Rosseland Definitif 3
Rosseland Definitif 3
1 ∂I
+ ω · ∇x I + Σ(T, ν)(I − Bν (T )) = 0
c ∂t
(1) Z +∞
∂E
Σ(T, ν)(I¯ − Bν (T )) dν,
(T ) =
∂t 0
where
αν 3
Z
1
Bν = I¯ = I(t, x, ω, ν) dω.
ehν/kT − 1 4π S2
1
2 F. GOLSE AND F. SALVARANI
The parameter c > 0 is the speed of light and Bν (T ) is Planck’s formula for
the black-body radiation at temperature T , the parameters h and k being
the Planck and the Boltzmann constants respectively.
We have introduced the normalization constant
4
h
α = 15 ,
kπ
so that Stefan’s law assumes the simple form:
Z +∞
Bν (T ) dν = T 4 .
0
The function Σ = Σ(T, ν), is the opacity — or absorption cross-section
— of the host medium. In general, the opacity takes into account how
energy is transferred from the incoming radiation to the electrons in the
background medium. Its expression is complicated in general, as it must
take into account the contributions of bound electrons, of free electrons as
well as transitions from bound to free states. The respective importance of
each such transition in the opacity strongly depends upon the nature of the
background and the temperature. (For instance, bound-bound transitions
are more important for heavy elements; however, their importance decreases
at high temperatures as most electrons are in free states). We refer to
chapter VII of [18] and to [1] for more information on this subject.
Thus we seek to establish the validity of the Rosseland approximation
under the mildest possible assumptions on the opacity, since so little is
known about it in general.
There are however, two particular features of the opacity that complicate
the Rosseland approximation, and that we wish to take into account in the
present paper.
Specifically, our proof of the Rosseland approximation allows considering
opacities that may tend to infinity as T → 0+ . As we shall see, such opacities
lead to a limit equation that is parabolic degenerate, and hence has singular
solutions.
This kind of behavior of the opacity can be seen on the example of the
Kramers opacity (an explicit formula for the opacity in the case of free-free
transitions and for a hydrogen-like atom). The Kramers opacity is of the
form
hν
1 − e− kT
kT
ΣK (T, ν) = C 1 + O
(hν)3 (kT )1/2 hν
— see fla (7.72) on p. 174 in [18]. In this example Σ(T, ν) ∼ T −1/2 .
Otherwise, the opacity can oscillate strongly because of bound-free and
bound-bound transitions.
In view of the above considerations, we only assume in this paper that
the opacity satisfies Σ(T, ν) ∼ T −λ as T → 0, with λ ∈ (0, 1].
The other important nonlinear term in system (1) is the internal energy
E = E(T ). This quantity depends on the temperature, but the experimental
ROSSELAND LIMIT FOR RADIATIVE TRANSFER 3
curves suggest that its dependence upon T is simpler than that of the cross
section.
Henceforth, we assume that E is increasing and such that E 0 (T ) ∼ T γ
as T → 0, where γ ∈ [0, 2]. This assumption includes the perfect gas law,
E(T ) = cV T . Without loss of generality we further assume that E(0) = 0.
In some relevant physical situations, for example in the external layers of a
stellar atmosphere [15], the opacity is independent of the photons frequency.
It is therefore convenient to integrate system (1) over frequencies and obtain
the following system of equations, called the gray model:
1 ∂u
4
c ∂t + ω · ∇x u + Σ(T )(u − T ) = 0
(2) Z
∂E 1
(T ) = Σ(T )(ū − T 4 ),
ū = u(t, x, ω) dω.
∂t 4π S2
ε ∂uε 1
4
c ∂t + ω · ∇x uε + ε Σ(Tε )(uε − Tε ) = 0
(3) Z
∂E 1
ε2 (Tε ) = Σ(Tε )(ūε − Tε4 ),
ūε = uε (t, x, ω)dω.
∂t 4π S2
The first step in obtaining the formal limit is the definition of the macro-
scopic density
Z
1
ρε (t, x) = ūε (t, x, ω) = uε (t, x, ω) dω
4π S 2
(1) (2) (3)
and the flux vector jε (t, x) = (jε (t, x), jε (t, x), jε (t, x)):
Z
1
jε(i) (t, x) = ω (i) uε (t, x, ω) dω,
4πε S 2
where ω (i) denotes the i-th component of the vector ω.
Then, by integrating the first equation of System (3), it is easy to deduce
that
∂ h ρε i
+ E(Tε ) + ∇x · jε = 0.
∂t c
Moreover, integrating the first equation of System (3) with respect to ω in
S 2 after multiplying each side of this equation by ω (i) gives
(i)
ε2 ∂jε
Z
1
+ ω (i) ∇x · (ωuε ) dω + Σ(Tε ) jε(i) = 0.
c ∂t 4π S 2
Assume the existence of the two limits:
ρ = lim ρε , j = lim jε .
ε→0 ε→0
(4) E is continuous in [0, +∞) and belongs to the class C 1 ((0, +∞)).
Moreover E(0) = 0 and E 0 (y) ≥ 0 for all y > 0.
(5) E 0 (y) ∼ y α for y → 0, where α ∈ [0, 2].
6 F. GOLSE AND F. SALVARANI
be smooth. This is not the case here, as temperature may fail to be differ-
entiable at interfaces with cold matter. Here, as in the case of Carleman’s
kinetic model [10], we use both relative entropy and a compactness argument
that allows for possibly singular solutions of the target equations. Relative
entropy is not used all the way through the proof to control the distance
between the radiative intensity and the (fourth power of) the temperature
that solves the Rosseland limiting equation, but only to control the radiation
flux. This control is in turn one essential step in the following compactness
argument.
That it is possible to use relative entropy within a compactness argument
to establish the validity of such a macroscopic (or hydrodynamic) limit seems
to be special to the case of limits leading to a pure diffusion equation (de-
generate or not), but without streaming term. In particular, the strategy
presented here does not seem to apply to the derivation of the incompressible
Navier-Stokes equations from the Boltzmann equation.
From a mathematical point of view, it will be more convenient to work
with two new unknowns: the photon density u = u(t, x, ω) and the normal-
ized temperature θ = θ(t, x), the last one being defined as θ = T 4 . Moreover,
we will set c = 1.
We write now the radiative transfer equations (3) with respect to the two
new unknowns.
We first define the normalized cross section σ and the internal energy E
as
σ(θ) = Σ(θ1/4 ), E(θ) = E(θ1/4 ).
The pair (σ, E), must satisfy the same constraints we have imposed on
the pair (Σ, E). The precise meaning of admissible cross section and energy
is given in the following definition:
Definition 2.4. The normalized opacity
σ : R+ → R+
and the internal energy
E : [0, +∞) → [0, +∞)
are admissible if and only if
(1) σ(y) is strictly positive for any y > 0 and is of class C 1 ((0, +∞));
(2) lim σ(y) = 0;
y→+∞
3. Existence proof
This section is devoted to the proof of Theorem 2.5. In order to simplify
the notation, in this section we will specialize system (4) by considering
ε = 1, and by eliminating all the subscripts in the unknowns. Obviously,
the results are valid for all parameters ε > 0.
10 F. GOLSE AND F. SALVARANI
3.1. The maximum principle. As a first step, we prove that (4) preserves
the non-negativity of the solutions, which are uniformly bounded in time.
This property is proved in the following proposition:
Proposition 3.1. Let us suppose that there exists a solution (u, θ) for the
radiative transfer system (4) in (t, x, ω) ∈ (0, τ ) × Ω × S 2 , with bounded
initial conditions
uε (0, x, ω) = uin (x, ω) ∈ L∞ (Ω × S 2 )
θε (0, x) = θin (x) ∈ L∞ (Ω)
and bounded boundary data
uε (t, x, ω)|Γ− = ub (t, x, ω) ∈ L∞ ((0, τ ) × ∂Ω × S 2 ).
Then there exists a constant M , which depends only on the initial and
boundary data, such that 0 ≤ u, θ ≤ M a.e. (uniformly in x, ω and t).
Proof. We develop first the part of the proof which establishes the bound
from above. We define
n o
M = max kuin kL∞ (Ω×S 2 ) , kθin kL∞ (Ω) , kub kL∞ (Γ− ×S 2 ×(0,τ ]) .
and
Z Z
d −
(E(θ)) dxdω = σ(θ)(u − θ) sign− (θ) dxdω = 0,
dt Ω×S 2 Ω×S 2
∂un
∂t + ω · ∇x un + σn (θn )(Un − θn ) = 0
(7) Z
1
Et (θn ) = σn (θn )(Ūn − θn ), Ūn = Un (t, x, ω)dω,
4π S2
where
σn (y) = min{σ(y), n}, Un (t, x, ω) = max{0 , min{un , M }}, n ∈ N.
System (7) has the form of a Lipschitzian perturbation of the free trans-
port equations. Therefore, thanks to Theorems 6.1.2 and 6.1.6 in [17],
System (7) admits a strong solution (un , θn ) ∈ C 0 ([0, τ ]; L2 (Ω × S 2 )) ×
C 0 ([0, τ ]; L2 (Ω)).
By applying the same strategy of Lemma 3.1, we can prove that, if we
suppose that there exists a solution (un , θn ) for the truncated approximation
(7) in (t, x, ω) ∈ (0, τ )×Ω×S 2 , with the same initial and boundary conditions
of Theorem 2.5, then there exists a constant M , which depends only on the
initial and boundary data, such that 0 ≤ un , θn ≤ M a.e. in Ω, uniformly
in x, ω and t.
Thanks to the maximum principle for System (7), the solution of such set
of equations coincides therefore with the solution of the following system:
∂un
∂t + ω · ∇x un + σn (θn )(un − θn ) = 0
(8) Z
1
Et (θn ) = σn (θn )(ūn − θn ), ūn = un (t, x, ω)dω,
4π S 2
where
σn (y) = min{σ(y), n}, n ∈ N.
We hence deduce the following Lemma:
Lemma 3.2. Consider System (8) for (t, x, ω) ∈ (0, τ ) × Ω × S 2 , with initial
conditions
un (0, x, ω) = uin (x, ω) ∈ L∞ (Ω × S 2 )
ROSSELAND LIMIT FOR RADIATIVE TRANSFER 13
By the first equation of System (8), we hence deduce that (∂t + ω · ∇x )un is
bounded in L1loc (t, x, ω).
Since un bounded in L∞ (t, x, ω), we then use simply the averaging lemma
proved in [9] to conclude.
As a next step we prove that the unknown θn of System (8) depends only
on ūn in a very special way:
Lemma 3.4. Consider System (8). Then there exists a . . . functional F :
Lploc ((0, τ ) × Ω) → Lploc ((0, τ ) × Ω), p ∈ [1, +∞) such that θn = F(ūn ).
Proof. Let us consider the second equation of (8). Since E is of class C 1 , we
can introduce the function Hn = Hn (θn ) defined in such a way that
E 0 (θn )
(9) Hn0 (θn ) = .
σn (θn )
14 F. GOLSE AND F. SALVARANI
p
(2) θn * θ in L∞ ∗
t,x,ω -weak and, by Lemma 3.4, that θn → θ in Lloc and
a.e..
Consequently, σn (θn ) θn * σ(θ) θ in in Lploc and a.e..
We can now apply the dominated convergence theorem to control the
nonlinear terms. We recall moreover that the only term which depends
on the angular variable ω is un . Therefore we can deduce that, modulo
extraction of a subsequence, there exists a measure µ such that
σn (θn ) un *∗ µ.
We only need to identify µ with the limit product σ(θ) u.
For any m < n, we have that
σm (θn ) un ≤ σn (θn ) un ,
and therefore, in the limit m, n → +∞
σ(θ) u ≤ µ.
On the other hand, the second equation of (8), says us that
Z Z Z τZ
in
E(θn )(τ, x) dx − E(θ )(x) dx + σn (θn ) θn dtdx =
Ω Ω 0 Ω
Z τ Z Z
1
σn (θn ) un dtdxdω.
4π 0 Ω S2
The l.h.s converges to
Z Z Z τ Z
in
E(θ)(τ, x) dx − E(θ )(x) dx + σ(θ) θ dtdx,
Ω Ω 0 Ω
whereas the r.h.s converges to
Z τ Z
µ dtdx.
0 Ω
ROSSELAND LIMIT FOR RADIATIVE TRANSFER 17
Classical theory on the Laplace equation ensures that f exists, is unique and
satisfies the strong maximum principle. In particular, 0 ≤ f ≤ M = kub k∞
in Ω̄ and f ∈ W 2,p (Ω) for any p ∈ [1, +∞). Obviously, here the time t acts
only as a parameter.
and
1 τ
Z Z
σ(θε )(uε − θε )2 dt dx dω ≤ J
ε2 0 Ω×S 2
uniformly for each ε > 0.
Thanks to Proposition 4.1, we can deduce the following corollary:
Corollary 4.2. Let (uε , θε ) as in Proposition 4.1. Then, if one of the se-
quences u , or θ , or ū admits a limit, then all the three sequences go to the
same limit in strong L2 -sense.
Proof. Since σ is bounded below by a strictly positive constant, we obtain
from Proposition 4.1 that
1
kuε − θε kL2t,x,ω = O(1)
ε
and therefore
lim ku − θ kL2t,x,ω = 0.
→0
Since
kūε − θε kL2t,x ≤ kuε − θε kL2t,x,ω
and
kuε − ūε kL2t,x,ω ≤ kuε − θε kL2t,x,ω + kūε − θε kL2t,x ,
we obtain that also
lim ku − θ kL2t,x,ω = lim kθ − ū kL2t,x = 0.
→0 →0
Because of the regularity on E (see Definition 2.4) and because (θ − ū ) →
0 strongly in L2 ((0, τ ) × Ω), then E(θε ) → E(ρ) (strongly) in L2 ((0, τ ) × Ω).
Let us now take into account the flux equation
(i) Z
∂jε 1
(22) ε2 + ω (i) ∇x · (ωuε ) dω + σ(θε ) jε(i) = 0.
∂t 4π S 2
Since σ is strictly positive, we can state that
Z
1/2 −1/2 2 ∂jε 1
σ(θε ) jε = −σ(θε ) ε + ∇x · (ω ⊗ ω)uε dω .
∂t 4π S2
Now, by Proposition (4.1), we have that
kσ(θε )1/2 jε k2L2 ((0,τ )×Ω) ≤ kσ(θε )1/2 (uε − θε )k2L2 ((0,τ )×Ω×S 2 ) +
In order to verify that the initial and boundary conditions are the correct
limits of the conditions imposed on the radiative transfer system, we will
use a regularity argument.
24 F. GOLSE AND F. SALVARANI
Let us now consider the continuity equation (19): since the flux jε is
bounded in L2 ((0, τ ) × Ω), we deduce that
(ρε + E(θε ))t is bounded in L2 (0, T ; H −1 (Ω)) ;
which, joint to the bound on ρ and E, implies that
(24) ρ + E(θ ) is relatively compact in C([0, τ ]; H −1 (Ω))
by Arzela-Ascoli’s theorem and therefore the target equation recovers the
correct initial data.
Because of the initial condition of the radiative transfer system (4), one
has
Z
−1 1
(25) ρ ∈ C([0, τ ]; H (Ω)) and ρ t=0 = uin (x, ω) dω.
4π S 2
Let us now care about the boundary conditions: we consider equation
(23) and perform the limit as → 0. We note immediately, thanks to the
identity
σ(θε )(uε − θε )uε = σ(θε )(uε − θε )2 + σ(θε )uε θε − σ(θε )θε2 ,
that ω·∇x uε is bounded in L2 ((0, τ )×Ω×S 2 ), we can deduce that, thanks to
Cessenat’s results [4, 5], uε has a trace in L2 -sense and that uε * u weakly
in L2 ((0, τ ] × ∂Ω × S 2 ), and therefore the trace of the limit exists. Thus the
proof of the nonlinear diffusion limit is competed.
Acknowledgment: This paper has been written when the second author
was visiting the Centre de Mathématiques Laurent Schwartz of the Ecole
Polytechnique, France. FS is very grateful to the Ecole Polytechnique for
the invitation and the facilities provided to him.
Financial support of the Italian National Group for Mathematical Physics
(GNFM) is acknowledged.
References
[1] Alexander, D. R.; Ferguson, J. W. Low temperature Rosseland opacities. Astro-
physical Journal 437, 2 (1994) 879-891
[2] Bardos, C.; Golse, F.; Perthame, B. The Rosseland approximation for the radia-
tive transfer equations. Comm. Pure Appl. Math. 40, 6 (1987) 691–721.
[3] Bardos, C.; Golse, F.; Perthame, B.; Sentis, R. The nonaccretive radiative
transfer equations: existence of solutions and Rosseland approximation. J. Funct. Anal.
77, 2 (1988) 434–460.
[4] Cessenat, M. Théorèmes de trace Lp pour des espaces de fonctions de la neutronique.
C. R. Acad. Sc. Paris 299, 16 (1984), 831–834.
[5] Cessenat, M. Théorèmes de trace pour des espaces de fonctions de la neutronique.
C. R. Acad. Sc. Paris 300, 3 (1985), 89–92.
[6] Dogbe, C. The radiative transfer equations: diffusion approximation under accretive-
ness and compactness assumptions. Comput. Math. Appl. 42 (2001), 783–791.
[7] Golse, F. The Milne problem for the radiative transfer equations (with frequency
dependence). Trans. Amer. Math. Soc. 303, 1 (1987) 125–143.
[8] Golse, F., Perthame, B. Generalized solutions of the radiative transfer equations
in a singular case. Comm. Math. Phys. 106 (1986), 211–239.
ROSSELAND LIMIT FOR RADIATIVE TRANSFER 25
[9] Golse, F.; Saint-Raymond, L. The Navier-Stokes limit of the Boltzmann equation
for bounded collision kernels. Invent. Math. 155 (2004), 81-161
[10] Golse, F.; Salvarani, F. The nonlinear diffusion limit for generalized Carleman
models: the initial-boundary value problem Nonlinearity 20 (2007), 927-942
[11] Goudon, T.; Poupaud, F. Approximation by homogenization and diffusion of ki-
netic equations. Comm. Partial Differential Equations 26 (2001), 537–569.
[12] Larsen, E.; Pomraning, G. ; Badham, V. Asymptotic analysis of radiative transfer
problems. J. Quant. Spectros. Radiat. Transfer 29 (1983), 285–310
[13] Marcati, P.; Milani, A. J. The one-dimensional Darcy’s law as the limit of a
compressible Euler flow. J. Diff. Eq. 84 (1990) 129–147.
[14] Mercier, B. Application of accretive operators theory to the radiative transfer equa-
tions. SIAM J. Math. Anal. 18 (1987) 393–408.
[15] Mihalas, D.; Weibel Mihalas B. Foundations of Radiation Hydrodynamics, Ox-
ford University Press, New York 1984.
[16] F. Murat. Compacité par compensation, Ann. Scuola Norm. Sup. Pisa Cl. Sci. 5
(1978) 489–507.
[17] Pazy, A. Semigroups of linear operators and applications to partial differential equa-
tions, Springer-Verlag, New York 1983
[18] Pomraning, G. The Equations of Radiation Hydrodynamics, Dover Publications,
Mineola NY, 2005.