q2) exp#

−q2(v, x, t)$ . Because M1 ≤M2, we know thatq1(v, x, t)≥q2(v, x, t).

We will construct local Maxwellians MU and ML in the form MU(v, x, t) =γ(t) m2

(2π)D

.det(Q2) exp ,

−q2(v, x, s+t) η(t)

- , (3.14a)

ML(v, x, t) = 1 γ(t)

m1

(2π)D

.det(Q1) exp ,

− ,

2− 1 η(t)

-

q1(v, x, s+t) -

, (3.14b)

where the functionsγ(t) and η(t) satisfy

γ(0) =η(0) = 1, γ"(t)>0, η"(t)≥0. Clearly, inequality (3.11b) is satisfied because

ML!

!t=0 =M1(s)≤M2(s) = MU!

!t=0. By direct calculation we find that

#∂t+v·∇x$ log#

MU$

= γ"(t)

γ(t) +η"(t) η(t)

q2(v, x, s+t) η(t) ,

−#

t+v·∇x$ log#

ML$

= γ"(t)

γ(t) +η"(t) η(t)

q1(v, x, s+t) η(t) . Because q1(v, x, s+t)≥q2(v, x, s+t), we see that

−#

t+v·∇x$ log#

ML$

≥#

t+v·∇x$ log#

MU$ , therefore criterion (3.11) will be satisfied if

(3.15) γ"(t)

γ(t) + η"(t) η(t)

q2(v, x, s+t) η(t) ≥A#

MU −ML$ .

Similarly, if ML= 0 while MU is given by (3.14a) then criterion (3.12) will be satisfied if

(3.16) γ"(t)

γ(t) + η"(t) η(t)

q2(v, x, s+t) η(t) ≥A#

MU$ .

This extends the basic inequality derived in [7] to a more general class of local Maxwellians.

We can treat criterion (3.16) as criterion (3.15) with ML = 0,

(12)

The right-hand sides of (3.15) and (3.16) contain expressions of the form A(M) where M is some local Maxwellian. If ρ(x, t), u(x, t), and θ(x, t) are the mass density, bulk velocity, and temperature associated withM. Then

(3.17)

A(M) =

" "

SD−1×RD

Mb(ω, v−v) dωdv

=%

%ˆb%

%L1(dω)

"

RD|v−v|β ρ

(2πθ)D2 exp ,

−|v−u|2

- dv

=ρ θβ2 a

,v−u

√θ -

,

where the attenuation coefficient a(w) is defined by

(3.18) a(w) =%

%ˆb%

%L1(dω)

1 (2π)D2

"

RD|w−w|βexp#

12|w|2$ dw.

By rotation invariancea(w) is a function of only|w|. It is a bounded function whenβ ∈(−D,0], and is an unbounded function when β ∈ (0,2]. This fundamental difference in the behavior of a(w) requires a different analysis for each of the two cases. These will be presented in the subsequent two subsections.

Remark. For any local Maxwellian M with mass density ρ(x, t), bulk velocity u(x, t), and temperature θ one has

(3.19)

#∂t+v·∇x$

log(M) = ∂tρ+u·∇xρ+ρ∇x·u

ρ + (v−u)·ρ(∂tu+u·∇xu) +θ∇xρ ρθ

+

,|v−u|2 2θ − D2

-∂tθ+u·∇xθ+ D2θ∇x·u θ

+

,(v−u)×(v−u)

θ − D1

|v−u|2

θ I

- :∇xu +

,|v−u|2 2θ − D2

-(v−u)·∇xθ

θ .

For this to be positive everywhere, we must impose the constraints

xθ = 0, ∇xu+ (∇xu)TD1x·uI = 0. For this to be a function of|v−u|, we must impose the constraint

ρ(∂tu+u·∇xu) +θ∇xρ= 0. The form of (3.19) thereby simplifies to

#∂t+v·∇x$

log(M) = ∂tρ+u·∇xρ+ρ∇x·u

ρ +

,|v−u|2 2θ − D2

-∂tθ+D2θ∇x·u

θ .

This quadratic function of |v −u| is similar to those obtained in (3.15) and (3.16). Moreover, the constraints we imposed above imply that the spatial dependence of ρ, u, and θ has the same form as for a global Maxwellian. This means the forms that we have assumed in (3.14) are close to being general.

(13)

3.4. Soft and Maxwell Cases. We now complete the construction of Section 3.3 for cases when the collision kernel has the separable form (1.4) with β ∈ (−D,0]. In other words, for cases when the kernel arises from either soft or Maxwell potentials. Because (3.18) shows the right-hand sides of (3.15) and (3.16) are bounded functions of v when β ∈ (−D,0], we take η(t) = 1 in the forms of MU and ML given by (3.14). Criterion (3.15) then reduces to

(3.20) γ"(t)

γ(t) ≥A#

MU −ML$

, γ(0) = 1. Below we show how this criterion can be met in certain cases.

3.4.1. Near Vacuum Case. The easiest case to treat is when the lower bound is the vacuum.

Proposition 3.2. Let M(t) be the global Maxwellian in the form (1.14) given by m >0 and (a, b, c, B)∈Ω. Let Fin(v, x) be any initial data such that

(3.21) 0≤Fin≤M(s) for some s∈R.

Let β ∈(−D,0]. Let

(3.22) Ns(t) =(a(L(dw)

3 det# 1

Q$

" t 0

θ(s+t")D+β2 dt", where Q= (ac−b2)I+B2 and θ(t) is given by (1.17).

Then there exists a mild solutionF(v, x, t)to the initial-value problem (1.1) for the Boltzmann equation over [0, T) that satisfies the bounds

(3.23) 0≤F(t)≤γ(t)M(s+t) over RD×RD×[0, T), where

(3.24) γ(t) = 1

1−mNs(t) over[0, T), T = sup{t >0 : mNs(t)<1}. In particular, this solution is global when β ∈(1−D,0] and

(3.25) m Ns≤1, where Ns = lim

t→∞Ns(t).

and is globally bounded by a global Maxwellian when strict inequality holds.

Remark. The local result above is in the spirit of Kaniel and Shinbrot [7], while the global result is in the spirit of Illner and Shinbrot [6]. The bounds obtained here are different because of our use of the global Maxwelian family (1.14).

Proof. Set ML = 0 andMU =γ(t)M(s+t). Then criterion (3.20) reduces to γ"

γ ≥γA(M(s+t)), γ(0) = 1. Because by (3.18) and (1.19)

A(M(s+t)) =mθ(s+t)D+β2 3

det# 1

Q$ exp

,

−θ 2xTQx

- a

,v −u

√θ -

≤mθ(s+t)D+β2 3

det# 1

Q$

(a(L(dw) =mNs"(t), this criterion will be satisfied if

γ"2mNs"(t), γ(0) = 1.

(14)

But (3.24) gives the solutionγ(t) of this initial-value problem along with its interval of existence [0, T), whereby criterion (3.20) is satisfied over this interval. The bounds (3.23) thereby follow from (3.9) of Proposition 3.1.

Because θ(t) given by (1.17) decays like t−2 as t → ∞, the integrand in definition (3.22) of Ns(t) decays like t−(D+β) as t → ∞. It is therefore clear that Ns = limt→∞Ns(t) <∞ if and only ifβ ∈(1−D,0]. The bound (3.25) on m then follows from (3.24). ! 3.4.2. Near Global Maxwellian Case. The next easiest case to treat is when the lower and upper bounds are proportional to the same global Maxwellian.

Proposition 3.3. Let M(t) be the global Maxwellian in the form (1.14) given by m = 1 and some (a, b, c, B)∈Ω. Let m2 > m1 >0. Let Fin(v, x) be any initial data such that

(3.26) m1M(s)≤Fin≤m2M(s) for somes ∈R. Let γmin =.

m1/m2. Let β ∈(−D,0]. Let Ns(t) be given by (3.22).

Then there exists a mild solutionF(v, x, t)to the initial-value problem (1.1) for the Boltzmann equation over [0, T) that satisfies the bounds

(3.27) m1

γ(t)M(s+t)≤F(t)≤γ(t)m2M(s+t) over RD×RD×[0, T), where

(3.28) γ(t) =γmin

1 +γmin+ (1−γmin)eminm2Ns(t)

1 +γmin−(1−γmin)eminm2Ns(t) over [0, T), T = sup{t >0 : (1−γmin)eminm2Ns(t)<1 +γmin}. In particular, this solution is global when β ∈(1−D,0] and

(3.29) m2Ns ≤ 1 2γmin

log

,1 +γmin

1−γmin

-

, where Ns= lim

t→∞Ns(t), and is globally bounded by a global Maxwellian when strict inequality holds.

Remark. Condition (3.29) yields global solutions with larger mass by pickingγmin closer to 1.

Remark. This result is in the spirit of Toscani [11]. The bounds obtained here are different because of our use of the global Maxwelian family (1.14). In particular, we can treat initial data with significant rotation that are excluded by the hypotheses in [11].

Proof. Set ML = γ(t)m1M(s+t) and MU =γ(t)m2M(s+t). Criterion (3.20) then reduces to γ"

γ ≥ ,

γ− γmin2 γ

-

m2A(M(s+t)), γ(0) = 1. Because A(M(s+t))≤Ns"(t), this criterion will be satisfied if

γ" =#

γ2−γmin2 $

m2Ns"(t), γ(0) = 1.

But (3.28) gives the solutionγ(t) of this initial-value problem along with its interval of existence [0, T), whereby criterion (3.20) is satisfied over this interval. The bounds (3.27) thereby follow from (3.9) of Proposition 3.1.

By arguing as in the proof of Proposition 3.2, we see that Ns = limt→∞Ns(t) <∞ if and only ifβ ∈(1−D,0]. The bound (3.29) on m2 then follows from (3.28). !

(15)

3.4.3. Between Global Maxwellians Case. We now treat the case is when the lower and upper bounds are proportional to possibly different global Maxwellians. The question addressed below is whether or not the only bounds on global solutions that one can obtain are the ones already recovered by the arguments in the foregoing subsections.

Let β ∈ (−D,0]. Let M1 and M2 be the global Maxwellians in the form (1.14) that are given by some (m1, a1, b1, c1, B1) and (m2, a2, b2, c2, B2) inR+×Ωrespectively. We assume that M1 ≤M2. This implies thatm1

.det(Q1)≤m2

.det(Q2), where Q1 = (a1c1−b12)I+B12 >0 and Q2 = (a2c2−b22)I+B22 >0. It also implies that q1(v, x, t)≥q2(v, x, t), which means that a1 ≥a2, c1 ≥c2, and

(b1−b2)2+(B1−B2(2 ≤(a1−a2)(c1−c2). Because by (1.17) we have

θ1(t) = 1

a1t2−2b1t+c1 >0, θ1(t) = 1

a2t2 −2b2t+c2 >0, while by (2.12) we have a1 ≥a2 and (a1−a2)(c1−c2)≥(b1−b2)2, we see that

1

θ1(t)− 1

θ2(t) = (a1−a2)t2−2(b1−b2)t+ (c1−c2)≥0. Hence, θ1(t)≤θ2(t). It is also clear that ρ1(x, t)≤ρ2(x, t).

Now set ML= γ(t)1 M1(s+t) and MU =γ(t)M2(s+t). Then because

#∂t+v·∇x$

log(MU) =−#

t+v·∇x$

log(ML) = γ"

γ , A#

MU−ML$

=γA(M2(s+t))− 1

γ A(M1(s+t)), criterion (3.20) becomes

γ"

γ ≥γA#

M2(s+t)$

− 1 γ A#

M1(s+t)$

, γ(0) = 1. The above differential inequality has the form

γ" ≥ 4

γ2− A#

M1(s+t)$ A#

M2(s+t)$ 5

A#

M2(s+t)$ .

By using (3.18) followed by (1.17) and (1.19), we find that the ratio above is A(M1)

A(M2) = ρ1θ

β 2

1 a

,v−u1

√θ1

-

ρ2θ

β 2

2 a

,v−u2

√θ2

- =

m1

.det(Q1

D+β 2

1 exp

,

−θ1

2xTQ1x -

a

,v −u1

√θ1

-

m2.

det(Q2

D+β 2

2 exp

,

−θ2

2xTQ2x -

a

,v −u2

√θ2

-.

We would like to know when this ratio is bounded below by a positive number.

The ratio is bounded below by a positive number when (a1, b1, c1, B1) = (a2, b2, c2, B2), in which case

A#

M1(s+t)$ A#

M2(s+t)$ = m1

m2

. This is just the “Toscani case” treated in the last subsection.

Références

Documents relatifs

The following lemma allows us to compare the results at the global scale (in variables (t, x)) and at the local scale (in variables (s, y)) in order to carry the desired result from

2 Key words; anyon, Haldane statistics, low temperature kinetic theory, quantum Boltzmann equation.... Here dω corresponds to the Lebesgue probability measure on

We complete the result in [2] by showing the exponential decay of the perturbation of the laminar solution below the critical Rayleigh number and of the convective solution above

We show how the singularities are propagated for the (spa- tially homogeneous) Boltzmann equation (with the usual angular cuto of Grad) in the context of the small solutions

More precisely, we show that if the initial data is non-negative and belongs to a uniformly local Sobolev space in the space variable with Maxwellian type decay property in the

In 1980s, the existence of weak solutions to the spatially homogeneous case was proved by Arkeryd in [17] for the mild singular case, that is, when 0 &lt; s &lt; 1 2 , and by using

Finally, we would like to mention that there are also some interesting results on the uniqueness for the spatially homogeneous Boltzmann equation, for example, for the Maxweillian

Furthermore, we apply the similar analytic technique for the Sobolev space regularity to the nonlinear equation, and prove the smoothing property of solutions for the