• No results found

Half-Space Problems for a Linearized Discrete Quantum Kinetic Equation

N/A
N/A
Protected

Academic year: 2021

Share "Half-Space Problems for a Linearized Discrete Quantum Kinetic Equation"

Copied!
25
0
0

Loading.... (view fulltext now)

Full text

(1)

Postprint

This is the accepted version of a paper published in Journal of statistical physics. This paper has been peer-reviewed but does not include the final publisher proof-corrections or journal pagination.

Citation for the original published paper (version of record):

Bernhoff, N. (2015)

Half-Space Problems for a Linearized Discrete Quantum Kinetic Equation.

Journal of statistical physics, 159(2): 358-379 https://doi.org/10.1007/s10955-015-1190-4

Access to the published version may require subscription.

N.B. When citing this work, cite the original published paper.

Permanent link to this version:

http://urn.kb.se/resolve?urn=urn:nbn:se:kau:diva-35795

(2)

Niclas Bernhoff

Half-Space Problems for a Linearized Discrete Quantum Kinetic Equation

Received: date / Accepted: date

Abstract We study typical half-space problems of rarefied gas dynamics, includ- ing the problems of Milne and Kramer, for a general discrete model of a quan- tum kinetic equation for excitations in a Bose gas. In the discrete case the plane stationary quantum kinetic equation reduces to a system of ordinary differential equations. These systems are studied close to equilibrium and are proved to have the same structure as corresponding systems for the discrete Boltzmann equation.

Then a classification of well-posed half-space problems for the homogeneous, as well as the inhomogeneous, linearized discrete kinetic equation can be made. The number of additional conditions that need to be imposed for well-posedness is given by some characteristic numbers. These characteristic numbers are calcu- lated for discrete models axially symmetric with respect to the x-axis. When the characteristic numbers change is found in the discrete as well as the continuous case. As an illustration explicit solutions are found for a small-sized model.

Keywords Bose-Einstein condensate · Low temperature kinetics · Discrete kinetic equation · Milne problem · Kramer problem

1 Introduction

Half-space problems have an important role in the study of the asymptotic be- havior of the solutions of boundary value problems of kinetic equations for small Knudsen numbers [4, 15, 16, 26, 27]. In this paper we study half-space problems related to a quantum kinetic equation [23, 33], for the distribution function of ex- cited atoms interacting with a Bose-Einstein condensate. Motivated by the work N. Bernhoff

Department of Mathematics and Computer Science, Karlstad University, 651 88 Karlstad, Swe- den

Tel.: +46-54-7002024 Fax: +46-54-7001851

E-mail: niclas.bernhoff@kau.se

(3)

of Arkeryd and Nouri [2] we are interested in the equation

p1dF

dx = C12(F) + Γ C22(F) , F(0, p) = F0(p) for p1> 0,

(1)

where F = F (x, p) denotes the distribution function of the excitations, Γ ∈ R+= {x ∈ R |x ≥ 0 } is constant, x ∈ R+, p = p1, p2, p3 ∈ R3, and F0= F0(p) is given, with the collision integrals

C12(F) = n Z

δ0δ3(1 + F) F0F0− F 1 + F0

1 + F0 dpdp0dp0, with

δ0= δ p− p0− p0 δ



p2+ n − p02

− p02 and δ3= δ (p− p) − δ p0− p − δ p0− p ,

and

C22(F) = Z

δ1(1 + F) (1 + F) F0F0− FF 1 + F0

1 + F0 dpdp0dp0 with

δ1= δ p + p− p0− p0 δ



p2+ p2− p02

− p02 .

Here and below we use the notation F0= F (x, p0) etc.. The density of the conden- sate, nc, is assumed to be constant, nc= n (cf. [2]). In the Nordheim-Boltzmann [24] (or the Uehling-Uhlenbeck [28]) collision integral C22(F) binary collisions between excited atoms are considered, while in the collision integral C12(F) bi- nary collisions involving one condensate atom are considered [33].

If the distribution function F is close to an equilibrium distribution, i.e. a Planckian

P= 1

eα(|p|2+n)+β ·p− 1

= 1

eα(|p−p0|2+n0) − 1, with α > 0, β ∈ R3, p0= −β

2, and n0= n − |p0|2, cf. [2], then the non-linear equation (1) can be approximated by the linearized equation

p1d f

dx+ L f = 0, f = f (x, p) f(0, p) = f0(p) for p1> 0,

(2)

where

F= P + (P(1 + P))1/2f, F0= P + (P(1 + P))1/2f0, and L = L12+ Γ L22, with

L12f = Z

δ0δ3 h

P− P0 (P0(1 + P0))1/2f0+ P− P0 (P0(1 + P0))1/2f0+ 1 + P0+ P0 (P(1 + P))1/2f

i

dpdp0dp0

(4)

and L22f =

Z δ1

h

PP− P0(1 + P + P) (P0(1 + P0))1/2f0+ PP− P0(1 + P + P) (P0(1 + P0))1/2f0+ P(1 + P0+ P0) − P0P0 (P(1 + P))1/2f+

P(1 + P0+ P0) − P0P0 (P(1 + P))1/2fi

dpdp0dp0. It can be shown (cf. [2] for L12and, for example, [16] for the linearized Boltzmann operator) that the linearized operators L12and L22, and so also L, (all acting in the velocity space) are symmetric and positive semi-definite operators on L2.

In the paper [2], Arkeryd and Nouri studied the Milne problem for the lin- earized equation (2), with Γ = 0, F = P (1 + f ), and a cut-off at λ > 0 in the integrand of L, such that |p| , |p| , |p0| , |p0| ≥ λ . The corresponding linearized half-space problems for the Boltzmann equation is well-studied [3, 17, 20], see also [4] and references therein.

In this paper we discretize the variable p and obtain a general discrete model for Eq.(1), which is similar to the discrete Boltzmann equation (a general discrete velocity model, DVM, for the Boltzmann equation) [14]. It is a well-known fact that the Boltzmann equation can be approximated up to any order of DVMs [11, 25, 18], which motivated us to introduce discrete models also for this equation. By the discretization, Eq.(1) reduces to a system of ordinary differential equations.

We find that the discrete linearized quantum kinetic equation (the discrete version of Eq.(2)) has the same structure as the linearized discrete Boltzmann equation.

This means that the linearized operator is symmetric and positive semi-definite, and that the null-space is non-trivial. One difference is that the mass flow is not constant (with respect to the variable x) as for the discrete Boltzmann equation.

However, this cause us no difficulties, in difference to in the continuous case in [2], since the structure will still be the same. A classification of well-posed half-space problems for the homogeneous, as well as the inhomogeneous, linearized discrete Boltzmann equation has been made in [5] (which is a continuation of the paper [9]), based on the dimensions of the stable, unstable and center manifolds of the singular points (Maxwellians for DVMs). We establish similar results in our case.

This means, that we, in addition to adding the Nordheim-Boltzmann (or Uehling- Uhlenbeck) collision integral C22(F), also can introduce an inhomogeneous term and more general boundary conditions. Similar results can also be established for the discrete Nordheim-Boltzmann (or Uehling-Uhlenbeck) equation and the discrete anyon Boltzmann equation (see Remark 6 and 7 in Section 4).

Furthermore, we have, for axially symmetric discrete models with respect to the x-axis, made a table of some characteristic numbers, from which we, by The- orem 1, can obtain the dimensions of the stable, unstable and center manifolds of the singular points (Planckians in our case). This includes determining when the characteristic numbers change, not only in the discrete, but also in the continuous case (cf. [7, 5] for DVMs and [17] for the Boltzmann equation).

Nonlinear half-space problems for the Boltzmann equation have also been studied for small perturbations of the singular points (Maxwellians for the Boltz- mann equation), see for example [6, 21, 22, 29] for the discrete Boltzmann equation

(5)

and [30, 19, 32] for the continuous Boltzmann equation. In the discrete case similar results to the ones in [6] can be obtained for the quantum kinetic equation (1).

We want to make clear that the aim of the paper is not the study of the gen- eral half-space problem we obtain, since it is already well studied for the discrete Boltzmann equation [9, 5]. The novelty of the paper is instead the introduction of discrete models for the equation for the distribution function of excited atoms in- teracting with a Bose-Einstein condensate and the studies of those models. These studies includes that we by the right linearization end up with a system having sim- ilar properties as the one obtained for the discrete Boltzmann equation. It makes it, as mentioned above, possible to extend the results in [2] obtained for the con- tinuous equation. The same is true also for the discrete Nordheim-Boltzmann (or Uehling-Uhlenbeck) equation and the discrete anyon Boltzmann equation (see Re- mark 6 and 7 in Section 4). However, in concrete situations, it will look different depending on which equation we study. One difference is the characteristic num- bers, studied for axially symmetric models in Section 5. Our experience from the Boltzmann equation, make us believe that these numbers (also calculated in the continuous case) are as important in the continuous case as in the studied discrete case. Another difference between our equation and the Boltzmann equation, is that in our case we will have a non-constant mass-flow. To illustrate this we created a model that we solved explicitly and was able to give an explicit expression for the non-constant mass flow for (see Section 6).

The remaining part of this paper is organized as follows. In Section 2 we in- troduce a general discrete model for Eq.(1) and derive some of its properties. By a transformation around a Planckian, we obtain a linearized operator and a non- linear part presented in Section 3. It is shown that the system has the same struc- ture (the linearized operator and the non-linear part have similar properties) as the corresponding system for DVMs of the Boltzmann equation. Then some results for the linearized discrete Boltzmann equation can be applied for the problem of our study. These results are presented in Section 4. In Section 5 some character- istic numbers, from which we, by Theorem 1, can obtain the dimensions of the stable, unstable and center manifolds of the singular points (Planckians), are ob- tained for axially symmetric discrete models with respect to the x-axis. When the characteristic numbers change, are determined both in the discrete as well as the continuous case. A linearized half-space problem (with Γ = 0) is explicitly solved for a small-sized discrete model in Section 6.

2 Discrete model

We introduce a general discrete model for Eq.(1) p1idFi

dx = C12i(F) + Γ C22i(F) , x ∈ R+, i = 1, ..., N, (3) whereP = {p1, ..., pN} ⊂ Rd is a finite set, Fi= Fi(x) = F (x, pi), where F = F(x, p) is the distribution function of the excitations, and Γ ∈ R+ is constant.

For generality, we allow p to be of dimension d, rather than of dimension 3. We assume that

p1i 6= 0, for i = 1, ..., N.

(6)

The collision operators C12i(F) are given by C12i(F) =

N

j,k,l=1

il− δi j− δikjkl((1 + Fl) FjFk− Fl(1 + Fj) (1 + Fk)) , where

δi j= 1 if i = j 0 if i 6= j , with Γjki = 1 if

pi= pj+ pkand |pi|2= pj

2+ |pk|2+ n, (4)

and Γjki = 0 otherwise. Furthermore, the collision operators C22i(F) are given by C22i(F) =

N

j,k,l=1

Γi jkl((1 + Fi) (1 + Fj) FkFl− FiFj(1 + Fk) (1 + Fl)) , with Γi jkl= 1 if

pi+ pj= pk+ pland |pi|2+ pj

2= |pk|2+ |pl|2, (5) and Γi jkl= 0, otherwise.

Remark 1 For a function g = g(p) (possibly depending on more variables than p), we will, as we consider the discrete case, identify g with its restrictions to the points p ∈P, i.e.

g= (g1, ..., gN) , with gi= g (pi) . Then Eq.(3) can be rewritten as

BdF

dx = C12(F) + Γ C22(F) , with x ∈ R and B = diag(p11, ..., p1N). (6) The collision operator C12(F) in (6) is also given by the expression

C12(F) = neLF+ n eQ(F, F), (7) where

 eLF

i =

N

j,k=1

i jkFk− ΓjkiFiand

Qei(F, G) =

N

j,k=1

ΓjkiQijk(F, G) − 2Γi jkQki j(F, G), with

Qijk(F, G) =1

2(FjGk+ GjFk− Fi(Gj+ Gk) − Gi(Fj+ Fk)) , and the collision operator C22(F) in (6) is given by the expression

C22(F) = Q(F, F) + bQ(F, F, F), (8)

(7)

where

Qi(F, G) =1 2

N

j,k,l=1

Γi jkl((GkHl+ HkGl) − (GiHj+ HjGi)) and

Qbi(F, G, H) = 1 2

N

j,k,l=1

Γi jkl((Fi+ Fj) (GkHl+ HkGl) − (Fk+ Fl) (GiHj+ HjGi)) . A function φ = φ (p) is a collision invariant, if and only if,

φi= φj+ φk, (9)

for all indices such that Γjki 6= 0, if Γ = 0, with the additional condition

φi+ φj= φk+ φl, (10)

for all indices such that Γi jkl 6= 0, if Γ 6= 0. We have the trivial collision invariants (”the physical collision invariants”)

φ1= p1, ..., φd= pd, φd+1= |p|2+ n (11) including all linear combinations of these. We want to stress that by Remark 1 and in correspondence with Eqs.(9) , (10) the collision invariants φi= φi(p) in Eq.(11) are vectors.

In the discrete case, in difference to the continuous case, there can be spurious (or non-physical) collision invariants. We consider below (even if this restriction is not necessary in our general context) only normal discrete models. That is, discrete models without spurious collision invariants, i.e. any collision invariant is of the form

φ = φ (p) = −α

|p|2+ n

− β · p (12)

for some constant α ∈ R and β ∈ Rd. Construction of normal discrete kinetic models and especially DVMs have been extensively studied, see for example [10, 12, 13] and references therein. A Maxwellian distribution or just Maxwellian is on the form

M= eφ= e−α(|p|2+n)−β ·p or

Mi= eφi= e−α(|pi|2+n)−β ·pi, i = 1, ..., N,

where φ = (φ1, ..., φN) is a collision invariant, and a Planckian distribution or just Planckian is given by

P= M

1 − M = 1

M−1− 1= 1

eα(|p|2+n)+β ·p− 1

(8)

or

Pi= Mi

1 − Mi

= 1

eα(|pi|2+n)+β ·pi− 1for i = 1, ..., N.

One can easily see that hH,C12(F)i =

n

N

i, j,k=1

Γjki((1 + Fi) FjFk− Fi(1 + Fj) (1 + Fk)) (Hi− Hj− Hk) , (13) and so

 log F

1 + F,C12(F)



= n

N

i, j,k=1

Γjki(1 + Fi) (1 + Fj) (1 + Fk)

 Fj 1 + Fj

Fk 1 + Fk

Fi 1 + Fi

  log Fi

1 + Fi

− log

 Fj 1 + Fj

Fk 1 + Fk



≤ 0, with equality if and only if

Fi 1 + Fi

= Fj 1 + Fj

Fk 1 + Fk

, (14)

for all indices such that Γjki 6= 0. Here and below, we denote by h·, ·i the Euclidean scalar product in Rn. Hence, there is equality in Eq.(14), if and only if, F

1 + F is a Maxwellian, or equivalently, if and only if, F is a Planckian.

Moreover, one can easily obtain that

hH,C22(F)i =1 4

N

i, j,k,l=1

Γi jkl(Hi+ Hj− Hk− Hl)

((1 + Fi) (1 + Fj) FkFl− FiFj(1 + Fk) (1 + Fl)) , (15) and so

 log F

1 + F,C22(F)



=1 4

N

i, j,k=1

Γi jkl(1 + Fi) (1 + Fj) (1 + Fk) (1 + Fl)

 log

 Fi 1 + Fi

Fj 1 + Fj



− log

 Fk 1 + Fk

Fl 1 + Fl



 Fk 1 + Fk

Fl 1 + Fl

Fi 1 + Fi

Fj 1 + Fj



≤ 0, (16) with equality if and only if

Fi 1 + Fi

Fj 1 + Fj

= Fk 1 + Fk

Fl

1 + Fl, (17)

(9)

for all indices such that Γi jkl6= 0. There is equality in Eq.(16), if and only if, F 1 + F is a Maxwellian, or equivalently, if and only if, F is a Planckian.

By the relations (13) and (15)

hφ ,C12(F) + Γ C22(F)i , (18)

is zero, independently of our choice of non-negative function F, if and only if, φ is a collision invariant, and so (for normal models) the equation

hφ ,C12(F) + Γ C22(F)i = 0, has the general solution (12).

3 Linearized operator Given a Planckian

P= 1

eα(|p|2+n)+β ·p− 1

= 1

eα(|p−p0|2+n0) − 1, (19) with α > 0, β ∈ Rd, p0=β

2 and n0= n − |p0|2, we denote

F= P + R1/2f, with R = P(1 + P), (20) in Eq.(6), and obtain

Bd f

dx+ L f = S ( f ) ,

where L = L12+ Γ L22is the linearized collision operator (N × N matrix) given by L12f = −2nR−1/2Q(P, Re 1/2f) − neLR1/2f (21) and

L22f = −R−1/2

2Q(P, R1/2f) + bQ(R1/2f, P, P) + 2 bQ(P, R1/2f, P)

. (22) The nonlinear part S ( f ) = S12( f , f ) + Γ S22( f , f , f ) is given by

S12( f , g) = nR−1/2Q(Re 1/2f, R1/2g) (23) and

S22( f , g, h) = R−1/2

Q(R1/2f, R1/2g) + bQ(P + R1/2f, R1/2g, R1/2h)+

Q(Rb 1/2f, P, R1/2h) + bQ(R1/2f, R1/2g, P) . (24) In more explicit forms, the operators (21) and (23) read

(L12f)i= n

N j,k=1

ΓjkiLijkf− 2Γi jkLi jk f R1/2i

, i = 1, ..., N, (25)

(10)

where

Lijkf = (1 + Pj+ Pk) R1/2i fi− (Pk− Pi) R1/2j fj− (Pj− Pi) R1/2k fk, and

S12i( f , g) = n

N

j,k=1

ΓjkiSijk( f , g) − 2Γi jkSki j( f , g) R1/2i

, i = 1, ..., N, with

Sijk( f , g) =1 2



R1/2j R1/2k ( fjgk+ gjfk) − R1/2i R1/2j ( figj+ gifj) −

R1/2i R1/2k ( figk+ gifk) . Moreover, the operators (22) and (24) read, in more explicit forms,

(L22f)i=

N

j,k,l=1

Γi jkl R1/2i

(Pi jklfi+ Pkljifj− Pkli jfk− Plki jfl), i = 1, ..., N (26)

where

Pi jkl= (Pj(1 + Pk+ Pl) − PkPl) R1/2i , and

S22i( f , f , f ) =

N

j,k,l=1

Γi jkl R1/2i



Skli j( f , f , f ) − Si jkl( f , f , f )

, i = 1, ..., N,

with

Skli j( f , f , f ) = (1 + Pi+ Pj) R1/2k R1/2l fkfl+

R1/2i fi+ R1/2j fj



PkR1/2l fl+ PlR1/2k fk+ R1/2k R1/2l fkfl . By Eqs.(4),(25), and the relations

Pj(1 + Pj)(Pk− Pi) = Pk(1 + Pk)(Pj− Pi) = Pi(1 + Pj)(1 + Pk), Pi(1 + Pj+ Pk) = P2P3= Pi(1 + Pj)(1 + Pk)

for Γjki 6= 0, we obtain the equality

hg, L12f i= n

N

i, j,k=1

ΓjkiPi(1 + Pj) (1 + Pk)

fi R1/2i

fj

R1/2j

fk R1/2k

gi R1/2i

gj R1/2j

gk R1/2k

.

(11)

Similarly, by Eqs.(5),(26), and the relations

PiPj(1 + Pk)(1 + Pl) = PkPl(1 + Pi)(1 + Pj), Pi jkl = PkPl(1 + Pj)

1 + Pi

Pi

for Γi jkl6= 0, we obtain the equality

hg, L22f i= 1 4

N

i, j,k,l=1

Γi jklPiPj(1 + Pk)(1 + Pl)

fi R1/2i

+ fj R1/2j

fk R1/2k

fl R1/2l

gi R1/2i

+ gj R1/2j

gk R1/2k

gl R1/2l

.

It is easy to see that the matrix L is symmetric and positive semi-definite, i.e.

hg, L f i = hLg, f i and h f , L f i ≥ 0, for all functions g = g(ξ ) and f = f (ξ ).

Furthermore, h f , L f i = 0 if and only if fi

R1/2i

= fj R1/2j

+ fk R1/2k

(27)

for all indices satisfying Γjki 6= 0 if Γ = 0.

If Γ 6= 0, h f , L f i = 0 if and only if also, additionally to Eq.(27), fi

R1/2i + fj

R1/2j

= fk R1/2k

+ fl R1/2l

(28)

for all indices satisfying Γi jkl 6= 0. We denote f = R1/2φ in Eq.(27) and Eq.(28) and obtain Eq.(9) and Eq.(10) respectively. Hence, since L is semi-positive,

L f = 0 if and only if f = R1/2φ , where φ is a collision invariant (12).

Then also D

S( f ) , R1/2φ E

= hC12(F) + Γ C22(F) , φ i +D

F, LR1/2φ E

= 0 for all collision invariants φ .

The system (6) transforms in Bd f

dx+ L f = S( f ). (29)

(12)

The diagonal matrix B (6) (under our assumptions) has no zero diagonal elements and is non-singular. We denote f (0) = f0. Then the formal solution of Eq.(29) reads

f(x) = e−xB−1Lf0+

x Z

0

e(σ −x)B−1LB−1[S ( f )] (σ ) dσ .

As in the case of DVMs for the Boltzmann equation, the linearized operator Lis symmetric, positive semi-definite, and have a non-trivial null-space, and by assumption, the matrix B is non-singular. Therefore, we can apply a result obtained by Bobylev and Bernhoff in [9] (see also [5]), that we will present below.

We denote by n±, where n++ n= N, and m±, with m++ m= q, the num- bers of positive and negative eigenvalues (counted with multiplicity) of the matri- ces B and B−1Lrespectively, and by m0the number of zero eigenvalues of B−1L.

Moreover, we denote by k+, k, and l the numbers of positive, negative, and zero eigenvalues of the ρ × ρ matrix K (ρ = d + 1 for normal discrete models), with entries

ki j=yi, yj

B=yi, Byj ,

such thaty1, ..., yρ is a basis of the null-space of L, i.e. in our case span y1, ..., yρ = N(L) = span

R1/2p1, ..., R1/2pd, R1/2(|p|2+ n) . Here and below, we denote h·, ·iB= h·, B·i and by N(L) the null-space of L.

In applications, the number ρ of collision invariants is usually relatively small compared to N (note that formally N = ∞ for the continuous equation whereas ρ ≤ 4). Also, the matrix B is diagonal and therefore all its eigenvalues are known.

This explains the importance of the following result [9]. The theorem is valid for any real symmetric matrices L and B, such that L is semi-positive, B is invertible, and dim(N(L)) = ρ ≥ 1.

Theorem 1 The numbers of positive, negative and zero eigenvalues of B−1L are given by

m+= n+− k+− l m= n− k− l m0= ρ + l.

In the proof of Theorem 1 a basis

u1, ..., uq, y1, ..., yk, z1, ..., zl, w1, ..., wl

(30) of RN, such that

yi, zr∈ N(L), B−1Lwr= zrand B−1Luα= λαuα, (31) and

uα, uβ

B= λαδα β, with λ1, ..., λm+> 0 and λm++1, ..., λq< 0, yi, yj

B= γiδi j, with γ1, ..., γk+> 0 and γk++1, ..., γk< 0, huα, zriB= huα, wriB= huα, yiiB= hwr, yiiB= hzr, yiiB= 0,

hwr, wsiB= hzr, zsiB= 0 and hwr, zsiB= δrs, (32) is constructed.

(13)

4 Half-space problems

We consider the inhomogeneous (or homogeneous if g = 0) linearized problem Bd f

dx + L f = g, (33)

where g = g(x) ∈ L1(R+, Rn), with one of the boundary conditions (O) the solution tends to zero at infinity, i.e.

f(x) → 0 as x → ∞;

(P) the solution is bounded, i.e.

| f (x)| < ∞ for all x ∈ R+; (Q) the solution can be slowly increasing at infinity, i.e.

| f (x)| e−εx→ 0 as x → ∞, for all ε > 0.

In case of boundary condition (O) we additionally assume that

g(x) ∈ N(L)for all x ∈ R+. (34) Remark 2 The boundary condition (O) corresponds to the case when we have made the expansion (20) around a Planckian P, such that F → P as x → ∞. The boundary conditions (P) and (Q) are the boundary conditions in the Milne and Kramers problem respectively.

We can (without loss of generality) assume that B= B+ 0

0 −B



, (35)

where

B+= diag p11, ..., p1n+ and B= −diag p1n++1, ..., p1N , with p11, ..., p1n+> 0 and p1n++1, ..., p1N< 0.

We also define the projections R+: RN→ Rn+and R: RN→ Rn, by R+s= s+= (s1, ..., sn+) and Rs= s= (sn++1, ..., sN) for s = (s1, ..., sN).

At x = 0 we assume the boundary condition

f+(0) = h0, (36)

where h0∈ Rn+.

The solutions of the system (33) with one of the boundary conditions (O) (together with condition (34)), (P), and (Q) reads

f(x) = Ψ+(x) +Ψ(x) + Φ+(x) + Φ(x),

(14)

where (in the notations of (30)-(32))

Ψ+(x) =

m+ r=1

ur

βr(0)e−λrx+

x Z

0

e(σ −x)λrhg (σ ) , uri λr

,

Ψ(x) = −

q

r=m++1

ur

Z

x

e(σ −x)λrhg (σ ) , uri λr

dσ ,

Φ+(x) =

k+ i=1

yi

µi(0) +

x Z

0

hg (σ ) , yii hyi, yiiB

+

l

j=1

zj

ηj(0) +

x Z

0

g (σ) , wj dσ − x

αj(0) +

x Z

0

g (σ) , zj

,

Φ(x) =

k

i=k++1

yi

µi(0) +

x Z

0

hg (σ ) , yii hyi, yiiB

+

l

j=1

wj

αj(0) +

x Z

0

g (σ) , zj

,

with for the case with boundary condition (O) (note that hg (x) , yii =g (σ) , zj = 0 for i = 1, ..., k and j = 1, ..., l, by condition (34))

µi(0) = αj(0) = 0 for i = 1, ..., k and j = 1, ..., l, and ηj(0) = −

Z

0

g (σ) , wj dσ for j = 1, ..., l,

and for the case with boundary condition (P)

αj(0) = −

Z

0

g (σ) , zj dσ for j = 1, ..., l.

By the boundary condition (36), or equivalently

R+Ψ+(0) + R+Φ+(0) = h0− R+Ψ(0) − R+Φ(0), and that

{R+u1, ..., R+um+, R+y1, ..., R+yk+, R+z1, ..., R+zl}

References

Related documents

Keywords: Nonlinear observability, identifiability, observability rank condition, nonlinear systems, metabolic model, kinetic model, metabolism, glycolysis, Saccharomyces

The results are in accordance with corresponding results for the continuous Boltzmann equation obtained in the non-degenerate case, with boundary conditions of Dirichlet type [48]..

These results include well-posedness results for half- space problems for the linearized discrete Boltzmann equation, existence results for half-space problems for the weakly

We consider some extensions of the classical discrete Boltzmann equation to the cases of multicomponent mixtures, polyatomic molecules (with a finite number of different

Here we consider a half-space problem of condensation for a pure vapor in the presence of a non-condensable gas by using discrete velocity models (DVMs) of the Boltzmann equation..

In this paper we present a general discrete velocity model (DVM) of Boltzmann equation for anyons - or Haldane statistics - and derive some properties for it concerning:

Therefore, regarding the computational complexity (i.e. number of floating point operations needed), memory requirements and time consumption, iterative methods become nearly

By means of a literature review, this article analyses how research on benchmark- ing e-government has evolved over time, with a particular focus on its main findings and the