Radiative Transfer An Introduction To Exact And Asymptotic Methods Hlne Frisch instant download
Radiative Transfer An Introduction To Exact And Asymptotic Methods Hlne Frisch instant download
https://ebookbell.com/product/radiative-transfer-an-introduction-
to-exact-and-asymptotic-methods-hlne-frisch-51985332
https://ebookbell.com/product/radiative-transfer-in-coupled-
environmental-systems-an-introduction-to-forward-and-inverse-
modeling-1st-edition-knut-stamnes-5318032
https://ebookbell.com/product/an-introduction-to-radiative-
transfer-1st-annamaneni-peraiah-1460988
https://ebookbell.com/product/remote-sensing-of-the-environment-and-
radiation-transfer-an-introductory-survey-1st-edition-anatoly-
kuznetsov-2519296
https://ebookbell.com/product/radiative-transfer-in-participating-
media-rahul-yadav-c-balaji-45334314
Radiative Transfer In The Atmosphere And Ocean Second Knut Stamnes
https://ebookbell.com/product/radiative-transfer-in-the-atmosphere-
and-ocean-second-knut-stamnes-9950782
https://ebookbell.com/product/radiative-transfer-in-stellar-and-
planetary-atmospheres-lucio-crivellari-10679910
https://ebookbell.com/product/radiative-transfer-in-the-atmosphere-
and-ocean-gary-e-thomas-1659948
https://ebookbell.com/product/3d-radiative-transfer-in-cloudy-
atmospheres-physics-of-earth-and-space-environments-1st-edition-
alexander-marshak-2115158
https://ebookbell.com/product/heat-transfer-volume-2-radiative-
transfer-michel-ledoux-abdelkhalak-el-hami-46563032
Hélène Frisch
Radiative
Transfer
An Introduction to Exact
and Asymptotic Methods
Radiative Transfer
Hélène Frisch
Radiative Transfer
An Introduction to Exact and Asymptotic
Methods
Hélène Frisch
Observatoire de la Côte d’Azur, CNRS,
Laboratoire Lagrange
Université Côte d’Azur
Nice Cedex 04, France
© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland
AG 2022
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
To Uriel, Anne, and Thomas
Preface
vii
viii Preface
constructed by (Wiener and Hopf 1931) for the Milne problem, which describes the
temperature profile of a stellar atmosphere in radiative equilibrium. Their method
of solution, which now carries their name, is one the greatest achievements in
mathematical physics in the first half of this century. It relies on the properties of
analytic functions in the complex plane. In contrast, when the question is posed to
find the distribution of photons that have been reflected or transmitted through a
scattering medium of finite extension, a slab for example, then there is no explicit
expression for the radiation field, inside or outside the medium. However, there
are many interesting exact relations and integral equations, which can be solved
numerically. This topic is only briefly mentioned in this book.
Following the work of Wiener and Hopf, several methods have been developed
to tackle semi-infinite medium problems. One of the best known method in
astrophysics is the construction of a nonlinear integral equation for the famous H -
function, a sort of special function for half-space transport problems. Proposed for
the first time by Ambartsumian (1942), it largely avoids complex plane analysis and
relies on invariance properties of the radiative transfer process. Plasma physicists
became strongly interested in the subject in relation to the transport of neutrons
in nuclear reactors. New methods were developed, such as the Case singular
eigenfunction expansion method (Case 1960), which leads to linear singular integral
equations with Cauchy-type kernels. Methods of solutions for these equations,
also based on complex plane analysis, rely largely on a technique introduced by
Carlemann (1922) and developed by Muskhelishvili (1953). These methods lead
to boundary value problems in the complex plane, known as Riemann–Hilbert
problems, and provide explicit expressions for the H -function, as does the Wiener–
Hopf method.
A key motivation for writing this book was to present in the same volume several
of the methods leading to exact results for semi-infinite media, usually presented
in separate books or articles, for both scalar and polarized radiation fields, and also
connections, that are not always clearly apparent, between the various methods.
When the number of scatterings in the host medium is very large, asymptotic
techniques may be employed to analyze the large-scale behavior of the random
walk of the photons and to explain, for example, why monochromatic scattering has
all the characteristics of an ordinary diffusion process, while complete frequency
redistribution of spectral lines has in contrast those of a Lévy walk. This asymptotic
approach is presented in Part III of the book, while exact methods are presented in
Parts I and II. Part I deals with unpolarized radiation and Part II with monochro-
matic radiation, polarized by Rayleigh scattering, and spectral lines, polarized by
resonance scattering and the Hanle effect, which is a modification of resonance
polarization by a weak magnetic field. Details on the organization of each part are
given in Chap. 1.
The intended readership for the book ranges from first-year graduate students
to professional scientists. Astrophysicists can find in this book exact methods of
solutions used in radiative transfer, but also applied in distant fields, such as financial
mathematics. A major effort has been made in the organization of the material
to give a synthetic view on exact and asymptotic methods in radiative transfer.
Preface ix
Although oriented towards methods of solutions, this book also provides exact
expressions for the radiation field intensity and polarization for a number of standard
problems.
I am deeply grateful to many colleagues for their help and encouragement in
this undertaking. My sincere thanks go particularly to V. Bommier, V. V. Ivanov,
B. Rutily, and P. Zweifel, who have read significant parts of the book and have
taken the time to share with me their intimate knowledge of the field. I have adopted
many of their suggestions. My thanks go also to my family, especially to my mother
D. Piron-Lévy, for lasting encouragements; to my husband, Uriel, for reading the
full manuscript and playing the role of the non-specialist; and to L. Anusha, E. Lega,
and M. Sampoorna for their very generous help in the preparation of the figures. This
book would not exist without the continuous support of Springer teams in Europe
and India. I owe them a thousands thanks.
References
xi
xii Contents
Fig. 2.1 Left panel: the space variable τ and the inclination angle
θ of a ray with direction n. The axis z is normal to the
surface of the medium. Right panel: the heliocentric
interpretation of the angle θ . The Sun is viewed from the
right. Observations at disk center correspond to θ = 0 and
those at the limb to θ = π/2 . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 14
Fig. 2.2 The Doppler profile, and several Voigt profiles with
different values of the parameter a. The left panel shows
ϕ(x) and the right panel log ϕ(x). In the right panel, the
Lorentzian wing regime characterized by ϕ(x) ∼ 1/x 2
can be observed for |x| larger than (− ln a)1/2 .. . . . . . . . . . . . . . . . . . . . 21
Fig. 2.3 Complete frequency redistribution. The integration
domains (ξ, x) and (x, ξ ), with ξ = μ/ϕ(x) . . .. . . . . . . . . . . . . . . . . . . . 29
Fig. A.1 Integration contour for the Plemelj formulae. The
contribution from the singular point is obtained by letting
ρ → 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 65
Fig. 5.1 The function g(ξ ) for the Doppler profile and the Lorentz
profile, 1/(π(1 + x 2 )), in linear scales in the upper panel
and in log-log scales in the lower panel. The value of
g(ξ ) is constant for ξ ≤ 1/ϕ(0). The algebraic behavior
of g(ξ ) for ξ → ∞ given in Eqs. (5.8) and (5.9) can be
observed in the lower panel. The Voigt profile has the
same algebraic behavior as the Lorentz profile .. . . . . . . . . . . . . . . . . . . . 71
Fig. 5.2 The function k(ν) for the Doppler profile and the Lorentz
profile, 1/(π(1 + x 2 )), in linear scales in the upper panel
and in log-log scales in the lower panel. The angular
point is located at ν = ϕ(0). For the Doppler and Lorentz
profiles, k(ν) tends to zero as 1/ν for ν → ∞. For
ν → 0, one can observe in the lower panel the algebraic
behaviors given in Eqs. (5.8) and (5.9). The Voigt profile
has the same algebraic behaviors as the Lorentz profile . . . . . . . . . . . 72
xxi
xxii List of Figures
Fig. 12.3 The figure shows the strips of analyticity of V (z), the
dispersion function, and of V ∗ (z) defined in Eq. (12.34).
It shows also the analyticity half-planes of Vu (z) and
Vl (z), which satisfy V (z) = Vu (z)/Vl (z), and of Vu∗ (z)
and Vl∗ (z), which satisfy V ∗ (z) = Vu∗ (z)/Vl∗ (z) . . . . . . . . . . . . . . . . . . 225
Fig. 12.4 Analyticity half-planes of the inhomogeneous term
Q̂∗ (z) and of the functions Gu (z) and Gl (z) defined in
Eq. (12.47). The thick solid line indicates the branch cut
of Vl (z) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 229
Fig. 12.5 The analyticity half-planes of the functions in the left
and right hand-sides of Eq. (12.16) and their common
analyticity strip . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 230
Fig. 12.6 The Wiener–Hopf method for the Milne problem. The
half-plane analyticity domains of the functions in the left
and right hand-sides of Eq. (12.73) and their common
strip of analyticity. The function Vu (z) has a double zero
at the origin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 235
Fig. 12.7 The Sommerfeld diffraction problem: (a) the geometry;
(b) the boundary conditions for the field ϕ(x, y) and
for ϕ (x, 0± ) = ∂ϕ(x, 0± )/∂y, the normal derivative of
ϕ(x, y) on the axis y = 0. The notation 0± means that
y → 0 by positive or negative values. The continuity
across the axis y = 0 is indicated by a vertical line with
two arrows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 241
Fig. 12.8 The Sommerfeld diffraction problem: the factorization of
γ = (k − λ)1/2 (k + λ)1/2 . The principal determination of
(k + λ)1/2 is analytic in (k) > 0 and that of (k − λ)1/2
in (k) < 0 .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 245
Fig. I.1 Monochromatic scattering. (a) Contours for obtaining an
explicit expression of Vu (z) in terms of the logarithm
of the dispersion function. The point z lies in the upper
half-plane (z) > δ > −ν0 . (b) Contour for obtaining the
relation between Vu (z) and X(z). The point z lies in the
upper half-plane (z) > −1 . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 247
Fig. 13.1 Ellipse described by the tip of the electric field vector in
the plane orthogonal to the direction of propagation. The
angle (0 ≤ ≤ π) specifies the orientation of the
ellipse and the angle χ (−π/4 ≤ χ ≤ π/4) the ellipticity
and the sense in which the ellipse is being described . . . . . . . . . . . . . . 257
Fig. 13.2 Rotation of the reference direction by an angle α . . . . . . . . . . . . . . . . . . 259
Fig. 13.3 Geometry of a single scattering. The components Ir
and Il are respectively perpendicular and parallel to the
scattering plane (, ) . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 262
List of Figures xxv
xxix
xxx List of Tables
Table 18.1 The matrices G̃(z), H(z), G̃R (z), and HR (z) at the origin
and at infinity. The matrix R = GR (∞) is a rotation
matrix, which can be calculated by solving numerically
the nonlinear integral equations for G̃R (z) or HR (z). The
matrix E = diag[, Q ] is equal to L(0). The I-matrix is
defined by I(0, z) = HR (z) . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 439
Table 19.1 The asymptotic behaviors of K(τ ) for τ → ∞, g(ξ ) for
ξ → ∞, k(ν) for ν → 0, and K̂(k) for k → 0. The
kernel is an even function of τ . Here τ , ξ , and ν are
positive. For monochromatic scattering, k(ν) is zero in
the interval [0, 1[ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 465
Table 19.2 The index α for the Doppler and Voigt profiles and the
asymptotic expressions of f (y) and f (y) for y large.
The function f (y) is the inverse function of the profile
ϕ(x) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 466
Table 22.1 Infinite medium. Behavior at large optical depths of the
resolvent function for small values of , the destruction
probability per scattering. The constant a is the Voigt
parameter of the line. The nearly conservative zone
corresponds to 1 τ < τeff and the strong absorption
zone to τ > τeff . For monochromatic scattering τeff ∼ ν0 . . . . . . . . 509
Table 22.2 Semi-infinite medium. Asymptotic behavior of the
resolvent function in the nearly conservative and strong
absorption zones. The thermalization lengths τeff are
given in Table 22.1. For monochromatic scattering
τeff ∼ ν0 . For = 0, the monochromatic √ resolvent
function has exactly the value 3 at infinity . .. . . . . . . . . . . . . . . . . . . . 513
Table 24.1 Moments of Hl (μ) and Hr (μ) and other numerical
constants. The two first moments of Hl (μ) and Hr (μ),
the constants q and c, and the values of Hl (1) and Hr (1)
are from Chandrasekhar (1960, p. 248). This reference
provides also tables for Hl (μ) and Hr (μ), μ ∈ [0, 1],
used to calculate numerically the third and fourth order
moments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 544
Table 25.1 Scaling laws for complete frequency redistribution
with a Doppler profile. In the first column, the small
expansion parameter; in row 1: β = 0 and T = ∞, in
row 2: = 0 and T = ∞, and in row 3: β = = 0. The
four subsequent columns show τeff , the characteristic
scale of variation of the radiation field, xc , the
characteristic frequency defined by τeff ϕ(xc ) ∼ 1, the
mean number of scatterings N, and the mean path
length L .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 559
xxxii List of Tables
(Königsberg-Göttingen).
1.1 Part I: Scalar Radiative Transfer Equations 3
solved in Chap. 5. One of the key ingredient of the Wiener–Hopf method makes
its applicable to partial differential equations. We give, with a wave diffraction
problem, an example in Sect. 12.6.
The second Part is devoted to radiative transfer problems for fields linearly
polarized by scattering processes sharing the same Rayleigh polarization phase
matrix, namely the Rayleigh scattering of monochromatic radiation, the resonance
polarization of spectral lines, formed with complete frequency redistribution, and
the associated Hanle6 effect, which takes into account the effects of a weak magnetic
field.
An important feature of the Rayleigh polarization phase matrix is that it can be
factorized as the product of two matrices depending separately on the directions
of the incident and scattered beams. We make use of this property to introduce
a representation of the polarized radiation field based on the irreducible spherical
tensors for polarization introduced by Landi Degl’Innocenti (1984). With this
representation, which can be introduced for one-dimensional and multi-dimensional
media, the scattering term (the integral over the directions of the incident beams
weighted by a polarization matrix) in the radiative transfer equation depends only
on the position inside the medium. The radiative transfer equations have thus the
same structure as those of Part I, except that the radiation field and the source term
are now vectors. For a one-dimensional medium, the convolution equation for the
source term has also the same structure as in the scalar case, but its kernel is a matrix.
The Green function, the resolvent function, the H -function also become matrices.
The scalar Riemann–Hilbert problem becomes a matrix problem and the nonlinear
H -equation becomes a nonlinear matrix equation. The matrix structure makes the
situation significantly more complicated and it is in general impossible to construct
explicit solutions for these equations.
In some very special cases, homogeneous matrix Riemann–Hilbert problems
have an explicit solution, conservative monochromatic Rayleigh scattering is
one of them. The reason is that the corresponding matrix problem can actually
be transformed into two scalar ones. This property was first demonstrated by
Chandrasekhar7 (1946), by a detailed analysis of the properties of the matrix kernel
and has allowed him to construct exact solutions for the Milne and the diffuse
reflection problems in terms of two functions Hl and Hr with exact expressions.
There is no such decomposition for non-conservative Rayleigh scattering, however
an exact solution was obtained by Siewert et al. (1981, see also Aoki and Cercignani
1985) with very advanced mathematics, out of the scope of this book. It is a
real breakthrough but the solution that has been obtained has a very complicated
form, not easily amenable to a numerical solution. Here we treat non-conservative
Rayleigh scattering in the same way as the resonance polarization of spectral lines
and the Hanle effect. For spectral lines, there is no exact solution, even when are
formed with complete frequency redistribution and the scattering is conservative.
We show, by different methods, how to construct a nonlinear matrix H-equation,
which can be solved numerically.
Part II is organized as follows. Chapter 13 is devoted to various aspects of
scattering physics. After recalling the classical electromagnetic description of
a polarized beam of light, we present polarization matrices for the Rayleigh
scattering,8 the resonance polarization of spectral lines and the Hanle effect. For
spectral lines, the polarization matrix depends on the frequencies of the incident
and scattered photons and will be referred to as the redistribution matrix. Wandering
from our main subject, we described in some details, for a simple-two level atom,
the spectral profile of the scattered radiation. The purpose is to place the assumption
of complete frequency redistribution in a broader context and also to offer a brief
introduction to the literature on the scattering polarization of spectral lines.
In Chap. 14, we show how the radiative transfer equations for the Stokes
parameters of a field linearly polarized by one of the three mechanisms mentioned
above, can be transformed into new radiative transfer equations, in which the source
term depends only on the position in the medium, as in the scalar case. For a one-
dimensional medium, these new equations can then be recast as vector or matrix
Wiener–Hopf integral equations and used as starting points to generalize some of
the methods of solution described in the scalar case. In Chap. 15 we generalize to
a polarized radiation√ the real-space methods described in Chap. 11 for the scalar
case. We construct -laws for Rayleigh scattering, resonance polarization, and the
Hanle effect, and also the corresponding nonlinear matrix H-equations. We also
introduce matrix singular integral equations that will be used in the subsequent
chapters.
In the remaining three chapters we show how vector and matrix singular integral
equations with Cauchy-type kernels can be treated with Hilbert transform methods.
For conservative Rayleigh scattering, we explain in Chap. 16 how the matrix
Riemann–Hilbert problem can be transformed into two scalar problems and how the
polarized Milne and the diffuse reflection problems can be solved exactly in terms of
the two scalar functions Hl and Hr . For polarization problems with no exact solution
(non-conservative Rayleigh scattering, resonance polarization, and Hanle effect), we
establish in Chap. 17 homogeneous matrix Riemann–Hilbert problems and discuss
in detail the asymptotic behavior of their solutions at infinity (in the complex plane).
Although they contain undetermined constants, these solutions can be used, as we
show in Chap. 18, to recover the nonlinear H-equations, established in Chap. 15
with a real-space method. This second method has the advantage of proving that
the existence of solutions to these nonlinear equations. We also discuss in the same
8 The expression Rayleigh scattering always implies that the scattering is monochromatic.
6 1 An Overview of the Content
We describe in Part III asymptotic methods based on the existence of small or large
parameters such as , the destruction probability per scattering, or T the optical
thickness of the medium. They are applicable when photons undergo a very large
number of scatterings and greatly contribute to the physical understanding of the
formation of a radiation field by multiple scatterings.
For example, it will be shown that for monochromatic scattering photons perform
an ordinary diffusion random walk. In particular, that the averaged distance travelled
after n scatterings (n large) goes as n1/2 . For complete frequency redistribution, the
random walk of photons has an anomalous behavior, performing a so-called Lévy9
walk, because of the changes of frequency within the spectral line. The averaged
distance travelled by photons varies as n1/α with α smaller than 2. The difference
between ordinary and anomalous diffusion has its origin in the asymptotic behavior
of the kernel. These scaling laws are derived in Part III with an asymptotic analysis
of the integral equation for the source function and also with a statistical analysis of
the random walk of the photons.
Asymptotic methods can address radiative transfer problems, which have no
exact solution and are usually solved numerically. They provide equations that
are simpler than the original equations. Their solutions are approximations to the
original problem, but their domain of validity and accuracy are controlled by
the small expansion parameter. We give examples with non-conservative Rayleigh
scattering and partial frequency redistribution of spectral lines.
Another strength of asymptotic methods is that they provide spatial and fre-
quency scales of variation for the radiation field, which describe the spreading of
photons in space and frequency (in the case of spectral lines). For example the
spreading in space, which is known in the astronomical literature as thermalization
length, henceforth denoted τeff , can be identified with the characteristic scale for the
spatial variations of the radiation field.
Part III is organized as follows. The somewhat technical Chap. 19 contains a
detailed analysis of the asymptotic behavior of the kernel K(τ ) for large values of τ
and of its Fourier transform K̂(k) for k → 0. In Chap. 20 we show how to perform a
large scale asymptotic analysis of the integral equation for the source function in an
infinite medium, by a proper rescaling of the optical depth. This analysis provides
scaling laws for the thermalization length τeff for monochromatic scattering and
References
Aoki, K., Cercignani, C.: On the matrix Riemann–Hilbert problem relevant to Rayleigh scattering.
J. Appl. Math. Phys. (ZAMP) 36, 61–69 (1985)
Chandrasekhar, S.: On the radiative equilibrium of a stellar atmosphere. X. Astrophys. J. 103,
351–370 (1946)
Hummer, D.G.: Non-coherent scattering I. The redistribution functions with Doppler broadening.
Mon. Not. R. Astr. Soc. 125, 21–37 (1962)
Ivanov, V.V.: Generalized Rayleigh scattering. III. Theory of I-matrices, Astron. Astrophys. 307,
319–331 (1996)
Landi Degl’Innocenti, E.: Polarization in Spectral lines III : Resonance polarization in the non-
magnetic, collisionless regime. Sol. Phys. 91, 1–26 (1984)
Siewert, C.E., Kelley, C.T., Garcia, R.D.M.: An analytical expression for the H-Matrix relevant to
Rayleigh scattering. J. Math. Anal. Appl. 84, 509–518 (1981)
Part I
Scalar Radiative Transfer Equations
Chapter 2
Radiative Transfer Equations
In this chapter, we present the scalar radiative transfer equations used in Part I to
illustrate exact method of solutions for radiative transfer equations in semi-infinite
media. We also present different types of integral equations that can be derived
from the integro-differential equations. All the equations are time-independent and
one-dimensional. They describe the formation of continuous spectra and spectral
lines. For continuous spectra, there is no change in frequency at each scattering.
We refer to it as monochromatic scattering. For a spectral line, photons change
their frequency at each scattering. They can be fully decorrelated, a situation known
as complete frequency redistribution, or partially correlated. One says that there is
partial frequency redistribution. An example of partial frequency redistribution is
presented in Appendix J of Chap. 13 and some asymptotic results on the large scale
behavior of the radiation field in Chap. 26. Otherwise, we consider only spectral
lines formed with complete frequency redistribution (also known as noncoherent
scattering) for which exact results can be obtained.
The integro-differential radiative transfer equations are presented in Sect. 2.1
for monochromatic scattering and complete frequency redistribution. Section 2.2
is devoted to the source function, sum of the scattered radiation and of a primary
source of photons. With our assumptions on the frequency redistribution, it depends
only on the position in the medium. We show that it satisfies an integral equation
of the convolution type, with a translation invariant kernel. Section 2.4 is devoted to
the Green function and some associated functions, such as the resolvent function,
regular part of the surface Green function. These functions, which characterize the
scattering process and the geometry of the medium, independently of the primary
sources, obey integral equations identical to those of the source function, except for
inhomogeneous terms. In this same Sect. 2.4, we prove some fundamental properties
of the Green function and of convolution type integral equations for semi-infinite
media.
Radiative transfer equations are linear Boltzmann equations describing the propaga-
tion of an ensemble of photons inside a host medium and their interactions with the
particles of the medium. Each photon is treated as a particle of zero mass, moving
with the velocity of light c. At an instant t, photons are characterized by a position r,
a direction of motion n, and an energy E = hν, with h the Planck constant and ν the
frequency. A systematic derivation of radiative transfer equations from the Maxwell
equations is outside the scope of this book. The electro-magnetic description of the
radiation field is introduced briefly in Part II when dealing with polarized radiation
fields. A systematic electro-magnetic description of multiple scattering of light can
be found in, e.g., Mishchenko et al. (2006) and Mishchenko (2008) (see also Landi
Degl’Innocenti and Landolfi 2004, Chap. 5).
In this book we consider exclusively interaction processes relevant to the forma-
tion of spectral lines and continuous spectra. These processes are (i) creation, (ii)
absorption followed by a destruction and (iii) absorption followed by a reemission.
The third type of interaction is referred to as a scattering process. Creations
correspond to a radiative decay of a level (bounded or unbounded) that has been
first excited by a collision, say, with electrons. Destruction is the opposite process:
a level excited by photon absorption is deexcited by collision with some particles of
the medium. These three processes, together with the photons entering or escaping
the medium participate in the formation of the radiation field.
The time evolution of an ensemble of photons can be described by a distribution
function f (t, r, n, ν) dV d dν, which is the number of photons at time t in an
infinitesimal volume dV around r, a solid angle d around n and a frequency
interval dν around ν. In Astronomy, the custom is to work with the specific intensity,
I (t, r, n, ν), related to the photon distribution function by
The specific intensity is proportional to the density of energy of the radiation field.
In the traditional radiometric point of view, it is defined in terms of the amount
of radiant energy dE in a frequency interval dν around ν, transported across an
element of area dσ in an infinitesimal solid angle d around a direction n, during a
time dt. By definition
where θ is the angle between the direction n and the outward normal to dσ . The
specific intensity is a positive quantity. The description of the radiation field with
the specific intensity, or with a scalar distribution function, is sufficient as long as
the polarization of the radiation is ignored. In Astronomy, polarized radiation is
usually described by its four Stokes parameters (see Chap. 13).
2.1 The Integro-Differential Radiative Transfer Equations 13
1 ∂I (t, r, n, ν)
+ n · ∇I (t, r, n, ν) =
c ∂t
− [χd (t, r, n, ν) + χs (t, r, n, ν)]I (t, r, n, ν)
∞
d
+ (t, r, n, ν, n , ν )I (t, r, n , ν ) dν
0 4π
+ Q∗ (t, r, n, ν), (2.3)
with
∞ d
χs (t, r, n, ν) = (t, r, n, ν, n , ν ) dν . (2.4)
0 4π
The right hand side in Eq. (2.3) describes the sinks and sources of photons. The
coefficient χd corresponds to an absorption followed by a destruction and the
coefficient and χs to an absorption followed by a reemission. The integral term
describes the scattering of an incident beam of radiation with direction n and
frequency ν into a scattered beam with direction n and frequency ν. The integration
is over all possible directions and frequencies of the incident beam. The last
term Q∗ (t, r, n, ν) represents a source of primary photons. An analysis of the
interaction between the radiation field and the matter of the host medium is needed
to determine the coefficients of the radiative transfer equation. For problems relevant
to Astronomy and more specifically to the formation of stellar spectra, this subject
is developed in several text books such as Aller (1963), Jefferies (1968), Mihalas
(1978), Ivanov (1973), Landi Degl’Innocenti and Landolfi (2004), and Hubeny and
Mihalas (2015).
The equations considered in this book are much simpler than Eq. (2.3) but are
not mathematical oddities. In particular they can describe the formation of spectral
lines and other spectral features such as atomic continua in a stellar atmosphere. The
simplifications come from several assumptions:
(i) The radiation field is assumed to be independent of time. The time dependence
plays a role only for very fast evolving phenomena, namely those with a
characteristic time of variation t such that t l/c, with l the scale
of variation in the physical space.
(ii) The medium is assumed to be one-dimensional. This assumption holds for
atmospheres with a thickness small compared to the radius of the star, such as
the solar atmosphere. The one-dimensional hypothesis implies that the physical
properties of the atmosphere (temperature, density, chemical composition,
and so on) have no horizontal variations. The one-dimensional variable, here
denoted z, is measured along the normal to the medium (see Fig. 2.1). This
one-dimensional assumption is raised for some problems considered in Parts II
and III.
14 2 Radiative Transfer Equations
+∞ z
z θ
n
n z
θ n
θ
τ n
z
−∞
+∞
Fig. 2.1 Left panel: the space variable τ and the inclination angle θ of a ray with direction n. The
axis z is normal to the surface of the medium. Right panel: the heliocentric interpretation of the
angle θ. The Sun is viewed from the right. Observations at disk center correspond to θ = 0 and
those at the limb to θ = π/2
(iii) In Part I, we also assume that the radiation field has a cylindrical symmetry.
The direction of the field is fully defined by the colatitude θ ∈ [0, π], i.e.
inclination of the ray with respect to the z-axis (see Fig. 2.1) or with μ = cos θ ,
μ ∈ [−1, +1]. For solar observations, n being the direction of the line-of-
sight, then θ is the so-called heliocentric angle. Observations at disk center
correspond to θ = 0 (μ = 1) and those at the limb to θ = π/2 (μ = 1) (see
Fig. 2.1).
Radiation fields without a cylindrical symmetry are considered in Part II,
when we discuss the polarization of radiation fields.
(iv) The absorption coefficient is assumed to be independent of the direction n. It
may then be written as
This form of coefficient holds for continuous spectra and spectral lines. The
coefficient kd has the dimension of the inverse of a length. It is proportional to
the number of scattering atoms per unit volume multiplied by a cross-section.
The function f (ν) gives the probability density that a photon is absorbed in a
frequency interval [ν, ν + dν].
(v) Concerning the scattering process, we consider monochromatic scattering and
complete frequency redistribution.
For monochromatic scattering
Equation (2.4), which expresses the detail balance between the absorptions and
emissions takes then the form
(vi) For spectral lines and continuous spectra, the primary source of photons often
comes from the radiative decay of an atomic level (bound or free) that has been
excited by a collision. When the colliding particles have a Maxwellian velocity
distribution, the primary source has the form
where Q∗ν (r) stands for the Planck function at the point r and at the
frequency ν.
With the assumptions listed above, Eq. (2.3) becomes
∂
μ I (z, ν, μ) = −κ(z)ϕ(ν)I (z, ν, μ)
∂z
1 +1 ∞
+ ks (z) R(ν, ν )I (z, ν , μ ) dν dμ + kd (z)ϕ(ν)Q∗ν (z), (2.11)
2 −1 0
16 2 Radiative Transfer Equations
where
The use of the optical depth as space variable is quite popular in radiative transfer
because it provides a scale for the extinction of the radiation field. For a semi-infinite
medium, it is positive inside the medium with τ = 0 at the surface (see Fig. 2.1).
Equation (2.11) can thus be rewritten as
∂
μ I (τ, ν, μ) = ϕ(ν)I (τ, ν, μ)
∂τ
1 +1 ∞
− [1 − (τ )] R(ν, ν )I (τ, ν , μ ) dν dμ − (τ )ϕ(ν)Q∗ν (τ ), (2.15)
2 −1 0
where
kd (τ ) kd (τ )
(τ ) = = . (2.16)
κ(τ ) kd (τ ) + ks (τ )
The coefficient [1 − (τ )] gives the probability that a photon that has been absorbed
will be reemitted and (τ ) that it will be destroyed (its energy turned into thermal
energy). The notation for the destruction probability per scattering can already by
found in Eddington (1929) and has been continuously in use (e.g. Thomas 1957).
For continuous spectra (monochromatic scattering), kd (τ ) is due to bound-free
and free-free transitions and ks (τ ), the scattering coefficient, is due to, e.g., Rayleigh
or Thomson scattering. For a spectral line, kd (τ ) is the absorption coefficient from
the lower level of the transition and ks (τ ) is related to the spontaneous radiative
deexcitation from the upper level of the transition. For a two-level atom, the
destruction coefficient may be written as
C21 (τ )
(τ ) = , (2.17)
A21 + C21 (τ )
where A21 is the Einstein coefficient for spontaneous emission, typically around
107 − 108 s−1 , and C21 (τ ) a rate of de-population of the excited level by inelastic
2.1 The Integro-Differential Radiative Transfer Equations 17
Here Q∗ (τ ) stands for Q∗ν (τ ). Equation (2.18) with = 0 describes the radiative
equilibrium of a grey stellar atmosphere (see Sect. 2.2.2).
It is convenient to rewrite the radiative transfer equation as
∂I
μ (τ, μ) = I (τ, μ) − S(τ ), (2.19)
∂τ
1 For a resonance line, the lower level is the atomic fundamental level.
18 2 Radiative Transfer Equations
The first term expresses the contribution of the photons that have been scattered at
least once and the second term yields the primary source of photons. The source
function depends only on the space variable and satisfies an integral equation of the
convolution type (see Sect. 2.2). For this reason, it plays a fundamental role in the
solution of radiative transfer equations.
ν − νc
x= , (2.21)
νD
where νc is the frequency of the line center and νD , the Doppler width. The
Doppler width measures the broadening of the absorption line profile by small scale
velocity fields, thermal and turbulent. It has the form
νc 2 2
νD = (v + vturb )1/2 , (2.22)
c th
with vturb the turbulent velocity and vth = (2kT /M)1/2, the thermal velocity. We
recall that k is the Boltzmann constant, T the temperature of the medium, M the
mass of the scattering atom and c the speed of light. The Doppler width varies
in general slowly inside a stellar atmosphere. Here we take it as a constant. The
frequency νc being very large for spectral lines in the optical domain (νc 1015 −
1016 Hz), we make the usual assumption that x varies from minus infinity to plus
infinity, with x = 0 at line center.
Introducing now R(ν, ν ) = ϕ(ν)ϕ(ν ) into Eq. (2.15) and the frequency variable
x, we can write the radiative transfer equation as
∂I
μ (τ, x, μ) = ϕ(x)I (τ, x, μ)
∂τ
+∞ +1
1
−(1 − ) ϕ(x) ϕ(x )I (τ, x , μ ) dμ dx − ϕ(x)Q∗x (τ ). (2.23)
2 −∞ −1
Over the width of a spectral line, the frequency variation of the Planck function
is in general very small and the value of Q∗x (τ ) can be taken at the line center
frequency. Here we assume that Q∗x (τ ) is independent of frequency and denote it
Q∗ (τ ). For the standard two-level atom model (see e.g. Mihalas 1978), is given
by Eq. (2.16) and may take very small values, say, 10−4 or less. The normalization
of the absorption profile in Eq. (2.8) becomes
+∞
ϕ(x) dx = 1. (2.24)
−∞
The optical depth τ is known as the frequency integrated or frequency averaged line
optical depth.
For spectral lines in the optical range, the absorption profile is in general a Voigt
function. A Voigt function is the convolution of a Lorentz function and of a Doppler
function, namely
+∞ e−y
2
a
U (x, a) = √ dy. (2.25)
π π −∞ a 2 + (x − y)2
20 2 Radiative Transfer Equations
√
The function H (x, a) = π U (x, a) is the so-called Voigt function. In contrast to
U (x, a) it is not normalized to unity. Here we refer to U (x, a) as the Voigt function.
The Lorentz function,
1 a
ϕL (x) = , (2.26)
π a + x2
2
takes into account the natural width of the energy levels and their broadening by
collisions with electrons and other species. The parameter a is a positive constant,
called the Voigt parameter of the line, defined by a = /4πνD , where , measured
in s −1 , is the line damping rate (see, e.g., Landi Degl’Innocenti and Landolfi 2004,
p. 590). In stellar atmospheres, typical values for a are 10−1 –10−3 . The Doppler
function,
1
ϕD (x) = √ e−x ,
2
(2.27)
π
1 a 3
√ e−x +
2
U (x, a) [1 + 2 + . . .]. (2.28)
π πx 2 2x
Voigt profiles for three different values of the parameter a smaller than one are
shown in Fig. 2.2. The Doppler profile corresponds to a = 0.3
The exact methods of solutions presented in this book hold if ϕ(x) is an even
function of x, is decreasing to zero monotonously as |x| → ±∞ and is integrable,
that is, satisfies some normalization condition such as Eq. (2.24).
For complete frequency redistribution, Eq. (2.23) can also be written as
∂I
μ (τ, x, μ) = ϕ(x)[I (τ, x, μ) − S(τ )]. (2.29)
∂τ
The source function S(τ ), which contains the primary sources of photons and the
contribution from the scattered photons, is defined by
+1 +∞
1
S(τ ) ≡ (1 − ) ϕ(x)I (τ, x, μ) dμ dx + Q∗ (τ ). (2.30)
2 −1 −∞
Fig. 2.2 The Doppler profile, and several Voigt profiles with different values of the parameter a.
The left panel shows ϕ(x) and the right panel log ϕ(x). In the right panel, the Lorentzian wing
regime characterized by ϕ(x) ∼ 1/x 2 can be observed for |x| larger than (− ln a)1/2
2hνc3 gl nu
S(τ ) = , (2.31)
c2 gu nl
where νc is the frequency at line center, h the Planck constant, c the speed of
light, gl , nl and gu , nu the statistical weights and the populations of the lower
and upper levels of the transition. The populations of the levels satisfy statistical
equilibrium equations, which describe the balance between population and de-
population by radiative and collisonal processes. When collisions dominate over
radiative processes, the populations of the levels follow the Saha–Boltzmann law
and the source function reduces to the Planck function. This situation is known
in Astronomy as Local Thermodynamic Equilibrium (LTE). In contrast, when
radiative processes have a dominant role, which is the case for strong resonance
lines, the source function will depend on the radiation field. For simple atomic
models it will take the form written in Eq. (2.30), where Q∗ (τ ) is the Planck
function at the frequency of the line center. This physical picture, known as Nonlocal
Thermodynamic Equilibrium (NLTE) was stressed in the late fifties by J. T. Jefferies
and R. N. Thomas (for historical references, see Hubeny and Mihalas 2015, p. 13).
For spectral lines formed in NLTE, the determination of the radiation field demands
that statistical equilibrium equations and radiative transfer equations be solved
simultaneously (see, e.g., Landi Degl’Innocenti and Landolfi 2004; Hubeny and
Mihalas 2015).
22 2 Radiative Transfer Equations
Primary photons can be created inside the medium, but can also come from outside.
It is a typical situation for planetary atmospheres, for which the main and usually
only source of energy comes from an outside illumination. The radiation field inside
the atmosphere and the reemitted fraction, known as the albedo, can be derived
from the radiative transfer equation. There are two different ways of handling the
incident radiation: as a boundary condition for the radiative transfer equation, or by
introducing a diffuse field. We describe this method, for a one-dimensional medium,
say a semi-infinite one, first for monochromatic scattering and then for complete
frequency redistribution. Diffuse fields are used in several chapters, for unpolarized
and polarized radiation.
We assume that a field I inc (μ) is incident on the surface at τ = 0. The idea of
the method is to separate the field in two parts: a direct field and a diffuse field. The
diffuse field denoted I d (τ, μ) is defined by
With our convention, μ < 0 for the radiation incident on the surface and μ > 0
for the radiation emerging from the medium. The diffuse field is zero at the surface
for incoming directions. It is equal to the total field for the outgoing directions and
at τ = 0 yields the emerging radiation. Introducing Eq. (2.32) into Eq. (2.18), one
obtains for I d (τ, μ) the equation
+1
∂I d 1
μ (τ, μ) = I d (τ, μ) − (1 − ) I d (τ, μ ) dμ − Q∗t (τ ), (2.33)
∂τ 2 −1
with
1
1
Q∗t (τ ) ∗
= Q (τ ) + (1 − ) I inc (−μ)e−τ/μ dμ. (2.34)
2 0
The second term in the right-hand side describes the scattering of the incident field
inside the medium. The important point is that it depends only on τ . The source
function for the diffuse field is
+1
1
S(τ ) = (1 − ) I d (τ, μ) dμ + Q∗t (τ ). (2.35)
2 −1
It satisfies Eq. (2.23) with the internal primary source Q∗ (τ ) replaced by
1 +∞
1
Q∗t (τ ) = Q∗ (τ ) + (1 − ) ϕ(x)I inc (x, −μ) e−τ ϕ(x)/μ dx dμ.
2 0 −∞
(2.38)
The second term in the right-hand side describes the scattering of the incident
photons and depends only on τ , just as with monochromatic scattering.
Radiative transfer problems with no internal source but an incident radiation
are usually referred to as diffuse reflection problems. Thanks to the introduction
of a diffuse field, any exact method of solution applicable to a problem with an
internal primary source, can also be employed when the the medium is illuminated
from outside. With the Case singular eigenfunction expansion method described in
Chap. 10, it is actually preferable to work with the original radiation field and the
associated boundary condition.
We now construct integral equations of the convolution type for the monochromatic
and for the complete redistribution source functions. The method is quite simple.
It suffices to introduce the solution of the transfer equations, expressed in terms of
S(τ ), into the defining equations for S(τ ). Monochromatic scattering and complete
frequency redistribution lead to similar integral equations.
where T and τ are two points inside the medium. When T > τ , Eq. (2.39) holds for
μ > 0 and when T < τ , it holds for μ < 0. To calculate the radiation field we need
boundary conditions. They depend on the geometry of the medium.
For an infinite medium, physically meaning full solutions of the radiative transfer
equation should grow less rapidly than an exponential at plus and minus infinity.
This condition, applied at T = +∞ for μ > 0 and at T = −∞ for μ < 0, implies
that the first term in the right-hand side of Eq. (2.39) is zero. Hence, in an infinite
Random documents with unrelated
content Scribd suggests to you:
The Project Gutenberg eBook of Studies in
Greek Scenery, Legend and History
This ebook is for the use of anyone anywhere in the United
States and most other parts of the world at no cost and with
almost no restrictions whatsoever. You may copy it, give it away
or re-use it under the terms of the Project Gutenberg License
included with this ebook or online at www.gutenberg.org. If you
are not located in the United States, you will have to check the
laws of the country where you are located before using this
eBook.
Language: English
BY
PAGE
1. Pausanias and his Description of Greece 1
2. Oropus 160
3. Rhamnus 163
4. Marathon 165
5. Prasiae 174
6. Mount Hymettus 178
7. Mount Pentelicus 182
8. Phyle 187
9. The Port of Athens 191
10. The Sacred Way 209
11. The Hall of Initiation at Eleusis 214
12. Eleutherae 216
13. Megara 219
14. The Scironian Road 220
15. The Isthmus of Corinth 223
16. The Bath of Aphrodite 226
17. The Prospect from Acro-Corinth 227
18. The Capture of Corinth by Aratus 228
19. Sicyon 232
20. Phliasia 233
21. Nemea 237
22. The Pass of the Tretus 238
23. Mycenae 242
24. The End of the Mycenaean Age 245
25. Mount Arachnaeus 248
26. Epidaurus 249
27. The Temple in Aegina 251
28. The Sanctuary of Poseidon in Calauria 252
29. Troezen 253
30. From Troezen to Epidaurus 255
31. Methana 260
32. Nauplia 261
33. The Springs of the Erasinus 263
34. The Lernean Marsh 266
35. The Anigraean Road 269
36. The Battlefield of Sellasia 270
37. Sparta 271
38. Mistra 274
39. On the Road from Sparta to Arcadia 278
40. Cape Malea 279
41. Monemvasia 281
42. Maina 282
43. Pharae and the Messenian Plain 284
44. Messene 285
45. On the Road to Olympia 287
46. Olympia 290
47. Phidias’s Image of Olympian Zeus 292
48. The Hermes of Praxiteles 293
49. Lasion 296
50. The Erymanthus 299
51. The Monastery of Megaspeleum 300
52. The Gulf of Corinth 301
53. On the Coast of Achaia 303
54. Pellene 304
55. The Road from Argos to Arcadia 306
56. Mantinea 308
57. The Road to Stymphalus 310
58. The Lake and Valley of Stymphalus 312
59. The Lake of Pheneus 315
60. From Pheneus to Nonacris 320
61. The Fall of the Styx 324
62. The Valley of the Aroanius 330
63. The Springs of the Ladon 331
64. The Gorge of the Ladon 333
65. Aliphera 336
66. Dimitsana 337
67. Gortys 338
68. The Plain of Megalopolis 342
69. The Cave of the Black Demeter 343
70. The Temple of Apollo at Bassae 345
71. The Temple of Artemis at Aulis 346
72. Glaucus’s Leap 347
73. Evening on the Euripus 349
74. The Copaic Lake 349
75. The Great Katavothra 355
76. The Vale of the Muses 357
77. Hippocrene 358
78. Lebadea 359
79. The Boeotian Orchomenus 361
80. The Plain of Chaeronea 364
81. Panopeus 364
82. Near Hyampolis 365
83. Tithorea 366
84. From Amphissa to Gravia 369
85. Daulis 371
86. The Cleft Way 373
87. Delphi 374
88. Aeschines at Delphi 378
89. The Pythian Tune 379
90. The Lacedaemonian Trophy at Delphi 380
91. The Gods in Battle 382
92. The Sibyl’s Wish 384
93. Orpheus in Hell 386
94. The Acheron 387
95. A Ride across Parnassus 389
96. Pericles 392
ebookbell.com