Sei sulla pagina 1di 312

Contents

PREFACE xv
PREFACE TO THE FIRST EDITION xix

l. The Radiation Field


1-1. The Specific Intensity 2 .
Macroscopic Definition 2
Photon Distribution Function 4
Invariance Properties 4
Observational Significance 5 ........

1-2. Mean Intensity and Energy Density 5


Macroscopic Description S
Photon Picture 6 ·
Equilibrium Value 6
Electromagnetic Description 7

1-3. The Flux 9


Macroscopic Description 9
Photon Energy Flux I0
The Poynting Vector 11
Observational Significance 11

1-4. The Radiation Pressure Tensor 12.


Macroscopic Description and the Photon Momentum
Flux 12
Relation of the Pressure Tensor to Volume Forces 13
The Maxwell Stress Tensor 14
Limiting Cases: Symmetry, Isotropy, Equilibrium,
Plane Waves 16
Variable Eddingto1. Factors 18
viii Contents Contents ix

2. The Equation of Transfer 19 4-5. Continuum Scattering Cross-Sections 105


Thomson Scattering I06
2-1. The Interaction of Radiation with Matter 20 Rayleigh Scattering I06
Distinction between Scattering and Absorption-Emission
Processes 20
.
The Extinction Coefficient 23
5. The Equations of Statistical Equilibrium 108
The Emission Coefficient 25 5-1. Local Thermodynamic Equilibrium 109
2-2. The Transfer Equation 30 The Maxwellian Velocity Distribution 110
The Boltzmann Excitation Equation 110
Derivation 30
The Transfer Equation as a Boltzmann Equation 32 The Saha Ionization Equation 112
Spherical Geometry 33 5-2. The LTE Equation of State for Ionizing Material 114
Optical Depth and the Source Function 34 Charge and Particle Conservation 115
Boundary Conditions 36 Solution by Linearization 117
Simple Examples 37
Formal Solution 38 5-3. The Microscopic Requirements of LTE 119
The Schwarzschild-Milne Equations 40 Detailed Balance 119
The Nature of the Radiation Field 120
2-3. Moments of the Transfer Equation 43 The Electron Velocity Distribution 121
2-4. The Condition of Radiative Equilibrium 47 The Ionization Equilibrium 123
2-5. The Diffusion Approximation 49 The Excitation Equilibrium 126
5-4. The Non-LTE Rate Equations 127
General Form 127
3. The Grey Atmosphere 53 Radiative Rates 128
3-1. Statement of the Problem 53 Collisional Rates 131
3-2. Relation to the Nongrey Problem: Mean Opacities 56 Autoionization and Dielectronic Recombination 134
Flux-Weighted Mean 56 Complete Rate Equations 137
Rosseland Mean 57 5-5. The Non-LTE Equation of State 140
Planck and Absorption Means 59 Limiting Cases 140
Summary 60 Linearization 143
3-3. Approximate Solutions 60
The Eddington Approximation 60 6. Solution of the Transfer Equation 146 I\)
Iteration: The Unstild Procedure 63
The Method of Discrete Ordinates 64 6-1. Iteration: The Scattering Problem 147 I
6-2. Eigenvalue Methods 150
3-4. Exact Solution 71 6-3. The Transfer Equation as a Two-Point Boundary Value
3-5. Emergent Flux from a Grey Atmosphere 73 Problem 151
3-6. Small Departures from Greyness 74 Second-Order Form of the Equation of Transfer 151
BounC,ary Conditions .153
Difference-Equation Representation 153
4. Absorption Cross-Sections 77 The Feautrier Sollltion 156
The Rybicki Solution 158
4-1. The Einstein Relations for Bound-Bound Transitions 77 Computation of the Flux 161
4-2. The Calculation of Transition Probabilities 81
The Classical Oscillator 81
Quantum Mechanical Calculation 84 7. Model Atmospheres 162
Application to Hydrogen 88 ·
Transition Probabilities for Light Elements 91 7-1. The Classical Model-Atmospheres Problem: Assumptions and
Restrictions 162
4-3. The Einstein-Milne Relations for the Continuum 94 7-2. LTE Radiative-Equilibrium Models 164
4-4. Continuum Absorption Cross-Sections 96 The Opacity and Emissivity: Continua and Line-Blanketing 165
Hydrogen 98 Hydrostatic Equilibrium 170
The Negative Hydrogen Ion 102 Radiative Equilibrium: Temperature-Correction
Other Ions of Hydrogen I04 Procedures 171
Helium 104 A Linearization Method 180
x Contents
Contents xi
7-3. Convection and Models for Late-Type Stars 185 10. Classical Treatments of Line Transfer 308
The Schwarzschild Stability Criterion 186
Mixing-Length Theory 187 I 0-1. Characterization of the Problem 308
Convective Model Atmospheres 190 10· 2 · The Milne-Eddington Model 310
Definition 310
7-4. Results of LTE Model-Atmosphere Calculations for Early-Type
Scattering Lines 311
Stars 192 Absorption Lines 312
Emergent Energy Distribution 193 Center-to-Limb Variation 314
Temperature Structure 205 Schuster Mechanism 315
7-5. Non-LTE Radiative-Equilibrium Models for Early-Type Stars 216 10-3. The Tl1eoretical Curve of Growth 316
Solution by Iteration: Detailed Balance in the Lines 218 10-4. The Empirical Curve of Growth 321
Formation of the Lyman Continuum 222
The Complete Lincarization Method 230 10-5. LTE Spectrum Synthesis with Model A~mospheres 328
Non-LTE Effects on Energy Distributions 234 11.
Temperature Structure: The Cayrel Mechanism and Non-LTE Line Transfer: The Two-Level Atom 332

7-6.
Line Effects 239
Extended Atmospheres 243
::·!·- .
~tusion, Destruction, Escape, and Thennalization 332
e Two-Level Atom without Continuum 336
Spherical Grey Atmospheres 245 The Source Function 336
Solution of the Transfer Equation in Spherical Geometry 250 Solution of the Transfer Equation 338
Extended Models for Early-Type Stars 255 The Thermalization Depth 341
Boundary-Value and Depth-Variation of the Source
7-7. Scmicmpirical Solar Models 258 Funcllon 343
Finite Slabs 346
The Effects of an Overlapping Continuum 350
8. The Line Spectrum: An Overview 268
Eff~ts of Depth-Variable Thermalization Parameters and
Line Profiles 355
8-1. Observational Quantities 269
8-2. The Physical Ingredients of Linc-Formation 271 11-3. The Two-Level Atom with Continuum 358
The Source Function 358
Classification of Lines 36 I
9. The Line'Absorptiori Profile 273
Line-Formation in the Presence of a Chromosphere 362
9-1. The Natural Damping Profile 274 11-4. Static Extended Atmospheres 367
Energy Spectra, Power Spectra, and the Autocorrelation 11-5. Comments on LTE Diagnostics 371
Function 274
The Damped Classical Oscillator 276
Quantum Mechanical Calculation 277
12. Non-LTE Line Transfer: The Multilevel Atom 374
w
I
9-2. Effects of Doppler Broadening: The Voigt Function 279 12-1. The EquivalentsTwo-Level-Atom Approach 376
9-3. Collision Broadening: Classical Impact Theory 281 Formulation 376
The Weisskopf Approximation 281 Application 380
The Lindholm Approximation 284 12-2. Eft'ec~s of Level Coupling: Source Function Equality in
Specific Cases 286 Multiplets 384
Validity Criteria 288
Photon Degradation and Conversion 385
9-4. Collision Broadening: Statistical Theory 289 Obsei:vational Indications of Source Function Equality 388
The Nearest-Neighbor Approximation 290 Solu11on of the Transfer Equation in Multiplets 391
Holtsmark Theory 291 12-3. T~e Complete Linearization Method 396
Dcbye Shielding and Lowering of the Ionization Potential 292
The Quasi-Static Ion Broadening of Hydrogen Lines 295 12-4. Light-Element Spectra in Early-Type Stars 401

9-5. Quantum Theory of Linc Broadening 297 13. Line Formation with Partial Frequency Redistribution 411
The Linc Profile 297
The Classical Path Approximation 299 13-1. Redistribution in the Atom's Frame 41 2
The Impact Approximation 301 13-2. Doppler-Shift Redistribution in the Laboratory Frame 415
Application to Hydrogen 303 General Formulae 415
Hydrogenic Ions 305 Resu Its for Speci fie Cases 4 I8
Neutral Helium Lines 306 Symmetry Properties 420
Other Light Elements 307 Applications 422
xii Contents
Contents xiii
13-3. Angle-Averaged Redistribution Functions 422
General Formulae 424 The Flui~-Frame Momentum Equations 545
Results for Specific Cases 427 The Inertial-Frame Equations 546
Symmetry Properties 432 15-4. Radiatively Driven Winds 549
13-4. Radiative Transfer with Partial Redistribution 433 Obsservational Evidence for Transsonic Winds in Early T
Formulation for a Two-Level Atom 433 tars 550 • ype
Methods of Solution 436 Basic D~namic~ of Radiation-Driven Winds 553
Results from Idealized Models 438 Lme-J?riven Winds in Of Stars 559
Application to Solar and Stellar Resonance Lines 442 Frontiers 566
REFERENCES 569
14. Radiative Transfer in Moving Atmospheres 447
GLOSSARY OF PHYSICAL SYMBOLS 587
14-1. The Transfer Equation in the Observer's Frame 449 INDEX 609
Formulation and Solution of the Transfer Equation 449
Line-Formation with Systematic Macroscopic Velocities in
Planar Atmospheres 453
Spherical Atmospheres: Low-Velocity Regime 459
Effects of Lines on Energy Balance in Moving Media 461
Linc-Formation in Turbulent Atmospheres 463
14-2. Sobolev Theory 471
Surfaces of Constant Radial Velocity 472
Escape and Thermalization in an Expanding Medium 478
Line Profiles 482
Multilevel Atoms: Application to Wolf-Rayet Stars 485
14-3. The Transfer Equation in the Fluid Frame 490
The Local Frequency Transformation 491
Lorentz Transformation of the Transfer Equation 493
Transformation of Moments of the Radiation Field 497
The Comoving-Frame Equation of Transfer 499
Solution for Spherically Symmetric Flows 503

15. Stellar Winds SI I


I 5-1. The Equations of Hydrodynamics for an Ideal Compressible
Fluid 512
Kinematics 512
The Equation of Continuity 515
Momentum Equations 516
Energy Equation 517
Sound Waves 518
The Rankine-Hugoniot Relations for Stationary Shocks 519
15-2. Coronal Winds 521
Expansion of the Solar Corona 523
One-Fluid Models of Steady, Spherically Symmetric Coronal
Winds 525
Transition to the Interstellar Medium 533
The Magnetic Field and Braking of Stellar Rotation 534
Detailed Physics of the Solar Wind 536
Stellar Coronae and Winds 538
15-3. Radiation Hydrodynamics 540
The Material Stress-Energy Tensor and the Radiating-Fluid
Equations of Motion 541
The Fluid-Frame Energy Equation 544
Preface

Since the appearance of the first edition of this book, there has been con-
tinuing rapid development of our understanding of stellar atmospheres, and
it has been clear to me for some time that a new edition was needed. One
of the major motivations for producing a new version of the book at this
time is the desire to describe the major advances that have been made-in
developing methods to solve the transfer equation in moving media, and in
the theory of stellar winds. As was true in the first edition, I have not
attempted to cover every possible aspect of the subject, but have again
treated a limited number of problems in some depth.
It was clear from the outset that, in view of the great demands made upon 01
the student's time in the now-crowded astrophysics curriculum (resulting I
from the explosion of our knowledge about the Universe), it was pointless
to write a book significantly longer than the first edition. Thus, to add new
material, it has been necessary to economize the presentation of the old
material, and to omit topics that are of specialized interest or that lie outside
the mainstream of the developments of primary importance to the book. In
particular, given that today's student is most likely to learn what he knows
about radiative transfer in a stellar-atmospheres course, but will be interested
in applying it to other physical situations, I have purposely shifted the
emphasis away from strictly stellar applications, and have developed the
transfer theory more generally and completely. I believe that a thorough
understanding of the radiative transfer theory presented in this book will
equip the student to attack a wide variety of transfer problems, whether in the
laboratory, the atmospheres of stars and planets, the interstellar medium, X-
ray sources, or quasars. Further, I have added exercises in which the student
is asked to fill in missing steps of derivations, or to apply the theory himself
to simple examples. In most cases the exercises are quite straightforward
xvi Preface
Preface xvii
and should require only a few minutes work; but some of the exercises in
Chapter 7 require substantial effort and would make good class projects. Direc~or of H.A.O., the scientific environment in which this book could
Ideally the material in this book should be taught in a course lasting two be wr~tten. I also thank Paulina Franz for converting hundreds of a f
0
my ~p1dery handwriting into smooth typed copy, Kathlyn Auer~ p ge~
quarters, covering Chapters 1- 7 in the first quarter, and Chapters 8-15 in
the second. If an entire year (two semesters) is available, the book should be the index, and Pat _B_rewer of W. H. Freeman and Company for ~:~:i:c:'i~!
and _careful superv1s1on of the production process.
supplemented with extra material on subjects of interest to the instructor
and students, perhaps drawn from problems of solar physics, stellar spec-
troscopy, pulsating atmospheres, peculiar stars, abundance analyses, or
_Finally, I th~nk. my_ father, M. D. Mihalas, for his unintentional b
pncel~ss) co~tnbut,~n in teaching me, through the example of his Ii~ (
meaning ofo.1>w1tr.no1011a1a and </nJ..or1111rx. e, t e
t
many others. If only one semester is available, I recommend omitting, first,
Chapters 4 and 9 (which are more physics than astrophysics); next, Chapters 3 Oxford, England
and IO (which are fairly elementary and may well have been covered in an October, 1977 Dimitri Mihalas
earlier course); and, finally, if necessary, Chapter 13 (which is not absolutely
essential for a basic understanding of line-formation).
In any case, many fascinating subjects will inevitably be omitted, and
teacher and student alike may feel frustrated, as I have been in writing the
book, that a more complete coverage is not possible. Again and again I
have felt like the traveler in Frost's "The Road Not Taken" (388, 105)*,
in choosing one of two equally fair paths, knowing full well that way would
lead on to way, and that I should not return to the other. I only hope that
the students will discover for themselves these other charming paths and
will spend a pleasant lifetime in their exploration.
It is no longer possible for me to acknowledge fairly the many people
who have helped me learn about stellar atmospheres and line:formation,
and I shall not try here, beyond offering a sincere thanks to all in whose
debt I am. But I would be remiss if I did not specifically thank Lawrente
Auer, David Hummer, and George Rybicki, who (as colleagues, critics,
teachers,· collaborators, and friends) have greatly deepened and enlarged O>
my understanding of the material in this book. Further, I wish to record my I
great debt to Professor W.W. Morgan of Yerkes Observatory. His en-
couragement has stimulated much of the work I have done in the past
several years, and his wise counsel has greatly enhanced its value. I also
thank him for sharing with me a few glimpses of his perception of the nature
of scientific method from the lofty point at which he can view it.
I wish in addition to thank the people who have helped with the writing
of this book: Barbara Mihalas, for reading and correcting the manuscript
and the typescript; Tom Holzer and Richard Klein for reading and com-
menting upon Chapter 15; and David Hummer and Paul Kunasz for reading
the typescript and offering many corrections and suggestions. Thanks also
are due to Gordon Newkirk for helping to provide, through his labors as
• A NOTE ABOUT REFERENCES: Referencesarelistedseriallyattheendofthetext,andaredenoted
in the text with boldface numbers-e.g., (105). Additional information, such as a page or chapter
citation, will be indicated following the refer11nce number-.g., (105, 27) or (105, Chap. 4).
Citations to two or more references are separated by semicolons---(1.g., (105, 27; 388, I05).
Preface

Since the appearance of the first edition of this book, there has been con-
tinuing rapid development of our understanding of stellar atmospheres, and
it has been clear to me for some time that a new. edition was needed. One
of the major motivations for producing a new version of the book at this
time is the desire to describe the major advances that have been made-in
developing methods to solve the transfer equation in moving media, and in
the theory of stellar winds. As was true in the first edition, I have not
attempted to cover every possible aspect of the subject, but have again
treated a limited number of problems in some depth.
It was clear from the outset that, in view of the great demands made upon
the student's time in the now-crowded astrophysics curriculum (resulting --..J
from the explosion of our knowledge about the Universe), it was pointless I
to write a book significantly longer than the first edition. Thus, to add ne..y
material, it has been necessary to economize the presentation of the old
material, and to omit topics that are of specialized interest or that lie outside
the mainstream of the developments of primary importance to the book. In
particular, given that today's student is most likely to learn what he knows
about radiative transfer in a stellar-atmospheres course, but will be interested
in applying it to other physical situations, I have purposely shifted the
emphasis away from strictly stellar applications, and have developed the
transfer theory more generally and completely. I believe that a thorough
understanding of the radiative transfer theory presented in this book will
equip the student to attack a wide variety of transfer problems, whether in the
laboratory, the atmospheres of stars and planets, the interstellar medium, X-
ray sources, or quasars. Further, I have added exercises in which the student
is asked to fill in missing steps of derivations, or to apply the theory himself
to simple examples. In most cases the exercises are quite straightforward
xx Preface to the First Edition

students at Princeton University, the University of Colorado, and the Uni-


versity of Chicago. It represents what I feel is a minimum background for a
student who wishes to understand the literature and to do research in the
field. Naturally, it has been necessary to be selective in the material presented.
In writing this book, I had in mind the goal of providing a basic synopsis
of the theory that can be covered in two quarters, with the hope that the
content of the third quarter of the nonnal academic year will be drawn
by the instructor (and the students) from the current literature on topics of
special interest to them. Although emphasis is given to the more modern
approaches, I have also attempted to give a coherent review of the older
methods and results. I feel it is important for students to be familiar with
these classical approaches so that they will be aware of the limitations of
such approaches and the conclusions based upon them.
It has been tempting to include a wider range of subjects, but I have avoided Stellar Atmospheres
doing so in the belief that it is more worthwhile for the student to consider a
smaller number of topics in depth than attempt to survey tire entire field super-
ficially. In this vein, I have purposely limited the comparison of theory with
observation to a few of the more crucial and illustrative examples. Moreover,
I have restricted most of the theoretical discussion to what may be called the
classical stellar-atmospheres problem-Le., atmospheres in hydrostatic,
radiative, and steady-state statistical equilibrium. This is ample material for a
two-quarter course and is understood well enough to require little speculation.
Even within this problem, I have limited the variety of techniques treated. For
example, I personally favor using differential equations over using integral ,
equations to solve transfer problems. Thus, although the latter method has,
enjoyed wide application and good success, particularly in the hands of the en
Harvard-Smithsonian Astrophysical Observatory group, there is little discus- I
sion ofit in this book. Thisomissionisnotarbitrary, however, but is based upon
the view that, since the two methods are mathematically equivalent, discussion
of one suffices and, in addition, that the one I have chosen seems to offer more
promise in future applications-for example, to situations involving hydro-
dynamics (wherein lies the real frontierofthesubject). On the other hand, in my
experience, the physics background of astronomy students is often uneven; I
have, therefore, not hesitated to develop those aspects of physical theory that
are of special interest to the atmospheres problem. In any case, I hope that users
of this book will find it a helpful outline, which they can edit, alter, and enlarge
upon as their needs dictate.

Williams Bay, Wisconsin Dimitri Mihalas


November 1969
1
The Radiation Field

co
From quantitative examination of the spectrum of a star, information can I
be obtained about the frequency distribution of the emergent radiation field.
We observe both broad, smooth expanses of continuum and spectrum lines,
where the frequency variation is quite abrupt. The entire spectrum contains
an enormous wealth of information, and the primary goal of the theory of
stellar atmospheres is to develop methods that can recover this information.
To this end we must be able to describe the flow of energy through the
outermost layers of a star, and to predict the observable characteristics of
the emergent radiation. We apply known physical laws that specify the
interaction of radiation with stellar material, and derive mathematical models
from which we compute theoretical estimates of observables. We then com-
pare theory and observation, and attempt to infer the physical conditions
in stellar atmospheres. Such analyses can provide information about the
structure of the envelope (important as a boundary condition for studies of
stellar structure), modes of energy transport in the atmosphere, chemical
abundances, rates of mass loss, and calibrations for converting observational
parameters (e.g., M. and B - V) into theoretically interpretable numbers
(luminosity and temperature). By studying large numbers of stars we can
2 The Radiation Field 3

n
establish relations of, say, chemical composition to stellar distributions,
kinematics, and dynamics; this information provides clues in developing an
understanding of the structure and dynamics of the Galaxy as a whole.
The program outlined above is ambitious, and it is not an easy one to
carry out successfully. The observational data are often difficult to acquire,
FIGURE 1-1
have limited precision, and are the results of very complicated physical Pencil of radiation used to define specific
structures. Often oµr physical theories are only primitive, and yet even intensity. The vector n is the direction of
these may lead to extremely complicated mathematical systems. But the propagation of the radiation, while sis the
key issue is that the information we deduce from stellar spectra will be a unit vector perpendicular to the element of
close approximation to reality only if the underlying physical theory is area dS.
sound and comprehensive. We must, therefore, devote c0nsiderable attention
to the development of an approach that correctly includes the essential
physics.
In this chapter we introduce the basic definitions required to charac- dS
terize the radiation field itself. The radiation field is treated from three
points of view-using macroscopic, electromagnetic, and quantum descrip-
tions. Each of these approaches yields useful information and, taken together, energy transported by radiation of frequencies (v, v + dv) across an element
they provide a full picture of the nature of the field. We ignore polarization, of area dS into a solid angle dw in a time interval dt is
but carry along an assumed time-dependence so that in later work we can
derive equations of radiation hydrodynamics. In subsequent chapters we bit = /(r, n, v, t) dS cos 0 dw dv dt (1-1)
shall consider how the radiation interacts with the material an~ is transported
through the atmosphere (Chapter 2), and shall write down detailed,descrip- where 0 is the angle between the direction of the beam and the normal to
tions of the atomic parameters that specify the absorptivity of the material the surface (i.e., dS cos 0 = n · dS); see Figure 1-1. The dimensions of/ are
(Chapter 4) and the mechanisms that determine the distribution of atoms ergs cm - 2 sec- 1 hz- 1 sr- 1 • As it has just been defined, the specific intensity
over available bound and free states (Chapter 5). After consideration of the-' provides a complete description of the radiation field from a macroscopic
grey probleQl, which supplies an ideal testing ground of methods and shows point of view. 0
clearly the overall approach used (Chapter 3) and development of general In this book consideration will be given only to one-dimensional problems I
mathematical techniques for solving transfer equations (Chapter 6), we in planar or spherical geometry; that is, the atmosphere will be regarded as
discuss the central problem of the book: the construction of model atmo- composed of either homogeneous plane layers·or homogeneous spherical
spheres (Chapter 7). We then examine the physics of line fo~mation f~r a shells. In planar geometries we employ Cartesian (x, y, z) coordinates with
given (static) model (Chapters 8-13), and the methods used to mfer che~m~al planes of constant z being the homogeneous layers; we can then ignore the
abundances in and physical characteristics of stellar atmospheres. Radiative (x, y) dependence of all variables, as well as derivatives with respect to x
transfer in moving atmospheres is then analyzed (Chapter 14) and, finally, and y. It is convenient to introduce polar and azimuthal angles (0, q,) to
all of the preceding developments are applied in a discussion of stellar specify n; we then haven· k = cos 0, n · i = sin 0 cos q,, n · j = fin 0 sin q,.
winds (Chapter 15). For one-dimensional planar geometry / will clearly be independent of <J,;
hence we can write I = I(z, 0, v, t); z is measured as positive upward in the
atmosphere (opposite to the direction of gravity). In spherical geometry
spatial location is specified by (r, e, <I>); but for spherical symmetry, I will
1-1 The Specific Intensity depend upon r only. The direction of the radiation can be specified in terms
of the polar and azimuthal angles (0, <J,), now measured with respect to a
MACROSCOPIC DEFINITION
unit vector f in the radial direction. Spherical symmetry again implies
The specific intensity /(r, n, v, t) of radiation at position r, traveling in azimuthal invariance, and we can now write / = /(r, 0, v, t). We shall often
direction n, with frequency v, at time t is defined such that the amount of replace the variable 0 with µ = cos 0.
1-2 Mean Intensity and Energy Density s
4 The Radiation Field

Exercise. I-/: By use of Snell's law, n 1(v) sin 0 1 = n 2(v) sin 0 2 , in the calculation
P to P'. Thus it immediately follows from equation (1-3) that I. = I~. Note
of the energy passing through a unit area on the interface between two dispersive also that equation (1-3) implies that the energy received per unit area falls
media with differing indices of refraction, show that I .n.- 2
is a constant. off as the inverse square of the distance between P and P'.

PHOTON DISTRIBUTION FUNCTION OBSERVATIONAL SIGNIFICANCE

The radiation field can also be described in terms of a photon distribution The spatial invariance of the specific intensity implies that the actual
function fR which is defined such that fR(r, n, v, t) dw dv is the number of value of I at the source can be obtained from measurements of the amount
photons per unit volume at location r and time t, with frequencies on the of energy falling, in a given time, within a specified frequency band, onto a
range (v, v + dv), propagating with velocity c in direction n into a s~lid receiver of known collecting area (and detection efficiency) from a source
angle dw. Each photon has an energy hv. The number of photons crossing subtending a definite solid angle. The requirement that dw must be specified
an element dS in time dt is fR(c dt)(n · dS)(dw dv), so that the energy trans- limits the determination of J to sources that are spatially resolved-e.g.,
ported is {)8 = (chv)fR dS cos 0 dw dv dt; comparison of this expression nebulae, galaxies, the sun, planets, etc.
with equation (1-1) shows that In particular, for the sun, the radiation at a given point emerges-at a known
angle relative to the local normal (in a one-dimensional model); hence
/(r, n, v, t) = (chv)fR(r, n, v, t) (1-2) measurement of the center-to-limb variation of the radiation allows us to
determine the angular variation of/. Note that we do not, in general, see to
INVARIANCE PROPERTIES the same depth in the atmosphere along all rays; hence we do not obtain
the angular variation of I at some definite position (z) inside the atmosphere,
An important property of the specific intensity is that it has been defined
but rather at some point robs outside the atmosphere.
in such a way as to be independent of the distance between the source and
the observer if there are no sources or sinks of radiation along the line of Exercise 1-2: The angular diameter of the sun is 30'. Suppose that atmospheric
sight. Thus, consider that pencil of rays which ·passes through.both the seeing effects limit resolution to I"; show that this sets a lower bound on theµ
element of area dS at point P and the element dS' at P' (see Figure 1-2). for which we can infer/(µ) accurately, and determine this µ,. 10 •
Then the amount of energy /JI passing through both areas is
lJ8 = I, dS cos 0 dw dv dt = /JI' = I~ dS' cos 0' dw' dv dt (1-3) 1-2 Mean Intensity and Energy Density
where dw is the solid angle subtended by dS' as seen from P, and dw' is the MACROSCOPIC DESCRIPTION
solid angle subtended by dS as seen from P'. From Figure 1-2 we see that
dw = r- 2 dS' cos 0' while dw' = r- 2 dS cos 0, where r is the distance from In both the physical and the mathematical ·description of a radiation
field it is useful to employ various angular averages, or moments. Thus we
define the mean intensity to be the straight average (zero-order moment) of
the specific intensity over all solid angles, i.e.,

J(r, v, t) = (41t)- 1 f J(r, n, v, t) dw (1-4)


r
The mean intensity has dimensions ergs cm- 2 sec- 1 hz- 1• The element of
solid angle dw is given by dw = sin 0 d0 d</> = - dµ d<f>. If we consider one-
dimensional atmospheres, I is independent of</>, hence

FIGURE 1~2
Geometry used in proof ofinvariance or specific intensity. The points J(z, v, t) = (41t) _1 Jor2n d<f> f'_ 1
dµ l(z, µ, v, t) f'
= 21 _1 J(z, µ, v, t) dµ (1-5)
p and P' arc separated by a distance r. Area dS subtends a solid angle
dw' at P', and the area dS' subtends dw at P; I and I' are unit vectors The same result applies in spherical geometry with z replaced by r.
normal to dS and dS'.
6 The Radiation Field 1-2 Mean Intensity and Energy Density 7

To calculate the energy density in the radiation field on the frequency given by the Planck function Bv(T) = (2hv 3/c 2)(eh•/kT - 1)- 1 [see (S20),
range (v, v + dv), consider a small volume V through which energy flows (392, 365)]. Thus, in thermal equilibrium the monochromatic energy density
from all solid angles. The amount flowing from a particular solid angle dw is E~(v) = (4n/c)B,(T), and the total energy density is given by Stefan's law:
through an element of surface area dS of this volume is

<J8 = /(r, n, v, t)(dS cos 0) dw dv dt


E~ = (8nh/c 3 ) f
0
"' (~•/kT - W 1 v3 dv = aR T4 (1-10)

where aR = 8n 5 k4 /(15c 3 h3 ). Here, as elsewhere in this book, we denote a


Now consider only those photons in flight across V; if the path length across quantity computed from thermodynamic equilibrium relations with an
V is I, then the time they will be contained within V is dt = 1/c. Further, asterisk.
I dS cos 0 = dV, the differential element of V through which they sweep.
Hence the energy in dV coming from dco is lJ& = c- 1 /(r, n, v, t) dco dv dV; Exercise 1-3: Derive Stefan's law by substituting x = hv/kT, and expanding
by integrating over all solid angles and over the entire volume, we find the (e' - 1)- 1 = e-•o - e-•)- 1 as a power series in e-•. The sum obtained from
total energy contained in V, namely: the term-by-term integration is related to the Riemann zeta-function [see (4, 807)].

&(r, v, t) dv = c- 1 [f vdV~ dco /(r, n, v, t)] dv (1-6)


Stefan's law is valid in the interior of a star, and in the deeper layers of stellar
atmospheres, where thermal gradients over a photon mean-free-path are
extremely small, and the radiation becomes isotropic and thermalizes to its
But if we pass to the limit of infinitesimal V, I becomes independent of equilibrium value. At the surface, the radiation field becomes very anisotropic
position in V, and the integrations can be carried out separately. The mono- and has a markedly non-Planckian frequency distribution, as a result of
chromatic energy density, ER(r, v, t) = &(r, v, t)/V is thus steep temperature gradients and the existence of an open boundary through
which photons escape into interstellar space; here Stefan's law becomes
ER(r, v, t) = c- 1 p /(r, n, v, t) dw = (4n/c)J(r~ v, t) (1-7) invalid.

ER has dimensions of ergs cm- 3 hz- 1• The total energy density (dimensions: ELECTROMAGNETIC DESCRIPTION
ergs cm- 3 ) is found by integrating over all frequencies:
Electromagnetic theory provides an alternative description of the radi-
ER(r, t) ,;,. f
0
"' ER(r, v, t) dv = (4n/c) f0"' J(r, v, t) dv = (41t/c)J(r, t) (1-8) ation field; we shall show how a one-to-one correspondence can be made
between the macroscopic and electromagnetic descriptions of the radiation
I\)
I
field. The electromagnetic field is specified by Maxwell's equations [see,
PHOTON PICTURE
e.g., (331, Chap. 6)] which, in Gaussian units, are
It is easy to show that the results derived above are consistent with the
photon picture of the radiation field. By definition, fR(r, n, v, t) is the number
V· D = 4np (l-1 la)
of photons, per unit volume, of energy hv propagating in direction n into V·B =0 (1-11 b)
intervals dv dco. The energy density clearly is just this number, multiplied
by the energy per photon, summed over all solid angles: i.e., (V x E) + c- 1(oB/ot) = 0 (1-1 lc)

ER(r, v, t) = hv pfR(r, n, v, t) dco (1-9) and (V x H) - c- 1(0D/ot) = (41t/c)j (1-1 ld)

The electric field E is related to the electric displacement D in terms of the


But from equation (1-21 hvfR = c- 1 /, hence equation (1-9) is seen to be permittivity e, namely D = eE. Similarly, the magnetic induction B can be
identical with equation (1-7). expressed in terms of the magnetic field H and the permeability µ by the
relation B = µH. For vacuum, e = µ = 1. In equations (1-11), p is the
EQUILIBRIUM VALUE charge density and j is the current density j = pv associated with charges
In thermal equilibrium the radiation field inside an adiabatic enclosure moving with velocity v. The electric field and magnetic induction can be
is uniform, isotropic, time-independent, and has a frequency distribution derived from a scalar potential q, and a vector potential A, which are defined
8 The Radiation Field 1-3 The Flux 9

such that denotes the usual Dirac function. Substitution into equation (1-7) yields the
B=VxA (l-12a) energy density ER = c- 1 I 0 , a result that is intuitively obvious for a plane
wave propagating with velocity c. Therefore we obtain a correspondence
and E = - Vq, - c- 1(oA/ot) (1-12b) between the two descriptions by making the identification
Equation (1-12a) satisfies equation (1-llb), while (1-12b) satisfies (1-llc).
Because B is defined as the curl of A, the divergence of A may be specified
10 = cE//81t (1-18)
arbitrarily; one of the most convenient choices is to impose the Lorentz It will be shown below that this choice yields consistent relations between the
condition Poynting vector and Maxwell stress tensor and their macroscopic counter-
(1-13) parts. The results derived here apply, strictly, onty to a monochromatic
With this choice, Maxwell's equations can be reduced to plane wave, but are easily generalized to fields having arbitrary angle and
frequency distributions by summing over suitably chosen elementary plane
V2 q, - c- 2(o 2 q,/ot 2 ) = -41tp (l-14a) waves.

and V A - c- 2(o 2 A/ot


2 2
) = -(41t/c)j (l-14b)

The solutions of these equations can be written as [cf. (331, Chap. 6), (494,
1-3 The Flux
Chap. 19)],
f Irp(r',- t')r'I d3r'
q,(r, t) = (1-15a)
MACROSCOPIC DESCRIPTION

We define the flux of radiation sr;(r, v, t) as a vector quantity such that


s,; · dS gives the net rate of radiant energy flow across the arbitrarily oriented
A( ) _ 1 f p(r', t')v(r', t') d 3 , surface dS per unit time and frequency interval. Noting that n · dS = dS
and
t. c Ir - r'I
t - r
(1-15b)
cos 0, where 0 is the angle between the direction of propagation n and the
normal to dS, we immediately recognize that the flux can be derived from the
where, as indicated, p and v at r' are evaluated at the retarded time t' = specific intensity via equation (1-1), for ol as written there is, in fact, nothing
t - c- 1lr' - rl which takes into account the finite speed of propagation df more than the contribution of the pencil of radiation moving in direction n
electromagnetic waves. __.
to the net energy flux. Thus we merely sum over all solid angles and obtain
One of the most important solutions of Maxwell's equations is that for w
monochromatic plane waves in vacuum, propagating in direction n0 with
velocity c:
sr;(r, v, t) = f l(r, n, v, t)n dw (1-19) I

E(r, t) = E 0 cos[21t(kn 0 • r - vt)] (1-16a) The flux has dimensions: ergs cm- 2 sec- 1 hz-: 1• Note thats,; is the first
moment of the radiation field with respect to angle.
and H(r, t) =H 0 cos[21t(kn 0 • r - vt)] (1-16b)
In cartesian coord.inates we have
where k = ,1.- 1 = c- 1v. The vectors (E0 , H 0 , n0 ) form an orthogonal triad
with H0 = n0 x E0 , so it follows that IHol = IEol, The result obtained from
electromagnetic theory for the instantaneous energy density W(t) in the
(s,; "' :F 1 , :F,) = (p In" dw, p ln1 dw, p In, dw) (1-20)

field is
where dw = -dµ di/>, n" = (1 - µ 2 )¼ cos q,, n1 = (1 - µ 2 )¼ sin¢, n, = µ.
W(t) = (E · D + B · H)/81t (1-17)
If the radiation field i.s symmetric with respect to an axis, then there will be
Averaging in time over a cycle introduces a factor of (cos 2 wt)T = ½, and a ray-by-ray cancellation in the net energy transport across any surface whose
using the relations IEol = IHol andµ = e = 1 (for vacuum), equation (1-17) normal is perpendicular to that axis, and the net flux will be identically zero
<
reduces to W = W(t))T = E 0 2/81t. In terms of the macroscopic picture, a across this surface. In particular, for a planar atmosphere homogeneous in
x and y, only $If, can be nonzero; we shall therefore require only this com-
monochromatic plane wave propagating in direction n0 [specified by angles
(0 0 , ¢ 0 )] has a specific intensity I(µ, q,) = I O o(µ - µ 0 ) o(q, - 1/> 0 ) where o ponent of the flux, and shall refer to it as "the" flux, as if it were a scalar,
10 Tire Radiation Field
1-3 The Flux II

and write THE POYNTING VECTOR

:F(z, v, t) = 2n f~ 1 l(z, µ, v, t)µ dµ (1-21) In_ electromagnetic theory, the energy flux in the field is given by the
Poyntmg vector
Exercise 1-4: (a) Show that,,, and,, vanish in an atmosphere with azimuthal S = (c/41t)(E x H) (1-25)
(,J,) independence of I. (b) Show that in a spherically symmetric atmosphere
only,, is nonzero and is given by equation (1-21) with z replaced by r. (c) Eval- Considering a plane wave as in §1-2, the average power over a cycle is
uate F for/(µ) = :EI.µ•; show that only the odd-order terms contribute to F. ~S)r = c(E x H)r/41t = (c(E 2 )rn 0 )/41t = (cE0 2 )n 0 /81t. On the other hand
m terms of macroscopic quantities, the flux associated with a plane wave i~
In astrophysical work it is customary to absorb the factor of n appearing in
equation (1-21), and to write the astrophysical.flux as F(z, v, t) = n- 1:F(z, v, t).
Further, regarding the flux as one ofa sequence of moments with respect toµ,
F =j In dw = p1 c5(n -
0 n0 )n dw = /0 n0 (1-26)

one may define the Eddington flux . N~w using equatio~ (1-1~), it is clear that F defined by equation (1-26) is
1dent1cal to ~S~r· ~gam, this resul_t can be generalized to arbitrary angle and
H(z, v, t) = (41t)- 1:F(z, v, t) = 1f I
l(z, µ, t)µ dµ (1-22)
frequency d1stnbut1ons of the radiation field.

OBSERVATIONAL SIGNIFICANCE
which is in a form similar to equation (1-5) for the mean intensity.
. The energy received from a star by a distant observer can be related
PHOTON ENERGY FLUX
directly to the flux :F • emitted at the stellar surface. Assume that the distance
D between star and observer is very much larger than the stellar radius r so
The same results for the energy flux may be obtained from the description that all ray_s from star. to observer may be considered to be parallel. •The
of the radiation field in terms of photons. The net number of photons passing, energy received, per unit area normal to the line of sight, from a differential
with velocity c, through a unit surface oriented at angle 0 to the beam, pt.,r area on the _star is d/._ = _I. dw _where dw is the solid angle subtended by the
unit time, is clearly area, and I. 1s the spe~1fic 1_ntens1ty emergent at the stellar surface. Considering
the geometry shown m Figure 1-3 we see that r = r• sin0 so that the area of
N(r, v, t) =c pfR(r, n, v, t) cos 0 dw (1-23)

Each photon has energy hv, so the net energy transport must be

F(r, v, t) = (chv) pfR(r, n, v, t)n dw (1-24)


To observer
In view of equation (1-2), equation (1-24) is obviously identical to equation
(1-19).
Furthermor~, photons of energy hv propagating in direction n have mo-
mentum hvn/c. Thus it is clear that c- 1F · dS dt gives the net momentum
transport across the surface dS in time dt, by particles moving with velocity c.
It therefore follows that the momentum density associated with the radiation
field is GR = c- 2, ; we shall find further significance of this result in §2-3
and shall use it in §14-3. FIGURE 1-3
Geometry of measurement of stellar flux. The annulus on the surface
E.-cercise 1-5: Verify the assertion that c- 2 gp represents a momentum density; of the star has an area dS = 2irr dr = 2,r~ sin O cos 0 dO normal to
the line of sight; this area subtends a solid angle dw = dS/D 1 as seen
check units for consistency. by the observer.
12 The Radiation Field 1-4 The Radiation Pressure Tensor 13

a differential annulus on the disk is dS = 21tr dr = 2nr.2 µ dµ, and dw = specific intensity with the photon distribution function JR, we see that
21r(r./D) 2µ dµ. The radiation emitted from this annulus in the direction of
the observer emerged at angle 0 relative to the normal; hence the appropriate (1-30)
value of the specific intensity is /(r *' µ, v). Integrating over the disk, we find

/v = 21t(r./D) 2 f: I(r., µ, v)µ dµ = (r./D) 2 ~(r., v) = i o:,/~(r., v) (1-27)


The above expression clearly gives the net flux of momentum, in the j-
direction, per unit time, from radiation of frequency v, through a unit area
oriented perpendicular to the i-direction; this is precisely the definition of
where o:. is the angular diameter of the star. [In the above calculation we have pressure in any fluid, and hence justifies the term "radiation pressure".
assumed there is no radiation incident upon the surface of the star; i.e., The average of the diagonal components of P may be used to define a mean
/(r •• - µ, v) = O.] For unresolved objects (e.g., stars), we can measure only radiation pressure: ·
the flux. The energy received falls off as the inverse square of the distance (1-31)
(because the solid angle subtended by the disk varies as v- 2 ). If the angular
diameter is known, then the absolute energy flux measured at the earth can
be converted to the absolute flux at the star. But (n./ + n/ + n.2) = 1 for any unit vector n, hence in general
Exercise J-6: Show that the flux emergent from a small aperture in an adiabatic
enclosure (blackbody) is .F88(v) • nB,(T). Show that the integrated flux is 3-aa =
P(r, v, t) = (3c)- 1 f /(r, n, v, t) dw = ~ ER(r, v, t) (1-32)
aR T'' where a" = (c/4)a,t = 2n 5 k4 /(15h 3c 2) = 5.67 x 10- 5 ergcm- 2 sec- 1 deg- 4
is the Stefan-Boltzmann constant. However, it must be emphasized that despite the generality of this result, P
does not give the actual radiation pressure unless the radiation field happens
to be isotropic. In general the radiation field in stellar atmospheres is far from
isotropic, and ordinarily the numerical factor relating PR (a scalar parameter
1-4 The Radiation Pressure Tensor that can be used to calculate radiation forces) and the energy density ER
MACROSCOPIC DESCRIPTION AND THE PHOTON
exceeds½ (see below).
MOMENTUM FLUX
RELATION OF THE PRESSURE TENSOR TO VOLUME FORCES 01
The ~ean intensity and flux are the scalar and vector quantities given by
I
the zero and first angular moments of the specific intensity against the direc- Let us now examine the relation of the radiation pressure tensor to
tion cosines between the direction of propagation and an orthogonal triad. volume forces exerted by the radiation field. Consider an element of area dS;
The second moment yields a tensor quantity that we shall identify as the the flow, per unit time, of the i-component of momentum in the radiation
radiation pressure tensor (or radiation stress tensor), namely field across this element is Li PiJni dS, where the n/s are the dfrection cosines

P(r, v, t) = c- 1 f /(r, n, v, t)nn dw (1-28)


of the normal to dS. Now integrating over a closed surface S, and applying
the divergence theorem, we find

or, in component form,


Ps L Pljnj dS = fv L (0PIJ/ox1) dV = Si, (V · P), dV (1-33)
PIJ(r, v, t) = c- 1
f /(r, n, v, t)n 1n1 dw (1-29)
j j

where V is the volume enclosed by S. The integral on the left gives the net
The dimensions of P are ergs cm - 3 hz- 1• It is obvious that P is symmetric; flow, per unit time, of the i-component of momentum out of the volume
i.e., P,1 = P11• through the surface S; thus from the integral on the right we see that (V · P)1
The physical interpretation of P follows directly from the description of the must be the rate at which the i-component of the momentum density in the
radiation field in terms of photons. Thus, using equation (1-2) to replace the field decreases; i.e., c- 2 (ofF/ot) 1•
14 The Radiation Field IS

Hence for the radiation field alone (i.e., in the absence of absorbing or
emitting material) we have

Equation (1-34) is essentially identical to the usual momentum equation of


hydrodynamics for an ideal fluid with no applied forces (cf. §15-1). We shall
generalize this result to include interactions with material in §2-3.

THE MAXWELL STRESS TENSOR

In electromagnetic theory, the stress in the field is described by the


Maxwell stress tensor, which is defined such that
(aG •..Jat) = v •TM (1-35)
Here G.m is the momentum density associated with the electromagnetic
field. The components ofTM are:
X

T,1M = [ E1E1 + H,H1 - ~ olJ(E2 + H 2) ] / 41t (1-36) FIGURE 1-4


The plane electromagnetic wave generated by E and H
is propagating along the Poynting vector S in direction
where oil denotes the usual Kronecker o-symbol. By comparison of equations n0 • The angle 1/1 0 measures the rotation of E around S
(1-34) and (1-35) we see that the Maxwell stress tensor should be equal to out of the plane defined by n0 and k, the unit vector in
the z-direction.
the negative of the radiation pressure tensor; it is instructive to verify this
conclusion by direct calculation. __.
Consider a plane wave propagating in direction n0 ; from the macroscopic· Substitution of equations (1-38) and the corresponding equations for H into
definition of radiation pressure we have CJ)
equation (1-36) yields components of TM; for example, for T .. M we find
I
P = c- 1 p1nndw = c- 1 10 f o(n - n0 )nndw = c- 1 10 0 0 0 0 T.,M = [ E,2 + H,2 - ~(E 2 + H 2 ) ] /41i = E0 2 (sin 2 00 - 1)/41t
= (E//81t)n0 n0 (1-37)
(1-39)
which should equal TM for a plane wave. Choose the electromagnetic field
to yield a Poynting vector S along n0 ; in addition to (0 0 , cf, 0 ) we must also Averaging over time yields (T,.M)r = -(E 0 2/81t) cos 2 00 which is indeed
specify the polarization of the wave via the angle .J, 0 • Here I/Jo measures the -P.. ; note that the final result is independent of.J, 0 •
angle of rotation of E around S from the plane through n0 and k (the unit
vector in the z-direction); see Figure 1-4. It is easy to see that
Exercise 1-8: Calculate the remaining components of T"' and show that T"' =
-P, independent ofi/1 0 •
E" = E0(sin 1/1 0 sin cf, 0 - cos 1/1 0 cos cf, 0 cos 00 ) (1-38a)
The above results demonstrate that a complete correspondence exists
E1 = - E0(sin 1/1 0 cos cf, 0 + cos 1/1 0 sin cf, 0 cos 00 ) (1-38b) between electromagnetic theory and the macroscopic or photon descriptions
of the radiation field; we shall exploit this correspondence in a useful
E, = E0 cos 1/1 0 sin 00 (1-38c) way in §§14-3 and 15-3 where we will be able to use the known Lorentz-
transformation properties of electromagnetic field quantities to establish·
Exercise J-7: Derive expressions analogous to equation (1-38) for (H., H,, H,). those of their radiation-field-description counterparts.
16 The Radiation Field 1-4 The Radiation Pressure Tensor 17

LIMITING CASES: SYMMETRY, ISOTROPY, EQUILIBRIUM,


In general, p11 as defined by equation (1-41) will not equal P defined by
PLANE WAVES
equation (1-32), but the two become equal if the radiation field is isotropic.
In a one-dimensional planar or spherically-symmetric atmosphere the For an isotropic field,/ is independent ofµ, and equations (1-7) and (1-41)
radiation field is azimuthally invariant, hence the pressure tensor becomes immediately yield PR = ¼ER, so that equation (1-40) reduces to
diagonal

PR O O) l (3pR - ER
P(r, V, t) = (~R ~R ~ ) (1-44)
P(r, v, t) = 0 PR O - 0 ~) (140)
0 0 PR
( 2
0 0 PR 0 That is, when the radiation field is isotropic, the· radiation pressure tensor
is diagonal and isotropic, and may be replaced, for all purposes of computa-
where PR(z, v, t) = (4n/c)K(z, v, I) (1-41) tion, by a scalar hydrostatic pressure, or even eliminated entirely in terms of
and, in turn, ER. If in addition to being isotropic the radiation field has its thermal

K(z, v, t) -= 1 JI
2 _1 /(z, µ, v, t)µ
2dµ (1-42)
equilibrium value, then the monochromatic radiation pressure is

p~(z, v, t) = 31 E~(z, v, t) = (4n/3c)B.(T) (1-45)


is the second moment of the radiation field in Eddington's notation.
and the total radiation pressure is
Exercise 1-9: (a) Derive equation (1-40) for the conditions stated. (b) Show I
p~(z, t) = 4
(1-46)
that the same expression ror P is obtained in spherical symmetry relative to the
orthogonal triad (cf,, 0, r). (c) Veriry that the tensor shown in equation (1-40) is 3 aT
consistent with equation (1-32). a result first obtained by thermodynamic arguments (160, 55; 565, 123).
From equations (1-7) and (1-41) we see that PR is an average of/(µ) weighted
It is clear from equation (1-40) that, for a one-dimensional atmosphere, by µ 2 whereas ER is a straight average. If the radiation field becomes peaked
only two scalars (PR and ER) are sufficient to specify the full radiation pressure in the direction of radiation flow out of the atmosphere, the larger values
......
tensor. Further, for such atmospheres, derivatives·with respect to (x, y) or ofµ become more heavily weighted in PR (recall µ = 1 for 0 = 0) and PR
(0, cf>), in the planar and spherical cases respectively, are identically zero, and will exceed its isotropic limit of ½ER. The most extreme departure from
the only non vanishing components of the divergence of the radiation pressure isotropy occurs when the radiation flows in a plane wave. For a wave in the
tensor are outward direction we can write /(z, v, µ) = l(z, v) b(µ - 1); then K(z, v) =
(l-43a) J(z, v) = H(z, v), and PR(z, v) = ER(z, v). This extreme limit is approached
in the outermost layers of very extended stellar envelopes (or in nebulae).
in planar geometry-or, in spherical geometry,
in which the radiation field originates from a stellar surface that occupies
only a very small solid angle as seen from the point in question.

In this book we shall confine attention strictly to one-dimensional problems, £x£'rcis£' /./I: (a) Show that for a plane wave moving along one or the coordinate
and with the exception of further formal development of the equations of axes, the radiation pressure tensor has only one non-zero component. (b) Show
radiation hydrodynamics in §15-3, the full tensor description of the radiation that the pressure tensor is isotropic ir the angular dependence or the radiation
field will not be required; for ease of expression, we shall therefore refer to field is given by /(µ) = /0 + / 1µ. This result is important because the radiation
field is accurately described by an expression or the stated form in the diffusion
the single scalar PR as "the" radiation pressure.
limit which obtains at great depth in the atmosphere (cf. §2-5).
Ex£'rcise 1-12: Suppose an observer is at distance r from the center of a star of
Exercise I-JO: Show that, for any diagonal tensor A, in spherical coordinates radius ,.* which has a uniformly bright surrace (i.e., I is independent ofµ). Derive
(V · A), = (oA.,/or) + (2A., - A00 - A,i,,i,)/r; use this result to derive equation analytical expressions for J, H, and K in terms or 0* = sin - 1(r.Jr), and show
(l-43b). that in the limit (r/r *) .... oo. J = H = K.
18 The Radiation Field

VARIABLE EDDINGTON FACTORS

From the results derived above it follows that the ratio PR(r, v, t)/ER(r, v, t)
or K(r, v, t)/J(r, v, t) is a dimensionless number whose value is fixed by the
degree of isotropy of the radiation field, and typically ranges between½ and 1.
2
It will be shown later (§6-3) that this ratio can be used in certain numerical
methods to reduce the number of independent variables in the transfer
problem; further, it may be used to effect a closure of the system of moment
equations derived from the transfer equation. It is useful, therefore, to define
The Equation of Transfer
the variable Eddington factor
f(r, v, t) = K(r, v, t)/J(r, v, t) (1-47)

or, in abbreviated notation J. = K./J •.


Exercise 1-/3: (a) Consider an expansion of the form/(µ) = 10 + L,, I.µ"; show
that f = ¼if the sum includes only odd powers n. (b) Suppose that /(µ) = I 1
for (0 ,i;; µ ,i;; l) and /(µ) a I 2 for ( - I ,i;; µ ,i;; 0); show again that f = ¼- This
representation of 1 provides a rough description (the two-stream approximation)
of a stellar radiation field, for we may let 12/ 11 -> 0 at the surface and 1zl11 -> l
at depth. (c) Show that, for a slab of infinite extent in (x, y) and finite extent in
z, f may drop below¼,

___.
(X)
I

As radiation passes through the gas composing a stellar atmosphere, it


interacts with the material and is absorbed, emitted, and scattered repeatedly.
These phenomena determine how radiative transfer occurs in the atmosphere,
In this chapter, macroscopic quantities that define radiation-matter inter-
actions are introduced (§2-1), and the equation of transfer (which describes
the transport of radiation through the medium) is developed (§2-2). Using
this equation, we can compute the emergent spectrum from a star, and
calculate how the angle-frequency variation of the radiation field changes
with depth in the atmosphere. The time-dependent equation of transfer will
be derived in order to obtain moment equations (§2-3) that describe the
dynamical behavior of the radiation field, but the discussion will then be
restricted to static atmospheres in all subsequent work through Chapter 13.
In Chapters 14 and 15 radiative transfer and its dynamical effects in steady
(i.e., time-independent) flows will be considered. ·
20 The Equation of Transfer 2-/ The Interaction of Radiation with Matter 21

2-1 The Interaction of Radiation with Matter processes allow photons to move from one part of the atmosphere to some
other part without coupling to local conditions, and thus tend to delocalize
DISTINCTION BETWEEN SCATTERING
the control of the gas-radiation equilibration process, and to introduce
AND ABSORPTION-EMISSION PROCESSES
global properties of the atmosphere (e.g., the presence of boundaries) into
In the interaction of radiation with matter, energy may be removed from, the problem.
or delivered into, the radiation field by a wide variety of physical processes. To illustrate the ideas developed above, let us consider the following as
For the present, it is adequate to characterize these processes by macroscopic typical examples of scattering processes.
coefficients; as will be seen in Chapters 4 and 7, these coefficients are specified (a) The interaction of a photon with an atom in bound state a leading
by atomic cross-sections and occupation numbers of energy levels of the to the excitation of a higher bound state b (the photon's energy being con-
constituents of the stellar material. It is worthwhile, from the outset, to make verted to internal excitation energy of the atom), .followed by a direct return
a distinction between "true" absorption and emission on one hand and the to state a with the emission of a photon. In general the emitted photon will
process of scattering on the other for, as we shall see repeatedly in the propagate in a different direction from that of the incident photon. Further,
development of the theory, the physical nature of the interaction between both the lower and upper states a and b of an atom in a radiating gas will
the atmospheric material and radiation is quite different in these two cases. not be perfectly sharp, but will have finite energy widths arising, for example,
However, it is also important to realize that in spectral lines the dichotomy from _the finite lifetime of each state produced by radiative decay, or from
between these processes can be established uniquely only when we consider interactions of the atom with other particles of the plasma in which it is
a transition between two specified atomic states, with no coupling to any imbedded. Each of the bound states can, therefore, be considered to consist
other states allowed. As soon as sequences of transitions among several of a distribution of substates, with radiative transitions possible from any
interacting states are considered, fundamental ambiguities arise, and it is substate of one level to any substate of the other. Thus if the decay of the
no longer possible to describe a given line as an "absorption" or a "scattering" upper level occurs to a different substate of the lower level than that from
line in a rigorous way; nor would it be important or useful to do so. Neverthe- which the excitation occurred, or if there is a redistribution of the excited
less, it is fruitful to have at least an intuitive notion of the contrast between electron from the original excited substate to some other substate (because,
these two basic processes, obtained by consideration of some definite say, the atom suffers an elastic collision with another particle) then the
examples. emitted photon's energy may be slightly different from the incident photon's.
We may identify as scattering processes those in which a photon interacts Similarly, motions of the scattering centers with respect to the fixed labora- .......
with a scattering center (perhaps producing a change in the scatterer's internal tory frame can change the emitted photon's energy from the incident energy
excitation state) and emerges from the interaction in a new direction with if the projection of the scatterer's velocity along the direction of propagation
co
I
(in general) a slightly altered energy. The essential point is that in this process, is different for the two photons, for then a differential Doppler shift can occur.
the energy of the photon is not converted into kinetic energy of particles in (Example: imagine the incident photon to be moving in the same direction
the gas. In contrast, we shall identify absorption processes as those in which the as v of the scatterer and the emitted photon to move in the opposite direction.
photon is destroyed by conversion of its energy (wholly or partly) into the The emitted photon will be redshifted by an amount ~v = -2v0 v/c relative
thermal energy of the gas. In this process we say that the photon has been to the incident photon). Changes in photon direction and frequency during
thermalized. The crucial physical point to note is that the local rate of energy scattering are described by redistribution functions (see below). Note that in
emission in scattering processes depends mainly upon the radiation field this process no significant part of the photon energy is imparted to the
(which may have originated at some other remotely situated point in the material.
atmosphere) and has only a weak connection with the local values of the (b) Scattering of a photon by a free electron (Thomson or Compton
thermodynamic properties (e.g., temperature) of the gas. Absorption processes, scattering) or by an atom or molecule (Rayleigh scattering). Thomson
on the other hand, feed photon energy directly into the thermal kinetic energy scattering may be viewed as the result of the free charge oscillating in the
of the gas, and hence are more intimately coupled to local thermodynamic electromagnetic field of the radiation, Compton scattering as a collision of a
properties of the material. Conversely, the inverse of absorption, thermal photon with a free charged particle, and Rayleigh scattering as a resonance
emission, transfers energy from the thermal pool of the gas directly into the of a permitted "oscillation" of the bound system with the field. The remarks
radiation field. Thermal absorption and emission processes thus tend to made in (a) above concerning redistribution and lack of coupling of radian.t
produce local equilibrium between the radiation and material; but scattering energy to the thermal pool apply here as well.
22 The Equation of Transfer 2-I The Interaction of Radiation with Maller 23

Similarly, we may consider the following to be examples of thermal that the nonlocal behavior of the radiation field has been altered. On the
absorption processes (and their inverses as thermal emission). other hand, no contribution has been made to the thermal energy of the gas.
(a) A photon is absorbed by an atom in a bound state, and ionizes Alternatively, consider the same process, but now with a collisional de-
the bound electron, allowing it to escape with finite kinetic energy into the excitation c -> b followed by an emission b -> a. The original photon can
continuum. In this process of photoionization or bound-free absorption, the be said to have been destroyed (absorbed); but is the emitted photon
photon is destroyed and the excess of its energy over the electron's binding "thermally" emitted when most of the original energy was derived from
energy goes initially into the electron's kinetic energy, and ultimately into the radiation field? Many other more complex and subtle cases may be
the general thermal pool after the electron suffers elastic collisions that constructed, which, taken together show the limits of usefulness of the
establish a thermal velocity distribution for the particles. The inverse process, absorption-vs-scattering description.
of a free electron dropping to a bound state with the creation of a photon In fact, a truly consistent picture emerges only when we write down the
whose energy equals the sum of the electron's kinetic and binding energies, full equations of statistical equilibrium (cf. Chapter 5), which describe all
is called direct radiative recombination. These processes clearly transfer possible processes (both radiative and collisional) that couple an arbitrary
energy back and forth between the radiation field and the thermal pool of state i to some other state j, and solve these together with the equations
the material. (transfer equations) that describe how the radiation is absorbed, emitted,
(b) A photon is absorbed by a free electron moving in the field of an ion, and transported through the atmosphere. To do this is, in general, quite
resulting in an alteration of the electron's kinetic energy relative to the ion. difficult, and formulation of successful methods of solution of the prob-
The electron then, classically speaking, moves off on a different (hyperbolic) lem will occupy the bulk of this book. (The full import of these comments
orbit around the ion. This process is known as free-free absorption because will emerge only when the student has studied the material through
the electron is unbound both before and after absorbing the photon. The Chapter 12; nevertheless they should be borne in mind at all stages of
inverse process, leading to the emission of a photon, is referred to as subsequent development).
bremsstrahlung.
(c) A photon is absorbed by an atom, leading to a transition of an electron THE EXTINCTION COEFFICIENT
from one bound state to another; this process is called photoexcitation or
bound-bound absorption. The atom is then de-excited by an inelastic collision To describe the removal of energy from the radiation field by matter I\)
with another particle. EJ\ergy is put into the motion of the atom and the let us introduce a macroscopic coefficient x(r, v, t) called the extinction 0
collision partner arid thereby ends up as part of the thermal pool. The photon coefficient, or opacity, or sometimes (loosely) the total absorption coefficient. I
is said to have been destroyed by a collisional de-excitation. The inverse This coefficient is defined such that an element of material, of cross-section
process leads to the collisional creation of a photon at the expense of the dS and length ds, removes from a beam with specific intensity J(r, n, v, t),
thermal energy of the gas. incident normal to dS and propagating into a solid angle dw, an amount of
(d) Photoexcitation of an atom with subsequent collisional ionization of energy
the excited atom into the continuum. Photon energy again contributes to ~E = x(r, n, v, t)J(r, n, v, t) dS ds dw dv dt (2-1)
the thermal energy of particles. The inverse processes is called (three-body)
collisional recombination. within a frequency band dv in a time dt. The extinction coefficient is the
To illustrate the conceptual limitations of the kinds of arguments given product of an atomic absorption cross-section (cm 2 ) and the number density
above, let us now consider some ambiguous cases. Suppose an atom has of absorbers (cm - 3 ) summed over all states that can interact with photons
~hree bound levels a, b, and c, in order of increasing energy, and a photo- of frequency v. The dimensions of x are cm - 1, and (1/x) gives a measure of
excitation from a to c occurs. Then suppose that c decays radiatively to b, the distance over which a photon can propagate before it is removed from
and b then decays radiatively to a; this process is calledfiuorescence. Here the beam-i.e., a photon mean-free-path (cf. §2-2).
a single photon of energy hv.,, = E, - E., is degraded into two photons of The frequency variation of x may be extremely complicated, and may
energies hvab = Eb - E., and hvbc = E, - Eb. Was the original photon include thousands or millions of transitions (bound-bound, bound-free, and
scattered or absorbed? By our original definition it has not been "scattered" free-free). For static media in which there are no preferred directions imposed
and, moreover, the new photons may have vastly different properties (e.g., on an atomic scale (e.g., by a magnetic field), the opacity is isotropic. For
probability of escape through the boundary surface) from the original, so moving media, the opacity has an angular dependence introduced by the
24 The Equation of Transfer 2-1 The Interaction of Radiation with Matter 25

Doppler shift that radiation experiences in the fluid frame relative to its by the material, and thus by the nature of the radiation field and its response
original frequency in the stationary laboratory frame; this Doppler shift to the global properties of the atmosphere (e.g., boundaries, scattering,
obviously depends on the projection of the velocity vector onto the direction gradients, etc.). The remarks made in this paragraph carry over with equal
of the incident beam. In what follows we consider only static atmospheres. force in the macroscopic description of emission. Again, the full significance
As outlined earlier in this section, it is sometimes useful to distinguish of these comments will emerge only with considerable further development
between "absorption" and "scattering"; hence we introduce volume coeffi- (see particularly §§5-1 and 5-3).
cients K(r, v, t) and u(r, v, t) that describe [ via equation (2-1)] the rate at which
energy is removed from the beam by "true absorption" and "scattering," THE EMISSIUN COEFFICIENT
respectively. The total extinction is given by
To describe the emission of radiation from the stellar material, we intro-
X(r, V, t) = K(r, V, t) + u(r, V, t) (2-2) duce a macroscopic emission coefficient or emissivity 11(r, n, v, t) defined such
that the amount of energy released from an element of material of cross-
That is, both processes are assumed to occur independently and to add section dS and length ds, into a solid angle dw, within a frequency band dv,
linearly. In actual practice x is sufficient to describe energy removed from in direction n in a time interval dt, is
the beam; the distinction between " and u is useful mainly in defining the
emission coefficient. oE = 11(r, n, v, t) dS els dw clv cit (2-3)
In the calculation of x it is necessary to include a correction for stimulated
emission (see §§4-1 and 4-3). This is a quantum process in which radiation The dimensions of '7 are ergs cm- 3 sr- 1 hz- 1 sec- 1 • As was true for the
induces a downward transition from the upper state at a rate proportional opacity, thermal emissivity is isotropic for static media (without imposed
to the product of a cross-section, the upper-state population, and the specific preferred directions) but is angle-dependent for moving material owing to
intensity. Because the process is proportional to l(r, n, v, t) and effectively Doppler-shift effects. For radiation emitted in scattering processes, there is
cancels out some of the opacity, it is convenient to include it in the definition normally an explicit angle-dependence, even for static media. The emissivity
of x. Stimulated emission occurs only when the emitting system exists in a is calculated by summing products of upper-state populations and transition
cl~/inite upper state (whether bound or free). There is thus no stimulated probabilities over all relevant processes that can release a photon at frequency I\)
emission in Thomson scattering (free electrons) or Rayleigh scattering v. In writing the transfer equation we shall usually use the unembellished
(involves virtual states) but there is stimulated emission in spectrum lines, symbol '7 to denote the total emissivity; if electron scattering terms appear
even if they are described with a "scattering" coefficient. explicitly in the same equation, '7 will then denote all other emission. Sub-
H we know the value of x(r, v, t) [ or of 11:(r, v, t) and u(r, v, t)], we have a scripts "c" and "/" may on occasion be used to denote continua and lines,
complete macroscopic ·description of the rate at which material removes respectively. Again, we must realize that the simplicity of this description is
energy from a beam of radiation. But it is crucial to emphasize that the deceptive, for the reasons given above in the discussion of the extinction
"completeness" of the description is illusory. The reason this unpleasant coefficient.
comment must be made is that the simple picture we obtain from equation An important relation exists between ttie emission and absorption coeffi-
(2-1) glosses over the fact that the level populations, which "determine" the cients in the case of strict thermodynamic equilibrium (T.E.). Ifwe consider an
rate of energy removal from the radiation field by their contribution to X, adiabatic enclosure in steady-state equilibrium containing a homogeneous
are, in turn, determined by the radiation field via photoexcitations, photo- medium, we know that the material will have the same temperature T
i.onizations, radiative emission, radiative recombination, and related pro- throughout (otherwise it would be possible to devise processes to extract
cesses. Thus, in reality, the interaction of the radiation field with the absorbing work from the temperature gradient, in violation of the second law of
material is nonlinear. The problem just described still remains (though more thermodynamics). Further, we may expect the radiation field to be isotropic
subtly) even if it is assumed that we can calculate level populations by local and homogeneous throughout the enclosure (including at the surface of the
application of thermodynamic equilibrium relations that depend only on walls), for if it were not, beams traveling in opposite directions would not
the density and temperature. (This is the so-called local thermodynamic be exactly similar and a directional transport of energy would result, from
equilibrium or LTE approximation.) The reason is that the temperature is which work could be extracted, again in violation of the second law of
determined by overall balance between energy emitted and energy absorbed thermodynamics. Consideration of the energy absorbed and emitted in
26 The Equation of Transfer 2-J The lnteractiun of Radiatio~ Matter 27

angle-frequency ranges dw dv by an element of material, in time dt, now in which there are large gradients of material properties, and an open
shows that if a steady-state thermal equilibrium is to be achieved (no net boundary through which radiation freely escapes; the radiation field is
gain or loss of energy by the matter), the thermal emission must be given by therefore highly anisotropic and has a markedly non-Planckian character.
One might argue for using LTE, even though the radiation field is obtained
171(v) = ,c(v)J(n, v) (2-4) from solving a nonlocal transfer equation, if one could show that some
mechanism, specifically collisions among the particles, enforced LTE popu-
which is known as Kirchhoff's law. For an enclosure in strict T.E. at temper-
lations. As already mentioned above, and as we shall see in detail in §5-3
ature T the intensity of the radiation field is given by the Planck function
and in Chapters 7 and 11-14, this will not, in general, be the case. Rather,
Bv(T), so that
the radiation field determines the state of the material, and hence equation
17*(v) = ,c*(v)B.(T) (2-5)
(2-6) becomes invalid; in the end we must carry through a general analysis
which is the Kirchhoff-Planck relation. This result has been obtained without in which we specify the thermodynamic state of the gas and the distribution
reference to the composition of the material and is valid (in T.E.) for all function of the radiation field simultaneously by sofving the coupled equa-
materials. [See the excellent discussion of the interaction of matter and tions of transfer and statistical equilibrium.
radiation in T.E. in (160, 199-206) and in an article by Milne in (416, 93-96).] Let us now consider radiation scattered by the material. For simplicity of
Strictly speaking, the Kirchhoff-Planck law applies only in the case of a notation, we suppress explicit reference tot, though all of the quantities may
system in T.E. But if the material is subject only to small gradients over the be time-dependent. As described earlier, in scattering processes both the
mean free path a photon can travel before it is destroyed and thermalized direction and frequency of a photon may change. These changes are described
by a collisional process (as is true, e.g., in the interior of a star), then we could by a redistribution function
expect equation (2-5) to be valid to a high degree of approximation at local
R(v', n'; v, n) dv' dv(dw'/4rc)(dw/4rc)
values of the thermodynamic variables specifying the state of the material.
In such a case we write which gives the joint probability that a photon will be scattered from direction
n' in solid angle dw' and frequency range (v', v' + dv') into solid angle dw ·
17 1
(r, v, t) = ,c*(r, v, t)B.[T(r, t)] (2-6) in direction n and frequency range (v, v + dv). We shall derive redistribution
I\)
functions and discuss them in detail in Chapter 13, but it is helpful to mention
The hypothesi~ of local thermodynamic equilibrium (or LTE) just mentioned I\)
some of the general properties of these functions at the present time. We shall
makes the assumption that the occupation numbers of bound and free states I
normalize R such that
of the material, the opacity, the emissivity-indeed all of the thermodynamic
properties of the material-are the same as their T.E. values at the local
values of T and density, throughout the entire atmosphere, out to the outer-
most regions. Only the radiation field is allowed to depart from its T.E. value
of Bv[T(r)], and is obtained from a solution of the t~ansfer equation. Such The redistribution function contains within it both a normalized absorption
an approach is manifestly internally inconsistent, although LTE expressions profile <J,(v), and a normalized emission profile 1/,(v) for the scattering process.
remain valid for certain quantities even in the general case. For example, From the physical definition given above, it is evident that if we integrate
equation (2-6) is a valid expression for the continuum emission coefficient over all emitted frequencies and angles we must obtain the probability for
even in the presence of departures from LTE so long as the velocity distri- absorption from solid angle dw' and frequency range v'-i.e., <J,(v') dv' dw'/4rc.
bution of recombining (or, for free-free emission, colliding) particles is Thus
Maxwellian; the equation is not valid for line emission, and further, the LTE
formula for the opacity is not correct. The use of LTE is a computational
p f
<J,(v') = (4rc)- 1 dw 0"' dv R(v', n'; v, n) (2-8)

expedient that simplifies the calculation of models of stellar atmospheres, J


which by virtue of equation (2-7) is normalized such that q,(v') dv' = 1. If
and has been widely applied. (We shall employ it in places to provide a u 0 (r) denotes the total scattering coefficient, we may write u(r, v') = u0 (r)<J,(v').
prototype with which we can introduce basic mathematical techniques of The joint probability that an amount of energy u0 (r)J(r, n', v') will be removed.
solving transf~r problems, and will discuss models built assuming LTE in from the beam in solid angle dw' at frequency v', and scattered into dw at
§§7-2 through 7-4.) But it must be stressed that stellar atmospheres are regions frequency v, is u 0 (r)R(v', n'; v, n)J(r, n', v') dv' dv(dw'/4rc)(dw/4rc). Hence if we
28 The Equation of Transfer 2-1 The lnter~of Radiation with Malter 29

integrate over all incoming frequencies and angles, we find the total amount which shows that the distribution of emitted photons depends upon the
of energy emitted at frequency v into solid angle dw, namely frequency profile of the incoming radiation.
In the limiting case that the intensity is independent of frequency, we
,7s(r, n, v) dv(dw/4n) = u0 (r) dv(dw/4n) p(dw'/4n) f 0
00
dv' R(v', n'; v, n)J(r, n', v') obtain natural excitation, with

(2-9) if,*(v) = f0 00
R(v', v) dv' (2-14)

J
Exercise 2-1: Show that the total emission rate c/v f clw 17(r, n, v) equals the If we have R(v', v) = R(v, v'), as occurs in most cases of interest (cf. §13-3),
J
total energy removed from the beam 4na 0(r) tf,(1.')J(r, v') c/v'-i.e., that the then if,*(v) = ¢(v); that is, for natural excitation, the emission and absorption
scattering process is conservative. profiles are identical (a result that is not true in general!). Natural excitation
prevails, of course, in T.E., and it is usually thought of in that context. There
Equation (2-9) gives the full angle-frequency dependence of the emission are, however, other physical circumstances in which the result if,(v) = ¢(v) is
profile. It is usually difficult to treat radiative transfer problems in the degree recovered. In particular, suppose that there is a complete reshuffling of atoms
of generality implied here, and useful simplifications of the problem can be in their excited state in such a way that there is no correlation between the
made. For example, if we are primarily interested in redistribution in fre- frequencies of incoming and scattered photons; then both have frequencies
quency and not in angle, we could assume that J(r, n, v) is nearly isotropic, inclepenclently distributed over the absorption profile ¢(v). This situation is
and replace it in equation (2-9) with J(r, v). Then the emission into dv dw is referred to as complete redistribution, or complete noncoherence. A good
approximation to this case occurs, for instance, when atoms are so strongly
(2-10) perturbed by collisions during the scattering process that the excited electrons
are randomly redistributed over substates of the upper level. In this case,
where the angle-averaged redistribution function both the absorption and emission probabilities independently are propor-

R(v', v) = (4n)- 1 pR(v', n'; v, n) dw' = (41t)- 1 p R(v', n'; v, n) dw (2-11)


tional to the number of substates available at each frequency within the
line [i.e., to ¢(v) itself], and the joint absorption-emission probability
R(v', v) is the product of these two independent distributions-Le., R(v', v) =
gives the redistribution probability from (v', v' + dv') to (v, v + dv) and is
¢(v')¢(v). For complete redistribution, the emissivity is
I\)
normalized such that w
(2-12)
'ls(r, v) = u0 (r)¢(v) f
0
00
¢(v')J(r, v') dv' (2-15) I
from which we see clearly that the emission and absorption profiles are
The function R(v', v) is rendered independent ofangle as a result ofintegrating identical. Complete redistribution is also a good approximation within the
over either dw' or dw; this follows from the fact that (cf§l3-2) R(v', n'; v, n) Doppler core of a spectrum line, and actually provides an excellent firs\
depends only on the angle between n' and n. Equation (2-10) provides an approximation in line transfer problems. We shall, in fact, assume complete
extremely useful approximation in line transfer problems because the crucial redistribution in our discussion ofline formation until Chapter 13.
phenomenon there is the frequency diffusion of photons from the opaque Another class of problems arises when we focus attention on the angular
line core (where they are trapped) to the more transparent line wings (whence redistribution of the emitted radiation, but assume that the scattering is
they may escape from the atmosphere at depths where I is, in fact, very nearly essentially coherent (i.e., v' = v). This is the situation of interest, e.g., in
isotropic). In the angle-averaged approximation the emission profile scattering of light by large particles in a planetary (including earth's) at•.
mosphere. We can then write
= 11(r, v) / f0
00
1/,(v) 11(r, v) dv R(v', n'; v, n) = g(n', n)¢(v') <5(v - v') (2-16)
is given by
where <5 is the Dirac function and g is an angular phase function normalized

1/,(v) = f
00

0
R(v', v)J(r, v') dv' / f 0
00
¢(v')J(r, v') dv' (2-13)
such that
(41t)- 1 f g(n', n) dw' = 1 (2-17~
30 The Equation of Transfer 31

Two important phase functions are those for isotropic scattering

g(n', n) =1 (2-18)
and the dipole phase function (which applies for Thomson and Rayleigh
scattering)
3
g(n', n) = 4(1 + cos 2 <I>) (2-19)
FIGURE 2-J

where cos <I> = n' · n. The phase functions for scattering by large particles Element of absorbing and emitting material considered in derivation
of transfer equation.
(i.e., whose size is comparable to a wavelength of light) are often extremely
complicated and show large, rapid variations as a function of angle (312;
359, Chap. 4). direction (Doppler shift and aberration) resulting from the transformation
For coherent scattering, equation (2-9) reduces to between the laboratory frame and the fluid frame. These effects depend upon
n · v; hence both x and ti will have an explicit angle-dependence in this case.
t1 5(r, n, v) = a(r, v) f l(r, n', v)g(n', n) !:' (2-20) Now calculate the energy in a frequency interval dv, passing in a time dt
through a volume element of length ds and cross-section dS oriented normal
to a ray traveling in direction n into solid angle dw (see Figure 2-1). The
In a spectrum line, coherent scattering would occur only if the lower state of difference between the amount of energy that emerges (at position r + M
the line were sharp, the upper state were not perturbed during emission, and at time t + Llt) and that incident (at r and t) must equal the amount created
the scattering atoms were at rest in the observer's frame. This is not the case, by emission from the material in the volume minus the amount absorbed.
however, and line scattering is much more accurately described by com~lete That is,
redistribution (except in the far line wings where ct, varies slowly over the
range corresponding to a Doppler shift). On the other hand, for continuum [l(r + M, n, v, t + Llt) - l(r, n, v, t)] dS dw dv dt
scattering (e.g., by electrons) the frequency distribution ofradiation is smooth
and essentially constant over the typical frequency shifts occurring in the = [t1(r, n, v, t) - x(r, n, v, t)l(r, n, v, t)] ds dS dw dv dt (2-22)
I\)
scattering process. For this reason continuum scattering processes are Lets denote the path-length along the ray; then !l.t = !l.s/c, and ~
customarily treated as ifthey were coherent (though this may be inadequate I
near a spectrum line). Moreover, as the angular redistribution effects from l(r + M, n, v, t + Llt) = l(r, n, v, t) + [c- 1(81/ot) + (ol/os)] ds (2-23)
a dipole phase function are usually very small in a stellar atmosphere, it is
customary to assume that continuum scattering is also isotropic, and to write Substituting equation (2-23) into equation (2-22) we have the transfer equation

t1 5(r, v) = a(r, v)J(r, v) (2-21) [c- 1(0/ot) + (ofos)]l(r, n, v, t) = t1(r, n, v; t) - x(r, n, v, t)l(r, n, v, t) (2-24)
The derivative along the ray may be re-expressed in terms of an orthogonal
2-2 The Transfer Equation coordinate system:

DERIVATION

Let us now consider the problem of radiative .transport. Choose an in-


ertial coordinate system and examine the flow of energy through a fixed
volume element in a definite time interval. Let us assume that the radiation
where (nx, ny, n,) are the components of the unit vector n. We may thus re-
field is, in genera~ time-dependent. If we suppose the material to be at rest, write equation (2-24) as
then both xand ti will be isotropic (unless we consider anisotropic scattering).
In moving material one must account for changes in photon frequency and [c- 1(8/ot) + (n · V)]l(r, n, v, t) = t1(r, n, v, t) - x(r, n, v, t)l(r, n, v, t) (2-26)
32 The Equation of Transfer 2-2 The Transfer Equation 33

For a one-dimensional planar atmosphere n, = (dz/ds) = cos 0 = µ; number density within a phase-space element must equal the net number
further, the derivatives (o/ox) and (o/oy) are identically zero, and we obtain introduced into the element by collisions; i.e.,

[c- 1(8/ot) + µ(o/oz)]l(z, n, v, t) = 11(z, n, v, t) - x(z, n, v, t)l(z, n, v, t)


(2-27)
of + (ax) (of) +
ot ot ax
(ay)
(of) +
at ay
(of)
ot oz
(az)
or, for the time-independent case,
+ F (of)+ F (of)+ F (of)- (DI)
X OPx y Opy • op, - Dt coll
(2-29)
µ[ol(z, n, v)/oz] = 11(z, n, v) - x(z, n, v)/(z, n, v) (2-28)
or, in more compact notation
Equation (2-28) is the "standard" transfer equation for plane-parallel model-
atmospheres calculations; the coordinate z increases upward in the atmo- (of/ot) + (v · V)/ + (F · Vp)/ = (Df/Dtlcoll (2-30)
sphere (i.e., toward an external observer). For static media, the specification
(z, n, v) in I'/ and x may be reduced to (z, v) only. Note that if I'/ and x are given, For a "gas" consisting of photons (with rest-mass zero), in the absence of
equation (2-28) is an ordinary differential equation, which may be solved for general relativistic effects, F = 0, and photon propagation in an inertial
all relevant choices ofµ and v. When I'/ includes scattering terms, the transfer frame occurs in straight lines with v = en, while the frequency remains
equation becomes an integro-differential equation containing angle and constant. The distribution function fR can be written in terms of the specific
frequency integrals of I. · intensity by means of equation (1-2). The analogues of"collisions" are photon
interactions with the material, and the net number of photons introduced
THE TRANSFER EQUATION AS A BOLTZMANN EQUATION into the volume will be the energy emitted minus the energy absorbed,
divided by the energy per photon. Thus for photons equation (2-30) becomes
The basic equation describing particle transport in kinetic theory is the
Boltzmann equation; we shall now show that the transfer equation is just the' (chv)- 1 [(ol/ot) + c(n · V)/] = (I'/ - xl)/(hv) (2-31)
Boltzmann equation for photons. Suppose we have a particle distribution
which is identical to the transfer equation (2-26). In effect the transfer equation
function f(r, p, t) that gives the number density of particles in the phitse
is a Boltzmann equation for a fluid that is subject to no external forces but I\)
volume element (r, r + dr), (p, p + dp). We follow the evolution off within
which suffers strong collisional effects. As will be seen in §2-3, the moments 01
u particular phase-space element for a time interval dt, in which r -> r + v dt
of the transfer equation yield dynamical equations for the radiation field, just I
and p -> p + F dt, where F denotes externally imposed forces acting on the
as moments of the Boltzmann equation for a gas lead to equations of hydro-
particles. The phase-space element evolves from
dynamics.

SPHERICAL GEOMETRY
where J is the Jacobian of the transformation.
In a spherically symmetric medium, the specific intensity will be inde-
Exercise 2-2: Show that to first order in cit the Jacobian or the transformation pendent of the coordinates 0 and <I> of the triplet (r, 0, <I>) which specifies
or a phase volume element is J = I. a position in the atmosphere, and of the azimuthal angle</> of the pair (0, </>)
which specifies the direction of the beam relative to the local outward normal
In ,view of the result of Exercise 2-2 we see that the phase-space element is r. Thus /(r, n, v, t) reduces to /(r, 0, v, t). In writing the transfer equation,
clejormecl, but its phase volume is unchanged. If all external forces F are starting from the general form of equation (2-24), we must now account for
continuous, then the deformation of the phase-space element is continuous, the variation of 0 along a displacement, and employ the general form ds =
and all particles originally within the volume remain there; as the volume dr r + r dO 9. As is clear from the geometry of the situation (see Figure 2-2),
itself is unchanged, the particle density is unchanged. But if, in addition, cir = cos 0 ds, while r c/0 = - sin 0 ds (note that dO ~ 0 for any ds), so that
collisions occur, individual particles may be reshuffled from one element of (o/os) _. cos O(o/or) - ,.- • sin 0(o/o0) = µ(o/or) + r- 1(1 - µ 2 )(8/oµ)
phase space to another "discontinuously"; i.e., their neighbors may be totally
unaffected during the same time interval. Therefore, the change in the particle (2-32)
34 2-2 The Transfer Equation 35

optical depth scale r(z, v) wpich gives the integrated absorptivity of the
material along the line of sight as

f.
Zma,
r(z, v) = • x(z', v) dz' (2-34)

The negative sign is introduced so that the optical depth increases inward
FIGURE 2-2
Geometric relations among into the atmosphere from zero at the surface (where z = Zmaxl, and thus
variables used in derivation of provides a measure of how deeply an outside observer can see into the
transfer equation in a spherically material [cf. equations (2-47) and (2-52)]. Recalling that x- 1 is the photon
symmetric medium. mean-free-path, we recognize that r(z, v) is just the number of photon mean-
free-paths at frequency v along the line of sight from Zmax to z. In addition,
we define the source function to be the ratio of the total emissivity to total
opacity,
S(z, v) = '1(z, v)/x(z, v) (2-35)
To simplify the notation we shall for the present suppress explicit reference
where, as usual, µ = cos 0. Hence the transfer equation for a spherically to z andµ, and denote frequency dependence with a subscript v. The equation
symmetric atmosphere is of transfer may then be written in its standard form

[ c- 1(o/ot) + µ(o/or) + r- 1(1 - µ 2 )(o/oµ)]l(r, µ, v, t) (2-36)


= '1(r, v, t) - x(r, v, t)l(r, µ, v, t) . (2-33) From the discussion of §2-1 we can write prototype expressions for the
source function which we use to study methods of solving equation (2-36).
which simplifies in an obvious way in the time-independent case. Note that
Suppose first we have strict LTE. Then from equation (2-6) we have
now, even with '1 and x specified, equation (2-33) is a partial differential
equation in rand µ. However, this added complexity can be avoided by using (2-37) f\)
the (straight-line) characteristic paths that reduce the spatial operator to a CJ)
single derivative with respect to pathlength (see §7-6). In fact, equation (2-33) If we have a contribution from thermal absorption and emission plus a I
is not structurally different from equation (2-28) and can be solved almost contribution from a coherent, isotropic, continuum scattering term (say from
as easily. Thomson scattering by free electrons or from Rayleigh· scattering) then we
could write
Exercise 2-3: (a) Consider an atmosphere which is axially symmetric but not (2-38)
spherically symmetric (e.g., a rotationally flattened star). Show that now /(r, n) =
/(r, 0, 0, cf,). (b) For the general case where /(r, n) = /{r, 0, <I>, O, cf,) (e.g., a and (2-39)
rotationally flattened star illuminated by a companion), show that the transfer
equation in spherical coordinates, accounting for all the spatial derivatives, is For a spectrum line with an overlapping background continuum we have

c· 1(iJ//cl) + µ(aJJar) + (y/r)(oJ/a0) + (u/r sin 0)(a//a<I>) Xv = X, + x,(v) = X, + X,</>. (2-40)


+ ,- 1(1 - µ2)(o//aµ) - (u cot 0/r)(o//act,) = '1 - xI where X, and x, denote the continuum and line opacities, respectively. If we
where 1• = cos cf, sin Oand u = sin cf, sin 0. assume that a fraction e of the line emission occurs from thermal processes
and the remainder is given by angle-averaged complete redistribution
OPTICAL DEPTH AND THE SOURCE FUNCTION [equation (2-15)], we can write

For the remainder of§2-2 let us confine attention to the time-independent


planar transfer equation (2-28). Writing dr(z, v) = - x(z, v) dz, we define an
l'/v = x,B. + x,cp. [ eB. + (1 - e) f cp,J. dv] (2-41)
36 The Equation of Transfer 2-2 The Transfer Equation 37

and range O ,;;; r ,;;; r•• (r • ,;;; r,) with intensity / 0 (µ', v) where µ' is the angle cosine
at the opaque surfe1ce; this case simulates an envelope around a star. (d) Show
e::) B. + [<lr ~ e;~•] f <J>.J. dv e.B. + (1 - e.) f <J,,J. dv
0

s. = (rr: = in case (a) ~bove that the flux is identically zero at all points r ,;;; r,.

(2-42) In the semi-infinite case (planar or spherical), the radiation field incident
upon the upper boundary must be specified by an equation of the form of
where r = x,/x,. (We may ignore the frequency variation of the continuum (2-43); for stellar atmospheres work, it is customary to assume r = 0
over a line-width.) In equations (2-39) and (2-42) we have examples of the (clearly this would not be done, e.g., in a binary system). For a lower boundary
explicit appearance of integrals of the radiation field over angle and fre- condition one may replace equation (2-44) by a boun(ledness condition in
quency, which shows the integro-differential nature of the transfer equation. analytical work where the limit t. -+ oo is taken. Specifically we impose the
It must be stressed that the source functions in equations (2-37), (2-39), and requirement that
(2-42) are only illustrative; they are based on essentially heuristic arguments, lim /(t., µ, v)e-••1µ = 0 (2-45)
r-oo
and a physically rigorous formulation can be provided only after the equa-
tions of statistical equilibrium (Chapter 5) are developed. The reasons for this particular choice will become clear in the discussion
?elow. Alternatively, at great depth in the atmosphere we may write J(t., µ, v)
BOUNDARY CONDITIONS m terms of the local value of s. and its gradient, or may specify the flux; these
conditions follow naturally from physical considerations in the diffusion
Solution of the transfer equation requires the specification of boundary limit where the photon mean free path is much smaller than its optical depth
co11clitio11s. Two problems of fundamental astrophysical importance are from the surface (see §2-5).
(a) the.finite slab (in planar geometry) or shell (spherical geometry), and (b) a
medium (e.g., a stellar atmosphere) that has an open boundary on one side SIMPLE EXAMPLES
but is so optically thick that it can be imagined to extend to infinity on the
other side-the semi-infinite atmosphere. Before writing the formal solution of the equation of transfer, it is
For the finite slab of total geometrical thickness Zand total optical depth instructive to consider a few simple examples in planar geometry. I\)
T,, (defined to be zero on the side nearest the observer), a unique solution is (a) Suppose no material is present. Then x. = 11. = 0, and equation (2-28)
obtained if the intensity incident on both faces of the slab is specified. Writing reduces to (oI ./az) = 0 or I. = constant. This result is consistent with the -..J
0 for the angle between a ray and the normal directed toward the observer, proof in §1-1 of the invariance of the specific intensity when no sources or I
andµ = cos 0, we must specify the two functions 1+ and r such that sinks are present.
(b) Suppose that the material emits at frequency v, but cannot absorb.
l(t. = 0, µ, v) = r(µ, v), ( -1 ~ JI ~ 0) (2-43)
Then equation (2-28) is µ(oIJaz) = 17,., and for a finite slab the emergent
at the upper boundary, and intensity is given by

(2-44) l(Z, Jl, v) = µ


- Jorz 17(z, v) dz + J+(0, JI, v)
1
(2-46)

at the lower boundary. For a shell of outer radius R and inner radius r,, The physical situation described above occurs in the formation of optically
,equation (2-43) still applies at r = R. forbidden lines in nebulae. Atoms may be excited to metastable levels by
collisions; because nebular densities are so low, the chances of a second
Exercise 2-4: (a) At the inner boundary of a spherical shell, r = r,, show that collision leading to de-excitation are very small, so the atoms can remain un-
the lower boundary condition is given by /(r,, +JI, 1') = /(r,, - µ, v) if the central perturbed in these levels for long periods of time, and large numbers of atoms
volume r ,;;; r, is void; this is Milne's "planetary nebula" boundary condition. (b) If may accumulate in these states. Eventually some of the atoms decay via
the volume contains a poinr source (a star) of intensity I0 , show that the result in "forbidden" transitions, which have very small but nonzero transition proba-
(a) must be augmented by /(r,, + µ, v) = /0 {v) t5(µ - I). (c) Extened results (a) bilities, and emit photons. Because the line is forbidden, the probability of
and (b) to the case of the volume being partially filled by an opaque source on the reabsorption is negligible, and the photon escapes. Thus photons are created
38 The Equation of Transfer 2-2 The Transfer Equation 39

at the expense of the energy in the thermal pool, and none are destroyed by The physical significance of equation (2-52) is that the emergent intensity is a
absorption. weighted average of the source function along the line of sight; the weighting
(c) Suppose that radiation is absorbed but not emitted by the material. function is the fraction of the energy emitted at each depth that penetrates
Then 11(aI./8z) = -x,I,, and defining dtv = -x. dz, the emergent intensity to the surface along a ray whose optical slant-length is (t/µ). Mathematically
from a finite slab of total optical thickness T v is speaking, (equatio11 (2-52) shows that the specific intensity is the Laplace
transform ~f the source function, a property that may be used to solve for S
J(tv = 0, µ, v) = J+(T., µ, v) exp(-Tvfµ) (2-47) in certain classes of problems.
Considerable insight can be gained by supposing the source function is a
Equation (2-47) applies, for example, to radiation passing through a filter in linear function of depth: S.(r.) = Sov + S 1.t.; then equation (2-52) yields
which photons are absorbed and degraded into photons of some very differ-
ent frequency (e.g., extreme far-infrared heat radiation) before being re- (2-53)
emitted, or are destroyed and converted into kinetic energy of the particles
in the absorbing medium. which is known as the Eddington-Barbier relation. This result states that the
emergent intensity is characteristic of the value of the source function at about
optical depth unity along the line of sight (recall that t, represents the normal
FORMAL SOLUTION
optical depth, so that slant optical depth unity for a ray penetrating at angle
Let us now obtain a formal solution of the equation of transfer; we con- cos-• µ occurs at r, = µ). The Eddington-Barbier relation has been widely
fine attention exclusively to planar geometry. Regarding Sas given, equation applied in empirical analyses of the solar and stellar spectra, and provides a
(2-36) is a linear first-order differential equation with constant coefficients, basic conceptual framework for many interpretive methods. In the case of
and must therefore have an integrating factor. The integrating factor is easily the sun, when we observe, at a fixed frequency, the variation in intensity from
shown to be exp(-t,/µ), so that disk center(µ = 1) to the limb(µ .... 0) we can (in principle) infer information
about S,(tvl for O ~ tv ~ 1. For stars, we cannot observe the center-to-limb
iJ[J,exp(-t,/µ)]/i!t, = -µ- 1S,exp(-t,/µ) (2-48) variation, but it is clear that ifwe observe at different frequencies (e.g., within
a spectral line), we encounter unit optical depth in higher layers for fre-
Integration of equation (2-48) yields f\)
quencies with high opacity (e.g., line-center) and iri deeper layers for frequen-
co
/(ti,µ, v) = l(t 2 , µ, v)e-<• 2 - • 1>111 + µ-• J.' S,(t)e-
1
2 111
0 -• 1> dt (2-49)
cies with low opacity (e.g., line-wings). Ifwe then know something about the
frequency-variation of s., we can infer information about its depth-variation; I
for example, in LTE, s. = B. and the frequency-variation over a narrow line
If S,. is given, equation (2-49) provides a complete solution of the planar can be ignored, so that one can, in principle, infer the ruri of the temperature
transfer problem. with depth. Although it is extremely useful conceptually, the Eddington-
We may apply equation (2-49) at an arbitrary interior point in a semi-infinite Barbier relation should not be applied indiscriminately and used literally to
atmosphere. Forµ ;.i, 0 (outgoing radiation), set t 1 = t, and t 2 = oo, and argue that J(0, µ, v) is identical to S.{t. = µ), because (a) there are always
impose the boundedness condition of equation (2-45); the result is significant contributions to I ,(0) from other depths (i.e., there is an intrinsic
l(t,, µ, v) = J.•• S,(t)e-<r-,.,,,. dt/µ,
00
(0 ~ µ ~ 1)· (2-50)
"fuzziness" in the problem), and (b) the assumptions from which it follows
may not be valid. A detailed critique of the limitations of the Eddington-
Barbier relation can be found in (18, 121-130) and (20, 20-30).
For incoming radiation, (-1 ~ µ ~ 0), set t 2 =0 and apply the upper
boundary condition = 0; we then obtain
r Exercise 2-5: Suppose the source function is to be represented by a power-series
expansion about the point r.; i.e., S(r) ~ S(i.) + S'(r.)(r - r.) + ½S"(r.)(r - r.)2 •
J(t., µ, v) = f;• S,(t)e-<•.-•>t<- 11>dt/(-µ), (-1 ~ µ ~ 0) (2-51) Calculate the emergent intensity and show that the choice r. = µ is "optimum"
in the sense that it eliminates the contribution of S' and minimizes the contribution
One of the most important applications of equation (2-50) is the expression of S" to 1(0, µ).
for the emergent intensity seen by an external observer (t = 0),
Another instructive example to consider is a finite slab of optical thickness
1(0, µ, v) = f0
00
S,(t)e-'111 dt/µ (2-52) T, within which Sis constant, and upon which there is no incident radiation.
40 The Equation of Transfer 2-2 The Transfer Equation 41

The normally emergent intensity is 1(0, 1) = S(l - e- 7 ). For T » 1, I = S. been abbreviated to an operator notation:
This result is sensible physically, for the radiation that emerges consists of
those photons emitted over a mean-free-path from the surface; the emission (2-58)
rate is 17 and the mean-free-path is x- 1, hence I = 17/x = S. For T « I,
e-r ~ 1 - T, so I ~ ST. Again this result is physically reasonable, for in the
optically thin case we can see through the entire volume; hence the energy Exercise 2-6: (a) Show that equations (2-50) and (2-51) are equivalent to
emitted (per unit area) must be the emissivity per unit volume 17 times the
total path-length Z, or I = 17Z = ('7/x)(xZ) = ST. Note that this result is
l(r, n) = s:-· 11(r') exp[ - t(r, r1] tilr' - rl

consistent with equation (2-46). where r'(s) == r - sn, t(r', r) == J~'-•I x[r'(s)] ds, and sm.. _is the distance along the
ray to any boundary surface in the direction ( - n); sm.. = oo for outward-directed
THE SCHWARZSCHILD-MILNE EQUATIONS rays in a semi-infinite medium. (b) Substitute the above result into the definition
of J(r) [equation (1-3)) to derive Peierls's equation:
By integration of the formal solution for the specific intensity over angle,
concise expressions for the moments of the radiation field may be derived. J(r) = (41t)- 1 fv {11(r') exp[ -t(r', r)]/lr' - rl 2 } d3 r'
Thus by substitution of equations (2-50) and (2-51) into equation (1-4) we where V denotes the entire volume containing material.
have for the mean intensity
By an analysis similar to that used to obtain equation (2-57) we can derive
J,(,.) = ~ f~ 1 1,(,., µ) dµ = ~ [f01 dµ f.~ S,(t)e-<•-••11 ~ µ
expressions for F, and K. first obtained by Milne (416, 77):

+ f o dµ Jof•• S•(t)e-1t.-,11<-µ> ~ ] (2-54)


:.1 (-µ)
- 2 f;• S,(t.)Ei(.. - t,) dt, (2-59)
Equation (2-54) is reduced to a more useful form by interchanging the order of
integration, and making the substitution w = ± l/µ in the first and seco!:ld ~ fo
00
and K,(,.) = s.(t.)E3lt. - •• 1 dt. (2-60)
integrals respectively. Then I\)

1 [s.' foo co
J.(,.) = 2 ••
00

dt S,(t) Ji dw
e-w(t-,,)
w + f/ dt S,(t) Ji
f 00

dw
e-w(,,-r)]
w
We also define the corresponding operators
I
<l>,[f(t)] = 2 f, 00
.f(t)E2 (t - t) dt
(2-55)
The integrals over ware of a standard form and are called the first exponen- - 2 J; f(t)Ei(. - t) dt (2-61)
tial integral. In general, for integer-values of n, the nth exponential integral is
defined as and X ,[f(t)] =2 J0
00
f(t)E 3 lt - •I dt (2-62)
(2-56)
Exercise 2-7: Derive equations (2-59) and (2-60).
In terms of E 1(x), equation (2-55) can be rewritten as
The mathematical properties of the exponential integrals are discussed in

J,(,,) =~ f
0
00
S,(t,)Edt, - ,,ldt, (2-57)
detail in (4, 228-231) and (161, Appendix I), and the properties of the A, <l>,
and X operators are discussed in (361, Chap. 2). A few of the most important
results are mentioned in the following exercise.
Equation (2-57) was first derived by K. Schwarzschild and is named in his
honor; Schwarzschild's paper (416, 35) is one of the foundation stones of Exercise 2-8: (a) Differentiate equation (2-56) to prove E~(x) = -E.- 1(x).
radiative transfer theory, and merits careful reading. Because the integral (b)lntegrateequation(2-56)bypartstoshowthatE.(x) = [e-x - xE.- 1(x)]/(n - I)
appearing in equation (2-57) occurs so often in radiative transfer theory it has forn > I. (c) Show that the asymptoticbehavior(x » I) of E.(x) is E.(x) ~ e-x;x.
42 The Equation of Transfer 2-3 Moments of the Transfer Equation 43

Some interesting physical insight can be gained by considering a linear Exercise 2-10: Assume a source-function or the form or equation (2-39). Show
source function S(t) = a + bt; with the results of Exercise 2-8, it is easy to that J .(t.) satisfies the integral equation
show that

1
+ bt) = (a + bt) + (2-63) where P. = u./(K. + u.)
A,(a
2 [bE 3(t) - aE 2(r)]

Comwlications of this kind are also introduced by other physical con-


4
<l>,(a + bt) = b + 2[aE 3(t) - bEh)] (2-64) straints-6\1 the solution. For example, we shall see in §2-4 that the requirement
3 of energy balance couples together the radiation field and the temperature
structure of the atmosphere. Thus even if strict LTE is assumed, and Sv is
4
+ bt) =
3 (a + bt) + 2[bE 5(t)
and X,(a - aEir)] (2-65) set equal to B, and presumed to be independent op., we are not in general
able to specify the run of T(r.) and hence of B,[T(r.)] until the radiation
field is already known. Here again the source function required to compute
Exercise 2-9: Verify equations (2-63) through (2-65). the radiation field depends upon the field itself. When the more general
non- LTE situation is considered, these problems become more subtle and
From the expressions written above (which will be applied in Chapter 3 complex, for now material properties (opacity, emissivity, etc.) depend di-
and later work), we can note some important results. First, because the rectly upon the radiation field, for it determines occupation-numbers in
exponential integrals decay asymptotically as e-"/x, it is evident from equa- atomic states. The transfer problem, in short, is fundamentally nonlinear
tion (2-63) that J(r) = A,(S) strongly approaches the local value of S(r) for in real stellar atmospheres. The development of methods to treat complicated
t » I; that is, the A-operator tends to reproduce the source function at interactions of the types just described will occupy a major part of this book.
depth. In contrast, at the surface, E 2 (0) = 1 and E 3 (0) = ½, and we obtain
A0 (S) = ½a + ¼b; in particular, if b = 0, then J(0) = ta = ts.
Physically
this reflects the fact that J at the surface is the average over a hemisphere
containing no radiation (none incident from empty space) and a hemisphere
in which I = S (no gradient in S). When a gradient is present, J(0) may lie 2-3 Moments of the Transfer Equation w
above or below S(O) depending on the sign and magnitude of the gradient.
· The angular moments of the transfer equation yield results of deep physical 0
In a general way, J will depart most strongly from Sat the surface. Second,
significance and great mathematical utility. The basic time-dependent trans- I
we see from equation (2-64) that for t » 1, H ➔ ½b, That is, the flux at
depth depends only on the gradient of the source function (see also §2-5). fer equation (2-26) may be rewritten as
The flux at the surface is H(0) = ¼a + lb; clearly the surface flux will be c- 1 [8/(r, n, v, t)/8t] .+ L 111[aJ(r, n, v, t)!ax1]
larger, the faster the source function rises inward. Note also that the effect J
of a gradient, relative to the case b = 0, is larger for H(0) than for J(0). = ,i(r, n, v, t) - x(r, n, v, t)J(r, n, v, t) (2-66)
(Why?)
Finally, it must be stressed that the solution of the transfer equation as where x1 denotes the jth cartesian coordinate. To obtain the zero-order
given by equations (2-50) and (2-51) or (2-57) and (2-59) is only formal, and moment equation, we integrate equation (2-66) over all solid angles dw, and
• its apparent simplicity is illusory. For example, suppose the source function use the definitions in equations (1-4) and (1-19) to write
contains a scattering term as in equation (2-39) or (2-42). Then it is clear
that the source function, which is required to compute the radiation field, (4n/c)[8J(r, v, t)/8t] + V · ~(r, v, t)
depends upon the radiation field, and hence upon the solution of the transfer
equation. In such cases we cannot calculate / or J by simple quadrature,
= f [,i(r, n, v, t) - x(r, n, v, t)/(r, n, v, t}] dw (2-67)
but must find the solution of an integral equation or of the corresponding where we have allowed for the possibility that x and '1 may depend upon
differential equation. angle (as they will for moving media). If we integrate equation (2-67) over
44 The Equation of Transfer 2-3 Moments of the Transfer Equation 45

all frequencies, and recall equation (1-8), we find or, in vector notation,
[oER(r, t)/ot] + V. §(r, t) c- 2 [o1F(r, v, t)/ot] + V · P(r, v, t)
f
= 0 dv ~ dw[17(r, n, v, t)
cx, - x(r, n, v, t)/(r, n, v, t)] (2-68) = c- 1 ~ [17(r, n, v, t) - x(r, n, v, t)/(r, n, v, t)]n dw (2-74)
This is an energy equation for the radiation field, and it closely resembles
Recalling that the momentum density in the radiation field is GR = c- 2 §
the standard energy equation for a moving fluid (cf. §1S-1). The terms in
(see §1-3), we see that equation (2-74) is analogous to the hydrodynamical
the equation have a straightforward physical interpretation: the time rate
equations of motion, and may be viewed as a dynamical equation for the
of change of the radiation energy density equals (a) the total energy emitted
momentum in the radiation field at frequency v. When we integrate over all
into the field by the material, (b) minus the total absorbed from the field
frequencies, we obtain a total momentum equation for the radiation, which
by the material, (c) minus the net flow of energy through the surface con-
we shall use in the equations of radiation hydrodynamics (cf. Chapter 15):
taining a volume element (the divergence of the flux). If the medium is static,
x and 17 are isotropic and the integrals on the righthand side simplify to yield c- 2 [o1F(r, t)/ot] + V · P(r, t)
[oER(r, t)/ot] + V · §(r, t) = 4,r Socx, (17(r, v, t) - x(r, v, t)J(r, v, t)] dv (2-69) = c- 1 f dv pdw[17(r, n, v, t) -
0
cx, x(r, n, v, t)/(r, n, v, t)]n (2-75)

For a time-independent radiation field in a one-dimensional planar static Equation (2-75) states that the time rate of change of the total momentum
medium, equation (2-67) reduces to the "standard" result density in the radiation field is equal to the negative of the volume force
exerted by radiation stresses (cf. §1-4) plus a term that must represent the net
[oH(z, v)/oz] = 17(z, v) - x(z, v)J(z, v) (2-70)
momentum gain (or loss) from interacting with the material [see further
or, abbreviating the notation and using equations (2-34) and (2-35), discussion following equation (2-76)]. As in equation (2-68), equation (2-75)
allows for the possibility of material motions. If the medium is at rest, then
caH./o-r,) = 1. - s. (2-71) the integral over '7 vanishes (which merely states that the net momentum loss
from the material by isotropic emission is zero, as is physically obvious), w
For spherical geometry (in a time-independent static atmosphere), by use while the second term reduces to an integral over the flux:
of the appropriate expression for the divergence, we find
(2-72) c- 2 [o.9F(r, t)/ot] + V · P(r, t) = -c- 1 f
0
00
x(r, v, t)1F(r, v, t) dv (2-76)

Equations (2-70) through (2-72) will be employed repeatedly to obtain The physical significance of the integral on the righthand side of equation
solutions of the transfer equation, and equation (2-69) will be used to develop (2-76) can readily be seen by the following argument. Consider a beam of
the equations of radiation hydrodynamics (cf. §1S-3). specific intensity /, entering an element of absorbing material of surface area
dS, at an angle 0 relative to the normal. The energy absorbed by material of
Exercise 2-1 I: Derive equation (2-72) directly from equation (2-33). opacity x from a solid angle dw and frequency band dv in time dt is

· The first-order-moment equation with respect to the ith coordinate axis is


dE = xi dS cos 0 ds dw dv dt
obtained by multiplying equation (2-66) by n1 and integrating over (dw/c),
where ds = (dz/cos 0) is the slant-length of the ray through the element of
which yields
thickness dz along the normal. The component of momentum deposited in
c- 2 [o§'i{r, v, t)/ot] + L, [oPu(r, v, t)/ox1] the material along the direction of the normal is c- 1 aE cos 0; hence the
J momentum deposition per unit volume per unit time is
= c- 1 p[17(r, n, v, t) - x(r, n, v, t)l(r, n, v, t)]n 1 dw (2-73) (c- 1 cos 0 dE)/(dz dS dt) = c- 'xi cos 0 dw dv
46 The Equation of Transfer 2-4 The Condition of Radiative Equilibrium 47

lfwe sum over all angles and frequencies, we obtain precisely the integral in the variable Eddington factor .f,-i.e., K, = f.J •. The factor f. is obtained by
equation (2-76). We have thus shown that the integral is the radiation force, iteration and allows us to effect an approximate closure of the exact system
per unit volume, on the material; this interpretation is clearly compatible with (if the iteration conve,·ges, the closure is also exact). Alternatively, it will be
the overall physical meaning of the equation. shown in §6-3 that the transfer equation can be rewritten in terms of angle-
For a time-independent radiation field in a one-dimensional static planar dependent mean-intensity-like and flux-like variables, and that exact closure
medium, equation (2-74) reduces to of an angle-dependent equation that resembles a moment equation can be
obtained. This equation is easily discretized and solved. In sum, the solution
[opR(z, v)/oz] = -(41t/c)x(z, v)H(z, v) (2-77a) of the transfer equation in terms of moments or equivalent variables can be
effected by differential-equation techniques of great generality and power.
or, integrating over all frequencies

[opR(z)/oz] = -(41t/c) So"' x(z, v)H(z, v) dv (2-77b)


2-4 The Condition of Radiative Equilibrium
Alternatively we can write
Deep within the interior of a star, nuclear reactions release a flux of energy
[ oK(z, v)/oz] = - x(z, v)H(z, 11) (2-78) that diffuses outward, passes ultimately through the star's atmosphere, and
emerges as observable radiation. In normal stars there is no creation of
or (oK,/iJt,) = H. (2-79) energy within the atmosphere itself; the atmosphere merely transports out-
ward the total energy it receives. In a time-independent transport process,
In spherical geometry, under the same assumptions, equation (2-74) reduces, the frequency distribution of the radiation, or the partitioning of energy
with the aid of the expression derived in Exercise 1-10 for (V · P),, to between radiative and nonradiative modes of transfer, may be altered; but
(2-80)
the energy flux as a whole is rigorously conserved.
There are two basic modes of energy transport in those atmospheric layers
Exercise 2-12: Derive equation (2-80) directly from equation (2-33). in which spectrum-formation takes place: radiative and convective (or $Orne
other hydrodynamical mode). In these layers conduction is ineffective and
Thus far we have examined the moment equations primarily from the can be ignored (it becomes important in coronae at temperatures of the order
standpoint of their dynamical significance; but in the time-independent case of I 0 6 °K). When all of the energy is transported by radiation, we have what is
w
I\)
they may also be used as tools to solve the transfer equation. By the introduc- called radiative equilibrium; conversely, pure convective transport is called
tion of moments, the angle-variable is eliminated and the dimensionality of convective equilibrium. Whether or not radiative transport prevails over
I
the system to be solved is reduced. As we have seen in §2-2, the mean intensity convection is determined by the stability-of the atmosphere against convec-
can be determined from the solution of an integral equation (see Exercise tive motions.
2-10). This gives the source function, from which the higher moments (e.g., The criterion for the stability of radiative transport was first enunciated
the flux) can be determined by quadrature. The question now arises whether by K. Schwarzschild (416, 25)in another of the fundamental papers of radia-
we can solve the moment equations as differential equations. Examination of tive transfer theory. Schwarzschild was able to demonstrate convincingly
equations (2-71) and (2-79) immediately reveals an essential difficulty: the that the dominant mode of energy transport in the photosphere of the sun is
moment equation of order n always involves the moment of order n + 1, radiative. Since his work, a number of analyses of radiative stability have
hence there is always one more variable than there are equations to deter- been carried out for a variety of stellar types; results are summarized in
mine them! This difficulty is known as the closute problem: one additional (638,215; 11,449; 654,432). The basic picture that emerges for a star like the
relation among the moments must somehow be obtained to "close" the sun is that radiative equilibrium obtains to continuum optical depths of order
system. For solving transfer equations, a variety of methods exist that employ unity, and that below this depth the atmosphere becomes unstable against
moments of arbitrarily high order, and introduce ad hoc closure relations convection. Convection zones exist below the outer radiative zone in all stars
[see, e.g., (361, 90-101; 365)]. However, in this book attention will be con- of spectral type later than about F5. For earlier spectral types, radiative
fined entirely to the moments J •• H., and K, (an exception appears in'§14-3), equilibrium prevails throughout the entire outer envelope of a star. I_n. this
and the system will be closed (see §6-3) by eliminating K, in terms ofJ. and book primary concern will be given to the early-type stars, and accordin•g~y
'•
48 The Equation of Transfer 2-5 The Diffusion Approximation 49

emphasis will be given to the radiative equilibrium regime. The theory of Although the radiation emerging from a star is by no means Planckian, it
convective transport is not, at present, in a fully satisfactory form, and only is nevertheless customary to define the effective temperature as the temper-
the mixing-length theory (a phenomenological approach that has been widely ature a black body would have in order to emit the actual stellar flux-i.e.,
applied in astrophysics) will be described (§7-3).
Let us now consider some of the implications of radiative equilibrium;
assume the medium is static and the radiation field is time-independent.
uR T:rr = f
0
00
F, dv = 41t f
0
"" H, dv = L/(4nR 2) (2-85)

From the discussion in §2-1, it is clear that the total energy removed from the Here Lis the total luminosity and R is the radius of the star; the atmo-
beam is spheric thickness is assumed to be negligible compared to R. Although T.rr

f dv pdw x(r, v)l(r, n, v) = 41t f


0
00

0
00
x(r, v)J(r, v) dv (2-81)
has only an indirect physical significance, it is a convenient parameter with
which to characterize the atmosphere, for typically the actual kinetic tem-
perature Twill equal T.rr near the depth from which the continuum radiation
where x denotes the total extinction coefficient. The total energy delivered emerges (i.e., unit optical depth at frequencies where the opacity is lowest).
by the material to the radiation field is In spherical geometry equation (2-84) implies

S dv pdw t7(r, v) = 4,r S x(r, v)S(r, v) dv


0
""
0
"" (2-82) r2F = constant = L/4,r (2-86)
In an extended atmosphere it is no longer really possible to choose a unique
where equation (2-35) has been used. The condition of radiative equilibrium radius R for the star and to define a unique value of T.rr; rather, L or the
demands that the total energy absorbed by a given volume of material must quantity r2 F should be regarded as fundamental. However the identification
equal the total energy emitted; thus at each point in the atmosphere R = r(,R = ½) is sometimes made, and a value of "Terr" derived with this
radius; here 'R is the Rosse/and optical depth scale (cf. §3-2).
41t S [tf(r, v) - x(r, v)J(r, v)] dv = 0
0
"" (2-83a) Finally, it is important to return to the point raised in the discussion of
the formal solution. Suppose that the opacity K, is independent of T; then
4,r f x(r, v)[S(r, v) - J(r, v)] dv = 0 J
00
or (2-83b) for physically reasonable distributions of ,c., .the integral ,c,B,(,) dv (which
0
gives the total thermal emission) will be a monotone increasing function w
Exercise 2-/J: Suppose that s. is given by equation (2-39). Show that, in the of T. Thus when we fix the total thermal emission at some value, we fix the w
condition of radiative equilibrium, the scattering terms cancel out to yield local value of T. It is then clear from equation (2-83b) (and from the result I
of Exercise 2-13) that the local value of Tis determined by the mean intensity,
f
0
"' ic,B,(T) dv = f 0
"' ,c,J• dv which depends upon the global properties of the atmosphere because it
follows from a solution of the transfer equation. Thus the temperature at a
Using equation (2-83) in equation (2-69) we have, alternatively, given point in the atmosphere is to some extent determined by the temper-
ature at all other points and, at the same time, helps to establish the tempera-
(2-84) ture structure elsewhere. This nonlocalness in the problem is a result of
Hence in planar geometry the condition• of radiative equilibrium is equivalent radiative transfer, through which photons moving from one point in the
to the requirement that the depth derivative of the flux is zero-i.e., the flux medium to another lead to a fundamental coupling (i.e., interdependence)
rs constant. Physically, equations (2-83) and (2-84) have the same meaning, of the properties at those points.
but mathematically the requirement F = constant is rather different from
the expressions in equations (2-83); either form of the constraint may be
used in constructing model atmospheres. 2-5 The Diffusion Approximation
Because the total flux is constant in a planar atmosphere, it may be used
as a parameter that describes the atmosphere; an equivalent quantity often At great depths in a semi-infinite atmosphere the properties of the radiation
employed is the effective temperature. From Exercise 1-6 we know that the field and the nature of the transfer equation become extremely simple. We
integrated flux from a black body of absolute temperature Tis Fee = uR T 4 • can obtain in a straightforward way an asymptotic solution that applies
50 The Equation of Transfer
2-5 The Diffusion Approximation SI

throughout the interior of a star (except, of course, in convective zones),


i.e., Id" B,/dT:I ~ B,/T:. Then it is clear that the ratio of succe_ssive terms in
and that provides a lower boundary condition on the transfer problem in the series is of order O(1/T, 2 ) or O(1/(x.) 2 ~z 2 ) where (X,) 1s the average
the stellar atmosphere. Consider first the properties. of the radiation field. opacity over the path-length ~z. In terms of the photon mean-free-path
At depths in the medium much larger than a photon mean-free-path, the /,. ~ 1/x., the convergence factor is O(/,.2/~z 2). It is clear that the convergence
radiation is effectively trapped, becomes essentially isotropic, and (eventually) is quite rapid; indeed the estimate just given turns out to be conservative.
approaches thermal equilibrium so that S, ➔ B,. Choose a reference point Also, it is obvious that convergence will be most rapid at frequencies where
'• » 1, and expand S, as a power-series: the material is quite opaque. For a star as a whole one expects ~z to be some
CX) significant fraction of a stellar radius, say ~z ~ 10 10 cm, while (x) ~ l
S,(t,) =L [d"B,/dt:](t. - ,,)"/n! (2-87) (which implies a photon mean-free-path of 1 cm), so that the convergence
... o factor of the series is of the order 10- 20 • It is clear that in the deep interior of
Calculating the specific intensity for O ~ µ ~ l from this source function the star only the leading terms are required.
with equation (2-50) we have In the limit of large depth we may therefore write

~ • d" B, d2 B,
I.(,.,µ) ~ B.{t,) + µ(dB,/dT.) (2-90a)
!,(,.,µ)= '- µ - d• = B,t,)+µ-d
( dB,
+µ 2 - d 2 +"· (2-88)
•'"0 T, '• T, J.{t,) ~ B.{t,) (2-90b)
A similar result for -1 ~ µ ~ 0 follows from equation (2-51) and differs l
H .(,.) ~
from equation (2-88) only by terms of order e-•h•; in the limit of great depth
the latter vanish and equation (2-88) applies on the full range -1 ~ µ ~ 1.
3 (dB,/d,,) (2-90c)

By substitution of equation (2-88) into the appropriate definition~ we find 1


~
for the moments K,(,,.)
3 B,(t,) (2-90d)
CX)

J.{t.) = L (2n + l)- 1(d 2"B,/dt.2") In equation (2-90a) we have retained two terms so as to account for the
n•0 nonzero flux [ cf. equation (2-90c)]. Note that equations (2-90b) and (2-90d)
show that Jim,_"' [K..{t.)/J.(,.)] = ¼, which is what we would expect for
(2-89a) isotropic radiation; we shall show below that the ratio of the anisotropic
to isotropic term in /(T, 11) becomes vanishingly small as , .... oo, so that the
CX) limit just found is appropriate. Insofar as the specific intensity as given
H.{t,) = L (2n + 3)- 1(d 2 "+ 1 B,/dt~•+l) by equation (2-90a) was computed from the formal solution of the transfer
n•0
equation, in effect using a source function S,(t,) = B.(t,) + (t. - ,.)~dB.fd,.),
1 equations (2-90) should obviate further use of the transfer eq?at1~n. It 1s
= 3 (dB,/dT,) + · · · (2-89b) easy to verify that this is so, for by inspection one sees that subst1tut1on from
equations (2-90) reduces both the transfer equation (2-36) and the zero-order
moment equation (2-71) to the single requirement d2 B,/dt.2 = 0 (already
CX)

and K.{t,) =L (2n + 3)- 1(d2"B,/dT/") assumed) while the first-order moment equation (2-79) is identical to (2-90c).
n=O
Thus, in 'effect, at great depth the transfer problem reduces to the single
1 1
= -3 B (T ) + -5 (d 2 B /d, 2
) + .. · (2-89c) equation
Mv)
V y y y •

H• _ 3l (aB.)
a,.
= _l ( ~
3 Xv
(dT)
aT. dz (2-91)
Note that only even-order terms survive in the even moments J, and K, and
only odd-order terms in H,. It is clear that equations (2-90) and (2-91) can be used, as mentioned in §2-2,
We now inquire how rapidly these series converge. The derivatives can be to set lower boundary conditions on the transfer equation in a semi-infinite
approximated, at least to order-of-magnitude, by appropriate differences- atmosphere.
52 The Equation of Transfer

Equation (2-91) [and also equations (2-90)] is referred to as the diffusion


approximation, primarily because of its formal similarity to other diffusion
equations, which are of the form
3
(flux) = (diffusion coefficient) x (gradient of relevant physical variable)

e.g., <I> = - " vr for heat conduction. The coefficient ¼x. - 1(oB,/oT) is, in
fact, sometimes called the radiative conductivity, a designation that is quite The Grey Atmosphere
appropriate in view of the fact that Xv - i = I, is the photon mean-free-path.
Note that equation (2-91) exhibits the essential physical content of our
earlier result that the flux computed by application of the <I>-operator to a
linear source function depends only on the gradient of S [ see discussion
following equation (2-65)]. Also it shows that the mere fact that energy
emerges from the star implies that the temperature must increase inward.
Indeed, replacing H with (L/41tR 2 ) and (dB/dz) with (uR T,4/nR), and taking
(x) ~ I, it is easy to show that the central temperature T, of the sun must
be of the order of 6 x 10 6 °K, a result consistent with our earlier statement
that the ultimate energy source in a star is thermonuclear energy-release at
the center.
In an intuitive picture of diffusion, one usually conceives of a slow leakage
from a reservoir of large capacity by means of a seeping action. These ideas
apply in the radiative diffusion limit as well. The diffusion approximation
becomes valid at great optical depth (i.e., many photon mean-free-paths
from the surface) whence many individual photon flights, with successive
absorptions and emissions, are required before the photon finally trickles to w
the surface, and issues forth into interstellar space. The grey atmosphere problem provides an excellent introduction to the study 01
Ifwe integrate equations (2-90a-c) over all frequencies, we obtain /(-r, µ) :::::: of radiative transfer in stellar atmospheres. The nature of the defining as- I
B(r) + 3µH. The ratio of the anisotropic to isotropic terms gives a measure sumptions is such that the problem becomes independent of the physical
of the "drift" in the radiation flow; this ratio is state of the material, and requires the solution of a relatively simple transfer
equation. At the same time, the grey problem demonstrates how the con-
Anisotro~ic term ~ 3H = ~ (O'R r:,r)/(O'R T ~ (T•,r)
4
)
4
2_92
straint of radiative equilibrium can be satisfied, and the solution can be
Isotropic term B 4 1t 1t T ( ) related to more general and more realistic physical situations. Furthermore,
an exact solution of the problem can be obtained, and this provides a com-
Clearly at great depth, where T becomes » T.rr, the "leak" becomes ever parison standard against which we can evaluate the worth of various approxi-
smaller. The same result is found by a physical argument from a slightly dif- mate numerical methods that can be applied in more complex cases.
ferent point of view. If nF is the energy flux carried from an element of
•material by photons of velocity c, the rate of energy flow per unit volume is
(nF /c); the energy content per unit volume is (41tJ /c) :::::: (41tB/c), so that (Rate
3-1 Statement of the Problem
of energy flow)/(Energy content) = (F/4B) = ¼(TerrfT)4. Again, we see that
diffusion, in the intuitive sense described above, occurs at great depth where The problem is posed by making the simplifying assumption that the opacity
T » T0 rr, while free flow of radiant energy occurs at. the surface where of the material is independent of frequency; i.e., x. = X· This assumption
T ~ T.rr· is of course unrealistic in many cases. Yet as we shall see in later chapters, the
opacity in some stars (e.g., the sun) is not too far from being grey and, in
54 The Grey Atmosphere
3-1 Statement of the Problem 55
addition, it is possible partially to reduce the nongrey problem to the grey
case by suitable choices of mean opacities. Thus the solution also provides a which implies the flux is constant, while the first moment gives
valuable starting approximation in the analysis of nongrey atmospheres. (dK/dt) =H (3-9)
If we assume x. = X, then the standard planar transfer equation (2-36)
becomes which, because H is constant, yields the exact integral
µ(81 ,/8t) = I• - S, (3-1)
1
Then by integrating over frequency, and writing K(t) = Ht + c = 4 Ft + c (3-10)

I = f0"' I, dv (3-2) To make further progress, we must relate J(t) to K(t). This is easily done on
the basis of the discussion in §2-5, where we showed that at great depth the
and similarly for J, S, B, etc., we have specific intensity is quite accurately represented by /(µ) = / 0 + / 1µ, which
produces a nonzero flux and also implies that, fort » 1, K(t) = ¼J(t). Thus
µ(8//8t) =I- S (3-3) the fact that K(t) -+ ¼Ft fort » 1 implies that at great depth
If we impose the constraint of radiative equilibrium [equation 2-83b)], we 3
(t » 1)
require J(t) -+
4 Ft (3-11)
(3-4)
That is, asymptotically the mean intensity varies linearly with optical depth.
On general grounds we expect the behavior of J(t) to depart most from
which, for grey material, reduces to J = S. Thus equation (3-3) becomes linearity at the surface [note equation (2-63)], which suggests that a reason-
µ(a1;at) =1- 1 (3-5) able general expression for J(t) is
3 3
which has the formal solution [equation (2-57)] J(,) = 4 F[, + q(t)] = 4 (uR r:rrM[t + q(t)] (3-12)

J(t) ~ A,[S(t)] = A,[J(t)] = ~ f


0
"' J(t)Edt - ti dt (3-6) The function q(t), known as the Hopf function, remains to be determined; w
from equation (3-6) it is clear that q(t) is a solution of the equation CJ)
Equation (3-6) is a linear integral equation for J known as Milne's equation; I
the grey problem itself is sometimes called Milne's problem. It is important to t f"' · [t + q(t)]Edt -:- ti dt
+ q(t) = 1 Jo (3-13)
recognize that, when a solution of equation (3-6) is obtained, it satisfies
2
simultaneously the transfer equation and the constraint of radiative equilib- Finally, we notice that because
rium. The determination of such solutions in the nongrey case will occupy
most of Chapter 7. 1 1
Ifwe now introduce the additional hypothesis ofLTE, thens. = B.(T)-
Jim [- J(,) - K(t)]
t"'.. 00 3
= -4 F lim [t
, ... a:>
+ q(t) - t - c] =0 (3-14)
which, from the condition of radiative equilibrium, implies that
we have c = q(oo) and hence can write equation (3-10) as
(3-7)

Thus, ifwe are given J(t), the solution of the integral equation (3-6), then the
K(t) = 41 F[t + q(oo)] (3-15)
additional premise of LTE allows us to associate a temperature with the
radiative equilibrium radiation field via equation (3-7). The solution of the grey problem consists of the specification of q(t). Given
Several important results may be obtained from moments of equation (3-5). q(,), the temperature distribution is obtained by combining equations (3-7)
Taking the zero-order moment and imposing radiative equilibrium we have and (3-12) into the relation

(dH/dt) =J - S =J - J =0 (3-8) (3-16)


56 The Grey A tmospl,ere 3-2 Relation to the Nongrey Problem: Mean Opacities 57

We shall derive approximate expressions for q(-r) in §3-3 and describe the where the second equality produces the desired identification with equation
exact solution in §3-4. First, however, it is useful to delineate the nature, and (3-19b) ifwe define
extent, of the correspondence between the grey and nongrey problems. (3-21)

The opacity XF is called the flux-weighted mean. Note that this choice does
3-2 Relation to the Nongrey Problem: not reduce the nongrey problem completely to the grey problem, for the
Mean Opacities monochromatic equation (3-18a) does not transform into equation {3-18b)
with.this choice ofx_. Further, there is the practical problem that H. is not
The opacity in real stellar atmospheres usually exhibits strong frequency known a priori, and therefore XF cannot actually be ca,lculated until after the
variations, at least when spectral lines are present. Although numerical transfer equation is solved. This latter difficulty can be overcome by an
methods now exist that allow a refined solution ofnongrey transfer equations iteration between construction of models and calculation of XF· Although
and an accurate determination of the temperature structure in a nongrey the desired goal has not been fully attained, the fact that the flux-weighted
atmosphere, the calculations are, at best, laborious, and it is important to ask mean preserves the K-integral is important, for it implies that the correct
whether a significant connection exists between the grey and nongrey cases. value is recovered for the radiation pressure [cf. equation {1-41)]. It also
We shall show in this section that such a connection, though limited in scope, follows that the correct value of the radiation force, which is the gradient of
does exist, and that, among other things, it permits the temperature distribu- the radiation pressure, is likewise obtained. Thus from equation (2-77)
tion of the deepest atmospheric layers to be determined quite accurately we have
from the grey solution.
Let us first compare side-by-side the grey and nongrey transfer equations. (dpR/dt) = - x.F - 1(dpR/dz) = (4n/cx.F) f0
00
x.H. dv = (41tH/c) = {<TT:rr/c)
Starting with the transfer equation and calculating the zero and µrst-order
moments we have, in the nongrey and grey cases respectively: (3-22)

µ(ol,/oz) = X,(S. - /,) (3-17a) µ(ol/oz) = x(J - /) (3-17b) so that use of the flux-mean opacity produces a simple expression for the
radiation pressure gradient. This is a result of practical value in the computa-
(oH./oz) = x.(s. - J,) (3-18a) (dH/dz) =0 (3-18b) tion of model atmospheres for early-type stars, because in these objects w
radiation forces strongly affect the pressure (or density) structure of the --..J
(3-19a) (dK/dz) = - xH (3-19b) atmosphere via the condition of hydrostatic equilibrium {or momentum I
balance in steady flow).
Here variables without frequency subscripts denote integrated quantities, as
in equation (3-2). We now ask whether it is possible to define a mean opacity
7., formed as a weighted average of the monochromatic opacity, in such a way ROSSELAND MEAN
that the monochromatic transfer equation, or one of its moments, when
Alternatively, suppose we wish to obtain the correct value for the inte-
integrated over frequency, has exactly the same form as the grey equation.
grated energy flux. From equations {3-19) it follows that this may be done if
Several possible definitions have been suggested.
x. is chosen such that

FLUX-WEIGHTED MEAN

Suppose we wish to define a mean opacity in such a way as to guarantee


or, equivalently,
an exact correspondence between the integrated form of equation (3-19a)
and the grey equation {3-19b). If such a mean can be constructed, then the
relation K{Y) = Ht + c will again be an exact integral, .as it was in the grey
X.- 1 = fo
00
Xv - 1(iJK.f oz) dv Ifo 00
(iJK./iJz) dv (3-24)

case. Integrating equation {3-19a) over all frequencies we have


Again we face the practical difficulty that K. is not known a priori, and hence
{3-20) the indicated calculation cannot be performed until the transfer equation is
58 The Grey Atmosphere 3-2 Relation to the Nongrey Problem: Mean Opacities 59

solved. But the mean defined in equation (3-24) can be approximated in the or the radiation force correctly in the outermost regions of the atmosphere.
following way: at great depth in the atmosphere, K.-+ ¼J. while J.-+ B•. This point must be recognized clearly, for it is precisely these layers in which
Thus may write (oK./oz) ~ ½(oB./oT)(dT/dz). We then define the mean spectrum-formation occurs, and hence which are of primary interest in the
opacity XR as analysis of stellar spectra.

PLANCK AND ABSORPTION MEANS

(3-25) Several other expressions for mean opacities may be chosen. For exam-
ple, if we demand that the mean be defined to. yield the correct integrated
value of the thermal emission, then we require
or (3-26)
Kp = [J
0
00
K.B.(T) dv]/ B(T) = 1t f
0
00
K.B.(T) dv / u8 T
4
(3-28)
The opacity XR is called the Rosse/and mean in honor of its originator. Note that only the true absorption is used, and scattering is omitted. The
Note that the harmonic nature of the averaging process gives highest weight opacity icp is known as the Planck mean; it has the advantage of being
to those regions where the opacity is lowest, and, where as a result, the greatest calculable without having to solve the transfer equation. On the other hand,
amount of radiation is transported-a very desirable feature. Again the use of icp does not permit a reduction of equation (3-18a) to (3-18b) nor of (3-19a)
XR or the mean defined by equation (3-24) does not permit a correspondence to (3-19b), and therefore it lacks the desirable features possessed by XF or
between equations (3-18a) and (3-18b), and hence does not allow the nongrey XR• Nevertheless this mean does have additional significance.
problem to be replaced by the grey problem. On the other hand, it is obvious In particular, near the surface of the star, the physical content of the
that the approximations made to obtain equation (3-26) are'precisely those condition of radiative equilibrium is expressed most directly by equation
introduced in the derivation of the diffusion approximation to the transfer (3-4). In view of this constraint, a correspondence between equations (3-18a)
equation (2-91); i.e., and (3-18b) can be made near the surface if ic satisfies the relation
H

= _!(~ oB•)(dT)
3 XvoT dz (3-29) w
(l)
Hence use of f 8 is consistent with the diffusion approximation. Therefore Once the material becomes optically thin (i.e., t. < l at all frequencies), J. I
on the Rosseland-mean optical-depth scale t 8 we must recover the correct becomes essentially fixed, and the integral above will be dominated by those
asymptotic solution of the transfer equation and the correct flux transport frequencies where "• » ic. If we represent B. by a linear expansion on a
at great depth. This implies that at great depth \tR » 1) the temperature 'f-scale, i.e.,
structure is quite accurately given by the relation T4 = ¾T:rr[t8 + q(tR)];
see equation (3-16). It is therefore clear why Rosseland mean opacities are B,(t) = B,(t) + (dB./dt)(t .:.. t) ~ B.(t) + (dB,/dt)(R/K,)(t, - t,),
employed in studies of stellar interiors. Note also that so long as the diffusion
then by application of the A-operator we find [cf. equations (2-57), (2-58),
approximation is valid, a simple expression can be written for the radiation
and (2-63)],
force, namely
1
(3-27) J,.(t) - B,.(r) ::::: -
2 B,(r)Ei{t,.)
Exercise 3-1: Derive equation (3-27).
+ (K/K.)(dB,./d"i) G E3(t.) + ~ t,E2(t.)] (3-30)
Although the diffusion approximation is nearly exact at great depth, and
provides the very useful results just discussed, it must of course break down In the limit t-> 0, E 2(t)-> 1, and E 3(t)-> ½, so the first term yields -½B.Jt)
at the surface, and exact flux conservation is not guaranteed in the upper while the second becomes least important when"• » ic [precisely the regiori.
layers by use of the Rosseland mean, nor will it give the temperature structure of highest weight in equation (3-29)]. Thus to satisfy equation (3-29), ~
60 The Grey Atmosphere
61
3-3 Approximate Solutions

should essentially fulfill the requirement that this relation is also valid for a variety of other situations, including the
f CJ)
0
K,B, dv =R f
0
CJ) B, dv (3-31)
two-stream approximation, which provides a rough representation of the
radiation field near the boundary. In view of these results, Eddington made
which shows that the Planck mean is the choice most nearly consistent with the simplifying assumption that J = 3K everywhere in the atmosphere. Then
the requirement of radiative equilibrium ni::ar the surface. the exact integral K =¼Ft+ c implies that in the Eddington approximation
Alternatively, we might demand that the mean opacity yield the correct J E(,:) = ¾Ft + c'. To evaluate the constant c' we calculate the emergent
total for the amount of energy absorbed. We then obtain the absorption flux and fit it to the desired value. Thus from equation (2-59) we have
mean
(3-32)
F(0) =2 f
0
"' (~ Ft + c') E (t) d,: = 2c'E {0) + ~· F [;- 2Ei0)]
2 3 (3-33)

Again only the true absorption is included, arid scattering is omitted. As so that, using the relation En(0) = 1/(n - 1) and demanding F(0) = F, we
was true forxF, we cannot calculate R1 until a sol'ution of the transfer equation find c' = ½F. Thus
has been obtained. Further, ic1 does not permit a strict correspondence
between the grey and uongrey forms of the transfer equation or any of its JE{t) = ~F(,: + ~) (3-34)
moments (as was also true for the Planck mean).
In Eddington's approximation q(t) = j. Imposing the constraint of radiative
SUMMARY
equilibrium and the assumption of LTE we have from equation (3-16)

We have seen that no one of the mean opacities described ab.ove allows,
in itself, a complete correspondence of the nongrey problem to the,grey
T 4 ::::: 3
4 Terr
4 (
,: + 3 2) (3-35)

problem. Yet mean opacities provide a useful first estimate of the temperature
structure in a stellar atmosphere if we assume, as a starting approximation, Equation (3-35) predicts that the ratio of the boundary temperature to the
effective temperature is (T 0 /T.rr) = (½) 114 = 0.841, which agrees fairly well
T(rR) = T 1,cy (fR), and then improve this estimate with a correction procedure
with the exact value (T0 /T.rr) = (3 112/4) 114 = 0.8114. Assuming S(t) = J,,;{t)
w
that is designed ,to enforce radiative equilibrium for the nongrey radiation (0
field. Indeed, the mean opacities XF, Kp, and R1 appear explicitly in some we may calculate the angular dependence of the emergent radiation field
in the Eddington approximation by substituting equation (3-34) into I
temperature-correction procedures.
From an historical point of view, it should be recognized that, before the equation (2-52) to obtain
advent of high-speed computers, the nongrey atmosphere problem required
far too much calculation to permit a direct attack, and the use of XR and
Rp provided a practical method of approaching an otherwise intractable
Ie(0, µ) = ~ Ff 0
"' ( t + D 1
µ- exp( - t/µ) dt = ~ F (µ + D (3-36)

problem. In fact, the answers obtained in this way often do not compare which yields a very specific form for the Eddington-Barbier relation [cf.
too unfavorably with more recent results despite the apparent crudeness of equation (2-53)]. The center of a star's disk, as seen by an external observer,
the approximation. Only some of the more basic properties of mean opacities corresponds toe = 0°, orµ = 1; the limb is ate = 90°, µ = 0. The ratio
have been mentioned here; further information may be found in (419) and e
1(0, µ)/1(0, 1), which gives the intensity at angle = cos- 1 µ relative to
(361, §§34-35). disk-center, is referred to as the limb-darkening law. In the Eddington
approximation the limb-darkening is
3-3 Approximate Solutions
THE EDDINGTON APPROXIMATION
le(0, µ)flE(0, 1) = ~(µ + ~) (3-37)

This result predicts the limb intensity to be 40 percent of the central intensity.
Jn fl•~ it
was shown that, at great depth in a stellar atmosphere, the Observations of the sun in the visual regions of the spectrum are actually
~'lltionJ • 3K holds; further, in §l-4(cf. Exercise 1-13), it was demonstrated in good agreement with this value and, in fact, it was precisely this agreement
3-3 Approximate Solutions 63
62 The Grey Atmosphere
§6- t, §7-2.) Further, even a second application of the A-operator to JV,>(r)
that led K. Schwarzschild (416, 25) to propose the validity of radiative introduces the functions A,[£11 (1)], which are cumbersome lo compute [sec
equilibrium in the outermost layers of the solar photosphere. (361), equations (14-50), (14-53), and (37-36) through (37-44)]. Therefore
Equation (3-35) predicts that T = r.,r
when t = ½. This result has given alternative methods for obtaining a solution must be developed.
rise to the useful conceptual notion that the "effective depth" of continuum
formation is t ~ ½: in fact, this is often a rather good estimate. In support Exercise 3-3: Using the results of Table 3-2, evaluate the percentage errors of
of this idea we may note that a photon emitted outward from t = ½ has a J E(r) and J~21(t) and display them in a plot. The required values of E.(r) can be
chance of the order of e- 0 •67 ~ 0.5 of emerging from the surface; this found in (4, 24S).
corresponds in a reasonable way with the place we would intuitively identify Exercise 3-4: Show that, although J E(r) was derived assuming F = constant,
with the region of continuum formation. the flux computed from JE(r) via equation (2-59) is not constant; make a plot of
the error t,.F(r)/F.
Exercise 3-2: The Eddington-Barbier relation shows that the intensity /(0, µ) is Exercbe 3-5: Apply the X-operator [cf. equations (2-62) and (2-6S)] to JE(r)
characteristic of S(t) at t(µ) ~ µ. Show then that the average depth that charac- to obtain K~ 1(t) = i\-F[!t + J - !E4 (r) + 2E 5(t)]. Use this result to write an
terizes the.flux is (t) = ;. analytical expression for the variable Eddington factor /(r) = K(t)/J(t). Show that
/(r = oo) = ½and/(t = 0) = H = 0.405. Using the results of Table 3-2, evaluate
Anticipating the exact solution given in Table 3-2, we can evaluate the the fractional error in f(r) [recall equation (3-1S)] and plot it.
accuracy of J E(t); one finds that the worst error occurs at the surface, where Exercise 3-6: Show that the improved estimate of the emergent intensity obtained
by using 1r1(r) is /~2>(0, µ) = ¾F(rt + ½µ + (½µ + ½µ 2) In[(! + µ)/µ]}. Compare
/:i.J/J = (JE - J....,)/J...., = 0.155. Both the size of the error and the fact this result and 1£10, µ) given by equation (3-36) with the exact result shown in
that it occurs at t = 0 are unsurprising when we recognize that the basic Table 3-1, and plot a graph of their fractional errors.
assumption upon which the derivation is based, namely J = 3K, is known
to be inaccurate at the surface. We know that J(t) must satisfy the integral
equation (3-6), and we know further that the A-operator has its largest ITERATION: THE UNSOLD PROCEDURE
effect at t = 0 [see equation (2-63} and associated discussion]. This suggests
The primary shortcoming of the A-iteration procedure is its failure to
that an improved estimate of J(t} can be obtained from
yield an improvement in the solution at great depth. UnsBld (638, 141)
~
proposed an ingenious alternative method that overcomes this inadequacy
2
J~ >(,) = A,[JE(t)]. = A, [43F (t + 32 )=]43F [t + 32-13 Ei{t) + 1 0
2 E3(t)] and can be generalized to the nongrey case. The basic idea is to start from an
initial estimate for the source function B(t), and to derive a perturbation
I
(3-38) equation for a change /:i.B(r) that more nearly .satisfies the requirement of
radiative equilibrium.
Recalling the properties of En(t), it is clear that the largest difference between If we calculate the flux from the initial guess B(t), we will find that it is a
J~21 (r) and JE(t) occurs at the surface, where we find J<j 1(0)/JE(0) = t, The function of depth, H(r), and not exactly constant unless B(t) happened to be
new estimate of T0 /T.,, is thus ('f6) 114 = 0.813 (exact value is 0.8114) and the exact solution of the problem. From the first-order moment equation (3-9)
t
q(0) drops from to -i7I = 0.583 (exact value is 1/.,/3 = 0.577). we then have
It is thus clear that an application of the A-operator has produced a
dramatic improvement in the solution near the surface. Note, however, that K(t) = s:
H(t') dt' + C (3-39)
•there is no improvement in the solution at t ➔ oo, where q remains at its
original value½, In principle, successive applications of the A-operator should
If we make the Eddington approximation J(,) = 3K(t) and evaluate C by
improve the solution, and, eventually, produce the exact solution. In fact, writing J(O) = 2H(0) [cf. equation (3-34)] we obtain
one can show that lim" ... 00 A"(l) = 0 [see (684, 31)] so that an initial error e
at any depth can ultimately be reduced to zero by repeated application of
J(r) ~ 3 s: H(t') dt' + 2H(0) (3-40)
the A-operator. In practice, however, the convergence is too slow to be of
But from the zero-order moment equation (3-8) we have
value, for the effective range of the A-operator is of order /:i.r = 1, so errors
at large depth are removed only "infinitely slowly." (We shall encounter this B(t) = J(r) - [dH(t)/dt] (3-41)
problem with A-iteration repeatedly in a wide variety of contexts! See, e.g.,
<,4 '11w Grey Atmosphere
3-3 Approximate Solutions 65
so that
B(t) ~3 J: H(t') dt' + 2H(0) - (dH(t)/dt] (3-42)
transfer equation to be solved may be written in the form

Equation (3-42) cannot be exact because of the approximations made in its µ[8/(r, µ)/8r] = /(t, µ) - f
~ ~ 1 I(t, µ) dµ (3-45)
derivation, but it can be used with sufficient accuracy to compute perturba-
tions. In particular, suppose AB(t) is chosen just so the flux computed from which is classified as an integro-differential equation. The essential difficulty
B(r) + AB(r) is constant; thus in obtaining the solution arises from the presence of the integral over angle.
However, definite integrals such as that in equation (3-45) may be evaluated
B(t) + AB(t) ~ 3 J: H dt' + 2H (3-43) numerically by means of a quadrature sum consisting of the function evaluated
at a finite set of points on the interval of integration times appropriate weights.
and by subtraction of equation (3-42) from (3-43) we obtain an expression Thus introducing {µ1} on [ -1, 1] we write, for any function/(µ),
for 68(-r), namely
1 f I 1 n

AB(t) =3 f: AH(t') dt' + 2AH(0) - [d 6H(t)/dt] (3-44) 2 - I /(µ) dµ ~ 2 l"f.:_• aJ!(µ1) (3-46)

Thus if we know the flux errors AH(r) = H - H(t) we can compute the The numbers {µ1} are called the quadrature points, {a1} the quadrature
correction AB(t); this correction is then applied and new values of the flux weights, and {!(µ 1)} the discrete ordinates. Having chosen a definite quad-
are computed, which give new (smaller!) errors AH; the process is iterated rature formula, we replace the integro-differential transfer equation (3-45)
until H becomes constant and AB ➔ 0 at all t. Equation (3-44) can be with a coupled system of 2n ordinary differential equations:
generalized for nongrey atmospheres: see equation (7-18). UnsBld's pro- 1 "
cedure is very powerful compared to A-iteration, for it provides a great
improvement in the solution at depth as well as at the surface; this result'is
µ,(8Id8t) = I, -
2l•L-n ai1, (i = ± 1, ... , ±n) (3-47)

demonstrated in the following exercise. where 11 denotes J(r, µ1). The radiation field is no longer represented as a
continuous function, but rather in terms of a set of pencils of radiation, each
Exercise 3-7: Assume a starting solution B(t) = 3H(t + c); i.e., q(t) = c.
of which represents the value of I(µ) over a definite interval. On physical __.
(a) Show that 6H(t) = H - H(t) = ½H[E4(t) - cE 3(t)]. Obtain expressions for
6H(0) and d(6H)/dt. (b) Apply Uns!:Sld's procedure and show that
grounds it is reasonable to expect the solution to become exact in the limit
n-+ co.
17
= 3H [ 24 1 1 3 3 ] The accuracy of the quadrature depends both upon the number of points,
6B(t) - c -
2cE2(t) + 2EJ{t) + 2cE4(t) - 2E5(t) and upon their distribution on the interval. If the points are distributed
(c) Show that, independent of the initial choice of c, the improved solution has uniformly on the interval we obtain a Newton-Cotes formula, of which
q(0) = -h = 0.583 (exact value 0.577) and q(oo) = ¼¾ = 6.708 (exact value = Simpson's rule with points at {µ 1} = (-1, 0, 1) is a familiar example. A
0.710). (d) Show that, in contrast, the A-operator acting on q = c gives q(0) = better choice is to use a Gaussian formula, in which the 2n points on [-1, l]
½c + ¼. which agrees with Unsl:lld's value only if c = ½, and q(oo) = c, which are chosen to be the roots of the Legendre polynomial of order 2n. It would
shows no improvement whatever at depth. take us too far afield to discuss the construction and accuracy of quadrature
formulae [see (161, Chap. 2)]; an important result that we shall merely state
THE METHOD OF DISCRETE ORDINATES is this: an m-point Newton-Cotes formula gives exact results for polynomials
of order m - 1 (for m even) or m (for m odd), but an m-point Gauss formula
The method to be described now furnishes a means of obtaining both
is exact for polynomials of order 2m - 1. For solving the transfer equation
approximate solutions and the exact solution of the grey problem. More
the double-Gauss formula is superior (619) to the ordinary (or "single")
important, it introduces the fundamental mathematical scheme that provides
Gauss formula. Here one chooses two separate n-point quadratures on the
the basis for practically all modern techniques of solving transfer equations. ranges (-1 :;;; µ :;;; 0) and (0 :;;; µ :.; l); on each range then points are given
Introducing the definition of J [equation (1-4)] into equation (3-5), the
by the 'roots of the Legendre polynomial of order n, scaled from [ -1, 1] to
66 The Grey Atmosphere 3-3 Approximate Solutions 61

the appropriate range. This approach has the advantage that /(-r, +µ)and where we have ordered {µi} such that µ 1 > µ1+ 1• Note that the largest µ 1
/(,, -µ) are approximated independently, and thus the integration formula must be less than unity, hence the smallest nonzero k2 must be greater than
can account, without difficulty, for the physical fact that /( - µ) = Oat , = 0 unity. In all there exist 2n - 2 nonzero values of the k's, in pairs of the form
while /( + µ) remains finite. In the single-Gauss formula, the discohtinuity in ±ki(i = 1, ... , n - 1).
/(µ)atµ = 0 when t = 0 introduces significant errors. In all of these formulae The general solution of the system (3-47) is therefore of the form
the points are chosen symmetrically about zero so that µ_ 1 = -µ 1, while
a_i = a1.
We now wish to solve the system of equations (3-47). Observing that the
/i{-t:) =b ["f
«=l
La(l + kaµ,)- le-k•• + "r,1L_a(l -
cz=l
kaµ,)- le+k••] (3-51)
system is linear and of the first order, we take a trial solution of the form
I, = g1 exp(-k-r), where g1 and k are to be specified. Substituting into equa- We must now seek the special solution corresponding to the root k2 = 0. In
tion (3-47) we find view of equation (3-11) which shows that J(t) must become linear in t at
depth, we examine trial solutions of the form / 1 = b(t + q1). Substituting
(3-48) this expression into equation (3-47) we obtain

so that g1 = c/( 1 + kµ 1). If we use this specific form for g1 and again substi- (3-52)
tute the trial solution into· equation (3-47), we find
Now observing that ifwe set/(µ) = µ in equation (3-46) the quadrature sum
(3-49) L aiµJ is zero, we see that equation (3-52) is satisfied by the simple choice
q1 = Q + µ 1• Hence the particular solution is /i(-i:) = b(t + Q + µ 1) and
the complete solution is
This is the characteristic equation, which is satisfied only by certain values of
k, called the characteristic roots (eigenvalues). Recalling that a_ 1 = a1 and
I'- 1 = - µ1, equation (3-49) can be used to define the characteristic function (3-53)
T(k 2 ), ~
n
We must now specify the 2n unknown coefficients b, Q, and L±«· This is I\)
· T(k
2
) =1- L a1(l - k2µ/)- 1 (3-50)
done by application of the boundary conditions. I
J•l
In the case of a.finite slab of total optical thickness T, both I_, - = 1(0, - µi)
The roots of T [i.e., those values of k for which T(k 2) = OJ are the desired and / 1+ = l(T, + µ1) are given functions ofµ. Thus we inay write 2n equations
characteristic roots. If we set /(µ) = 1 in equation (3-46) we see that for the 2n unknowns:

hence it follows that k2 = 0 is a solution of the characteristic equation; i.e., and


2
T(k = 0) = 0. ~here are an additional (n - 1) nonzero roots, which may
be seen as follows. Note that k2 = µ1- 2 is a pole of T, which becomes in- n-1 L e-k.T n-1 M )
finite for these values of k2 • For k2 = µ1- 2 - e, T(k 2 ) < 0, and by making [(T, +µ,) = b ( Q + T + µi + L 1a k + L 1- ~ = /, +
a=I + aµl a=I aµl
e arbitrarily small, T(k 2 ) ➔ - co. Similarly, for k2 = µ1- 2 + e, T(k 2 ) > o,
and as e ➔ 0, T ➔ + co. It is thus clear that T must pass through zero (3-54b)
somewhere on the interval between successive poles, hence the (n - 1) non-
zero roots must satisfy the relations where, to improve the numerical condition of the equations, M« = L_aek• 7.
Equations (3-54) may be solved by standard numerical techniques for linear
µ1-2 < k12 < µ2-2 < ... < k~-1 < µ.-2 equations.
3-3 Approximate Solutions 69
68 The Grey Atmosphere

In the case of a semi-infinite atmosphere in radiative equilibrium we have and thus the discrete-ordinate representation of q(t) is
the boundary conditions I_, - = 1(0, -µ 1) = 0, and demand that J(t) must n-1
not diverge exponentially as t -. oo. To satisfy the latter constraint we set q(t) =Q+ L L.e-k•• (3-61)
o=l
L_ 0 =0 (3-55)
Exercise 3-9: Derive equation (3-60).
and use the upper boundary condition to write the n equations
Numerical results for q(t) were obtained for n = 1, ... , 4 by Chandra-
•-1 sekhar (153) using a single-Gauss formula and by Sykes (619) using the
Q - µ1 + L L0 (l - k0 µ 1)- 1 = 0, (i = 1, ... , n) (3-56) superior double-Gauss formula. In every case the exact value q(0) = 1/Jj is
o•l obtained. The maximum percentage (absolute) error in J(t) for the single-
Gauss solutions is (9.0, 4.1, 2.7, 2.0) for n = 1, ... , 4, respectively, while for
Solution of equations (3-56) yields the n unknowns Q and L 0 • In addition, the double-Gauss it is (9.0, 1.8, 0.9) for n = 1, ... , 3, respectively. The
we require that the flux equal the nominal flux F. Thus we demand that double-Gauss estimates of q(oo), namely Q, are 0.71132 and 0.71057 for
n = 2 and 3; these values compare favorably with the exact value 0.710446.
(3-57) The double-Gauss solution with n = 3 gives the emergent intensity /(0, µ)/F
with a root-mean-square error of only 0.1 %, and is very accurate (0.02%) for
µ ~ 0.3. It should be stressed that the main importance of the discrete ordi-
Substituting equations (3-53) (with L_ 0 = 0), we have nate method is that in the limit 11 -+ oo it yields the exact solution and that
it affords an extremely powerful approximation technique for more com-
plicated problems.
Several very important results that will be useful in later work can be
established by analytical manipulation of the characteristic equation and
the boundary conditions. To simplify the notation we define x = 1/k and
In view of equation (3-46) the first sum is zero, the second equals ½, and X = 1/k 2 • The characteristic function defined in equation (3-50) may be
from the characteristic equation the fourth is found to be zero. rewritten in the equivalent forms

Exercise 3-8: Verify the. statement made above about the values of the sums in n a n n n 2

equation (3-58). T(X) =1- X L


J= I
X
-
j
µ}
2 = L al -
Jc 1
X L
J= 1 X
al
- µ}
2 = L1 µ}~JµJ- X
}=

Thus we find that b = ¾F, as would be expected from equation (3-11); note
(3-62)
also that the quadrature calculation yields a constant flux automatically. To clear T(X) of fractions, multiply through by TI'.i= 1 (µ/ - X); this yields
Finally, the complete solution for the semi-infinite atmosphere may be
written n n n
P(X) = fl (µ/ - X)T(X) = L a1µ,2 fl (µ/ - X) (3-63)
1 j= I I= 1 J;.f
l;(t) = 43 F [ t + Q + µ 1 + •-
•~• L.e-k••(l + k µ1)-
0
1]
, (i = 1, ... , n) which is clearly a polynomial of order (n - 1) in X. But we know that T(X)
(3-59) has the (n - 1) roots X 1 = 1/k/, ... , x._
1 = 1/k;_ i, so P(X) must have
the form C(X - X.) · · · (X - X •-.).To evaluate the constant, we note that
We may compute J(t) by substituting equations (3-59) into the quadrature the coefficient of the term in x•- 1 in equation (3-63) is (- 1)"- 1 1 a1µ/ = Li'=
formula; making use of the characteristic equation (3-49), we obtain (- 1r 1
½; this is simply C itself. We thus have

(3-60)
70 The Grey Atmosphere 3-4 Exact Solution 71

and therefore [note that unlike equation (3-37), the reference point here is the limb, not

T(X) = -31[n-1
TT (XJ - X) ]/nTT (µ/ - X) (3-64)
disk-center]. In the discrete-ordinate approximation we find from equations
(3-68) and (3-70)
J= I }• I

From equation (3-62) we see that T(X = 0) = 1, hence setting X


equation (3-64) we obtain the useful (see below) result that
= 0 in H(µ) = [CT (1 + µ, - µ)]/X( (1 + k,µ)
I (3-72)

By further analysis using S(µ) it is possible to write explicit expressions for


µ1µ2 · · • µnk1k2 ··•kn- I = 1//j (3-65)
the La's and Q in terms of the points {µ 1} and the roots {k«} [ see (161, 78-79;
Now consider the emergent intensity /(0, µ). Define a function S(µ) such 684, 25)]. ·
that Before leaving the discrete ordinate method, let us show that q(0) = 1/-/3

S(µ,) =Q - µ, + L
n-1 La(l - kaµ,)- 1 (3-66)
is the exact value. First, note that in the nth discrete approximation
«~1
(3-73)
The surface boundary conditions in equation (3-56) may then be written
while from equation (3-59) we find
1(0, - µ 1) = 43 FS(µ 1) = 0 (3-67)
(3-74)
We then generalize S(µ) to apply at all values ofµ and write
independent of the order n. Thus we conclude that in the exact solution J(O) =
/(0, µ) = 43 FS( - µ) ·(3-68) /(0, 0). But from equations (3-68) and (3-70)

forµ ~ 0. Note that, with this generalization, /(0, - µ) is not s 0, but will in I.(0, 0) = 43 FS(0) = 43 Fµ 1 · · · µ.k 1 · · · k._ 1 (3-75)
general have nonzero values for - µ ::/:: µ_ 1• By working with S(µ), we can
obtain an expression for /(0, µ) that does not involve the constants L« and Q
Then, combining equations (3-65), (3-75), and (3-73), we deduce that, inde-
explicitly. Clear fractions from equation (3-66) to obtain
pendent of the order n, q.(0) = 1/ J3;
hence this result is exact.
n-1 n-1 n-1 n-1
S(JI) TT (1 - kaµ) = (Q - µ) TT (1 - kaµ) + L La TT (1 - k,µ) (3-69)
,::al m• l «• l I 11

3-4 Exact Solution


The righthand expression is obviously a polynomial of order n in µ. But
S(µ) has the n roots µ 1, ... , µn; hence this polynomial must be of the form The exact solution for q(t) and H(µ) can be obtained by taking the discrete-
C(µ - µi) · · · (µ - µn). To find C we note that the coefficient µn on the ordinate method to the limit n ➔ ro (161, Chap. 5; 361, §27), by applying
righthand side of equation {3-69) is (- l)nk 1 ···kn- i, which is C itself. the principle of invariance (161, Chap. 4; 361, §28), or by direct Laplace
Therefore transform methods (361, §29; 684, Chap 3). Only certain important results
will be quoted here, and the reader should consult the references cited for
{3-70) details.
Several expressions for H(µ) exist (361, 186-187); a form convenient for
which, when inserted into equation {3-68), yields the desired expression for numerical computation is
1(0, µ). It is customary to define a limb-darkening function H(µ) as

H(µ) = 1(0, µ)/1(0, 0) (3-71)


H(µ ) = (µ + l)_ 112 ex
P
[! Jo
1t
r,.12 0tan 1
- (µ tan 0) de]
1 - 0 cot e
(3-76)
72 The Grey Atmosphere 3-5 Emergent Flux from a Grey Atmosphere 73

TABLE 3-1 be written (407) for q(T), namely


Exact Limb-Darkening Law for
Grey Atmospheres 1 fl e-,/u du
µ 1(0, µ)/F
q(T) = q(oo) - 2J3 Jo H(u)Z(u) (3-79)
µ 1(0, µ)/F

0.0 0.43301 0.6 0.95009 where H(u) is as defined above and


0.1 0.54012 0.7 1.02796
0.2 0.62802 0.8 1.10535 Z(u) = [ 1 - -1 u In (1-+-
2 1- u
u)] 2
+ -1 1t 2 u2
4
(3-80)
0.3 0.71123 0.9 1.18238
0.4 0.79210 1.0 1.25912
Results obtained from a numerical evaluation of equation (3-79) are given
0.5 0.87156
in Table 3-2.

Evaluation of this integral (152; 518) yields the results summarized in Table 3-5 Emergent Flux from a Grey Atmosphere
3-1.
The value of q(oo) can be obtained by noting from equation (3-15) that The basic physical assumption made in the grey-atmosphere problem is that
K(O) = Hq(oo), and thus, using equation (3-71), the opacity is independent of frequency. In this event, the constraint of
radiative equilibrium reduces to S(T) = J(T), and the problem simplifies to
that of obtaining the solution of equation (3-6). If, in addition, it is assumed
q(oo) = [ fo1 H(µ)µ2 dµ ] / fol H(µ)µ dµ (3- 77)
that LTE prevails, then we may equate B(t) = (<1R T 4)/7t to J(t), and thus
arrive at equation (3-16) for T(t). The radiation field has a dependence upon
From the known expressions for H(µ) one then obtains frequency because the source function, which we assume is B,(t), depends
upon frequency. Given the source function, the flux, also frequency-
6 1 f"/2 ( 3 1 )
q( 00 ) = 1t 2 + nJo 02 - 1 - 0 cot 0 dO (3-78) dependent, can be computed at any depth by means of equation (2-59),
which now reads ~
01
which yields q( oo) = 0.71044609 (519). Finally, a closed-form expression can F.(T) =2 f" B.[T(t)]E (t - 2 t) dt - 2 s: B,[T(t)]E 2(t - t) dt (3-81) I

TABLE 3-2 The temperature appears in the Planck function· only in the combination
The Exact Solution for q{t) (hv/kT); further, the ratio T(t)/Terr is a unique function of T [cf. equation
(3-16)]-say, 1/p(T). We may therefore simplify the equation by introducing a
t q(t)
parameter oc = (hv/kT.rd, in terms of which we can write (hv/kT) = ocp(t).
t q(t)

0.00 0.577351 0.8 0.693534 Expressing the flux in the same units by writing Fm(T) = F.(t)(dv/doc), and
0.01 0,588236 1.0 0.698540 using the relation F = (<1R r:rrV1t, we may rewrite equation (3-81) as
0.03
0.05
0.601242
0.610758
1.5
2.0
0.705130
0.707916 F.(T)
F =
( 41tk
4

h 3c2 <1R
)
(X
3 {J."'' exp[(Xp(t)]
E (t - t) dt
2
- 1-
f' E2 (t - t) dt }
Jo exp[(Xp(t)] - 1 3 82
( - )
0.10 0.627919 2.5 0.709191
0.20 0.6495S0 3.0 0.709806
The expression in the brackets is a function of oc and t only, and may be
0.30 0.663365 3.5 0.710120
calculated once and for all. A tabulation of Fm{t)/F is given in (161, 295),
0.40 0.673090 4.0 0.710270
and a plot of the function is displayed in Figure 3-1. The figure shows clearly
o.so 0.680240 s.o 0.710398 the degradation of photon energies as they transfer from depth to the sur-
0.60 0.685801 0.710446
00
face; for example, the most common photon energy at T = 0 is only about
74 3-6 Small Departures from Greyness 75

Let us now consider how we may solve the nongrey transfer problem using
a method of successive approximations. The nongrey transfer equation
0.20
(assuming LTE) is

µ(aI./fff) = (xvfi.)(/. - B.) = (1 + fl,)([. - B.) (3-85)


0.15
If we suppose that the departures fl. may be regarded as small, then a first
F.(r) approximation to the solution of equation (3-85) is obtained by setting
T 0.10
fl. = 0 initially. The transfer equation then returns to the equation for the
grey problem itself, namely

(3-86)
0.05
whose solution is already known. To obtain a second approximation we write

µ(an 2 >/8'f) = n2 > - B. + /3.(]~ 1 > - B.)


0 2 4 6 8 10 12
= J<;> - B, + fl,µ(oitl)/fff) (3-87)
Cl

FIGURE 3-1
by substitution from equation (3-86). If we demand that the radiation field
Frequency distribution of Hux at selected depths in a grey atmosphere. From
(lSS), by permission. resulting from this second approximation should satisfy the constraint of
radiative equilibrium, then we must have (dF(l)/d'f) = 0, where F< 2 > is the
75 percent of that at t = 1. This progressive reddening of photons in the integrated flux; then from equation (3-87)
outer layers results from the outward decrease in temperature produced by
the requirement of radiative equilibrium. J<l> _ B !!.. [ Jo1"' fl •Fo>
+ d"'i • dv] = 0 (3-88)
~
3-6 Small Departures from Greyness But note that equations (3-83) and (3-84) imply that 0)
I
By use of an appropriate mean opacity, it is possible to account for small 1"' (1 + fl.)F 0 > dv = -x F + -x 1"' fl
-x F = -xc Jo F 0V > dv (3-89)
departures from greyness, at least approximately, and thus to extend greatly
C .. y y Jo
C ' C V

the usefulness of the results obtained for a grey atmosphere. Suppose that the or that
frequency variation of the opacity is the same at all depths so that we can
(3-90)
write
(3-83)
Therefore the radiative equilibrium constraint for a nongrey atmosphere
treated by the above approximation scheme collapses to J< 2 > = B('f). This
where (3-84)
shows that the grey solution for T('f) on the Chandrasekhar-mean optical-
depth scale will automatically satisfy the condition of radiative equilibrium
In equation (3-84), F~1 > denotes the.flux in a grey atmosphere. The mean opac-
in the first approximation to the nongrey atmosphere. The method for
ity Xe defined in equation (3-84) is known as the Chandrasekhar mean; as we
obtaining higher approximations is described in (161,296 ff.)
shall see in what follows, this mean is constructed in a way that makes
At this first level of approximation we may compute the nongrey emergent
optimum use of the information contained in the grey solution. Unlike the
intensity as
flux mean [equation (3-21)] the Chandrasekhar mean can be computed
straightaway for any given frequency dependence (i.e., fl. or;,.) of the opacity
because F~1> is a known function.
I.(O, µ) = f0
"' B,[T('f)] exp(-y/f/µ)(y./µ) d'f (3-91)
76 The Grey Atmosphere

and the flux as

4
If we introduce the parameter ex = (hv/kT.rr) as was d"one in §3-5, and write
B.[T(f)] = B.(T0 ) [
exp(cxTcrr/T0 ) -
exp[cxp(f)] _ 1
1] = b _)
B.(To) a(t (3-93)
Absorption Cross-Sections
equations (3-91) and (3-92) reduce to the parametric forms

I.(O, µ) = B,(T0 ) f0
00
h«(f) exp(-y;t/µ)(y,/µ) df = B,(To)..l(rx, y,/µ) (3-94)

and

(3-95)

The function b«('f) is unique for a given value of rx, and hence the functions
f(rx, P) and :F(rx, P) may be computed once and for all; tables of these func-
tions are given in (161, 306-307).
The functions ..l(rx, P) and :F(rx, P) described above at one point played an
important role in the development of the theory of stellar atmospheres. By
analysing the observational material available for the sun, MUnch (473) was
able to find those values of y, that best reproduced the observed fluxes and
limb-darkening. These were shown to be compatible with the frequency ~
variation of the absorption coefficient of the negative hydrogen ion, H-, as In this chapter we outline the quantum mechanical calculation of atomic --...J
computed by Chandrasekhar. Analyses such as these led to conclusive absorption cross-sections. The discussion is meant to be self-contained, but I
identification of H- as the major opacity source in the solar atmosphere. limitations of space require that knowledge of the basic principles of quantum
We examine opacities in some detail in the next chapter. mechanics at the level of (392, Chaps. 2-8) be presupposed.

4-1 The Einstein Relations for Bound-Bound


Transitions
Let us first consider the absorption and emission of radiation by an atom
in a transition between two bound states. Assume that the lower state (i)
has statistical weight gi, and the upper state (j) statistical weight g1. There
are three basic processes involved, which are usually described in terms of
rate coefficients first introduced by Einstein (207).
The first process is the direct absorption of radiation, leading to an upward
transition from level i to level j. The rate at which this process occurs for
radiation of specific intensity /. can be written in terms of the Einstein
coefficient BiJ as
(4-1)
78 Absorption Cross-Sections
4-1 The Einstein Relations for Bound-Bound Transitions 79
where n1(v) is the number ofatoms per cm 3 in state i that can absorb radiation
by the Boltzmann law (cf. equation (5-5)]
at frequencies on the range (v, v + dv). In general the spectrum line corre•
sponding to the transition will not be sharp; rather, because of perturbation~
exerted by nearby atoms and ions, as well as the finite lifetime of the upper
(4-5)
state, it will have a spread in frequency that we can describe by an absorption
Moreover, in T.E., t/1. = 1/1: = <J,. [cf. equation (2-14)]. Now in strict T.E.,
J
profile, <J,., normalized such that <J,. dv = 1. Thus if the total number of each upward transition on the range (v, v + dv) must be exactly balanced
atoms in state i is n,. the number capable of absorbing at frequency v is
(detailed balancing) by an emission on that range. Hence, frequency by
n1(v) = n1<J,,. In making the transition from level i to levelj, the atom absorbs
frequency we must have
photons of energy hv1J = EJ - E1• Thus the rate at which energy is removed
from an incident beam of radiation is (4-6)
(4-2) Solving for B., we find

where a, denotes a macroscopic absorption coefficient (per unit volume),


uncorrected for stimulated emission (see below).
B, = n*B ~j:J~*B = (nAJ')[(g,BB1J) exp(hvlJ/kT) - 1]-1 (4-7)
I IJ J JI JI gJ JI
For atoms returning from level j to level i, two processes are possible.
The first of these is a spontaneous transition with the emission of a photon. But the correct expression for the Planck function as obtained from statistical
Writing the probability of spontaneous emission per unit time as AJ,, the rate m~chanical _arguments is B, = (2hv 3/c 2)(exp(hv/kT) - 1J-1, and to make
of emission of energy is this expression correspond to equation (4-7) we conclude that

(4-3) AJ1 = (2hv /c )BJ


3 2
1 (4-8)

Here the emission profile 1/t. specifies the number of atoms in the upper state and g1B1J = gJBJ1 (4-9)
that can emit photons on the frequency range (v, v + dv); it is normalized
We shall use these results repeatedly in later work.
such that Jt/1, dv = 1. The other possible return process is a transition ~
It should be not d that, although the Einstein relations in equations (4-8)
induced by the radiation field (stimulated emission). The rate at which such 7
and (4-9) were derived, for ease, from a thermodynamic equilibrium argu- (X)
emission occurs is assumed proportional to the intensity of the incident
ment, th~ Einstein coefficients are really properties of the atom only, and I
radiation field. The energy emitted may be written in terms of the Einstein
must be independent of the nature of the radiation field. Therefore we must
coefficient BJ1 as
(4-4) conclude that equations (4-8) and (4-9) are true in general. It is of interest
that historically a line of argument similar to that presented above led to the
realization that the stimulated emission process must occur in nature-a
In writing equation (4-4), use has been made of the result that the profile
fact not intuitively obvious at first sight.
for induced emission is the same as that for spontaneous emission, as can
be shown from general quantum mechanical considerations (197, §62). It
should be noted that spontaneous emission takes place isotropically. Induced Exercise 4-1: Show that for a Planckian radiation field at typical stellar temper-
atures ( ~ 104 °K) spontaneous emissions occur much more rapidly than induced
emission is proportional to and has the same angular distribution as J.;
emissions in the ultraviolet where hv/kT » I, while the reverse is true at far-
because of this, induced emissions are sometimes considered to be negative infrared and radio wavelengths where hv/kT « I.
absorptions, though this is not quite correct, for in general If,, will not be
identical to <J, ••
The coefficients AJ1t BJ,, and B1J are simply related, as can be shown by The microscopic formulation described above may immediately be incor-
calculating rates of absorption and emission in thermodynamic equilibrium. porated into the equation of transfer. If only the bound-bound process
In strict T.E., the radiation field is isotropic, and J. = B., the Planck function. occurs, then the appropriate transfer equation is
Furthermore, in T.E. the occupation numbers of levels i and j are related
80 Absorption Cross-Sections 4-2 The Calculation of Transition Probabilities 81

Here we have followed the usual practice of grouping together all terms into the source function, yielding what we shall refer to as an explicit form.
involving I•. In this way one can define a line absorption coefficient corrected As we shall see in Chapters 7 and 11 through 14, the latter form is by far
for stimulated emission, namely the more powerful and useful.

x,(v) = (n 1B1Jhv 1J)<J,.(l _ nJBJ11/1,) = (n1B1Jhv1J)·q,.(l _nnJgii/1•) (4


•ll)
41t n1BIJ<J,, 41t 1gJ<J,, 4-2 The Calculation of Transition Probabilities
and a line source Junction
Let us now turn to the calculation of the Einstein coefficients. Specifically,
S -
1
-
n1AJ11/1, - (2hv,/)
n1B1J<J,, - nJB11iJ,, - c2
[n'gl<J,•
n1g11/J,
-)]-1 (4-12)
we shall derive the direct absorption probability BIJ, as B11 and AJ1can be
obtained from BIJ by use of equations (4-8) and (4-9). The computation may
be made on three successively more accurate levels of approximation, as
The transfer equation then reduces to the standarct-form [eq~tion (2-36)]. follows.
In many cases of astrophysical interest, the simplifyin1/.pproximation (1) Classical Atom and Electromagnetic Field. The electron is considered
of complete redistribution is valid; then 1/1. = <J,. and equations (4-11) and to be a damped harmonic oscillator driven by the electromagnetic field.
(4-12) reduce to A unique absorption coefficient is derived, which is dimensionally correct
and accurate for very strong lines; for weak lines it may be wrong by orders

Xr
(v) = (n 1B11hv1J) <J,
41t •
(l _nig')
n,g1
(4-13)
of magnitude.
(2) Quantum Mechanical Atom and Classical Electromagnetic Field.
Here the correct values of BIJ and B11 can be derived, but A1, does not appear
and S, = (2hv,// c2 )[(n1gifnJg1) - 1J- 1 (4-14) in the formulation (although it is still given correctly by the Einstein relations).
(3) Quantum Mechanical Atom and Quantized Electromagnetic Field.
These expressions will be used through most of this book, except in'Chapter Here the correct results are automatically obtained for all three coefficients,
13 where a distinction between iJ,, and <J,. will be made. In the case of LTE, and this approach represents the complete theory.
Boltzmann statistics apply; hence (nifn 1) = (nifn,)• = (g1/g 1) exp(-hv,JlkT),' In this section we shall carry through the calculation by method (1) for
and the absorption coefficient becomes historical interest and general background, and by method (2) to obtain the
correct expression for Bli. Application of method (3) is more complicated
xf(v) = (n1B1Jhv 1if41t)<J,.[l - exp(-hv,JlkT)] (4-15) and does not need to be carried through if one is satisfied to use the Einstein
The factor [1 - exp( - hv/kT)] is usually referred to as the correction for relations. A complete discussion of the third method may be found, e.g.,
stimulated emission; but as is evident from equations (4-11) and (4-13), this in (197, Chap. 10; 293, §§7, 14, 17; and 418, Chap. 22).
expression for the correction factor is valid in LTE only. Similarly, in LTE
the source function becomes THE CLASSICAL OSCILLATOR

Sf = (2hv,//c2)[exp(hvlJ/kT) - 1]- 1 = B. (4-16) Consider first the electromagnetic radiation from a moving charge.
Suppose a particle has charge e, velocity v, and acceleration v. Then from
as expected from the Kirchhoff-Planck relation [equation (2-5)]. classical electromagnetic theory (331, Chap. 9; 343, Chap. 17; 494, Chap. 20)
Equation (4-14) contains implicitly the solution of the statistical equi- the electric and magnetic fields at a distant position r relative to the charge ··
librium equations (cf. Chapter 5), inasmuch as it makes reference to the ratio are found to be
of the populations of the levels involved. Only in the case of LTE can this E(r, t) = (eiJ/c 2 r-) sin 0 G (4-17)
ratio be expressed in terms of a single thermodynamic variable, T; in general
it will depend upon the temperature, density, and the radiation field (in all and H(r, t) = (eiJ/c 2 r) sin 0 cj, (4-18)
transitions of the atom). We shall therefore refer to equation (4-14) as an
implicit form of the source function. An alternative approach is explicitly where 0 denotes the angle between the acceleration v and f, 8 and cl, arb. ··
to introduce analytically the solution of the statistical equilibrium equations unit vectors in a spherical system defined by v and f, and i1 is evaluated at a ,·
82 Absorption Cross-Sections 4-2 The Calculation of Transition Probabilities 83

time t' = t - (r/c). The power radiated per cm 2 is given by the Poynting We now can calculate the scattering coefficient for a classical oscillator
vector [cf. equation (1-35)] in an imposed electromagnetic field. In the classical picture the interaction
is a ~onservative scattering process; hence we can compute the energy
S = (c/41t)(E x H) = (e2 1i2/41tc3r 2) sin 2 er (4-19) scattered out of a beam by calculating the energy radiated by an oscillator
driven by the electromagnetic field of the incident radiation. The equation
Now integrating over a sphere of radius r by forming S · dA where of motion for an oscillator of mass m and charge e driven by a field of
dA = (r2 dw)r = (r2 dµ d<J,)r and writing sin 2 0 = (1 - µ 2), we find the amplitude E 0 and frequency w is
total power radiated in all directions is
(4-27)

Taking a trial solution for x that is proportional to exp(iwt), we find for


the steady-state solution
In particular, for a harmonic oscillator we can write x(t) ~ x 0 cos wt,
v(t) = -wx 0 sin wt, and v(t) = -w 2 x 0 cos wt. Substituting in/.o equation x = Re [ (e/m)E 0 e "''
1
] (4-28)
(4-20) and averaging over one period of the oscillation by notin)? that (w 2
- w0 2
) + iyw
(cos 2 wt) = ½, we find
from which we derive
(4-21)
(4-29)
Because the oscillator is radiating away energy, the oscillation will even-
tually decay. We may describe the decay in terms of a damping force that
may be viewed as the force exerted on the moving particle by its own electro- Thus, substituting into equation (4-20) and averaging over a period, we have
magnetic field. To calculate the effective damping force, we assume that the
rate of work done by it accounts for the energy loss by the oscillator. Thus e4w4) Eo2
from equation (4•20) we write (P(w)) = (3m2c3 (w2 - w/)2 + y2w2 (4-30)
01
(4-22) which is to be identified with the total energy scattered out of the beam. To 0
Then calculate the energy scattered, we suppose that the scattering cross-section I
f, 12

1
(F,1 d • v) dt + (2e 2/3c 3) (I:: v· t - J,'. \I· v dt) = 0
2
(4-23)
is isotropic, and write /(µ, cf,) = /0 c5(µ - µ 0 ) c5(q, -. <J, 0 ); in §1-2 we found
that to produce a correspondence between the macroscopic and electro-
magnetic descriptions of radiation, we have / 0 = (cE//81t). Thus
Over a cycle the integrated term vanishes, therefore on the average

F,1d = (2e 2/3c 3)\I (4-24)


(P(w)) = o-(w) f IdQ = (!: f;" def, f~
u(w)
2
) 1
dµ c5(µ - µo) c5(<J, - <J>o)

To a good order of approximation we may calculate \I from its value for the
= (cE0 2/81t)o-(w) (4-31)
,undamped oscillator, namely \I = -w/v, and thus we can write Comparing equations (4-30) and (4-31) we see that the scattering coefficient is

F,1 d = -myv (4-25) (4-32)

where (4-26) Equation (4-32) can be simplified by noting that, because y « w for fre-
quencies corresponding to light, o-(w) is a sharply peaked function in the
The constant y is called the classical damping constant because of the formal neighborhood w ~ w 0 • To a good approximation we can replace
resemblance of the radiation reaction term as expressed in equation (4-25)
to a viscous damping term. (w 2 - w/) = (w + w 0 )(w - w0 ) ~ 2wo(w - Wo)
84 Absorption Cross-Sections 4-2 The Calculation of Transition Probabilities 85

and substitute equation (4-26) for y into equation (4-32) to find wave functions are solutions of Schriidinger's equation
(4-37)
a(w) =(:;)[cw - wo)~ + (y/2)~] (4-33)
where H is the total Hamiltonian of the system in operator form [see, e.g.,
(418, Chaps. 8-10) for a discussion of mathematical expressions for the
The total cross-section can be found by integrating equation (4-33) over all Hamiltonian]. The Hamiltonian operator is constructed from the classical
frequencies, namely
Hamiltonian according to the rule
a
101
=1te2
me
f"" (y/41t2) dv
Jo (v - v0 ) 2 + (y/41t) 2
=1te2 ! f""
me 1t - 00
~
1 + x2
=1te2
me
(4-34) (4-38)

where q, and p1 are space coordinates and momenta, respectively. The atom
where we have written x = 41t(v - v0 )/y and observed that, for all practical has certain stationary states (or eigenstates) in which its energy is constant.
purposes, -41tv 0 /y = - oo. The total cross-secttqn gives a measure of the For simplicity we shall assume that these states are nondegenerate. Thus if
efficiency with which energy is removed from the !,'earn. The factor in square HA is the Hamiltonian of an atom which is in a stationary state j of energy
brackets in equation (4-33) is thus a normalized profile function that is known E1, then
as the Lorentz profile (or damping profile). For our present purposes, attention (4-39)
will be confined to the total cross-section alone; the pro~nction will be which implies
discussed in detail in Chapter 9. v,j(t) = v,1(0) exp(-iE1t/h) (4-40)
The classical result derived above predicts a unique scattering efficiency
for all transitions; this is not surprising insofar as the theory makes no We may therefore write the general solution in the form
reference to the actual atomic structure or to the nature of the leVels between
which the transition occurs. The quantum mechanical treatment shows that l/tir, t) = cp1(r) exp(-iEif/h) (4-41)
cross-sections for different transitions may differ by orders of magnitude. where the time-independent solutions ¢1 satisfy the equation
A customary way of writing the quantum mechanical result for the total
01
cross-section is HA'PJ = E1</J1 (4-42)
(4-35) and are orthogonal so that

where / 11 is called the oscillator strength of the transition. In pictorial terms, . (4-43)
/iJ may be thought of as giving the "effective number" of classical oscillators
involved in the transition; only for the strongest lines does J11 approach A general state of the system at time t = 0 can be expanded in terms of the
unity. The oscillator strength is related to the Einstein coefficient B11 by the eigenstates (which form a complete set) by writing
expression
(4-36) l/t(0) = L a1cp1 (4-44)
J

where a = (cpjlv,(0)). For a system in a general state with wavefunction


QUANTUM MECHANICAL CALCULATION
t/,(0), th~ probability of finding the system in a specific state j ?Y a measure-
Let us now consider the calculation of B,1 when the atom is treated ment process is aja1 = ia1!2 • At an arbitrary time t, we can write the general
according to quantum mechanics and the radiation field according to clas- state as
sical electrodynamics. The atomic structure is described by a wave function
1/l(r 1 , r 2 , ... , rN, t) where ri, etc., are the positions of the bound electrons. v,(t) = L aj(t)l/ti(t) = L ai(t)cp1 exp(-iEif/h) (4-45)
j J
The quantity t/ll/t* dr 1 • • • drN is interpreted to be the probability of finding
the atom with the electrons in the volume element (r 1, r 1 + dri), etc. The and again the probability of finding the system in a specific state j is ja1(t)l2.
4-2 The Calculation of Transition Probabilities 87
86 Absorption Cross-Sections
The quantities dmn are called the dipole moment matrix elements. Substituting
If the atom is unperturbed (i.e., H = H,t), then the a/s are constant. If, equation (4-52) into equation (4-51) we have
however, the atom is perturbed by some potential V, then the a's will change
with time; this is interpreted as the atom undergoing transitions from one dm(t) = (ih)- 1 }: a.(t)hm•e'"'m•1(e'"'1 + e- 1"'1)
(4-53)
state to another. An example of such a perturbation is that exerted by an "
external electromagnetic field upon the atomic electrons. In the lowest order
We now make the simplifying assumption that, at time t = 0, the atom is
of approximation we can assume the atom is in a uniform, time-varying
in a definite eigenstate k, and we consider a time interval Tso short that this
electromagnetic field, E = (E0 cos wt)i. The assumption of uniformity is
state is not appreciably depopulated. That is, at t = 0, we assume ak(0) = 1
reasonable for light waves, which have wavelengths (A. ~ 10- 5 cm) that are
and a.(O) = 0 for all n ,t, k. Moreover, we choose Tsuch that ak(t) ~ 1 for all
large compared to a typical atomic dimension characterized by the Bohr
t ~ T. Then the sum in equation (4-53) may be replaced by a single term
radius(a 0 = 5 x 10- 9 cm). The potential of the atomic electrons in the field is
N
dm(t) = (ih)- 1 hmke'"'"'"1(e1"'1 + e- 1"'1) (4-54)
V =e L E · r, = E · d ""- (E 0 cos wt)(i · d) (4-46) Integrating equation (4-54) with respect to time we obtain
fs l

where d is the dipole moment of the atom. With a perturbing potential, a (t) = hmk {exp[i(wmk
ih
- w)t] - 1
(wmk - w)
+ exp[i(wmk + w)t]
(wmk + w)
- l} (4-
55
)
SchrMinger's equation becomes m

(H,t + V)tf, = ih(otf,/ot) (4-47) As we are interested in absorption processes we choose Em > Ek, so that
Wmk > 0. From the denominator of the first term in the braces, we see that
Substituting equation (4-45) for tf, we have the dominant contribution to am(t) will come when w :::: Wmk (i.e., radiation
near the line frequency is most effective in producing transitions). It is clear
(H,t + V) L an(t)t/ln(t) = ih L dnt/ln + ih L a.(ot/ln/ot} (4-48) that the second term can be neglected in comparison with the first. Then,
n n
writing x = (w - wmk), and forming laml 2 = a!am, we have
In view of equation (4-39), equation (4-48) reduces to

ih L dn'kn =Lan Vt/In


n n
(4-49) lam(t)l2 = 4h- 2 hm/x- 2 sin 2 G= xt) 1i- 2 E 0 2 li; dmkl 2 x- 2 sin 2 G xt) (4-56) (JI
I\)

We may isolate a particular coefficient d,,, by using the orthogonality of the Equation (4-56) gives the number of k - m transitions (per atom in the initial I
state k) produced in time t by radiation of frequency v = w/2n. To calculate
</>'s. Thus, multiply equation (4-49) by tf,! and integrate over all space. We
the total number of transitions, we must sum over all frequencies that can
then have
contribute. Suppose that the line has a profile</>, that falls sharply to zero over
ifi La. exp[i(Em - En)t/h](</>!l</>n) = L an(t) exp[i(Em - En)t/h](</>!IVl</>n) some characteristic frequency interval 6v, and that over this range (at least)
" " the intensity of radiation (and hence E0 2 ) is constant with a value J,. Then
(4-50) integrate over dv = dw/2n = dx/2n, and define u = ½xt, to obtain

Now writing Wmn =


(Em - En)/h and Vmn
(4-43), equation (4-50) reduces to
=(</>!IVl</>n), and using equation (4-57)

Now for thermal radiation, a characteristic frequency interval 6w over which


dm(t) = (ih)- 1 L an(tWmne 1'°"'" 1 (4-51) the intensity will be constant is of the order (kT/h) ~ 10 15 , while transition
times t ~ 10-s sec; hence the limits ± U on the integral may be extended
n

For the perturbing potential given by equation (4-46), we see that formally to ± oo. The value of the integral is then found in standard tables
to be n. Further, as was shown in §1-2, ER = (4nJ,/c) = (E//8:,;), therefore
Vmn = (Eo COS wt)i' (</>!ldl</>n) = (Eo COS wt)(i · dmn)
%km = (81t 2/h 2 c)li · dmkj 2 J,t (4-58)
= 2hmn cos wt = hmn(e + 1
"'' e- 1"'1) (4-52)
90 Absorption Cross-Sections 4-2 The Calculation of Transition Probabilities 91

which follows from the fact that the allowed values of I are O ~ I ~ n - 1; TRANSITION PROBABILITIES FOR LIGHT ELEMENTS
thoseofmare -1 ~ m ~ l;andeachnlmstatehastwopossiblespinorienta-
(a) H artree-Fock Met hod. When the atom has more than one electron,
tions s = ±½,
the wave equation can no longer be solved in closed form, and approxi-
Exercise 4-2: Derive equation (4-74). mations must be made. The actual Hamiltonian for an N-electron atom is
N N
Because the wave functions are known analytically, explicit expressions H = -(h 2/2m) I V,2 - I (Ze 2/r1) + I (e 2/lr, - r11) (4-80)
can be derived for the oscillator strengths: 1=1 1=1 all pain
(I, Jl

, I'
f( n, ; n, 1) =3
1( 1
n' 2 -
1)
n2 (ii'
max([, I') 2 , ,
+ l) q (n, I ; n, 1) (4-75) The first term represents the kinetic energy of the electrons, the second their
electrostatic potential with the nucleus of charge Z, and the third their mutual
and Coulomb repulsion. It is the last term that causes the principal difficulties.
J(n', n) =3
1
n
,2 (~ -
n
-;)
n
["'r,
1·-1
1
l'<J 2(n', /'; n, I' - I) One of the most important methods of deriving approximate wave func-
tions is Hartree's self-consistent field method. In this approach, the sum over

+ "'f (I' + l}<J


1'•0
2
(n', I'; n, I' + l)] (4-76)
electron pairs is replaced for each electron by its spherical average. An
excellent description of how this average is computed is given in (576,
Chaps. 3 and 9). Each electron then moves in a potential that depends only

where q
2
(n', I'; n, I) = (f 00

0
P•.,.(r)P.1(r)r drr (4-77)
upon its distance from the nucleus, and we make the replacement

L (e2/lr, - r1I) ... L l-;(r,) (4-81)


all pairs I
An explicit expression for <J~~as first derived by Gordon (254) and an explicit (i, J)
form for f(n', n) was deriveCNll,.Menzel and Pekeris (417). Extensive tabfes
of f(n', n) can be found in (417) and (257). A very convenient form for hydro- This results in the approximation of the actual potential by a central field.
gen oscillator strengths is obtained by expressing them in terms of the semi- With a central field potential, the angular fac~ors in the SchrMinger equation
classical value derived by Kramers (363), namely can be separated out in exactly the same way as for hydrogen, and for each 01
electron the wave function has the form ~

fK(n', n) = 3: ~ ()2 - :2 rl (n3~,5) (4-78)


I

which shows the principal dependences of J upon n' and n. It is then cus- where the normalizations given in equation (4-72) still apply. The functions
tomary to express the exact f-value in terms of Kramers' approximation U, are called electron orbitals. The radial equation for each orbital is of the
/,: by writing form (r in units of a0 , E in Rydbergs)
f(n', n) = g1(n', n)JK(n', n) (4-79)

where o,(n', 11) is called the Gaunt factor. The Gaunt factors are all numbers
of order unity; an extensive tabulation of g1(n', n) can be found in (60). Here Z,rr(r) is the "effective nuclear charge" sensed by an electron after
allowance is made for shielding by other electrons [ using the central fields
Exercise 4-3: Using the analytical expressions for hydrogen wave functions of equation (4-81)]. The atom is now considered to be made up of N such
given in texts on quantum mechanics [e.g., (392, 183)], calculate the f-values for orbitals, and these are used to construct the wave function for the entire
L-:x.(n' = 1 .... n = 2) and Ha(n' = 2 .... n = 3). Obtain values for each f(n', /'; n, I) configuration.
and combine these to find f(n', n). Compare your values with those given in tables Because of the Pauli exclusion principle, the set of four quantum numbers
( e.g.. (9, 70)]. (n, I, m, s) for each orbital cannot be identical for any two orbitals. Also, the
92 Absorption Cross-Sections 4-2 The Calculation of Transition Probabilities 93

wave function of the atom must be constructed so that it is antisymmetric where ~max = max(/, /'). The factor fl'(.,.({) is the strength of the multiplet,
under the interchange of the coordinates of any two electrons. In practice depending on nLS and n'L'S', and the factor Y'(!l') is the strength of the line
these conditions may be met by writing the wave function as a Slater deter- within the multiplet. Extensive tables of Y'(A) and Y'(!l') can be found in
minant (576, Chap. 12), (11, Chap. 8 and Appendix; 9, §§26-28; 250; 251), and general formulae for
computing these faclo:s are given in (534; 535; 572, §§ 10.8-10.10; and 191).
Generally, the most difficult part of the calculation is the determination of
o-2, but serious complications also occur when there are deviations from
(4-84) L-S coupling.
(b) The Coulomb Approximation. Because of the labor involved in
obtaining o- 2 from Hartree-Fock calculations, it i's desirable to have an
approximate method that can be applied easily. Such an approach was
where the numbers 1, 2, •.. , N denote the orbitals of electrons 1, 2, etc.,
developed by Bates and Damgaard (74), who pointed out that often the
while tx, {1, ..• , v stand for the space and spin coordinates of electrons
largest contribution to th~ radial integral comes from large values of r,
:x, ... , v, respectively.
:,vhere the electron moves m a nearly Coulomb potential. In this event, the
The solution for the wave function is carried out iteratively. Thus z.,r(r)
mtegr~I ~an be approximated using hydrogenic wave functions, provided
depends in an involved way on integrals over the electron orbitals, but in
the prmctpal quantum numbers are chosen to give the observed energy of
turn it determines those orbitals. Therefore, we start with an approximate
set of orbitals, compute Z,rr, solve for the P.,•s, recompute Zerr, and iterate
!n
the le_veL If Z is the charge the asymptotic potential, then the appropriate
effective quantum number ts nt = Z/e.,t where e.01 is the energy of the level
until the procedure converges. The calculations are time-consuming and
below the continuum, measured in Rydbergs; in general, nt is not an integer.
laborious, but are within the capabilities of modern computers, and a large Bates and Damgaard then show that one cari write
number of wave functions for a wide variety of atomic configurations are
now available. o-(nt-1, I - 1; nt, I)= F(nt, /)J(nt-i, nt, 1)/Z (4-87)
A s~cific term in an atomic spectr1,1m can be characterized by certain
quant numbers describing the atom as a whole. In light atoms these Exten~ive tables of the functions SF and J can be found in (74) and (483); an
describe he total orbital angular momentum L (the vector sum of the in- extension of the theory is given in (389). Because of the simplicity of the
dividual 1/s), the total spin angular momentum S (the vector sum of the s/s), metho~. it has been widely applied in astrophysical analyses; an extensive
01
and the total angular momentum J, which is the vector sum of L and S. tabulation of Coulomb approximation /-values is given in (264, 363-441). 01
This type of coupling of the individual momenta is called (L-S) or Russell- (c) Experimental Methods. In many cases the Coulomb approximation I
Sa1111c/ers coupling. As a given L, S, and J may result from more than one is inaccurate, while a more accurate quantum. mechanical calculation is
arrangement of the individual l's, m's, and s's of the orbitals, the complete simply too complicated to carry out. In these cases the fvalues must be
wave function will, in general, consist of a sum of Slater determinants, and determined by experiments, which in addition, provide a direct comparison
thus may be very complicated. standard to test the accuracy of various theoretical computations. A wide
In calculating transition probabilities, it is generally assumed that only variety of experimental techniques exist; brief descriptions of some of the
one orbital is different between the initial and final state-i.e., only one more useful methods are given in (11, 300-310; 261, 146-149;264, Chap. 15).
electron undergoes a transition. In this case the matrix element r 11 can be There is a huge literature containing both experimental and theoretical
split into factors, one coming from the initial and final radial wave functions, determinations of fvalues or transition probabilities; a complete biblio-
and another depending on the angular and spin wave functions. It is cus- graphy of this work is given in (454; 230; and 231). Compilations of critically
tomary, therefore, to write the expression for the line strength in the form evaluated (i.e., "best values") transition probabilities for many of the elements
of astrophysical interest are given in (453; 584; 670; 672; and 673).
S(n', L', S', J'; n, L, S, J) = a 0 2e 2o- 2(n', l'; n, /)Y,(.,.({)9'(!l') (4-85)
Exercise 4-4: Calculate f-values for the He I l5876 (2p 3 P-3d 3 D), He I l6678
1
(2p P-3d I D), and He I A.4471 (2p 3 P-4d 3 D) lines using equations (4-66) and
Here (4-8S) with Coulomb approximation values of o- 2 and multiplet- and line-strengths
from tables. Compare with standard values in (672).
94 Absorption Cross-Sections
4-3 The Einstein-Milne Relations/or the Continuum 95

4-3 The Einstein-Milne Relations city distribution [ cf. equation (5-2)]


for the Continuum
11,(v) dv = n,(m/2nkT)1 exp(-mv 2/2kT)4nv 2 dv (4-91)
The Einstein relations were generalized to bound-free processes by Milne
(461) in a paper of considerable interest and importance. We consider photo- Anticipating the results of Chapter 5 [cf. equation (5-14)], the T.E. relation
ionization processes that start with atoms (or ions) in a definite bound state between n~ and nT is
(not necessarily the ground state) and produce an ion in a definite state
(perhaps excited) of the next highest ionization stage plus a free electron
moving with velocity v. The inverse process is a recombination of a free
Using equations (4-88), (4-91) and (4-92) we find that
electron by a collision with an ion (in the particular state mentioned above)
to form an atom (in the proper state). The recombination process can occur n~p.m/nT n,(v)hG(v) = (h 2 g0 /8nm 2g 1 v2 )[p,/G(v)]eh•JkT (4-93)
spontaneously or can be induced by incident radiation. Let n0 be the number
density of the atoms, n 1 the density of the ions, and ne the density of free Thus to reduce equation (4-90) to the Planck function we must have
electrons. The electrons have a Maxwellian velocity distribution, and we
write n,(v) dv for the number with speeds on the range (v, v--+ dv). Let p. F(v) = (2hv 3 /c 2 )G(v) (4-94)
be the probability of photoionization of an atom by a photon in the frequency
range (v, v + dv); then the number of photoionizations in time dt on this and P, = (8nm 2 v2 gi/h 2g0 )G(v) = (4nc 2 m2 v2 g 1/h 3g0 v3)F(v) (4-95)
frequency range is n0 p,1, dv dt. The usual energy absorption coefficient IX, is These are the continuum analogues of equations (4-8) and (4-9). Again we
related top, by the expression IX, = p,hv. Further, let F(v) be the spontaneous recognize that, although these relations have bee·n derived from thermo-
recapture probability and G(v) be the induced recapture probability for dynamic equilibrium arguments, the quantities p., F(v), and C(L') must really
electrons in the velocity range (v, v + dv) by the ions; then the number.of depend only on the properties of the atom; hence equations (4-94) and (4-95)
recombinations by electrons of velocity v in time dt is are true in general.
The great importance of the results just derived becomes more clearly
n1 ne(v)[F(v) + G(v)l,]v dv dt manifest when we write the transfer equation assuming that at the frequency
under consideration only the particular photoionization and recombination CJl
The energy of the photon required to ionize the atom (and thus of the photon CJ)
emitted in the recombination process) is processes considered above occur. The generalization to a multilevel, multi-
atom case with several overlapping opacities and emissivities is trivial be- I
1 cause each term adds linearly and the conclusions we shall derive apply to
hv = x, + -mv
2
1 (4-88)
the sum. The transfer equation is

where x., is the ionization potential from the atomic to the ionic state (i.e.,
the energy difference between these states).
Now in thermodynamic equilibrium, the number of photoionizations must Neither n0 nor 11 1 necessarily has its LTE value in the above equation. Ifwe
exactly equal the number of recombinations. In T.E., I, = B., hence are to write the transfer equation in standard form, then it is clear that the
absorption coefficient corrected for stimulated emission must be
n~p.B, = ntn.(v)[F(v) + G(v)B,](h/m) (4-89)
(4-97)
where the asterisks denote T.E. occupation numbers, and equation (4-88)
has been used to write h dv = mv dv. Solving for B, we find Using equations (4-88), (4-91), (4-92) and (4-95), and recalling that IX, = p,hv,
we find
B, = [F(v)/G(v)]{[n~p.m/ntne(v)hG(v)] - q- 1 (4-90) (4-98)
This expression is to be compared with the standard expression for the Planck In equation (4-98), nt denotes the LTE value of 11 0 computed from equation
function, namely B,(T) = (2hv 3/c 1)[e'"'tT - 1J- 1• For a Maxwellian velo- (4-92) using the actual values of 11 1 and lie (i.e., the LTE population relative
96 Ah.l'Orption Cross-Sections 4-4 Continuum Absorption Cross-Sections 97

to the actual ion density). In the particular case or LTE where 11 0 = 116, bound transitions. Consider absorptions fro:n a bound state n, of statistical
weight g., to the continuum in a frequency interval av. The free states have
(4-99) wave functions characterized by E, the energy of the free electron, and are
normalized such that
As was true for bound-bound transitions, the term (1 - e-hv/kT) is usually
(E'IE> = b(E' - E) (4-102)
called the correction factor for stimulated emission; but it is clear that this
expression is correct only for LTE. Indeed we see from equation (4-98) that so that there are t.E states in the energy interval t.E. Thus by analogy with
the stimulated emission always occurs at the LTE rate (ff we understand n6 equation (4-65) and in view of (4-36) we can write
to have the meaning given above); this must be true because the recombina-
tion process is a collisional process involving particles with an equilibrium g.rxv 6v = (81t 2/3h 2 c) t.E(EJdln) 2 (hv/41t) (4-103)
(i.e., Maxwellian) velocity distribution. Note the contrast here with the result
given in equation (4-13) for bound-bound transitions, where in general the or IX, = (81t 3v/3cg.)(EJdln) 2 (4-104)
stimulated emission term does not have its equilibrium value. When depart-
ures from LTE affect the bound-free opacity, they change the direct absorp- The calculation of free-state wave functions will not be considered in this
tion term involving n0 (which in general will not equal n6). We shall use book; the reader should refer to standard texts on quantum mechanics-
these results both in calculating the stimulated emission rates in the equa- e.g., (197) or (418)-for information on this subject. Further, we shall not
tions of statistical equilibrium [ cf. equation (5-63)] and in writing a general consider the details of calculations based on equation (4-104), though results
expression for the opacity [cf. equation (7-1)]. of such work will be quoted. An approximate method for the evaluation of
Returning to equation (4-96) and examining the term involving F(v), it is IX, by means of equation (4-104), the quantum defect method, is described
clear that the emissivity is .,below.
An alternative formula for IX, can be written if we suppose each continuum
(4-100) state k to have an effective oscillator strength J.k for absorptions from the
bound state n. If there are t.k free states in the frequency interval t.v, then
which, with the help of equations (4-88), (4-91), (4-92), and (4-95); can be
written (4-105)
Y/v = (2hv 3/c 2)ntrx,e-h•fkT = ntrx,(1 - e-hv/kT)B, = ,c:B,(T) (4-101)
This formulation is useful for calculating the cross-sections of hydrogen. 01
Thus the continuum emissivity always occurs at the LTE rate (if n6 is de-
fined as above), which is what we would expect, for the recombination process
is collisional. Notice that this derivation recovers the Kirchhoff-Planck law,
The quantum defect method, developed by Seaton and Burgess (566; 120),
is the continuum analogue of the Coulomb approximation. This method
ex·ptoits the fact that the dominant contribution to the matrix element
"'
I

equation (2-6), and extends its validity somewhat. Again, notice the contrast (EJdJ11) 2 orten occurs in regions where the ~ave functions can be represented
with the bound-bound spontaneous emission where departures from LTE by Coulomb wave functions in the appropriate potential. Consider absorp-
enter directly if n1 is not identical to nf. These results will be exploited in tions from a bound state (Ii, /) to the continua (E, 1 ± 1) where Eis the energy
calculating spontaneous emission rates in the statistical equilibrium eq ua- of the free electron. Let / "' be the ionization energy of this state, expressed' in
tions [cf. equation (5-61)] and in writing a general expression for the emis- Rydbergs, and let Z be the charge on the ion after the electron is removed.
sivity [cf. equation (7-2)]. Then define the effective quantum number v.1 such that 1.1 = Z 2 /v.,2. In
general, v., will not be equal to the principal quantum number n of the shell
Exercise 4-5: Verify equations (4-93), (4-98), and (4-101). to which the electron belongs, and we can define a quantum defect µ(v, I) =.
n - v. 1• The quantum defect can be found for each level (n/SL) of a giver\.
spectroscopic type, defined by (ISL) (e.g., 3 P or 4 D), in a series. Defining
4-4 Continuum Absorption Cross-Sections e,, 1 = -1/v.,2, we can determine the behavior of µ(e.,, /) versus e. 1; in favor-
able casesµ is a simple function of e (say, constant, or linear in e). It is then
Cross-sections for bound-free absorption can be calculated quantum assumed that this variation of µ with e can be extrapolated into the con-
mechanically by essentially the same methods as used in §4-2 for bound- tinuum (i.e., for e > 0) to give µ'(e). This establishes the properties of fhe
98 Absorption Cross-Sections 4-4 Continuum Absorp1ion Cross-Seclions 99

continuum wave functions. The radial matrix element can then be evaluated The formula for the continuum oscillator strength follows from a general-
using hydrogenic wave functions, and the cross-section when the energy of ization of equations (4-78) and (4-79) to
the ejected electron is k 2 = Z 2e (in Rydbergs) can be written

et.(nl, k2) = 8.56 x 10- 19 [(]., + k2)//.,2] L C,,lg(vl; e/')12 cm 2 (4-106) fn'k = ( 37t32J3)(n'5k3
1 )( 1
n'2 + l
k2
)-J ,
gu(n, k)
(4-111)
l'sl± I

Here g(vl; el')= [C(v, /)J-tG(vl; el') cos{1t[v + µ'(e) + x(vl; el')]} (4-107) where g11 is the bound-free Gaunt factor. Formulae for the Gaunt factor
are given in (417) and an extensive numerical tabulation is given in (35~);
C(v, I) = 1 + 2v- 3[oµ(e)/oe] (4-108) q is a number of order unity at the ionization threshold, shows a slow nse
t~ about 1.10 (in the limit as n' -+ oo) at about 1 Rydberg above threshold,
and G(vl; el') and x(vl; el) are tabulated functions (503) [Norn: the notation and then decreases to small values in the X-ray region. The absorption
in the reference cited differs from that in (120), which is used here]. The cross-section can now be derived by substituting equation (4-111) into
coefficients C,. are algebraic factors obtained from the integrations over (4-105), noting from equation (4-110) that for n' fixed,
angular and spin coordinates and are tabulated in (120) for several important
cases; they are analogues of the factor 9'(.lf)9'(.5e) appearing in equation (dk/dv) = -(hk 3/29t) (4-112)
(4-85) for bound-bound oscillator strengths. We then find
The quantum defect method, despite its simplicity, often gives very good
cross-sections [see (120; 503)] and has been widely employed in astrophysical (4-113)
work. A number of quantum defects µ(e) are given in (503) where they are
used to calculate cross-sections and opacities for abundant elements in
stellar atmospheres [see also (502)]. For brevity, only absorption by hydro- which, in view of equation (4-69), reduces to
gen and helium, and their ions, will be discussed in this chi\pter; these are
the most abundant elements in stellar material, and usually dominate the 647t 4 me'0 ) 1 , gu(n', v) (4-114)
opacity. References to other opacity sources will be given in §7-2.
et.. = ( 3}3 ch6 n'5v3 gu(n' v) = .1( ~
where .1( = 2.815 x 10 29 • Thus bound-free absorption from level n com- 0,
HYDRC?GEN mences abruptly at the threshold frequency v. = (9t/hn 2 ) and falls off at (X)

A simple way of obtaining bound-free and free-free absorption cross- higher frequencies as v- 3 (neglecting the weak variation of the Gaunt factor). I
sections for hydrogen was suggested by Menzel and Pekeris (417). They The threshold cross-section is given by
introduced the formalism of representing bound states by real (integer) 2
quantum numbers, and free states by imaginary quantum numbers. The
et.(v., n) = 7.91 x 10- 19 ng11 (n, v.) cm
'
bound-state energies relative to the continuum are given by equation (4-68), The opacity per cm 3 of the stellar material can be computed by multiptt~g
and it follows that the energy of the transition (n' -+ n) is the cross-section for level n by the number of hydrogen atoms (per cm ) m
that level, and summing over all levels that can absorb at a given frequency v
(4-109) (i.e., all n such that v. ~ v). The bound-free opacity of hydrogen calculated
If a free state has the imaginary quantum number ik, then by analogy in this way has a jagged character, as shown in Figure 4-1. Except for ~he
hottest stars, most of the hydrogen is in the ground state, and the absorption
edge at l912 A(one Rydberg) is extremely strong. For 912 A ~ l ~ 36~7 A,
(4-110) absorptions from the ground state can no longer occur, and the d~mmant
opacity source is photoionization from the n = 2 level (Balmer continuum).
where the first term clearly represents the ionization potential from bound Similarly, for 3647 A ~ l ~ 8206 A, both n =
1 and n = 2 cannot absorb,
state n', and the second the energy of the free electron. Note that k -+ oo at and the dominant continuum is from n = 3 (Paschen continuum); and so~~-
the ionization limit and becomes small high in the continuum. Actually the opacity variation shown in Figure 4-1 is idealized in that there
100 4-4 Continuum Absorption Cross-Sections IOI

-17 and (4-116)

Assume that absorptions take place from a band of states dk into a band of
states di = (d//dv) 6v; then in equation (4-105) we replace fnk with 1,,.1 dk, and
-18 6k with di to obtain

o:(v, v) = (ne 2 /mc)!,,.1 dk(dl/dv) (4-117)


T - 2s,000°K
n =I as the absorption coefficient per ion and per electron moving with velocity v.
-19
The appropriate generalization of equations (4-78) and (4-79) is
log a,
J; = ~ _!_ (..!_ _..!_)- 3
g111(k, I) (4-118)
-20
kl 3njj gk k2 12 k3/3

where Yk is the statistical weight of a free electron, given by quantum statistics


as
(4-119)
-21
the second equality following from equation (4-115). Substituting into
equation (4-117) we find
-22'--__.l---l---'--.....L--L.:::.1-....1--,--'
0 2 4 6 8 10 12 ' 14 o:(v, v) = (ne
me
2
)(~)(
3nj3
h3k
3
16n9lm 2 v
)(ge)hv
3
gt.(v, (di) (4_120)
v)
k3 / 3 dv
1/l
FIGURE 4-1 and making use of the relation that fork (or v) fixed, (dl/dv) = (h/ 3/291) from
Opacity from neutral hydrogen at T = 12,SOO'K and equation (4-116), we obtain
T .. 25,000'K, in LTE; photoionization edges are labeled 01
with the quantum number or state from which they arise.
o:(v, v) =( 29lhe (g v v v))
2
) 111 ~,
(+121)
co
Ordinate: sum or bound-free and free-free opacity in cm 2/atom; 3 I
abscissa: 1/i. where i. is in microns. 3nJj m c
The total absorption cross-section per ion and per electron is obtained by
exists a series of lines converging on each photoionization threshold at the summing over all incident electron velocities, assuming a Maxwellian velocity
series limit. Near the limit the lines blend together smoothly and merge into distribution as given by equation (4-91); The result is
the continuum. Bound-free absorption from hydrogen is the dominant
continuum opacity source in stars of spectral types A and B.
(4-122)
Let us now consider the free-free opacity of hydrogen. In this process, a
free electron passing near a proton causes a transitory dipole moment, and
absorptions and emissions of photons (with a consequent change in the where use has been made of equation (4-69), and g111 is the thermal average
electron's energy) become possible. By analogy with the calculation of of the Gaunt factor
bound-free absorption, we introduce imaginary quantum numbers for both
the initial and final states, say ik and ii, such that, if v is the initial velocity (4-123)
of the free electron, and v is the frequency of the radiation absorbed, then
where u = (mv 2/2kT).
9lk- 2 = !2 mv 2 (4-115)
Exercise 4-6: Verify equations (4-122) and (4-123).
103
102 Absorption Cross-Sections
5
Inserting numerical values for the atomic constants, multiplying by the
electron and proton densities, and correcting for stimulated emission (notice
that because the process is collisional it is always in LTE, at the actual 4
electron and ion densities), we obtain the opacity coefficient "'
8
X
:::-- 3
I

Formulae for g111 are given in (417) and extensive tables can be found in (85) ">.
C

and (352). The free-free opacity plays an ever more important role at low ."'
frequencies compared to the bound-free, because of the decreasing number I
I
2

of photoionization edges that contribute as v -+ 0. Further, the free-free ::r:


~
becomes more important at high temperatures, for as can be seen from "
equation (4-92), in the limit (kTfx 10n) » I, the bound state populations vary
as 111 oc 11,11pr-t; hence the ratio of free-free to bound-free opacity rises
'l.. Tin the high-temperature limit. The free-free process is the dominant true
absorption mechanism in, e.g., the O-stars. 0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
A(A)/1000
THE NEGATIVE HYDROGEN ION FIGURE 4-2
Bound-free and free-free opacity from H" at T = 6300°K. Ordinate:
Hydrogen, because of its large polarizability, can form a negative ion cross-section ( x 1026 ) per neutral H atom and per unit electron
consisting of a proton and two electrons. This ion has a ~ingle bound state pressure p, = n,kT; abscissa: A/1000 where). is in A.
with a binding energy of0.754 eV. Because of its low binding energy, H- does
not exist at high temperatures (it is destroyed by ionization) but is prevalent bound free near 15000 A (1.5 µ) and increases towards longer wavelengths.
mainly in the atmospheres of solar-type and cooler stars. It was recognited The summed absorption coefficient (see Figure 4-2) has a minim~n:i at about CJ)
by Pannekoek and Wildt that H- could be an important opacity source in 1.6 µ; although other absorption processes act to wash out the mm1mum, the 0
such stars. As it turns out, the abosprtion cross-section of H- is large and, opacity for cool stars is smallest near this wavelength. . I
although only a small fraction of the hydrogen exists in this form, the opacity The determination of cross-sections for the two processes ment10ned above
from H - is the dominant one in the atmospheres of cooler stars. is difficult and has been attempted both theoretically and experimentally.
The negative hydrogen ion can absorb and emit radiation via both bound- Very elaborate wave functions are required to give the desire~ accuracy.
free and free-free processes; i.e., Pioneer calculations that gave fairly accurate values were carried out by
(4-125) Chandrasekhar and Breen (162). These were shown to be in accord with·
empirically deduced values for the absorptio_n coeffic!ent in the ~un, and led
where ¼mu 2 .. hv - 0,7S4 eV, and to the firm identification of H - as the maJor opacity source m the solar
atmo~phere (see §~-6). More aeciurat11 vah!<lll uo now av1dlablo ror both the
H + e(v) + hv ;: H + e(v') (4-126) bound-free (242) and free-free (604) cross-sections; these are in good agree-
ment with experimental values.
where ½mv' 2 = ½mv 2 + hv. In the free-free process, an electron passing near In LTE the number of H- ions per cm 3 is given by a Saha formula [ see
to a neutral hydrogen atom induces, by polari~tion, a temporary dipole equation (5-14)] that is of the form n*(H-) = n(H)p,<I>(T) wher~ n(H) is the
moment that can interact with the radiation field, leading to absorptions and number of hydrogen atoms per cm3, p, = n,kT, and <I>(T) contai~s the tem-
emissions. The bound-free absorption process has its threshold at about perature dependence of the ionization equilibrium. The LTE opacity can thus
16500 A (1.65 µ), corresponding to the detachment energy. It reaches a be written K:(H-) = cx,(H-)n(H)p,<I>(T)(l - e-hv/kT); de~artures fro~ L:fE;
maximum cross-section of about 4 x 10- 17 cm 2 at 8500 A and decreases may enter both in the calculation of n(H-) and in the stimulated em1ss10n
toward shorter wavelengths. The free-free cross-section is about equal to the
104 Absorption Cross-Sections 4-5 Co111i111111111 Sca//ering Cross-Sections

correction factor. Because K•(H -) is proportional to p., it is clear that it will is appreciably ionized before the excited states contribute to the opacity
be a more important opacity source in dwarfs than in giants. Also, because significantly. In a few stars the helium to hydrogen ratio is anomalous, and
the electron density in G-type and cooler stars depends upon the abundance approaches or exceeds unity; here helium can dominate the opacity.
of the metals, H- will be a much weaker opacity source in Population II stars Helium is a three-body system, and exact wave functions cannot be
(which have low heavy-element abundances). obtained. A number of special methods can be applied to obtain accurate
approximate wave functions [ see (87, §§24-32; 577, §§18.l - 18.3)]; varia-
OTHER IONS OF HYDROGEN
tional techniques applied to the ground state have been refined to the point
of yielding very precise wave functions. The ground-state absorption coeffi-
Hydrogen exists in two other forms that can contribute significantly to cient calculated from an accurate Hartree-Fock wave function is given in
the opacity in stellar atmospheres, namely H 2+ and H 2- . The positive ion (603); this agrees well with experimental values (311). Absorption cross-
H2 + consists of a single electron shared by two protons; absorption cross- sections from the 2 3 S, 2 3 P, 2 1S, 2 1P levels have been calculated using accurate
sections are given in (72) and (117). As the number density of H 2+ is propor- variational bound-state wave functions and close-coupling free-state wave
tional to n(H)np, H 2 + contributes significantly to the total opacity only for functions (332). For higher excited states, precise cross-sections have not
the temperature-pressure range where both neutral and ionized H-atoms been published, and here one may use the quantum defect method.
exist simultaneously in appreciable numbers; i.e., where the hydrogen is Ionized helium is a hydrogenic ion with Z = 2. As energies in such ions
about half-ionized. This range is characteristic of the A-stars, and H 2+ makes scale as Z 2, the frequencies of the ionization edges, "•• are larger by a factor
about a 10 percent contribution to the opacity in the visible part of the of four, and the ground-state edge occurs at )..227 A. This edge dominates the
spectrum (the H2 + absorption peak at ).1100 A is swamped by the Balmer far ultraviolet spectrum of O-stars, except at the very highest temperatures
continuum of hydrogen). where the helium becomes doubly ionized. The n = 2 edge of He II coin-
The negative molecular ion H 2 - exists only at relatively low temperatures, cides with the hydrogen Lyman continuum; higher edges from states with
characteristic of the M-stars, and its free-free continuum makes a significant even quantum numbers coincide with hydrogen edges from states with
contribution at long wavelengths (the bound-free process is negligible). In n = n(He Il)/2, while those from states with odd quantum numbers fall
this process, an electron passing near an H 2 molecule temporarily induces a between the hydrogen edges.
dipole moment by polarization effects, and this moment can interact with the Hydrogenic cross-sections can be used for He II, but the bound-free cross
radiation field. The H 2 - continuum tends to fill in the opacity minimum of sections are a factor of Z 4 larger, and the free-free are a factor of Z 2 larger. CJ)
H- at 1.6µ. The free-free cross-section is given in (592). The hydrogenic Gaunt factors apply if one evaluates them as functions of
(v/v.). He II affects the visible spectrum only in stars of types BO and hotter.
Finally, helium can give rise to a free-free opacity in cool stars. Cross-
HELIUM
sections for this process are given in (593; 340).
Helium is observed in stellar spectra in both its neutral and singly- Exercise 4-7: Calculate the photoionization cross-sections of Hel from the four
ionized states. Because the ionization potential of neutral helium is 24.58 eV n= 2 states by the quantum defect method and compare these results with t_he
it persists to temperatures characteristic of the B-stars, where hydrogen i~ more accurate values cited above.
already strongly ionized; in the O-stars, He II becomes a major opacity
source. The threshold for abosprtion from the ground state of He I is at
J._504 A; the ultraviolet spectrum for ,t < 504 Ais dominated by He I absorp- 4-5 Continuum Scattering Cross-Sections
tion for stars of types BO and cooler. The excited states of helium fall into two
groups, singlets and triplets, and each (n, /, s) state has a different ionization As mentioned in Chapter 2, continuum radiation may be scattered as well as
energy. Roughly speaking, the ionization energies lie close to tpe hydrogenic absorbed. In the latter case, photons are destroyed, and their energy con-
value at the same n; thus helium contributes several absorption edges near tributed at least partially to the thermal content of the gas. In a scattering
each hydrogen edge (for n ;;i, 2). Because the excitation energy of even the event, the photon is not destroyed, but merely redistributed in a_ngle, and
lowest excited state is so large (19.72 eV), helium adds to the opacity in the perhaps shifted slightly in frequency. Cross-sections for the two most impor-
visible regions of stellar spectra only in hot (B-type) stars. Generally, helium tant scattering processes in stellar atmospheres are derived in this section:
106 Absorption Cross-Sections 4-5 Continuum Scattering Cross-Sections 107

THOMSON SCATTERING
Rayleigh scattering can be important in the atmospheres of stars of
The scattering of light by free electrons is referred to as Thomson scat· moderate temperature (spectral types G and K). Here most of the hydrogen
tering. The classical formula for this process can be obtained directly from is neutral and in the ground state. The resonant frequencies corresponding
equation (4-32) by noting that for an unbound electron both the resonant to the Lyman lines (I -+ n) lie far in the ultraviolet, and visible photons
frequency w 0 and the damping parameter are zero. Thus we find interact with these transitions by the Rayleigh scattering mechanism. The
macroscopic scattering coefficient is obtained by summing over all lines and
u, = (81te4 /3m 2c4) = 6.65 x 10- 25 cm 2 (4-127) multiplying by the density of hydrogen in the ground state. Rayleigh scat-
tering from neutral hydrogen can dominate the total opacity at relatively
Note that this cross-section is independent of frequency. The Thomson low temperatures and high frequencies; see the graphs in (97) and (651 ).
cross-section has been verified by quantum mechanical calculations in the Moreover, in stars with low metal abundances (Population II stars), the
limit of tow photon energies; i.e., hv « mc 2 • At high photon energies (,l,. < number of free electrons (coming mainly from the metals) is greatly reduced;
I A, in the X-ray region) one must employ the Klein-Nishina formula accordingly the opacity from H- is diminished, and thus the importance of
(293, §22; 392, 433), which predicts a smaller cross-section; in practice the Rayleigh scattering is much enhanced. Molecular hydrogen, H 2 , may also
departure of u from u, can be ignored in stellar atmospheres work (except scatter radiation in an analogous fashion. The cross-section per molecule
for X-ray binaries). · (189) is comparable to the cross-section per atom for atomic hydrogen. At
In the derivation of equation (4-32) an angle averaging was performed, low temperatures (e.g., in M-stars), H 2 is much more prevalent than atomic H,
and thus the angular dependence of the scattering coefficient was suppressed; and thus molecular Rayleigh scattering dominates.
the correct angular dependence is given by the dipole phase function in It should be noted that in a continuum scattering process there is no
equation (2-19). In stellar atmospheres applications, this angular dependence analogue of the stimulated emission that occurs in absorption processes.
can almost always be ignored and the process considered to be isotropic. Thus in the macroscopic scattering coefficient there is no correction factor
Frequency redistribution in the laboratory frame caused by Doppler shifts for stimulated emission such as appears in equations (4-98) and (4-99).
from the electrons' motions have also been neglected; these will be considered
in Chapter 13 (see Exercises 13-5 and 13-6). In the continuum, the.frequency
redistribution just mentioned can be ignored; near a spectral line, it may be
necessary to take it into account. Electron scattering is one of the mo~t
important opacity sources in hot stars (e.g., the 0-stars). (J)
I\)

RA YLEJGH SCA.TTERING I

The term Rayleigh scattering refers to the scattering of radiation by


bound systems, such as atoms or molecules, at frequencies much lower than
characteristic transition frequencies of the system. Again using equation
(4-32), the process can be described by representing real transitions of the
scatterer with equivalent classical oscillators of appropriate strengths J; 1, and
resonant frequencies wlJ equal to the actual transition frequencies. Then for
w « wiJ, equation (4-32) simplifies to yield

u(w) = (81te4/3m 2 c4 )f,iro4 /(w1f - w2 ) 2 = u,f,/n4 /(wi/ - w2 )2 (4-128)

Far from the resonant frequency, u(w) varies as w 4 or ;,.- 4 , which leads to a
strong color dependence of the scattered radiati<;m; a well-known example
of this dependence is the blue color of the sky, resulting from sunlight scat-
tered by molecules of air.
88 Absorption Cross-Sections 4-2 The Calculation of Transition Probabilities 89

Now in terms of the Einstein coefficient Bkm• its wave functions and oscillator strengths. There are four quantum numbers
that specify a distinct state of hydrogen: n, the principal quantum number,
(4-59) which characterizes energy; I, the azimuthal quantum number, which charac-
terizes the orbital angular momentum; m, the magnetic quantum number,
hence (4-60) which gives the projection of the orbital angular momentum along a preferred
axis (taken to be the z-axis); ands, the spin quantum number of the electron,
In general we shall be interested in the absorptivity of bulk material. If we
equal to±½.
assume the atoms are oriented at random with respect to a beam of radiation, In most atomic systems, the energies of different (n, 1) states are distinct, but
then (Ii· dmkl 2) = dm/ (cos 2 0) = ¼dm/, so we may write, finally for hydrogen they depend upon the principal quantum number n only, and
(4-61)
E. = -Pll/n2 (4-68)
The spontaneous emission rate follows from equations (4-8) and (4-9), which where ,d,t is the Ryclherg constant
give
(4-62) (4-69)

In many cases the upper and lower states ofa line will be degenerate (or we Here µH is the reduced mass, given in terms of the masses of a proton, mp, and
may wish to group several levels belonging to a multiplet). It is then customary electron, m., by
to sum over all substates k of th~er state i and substates m of the upper µ11-1 =mp-I+ me-I (4-70)
state j, and to define a line strength S such that
The wave function has the form [cf. (392, Chap. 5; 418, Chaps. 9 and 10)]
S(i, J) = L dm/ (4-63)
mk (4- 71)
Then we may write
g;A 1, = (641t4 v3/3hc 3 )S(i,J) (4-64) where r,m is the spherical harmonic function expressible in terms of associated
or, equivalently, Legendre functions, and R.,
is the radial function, which can be expressed in 01
g1B,1 = (321t 4/3h 2c)S(i, J1 (4-65) terms of associated Laguerre polynomials and exponentials. These functions (JJ
are normalized such that I
and from equation (4-63)
(4-72a)
(4-66)

Finally, noting that S(i, J) is a sum over all upper and lower substates,
and ft clq, J: c/O[Y,m(O, </>)]*Y,,m'(O, </>) sin O = c5,,, c5mm' (4-72b)
equation (4-66) can be used to express the total oscillator strength of a "line"
connecting two degenerate levels (or of the entire multiplet connecting two In equation (4-72a), r is measured in units of the Bohr radius
sets of closely-spaced levels). Let n' be the principal quantum number of the
lower level, and label each sublevel with l'; let n and I correspond to the upper
ao = h2/(41t2e2µH) (4- 73)

level. Then It is often convenient to work with the function P.,(r) = rR.,(r) which is
g•. f(n', n) = L g•• ,.f(n', I'; n, I)
,.. , (4-67) defined such that P., 2
measures the charge density of the electronic wave
function.
Because all states with a given n are degenerate, we require the statistical
APPLICATION TO HYDROGEN weight g,. and normally will work with the oscillator strength f(n', 11) for all
transitions (n' .... 11). The statistical weight is
Hydrogen, the most abundant element in the Universe, has the simplest
atomic structure, and it is possible to obtain exact analytical expressions for 9n = 2n2 (4-74)
5-1 lom/ Thermodynamic Equilibrium 109

ionizations (and their inverses), and thus acts to help determine the occupa-

5 tion numbers of the gas; we shall show in fact, that radiative transitions
dominate the state of the gas. In this case the occupation numbers must be
determined from equations of statistical equilibrium, which specify all of the
microprocesses that produce transitions from one atomic state to another.
The fact that the state of the material depends upon the radiation field
introduces the essential difficulty of stellar atmospheres theory for, as we
The Equations of mentioned in Chapter 2, the radiation field, in turn, depends on the occu-
pation numbers via the absorptivity and emissivity and their effects upon
Statistical Equilibrium the transfer of radiation through the atmosphere. Thus what is required is
a completely self-consistent simultaneous solution of both the radiative transfer
and statistical equilibrium equations. This is a difficult problem in general,
and its solution occupies the bulk of Chapters 7, 11, and 12 of this book.
For the present we shall only show that there are strong expectations that
the state of the material will depart from that predicted by LTE, which is
therefore at best a computational expedient. If in any particular case the
occupation numbers obtained from the general analysis happen to agree
with those predicted by LTE, then one may legitimately use the LTE assump-
tion; but for a wide range of problems (line-formation in particular), such
agreement is not generally attained (nor can we accurately predict a priori
when it will be for most cases of interest!).

5-1 Local Thermodynamic Equilibrium


CJ)
Stellar atmospheres are regions of high temperature and low density. There- In thermodynamic equilibrium, the state of the gas (i.e., the distribution of w
fore the gas consists mainly of single atoms, ions, and free electrons; in atoms over bound and free states) is specified uniquely by two thermo- I
cooler stars molecules also form. Because of the low densities, the material dynamic variables (we shall choose the absolute temperature T and the
always behaves as a perfect gas. The state of the gas is specified when we total particle density N) via the well-known equilibrium relations of statistical
know the distribution of the particles over all available bound and free mechanics. These relations will not all be derived in this chapter, as they are
energy levels-i.e., when we know the occupation numbers of these levels. easily found in standard texts [see, e.g., (565, Chaps. 12, 14, and 15; 11,
We then have the information required to compute the gas pressure, mass Chap. 3)] but will be summarized in forms useful for further developments
density, opacity, emissivity, etc. of the material. in this book. The assumption of LTE asserts that we may employ these
To specify occupation numbers, we must deal with the phenomena of same relations in a stellar atmosphere at the local values T(r) and N(r)
excitation and ionization of each chemical species in the gas. One approach despite the gradients that exist in the atmosphere. This simple assumption
is to assume that we may apply the equilibrium relations of statistical me- is actually a very strong one, for it implies that we propose to calculate the
chanics and thermodynamics at local values of the temperature and density; above-mentioned distribution functions without reference to the physical
this is the local thermodynamic equilibrium (LTE) approach. As we shall see, ensemble in which the given element of material is found. Thus it is assumed
LTE provides an extremely convel)ient method for computing the particle that it is irrelevant whether the material is contained within an equilibrium
distribution functions. One of the fundamental properties of stellar atmo- cavity (the classical hohlraum), an atmosphere with a strong radiation field,
spheres, however, is the presence of an intense radiation field whose character or in the exhaust of a space vehicle, despite the obvious dissimilarities of
is very different from the equilibrium Planck distribution. This radiation these situations! In LTE, we have a purely local theory, which makes no
field interacts strongly with the material via radiative excitations and photo- allowance for coupling of the state of one element of gas with that of another,
110 The Equations of Statistical Equilibrium 5-1 Local Thermodynamic Equilibrium Ill

say by radiative exchange (except as may be imposed by certain global con- any excited level is
straints on the atmosphere-e.g., hydrostatic or radiative equilibrium).
Moreover, in LTE the absolute temperature T has a quite general signifi- (5-4)
cance. The same T applies in the calculation of the velocity distribution
where the subscript Odenotes the ground level and the superscript • denotes
functions of atoms, ions, and electrons; the distribution of atoms and ions
LTE. For any two excited levels / and m,
over all states (Boltzmann-Saha equations); and the distribution of thermal
emission (Planck function). In short, the full implications of the LTE assump-
(nmjd111jk)* = (llmJdll1Jk) exp[ -(xmjk - Xlik)/kT]
tion are quite sweeping. It is this very fact which makes it so effective in
reducing the complexity of the equat'ions, and at the same time so difficult = (YnoJkffhJk) exp( - hv1m/kT) (5-5)
to justify physically and so vulnerable to error.
where hv1m is the energy of a photon that equals the energy difference between
THE MAXWELLIAN VELOCITY DISTRIBUTION
the levels. In calculations of ionization equilibria, we typically wish to know
the total number of atoms in a particular ionization state, which can be
The probability, in thermodynamic equilibrium, that a particle of mass written as
m at temperature T has a velocity on the range (v, v + dv) is given by the
Maxwellian velocity distribution

f(v) dv" dv, dv, = ( 21t:T r exp[ - m(v"


2
+ v/ + v, 2)/2kT] dv" dv1 dv, (5-1)
where
= (n~1klll0Jk)U Jk(T)
U1k(T) = L ll1Jk exp(- X.lJtfkT)
(5-6)

(5-7)
or, in terms of speeds on the range (v, v + dv) i

is called the partition function. A form of equation (5-4) customarily used in


f(v) dv = ( 1t:T) i exp(-mv 2/2kT)4nv 2 dv (5-2) classical curve-of-growth analyses of spectra (see §!l I 0-3 and I 0-4) is
2
CJ)
These distributions may be characterized in terms of the most probable (5-8)
~
speed ·
v0 = (2kT/m)½ = 12.85(T/104 A)½ km/sec (5-3) The partition function is tedious to compute, and for some atoms and I
ions (e.g., of the rare earths) our knowledge of the term structure is so
where A is the atomic weight of the particle. Related parameters are the incomplete that we lack the needed data. Tabulations of partition functions
root-mean-square speed (v 2 )½ = (3kT/m)t, and the root-mean-square ve- (ti, I 15-117) and convenient analytical fits with approximation formulae
locity in one component (e.g., along the line of sight) (v/)½ = (kT/m)t. (103; 220) are available. Often a fair estimate is just g0 ik or a sum over a
few low-lying states. Note that formally the partition function in equation
THE BOLTZMANN EXCITATION EQUATION (5-7) diverges if the sum extends over all (an infinite number) of bound states,
for a lower bound to each term of the sum is exp( - XtJk/kT) (where XtJk is
In thermodynamic equilibrium at temperature T, atoms are distributed the ionization potential), which is nonzero. This problem is not a physical
over their bound levels according to the Boltzmann excitation equation. Let one, for in reality the highest levels cannot remain bound because they are
n1Jk denote the number density of atoms in excited level i of ionization state strongly perturbed by neighboring atoms and ions. To estimate the effect
j of chemical species k. Let J = 0 denote neutral atoms, J = 1 singly ionized we might suppose that the only levels that are bound are those contained
atoms, etc. Measure the excitation energy XiJt relative to the ground state within the average volume available to ions. For a particle density N, the
of the atom. Let YiJt denote the statistical weight assigned to the level to mean interatomic distance is r0 = (3/41tN)½, and for a hydrogenic ion of
account for degenerate sublevels (e.g., the 2J + 1 m-states in the absence of charge Z the radius of the state of principal quantum number n is r. = n2 a0 /Z
a magnetic field). Then, according to the Boltzmann law the population of where a0 is the Bohr radius= 5.3 x 10.- 9 cm. lfwe set r. = r 0 and choosd
112 The Equations of Statistical Equilibrium 5-1 Local Thermodynamic Equilibrium 113

a typical value of N ~ 10 15 we find n :::::: 30zt, so clearly the sum is finite. Now, summing over all final states by integrating over the electron velocity
A more accurate calculation (see §9-4) shows that in a plasma with charge distribution we obtain
density n, and temperature T, an ion of charge Z suffers a lowering of the
ionization potential by an amount t:.x :::::: 3 x 10-s Zn,tr-t eV. For hydro- (no, 1, kn,/no. o. k)*
genie energy levels this implies n!•• :::::: 4 x 10 8 Zn, -trt or nm•• ~ 60Z½
for n, ~ 10 14 and T 104.~ = 81tm 3 h- 3 (g 0 , 1, k/9o. o. k)(2kT/m)t exp( - x,. o. k/kT) f
0
""
2 2
e-"' x dx (5-11)

THE SAHA IONIZATION EQUATION or, evaluating the integral,

Above the discrete bound eigenstates ofan atom there exists a continuum nt 0 , k = nt 1• kn: ½(h 2/21tmkT)i(g 0 , u/g0 , 1 • k) exp(x,, 0 , ./kT) (5-12)
of levels in which the electron is unbound and has a nonzero kinetic energy.
which is a basic form of Saha's equation. Note that in the derivation we made
The energy above the ground level at which this continuum begins is called
no explicit reference to the ionization state of the initial "atom," hence we
the ionization potential Xt· The relative numbers of atoms and ions in succes-
may extend equation (5-12) to apply between any two successive stages of
sive stages of ionization can be computed from the Saha ionization equation,
which we shall derive as an extension of the Boltzmann formula to free states. ionization
Consider a process in which an atom of species k is ionized from its ground nt1k = nt J+ 1, kn: ½(h 2/21tmkT)i(g 01k/Yo. J+ 1, k) exp(x 11k/kT) (5-13)
level, resulting in an ion in the ground level plus a free electron in the con-
tinuum moving with speed v. The energy required to carry out this process is If, further, we apply Boltzmann's formula, equation (5-4), we obtain an
x,. o.• + ½mv2 (where we have used notation parallel to that used in our expression for the occupation number of any state of ion j in terms of the
discussion of the Boltzmann formula). The statistical weight of the initial temperature, electron density, and ground state population of ion j + 1,
state is Yo, o. •• and the statistical weight of the final state (ion + electron) namely
may be written g(v) = 9o, 1, k x 9clcclron· If we use no, I, k(v) to d~note the
number of ions in the ground level with accompanying free electron with "Gk= no.J+1,kn,(gukl90.1+1,k)C,T-t exp[(x11• - x,1.)/kT]
speed in the range (v, v + dv), we may apply equation (5-4) to write = no, J+ I. kn,<l>11k(T) (5-14)
CJ)
[no. , .• (v)/no. o, .]• = [g(v)/go, o,k] exp[-(x,. o. k + ½mv 2)/kT] (5-9) Equation (5-14) is the most useful form ofSaha's equation for the formalism
01
we shall employ, and will be used to define LTE populations in the full
We identify 9cicctron with the number of phase space elements available to the I
non-LTE equations of statistical equilibrium (for this reason the superscript*
free electron, which, according to quantum statistics, is
on no.J+1,k and n, has been omitted). The. constant has the value C, =
Yclcctron = 2(dx dy dz dp,. dp1 dp,)/h 3 2.07 x 10- 16 in cgs units.
By applying equation (5-6), we may rewrite equation (5-14) as
where the factor of 2 accounts for the two possible orientations of the
electron spin. We choose the space-volume element to contain exactly one "Gk= Nj+1,kn,[g 0 , 1+1,k/U1+1,k(T)]<l>11k(T) = Nj+1,k"•~IJk(T) (5-15)
free electron, and make the substitution dx dy dz = n, - 1 • We rewrite the Further, by summing over all bound levels of the lower ionization stage and
momentum volume element in terms of the electron's speed, again using equation (5-6), we obtain an equation for the ratio of the total
dp,. dp 1 dp, = 41tp 2 dp = 41tm 3 v2 dv number of atoms in successive stages of ionization:

Then equation (5-9) becomes (N 1k/N 1+1,k)* = n,[U1k(T)/U1+1,k(T)]C,r-t exp(x 11k/kT) = n,4>1k(T)
(5-16)
[ no. 1, k(v)/no, o, .]•
By recursive application of equation (5-15) between successive stages_ of
= 81tm 3 h- 3(g 0 , 1,J9o,o,k)n. -t exp[ -:(x,.o.• + ½mv 2)/kT] v2 dv (5-10)
ionization, we can obtain an expression for the fraction of atoms of chemical
114 The Equations of Statistical Equilibrium 5-2 The LTE Equation of State for Ionizing Material 115

species k in ionization stage j relative to the total number of atoms of that rapidly as a function of temperature. In fact, this was the basis upon which
species: the first understanding of the spectral sequence as a temperature sequence
was built by Saha (546; 547), Pannekoek (495), Cecilia Payne (501), and
Fowler and Milne (222; 223). In normal stellar atmospheres hydrogen is
by far the most abundant constituent, and helium is next most abundant
with N (He)/ N(H) :::::: 0.1. The heavier elements have much smaller abundances
relative to hydrogen [ see, e.g., (252) for element abundances in the solar
atmosphere]. At typical temperatures in the solar atmosphere (6000°K)
hydrogen is essentially neutral, and the electrons are contributed mainly
by the "metals" such as Na, Mg, Al, Si, Ca, and Fe. At higher temperatures,
characteristic of the A-stars (10,000°K), hydrogen ionizes and becomes the
dominant source of electrons. At very high temperatures, characteristic of the
0- and early B-stars, helium ionizes and makes an appreciable contribution
where Jt is the last ionization stage of species k considered. We observe of electrons.
the convention that the product term for / = J• (which formally becomes
void) is replaced by unity in both the numerator and denominator. CHARGE AND PARTICLE CONSERVATION
Consideration of the above results shows that, if we know (n., T), then we
may determine, for any chemical species k, the fraction in any chosen In calculations of stellar atmospheres we specify the gas pressure from
ionization stage from equation (5-17), and in any particular excitation state the equation of hydrostatic equilibrium. Thus, given p9 and T, we know the
from equation (5-15). If, in addition, we know the total number density of total number density N from the relation
atoms of this species, we can obtain absolute occupation numbers n11k. In
practice this procedure is useful in LTE calculations of line spectra where p9 = NkT = (Na,oms + N1ons + n,)kT = (NN + n,)kT (5-18)
we are given a model atmosphere that specifies ne(z) and T(z). In computation
of the model itself, however, we generally do not know n,(z), but rather the Here N N denotes the density of "nuclei"; i.e., atoms and ions of all types.
In equation (5-18) and subsequent equations of this section we suppress {J)
total particle density N(z); we must then determine n,, and as can be seen {J)
from equation (5-17), this implies we must solve a nonlinear set of equations. the"*" that denotes LTE for notational simplicity. We define the abundance
Let us therefore now consider methods of solving the nonlinear problem. IXk of chemical species k to be such that Nk = IXtNN where LtlXk = 1. Then I
(5-19)

5-2 The LTE Equation of State summarises the constraint of particle conservation (i.e., :rNk = NN). In
for Ionizing Material addition we require the plasma to be electrically neutral; then the number
of free electrons equals the total ionic charge, an(! the condition of charge
The Saha-Boltzmann relations allow a computation of the fraction of each conservation reads
chemical species in various stages of ionization, and the number of free h h h
electrons that each contributes to the plasma. Stellar atmospheres consist 11, =L L jN Jk = Lk N • L jf1.(n,, T) = (N - n,) Lk IX• L jf'Jk(n,, T)
of a mixture of elements with widely differing ionization potentials; in k j= I }= I j= I

general some of the species may be neutral while others are singly or multiply (5-20)
ionized. Usually the transition from one ionic stage to the next occurs
fairly abruptly with increasing temperature, and normally a particular As mentioned above, if we know (n,, T) we may calculate N and the hk
chemical species exists essentially entirely in two successive ionization stages. directly. But if we know (N, T), we must find n, from a nonlinear equation.
This provides a sensitive diagnostic tool to infer the temperature structure Before the availability of electronic computers this problem was solved by
of a stellar atmosphere, for it implies that ratios of line strengths of two constructing tables of log p9 (T, log p,) (here p, = n,kT), in which interpo-
successive ionic spectra (e.g., He I and He II, or Ca I and Ca II) will vary lations could be made to find Jog p,(T, Jog p9 ). Examples of such tables are
116 The Equations of Statistical Equilibrium 5-2 The LTE Equation of State for Ionizing Material 111

given i? (11, 130) and (638, 104). We shall develop a different procedure, A fairly good estimate of n, can be obtained from equations (5-24) and (5-25)
along Imes suggested by L. H. Auer, that is better suited for machine com- if IXM « 1 and XM « XH by writing n, :::::: n,(H) + n.(M).
putation, and that fits into the overall approach of Chapter 7 for the
computation of model atmospheres. But first consider an instructive example
SOLUTION BY LINEARIZATION
that yields physical insight in limiting cases.
Suppose that the gas consists only of hydrogen (XH = 13.6 eV) and one Let us now turn to the problem of determining n, for a given value of
metal "M" with a single ionization stage of much lower ionization potential (N, T) by means of an iterative linearization procedure (generalized Newton-
(say XM = 4 or 5 eV) and an abundance relative to hydrogen °'M « 1. At Raphson method). w~ shall describe the procedure in fair detail because it
hi~h temperatures where the hydrogen is appreciably ionized, it will con- is a simple example of the approach we shall use i~ more complicated cases
tribute most of the electrons; at lower temperatures the hydrogen is neutral, (e.g., the non-LTE rate equations and the transfer equation). The only
and n, is determined by fM, the ionization fraction of the metal. The number equation to be solved (contrast this with the non-LTE case, cf. §5.5 !) is
of particles of all types is equation (5-20) where /Jk(n,, T) is given by equation (5-17). Suppose that
we have an initial estimate of the electron density, n,°; suppose also we
N = nH(l + fH) + °'MnH(l + fM) (5-21) find that using n, 0 to evaluate the righthand side of equation (5-20) yields
while the number of electrons is a density n, 1 :/; n,°. It is then clear that the true density differs from n.°,
so we write n, = n, 0 + on, where on, is to be determined in such a way
n, = nP + nM+ = nH(fH + °'MfM) (5-22) as to satisfy equation (5-20) exactly. Because the equation is nonlinear, we
cannot determine this on, exactly, but on the supposition that on,/n, 0 « l,
Then pefp, = {JH + IXMfM)/[l + fH + IXM(l + fM)] (5-23) we can estimate on. by expanding all terms to first order and solving for
~t high _enough temperatures, fH-+ 1, and as °'M « t (p,/p_g)-+ ½. At on,. Then we have
mtermed1ate temperatures, where °'M « fH « 1, and at the same time fM :::::: 1, n,° + on, : : : [(N - n,° - on,)· t(n,°, T)] + (N - n,°)[of(n,, T)/on,] ••o on,
(p,/p,) :::::: fH• At low temperatures, fH -+ 0 while JMifH » 1, hence (p,/p,) -+
°'MfM• We thus see that at high temperatures the metals are essentially (5-26)
irrelevan~ to the determination of p.fp,, while at low temperatures they
play a crucial role. In particular, note that the metal abundance enters or on,:::::: [(N - n,°)t - n,°][l + t - (N - n,°)(of/on,)]- 1 (5-27) (j)
-..._J
directly in fixing p. in cool stars; this is important because the dominant 1.
opacity source in cooler atmospheres is absorption by the H- ion, and where l:(n,, T) = L °'ksk - 1(n,, T) L jP1k(n., T) (5-28) I
n(H-)/n(H) is proportional ton,. Thus in these stars the metal abundance k Jc!
fixes the opacity as well.
For a pure hydrogen gas, equations (5-16), (5-19), and (5-20) may be solved Note that we may rewrite the functions P(n., T) and S(n., T) as
analytically to obtain

(5-24)
which shows that at low degrees of ionization, n, ~ N½ for a given T. and
Exercise 5-1: Derive equation (5-24).
The value of on, given by equation (5-27) will not be exact, so we iterate
If only the metal in our two-component gas described above is ionized the procedure by using a new estimate n,°(new) = n,°(old) + on, to re-
UH « °'M) then we have evaluate t and of/on., and to compute yet another value of on,.
The convergence of this procedure is quadratic (if our original estimate
lies within the range of convergence) so, if the first fractional error on,/n.:.
(5-25) is £, subsequent iterations will produce corrections of order £ 2, e4, £ 8 , etc.,
118 The Equations of Statistical Equilibrium 5-3 The Microscopic Requirements of LTE 119

which implies that one can obtain the result to the desired accuracy quickly. Exerci.w! 5-2: Obtain expressions for the coefficients in equation (5-35) in terms
It is also worth noting that the derivative iJ'r./on, can be evaluated of PJk• St, and <f>ut• and their derivatives.
analytically:
Equations (5-34) and (5-35) provide the information we shall require in
§7-2 to find the response of the opacity and emissivity (bx, D17) to changes
in the model structure (t,N, {JT).

where (iJP1k/iJn,) and (iJSk/iJn,) follow immediately from equations (5-29) Exerci.w! 5-3: Show that <1NJh for the last ionization stage of element k, has a
and (5-30) and produce a compact expression for (5-31). In general, the particularly simple form because hk involves only s,. Then show that expressions
derivatives appearing in linearization procedures can be estimated numer- for oN1k of lower ions can be evaluated recursively from equation (5-16), and that
ically; however, we shall usually be able to obtain analytical derivatives, these lead from equation (5-15) to simple expressions of the form of equation
and experience has shown that in this way we obtain better control of the (5-35) for 01111,.
calculation.
Finally, having obtained a satisfactory value for n., and, as a byproduct
the Jj., we may calculate any particular occupation number from equa- 5-3 The Microscopic Requirements of LTE
tion (5-15)
Before we develop the equations of statistical equilibrium, it is worthwhile
n11k = N1+ 1,1n,4>,Jk(T) = rx1(N - n,)n,fi+ 1, k(n,, T)<l> 11k(T) (5-32) to discuss qualitatively the microscopic requirements ofLTE. An interesting
commentary on these requirements by K. H. Bohm may be found in (261,
this completes the computation of the LTE equation of state. Chap. 3); we shall summarize and discuss this analysis here along with other
The procedure outlined above has a larger significance than indicated material of relevance.
thus far. We have assumed that N and T are given. 'But these quantities
follow from constraints of pressure and energy balance, and in general are DETAILED BALANCE
known only approximately at any particular stage of calculation of a model.
As we shall see in Chapter 7, we may apply the linearization procedure to In thermodynamic equilibrium, the rate at which each process occurs is
all the variables involved, and hence we shall need to evaluate the response exactly balanced by the rate at which its inverse occurs, for all processes; i.e., O>
of the ·occupation numbers to the perturbations t,N and t,T. Perturbing each process is in detailed balance. This is a very strong requirement, and it (X)
equation (5-20) we obtain proves to be very useful in constructing relations among rate coefficients I
(recall the use of this procedure in Chapter 4). We may classify processes
n, + bn, = (N + {JN - n, - bn,)t that produce transitions from one state to another (bound or free) into
+ (N - n,)[(iJt/iJn,) bn, + (iJ'r./iJT) t,T] (5-33) two broad categories: radiative and collisional. Collisional processes are
the processes invoked in statistical mechanics to establish equilibrium, and
or, assuming that n, is a solution of equation (5-20) at the current values can be expected to be in detailed balance whenever the velocity distribution
of(N, T), . of the collidino particles is the equilibrium (i.e., Maxwellian) distribution. We
shall show below that this can be expected to be the case in stellar atmo-
bn, = [I + t - (N - n,)(o'r./on,)]- 1 [t t,N + (N - n,)(iJt/iJT)bT] spheres. Furthermore, we may make the same statement about processes
= (iJn,/iJN)r t,N + (iJn,/iJT)N {JT (5-34) which are essentially collisional in character, even though a photon is
emitted (e.g., free-bound radiative recombination and free-free emission);
where again ot/iJT may be evaluated analytically. Further, from equation we can therefore use detailed balancing arguments to calculate the rates or
(5-32) we may develop an expression for bn11• of the form bn11k = ~, t,N + these processes when convenient to do so. In contrast, radiative processes
~2 t,T + ~3 bn,, which can be collapsed down by use of equation (5-34) (e.g., photoexcitation, photoionization) depend directly upon the character
to an expression of the form of the radiation field, and will be in detailed balance only if the radiation
field is isotropic and has a Planck distribution. We shall show below tqat:,
(5-35) this is not the case in stellar atmospheres. ·
120 The Equations of Statistical Equilibrium 5-3 The Microscopic Requirements of LTE 121

If son_ie processes ar~ in detailed balance while others are not, the final less because of limb-darkening) but in an extended stellar envelope W « I
occupation numbers will be determined by a competition among them and (and in a planetary nebula, W ~ 10- 14). Thermodynamic equilibrium
m~y depar! ~ore or less strongly from an equilibrium distribution. LTE requires that W = l, so it is clear that detailed balancing in radiative
wdl be vahd in the deepest layers of stellar atmospheres where densities transitions cannot in general occur in a stellar atmosphere.
are high and the collision rates become large, and the optical depth is so In addition to being dilute, the stellar radiation field has a markedly
large that no photon escapes from the atmosphere before being thermalized non-Planckian frequency distribution. As we know from the Eddington-
so that the radiation field approaches the Planck function. But in the observ~ Barbier relation, the emergent specific intensity at frequency v is approx-·
able layers, precisely the opposite; regime is found.
•,
imately equal to the source function s. at t, = l. Even if s. were B., the
fact that the material is vastly more opaque at some frequencies than at
others (line to continuum ratios are often 10 3 and may reach much larger
THE NATURE OF THE RADIATION FIELD
values) implies that the radiation will emerge from greatly differing depths
A s_tellar atmosphere is not in any sense a closed system in equilibrium at substantially different temperatures; the radiation field is therefore a
at a uniform temperature. Indeed the opposite situation prevails: radiation composite of widely differing radiation temperatures. The effects of the
flo':"s f~eely_ from the surfa~ ~ayers of_the st_ar into essentially empty space, temperature gradient become extreme when hv/kT > 1, for then the Planck
which 1mphes that the radiation field 1s decidedly anisotropic, and that the function varies as exp( - hv/kT) and becomes very sensitive to small changes
at~osphere ?as a large temperatu~e _gradient. The radiation field at any in T. Ifwe were to parameterize the radiation field by introducing a radiation
point_ 1s the integrated result of em1ss1ons and absorptions over the entire temperature TR(µ, v) such that for µ ;;i,: C, J(r •• µ, v) = W B,[T R(µ, v)], we
(po_ss1_bly large) volume. within which a photon can travel from its point of would find marked variations of TR with both v and µ. For example, in
em1ss1on to the test point. 1:his volume may include the boundary surface the solar spectrum, TR ranges from 4800°K in the visible to ~25,000°K
and empty s~ace beyond, with a consequent reduction of intensity, as well in the ultraviolet in the ground-state continuum of He+. In sum, the radiation
as _l~yers of higher ~e~peratures and densities from which intense radiation field displays an extremely complex behavior, and the conditions required
ongmates. The rad1at1on field therefore is distinctly nonlocal in nature and to assure LTE a priori are simply not met.
has an absolute intensity, directional distribution, and frequency spec~rum
that may h~ve_ no resemblance whatever to the local equilibrium distribution THE ELECTRON VELOCITY DISTRIBUTION
CJ)
B.(T). Rad1at1ve rates may therefore be far from their equilibrium values
and thus tend to drive the material away from LTE. ' In stellar atmospheres, the free electrons are produced by photo- (()
The radiati~n field is plainly anisotropic because the radiating surface ionization and collisional ionization. The inverse processes are radiative I
subtends a sohd angle less than 41t, and essentially no radiation enters from recombination and three-body collisions, which lead to recaptures of elec-
the_su~rounding void. We may describe this geometrical effect by introducing trons into bound states. While in the continuum, an electron may undergo
a d1lut1on factor W defined to be w./41t where w is the solid angle subtended elastic collisions with other electrons and inelastic collisions (leading to
by the stellar disk. • excitation or ionization of bound electrons) with atoms and ions. The elastic
collisions redistribute energy among the electrons and tend to lead to an
Exercise 5-4: Show that
equilibrium partitioning-hence a Maxwellian velocity distribution. If a
Maxwellian velocity distribution is in fact attained, we may define the local
temperature to be the kinetic temperature of the electrons. On the other
(5-36) hand, inelastic collisions and recombinations disturb the achievement of a
Maxwellian velocity distribution, for the inelastic collisions involve electrons .
where r• is the radius of the radiating surface and r denotes the position of the only in certain velocity ranges and tend systematically to shift them to much
observer. Show that for r./r « 1, w = ¼(r./r) 2 lower velocities, while recombinations remove electrons from the continuum
and prevent further elastic collisions. Whether or not the Maxwellian velocity
As defi~e~, W clea~ly measures the factor by which the energy density in distribution is established hinges upon how rapidly thermalization by elastic
t~e rad1at1on field 1s reduced as the source of radiation moves to a large collisions occurs compared to the perturbing processes: if it occurs much
distance. At the "surface" of a star, it is obvious that W = ½(actually a little more rapidly, the velocity distribution will be very nearly Maxwellian.
5-3 The Microscopic Requirements of LTE 123
122 The Equations of Statistical Equilibrium

The thermalization rate can be measured in terms of the relaxation time population is far from its equilibrium value; these conditions can occur in
of the system, which, for particles interacting with themselves, is the solar chromosphere.
Finally, one may ask if the atoms and ions in the atmosphere also have a
(5-37) Maxwellian velocity distribution, and if their kinetic temperature T k = T •.
An analysis of this question (88) for a pure hydrogen atmosphere_ of ato~s,
(see 598, Chap. 5). Here D is the Debye radius (see §9-4) D = (kT/8rce 2 n,)t ions, electrons, and radiation, demanding a steady-state solution, "'.'h1le
and p0 = e2 /mv 2 is the impact parameter for a 90° collision. Now consider allowing for energy exchange among the four components of the medium.
recombinations; if u is the average cross-section for the process then the shows that ifn, > 10 10 (a condition easily met in the bulk of the atmosphere)
mean time between recombinations is and 5 x 10 3 < T. < 10 5 , then jTk - T.I ~ 10- 3 T •. It thus appears safe
to conclude that a unique local kinetic temperature applies to all the particles
(5-38)
in most atmospheric regions.
where N is the density of the particles with which recombination occurs.
Two astrophysically important processes are (a) H + e ➔ H- and (b) H + + THE IONIZATION EQUILIBRIUM
e ➔ H. At T ~ 6000°K (a typical solar temperature) u"- ~ 3 x 10- 22
cm 2 and n./N" ~ 10- 4 , At T ~ 10000°K, <1" ~ 6 x 10- 21 cm 2 and The degree of ionization of stellar material is determined_ b~ the balanc_e
n./np ~ 1. Substituting these values into equations (5-37) and (5-38), we of photoionizations and collisional ionizations against rad1at1ve re~omb1-
find t,/t, ~ 10 5 for process (a) and t,/t, ~ 10 7 for process (b). We conclude, nations and three-body collisional recombinations. Let us first examme the
therefore, that under representative conditions in stellar atmospheres a relative rates of photoionization and collisional ionization; it suffices to
free electron will undergo an enormous number of elastic scatterings between obtain only an order-of-magnitude estimate. . .
recombinations, and that the latter will not seriously hinder equilibration The energy absorbed by an atom in bound state i at frequency v m mterval
to a Maxwellian distribution. dv is 4rcJviX 1(v) dv; each photon has energy hv, hence the total number of
Let us now consider inelastic collisions. Collisions of-electrons with the photoionizations is
most abundant element, hydrogen, occur frequently, but the excitation
energy of hydrogen is 10 eV while the thermal energy of the electrons (5-39)
is 1 eV. Thus only 3 x 10- 5 of the electrons have sufficient energy, to
induce the excitation, and only a fraction of these will be effective. Using To estimate Ri,• we adopt a hydrogenic cross-section --...J
typical excitation cross-sections one finds that (at 10,000°K) the rate of 0
inelastic excitations is of the same order as the recombination rate-i.e., I
very small compared to the elastic collision rate. One must also consider
collisions with other elements, which may be grouped as follows: (a) the where J. is the integrated oscillator strength for the continu_um. Further,
alkalis, which have large cross-sections but low abundances (10- 6 ); (b) Fe, we write
which ha~ .numerous low-lying levels and moderate abundance (4 x 10- 5); co
and (c) C, N, and 0, which have small cross-sections but large abundance Jv = WB.(TR) = W(2hv 3 /c 2 )}: exp(-nhv/kTR)
n=l
(10- 3). Most of the levels for groups (b) and (c) are metastable, so that most
co
o(the inelastic. excitations are subsequently cancelled by collisional de- (5-40)
excitation; ignoring this effect we overestimate the number of inelastic Then R1, = (16rc 2 e2 v//mc 3 )J.W }: E 1(nhvo/kT R)
n=I
excitations. Taking the various factors into account and ignoring compen-
sating de-excitation, Bohm estimates (elastic collisions/inelastic collisions) ~ The rate of collisional ionizations can be computed from u(v}, the collisional
10 3 and hence concludes that a Maxwellian velocity distribution that defines ionization cross-section for electrons of velocity v:
T, is established. Recent work (573) suggests that departures from a Max-
wellian distribution in a pure hydrogen gas can occur in the high-energy n-C = n1n. Juo
rco u(v}f(v)v dv (5-41)
1K
tail if(a) the ionization level is very low (n./n" :$ 0.01) and (b) the ground-state i
124 5-3 The Microscopic Requirements of LTE 125

TABLE 5-1 We may carry out similar estimates for the rates of radiative recombination
Ratio of Radiative to Collisional Ionization and three-body collisional recombination; these processes are both essen-
Rates
tially collisional and hence per ion occur at the LTE rate. We can then use
Star Xion = SeV ;i: 100 = leV detailed balancing arguments to compute the rates in terms of the equilibrium
values of the upward rates. We may still use equation (5-44), except that for
Sun 103 2 the radiative recombinations the appropriate temperature is now T, not
O-star 20 0.2 TR• and W = 1. We then find that radiative recombinations always outweigh
collisional recombination, both in the photosphere and corona (in the
SOURCE: From data by K. H. B6hm, in Stellar
Atmosph,res, ed. J. L. Greenstein, Chicago: corona, yet another mechanism-dielectronic recombination-outweighs
University of Chicago Press, 1960, by permission. radiative recombination).
The ionization balance is thus determined by photoionizations and radia-
tive recombination; to establish the equilibrium the numbers of ionizations
To obtain an estimate, we adopt the semiclassical Thomson formula [cf. nr
and recombinations are equal: n1R1" = n"R"' = Rt where the last equality
(684, 120)] follows by a detailed-balance argument. Hence for the ground state,
(5-42)

where E = ½mv 2 is the energy of the incident electron. Substituting equations


(n~Jin 01 ) = 4nW f.: (hv)- 1a..B,(T R) dv / 4n f.: 1
(hv)- a..B,(T.) dv
(5-2) and (5-42) into (5-41) and integrating we obtain

(5-43)
= W J,.:.tkTR e-"x- 1 dx/ J,.:,kr. e-"x- 1dx
WE 1(hv 0 /kTR)/E 1(hv 0 / kT,)
where u0 = hv0 /kT •. In the limit that hv0 » kT Rand hv0 » kT, we retain
only the first term of equation (5-40), and use the asymptotic result that for :::: W(T R/T,)[exp( -hv0 /kT R)/exp( -hv 0 /kT,)] (5-45)
x » 1, E 2(x) -+ E 1(x) -+ e-"/x to obtain
where we have again used hydrogenic cross-sections. If we substitute for
R," ~ 4(2n k)*hv
c,"
3
2 3
3m*e c
3
0 (WTR)ex [hv (-1- __
O
n,T.* p kT.
l_)]
kTR
(5-44)
n~ from the Saha equation (5-13) we may obtain the approximate ionization
1
ec1uation
----.J

Forphotosphericlayerswecouldadopt W ~ ½, TR ~ T,. Bl:lhmcalculates (n,n 0 , J+ i/n 0 , 1)


estimates of R1"/C," for representative cases of levels with ionization poten- = W · (2g 0 , J+ ifg 0 , 1)(2nmkT Rfh 2 )t ··(T ,/TR)t · exp(- xdkT R) (5-46)
tials of 1 eV and 8 eV for conditions characteristic of the outer layers (-r ~ 0.05)
of the sun and an O-star. In particular, for the sun he adopts n, :::: 3 x 10 12 , which has been extensively applied-e.g., in analyses of gaseous nebulae.
and T ~ 5 x 103 °K while for the O-star he uses n, :::: 3 x 10 14, T :::: To analyze the ionization balance in stellar atmospheres we now must
3.2 x 104 °Kand finds the values for R1"/C1" listed in Table 5-1. It is clear decide (a) how to choose J., and (b) which levels dominate. Bl:lhm suggested
that in stellar photospheres, the radiative rates dominate, except for high- J J
comparing the values of 41t (hv)- 1 K.}, dv with 41t (hv)- 1 K.B. dv, where
lying levels at high temperature and densities. In fact, for O-stars the im- "• is the total opacity from all overlapping continua and J, is the mean
portant levels have even larger values of Xion than those listed in Table 5-1 intensity obtained from LTE model-atmosphere calculations. If these num-
(e.g., the ground state ofH at 13.6 eV and the ground state of He I at 24.5 eV), . bers are equal the claim is made that LTE is self consistent. Bl:lhm examines
and are even more markedly radiatively dominated. Thus the ionization equi- the Fel .... Fe II equilibrium in a model solar atmosphere and finds that
librium is vulnerable to departures from LT E if J. departs from B•. Note in the rates mentioned above have a ratio of 2.9 at "'r = 0.01, 1.3 at "'r = 0.05,
passing that in the corona of a star where T, ~ 2 x 106 °K and TR ~ and essentially unity at T ~ 0.1; from this one is tempted to conclude that
6 x 103 °K (for the sun), while the relevant values of hv0 are around 300 eV, the Saha ionization formula is valid below "'r = 0.1.
the exponential factor in equation (5-44) becomes very small and collisional There are, however, flaws in this argument. First, it is clear that integr~tep
ionizations dominate. rates summed over all continua of an atom may be subject to cancellattoris
126 127
The Equations of Statistical Equilibrium

and compensations, and}t is not at all clear what a given departure between TABLE 5-2
Ratio of Collisionc,/ to Radiative Excitation Rates
the two integrals implies for any particular level (i.e., some levels may be
overpopulated and others underpopulated and the integrals could balance). Star ;.(Al = 3000 4000 5000 6000 7000 8000 9000
Second, and far more important, the reasoning is circular if from the outset
we use B, as S, to calculate J., for we know that 0.003 0.007 0.017 0.035 0.061 0.099 0.15
Sun
0.19 0.44 0.85 1.4 2.2 3.1 4.2
O-Star

SOURCE: K. H. B<ihm, in Stellar Atmo,pheres, ed. J. L. Greenstein, Chicago: University


i.e., J. is forced to B. at 't'v c:=: 1, and the two integrals automatically become or Chicago Press, 1960, by permission.
equal artificially.
As we shall show below (cf. §7-5 and Chapters 11 and 12) the characteristic J. with WB,.. We again use equation (5-43) to calculate the collision rate
feature ofnon-LTE transfer is that the source function contains a dominant with.Jc replaced by flJ = B1ihvmc/4rc 2 e2 • We then find
scattering term that is only weakly coupled to local thermal parameters. In
such cases, s. may differ greatly from B. over large ranges of optical depth
(essentially to the depth from which a photon cannot escape without being
S =[
Rii
3e2mt).3 ]
2h(2rc 3 kT) 1
(n•)
W
E2 (x)(e" - 1) (5-47)
thermalized despite the high probability of scattering and low probability
of destruction). Furthermore, if we attempt to find s. by starting with B., where x = hviJ/kT. BBhm estimates this ratio for t~pical conditions in ~he
computing J., using this to re-evaluate S, and iterating, it is found (cf. §6-1) solar atmosphere and the atmosphere ofan O-star (usmg W = 1) and obtains
that the convergence rate is extremely slow; with a single iteration one the results shown in Table 5-2. We see that the radiative rates dominate
inevitably obtains an estimate of s. that is very close to LTE, but that is except in the red and infrared of hot stars. The same remarks made above
false. Further iterations (perhaps thousands!) are required to propagate about the non-Planckian nature of the radiation field apply (even more
information about the existence of a boundary to the deeper layers by the strongly!) in the lines; hence again we conclude that the statistical equilibrium
inefficient iteration procedure; each iteration will show a continuing, pro- and transfer equations must be solved self-consistently. One might be tempted
gressive departure from LTE, and when strict consistency between s. and to conclude that LTE must prevail in the long-wavelength line spectrum on
J. is obtained, the departures are much larger and extend far deeper than the basis of the dominance of collision rates. But as we shall see in §12-4 this
the first iteration indicates. In short, experience has shown that estimates 0£ is not true, and, in fact, these lines often show the largest effects of departures ---.I
the kind outlined above, based upon a single iteration away from LTE are from LTE! I\)
worthless. The fact that the seemingly plausible arguments based on such I
estimates are false was not realized in much of the classical work on stellar
atmospheres, and errpneous conclusions about the validity of LTE were 5-4 The Non-LTE Rate Equations
drawn; we shall return to this crucially important point in Chapters 7, 11,
and 12-further discussion has also been given by Thomas (626, 141-147). Let us now consider the equations of statistical equilibrium (or rate equation.~)
In summary, we have shown that in the ionization process radiative rates by which we calculate the actual occupation numbers ?f b~u?d and fr.ee·
dominate collisional, and given the nonequilibrium character of the radiation states of atoms in stellar atmospheres. We shall make the s1mpltfymg assump-
field we must expect that LTE will not be valid, and therefore from the tion of complete redistribution in the lines (i.e., the emission and_ abso,:Pti~11
outset we must perform a simultaneous solution ofboth the statistical equilib- profiles are taken to be identical); the equations developed on this basis will
rium and transfer equations. Only when a strictly self-consistent solution be used in our discussion of line formation through Chapter 12, and con-
is obtained is it possible to decide in which regions LTE actually prevails. sideration of partial redistribution effects will be deferred until Chapter 13>

THE EXCITATION EQUILIBRIUM GENERAL FORM

As in the case of ionizations, we ·again ask whether any particular Consider a volume element in a moving medium. The number density
transition is dominated by collisional· or radiative processes. The radiative of particles of a given (bound or free) state i of chemical species k will cha1w1
J
excitation rate is given by BIJ <f>.J. dv, which we shall estimate by replacing in time according to the net flux of particles through the volume and the
128 The Equations of Statistical Equilibrium 5-4 The Non-LTE Rate Equations 129

net rate at which particles are brought from other states, j, by radiative and isotropic, so integrating over all angles and frequencies we have the number
collisional processes; that is, of absorptions in the line

niRiJ = niBii f <f,.J. dv = niB1iJIJ n14na.lJJli/hviJ


=

Here P11 denotes the total rate from level i to j. The first term on the right- = n 4n f a.lJ(v)(hv)- J. dv
1
1
(5-53)
hand side can be shown, by use of the divergence theorem, to be the net
number of particles streaming into and out of the unit volume. Ifwe sum over In moving media (Chapter 14) we may consider either the comoving frame,
Lt
all states ofspecies k and write Nt = n1t, then we have a continuity equation in which case equation (5-53) remains valid if J. is the mean intensity as
for this species measured by an observer at rest with respect to the moving fluid, or the
(5-49) observer's frame, in which case <f,. now has an angular dependence, ~n_d a
double integration over angles and frequencies must be earned out exphc1tly.
Multiplying equation (5-49) by mt, the mass of species k, and summing over Similar remarks apply to other radiative rates given below.
all chemical species we obtain the standard hydrodynamical continuity The number of stimulated emissions is
equation
(op/ot) + V · (pv) = 0 (5-50) n1B11 f <f,J. dv = n1B11 J 11 = ni(g1B ifg1)]11 = ni(4n/hv11)(g1a.11/01)J1J
1 (5-54)
where p = L mkNt, For a steady state, equation (5-48) simplifies to The number of spontaneous emissions is
n,t L P11 - L "1tPJ1 = -V · (n1tv) (5-51)
f
n1A1i <f,. dv = ni(2hv1//c 2 )B11 = n1(2hv,//c 2 )(4n/hv11)(g1a.1i/01) (5-55)
J•I J•I

and for a static atmosphere we have (suppressing the subscrii,t k) The total downward rate is the sum of the spontaneous and stimulated rates:
n1 L P11 - L n1P11 = 0
J•I )"I
(5-52) n R11 = ni(AJi + B11 J 11 ) = n1(4n/hv 11 )(g,a. 1ifg1)[(2hv,//c2 ) + 111] (5-56)
1
As we consider only static media through Chapter 13, we shall deal almost A prime has been added to R11 so as to r~serve the unadorn~d symbol for a --..J
exclusively with equation (5-52). We shall show how to handle moving media different use below. We may rewrite equation (5-56) by factonng out the term w
in Chapter 14, and will mention some implications of the righthand side of .(ndn1)* = g1 exp(hv11 /kT)/g1 from the righthand side, and express the total I
equation (5-51) in Chapter 15. The total rate PIJ in general contains both downward rate as
radiative and collisional terms; let us now write these out in detail.
' -= ("i)* R11 -- "1 ("i)* [4n s· a.lJ(v)
n1R11 "1 -
~ ~
cl +
hv (2hv3
.
J )
• e-hv/kT dv] (5-57)
.
RADIATIVE RATES
At first sight equation (5,57) appears very cumbersome, for clearly the terms
(a) Bound-Bound Transitions. We shall develop two notations (which (hv)- 1, v3, and exp( - hv/kT) can all be taken out.side the integral owing ~o
have identical physical content) by writing bound-bound rates in terms of the swift decrease of <f,. away from v11 . We have written the downward rate m
Einstein transition probabilities or in terms of energy-absorption cross- this particular way because it is then of exactly the same fo~ _as the dow~-
sections; the former is useful for analytical manipulation with simplified ward rate in the continuum; moreover, the downward colhs1on rates _will
atomic models, while the latter allows bound-bound and bound-free rates also have a factor of (n,ln 1)• appearing_ explicitl~. In the ~nd we achieve
to be written in an identical form well-adapted to model-atmosphere com- notational economy in the full rate equations by usmg equation (5-57) rather
putation. Consider transitions from bound level i to a higher bound level j, than the simpler Einstein probability form. . .
in a line with absorption (and emission) profile <f, •. The number of transitions Finally, it is sometimes useful to work with the net rate from levelJ to level 1,
produced by incident intensity /. in the frequency interval dv and solid angle
dw is n1B11<f,J. dv dw/4n or n1(a. 11 /hv)<f,.I • dw dv. In a static medium, <f,. is nj(A 1, + B11J IJ) - n1 B1J J IJ = 111A11 ZJt (5•58).
130 The Equations of Statistical Equilibrium 5-4 The Non-LT£ Rare Equarions 131

where the term Z 11 is called the net radiative bracket (NRB). Net radiative Recall from equation (5-14) that (ni/11.)* = ne<I>1K(T), which shows that the
brackets are useful notational devices that we shall employ in Chapters 11 spontaneous recombination rate depends on the product of the electron and
and 12. Further, Z 11 can be rewritten as ion densities and a function of the temperature (which itself depends on
atomic properties through the cross-section).
Z 11 = 1 - ] 11 (n 1BIJ - n1B11 )/(n1A11 ) = 1 - (JIJ/Sii) (5-59) The number of stimulated recombinations may be calculated by a similar
procedure; in T.E.,
where S11 is the frequency-independent line source function. Because the NRB
contains only the ratio of J to S, it is often true that it is known to much higher R, 1ls1im
(n •• * = n,• • 47t f,""•·o oc,. (V)(h V)- 1 B ,.e -h\-/kT dV (5-63)
accuracy in an iterative procedure than either S or J themselves. Under
favorable conditions, use ofNRB's can significantly enhance the convergence To generalize this result for the non-LTE case, we (a) replace the equilibrium
of certain types of solutions of multilevel line-formation problems. If a radiation field B,. by the actual value J,., and (b) use the actual ion density 11.:
particular line i --+ j is in radiative detailed balance, then Z 11 = 0, and we may
cancel the corresponding terms out of the rate equations analytically (i.e.,
omit RIJ and R11 ); this situation occurs when a particular line thermalizes, and
(11.R~;)slim = 11.(111/nK)*41t f.:: oc,K(v)(ltv)- 1
J,.e·hv/kT dv (5-64)
the cancellation procedure is of great use in simplifying the rate equations The total number of recombinations is, therefore,
(cf. §7-5).
(b) Bound-Free Transitions. Let us now calculate the radiative rates from
a bound level i to the continuum "· Let oc1K(v) be the photoionization cross-
section at frequency v; then the number ofphotoionizations is calculated by
dividing the energy absorbed in interval d11 by the appropriate photon energy
'11•, and summing over all frequencies. Thus the number of photoionizations is (5-65)

(5-60) The number of recombinations is sometimes expressed in terms of a recom-


bination coefficient ocRR(T) defined such that total recombination rate as given
We may calculate the number of spontaneous recombinations by use of a by equation (5-65) is nKn,ocRR(T).
detailed-balancing argument. In thermodynamic equilibrium, the number of By comparison of equations (5-53) and (5-60) we see that in writing com-
spontaneous recombinations must equal the number of photoionizations plete rate equations we can systematize our notation and write all upward
calculated from equation (5-60) when (a) J. has its equilibrium value (i.e., radiatit>e rates (i --+ j), for j bound or free, as n1RIJ where
B,.) and (b) we correct for stimulated emissions at the T.E. value by multiply-
ing by a factor of (1 - e-hv/tT) (c.f. §4-3). Thus if nK denotes the ion density, (5-66)

(5-61) and, by comparison of equations (5-57) and (5-65), all downwar,I radiative
rates (j--+ i) as nJln;/nJ)*R 11 where
The recombination process is a collisional process involving electrons and
ions, and therefore _is proportional to nK • n•• For a given electron density (5-67)
and a given T., which by definition describes the electron velocity distri-
bution, the rate just calculated above must still apply per ion, even out of Note th~t in equilibrium, R:'i = Rj,.
T.E. Hence to obtain the non-LTE spontaneous recombination rate we
need correct equation (5-61) only by using the actual ion density nK. Then
COLLISIONAL RATES

(nKR~1)1pon = nK(ni/nK)* · 4,r J, 00

YO
oc1K(v)(hv)- 1B.(1 - e-hv/kT) dv . The gas in stellar atmospheres is a plasma consisting of atoms, ions, and
electrons, among which a wide variety of collisions may occur and produce:
= nK(nifnK)*47t J.: oc1K(v)(hv)- 1(2hv3/c 2
)e-h•/kT dv (5.-62) excitations and ionizations. In cool stars, where the material is primarily
132 The Equations of Statistical Equilibrium 5-4 The Non-LTE Rate Equations 133

neutral, collisions with neutral hydrogen atoms are numerous and important. equation (5-2) into (5-68), we find
As the material becomes appreciably ionized, however, collisions with
charged particles predominate owing to the long-range nature of Coulomb q;i(T) = C 0 rt f"' QIJ(ukT)ue-• du (5-72)
Juo
interactions. Moreover, because the collision frequency is proportional to
the flux of impinging particles, and hence to their velocity, we normally need where u = E/kT, and C0 = 7ta0 2(8k/mn)½ = 5.5 x 10- 11 • Writing x =
to consider only collisions with electrons, for in thermal equilibrium their (u - u0 ), where u0 = E0 / kr, we obtain
velocities are a factor of (m"A/m,)t ~ 43At larger than those of ions of
qiJ(T) = C 0 rt exp( - E0 /kT)f 1j(r) (5- 73)
atomic weight A.
If we denote the cross-section for producing the transition (i -+ j) (where j
may be bound or free) by collisions with electrons of velocity v as e1I j(v), then
where f1J(T) = f
0
"' Qu(Eo + xkr)(x + u0 )e-x dx (5-74)
the total number of transitions is
Exercise 5-5: Verify equations (5-72) through (5-74).
(5-68)
The advantage of writing the collision rate as in equation (5-73) is that the
where v0 is the velocity corresponding to E0 , the threshold energy of the principal sensitivity to the temperature has been factored out in the product
process-Le., ½m,v0 2 = E0 • The downward rate (j-+ i) can be obtained rt exp( - E0 /kT) while rlJ(T) is a slowly-varying function of r.
immediately on the basis of detailed-balancing arguments, for the electron Of course the main problem in application is to obtain reliable values of
velocity distribution is the equilibrium (i.e., Maxwellian) function; thus we QI1 • A characteristic difficulty for astrophysical work is that for many
must ha,,,C-- transitions of interest, kT « E 0 , so that the rate depends extremely sensi-
nrclJ = njC11 (5-69) tively ':1pon values of QI1 near threshold. Unfortunately, for E ~ E0 a great
computational effort is required to obtain accurate cross-sections because
from which it follows that the number of downward transitions is the simplifying approximations that are valid for E » E0 break down, and
because complicated variations of Q11 result from resonances in the collision.
n1C11 = n1(nifn1)•c11 = n1(ni/n1)•n,qlJ(r) (5-70) process. When values for QI1 can be ~btained, one typically fits them by
numerical procedures to simple analytical approximants, against which the --.J
As was the, case for radiative transitions, it is sometimes useful to introduce 01
integration in equation (5- 74) can be performed analytically.
the net collisional bracket Y,1 and write the net rate for collisions i -+ j(E1< E ) For the astrophysically important spectra of H, He I, and He II, accurate I
~ J
experimental cross-sections exist for excitation and ionization from the
n1C1Jfi1 = n1C11 - n1C11 = n1CIJ[l - (nJlnj)(nr/n 1)] (5-71) ground state. For transitions arising from excited states one must rely upon
The actual cross-sections e111 required to compute rates are found either theoretical calculations. For many atoms and ions of interest there may be·
experimentally or by rather complicated quantum-mechanical calculations· no detailed estimates whatever available, and one must have recourse to
it would ~ake us too far afield to describe these methods here, so we mere); rough methods to estimate rates. A very useful (though quite approximate)'
refer the interested reader to (410). There exists a vast literature containing expression for excitation rates in radiatively permitted transitions can be
results (theoretical and experimental) for a variety of transitions of astro- written (639) in terms of the oscillator strength f1J, namely
physical interest; bibliographies of this literature are issued from time to C I1 = C0 n,Tt[14.5Jji(/8 /E 0 ) 2]u0 exp(-u0 )f,(u0 ) (5-75)
time by the Information Center of the Joint Institute for Laboratory Astro-
physics of the University of Colorado and the National Bureau of Standards. where u0 = E0 /kT, 18 is the ionization energy of hydrogen, and for ions
(This center also maintains current literature references in an on-line com-
puter.) As indicated by equation (5-68) we are more directly interested in (5-76)
rates for a given cross-section, so let us examine q11 in a bit more detail.
Usually cross-sections are measured in units of na/, where a0 is the Bohr The parameter ?J is about 0. 7 for transitions of the form nl -+ nl', and about
radius; i.e., we write alJ = na/Q11 . Also, Q11 is usually tabulated in terms of 0.2 for transitions of the form nl -+ n'l', n' '# n (9S). For neutral atoms:
the energy of the exciting particle, so writing ½mv 2 = E, and substituting r,(u 0 ) has a different form [see (47)]. It is worth stressing that equations
5-4 The Non-LT£ Rate Equations 135
134 The Equations of Statistical Equilibrium

(5-75) and (5-76) provide, at best, rough values and should be applied with In particular, for an ion of chemical species X and charge Z, we consider
caution. In particular, collisions are not restricted by the dipole transition processes of the type
selection rules 6.1 = ± 1, and cross-sections for other values of I'll may be (5-80a)
x+<Z>(n, I)+ e(E, [" + 1) =; x+<Z-ll(n', I+ 1; n", I")
as large as for I'll = ± 1 despite JIJ being zero in the dipole approximation.
For collisional ionizations there exists a semi-empirical formula (402) followed by the stabilizing transition
O'iK(E) = na/[2.5W8 /Eo) 2
] ln(E/E 0 ){1 - b exp[ -c(E - E0 )/E0 ]}/(E/E0 ) x+<Z-ll(n', I+ 1; n", /")-+ x+<Z-l>(n, [; n", I")+ hv (5-80b)
(5-77)
which leaves the ion (Z - 1) in a bound excited state. As an example, for
which yields a rate
He+ we might have
C,K = Con,T*[2.5WHIEo) 2 ]uo[E 1(u 0 ) - be<u0 E1(u 1 )/u 1 ] (5-78) '
He+(ls) + e(E, I" + 1)-+ He 0 (2p; n"/")
where ,. b, and c are empirical quantities fitted to individual atoms, and He 0 (2p; n"I") -+ He 0 (1s; n"l") + hv
u1 = u0 + c. An alternative approximate formula can be obtained by ex-
pressing the collisional ionization cross-section in terms of the photo- If we denote the doubly excited state by d, the final bound state of the ion
ionization cross-section (73, 374) which yields a rate (334, 121) (Z - 1) by b, and the ground state of the ion Z as "• then the number of
dielectronic recombinations to state b from d can be written as n,.R4b = n4A,
C1,. = 1.55 x 10 13 n,r-tg,a(v 0 )exp(-u0 )/u 0 (5-79) where A, is the spontaneous transition probability for the stabilizing emis-
where a(v0 ) is the threshold photoionization cross-section, and g1 is of order sion; to a good approximation (particularly for large n") A, = A(n', I + 1; n, I)
0.1, 0.2, and 0.3 for Z = 1, 2, and >2, respectively (here Z is the charge on Jo,· the Z ion. In the limit of low radiation fields, the reverse process in equa-
the ion). The same caveats expressed about equation (5-75) apply to equations tion (5-80b) can be ignored, and if A. measures the transition probability
(5-78) and (5-79) as well. I for autoionization, 11 4 can be written (73, 258) in terms of its equilibrium
population 11:
AUTOIONIZATION AND DIELECTRONIC RECOMBINATION
nd = n: A./(A. + A,) (5-81)
---J
In complex atoms with several electrons, the ionization potential is nt = nKn,(gd/g,.)C 1 y-t exp(-XdK/kT) = nKn,<l>,.d(T) (5-82) (J)
where
determined by the lowest energy to which a sequence of bound states with
I
only one excited electron converge (to the ground state of the ion plus a free Here X.dK is the energy of state d above the ionization limit.
electron). If two electrons are excited within the atom, they will, in general,
give rise to states with energies both below and above the ionization potential Exercise 5-6: Derive equation (5-82) by applying the Saha equation (5-14) between
defined above. Subject to certain selection rules [(172, 371; 297, 173)] the the continuum ,-: and bound state b, and the Boltzmann equation (5-5) between
states above the ionization limit may autoionize to the ground state of the states b and d.
ion plus a free electron. The inverse process is also possible and, if an ion
in the ground state suffers a collision with an electron of sufficiently great Thus from each state d to each state b we have the number of dielectronic
energy, then a doubly excited state of the atom may be formed. In general, recombinations
this process will be of little interest because the compound state will imme- (5-83)
diately autoionize again (typical autoionization transition probabilities A
are in the range 10 13 -10 14 !), and its equilibrium population will be small~
In some cases, however, a stabilizing transition occurs in which one of the As in the case of radiative recombinations, it is often useful to define a di-
two excited electrons (usually the one in the lower quantum level) decays electronic recombination coefficient °'oR such that n,.Rdb = nKn,aoR· Note
radiatively to the lowest available quantum state, leaving a bound atom that the ratio of numbers of dielectronic to radiative recombinations depends
with a single excited electron. This process can provide an efficient recom- only on th~ ratio °'oRl°'RR• and hence is independent of density and is a
bination mechanism referred to as dielectronic recombi~ation. function of temperature only.
5-4 The Non-LTE Rate Equations 137
136 The Equations of Statistical Equilibrium

Dielectronic recombination plays an important role in stellar atmospheres produced by the radiation field in the stabilizing transition. If stated is char-
in two contexts. First, we can calculate a total rate for dielectronic recom- acterized as (n', I', L') and state b as (n, I, L) we have the total dielectronic
bination by summing over all states (n, /) and (n", I"); as was shown in a recombination rate
classic paper by Burgess (118), this process becomes extremely important at
high temperatures. For example, at temperatures T ~ 106 °K, the dielec- nKRdb = nKne L <I>n'l'L'(T)A(n', /', L'; n, I, L)[l + (2hv 3/c2)- 1J.TL,]
n', I', L', L
tronic recombination rate for He+ exceeds the radiative recombination rate
by a factor of 102 but drops below radiative recombination for T ~ 105 0 K. (5-84)
Burgess convincingly demonstrated that dielectronic recombination is the
dominant recombination process in the solar corona (where T ~ 2 x 10 6 °K where J i~ the mean intensity in the stabilizing t.ransition. Often the double
and n~ ~ 108), and that this mechanism establishes the coronal ionization excitation 'state is so broad (because of the very short lifetime against auto-
balance. In calculations for the total rate one must sum over vast numbers ionization) that the radiation field used to compute J can be fixed at the
of states, and the most important contributions come from states with n" » n' continuum value. If the L' dependence of Xn•, ,.. L' ls small, then one can define
and l" » 1 + 1. It is because these high states have large values of XdK that A*(n', I'; 11, /) = g1• - 1 LL', L g(L')A(n', 1, L'; n, 1, L) and replace <I>n'l'L' with
high temperatures are required to overcome the exponential factor in <I>•• ,, to obtain
equation (5-82) and produce significant dielectronic recombination [i.e., a
large value of kT is required for the electrons to have sufficient energy for nKRdb = nKne L <1>•• 1.(T)A*(n', I'; n, /)[l + (2hv 3/c 2 )- 1J•. ,.J (5-85)
the reaction in equation (5-80a) to occur]. In the sum one encounters a n', I'
divergence problem at large n", similar to that found for partition functions,
unless one includes both probabilities A, and A 0 in equation (5-83) and takes while the inverse (upward) rate b -+ d becomes
into account the fact that A, will dominate A 0 for sufficiently large n". The
calculations of total rates require the estimation of large numbers of stabili- nbRbd = nb L B*(n, I; n', /')]•. ,, (5-86)
n', I'
zation transition probabilities and collision cross-sections (to calculate A 0 by
detailed balancing arguments), and are at best difficult; an approximate where B*(n, /; n', I') = A*(n', I'; n, l)(g1,c2/2hv 3g1).
general formula that provides tolerably accurate values for most ions of
......J
interest has been developed (119).
......J
A further study (121) has shown that in the corona the effects of inverse COMPLETE RATE EQUATIONS
transitions in equation (5-80b) produced by incident photospheric radiation
I
are small, and that the doubly excited states can be destroyed by collisional Having examined all of the processes of .interest, we may now assemble
ionization if densities are sufficiently high, from which one concludes that the individual rates into a single complete equation of the form of equation
dielectronic recombination does not play an important role in the deeper (5-52) for each bound state i of each ionization stage of each chemical species
layers of the atmosphere (e.g., in the solar chromosphere and photosphere). in the material. We shall (a) ignore explicit mention of dielectronic recom-
A second situation where dielectronic recombinations are important arises bination because the rate has the same form as for radiative recombination;
for some ions (e.g., C III and N III) that have low-lying double-excitation and we shall assume that both are included; and (b) assume that all ioniza-
states (XdK < kT) that feed free electrons into selected bound states. A tions from bound states of ion j go to the ground state only of ion j + 1
striking example is afforded by the 2s2p( 1 P 0 ) 3d state of N III, which lies (generalization is easy but complicates the notation and discussion). We then
only 1.6 eV above the ionization potential to N IV 2s 2 1S, and which feeds may write
electrons directly into the N III 2s 2 3c/ levels and thereby produces the
famous N III ,1.4634-40 (3d-+ 3p) emission lines in Of-stars (115; 429; 440).
In such cases one finds that A0 » A., so that nd is given by its equilibrium
value (relative to actual ion densities), equation (5-82); furthermore one need K
sum over only a few states. On the other hand, these processes occur deep L n1.(nJn1,)*(Rl'i + Cw) = 0 (5-87).
enough in the atmosphere that one must account for the inverse transitions i'>I
5-4 The Non-LT£ Rate Equations 139
138 The Equations of Statistical Equilibrium

where the radiative rates are defined by equations (5-66) and (5-67), the or hydrogen, yielding Mu = Lu + 1 slates in all (including protons). Then,
collision rates by equation (5-68), and the LTE population ratios by equations using X's to denote nonzero elements, the rate matrix .9' has the form
(5-5) and (5-14) for bound states, and bound and free~ates, respectively. One
such equation may be written for each bound state. have one more vari- Row
able (the ion density n") than we have equations. If wrote down an Number
ionization equation I XX X X 0 0
2 ."( ."(
L n,(R," + c,I() - L (nifn")*(R"' + C11() = 0
l<K
""
i<K

we would find it to be redundant with the set (5-87).


(5-88)

Exercise 5-7: Show that equation (5-88) results from summing equation (5-87)
.'( ."( X X
0 0
0 0

L0 + I xx· \_ ......X 0
over all bound states. ·

We therefore invoke an additional physical constraint to complete the system.


For an impurity species (i.e., r:t.t/rx.H « 1), we close the system by demanding
that the total number of atoms and ions (of all ·kinds) of the species equal
0
XX
..
X .'I: •

I I
X

I
X

I
0
-y -y· . -}'
0

-y
the correct fraction of the number of all hydrogen atoms (including pro- I I I
tons); i.e., 0 X x· X X

(5-89) X x· X X

Alternatively, we can close the system by invoking charge conservation


(saving the total number conservation for use elsewhere) and write
0 0 0 X x· X X
--..J
L LiN1t + nP = n, (5-90) 0 . 0 I I I 2 0 O· 0 I (X>
k J I
L0 MH~
where N1t = Li niJ•· The final system, for all levels of all ions of all species
Column Number·
is written in the general form
(5-91)
The first L 0 rows correspond to equation (5-87) for He 0 , the next L+ rows
where n denotes a vector that lists all occupation numbers (say .JV of them) give equation (5-87) for He+, the M He th row gives the abundance equa~ion ·
whiled is an (.JV x .%) matrix and~ is a vector in which only one element (5-89), the next LH rows give equation (5-87) for H, and the last row gives
is nonzero [from equation (5-89) or (5-90)]. charge conservation. The vector n consists of elements
To make these. considerations more definite, let us consider a case that is
simple enough to be manageable and complicated enough to be of general n = [n 1(He 0 ), ••• , n1.o(He 0 ), n 1(He+), ... , nL.(He+),
applicability. Suppose we have an atmosphere composed entirely of hydro- n(He + +), n 1(H), ... , nL"(H), nPY (5-92)
gen and helium (of abundance Y, by number, relative to hydrogen). We
consider the helium to consist of a ladder of three ionization stages, He 0 , and ~ = (0, ... , 0, n,)r (5-92)
He+, and He++, and we suppose that these ions have L 0 , L +, and I levels
respectively. Further we write MH• = L 0 + L+ + I; i.e., MH• is the total For given values of n,, T, and the radiation field, equation (5-91) is a linear sys-
number of helium states of all kinds. Similarly we consider LH bound states tem in n, and may be solved by standard numerical methods (526, Chapter 9).
5-5 The Non-LTE Equation of State 141
140 The Equations of Statistical Equilibrium

LTE. as expected. Two comments are necessary here, however. (a) To


5-5 The Non-LTE Equation of State
obtain LTE in a multilevel atom J. must equal B. in all transitions. If any
From the results of the preceding sections, we see that in LTE each occupa- transition is transparent, then LTE will not be obtained (unless densities are
tion number at a specific point in the atmosphere is a function of only two so high that collisions dominate), not only for the particular levels involved
thermodynamic variables: i.e., n1 = n1(N, T) where Tis the absolute tempera- in the transition under consideration, but actually for all other levels as well
ture at that point. In contrast. in the non-LTE case. the full rate equations bec~·1use the radiation field in each transition influences the populations of
imply that n1 = n1(N, T, J,.) where J. denotes the frequency dependence of all le els (see below). (b) We have left unanswered the question of how large
the radiation field over the entire spectrum and Tis now a kinetic temperature is "ve rge" optical depth. As we have indicated earlier, t. ~ 1 is not
describing only the particle velocity distribution function. We now have as sufficient to guarantee J,. -+ B •. Rathert,. must exceed a therma/ization depth,
many new (fundamental!) thermodynamic variables as are required to specify for which precise estimates will be given in Ch~pters 7 and 11. In the low-
the distribution of radiation in frequency. [Note that if we could simplify the cfensity limit (e.g., in a nebula), equation (5-94) reduces to
description of this distribution-e.g., if we could write J. = WB,.( T )-then
the situation would be vastly simplified; but in general we may need to
consider perhaps hundreds of new variables.] As was the case for the LTE
(n 1/nT) = f.: (rx,.B..fhv) cfv / f.: (a.J ./hv) dv (5-95)

equation of state, the non-LTE statistical equilibrium equations are actually


nonlinear in the electron density n,, and we shall require a linearization which states that, if the recombination rate exceeds the photoionization rate,
procedure to solve for the occupation numbers; but now we shall have to the level is overpopulated; and it is underpopulated if the reverse is true.
extend the linearization to include changes in the radiation field as well. We Equation (5-95) is, of course, equivalent to equation (5-46) which is often
6
shall see in §7-5 that this approach provides a method for coupling the applied in nebular analyses. In the coronal case, we have T,( ~10 °K) »
transfer equations and statistical equilibrium equations together, and allows TR( ~6000°K), which implies that collisional ionizations exceed radiative
us to detennine the global response of the gas to the radiation field simul- [see equation (5-44) and related discussion] while radiative plus dielectronic
taneously with the reciprocal response of the radiation' field to material recombinations, both of which proceed at a rate specified by T,, exceed
properties. - collisional recombinations. Then
Before developing the linearization procedure required in the general case,
it is worthwhile to consider a few examples that illustrate clearly the essential
physical content of the statistical equilibrium equations.
(5-96)
so that
LIMITING CASES That is, the coronal ionization balance depends only on temperature and is
independent of the electron density, a fact that vastly simplifies analysis of
Consider first an atom consisting of a single bound level that can ionize
the corona. Both the coronal and nebular situations represent extreme
to its continuum. We then have one rate equation which states that (ignoring
stimulated emissions) departures from LTE. .
Let us now consider some multilevel problems. Suppose we have a volume
of pure hydrogen gas illuminated by a very dilute radiation field (i.e., a
(n 1/nT) = [ 4n J.: (a.B./hv)cfv + n,q 1" ] / [4n J.: (a,.J.fhv)cfv + n,qlK] (5-94) nebula). We anticipate that virtually all of the hydrogen will be in its ground
state, and we assume that all the resonance lines are completely opaque (and
We note first that, as the electron density becomes very large, so that colli- hence in detailed balance). Further, we assume that, after an atom is photo-
sional rates exceed the radiative rates, then ionized from the ground state, recombinations occur to all states, but the
populations of the upper states are so small and the incident radiation field
lim (nifnT) = lim (n,q 1"/n,q 1") = 1 so diluted that (a) we can ignore photoionization out of these states, and
"• ... x: "• ... -:c (b) electrons in any excited state cascade downward at rates determined by
i.e., LTE is recovered. Further, at very large optical depth, J. -+ B. and the Einstein coefficients A1•.. without reabsorption upward (i.e., the subordi-•
clearly n 1/nT -+ 1; i.e., if the radiation field is perfectly Planckian we recover nate lines are transparent). We further assume densities are so low that we
142 The Equations of Statistical Equilibrium 5-5 The Non-LT£ Equation of State 143

may neglect collisions. Then we have an ionization equation degraded from high energies (say far ultraviolet) to low (visible and infrared);
for example, in a nebula, Lyman continuum photons are degraded-e.g.,
I
into Balmer continuum photons plus Loe photons (state 1 = Is, state 2 = 2p,
n,RIK = n, L ocRR(i, T)
2
(5-97a)
state 3 = continuum). We may calculate the ratio R 1 ... 3 ... 2 ... i/R 1 ... 2 ... 3 ... 1
l=l
quite easily. The number of excitations I ..... 3 is n1B 13 W B(v 13 ). Of the excited
and a number conservation equation atoms in state 3, a fraction A 32 /(A 32 + A31 ) decays to state 2, and of the
I atoms in state 2 a fraction A 2 i/[A 21 + B23 W B(v 23 )] decays to state 1 (here
L n1 + n, = nH (5-97b) we have ignored stimulated emission). Thus
l•I

where nH is the (given) hydrogen density,/ is the total number of bound states (5-99)
considered, and n, = nP (for pure hydrogen). R 1" is assumed given in terms
of J, = W B,(TRl, as in equation (5-40). For any subordinate state we can By similar reasoning
, , terms of ,the branching ratios aJI -= AJI /~
calculate the population, in L,I < J AJh
an d t he cascade probabrl1t1es p1I which are defined recursively as p
d - ~)- I • • I+ I. I
= (5-100)
a;+ 1, 1, an P11 = a11 + LJ=I+ 1 PJka., for (J = 1 + 2, ... , /). Then for level i
we find
I so that
n, L A 11 = n,2ocRR(i, T) + )>I
l<I
L A1In1.

= n.2 [ ocRR(i, T) + f:
}>I
p1,ocRR(j, T)]
'
(5-98)
But using the relations among the Einstein coefficients and writing B, in the
Exercise 5-8: (a) Verify the expressions for p11 given above and derive equation Wien approximation (hv/kT » 1) one finds [B 11B(v1i>f A11 ] = (n1/n,)•, so
(5-98). (HINT: Start with level I and work downward.) (b) Show that equations equation (5-101) reduces to R 1... 2 ... 3 ... i/R 1... 3 ... 2 ... 1 = W < I, which proves the
(5-97) and (5-98) yield a quadratic equation in n. that allows the determination or theorem. The result clearly follows from the fact that in the cycle 1 ➔ 3 ➔
n.(nH, T), and hence all the n1's. (c) Show that p11 = 1 (J > 1); interpret this. 2 ➔ 1 the dilution factor enters only once, while in the reverse process it CD
result p~ysically. enters twice. In stellar atmospheres, Rosseland's theorem is relevant because 0
at certain depths one may have resonance lines that are opaque (i.e., W = 1) I
From equation (5-98) we may estimate ratios of occupation numbers and exciting atoms to upper states, from which the subordinate lines are trans-
hence ra!ios. of lin~ _intensities along a series. For example we can co~pute parent; in such cases we anticipate a systematic photon degradation.
the relative mtens1t1es of the Balmer lines (the Balmer decrement) as
LINEARIZATION
/(Ht)//(H1) = (ntAuhvu)/(n1A12 hv12 )
As mentioned before, the general system dn = 1A can be solved as a
~nd com?are the theoretical results with observation. The approach outlined linear system for n if n,, T, and J, are all specified. But, in practice, we do not
m equations (5-97) and (5-98) (with extensive elaboration and refinement!) know exact values for these variables in the course of a model-atmosphere
forms the basis forthe analysis of nebulae [ see (15, Chaps. 23-25; 10, Chap. 4; computation (recall the discussion for the LTE equation of state) but have
415, pp. 40-110; and 350, Chaps. 1-3)]. only current estimates in an overall iterative process. We expect all of these
. Fina~ly, consid~r an atom that consists of three states (1, 2, 3) in order of variables to change by amounts bn,. '5T, '5J., etc. to satisfy better the con-
1?creasmg energy ma rarefied medium (neglect collisions) and a dilute radia- straints of energy and pressure balance, and must evaluate the response of
tion field. t:,- famous result regarding such a system is Rosseland's theorem of n to these changes, in the form
cycles, which states that the number of radiative transitions in the direction K
I ..... 3 ➔ 2 ➔ 1 exceeds the number in the inverse direction 1 ➔ 2 ..... 3 ..... 1.
A consequence of this result is that energetic photons are systematically
6n = (8n/8n,) bn, + (8n/8T) '5T + L (8n/iJJk) Mk
k=I
(5-102:)
144 The Equations of Statistical Equilibrium 5-5 The Non-LTE Equation of State 145

Here Jk(k = 1, ... , K) is the mean intensity at discrete frequencies that from which we find
sample the spectrum. These frequencies are chosen such that all integrals
over frequency are replaced by quadrature sums-i.e., a;= [(iJ.91/iJJk) · n],

(5-103)
= (4nwk/hvd {rJ>i
o:ij(vk)[n, - n1'ndn1)"'e-h••/kT]

(5-106)
We obtain equation (5-102) by linearization of the original equations (5-91)
~and also can find parallel linearized equations that give, in essence, oJk(liT,
i5n., 6n) ~rom _the transfer equations; cf. §7-5]. Equations of the form (5-102)
We then can construct
are required m two contexts: (a) model atmosphere calculations where all (iJn/iJJk)1 = - L d,;.1am (5-107)
variables may change in an iteration cycle, and (b) multilevel statistical '"
equilibrium calculations for a given model (ne, T, and total particle density
It is particularly instructive to consider the case where only one (i -+ j)
fixed). Th~ procedure for case (b) will be deferred until Chapter 12, and we
shat~ consider onl~ case ~a) here. (In case (b) one may use a special technique transition can absorb at vk, and all the other o:,m(vd's are zero.
motivated by consideration of the computational methods of solving transfer
Exercise 5-9: Show that, for the case just described, equation (5-108) is valid.
equations, to be developed in Chapter 6.]
If x denotes any variable, then by linearization of equation (5-91) we have (iJn/iJJd, = (.~,"j 1 - d 11 1 )(4nwko:;i{vk)/hvk](n; - nj(n;/n1)"'e-h••lkT]
(5-108)
1
(5-104) This result, besides being simple, shows clearly that, because .91- must in
general be a full matrix, a change in the radiation field at any frequency vk
causes a change in the occupation numbers of every level /, even if I cannot
where we have assumed that n is the solution of the current system dn = Bl 1
absorb or emit photons of that frequency. Of course the elements dji, and
(we might introduce a subscript zero, or some similar device but it would .91 11 1 may be small, and the coupling weak, but the basic physical point
become unwieldy). An extremely important feature of this ap~roach is that (X)
remains true. Similar results for (iJ/iJne) and (iJ/iJT) are given in the references
every derivative in equations (5-102) and (5-104) can be written down
analytically (though th~ inverse .91- 1 must be computed numerically); this cited.
produces a system_ of high accuracy and reliability. To illustrate the proce-
dure, we shall write down some representative derivatives for the model
atoms discussed at the end of §5-4; more comprehensive collections of for-
mulae are given in (42) and (437). In what follows, we use the auxiliary vector
~ = (8.91/iJx) · n. Suppose we choose some frequency vk, and wish to calculate
cn/iJJ k· Except for the "special" rows of .91 that express abundance and charge
conservation, we will generally have

(5-105a)

(iJ.91/iJJk)u = (4nwk/hvk)[ L o: 1j(vk) + L o:11 (vk)(nJin 1)"'e-hl-•lkT]


J>I J<I
(5-105b)
and (iJ.91/iJJk)IJ = -[4nwko:IJ(vk)/hvk](n;/n1)"'e-h••1kr, (j > i) (5-105c)
6-1 Iteration: The Scattering Problem 147

6-1 Iteration: The Scattering Problem

6 One of the fundamental physical difficulties inherent in the solution of


transfer problems is the existence of scattering terms, which decouple the
radiation field from local sources and sinks, and involve global transport
of photons over large distances in the atmosphere. It is through these terms
that the presence of a free boundary makes itself felt even at great depth (t. »
Solution of the 1) in an atmosphere, and causes large departures of the mean intensity J. from
local values of the thermal source function B•. For ease of discussion we
consider a prototype source function that contains a thermal emission
Transfer Equation component and a coherent isotropic scattering term-i.e.,

where p. = <J ./(K. + <1 .). The solution of the standard transfer equation

µ(oI ,/at.) = 1. - s. (2-36)

can be written formally in terms of J. as [ cf. Exercise (2-10)]

If there were no scattering, p. = 0, then J• could be calculated, as a quadrature,


from B,; when p, '=I= 0, we must solve an integral equation for J,. One of the
first methods that comes to mind to effect such a solution is iteration. As we
know that J, -+ B, as t.-+ ex:,, let us deal with (J. - B.). Suppose p, were
The analysis of actual stellar spectra requires the calculation of the emergent everywhere zero; then (J, - B,) would equal (B, - B,) where B.(t,) =
CD
flux from a .model atmosphere by a solution of the transfer equation. In this A,.[Bv]. If p, is not zero, we could regard this value as a first approximation I\)
chapter we theref~re turn attention to numerical techniques for solving and write I
(J. - B.)< 1 > = (B. - B,) + A,.[p,(J, - B,) ]
0
transfer problems m terms of differential equations; we shall find that two
extremely general, flexible, and powerful methods result when we formulate = (B. - B.) + A,.[p.(B. - B.)]
the tr~nsfer equation as a tw?"'point boundary-value problem using difference
equations. Many methods exist for solving the transfer equation in terms of = (B. - B,) + 6< 1 > (6-1)
integral equations, but these will not be discussed in depth in this book as
they are adequate!~ described elsewhere ( 18, Chap. 8) and as they do not l~nd Then by iteration, we find
themselves so readtly to the treatment of moving atmospheres (Chapter 14). "
In the present chapter we shall restrict attention to static, one-dimensional (J, - B,)'") = (B, - B,) + L 6(n) (6-2)
l=I
plane-parallel atmospheres; more general problems will be considered in
Chapters 7 and 14. There are strong physical and mathematical motivations where 6<•> = A,.[p. l!,<•- 1 >]. In practice we continue the iteration until some
for using the techniques presented in§ 6-3 (or their integral-equation convergence criterion-e.g., 1161">/(J, - B.)'">ll ~ e, where e « 1-is satisfied.
analogues), which are specially designed to overcome certain difficulties It is clear that, if IIP,11 « 1, the iteration procedure of equation (6-2) can be
that are characteristic of radiative transfer in optically thick media in the expected to converge, for successive corrections !!,<•> must be of order IIP,11"
P:esenc: of scatter!ng _terms. We shall try to develop insight into these relative to (J. - B,). If, however, IIP,11 ::::: 1 over a large depth of the
d1fficult1es by constdermg other plausible, but useless, approaches to the atmosphere, the iteration method fails.
problem in §§6-1 and 6-2.
148 Solution of the Transfer Equation 6-1 Iteration: The Scallering Problem 149

The circumstance just mentioned actually occurs in stellar atmospheres, where the second equality follows from the form of B, assumed in equation
and the thermal coupling parameter l, = 1 - p, may be very small through- (6-3). The solution of equation (6-6) is
out a large part of the atmosphere. For example, in very hot stars the principal
source of continuum opacity in the outer layers is electron scattering, and J. - B, =ix.exp[ -(3l.)t-r.] + {3. exp[ +(3l,)t-r.] (6-7)
i.• may be of order 10- 4 very deep into the atmosphere (until finally, as the
density rises, free-free thermal absorption overwhelms the electron scat- As we demand that J. -> B. as t,. -> oo, we must have /3. = 0. To evaluate
tering). In cool stars of low metal abundance, the hydrogen is neutral in ix, we make use of the boundary condition J,(0) = J3
H .(0) = (dJ ,/dt,)o/J3
[the second equality following from equation (6-5) in the Eddington approxi-
the upper atmosphere and free electrons are scarce, so Rayleigh scattering
by H and H 2 dominates the H- opacity, and p, is nearly unity until great mation]. We thus find from equation (6-7)
depth (at some point the hydrogen rather abruptly becomes excited and
J ,(0) = a. + ix, = (dJ ,/dt) 0 /J3 = [b, .,.. ix,(3l,)½]/J3 (6-8)
ionized, and l, suddenly rises to unity). For lines, the corresponding thermal
parameters may be very small, l, ~ 10-s (see Chapter 11). Hence we obtain finally
The symptomatic behavior of the iteration method in these cases is that
the solution stabilizes, and although successive iterations differ fractionally J,(t,) = a. + b,t,. + (b. - J3 a.) exp[ -(3l.)t-r,]/[J3 + (3l,)½J (6-9)
only by some small value, the /l's are monotonic, and are nearly equal in
iteration after iteration. In such cases, although the fractional change per Equation (6-9) reveals the essential physics of the problem. First, it shows
iteration is e (e « 1), there is no guarantee that, say, 1/e more iterations may that J. may be markedly different from B, at the surface. For simplicity,
not actually be required to reach the final solution. The discussion thus far consider an isothermal atmosphere-Le., b. = 0 and B. = a,; then at t, = 0,
has been couched in terms of integral equations using the /\-operator, but J,(0) = J,.½a,./(1 + l}) = l}B./(1 + l}). Thus when l, « l: then J, is
it should be stressed that the same difficulties would arise with a similar much smaller than B, at the boundary. Second, we see that this departure
iterative solution of the transfer equation as a differential equation (we shall, extends deep into the atmosphere because the slow decay of the. exponential
in fact, refer to either procedure as "/\-iteration" even when we do not term implies that J,(t,)-> B,(t,) only at depths t, ~ (A.,)-t; in view of the
actually employ the /\-operator). The failure of /\-iteration to converge is small values quoted above, these are very large depths indeed. When J, has
a point of crucial importancewhose physical significance must be understood approached B,. arbitrarily closely, we say that the solution has thermalize~;
completely; to this end we may consider the following simplified analysis. we therefore refer to J,.-t as the thermalization depth (a concept that will
Suppose that the depth-variation of the Planck function can be represented be generalized in Chapters 7, 11, and 12). co
with sufficient accuracy by a linear expansion We may obtain an intuitive understanding of the thermalization depth w
from the following physical argument. The parameter ).• = K,./(K, + a,) I
(6-3) clearly is just the probability that a photon is destroyed (i.e., converted into
thermal energy) per scattering event. To assure thermal destruction, the
and that p,. is constant with depth. The zero-order moment of the transfer photon must be scattered about n = 1/A., times. If the photon progresses
equation can be written, using equations (2-71) and (2-39) through the atmosphere by a random-walk process, with mean free path 6.r
(which must be approximately unity), then the total optical thickness through
(6-4) which it may pass without destruction is nt 6.r = 6.r,1,,,-t ~ J, -t. Photons
while the first-order moment gives emitted at greater depths are unlikely to escape without being thermalized
(hence J • .... B.), while those emitted from shallower depths manage to
(oK./o-r.> = H. (6-5) escape and allow J. to fall below the thermal value (i.e., B,).
We now can understand why /\-iteration fails when we adopt J,. = B,.
If we use the Eddington approximation K,, = ½J. and substitute equation as an initial estimate. Each successive iteration can propagate information
(6-5) for H,. into equation (6-4) we obtain about the departure of J,. from B. only over an optical depth M ~ 1-:-i.e.,
a mean free path [recall that E 1(6.t) falls off as e- A•/6.t for M » I]. Thus
l 2 we must perform of order A., -t iterations to allow the effects of the boundary
2
3 (o 1.;a-r. > = l,.(J,. - B.> = 3l [a 2(J. - B.)/o-r.2] (6-6) to make themselves felt in the solution to a thermalization depth. When
6-3 The Transfer Equation as a Two-Point Boundary Value Problem 151
150 Solution of the Transfer Equation
ascending and descending exponentials. Therefore, in general. the terms in
.l.. « 1 such a procedure becomes computationally prohibitive, and we exp(kt) will be present; these are called parasite~, and they_ increase a~ a
conclude that any useful method must account for the scattering terms in rate of order exp(2kt) relative to the true solution. Thus 1f our starting
the source function from the outset and provide a direct solution for such values are wrong by an error e, the parasite will be of order e exp(2ktm•~) ':"
terms. lOktm., compared to the true solution at the other boundary, an_d it _is
6
obvious that, unless our initial choice is very good (e « 1), th~ parasite ~ill
swamp the true solution, which will then be lost. I_n f~ct, to retam any vesug~
6-2 Eigenvalue Methods of the real solution, we must employ n ::::: ktmax significant figures. if se~er:o
an le- uadrature points are used, some µ, « 1 an_d hence s?me » •
A characteristic mathematical difficulty that emerges in treating the transfer g ~th a moderate t ~ 10 we will Jose the solution on typical computers.
equation as a differential equation arises from the nature of the boundary even WI max ~ 3 4 • h J' h' ·h
At t ::::: 1 in the continuum, tmax may be ~ 10 to. 10 m t e mes, w _ic
conditions. Suppose we use the metho_d of discrete ordinates, replace the shows the hopelessness of this approach. In summary, the mathematical
angular integral for J. by a quadrature sum, and attempt to integrate structure of the problem requires that we employ a metho~ _that accounts
numerically the system explicitly for the two-point nature of the boundary conditions from the
1 • outset. We now turn to a discussion of such methods.
L ai1 -
µi(d/ 1/dt) = 11 - -
2
p
}"' -· (1 - p)B, (i = ± 1, ... , ± n)

To effect the integration, we require starting values for I,. for all values
(6-10)
Exercise 6-1: (a) Solve equation (6-10) with p =: 0, B = cons~, for It. with
_ ½ Sh that d2 J /d-c2 = 4(J - B) and wnte exact solut10ns for J, I+,
of i; these are fixed by the boundary conditions. As described in Chap- ~id~ ~ c~lcul~~ng constants of integration from boundary condit!ons. Suppose
-d h I ( ) - / (-c ) = B· evaluate the (false) solution and show
ter 2, the boundary conditions fall into two groups, namely / 1(0) = 0, one h a C osen + 'Cmox - - max ' 'fl B t
that the error & = B exp( -2tmax) at the tower boundary amph es to & = a
(i = -1, ... , -n), for incoming rays on the range -1 ~ µ1 ~ 0 and
the surface. (b) Generalize the discussion to the case where p i' 0 (but constant).
/ 1(tmax) = g(µ1) [e.g., g(µ) = B,], (i = 1, ... , n), for outgoing rays on the
range 0 ~ µ1 ~ 1. Here tm■x refers to the deepest point actually treated
in a semi-infinite atmosphere. The problem is this: suppose we wish to
start the integration at t = 0, and proceed step-by-step inward; we cann0t, 6-3 The Transfer Equation as a Two-Point
for we do not know the values of / 1(0). Similarly at tmax we lack values for Boundary Value Problem · Cl)
/ _,(tmax),· ~
Thus we face an eigenvalue problem of order n. We could, for example, In this section we shall derive two very general, flexible, and powerful I
guess a set of values for/ _,(tmax) and use these to integrate toward the surface. approaches for solving transfer problems. These approaches r~sult fr?m
When the integration reaches the surface, we would in general find/_ 1(0) ,t, 0. 't' the transfer equation as a second-order differential equation subject
In principle, we could then adjust the values of/ -,(tmax), and by successive ~:\;;.point boundary conditions. Most of the basic ideas were presented
trials find those values that forced / _,(0) = 0. In practice, however, this in an important paper by Feautrier (209). These methods _have proven to be
method is strongly unstable and can work only if tmax is not very large. stable and easy to implement; each offers advantages m c~mplementary
We can see this as follows. As we know from the grey problem, the discrete ranges of the parameters that set the scale of the computational effort to
ordinate method leads to exponential solutions of the form exp(±kt) where solve a given problem.
the k's are of order 1/µ. In cases where the coefficients (such asp,) are depth-
variable, the solution no longer consists of pure exponentials, but, neverthe-
SECOND-ORDER FORM OF THE EQUATION OF TRANSFER
less, still has an exponential character, perhaps /(t) exp(±kt) where f is
a weak function of t. In a semi-infinite atmosphere we must suppress the In plane-parallel geometry we may write two equations governing the
ascending exponentials. For the grey problem this can be done explicitly, outgoing and incoming radiation field at ±µ:
for we have an analytical form with which to work. But in the nongrey
variable-coefficient case, the solution is known only numerically, and unless ±µ[ol(z, ±µ, v)/az] = x(z, v)[S(z, v) - /(z, ±µ, v)] (6-11)
exactly the right choice of starting values is made, it contains both the
152 Solution of the Transfer Equation 6-3 The Transfer Equation as a Two-Point Boundary Value Problem 153

where we restrict µ to the half-range 0 ~ µ ~ I. We now define symmetric Note that in contrast to the moment equations, which do not close, equa-
and antisymmetric averages tion (6-17) [first derived by Feautrier (209)] yields exact closure of the system
in terms of the angle-dependent symmetric average u,.,. We shall see below
u(z, µ, v) = 21 [/(z, µ, v) + l(z, - µ, v)] (6-12) that it is sometimes advantageous to follow an intermediate course and to
use an approximate closure of the moment equations in terms of variable
1 Eddington factors.
and v(z, µ, v) = [/(z, µ, v) - /(z, - µ, v)] (6-13)
2
BOUNDARY CONDITIONS
which have, respectively, a mean-intensity-like and a flux-like character. In
terms of u and v we can construct a system of two first-order equations by Equation (6-17) must be solved subject to boundary c?nditions at t =. 0
adding the two equations (6-11) to obtain and at , = rmax [ which denotes the thickness (o~ hal_f-th1ckn~ss) of a fimt_e
slab, or a great depth where the diffusion approx1mat1on applies for a semi-
µ[ov(z, µ, v)/oz] = x(z, v)[S(z, v) - u(z, µ, v)] (6-14) infinite atmosphere]. At r = 0, /(0, - µ, v) = 0 which implies that v,..(O) =
u,.,.(0) so that
and subtracting them to obtain µ(ou,../ot,.)o = u,.,.(0) (6-20)

µ[ ou(z, µ, v)/oz] =- x(z, v)v(z, µ, v) (6-15) At r = rmax• we specify /(tm.., + µ, v) = I+(µ, v), and write v,..(tmax) =
Then substituting equation (6-15) into (6-14) we can eliminate v and obtain I +Cµ, v) - u,.,(rmax) so that
a single second-order system for u: (6-21)

1 ou(z, µ, v)]
a [ -(-) '
2
µ
-(-)-;-
X Z, V uZ X Z, V
,,
uZ
= u(z, µ, v) - S(z, v) (6-16) If the diffusion approximation is valid at tmox• then

or, defining d,, = d,(z, v) = - x(z, v) dz and abbreviating the notation, /(tmax• µ, v) = B,(tmax) + µ ( Xv1 laB,·I)
Tz •m,. (6-22)
ro
(6-17) 01
In writing equation (6-15) we have assumed that Sis symmetric in µ; this I
will be true for most of the source functions we shall consider-e.g., those of
(6-23)
the form
f
S, = a, <p,,J ,. dv' + {J, (6-18)
Exercise 6-2: (a) Generalize equation (6-20) when 1(0, - µ, v) # 0. (b) Show ·
or S, = a, f R(v', v)J,. dv' + {J, (6-19) that for a symmetric slab (infinite in x and y), of finite thickness (in z) tm••• _th~ lo"".er
boundary condition can be written at t = ½tm•• as (i!11.Jct,) = 0. This 1mphes
but may not be true if the redistribution is angle-dependent [in which case that we need consider only half the slab: 0 ~ t ~ ½tmax·
other techniques are required, cf. (460)] or if there are motions in the atmo-
sphere (see §14-1). In equations (6-18) and (6-19) the cx's essentially stand for
DIFFERENCE-EQUATION REPRESENTATION
scattering coefficients divided by the total opacity and the /J's represent
thermal terms. It must be stressed that these choices of are purely illus-s. We now convert the differential equation (6-17) into a set of difference
trative, in the sense that we shall later (cf. §§7-2 and 7-5) find similar-looking equations by discretization of all variables. Thus we choose a set of de_Pth
terms that involve the radiation field over the entire spectrum (imposed by a points {rd},(,/ = 1, ... , D) with r 1 < r 2 < · · · <. r 0 ; a set of angle points
radiative equilibrium constraint) or for the entire transition array for a multi- 'µ \ (m - l M)· and a set of frequency points {v.}, (n = 1,, .. , N).
\ mi• - ' ••• ' ' • I b
level model atom. The analysis given below still applies in such cases. For any variable g, we write g(zd, µm, v.) = 9dmll' We replace integra s Y
154 Solution of the Transfer Equation 6-3 The Transfer Equation as a Two-Point Boundary Value Problem 155

quadrature sums-e.g., for equation (6-18) we write where S41 has the form of equation (6-24) or (6-25), and indeed can be gener-
N M alized still further (to include, e.g., the entire spectrum; cf. §§7-2 and 7-5). As
sd. = ad. L a.<f>dn L bmudmn + /Jdn (6-24a) indicated, there is one such equation for each angle-frequency point i, at
n• 1 m• 1 each of D - 2 depth points.
Further, we group angles and frequencies into a single serial set of values If we now define the vector o4 , of dimension /, to consist of the angle-
subscripted i such that (µ,. v1) = (µm, v.) at i = m + (n - l)M, and hence frequency components at depth-point d-i.e., (o4) 1 = ud,then equation
reduce (6-24a) to (6-30) can be written as a matrix equation
I (6-31)
S41 = ad, l'•L1 w,,<f>41•U41• + f141 (i = 1, ... , /) (6-24b)
The(/ x /) matrices Ad and C4 are diagonal and contain the finite-difference
Similarly, for equation (6-19) we have representation of the differential operator. 8 4 is a full matrix that has the
differential operator down the diagonal plus off-diagonal terms that come
I from the quadrature sum representing the scattering integrals in equations
S41 = (Xdl I'•L1 aid, I', ,udl' + f141 (i = 1,' .. , /) (6-25) (6-24) and (6-25). Ld is a vector containing the thermal source terms. More
accurate difference representations than equation (6-30) may be written
Note in ~assing _th~t these s~urce functions are independent of angle, and using spline colocation (374), (442) or Hermite integration formulae (34),
hence this description contains redundant information (which can be re- but these do not change the general form of equation (6-31) (though~ and
moved :,v_hen we introduce variable Eddington factors). Equation (6-24b) has Cd may become full).
an add1t1onal redundancy because the scattering integral is independent To complete the system, we use the boundary conditions. At the surface
of v (?r of i); we shall exploit this later in Rybicki's methoa of solving the we could write
equations. ' (6-32)
Further, we replace derivatives by difference formulae, and write, e.g.,
which is only offirst-order accuracy; second-order accuracy can be obtained
(dX/dt)d+½ ~ (AXd+t/Atd+½) = (Xd+ 1 - X4)/(t4+ I - t4} (6-26) (30) from the Taylor's expansion u2 = u1 + Att(du/dt) 1 + ½At/{d2u/dt 2) 1, en
which, using equations (6-17) and (6-20), yields 0)
and 2 2
(d X/dt )4 .~ [(dX/dt)d+½ - (dX/dt)d-tJ/D<Atd+½ + At4-t>] I
µ 1(u 21 - u11 )/Att,1 = u 11 + GAtt.ijµ,)(u 11 - S 11 ) (6-33)
(6-27)
thus, defining
or, in matrix form
(6-28) (6-34)

1 Similarly, equation (6-21) at the lower boundary becomes


and At4, I = 2 (Atd-½, I + At4+½, ,) (6-29)

we rewrite equation (6-17) as


µ1(uDI - Uo-1, 1)/Ato-t, 1 = / 01 + - uDI - GAto-t, ,/ µ,) (uDI - S01 )

(6-35)
which, in matrix form is
(6-36)
(6-30)
Note that A 1 = 0 and C0 = 0.
156 Solution of the Transfer Equation 6-J '/1,e 'l'r,111.1/i•r /:'q11ario11 "·''" '/im-i'oi11t JJ011111/{lry Vaiut' /'rob/em 157

Exercise 6-.J: Derive equations (6-33) and (6-35), specialize the latter to the dif- has proven itself to be very stable, and has many desirable properties. Note,
fusion approximation using equation (6-23). for example, that at depth the system tends to become diagonal (the terms
in 1/l::,:r 2 ..... 0) and hence Jd --> Sd, as expected; in fact, we find J ..... S +
THE FEAUTRIER SOLUTION µ 2(d 2S/d-c 2 ), which recovers the diffusion approximation automatically. The
depth-discretization is commonly taken to give equally-spaced steps in
The set of equations (6-31), (6-34), and (6-36) have the overall structure log,, usually with 5 or 6 steps per decade of -c; such a choice has the advantage
that at different frequencies with widely differing opacities (e.g., a line-core

0 vs. nearby continuum) one has a reasonable distribution of depth-points.


We can estimate the amount of computing time required in a given prob-
lem by counting the number of multiplications. needed to solve the system;
the solution of a linear system of order n requires O(n 3) operations, so the
time required by Feautrier's method is TF = cD/ 3 = cDM 3N 3 where Dis
the number of depth-points, M the number of angle-points, and N the num-
Uo-1 Lo-1
ber of frequency points. It is clear that one pays a penalty for any unnecessary
Uo Lo redundancy in the angle-frequency information, and that the representation
(6-37) of these variables must be economized as much as possible. If we have a
Each element i_ndicated is either an (/ x /) matrix or a vector of length /; coherent scattering problem, N = 1, M is generally small, and Feautrier's
the grand ~atnx has a blo_ck _trid~agonal structure, and the solution proceeds method is optimum. However, in other problems the number of frequencies
by a~ efficient for~ard-ehmmat1on and back-substitution procedure (209). can be large because we must satisfy the constraint of radiative equilibrium,
~n this scheme_ we m e~ect express each u4 in terms of ud+ 1 and substitute or statistical equilibrium in several transitions; but the angular information
mto the following equation. Thus from equation (6-34) we can write is essentially unnecessary because only J., not u,.., enters these constraints.
We therefore eliminate the angular information by introducing variable
u, = B, - 1c1u2 + 8 1 - 1L 1 = D 1u2 + v1 (6-38) Eddington factors J. = Kv!J. (44).
By integration of equation (6-17) overµ we obtain
Substituting equation (6-38) into equation (6-31) for d =2 yields u2 =
D2u3 + v2 where 0 2 = (B 2 - A2 Di)- 1C2 Cl)
(6-42)
and ---.J
and the boundary conditions yield I
We therefore have in general
(6-43)
(6-39)
where D4 = (B4 - A,0 4 _ i)- 1C4 (6-40) and o<{::,·lj tmall
=~ (~l !··j)
Xv
0
tin11x
(6-44)

and v, = (B, - A404_ i)- 1(L4 + A4V4_ i) (6-41)


where h,. = H,.(0)/J .(0). Equations (6-42) through (6-44) may be differenced
for d = 1, ... , D. Starting at d = 1, we compute successive values for 0 4 in the same way as the angle-dependent equations, but the solution of this
and V4 through d = D - 1. At the last point, d = D, C0 = 0, hence system requires a time of only T,. = cDN3, which represents a considerable,
Do = 0, and Uo = vD [which still follows from equation (6-41)]. Having saving. To solve these equations we must know the depth-variation off-.'.
found uD we then perform successive back-substitutions into equation (6-39) at all frequencies. We proceed as follows. (a) From any given S, (e.g.,
to find U4, (d = D - 1, ... , 2, 1). Having found u4mn, we may then evaluate s. = B. as a first estimate) we can solve equation (6-17) for one angle u,..
J dn = r.!f.. 1 bmudmn• and the source function, which involves frequency in- and frequency at a time. In matrix form we have T 1u1 = S1 where T is tri-
s,.
tegrals of J .: e.g., Ln•
= 11.dn w•. <f,4•. J dn' + P,.. . diagonal, and u1 and S1 represent the depth-variation of u41 and S41 respectively ..
The. forward-backward sweep described above accounts explicitly for Solution of a single tridiagonal system of order n requires O(n) operations, so.'
scattering terms and the two-point boundary conditions. Feautrier's method the time required to evaluate the full angle-dependent radiation field for.
158 Solution of the Transfer Equation 6-3 The Transfer Equation as a Two-Point Boundary Value Problem 159

givens. is T,. = c'DMN. (b) Given u4mn we then calculate depth, and to solve depth-by-depth; in that method we may treat a fully

hn = ~ bmµm 2 Udmn /L bmudmn


frequency-dependent source function [e.g., equation (6-25)] with partial re-
distribution, but the computing time scales as the cube of number of frequency
points. In a beautiful paper, Rybicki (543) pointed out that, in the most
and h. = L,n bmµmU1m.lL b,,,ulmn· Note that even if the radiation field is commonly considered case of complete redistribution, much of this frequency-
known only with modest accuracy, the Eddington factor may be determined dependent information is redundant, for to specify the source function
with substantially better precision (e.g., if u is in error by a scale-factor, f is (equation (6-24)] we need only the single quantity J = f </J.J, dv. In a
still correct). (c) Now, given 1'4., we solve equations (6-42)-(6-44) for J. penetrating analysis, Rybicki showed how the solution in this case could
usi~g exflicit expressions of the form of equations (~-24) and (6-25) for s. be reorganized to yield a system of as great power and generality as the
(written m terms of J,). We then re-evaluate S, using the new values for J •. original Feautrier method but with very favorable computing time require-
(d) Because S, found in step (c) differs from that used in step (a), we iterate ments.
steps (a)-(c) to convergence. If L is the number of iterations the total com- Instead of describing the frequency variation of the radiation field at a
puting time is T,; = L(cDN 3 + c'DMN) « cDM 3 N 3 for moderate L. Ex- given depth, let us instead reverse the grouping and work with vectors that
perience with this method for a very wide variety of physical regimes in describe the depth-variation at a given frequency. That is, we now define
stellar atmospheres has always shown extremely rapid convergence (L U1 = (u11, U21, • • • , Uo1f (6-45)
usually is 3 or 4), and substantial economies (about a factor of ten) are
achieved. Finally, we note that additional equations can be added to the where i denotes a particular angle-frequency point. Similarly let
transfer equations at each depth-point d; these arise from other physical
constraints-e.g., statistical, hydrostatic, or radiative equilibrium (see §7-5).
J = (Ji, J2, · • •, Jof (6-46)
The basic form of equation (6-31) remains unaltered because these con- Then at angle-frequency point i, equations (6-30), (6-33), and (6-35) yield a
straints involve information only at one or two depth-points at a time. system
Thus if we have C constraints the total computing time becomes'T,; = (6-47)
L[cD(N + C) 3 + c'DMN]; this result bears on the question of whether it
is advantageous to use Feautrier's solution or Rybicki's solution, which we where T, is a (D x D) tridiagonal matrix representing the differential operator
shall discuss next. at frequency i, U 1 is a diagonal matrix containing the depth-variation of
the scattering coefficient (ix41 in equation (6-24)], and K1 is a vector that en
Exercise 6-4: This exercise requires access to a digital computer (of small capa- contains the depth-variation of the thermal term at frequency i. We have en
city). (a) Write a computer program to perform the formal solution of the one set of equations (6-47) for each angle-frequency point. In addition we I
transfer equation with a given S, for u,.., one angle at a time as described above, have D equations that define J4 , namely .
and to evaluate the variable Eddington factors at all depths. Use equally-spaced I
steps in A log T starting at T = 10· 3, up to T = 10 (Sor 6 steps per decade), and L wi'<Pdl'udl' - Jd = 0, (d = 1, ... , D) (6-48)
use a double-Gauss angle quadrature (4, 921); experiment with the number of I'= I
angle-points M to examine the sensitivity of the Eddington factors to the quadra-
ture. (b) Write a computer program to solve equations (6-42) through (6-44) Considering the grand matrix composed of all angle-frequencies and all
with given Eddington factors, assuming coherent scattering-i.e., S, = a.J, + p. depths, we have for the overall structure
Integrate the two programs and study the convergence of the iteration process
in cases with r,. = (1 - e), P = e, e « 1, starting with J, = 1, for e = 0.1,
0,01, 10-•.
0
T,
T2
'
U1
U2
U1
U2
K1
K2

THE RYBICKI SOLUTION

As we have seen above, the Feautrier solution organizes the calculation


0\\ T1 u. U1
=
K1
(6-49)

in such a way as to group all frequency information together at a given v. Vz ... v. E J p


160 Solution of the Transfer Equation 6-3 The Transfer Equation as a Two-Point Boundary Value Problem 161

where the V's are (D x D) diagonal matrices containing the depth variation of T 1 is markedly less costly than any other approach for generating A (34);
of the quadrature weights and profiles in equation (6-48). E would be the put another way, one may use integral-equation techniques if one wishes,
negative identity matrix and P would be void for equation (6-48) but, as but one should do so by means of Rybicki's method for generating A1.
we shall see in §7-2, more general systems of the form of equation (6-49)
arise in the computation of LTE model atmospheres. Comparison of equa- Exercise 6-5: Using a digital computer, write a program to solve the transfer
equation by Rybicki's method for a coherent scattering source functions. = (1.J,. +
tions (6-49) and (6-37) reveals that, in essence, the inner and outer structure
of the equations has been interchanged. p for the same values of e as were used in Exercise 6-4. Note that the Rybicki method
does not show its advantage here because only one frequency-point is involved.
The solution of system (6-49) is very, efficient. We reduce each "row" of
the grand matrix by solving for Finally, let us mention the effects of constraints in Rybicki's solution. For
1 1
u, = (T,- K,) - (T,- U1)J (i = 1, ... ' /) (6-50) each constraint that introduces essentially new information into the problem,
one requires an additional new variable similar to J, along with its defining
Then, multiplying by V1 and subtracting ftom the last "row" we cancel the equations. For example, in a multiplet problem (see §12-3) one requires a J
"element" in the ith column to zero. Carrying out this procedure for all for each independent transition, and in problems where one has introduced
values i we obtain a final system for J, namely WJ = Q where the full (D x D) the full set of statistical equilibrium equations by linearization a new variable
matrix Wis is required for each level of the model atom or every line in the transition
I array (see §12-4). If we have a total of C variables describing the constraints,
w = E- I v,T,- 1 u 1 (6-51) then each U matrix must consist of C diagonal (D x D) matrices side by
l=l
and the vector Q is side while each V matrix consists of C diagonal (D x D) matrices stacked
I into a column, and E becomes a matrix of dimension (CD x CD). In this
Q =p - L V,T,- 1
K, (6-52) case the computing time for a direct solution becomes
/st

The final system for J is solved; this yields sufficient information to calculate TR = c(D 2 • M · N · C) + c'(DC) 3
S (the run of the source function with depth). If desired, the full angle- for C » 1 this value exceeds the corresponding value for TF, and at first
frequency variation of the radiation field may be reconstructed using the sight Feautrier's method looks more attractive for dealing with systems
already-available quantities T1- 1 K, and T,- 1u1 in equation (6-50). involving many constraints (which is why we shall apply it in §7-5 for non-
The solution of the / tridiagonal systems in equation (6-50) requires LTE model construction). Nevertheless, for statistical equilibrium calcula-
en
0(D 2 /) = 0(D 2 MN) operations, and solution of the final system requires tions, Rybicki's method has been applied successfully even for large values
co
O(D 3 ) operations, so the overall computing time becomes TR = cD 2 MN + of C by using an iterative solution of the overall system (cf. §12-4).
I
3
c'D (NOTE: these e's are not numerically equal to those in the formulae
for TF, TE, etc.). Unlike the Feautrier system, in which the computing time COMPUTATION OF THE FLUX
scales as the cube of the number of angle-frequency points (i.e., M 3 N 3 ),
Rybicki's method is linear in MN. It is obvious that Rybicki's method is To compare with observations, we must calculate the emergent flux.
vastly more economical than Feautrier's (even with variable Eddington This may be done in a variety of ways. If Feautrier's method is used with
factors) when a large number of frequency-points is required. Recall, how- variable Eddington factors, h. is available, and hence H.(0) = h.J.(0) can
ever, that Rybicki's method works only if s. can be· written in terms of a be calculated directly. If Rybicki's method or the angle-dependent Feautrier
single quantity J in the scattering integral, while Feautrier's method works equations are used, we can calculate H,.(0) = Lm bmµmu(0, µm, v). Alterna-
for general scattering integrals. In principle one could use variable Eddington tively, having S,(t,) we can use the <l>-operator [equation (2-61)] to find
factors with Rybicki's method, but the advantage gained would likely be F,.(0) = <l> 0 [S(t,.)]; in practice this operation is done using a quadrature·
small (if any) because iterations would then be required. It should also be sum, for which several choices are available [see, e.g., ( 141; 246; 8, 33)]. If
emphasized that Rybicki's method is exactly equivalent to the integral the flux is required at points internal to the atmosphere one may apply the
equation approach in which one writes u1 = A1J + Mi> where the A matrix operator <l>, to S,., or one may compute v(td±½• µm, v.) from equation (6-15)
is generated by analytical integration of the kernel function against a set of and find Hd±t,n = Lm bmµmvd±½,m,n (note that this defines the flux at
basis functions representing J. In fact, T1- 1 is the A1 matrix, and inversion midpoints of the depth mesh).
7-1 The Classical Model-Atmospheres Problem 163

compared to the radius of the star, or (in 7-6) homogeneous spherical sh~/1s
7 when the thickness is an appreciable fraction of the radius. The assumption
of homogeneity makes the problem one-dimensional and thus greatly sim-
plifies the analysis; but it excludes many interesting phenomena involved in
small-scale structures seen in the solar atmosphere. For the stars we have
almost no information about the homogeneity of the atmosphere [see, how-
Model Atmospheres ever, (261, Chap. 11)] and we can only hope that one-dimensional models
yield some kind of "average" (in an ill-defined sense) information. However,
because the "averaging" process is nonlinear, the question is really an open
one and it is not at all clear whether such models always do yield meaningful
ave;ages, (e.g., in chromospheres), although they may be satisfactory for
some cases. In particular, in the solar atmosphere many of the inhomo-
geneities arise from hydrodynamic phenomena driven, ultimately, by the
convection zone; for stars without strong convection zones, the atmospheres
may indeed be homogeneous. (Counterexample: the Ap stars, which show
gross variations of physical properties over their surface, presumably asso-
ciated with the existence of strong magnetic fields).
(b) Steady state. We shall assume that the atmosphere is in a steady
state, and shall avoid discussion of all time-dependent phenomena-e.g.,
stellar pulsations, shocks, transient expanding envelopes (novae, su~rnovae),
heating by a binary companion, variable magnetic fields, etc. In this chapter
we consider only static atmospheres; in Chapters 14 and 15 we extend the
theory to steady flows (expanding atmospheres). We shall ass_ume that the
transfer equation is time-independent, and that level-populations are con-
stant in time and are specified by stati!ltical equilibrium equations (a special
7-1 The Classical Model-Atmospheres Problem:
Assumptions and Restrictions
case being LTE) that equate the number of atoms leaving a level by all co
microprocesses to the number that return. . 0
(c) Momentum balance. Having specified a st~ady state, we sha~l c?ns1~er I
The model-atmospheres problem refers to the construction of mathematical
either hydrostatic equilibrium in which the static gas pressure d1stnbution
models that provide a description of the physical structure of a stellar
just balances gravitational forces, or one-dimensional, lamin~r, steady flows.
atmosphere and of its emergent spectrum. In its greatest generality, the
Here we are ignoring the possibly large effects of magnetic forces: both
problem is one of enormous complexity, and presents both physical and
large-scale (as in the Ap stars) and small-scale (e.g., in sunspots or in the
mathematical difficulties that are beyond solution at the present time. It
concentrated knots of the general solar magnetic field). We further ignore
is therefore necessary to make a number of simplifications, and to deal with
the effects of small-scale motions such as waves, and larger scales such as
idealized models that are rather high-order abstractions from reality. Such
super-granulation flows, convective cells, etc., as well as major tidal dis-
abstractions are useful inasmuch as they enhance our insight without over-
tortions in close binaries.
whelming us with detail; yet it is important to state, at the outset, some of
(d) Energy Balance. Usually we shall assume that the atmosphere is in
the restrictions we have imposed, not only because this helps to define the
radiative equilibrium, which again implies that it is static; in §7-3 we shall
problem, but also as a reminder of the almost limitless numbers of fascinating
consider the effects of convection, but only in the roughest terms. In Chapter
research questions left to explore. The assumptions used in our work fall
into several broad categories: 15 we shall generalize to steady flow and include one-dimensional hydro-
dynamic work terms. The existence of c~mplicated motions in the solar
(a) Geometry. We assume that the atmosphere is composed of homo- atmosphere is well documented observationally [see, e.g., (694, Chaps. 9
geneous plane-parallel layers when the thickness of the atmosphere is small and 10) or (244, Chap. 5)] and, although data for stars are less complete. '
7-2 LTE Radiative-Equilibrium Models 165
164 Model Atmospheres

THE OPACITY AND EMISSIVITY:


there is little doubt that complex mass motions play an important role in
CONTINUA AND LINE-BLANKETING
the atmospheres of many stars (e.g., supergiants). But in its present state
the theory is unequipped to handle with full consistency the details of energy The frequency variation of the opacity and emissivity in stellar atmo-
exchange between the radiation field and hydrodynamic motions. Turbulent spheres plays a key role in determining the nature of the emergent spectru~.
dissipation in convection; wave generation, propagation, and dissipation; For example, the sharp decrease in flux shortward of about ..1.3650 A m
effects of shear in rotating atmospheres; magnetic field effects; and a variety A-stars can be explained by the large jump in the opacity caused by photo-
of other phenomena are all essentially overlooked! These are vital phenom- ionizations from the n = 2 state of hydrogen. Because the material becomes
ena, for without them we cannot account for chromospheres and coronae more opaque, we see less deeply into the atmosphere, and therefore receive
(in this book we shall approach these regions from a semiempirical diagnostic energy only from the outer, cooler, layers. We have alre_ady seen (Chapter _3),
view because we do not have an ab initio theoretical method). It remains that we cannot reduce the problem ofan atmosphere with a nongrey opacity
true that important limits on our understanding of stellar atmospheres are to the grey problem by any choice of average opacity, and we must, theref~re,
imposed by our inability to handle the intricate interchange of energy make allowance for the detailed frequency-dependence of the absorption
between radiative and nonradiative modes, and that development of a coefficient from the outset. At the very minimum we must treat the opacity
satisfactory theory to handle such interactions is probably the most vital variation in the continuum, which accounts for the gross features of the
research frontier in this field of astrophysics. energy distribution in the emergent spectrum; in more refined work we
It should be said, however (lest the reader receive an unduly gloomy must also include the effects of lines.
picture of our efforts to date), that progress has·been rapid, and continues The opacity at any given frequency contains contributions from_all possi~le
at an accelerating rate, so that we may reasonably expect at least some of transitions (bound-bound, bound-free, free-free) of all chemical species
the inadequacies of the present-day theory to be ameliorated in the near that can absorb photons at that frequency. From equations (5-53) and (5-60)
future. Moreover, the framework imposed above does appear to yield many we see that the direct absorption coefficient for process (i ..... j) from level
successful predictions of continuum features and line profiles for many i is n1o:IJ(v). Stimulated emissions return energy to the beam at a ~ate propor-
(perhaps most) stars. tional to 1v; hence (assuming identity of the emission and absorption profiles)
we correct the opacity by subtracting stimulated emissions from the absorp-
tivity. In view of equations (5-54) and (5-64), the correction is npu(v)G(v)
where G(v) = y;/g1 or G(v) = (nifn1)* exp(-hv/kT) for bound-bou?d or
7-2 LTE Radiative-Equilibrium Models bound-free processes respectively. Let nr denote the LTE population of
state; computed from the usual Saha-Boltzmann formulae (equation (5-14)]
......
In this section we develop the methods that can be used to construct planar, using the actual ion density. Then summing over all levels and processes
static, radiative-equilibrium models assuming LTE; the results of such cal- we have the non-LT E opacity
culations will be described in §7-4. As was discussed in Chapter 5, the assump-
tion of LTE vastly simplifies the computation (as one can see by comparing Xv =L L [n; - (gifg)n1]a1J(v) + L (111 - 11re·hv/kT)ct1,(v)
; j>; I
the methods of this section with those of §7-5). We criticized the use of LTE (7-1)
because it does not give an accurate description of the interactions of radia-
+ 'L, n,.n,a,K (.v, T)(l. - e -h,·/kT) + n,a.
tion and matter in stellar atmospheres, and is totally deficient in many

where the four terms represent, respectively, the contributions of bou~d-
important conceptual points (especially regarding line-formation). But on
bound, bound-free, and free-free absorptions, and of electron scattering
the pragmatic side, LTE models allows treatment of many effects (e.g.,
(other scattering terms-e.g., Rayleigh scattering-may also be added)._ To
line-blanketing) that are of importance in the application of the results of
calculate the spontaneous thermal emission (non-LTE) we use the rates denved
stellar atmospheres computations to the interpretation of photometric in-
dices, stellar temperatures and luminosities, etc., but that still lie beyond in equations (5-55) and (5-62) to write
the present capabilities of a non-LTE calculation. In a sense, then, the two
approaches are complementary: the non-LTE theory provides deep physical '7v = (2'1v3/c2)[L L nj(g;/g1)a1J(v) + ~ nfa;,(v)e•hv/kT
/ )>/ I
insight while LTE allows a preliminary assessment of complexities in the
models. Of course the end goal will be to have non-LTE models that are as + ,";;' n,n,a.K (v, T) e-h,•/kT] (7-2}
"refined" as any LTE model can be.
166 Model Atmospheres
7-2 LTE Radiative-Equilibrium Models 167

The three terms again describe bound-bound, bound-free, and free-free In addition to continua, the opacity of stellar material contains contribu-
process~s. Emission from continuum scattering terms will be written sep- tions from thousands to millions of spectral lines, both atomic and molecular.
arately m the transfer equations. Equations (7-1) and (7-2) apply in the non- Bound-bound opacities are significant for stars of all spectral types. For the
LTE case; if we assume LTE they simplify to earliest spectral types the resonance lines of H, He I, He II, and ions of light

x: = [r I
I }>I
nral)(v) + I nra,,c(v) + I n.n,c«,c,c(v,
I ,c
T)] elements dominate in the ultraviolet. For A-stars the hydrogen Lyman and
Balmer lines are important. For solar-type stars the important effects come
from lines of neutral and singly ionized metals and other atoms of moderate
x (1 - e-h•ftr) + n,u. (7-3) atomic weight. In later types, molecular bands (CN, CO, H 2 O, etc.) dominate.
and ,,: = (2hv3/c2)e-h•ttT Again, the chemical composition of the gas is a fundamental parameter in
setting the line opacity. In addition, parameters that determine linewidths-
x [r L
I }>I
nt«,1M + LI nta,,c(v) + L,c n,n,c«,c,c(V, T)] (7-4)
e.g., macroscopic velocities in the atmosphere (the so-called "microturbu-
lence"; see §10-3}-also enter.
w.. • • The effects of bound-bound absorptions upon a stellar atmosphere are
_ritmg K, = X, - n,u., we see that I'/: = K:B., as expected from the referred to as line-blanketing, and play a crucial role in determining both the
Kirchhoff-Planck relation [equation (2-6)]. emergent energy distribution and the physical structure of the atmosphere.
In LTE, the occupati?n n~mb7rs nr = nt(N, T), hence x: = x:(N, T) The temperature gradient in the atmosphere implies that, in the opaque lines,
and ,f:, = ,f:, (N, T),_ which s1mpl~fies the computation and allows great the layers from which we receive radiation will be cooler, and hence radiate
nu~bers of absorption and em1ss1on processes to be included easily. The less energy. The presence of numerous dark spectral lines within a given
basic_ free para~e~ers that enter t~e calculation are those describing the photometric band (established, e.g., by a filter) obviously directly affects the
chemical co'!'pos1t1~n of the material. In different spectral types, different measured flux. This effect is called the blocking effect. Because the total flux
absorbers w!ll dominate, depending upon the ionization aq_d excitation state in the atmosphere must be conserved, the flux blocked by the lines must
of the material. Thu~ for stars of solar temperature and cooler, th~ dominant emerge at other frequencies, and the energy emitted in the continuum bands
bound-free absorption process is from the H- ion· for A-stars it is from into which it is redistributed rises above the value it would have had in the
neutral H; in the B-stars He I begins to make significant contributions; in absence of lines. Furthermore, because the bandwidth of the spectrum in .
the 0-st?rs He II and numerous light-element ions (e.g., of C, N, O, Ne, Si) which energy transport occurs readily is restricted by lines, steeper tempera-
play ~n 1~portant role (101), (319). In the later-type stars a wide variety of ture gradients are necessary to drive the flux through; as a result, temper-
negative 1o~s of atoms and molecules contribute [(109), (73, Chap. 4), (644)]. atures in underlying layers rise, leading to the backwarming effect. Finally, (0
In general l~terally hun~reds or thousands of levels may contribute; only in the lines alter the temperature in the outermost layers of the atmosphere. We N
LTE can_ this much detail be handled, and even then extensive computations shall study these effects further in §7-4; it is clear from what has been said I
?r: required _[see, e.g., (504)]. Free-free absorption from He+, He and H that it is important to include bound-bound opacities in the calculation,
~s important m the O-stars; for the A-stars the main free-free contribution and we ask here how this may be done.
1~ from H; in th 7s~n it is from H-; and in the M-stars H 2 - free-free absorp- The most straightforward method of treating lines is the direct approach
tion becomes s1gmficant. Electron scattering is a major opacity source in in which one includes enough frequency points in the calculation to describe
!he 0-stars, while Rayleigh scattering by Hand H 2 contributes to the opacity the profiles of the lines under consideration. The full frequency and depth-·
m the atmospheres of intermediate-temperature stars (spectral types G and variation of the absorption coefficient is taken into account by this method,
K). Very elaborate and comprehensive opacity calculations have been made which may be applied when the spectrum is dominated by just a few lines.
by wo_rkers of the groups at Kiel and Los Alamos,· who have published The direct method suffers from the disadvantage that for many stars the
extensive graphs and tables of results [the reader should examine these line spectrum (e.g., in molecular bands that contain millions of lines) is so
carefully; see (651; 97; 638, 181-199; 184)]. The bulk of the Los Alamos complex that a detailed description is prohibitively expensive in computing
work pertains to stellar interiors, but some of the results are relevant to time, and we must therefore consider alternatives, which we may categorize
stellar atmospheres; a very complete discussion of the techniques used in as statistical methods. Here one attempts to replace the complicated fre-
these computations may be found in (14, Chap. 3). For continua the cal- quency variation of the line opacity in a given band (see Figure 7-1) by a
culations may be laborious, but are straightforward in principle. small number of parameters. The simplest possible description is to reduce
169
168

FIGURE 7-1
Schematic absorption coefficient or overlapping spectrum
FIGURE 7-2
Schematic opacity dis1ribution runction or the spectrum in
lines. A very large number or frequency points would be
Figure 7-1. A relatively small number or representative
required to describe the detailed variation or the opacity.
opacities sullice to describe this smooth distribution.

all the information to a single numbel', given by a mean opacity; in particular


we could consider either the Planck mean [ cf. equation (3-28)] or the Rosse- The main limitation of the opacity distribution function approach is that
land mean [cf. equation (3-26)]. As might be expected, this approach is not it implicitly assumes that the positions of the lines (in frequency) do not
really satisfactory (for precisely the same reasons it fails for the continuum). change markedly as a function of depth, measured in units of a photon
For example, while it is quite adequate to use the Rosseland mean at depth mean-free-path (i.e., unit optical depth in the continuum). It is crucial to the
where the diffusion approximation holds (as is done to include line absorption transfer problem whether a line in one layer of the atmosphere coincides in
in stellar interiors opacities), this method tends to underestimate the opacity frequency with a line or with a continuum band in an overlying layer, for
near the surface and yields poor approximations to the actual energy balanc~ photons might freely escape in the latter case, but not the former. Marked
there (126). The Planck mean fails to yield the diffusion limit for the flux at variations in the line spectrum, which invalidate the opacity distribution
(0
function approach, can occur in a number of situations-for example, the
depth, and. grossly overestimates the opacity at the surface; this leads to
substantial errors in predicted fluxes and temperature structure of the model. following. (a) Molecular bands of two species may overlap; one species may w
Recognizing the inadequacy of a single mean opacity we instead replace show a rapid decrease or increase with depth relative to the other. Even I
the detailed spectrum by a smooth opacity distribution function (which is a though the total opacity of the two bands together might not change, the
generalization of the classical picket-fence model described in §7-4), as in positions of the two sets of lines could be radically different. (b) A strong
Figure 7-2. We consider a narrow enough band to assure that the exact shock in the atmosphere might produce an abrupt change in the excitation-
position of the line is not important (i.e., other properties such as the con- ionization state of the gas over a small distance. The line spectra through th~
tinuum opacity or Planck function do not vary much over the band). We shock front might change radically. (c) Velocity shifts in expanding atmo-
then could, for example, find the fraction of the band covered by lines with spheres systematically move lines away from their rest positions; this strongly
opacity x,(v) > some chosen value X 1, and plot a graph of this fraction affects momentum and energy balance in the material (cf. §§14-1 and 15-4).
against X 1• The result is a smooth curve that can be well approximated by In such cases one must employ either the direct approach, or a generalization
a small number of subintervals (possibly of differing widths) containing of the statistical approach that in some way allows for the changes in the
constant opacities appropriate to the curve. This procedure may be carried frequency positions of the lines. An alternative approach, called the opacity
out for a mesh of temperatures and densities to produce a description of the sampling technique (based on a random-sampling procedure) has recently
variation of the line opacity through the atmosphere. A critical study of this been suggested (585); although this method appears computationally mo~e
approach (126) shows that opacity distribution functions yield excellent costly than the opacity distribution function method, it also appears that 1t
results, and reproduce both the emergent fluxes and physical atmospheric may not suffer from the limitations just described, and should be teste? _
structure given by detailed direct calculations to satisfactory accuracy. further. ·
170 Model Atmospheres 7-2 LTE Radiative-Equilibrium Models 171

HYDROSTATIC EQUILIBRIUM ter 15, for some O-stars the radiation forces on spectrum lines in stellar winds
exceed g and accelerate the material to very large velocities ( ~3000 km sec-
1
).
In a static atmosphere, the weight of the overlying layers is supported by
the total pressure, and it is this balance, in essence, that determines the
Exercise 7-/: Consider fully ionized stellar material of hydrogen and helium
density structure of the medium. Thus (abundance Y). (a) Show that n,r1,/p, which provides a lower bound on the
opacity, is r1,(1 + 2 Y)/mH(l + 4 Y). (b) Take advantage of the grey nature of
Vp = pg (7-5) electron scattering to show that, if gravity is to exceed radiation forces, then g
must be ;;i:g., 11 where g., 11 = r1,(l + 2Y)(r1RT:rrl/[cmH(l + 4Y)]. (c) Re-express
where the total pressure p = p, + PR (dynes cm- 2 ); the gas pressure
4
this result to show that the luminosity L of the star must be :i;;;L., 11 :::: 3.8 x 10
P9 =. NkT; the radiation pressure PR = (4n/c) fK. dv; g is the surface (,,./(/.II 0 )L0,
gravity _(regarded as a fundamental parameter describing the atmosphere).
Here p 1s the mass density (gm cm - 3) which, using the notation of 5-2, can For computational purposes we can rewrite (7-5) as a difference equation
be written connecting the depths specified by column masses m4 and m4 namely +"
P = (N - n.)mH L 0CtAt = (N - n.)ffl (7-6)
k N
N4kT4-N4-1kT4-1 +(4n/c) L Wn(hnJdn-fd-l,nJd-1,n)=g(m4-m4-1)
where mH is the mass of a hydrogen atom, and At is the atomic weight of n= I
chemical species k with fractional abundance 0Ct, If we define the column (7-9)
mass m (gm cm - 2 ) measured from the outer surface inward as our new in-
dependent variable-i.e., Here K. is expressed in terms of the mean intensity and a variable Eddington
dm = -pdz (7-7) factor; i.e., K. = J.J •. We can obtain a starting value from equation (7-8)
by assuming that the radiation force remains constant from the boundary
then we may rewrite equation (7-5) as dp/dm = g, which yield~ an exact surface upward, and thus
integral p(m) = gm + c. It is obviously advantageous to be able to write.
such a result, so we shall use m as the independent variable henceforth· the co
choice of m. instead of z has no significant effect on the transfer equa;ion. ~
Using equation (2-77b) for the radiation pressure gradient we can rewrite I
equation (7-5) in another useful form: Equations (7-5) through (7-10) are valid for both LTE and non-LTE atmo-
spheres.
(7-8) Notice that, if we knew the temperature structure T(m), and could either
(a) ignore radiation forces or (b) estim.ate them, using equation (7-8), as
(x./p)(aR r:rr/c) where x. is a suitable mean opacity, then we immediately
which shows that radiation forces tend to cancel gravitational forces, and could derive the density structure N(m). From this we could calculate
lead to a smaller pressure gradient in the atmosphere. Put another way, the
x:(N, T), ri:(N, T), solve the transfer equation, and thus determine all model
material tends to "float" upon the radiation field. As was shown in Chapter 1 properties of interest. Of course in general we do not know the temperature
the radiation force is related to the flux through the atmosphere, and we ca~ structure, and we must now address the issue of how it is to be determined.
thu_s see t~at_ for a given T.rr there. will be so°'!t lower bound on g below
~h1ch rad1at10? forces exceed gravity and blow the material away. Specif-
1cally, Underhill (633) showed that gravity forces will exceed radiation RADIATIVE EQUILIBRIUM:
forces only if g ~ 65 (T.rr/10 4 )4 cm sec- 2 • Clearly radiation pressure forces TEMPERATURE-CORRECTION PROCEDURES
~re negligible for the sun (T.rr ::::: 6 x 10 3 , g ::::: 3 x 10 4 ) but become very
important for an O-star (T.rr ::::: 4 x 104, g ~ 104 ) and in supergiants where For a given temperature distribution, the equation of hydrostatic equi-
g is quite low (and indeed approaches g0 , 11 ). In fact, as we shall show in Chap- librium can be integrated as described above, and opacities and emissivities:
172 Model Atmospheres 7-2 LTE Radiative-Equilibrium Models 173

derived. The radiation field then follows from a solution of the LTE transfer computed J,, = A,.[B,,(T0 )], hence the name of the method, and that
equation equation (7-12) is not satisfied. We then assume that the run of T(m) that
does satisfy the condition of radiative equilibrium is T(m) = T 0 (m) + 6 T(m),
82 (/.,J.)/8-r:.2 = J. - (11: + n,u.J,)/x: = (1 - n.a./x:)J. - (K:/x:)B. and require that
(7-11)
S K:B.(T
0
"" 0 + 6.T) dv = S0"" K:J, dv (7-14)
at all frequencies and depths using the techniques described in Chapter 6.
For an atmosphere in radiative equilibrium the total energy absorbed by Expanding B,.(T0 + 6 T) ~ B,,(T0 ) + 6 T(8B,/8T) we find
the material must equal that emitted, hence in LTE

4n f0
00
[11: - (x: - n,u,)J,] dv = 4n f
0
00
K:(B, - J,) dv =0 (7-12) !J.T::::: So"" K:[J,, - B,,(T0 )] dv/ S K:(aB,/8T)r dv
0
00
0
(7-15)

or, in discrete form, (and allowing for departures from LTE), It must be emphasized that J, in equation (7-15) denotes the value already
computed from B,(T0 ). If one carries through the process· and recomputes a
4n I wn[11" - (x. - n,u.>J"J =o (7-13) new model with the new temperature distribution, some improvement in
n
satisfying equation (7-12) usually will be found. However, the procedure
In radiative equilibrium the total flux 4nH = uR T!rr = constant, and we suffers from several severe defects.
may choose it (or Terr) as another fundamental parameter characterizing (a) BecauseJ, = A,.[B,(T0 )] = B,(T0 ) + 0(e-'•),itisclearthatatdepth
the model. Now in general we do not know the temperature distribution the temperature correction goes rapidly to zero, no matter how bad the
that produces radiative equilibrium, and using our present estimate of T(m) solution actually is at those points. We found a similar result in the grey
we will normally find that the radiation field does not satisfy equation (7-12) problem.
or (7-13). It is therefore necessary to adjust T(m) iteratively in such a way K:
(b) If the frequency variation of is such that the opacity is much larger
that the radiation field does ultimately satisfy the requirement, of energy (say several orders of magnitude) at some frequencies than at others, the
balance. The determination of T(m) is, in fact, the very heart of the problem method again fails. The reason is that in the opaque frequency bands the ,
of constructing LTE models. There are basically two strategies we may usi:: contribution to the numerator vanishes as -r, -+ 1 while the contribution to
(a) temperature correction procedures, and (b) solution of the transfer equation the denominator swamps that of all other bands. In effect the A-iteration (0
subject to p constraint of radiative equilibrium. In temperature correction pro- procedure is effective only over 6-r, ~ 1 for the most opaque frequencies.
01
cedures one attempts to use information about the radiation field calculated (c) .Equation (7-12) places a condition only on the flux derivatives; hence
I
from a given T(m) in an a posteriori fashion to estimate a change 6 T(m) that we have no way of specifying the actual value of the flux to which the solution
will cancel out the errors found in the flux and in the flux derivative [ equiva- converges (if it does). •
lent to equations (7-12) and (7-13); see equation (2-71)]. In the second ap- (d) The real failure of the A-iteration procedure is that it ignores the effect
proach, one attempts from the outset to formulate the transfer equation in that 6.T, computed at some depth -r, has on J,(-r') at all other depths (i.e., J,
such a way that the resulting radiation field will automatically satisfy radiative is presumed to be fixed). This oversight necessarily leads to spurious valu~s
equilibrium. The first approach (corrections) was historically the one orig- of 6 T. Actually J,.(-r~) = A,~[B,(T + 6 T)], which means we should, m
inally used to solve the nongrey atmospheres problem, and the methods are reality, solve an integral equation for 6 T; we shall return to this point later.
often quite ingeniously constructed. The second approach (constraints) is When the reasons for the failure of the A-iteration procedure were under-
more subtle and powerful, and overcomes inadequacies fatal to "correction" stood, it was realized that methods were needed that made use of information
procedures in the non-LTE case, thus allowing a deep penetration into about errors in the flux itself (which gives direct information about the
problems of considerable complexity. Ironically, the roots of the idea of temperature gradient at depth) as well as in the flux derivative. One method
using "constraints" are to be found in the methods used to solve the grey of doing this was suggested by Lucy (283, 93), who generalized the method
problem. We first consider temperature correction procedures. devised by Unsold for the grey problem (see §3-3) to the nongrey case. If
The first, and most obvious, method is the so-called lambda-iteration pro- we use a Planck mean [equation (3-23)] optical depth scale d-r = -Ki dz,
cedure. Here we suppose that from a given run of T 0 (m) we have, in effect, then exact frequency integrals of the moment equations (using quantities
7-2 LTE Radiative-Equilibrium Models 175
174 Mode/ Atmospheres

without subscripts to denote frequency-integrated variables) are and

(dH/dr:) = (,c1f,cJ)J - B (7-16)


T1 ={(I +ti) S x.(l
0
00 0
-p,.)[J. -B.(T0 )] dv
and (dK/dr:) = (xi/,cj)H (7-17)
- 3t(I - .1f/H 0 ) Soa) x.(1- p.)H .° dv
where ~1 is the absorption mean [ equation (3-32)] and xi is the flux mean
[eq~ation (3-21)]; note that the scattering coefficient is included in xi but 00

+ r'1 So"' [x~(I - pJ- x.p~][ J .°- B.(To)] dv} / So x.(1- p.)(8B.f8T)o dv
not .m the oth~r means. Then relating K to J with the Eddington approxi-
mation, ~quations (7-16) and (7-17) are combined to give an expression for (7-20)
B(t), which, tr~ated as a perturbation equation for a correction 6B(T) =
4o-R T 3 6 T/n gives finally where the prime denotes the derivative d/dt; quantities with superscript or
subscript zero denote current values; H 0 = f~ H.° dv; .1f denotes the
6B(t) = -d(6H)/dr: + (,c1f,cj) [3 s: <xi/,cj) 6H(r:') dr:' + 26H(O)] nominal flux (uT:rrf4n); and p. = o-./x. = u./(1<.. + u.). Equation (7-19) is
a linear first-order differential equation fort I that may be integrated straight-
(7-18) away starting with r: 1(0) = 0; given values for r: 1 and r'1 , one then computes
T from equation (7-20) and hence obtains T(t) = T 0(t) + T 1(t) at t =
Here ~H(t) = H - . H(r:). T~e first term on the righthand side of equation 1
t + t (t). Experience shows that the t 1 correction leads to revisions of the
1
(7-18) !s the correctio.n predicted by the A-iteration procedure; the other temperature scale at depth, while T I is most important at the surface.
terms introduce new mformation that gives nonnegligible values of 6B at Both the Unsold-Lucy and the Avrett-Krook procedures have been
depth and produces a response to flux errors at the surface. Experience has widely applied and have been proven quite successful in the construction
sho:,v~ the U~s.ol~-Lucy procedure to be quite effective in fOnstructing LTE ofLTE radiative-equilibrium models. With care in calculation, these methods
rad1at1ve-equihbrmm models (but it has no obvious generalization to the produce models with errors in l6F/Fl and Id In F/dtl of the order of 0.1
non-LTE case). · percent. But despite the fact that temperature correction procedures are
easy to use and work fairly well, they have a number of serious drawbacks
Exercise.7•2.: Deriv~ equations (7-16) and (7-17) and, applying reasoning simiiar that render them ineffective against non-LTE problems. We shall describe co
to th~t y1eldmg equation (3-44), derive equation (7-18). · these briefly here to motivate the discussion of the constraint-type procedures O>
which, although first developed to attack the non-LTE case, also are ex- I
Anot.her very clever and quite useful method of calculating temperature tremely effective for LTE models, and are now the preferred methods.
correc~ions was suggested by Avrett and Krook (SS), who introduce per- First, temperature-correction methods tend to stabilize rather than con-
turbations to both the temperature and the optical depth scale. That is, we verge ifthere are large variations in the frequency dependence of the opacity,
sup.pose that the curr~nt. te~perature distribution T 0(t) is related to the This failure is particularly troublesome when we attempt to construct models
desired. tempera~ure distribution T(t) (which yields radiative equilibrium) including spectral lines or a major continuum jump (e.g., the hydrogen
?Ya patr of relations: T = T0 + T 1 and t = t + r: 1. The transfer equation Lyman continuum). In such cases temperature correction procedures leave
1s then expanded to fi~st order in the perturbations t I and T 1, and by taking the temperature structure of the outermost layers essentially undetermined,
mom~nts of the r~sultmg first-order equation of the perturbation expansion, for the energy balance there is established entirely by these opaque transi-
equati~ns are derived for r: 1 and T 1. These equations [extended to allow for tions (the optically thin regions already have radiation fields that are fixed).
scattermg terms (421) and with an improved closure relation (351)] are Second, these methods tacitly assume that the temperature is the crucial
variable, and cannot deal effectively with cases where the radiation field is
r'1 + t1 (Jo"' x~H. 0
dv/ f
0
a) x.H.° dv)
only weakly coupled to the local thermal pool-e.g., in atmospheres where
scattering terms dominate or where non-LTE line formation occurs (we shall
see why this is so in §7-5 and Chapters 11 and 12). Finally, these methods
= (1 - .1f/H0 ) - rt f0 a) x.[J.° - B.(T0 )] dv/ f
0
00
x.H.° dv. (7-19) are not sufficiently accurate. Although errors of a tenth of a percent see"?
176 Model Atmospheres
7-2 LTE Radiative-Equilibrium Models 177

small, it must be realized that there is a close similarity in the requirements where the M 4, y represent the contribution of the interval (t 0 • Y' co) to the
of radiative equilibrium and statistical equilibrium, and that errors of this
size may be totally unacceptable in the latter context. Specifically, suppose
integral. Substituting equation (7-23) into (7-21), and (a) assuming K:
is
unchanged by /J. T, (b) writing By(T + !J. T) ~ B.(T) + (8By/8T) AT, and
we consider a ground-state continuum that dominates the opacity by orders (c) introducing a frequency quadrature {vn}, (n = 1, ... , N), we find a set
of magnitude. Then energy balance requires f: (Xy 1J y dv = f: (Xy I B, dv while oflinear equations for the values of AT4 :
the net radiative rate in the statistical equilibrium calculation is of the form
f~ (X![b1J, - B,.]v- 1 dv (where b1 denotes nifnT of the ground state). If
bi ~ 1, these two criteria differ only by the (weak!) "profile" function v- 1 •
In the limit that hv0 /kT » 1 and both B, and J y have a characteristic fre-
dt [JI WnK?,i8B/8T}4•i044• - Add'n)] AT4,

quency variation of exp( - hv/kT), both pairs of integrals are strongly


dominated by contributions from v ~ v0 , and hence one of them becomes = L w,,1,J,, [Md,. +
II
I (Add'" - 044•}B4•n]•
d'= I
(d = 1, ... , D) (7-24)
essentially redundant. Thus an error of a few tenths of a percent in energy
balance implies a similar error in the net radiative rate; because the radiative Exercise 7-3: Verify equation (7-24).
rates may exceed the collisional rates by orders of magnitude (recall the
discussion of §7-3), such errors may overwhelm all other terms in the rate The solution of this system yields the change in the temperature consistent
equation, and lead to false equilibration. We now turn to a discussion of with the global properties of the radiation field. Because our expansion of B,
methods that treat the condition of radiative equilibrium as a constraint; is only linear, the system has to be iterated to convergence by using the new
these overcome all of the shortcomings described above. temperatures to recalculate B., (8B./8T), IC:, etc., and re-solving the
The essence of the constraint approach is to· build the requirements of system; if assumption (a) is valid, we would expect quadratic convergence.
radiative equilibrium directly into the transfer equation in such a way as to There are some defects to this approach. (1) The computation of the A
solve both problems simultaneously. These methods deal directly with the matrix is cumbersome and costly, and (because IC: really is a function of T)
global nature of the radiation field; i.e., they account for the effect that a must be done again for each iteration. (2) It is possible to calculate the
temperature change at one point in the atmosphere produces upon the ra- response of the A matrix to changes in Ty (arising from changes in K:
caused
diation field at all other points, and vice versa. To simplify the discussio,n by the temperature change) but this again is extremely cumbersome and'
in the remainder of this subsection we shall ignore scattering terms in tqe costly (also, there are problems of stability) [see, e.g., (347; 575)]. The method
source function. One such procedure was proposed by E. Bohm-Vitense described in the final subsection of this section overcomes these difficulties. (()
(283, 99) who suggested that the integral equation for AT, (3) As originally formulated, and as described thus far, the method does not --..J
force convergence to a prespecified flux. This may be done by applying the I
f0
00
1C:By(T + AT) dv = f
0
00
IC:A,.[By(T + AT)] dv (7-21) diffusion approximation at the lower boundary and demanding the correct
flux transport (32). Thus, we write for Ty > to, y
be solved directly; we shall recast her discussion into slightly different terms
using current notation (see also 32). To solve an integral equation of this form
(7-25)
we first construct a matrix representation for the A-operator· we introduce a
discrete set of points {Ty, 4 }, (d = 1; ... ,
D), at which we wi~h to determine
~he solution, and we represent the variation of B.{ty) analytically in terms of
(7-26)
interpolation (basis) functions on this mesh. The integral

Integrating against µ, and over all frequencies, we find


(7-22)

is then calculated analytically for these basis functions to produce the system (7-27)
0
Jd,, = d'•L1 Add',YB,,,Y + Md,Y> (d = 1,,,,, D) (7-23) which fixes dT/dz in equation (7-25), and introduces the flux into the quanti-:
ties M 4•• ; note the similarity of this device to that used in the grey problem.
178 Model Atmospheres 7-2 LTE Radiative-Equilibrium Models 179

Exercise 7-4: (a) Evaluate the elements Add' by assuming a piecewise linear inter- number of frequencies, must be large (the angles can be eliminated in ter~s
polation for B,(t,) on a discrete grid; i.e., on of variable Eddington factors); if radiative equilibrium is the only constrat~t
[rd, td+ .], set B(t) = [BJ-rd+ 1 - t) + Bd+ 1(r - rd)]/(rd+ 1 - rd) involved, it is cheaper to use Rybicki's method, letting J denote the term m
the numerator of (7-28); this equation in effect, replaces equation (6-48).
(b) Evaluate Md using equations (7-25) and (7-27). Feautrier applied his scheme with good results for both LTE and non-LTE
Exercise 7-5: Using a digital computer, program the solution of equation (7-24)
continuum models. The basic drawback of the method is that it is not clear
for an opacity step function:"• = ,c 0 for (v :E;; v0 ), "• = o:,c0 for v > v0 ; parame-
terize the problem in terms of the value of p = hv 0 /kT.rr, which specifies the how to generalize it, as it focuses entirely on the temperature correction (which
frequency of the step. Notice that the A matrix and M vector are independent of is not sufficient in general).
frequency (though different) in the two ranges (v :E;; v0 ), (v > v0 ); hence the
integrals over frequency of 8, and iJB,/iJT can be done analytically and expressed Exercise 7-6: With the help of a computer, use the method just described to
in terms of elementary functions [by using the known result for the complete solve (a) the grey problem for q(t); start with q(t)= C and try several values for
interval (0, oo) and appropriate expansions for (0, v0 ) when p « I and for (v 0 , co) C; (b) the opacity-step problem of Exercise 7-5 [cf. (128; 605; 38)] .. In part
when p » I] or in terms ofDebye integrals (4,998). Solve the problem for several (a) use a quadrature for the frequency integral (not the exact results-this would
values of oc and p, starting from the grey temperature distribution (on the Rosscland make the problem trivial) and Rybicki's method for solving the final system.
mean scale); compare your results with those in (128; 605; 38).
An alternative method that also uses second-order difference equations
A second constraint procedure was suggested by Feautrier (283,108; 210). for the transfer equation was proposed by Auer and Mihalas (38); this method
Noting that .radiative equilibrium implies that is very easily generalized to extremely complicated problems, and forms the
basis of the methods described in the final subsection of this section and in
L WnK.1nJdnfL WnK.1nBdn =1 (7-28) §7-5 for non-LTE problems. If the temperature structure T*(m) of the atmo-
n n
sphere were precisely that which produced radi_ative equilibrium, and _B:
at all depths d, he solved the transfer problem [ equation (6-30) or (6-42)] with the corresponding Planck function, then the solution of the transfer equation
the source function
o2(f.J.)/ot.2 = J, - B: (7-30)

(7-29) with lower boundary condition [using equation (7-27)]


co
0)
where the J's are regarded as unknowns. Note the conceptual similarity (7-31) I
between this approach and that used to solve the grey problem! In contrast
to the integral operator approach described above, Feautrier's method is very
easy to formulate and solve using the difference-equation procedures de- would automatically satisfy the condition of radiative equilibrium
scribed in Chapter 6. In this method the "scattering" integral in equation
(6-24) now involves the entire frequency spectrum. This shows explicitly the LWnK.1nJdn = L WnK.3nB3n
n · n
(7-32)
physically important fact that the radiation field at any frequency actually
depends upon the field at all other frequencies. Using current estimate of B, In practice we do not know T*(m) but only a current estimate T(m); we there-
and 1<.~, equations (7-29) and the discretized form of (6-42) are solved for J. fore suppose that T*(m) = T(m) + 11 T(m) and, ignoring changes in 1<.!, expand
2
at all depths. These values are used in equation (7-28), which is solved for the B! = B,(T) + (oB,/oT) AT and (oB!/oT) = (8B,/oT) + (8 2 B,/8T ) 6. Tin
new temperature that satisfies it (linearizing, in principle, both B, and ": in equations (7-30) and (7-31), where 6. T must satisfy
terms of AT and iterating). Because opacities, etc., will be altered as a result
of the changes in T, the whole process must be iterated to convergence.
In his original analysis Feautrier did not introduce the desired flux
ATd = ~ WnK.tn(Jdn - Bdn) /r.wnK.3n(8Bnl aT)d (7-33)
explicitly into the problem; one may do so easily, however, by using equation
(7-27) to fix ldT/dzl in the lower boundary condition [ equation (6-44)]. If one We now can foliow two possible approaches. We could expand equations
uses Feautrier's method to solve the system, the cost is high because N, the (7-30) and (7-31) as described, eliminate 6.T using equation (7-33), and solve
7-2 LTE Radiative-Equilibrium Models 181
180 Model Atmospheres
Rosseland-mean optical depth scale; we know that this is asymptotically the
correct solution at depth. Here q(t) may be the grey solution or some other
function, based on previous results, which may differ from the grey value,
(7-34) particularly at the surface. We then integrate the hydrostatic equation (7-8)
in the approximate form
(dp,/dm) =g- (<1R r.rrfc)(''i.R/p) (7-35)
with the corresponding boundary condition, regarding J dn' as unknown.
Exercise 7-7: Write out the perturbed boundary condition with AT eliminated. simultaneously with the definition of the optical-depth scale

Eq~at~on (7-34) closely resembles Feautrier's system (6-42) plus (7-29), and dt R = ('x.RIp) dm (7-36)
again involves the entire frequency spectrum; the same remarks about physi-
cal content apply. The solution of (7-34) satisfies both the transfer equation step-by-step on a mesh {m,}, (d = 1, ... , D). This yields (N4 , T 4) at each
and the radiative equilibrium constraint (to first order) simultaneously. After point on the mesh, and using the method described in §5-2 we solve the LTE
the new intensities are found, the new temperature structure is evaluated equation of state for ne and nr(N, T) for all atomic and ionic levels. We then
using equation (7-33). Again ifwe wish to solve only for fl T, it is more efficient calculate x3n and 11:n from equations (7-3) and (7-4) on a frequency mesh {vn},
to use _Rybicki's. method, letting fl T be the constraint variable replacing J, and evaluate the mean intensities J dn and Eddington factors hn from a formal
and using equation (7-33) to replace (6-48) (this is the approach used for LTE solution of the transfer equation (6-30) with a given source function S,n·
atmospheres). The starting solution provides enough information to evaluate the radia-
tion pressure gradient in equations (7-9) and (7-10), which we can therefore
. Both Fea~trier's cons_traint pro~dure and the linearization method pro-
vide the equivalent of direct solution of the integral equation (7-21) but are integrate to find a new estimate of the run of the total number density N,,
simpler to implement than equation (7-24). Even though the linearization and new LTE occupation numbers nr. Further, knowing the variable
Eddington factors we can solve the discrete form of the transfer equations
~etho_d intr~duces a local _perturbation fl T, it defers knowledge of the mean
intensity until the system 1s solved; it thus yields global convergence and is (6-42) through (6-44), which in light of(7-27) we can write as
in no_ sense ~ l~mbda iteration. Further, the method is not inhibited by large (7-37a)
opacity variations, for these enter only as coefficients of linear algebrak
equations, and insofar as the correct inverses of these equations are obtained co
we obtain'the correct solution directly. ' co
I
Exe:cise 7-8: Repe~t Exer~ise 7-6 using the linearization method. Again check
(7-37b)
against the references m the literature, and use Rybicki's method to solve the system.

A LINEARIZATION METHOD and (at d = D),


Let us now draw together the various elements of the above discussion UonJ Dn - fo- I, nJ D-1, n)/flto-t, n
and outline an efficieni method for LTE model construction that experienc~
shows to be general, stable, and effective (35; 275). The basic thrust of the
= Hxon - 1(aB./oT)onf'L wn[XDn- 1(oB,/oT)on] (7-37c)
n
method is to write the system of transfer equations plus the constraints of
hydrostatic and radiative equilibrium in terms of a current solution (which (7-38a)
where
satisfies the constraints only imperfectly) and a perturbation of the funda-
mental _variables· (T, N) which, whe~ evaluated, more nearly satisfies the
1 (7-38b)
constraints. In each equation we allow for the change produced in all vari- fltdn =Z (fltd-½,n + fltd+½,n)
ables by these perturbations, and for the coupling of these changes from one
point in the atmosphere to another. First-order boundary conditions have been written for simplicity; it is easy
To begin, we need a starting solution for the structure of the atmosphere.
to include second-order terms.
We adopt a temperature distribution T 4 ('rR) = ¾T:rr['rR + q('rR)] on a
182 Model Atmospheres 7-2 LTE Radiative-Equilibrium Models 183

If we solved equations (7-37) we would find that the constraint of radiative Note that equations (7-39) through (7-46) apply both for LTE and non-LTE
equilibrium, equation (7-13), is not satisfied; we must therefore change the cases.
temperature T(m) in such a way as to more accurately satisfy the conditions
ofradiative equilibrium, and iterate. There are two difficulties: (a) the problem Exercise 7-9: (a) Derive equation (7-39). (b) Derive linearized expressions for
is nonlinear, and (b) the coupling is global. That is, any change l>Td implies a the upper and lower boundary conditions, equations (7-37a) and (7-37c). See
change l>N, (from hydrostatic equilibrium) and therefore <>x,, 0174 , and hence also (437).
l>J4 ,n at all d' and n throughout the atmosphere. To handle these problems, we In equation (7-39), assuming LT E, all material variations are expressed· in
linearize the equations, replacing each variable x by x 0 + l>x, and retain only terms of l>N and oT. Thus, from equations (7-6) and (7-3)
first-order terms in the o's. The power of this method is that (a) it may be
applied to a wide variety of constraints, and (b) it produces systems that opd = m(oNd - on,, d) (7-47)
allow for the effects of a change in a variable at a given point in the atmosphere
upon all other variables at all other points. In particular, the linearized ox:.= (ox:/oT)d oTd + (ox:/on,)d on•. ,+ LI (ox:/anr)d ontd (7-48)
transfer equations describe fully how a change in material properties or
with a similar expression for 011:d; in equation (7-48), o/oT applies to the
radiation field at any point propagates and affects the solution at every other
explicit appearances of Tin exp( -hv/kT), (X""(v, T), etc., and similarly for
point. We may use the linearized transfer equations frequency-by-frequency
o/on,. Now from equation (5-35) we have relations of the form
to eliminate the l>J's from the constraint equations (radiative and hydro-
static equilibrium), yielding a final system for the perturbations of the ontd = (onr/oT)Nld l>T., + (onr/oN)rld oNd (7-49)
"fundamental" variables l>N and l>T. Thus, linearizing the transfer equation
(assuming the Eddington factors remain unchanged) we have, away from the and similarly for on,, d• so that the linearized values of Pd, x:n and 11!. can
boundaries, and at each frequency Vn, all be collapsed down to expressions of the form

ox:. = (~x;tld oT., + Gfti., l>N.,


(7-50)

....
The end result is that equation (7-39) reduces to a formula involving the
perturbations at three adjacent points (d - 1, d, d + 1), of the general form 0
0
d+I d+ I d+I
I
I Tdd',noJd'n + L, udd',n oNd' + L
d'=d-1 d'=d-1 d'=d-1
Similarly, the linearized constraint of radiative equilibrium [ equation (7-13)]
is
where (Xdn = (J.,nJdn - h-1,nJd-l,n)/(l!,:r,_t,n L\tdn) (7-40)
I,w.(xdn - n,,da,)i5Jdn + I,w.(JdnOXdn - 0'1dn - a,Jd.on,,.,)
Y4n = (J.,.Jdn -J.,+1,nJd+l,n)/(L\td+t,nl!,.t,1n) (7-41) n n

= L, w.['1,n - (Xdn - n,, 4a,)J.,.] (7-52)


P,n = (Xdn + Ydn (7-42) n

a,.= [ex,.+~ P,.(£\t.,-t,./£\t,.>]/<w,-1 .• +w,.l (7-43)


and hydrostatic equilibrium [ equation (7-9)] yields

(4n/c) L, w.(J.,. oldn - fd-1. n Md- I..)


C,1n = [Ydn + ~ p.,.(Md+t,n/L\t,n)]/<w,n + W,1+1,n) (7-44)
"
+ k(Td oNd + Nd oTd - rd-• oNd- • - Nd-• or,_.)
(7-45) = g(md - md-il - NdkTd + Nd_ 1kTd-l

and (7-46)
- (4n/c) L, w.(J,i.Jdn - h- 1, .Jd- L .)
n
184 Model Atmospheres 7-3 Convection and Models for Late-Type Stars 185

In (7-52) ox and 017 are replaced with expressions of the form (7-50). Equa- where
tion (7-52) involves information only at one depth-point while (7-53) in-
volves two. R =C - I:X.T. - 1 u.;
n

Exercise 7-10: (a) Verify equations (7-52) and (7-53). (b) Linearize the upper
boundary condition on hydrostatic equilibrium.

The whole system, for all depths and frequencies, can be organized into The final system (7-59) is solved for 6N and 6T.
a form suitable for a Rybicki-method solution. Thus let Using 6N and 6T to revise the density and temperature, we can at each
point md solve for new values of all the nt ,'s and n,. 4 [ using equations (5-27)
6J.= (OJ In• OJ 2•• • • • , OJ Dnf, (n = 1, ... , N) (7-54) through (5-31) iterated to consistency] and hence new values of x:. and
11:., which are, in turn, used to obtain a formal solution of the transfer
6T = (0T 0T2, ... , 0T0 )7
11 (7-55) equation for new values of Jdn and hn (d = I, ... , D; n = 1, ... , N). We
use these new estimates to reconstruct equation (7-57), and iterate; as the
6N = (oNi, oN 2 , ••• , 0N 0 )7 (7-56) solution improves, K., L, and Mall ➔ 0, hence 6N and 6T ➔ 0.
The computing time per iteration scales with N and D (the number of
Then equations (7-39), (7-52) and (7-53) yield
frequencies and depths) as T = c(2N)D 2 + c'(2D) 3 , which is linear in N
T, 0 0 u, (so that many frequencies may be included-e.g., for line-blanketing).
V1 6J1 K1 Actually, experience has shown (35; 275) that the solution can be greatly
0 economized in most cases by assuming that the gas pressure p, will remain
T2 U2 V2 6J2 K2 unchanged during the linearization [as it will if the radiation-pressure

~~
terms in (7-9) are negligibly small]. We then rewrite all expansions as
= (7-57)
ox = (ax/aT)p, oT + (ax/ap,h op,, where
0 ... TN UN VN f,JN KN
W1 W2 WN A B 6N L
(ax/aT)p,= (ax/aT)N + (ax/aNh(aNJaT)P,' 0
X1 X2 XN C D 6T M and then explicitly assume op, = 0. This eliminates the last "row" of system
(7-57), and we solve only for 6T; the computing time then becomes T =
Each "element" in equation (7-57) is a matrix of dimension (D x D); the cND 2 + c'D 3 • The method just described has not yet been widely used,
first N "rows" represent transfer equations, the next-to-last "row" represents but its advantages are manifest; it is likely to be the preferred method in
radiative equilibrium, and the last, hydrostatic equilibrium. The T, U, and future work on LTE model atmospheres.
V matrices are tridiagonal; W, A, and Bare diagonal; X, C, and Dare bi-
diagonal. The vectors K, L, and M give the errors in the transfer and constraint
equations arising with the current estimates of the radiation field, tempera- 7-3 Convection and Models for Late-Type Stars
ture, and density. Equation (7-57) is solved by performing eliminations of
6J. from the nth "row" into the last two "rows," one frequency at a time. Thus The energy transport in a stellar atmosphere may proceed by radiative
we solve for transfer or by convection; the process that actually occurs is that which is.
6J. + (T. - 1u.) 6N + (T. - 1v.) 6T = T. - 1K. (7-58) more efficient. In general terms, radiative equilibrium prevails in spectral" .
types A and earlier, while convection becomes important in the middle
and eliminate 6J. to obtain a final system of the form F-stars and dominates in later types. The convective flow in stellar atmo-
spheres is turbulent [see, e.g., (90)] and consists of a complicated hierarchy
of "eddies" or "bubbles" moving and interacting in an extremely involved
(7-59) way. The situation poses many physical and mathematical problems :qf
7.3 Convection and Models for Late-Type Stars 187
186 Model Atmospheres

great complexity, and a definitive convection theory does not yet exist. We and hence the critical value of VR at which convection occurs. Thus for a
shall, therefore, consider only the phenomenological mixing-length theory, perfect monatomic gas VA = ½/i = 0.4, while for pure radi~tio_n ~r~ssure
which contains a number of the basic physical ingredients and provides a r = t, so vA = 0.25, and for conditions where hydrogen 1s 10mzmg r
framework for at least illustrating the effects of convection. may be only 1.1 so VA drops to 0.1 ! These results clearly sug~est that w~ ma_y
expect convection to occur in hydrogen ionization zones. This expect~tto~ 1s
THE SCHWARZSCHILD STABILITY CRITERION strengthened by noting that in the limit of the diffusion ~ppro~1!11~t1on
(-dT/dr) = (3nfxR)/(16aR T 3), which implies (from hydrostatic eqmhbnu~)
Suppose we have an atmosphere in radiative equilibrium. We then ask VR = (3nfxRP)/(16aRgpT 4 ). From this we see tha~ large values ~fthe opacity
whether an element of material, when displaced from its original position, require that the radiative gradient must be steep m order to drive the flux F
experiences forces that tend to move it farther in the direction of its displace- through the atmosphere. The opacity or-stellar material ~ecomes large when
ment. If so, the atmosphere is unstable against mass motions, and convection hydrogen is appreciably excited into its upper states; this happens at about
will occur; if not, the motion will be damped and will die out, and radiative the same conditions where ionization occurs and causes r to decrease. The
equilibrium will persist. The basic criterion for stability against convection two effects work together and imply that the radiative gradient does, in fact,
was established by K. Schwarzschild (416, 25). exceed the adiabatic gradient in the hydrogen ionization zone, so that
Consider a small element of gas whose position is perturbed upward by a convection occurs. The importance of these mechanisms and the existence
distance Min the atmosphere. We suppose that (a) the movement is so slow of extensive hydrogen convection zones in stellar envelopes was first
that the element remains in pressure equilibrium with its surroundings and recognized by Unsold (636). . . . .
(b) the element does not exchange energy with its surroundings (i.e., the In the earliest-type stars, hydrogen 1s essentially completely ionized
process is adiabatic). Because the pressure drops as the element rises, the throughout the envelope, and radiative equilibrium prevails (thin, weak
gas expands, and the density decreases by an amount (~P)e = (dp/dr)A M; convection zones associated with He 0 and He+ ionization exist, but transport
the subscript E denotes "element".· and A denotes "adiabatic." If, at its new only a tiny fraction of the flux). In the A-stars, thin hydrogen convection
position, the density of the element is less than that of its surroundings, it zones begin to develop at shallow depths (t ~ 0.2). In_the F-stars, the con-
experiences a buoyancy force and will continue to rise. Thar is, if (dp/dr)R vection zone starts somewhat deeper, and becomes thicker; by types F2 to
is the density gradient in the radiative surroundings, instability occurs if F5 convection will transport essentially all of the flux at some point within .....
the zone. For later and later types the zone extends ever deeper, and ~on-
(.1p)8 = (dp/dr)A M < (.1p)R = (dp/dr)R M (7-60) vection becomes more efficient; in the M-stars the convective envelope 1s so 0
I\)
(recall 'that dp/dr < 0). We may write equation (7-60) in a more convenient extensive that it determines the structure of the star as a whole (396).
form. In the adiabatic element (which we shall momentarily assume is a
I
perfect gas) the equation of state is In p = y In p + C, while in the radiative MIXING-LENGTH THEORY
surroundings (again assumed to be a perfect gas) In p = In p + In T + C'.
Using these relations to compute (dp/dr)A and (dp/dr)R, and demanding The basic physical picture used in the mixing-length theory is that ~he
the pressure gradients be equal, we find from equation (7-60) that the transport in the unstable layer is effected by turb~lent elements mo~mg
Schwarzschild condition for instability is upward and downward through a surrounding en~1ronment. The upward
(downward) moving elements have an excess (deficiency) of thermal en~r~y
[(y - 1)/y](-d In p/dr)R < (-d In T/dr)R (7-61) relative to the surrounding material. At the end of some characteristic
distance, the mixing-length, one supposes that these elements "dissolve"
or VR = (d In T/d In p)R > (y - 1)/y = (d In T/d In p)A = VA (7-62) smoothly into the surroundings, delivering any excess energy they possess
In stellar atmospheres the gas is not perfect because of the effects of ionization or absorbing any deficiency. A direct energy !ransport results: and the
and radiation pressure; we may account for this by generalizing y to r temperature gradient is decreased below that wh1c~ would occur if th~ only
(160, 57) and writing VA = (f - 1)/f where r will not, in general, equal transport mechanism were radiat~on. To c~ar_actenze _the process we intro-
its value for a perfect monatomic gas, namely y = (Cp/C.) = i, Convenient duce the following gradients: VR 1s the radiative gradient that would o~cur
formulae for the calculation of r, allowing for radiation pressure and if convection were suppressed; VA is the adiabatic gradient; Va is the grad~e.nt
ionization, have been given by several authors [ see, e.g., (638, §56; 643; 364)]. of the convective elements; and Vis the "true" gradient of the surroundmgs
These effects can be of major importance, and may drastically lower VA• in the final state where both radiation and convection transport the total
,
188 Model Atmospheres 7-3 Convection and Models for late-Type Stars 189

flux together. In general we will have One of the uncertaintities of this approach lies in the question of how to
specify the mixing-length I; usually it is assumed that I is simply some
(7-63)
multiple of H, say 1 or 2.
Consider now a rising element of material. If oT is the temperature difference To complete the theory, we need another relation that will allow us to
between the element and its surroundings, the excess energy delivered per express V and VE in terms of VR and VA; this may be done, following Unsold,
unit volume when the element merges into the surroundings is pCP oT. The by considering the efficiency of the convective transport. As an element
temperature difference arises from the difference between the gradients of the rises, its temperature exceeds that of the surroundings (which accounts for
element and the surroundings. Thus for elements traveling over a distance the energy transport); the temperature excess implies that it will lose some
6r with an average speed li, the energy flux transported is energy to its surroundings by radiation. This energy loss will diminish the
excess energy content of the element and therefore decrease the energy yield
1tF00nv = pCPlioT = pCPli[(-dT/dr) - (-dT/dr)E] llr (7-64) when the element "dissolves" at the end of its mixing length. We therefore
define the efficiency parameter as
At a given level in the atmosphere we will find elements distributed at
random over their paths of travel; averaging over all elements, we set t:.r = 1/2 excess energy content at time of dissolution
where I is the mixing length. Further, using the hydrostatic equation (dp/dr) = (7-70)
Y = energy lost by radiation during lifetime of element
- pg, and introducing the pressure scale height H = (- d In p/dr)- 1 = p/(gp)
we can rewrite (7-64) as
The excess energy content of the element is proportional to (V - Ve) [cf.
(7-65) equation (7-65)]; had the element moved adiabatically, the energy content
would have been proportional to (V - VA). Therefore the loss by radiation
To estimate 11, we calculate the work done by buoyant forces on an element is proportional to (V - VA) - (V - VE) = (VE - VA) so that
and equate this to its kinetic energy. If op is the density difference between
the element and its surroundings, then the buoyant force is Ji,= -g op. (7-71)
The equation of state yields In p = In p - In T + In µ, where µ is now
considered to be variable to allow for effects of ionization and radiation Alternatively, we may calculate the quantities in the numerator and denomi- ......
pressure. Thus we may write d()n p) = d(ln p) - Qd(ln T), where Q.= nator of equation (7-70) in terms of local variables. Thus for an element of
0
1 - (o In µ/o In T)p, and, demanding pressure equilibrium (op = 0), we volume V, with excess temperature oT, the excess energy content is pCP V oT.
have op = -Qp oT/T, so that The radiative loss depends on whether the element is optically thin or thick. w
In the thin limit the rate of energy loss will be 41tXR t:.B, from a volume V, I
f,, = (gQp/T) oT = (gQp/T)[(-dT/dr) - (-dT/dr)e] 6.r (7-66) over a lifetime (//ii). Assuming an average .excess of (oT/2) over this lifetime,
we have
The buoyancy force is thus linear in the displacement; integrating over a
total displacement 6., and setting 6. = l/2 to account for the average over Ythin = (pCPV oT)/[4n(40'R T 3 /n)(oT/2)(xR V)(lfv)]
all elements passing the point under consideration, we obtain the average
work done on the elements ".'" (pCpli)/(8aRT 3 t,) (7-72)
Iv = s: f,,(6.r) d(6.r) = (gQpH/8)(V - Ve)(I/H) 2 (7-67) where t, denotes the optical thickness of the characteristic element size /,
-r, = XRI. Equation (7-72) applies when t, « 1. At the opposite extreme,
We now suppose that about one-half of this work will be lost to "friction"
in pushing aside other turbulent elements and the other half will provide t, » 1, we apply the diffusion approximation to determine the radiative
the kinetic energy of the element (i.e., ½p1J 2 ::::: ½W) from which we find flux lost by an element of characteristic size I, with fluctuation oT, by writing·
(-dT/dr) ~ (oT/1). Assuming a surface area A and the same lifetime as
li = (gQH/8)i(V - Ve)i(I/H) (7-68) before, we now have
and, therefore, from equation (7-65), Ythlck = (pCPV oT)/[(16aRT 3 /3xR)(oT/l)A(lfii)] = (pCpv/16uRT 3 )3xR(V/~)
1tFconv = (gQH/32)-i(pCpT)(V - VE)¼(//H) 2 (7-69) (7-73)
190 Model Atmospheres 7-3 Convection and Models for !Ate-Type Stars 191

The choice of (V/A) is ambiguous and introduces another source of un- where A depends only on local variables. Adding (V - VE) + (VE - VA) to
certainty into the theory; if the elements are presumed to be spherical, both sides of equation (7-78) and using equation (7• 76) to eliminate (VE - VA),
(V/A) = 1/3 and we find a cubic equation for (V - VE)½ = x, namely
Ythlck = ½te(PC,v)/(80-R T )
3
(7-74)
A(V - VE)t + (V - VE)+ B(V - VE)½= (VR - VA) (7-79)
We interpolate between the two extreme cases by writing which may be solved by standard methods for the root x 0 • We thus obtain
the true gradient V = VA + Bx 0 + x 0 2 , and proceed with the integration,
(7-75) now regarding T as a function of p. If, at some point, convection ceases,
we revert to the original T(fR) relation (adjusted to match the current values
Combining equations (7-71) and (7-75), and substituting equation (7-68) for
of T and tR) and continue the integration into a radiative zone.
v we derive finally
In the case that the material is presumed grey [ or, for nongrey material,
the convection zone is really deep enough that the diffusion approximation
is correct, and the true nongrey temperature distribution is known near
the surface] the treatment described above is essentially exact. Using this
approach for grey atmospheres, Vitense (653) performed computations for a
The final requirement we place upon the theory is that the correct total flux wide range of effective temperatures and gravities; this work nicely delineates
be transported by radiation and convection together-Le., the role of convection in stellar atmospheres over much of the H-R diagram.
In a general way, the outermost layers can always be expected to be in radia-
1tF = 1tFrad + 7tFconv = O'R r:rr (7-77) tive equilibrium because densities and opacities a:re small and radiative
transport is more efficient than convective. In deeper layers, the opacity and
The mixing-length theory described above is the simplest (and most
density rise, ionization may occur, and convection may begin. Convection
widely used!) convection theory in astrophysics. Numerous refinements
will have its largest effects in stars oflow effective temperatures (in which the
have been proposed, attempting to introduce nonlocal information into the
hydrogen is essentially neutral in the outer layers) and high gravities (which
theory; it would take us too far afield to attempt to describe these here and
imply large densities and heat capacity, hence efficient thermal transport). _.
the interested reader should examine the literature. [See, e.g., (594; 595 ;' 450,
237; 479) and the references cited therein.]
When convection is efficient, it will transport practically all the flux, and V O
will be close to VA; indeed, in stellar interiors, convection (which it occurs) ~
is so efficient that one may set V = VA and dispense with the mixing-length I
CONVECTIVE MODEL ATMOSPHERES
theory entirely. When convection is inefficient, the true gradient V will lie
The computation of convective model atmospheres is more complicated close to VR• and a substantial part of the flux may be carried by radiation;
than for radiative models (even assuming the mixing-length theory) because in this regime the uncertainties of the mixing-length theory make themselves
there are two transport mechanisms that must be brought into a final balance felt fully.
to satisfy equation (7-77). We may proceed as follows. Suppose we assume When the convection zone lies close enough to the surface that the diffusion
some specification of the temperature distribution-e.g., the grey distribution approximation used to derive equation (7-78) is invalid, it is then necessary
on a Rosseland-mean optical-depth scale. We then carry out a step-by-step to calculate F,ad from the solution of the nongrey transfer equation, and
integration of equations (7-35) and (7-36), as before. At each point we may employ an iterative temperature-correction procedure. In any such procedure
calculate VR = VR(T, p, pg) and VA = VA(T, p, pg). If at some point we find it is essential to account for changes in both F,ad and Fconv induced by t~e
that the instability criterion is satisfied, we must determine the true gradient proposed alteration of the temperature structure. Methods for constructing
V, VR ~ V ~ VA• which satisfies equation (7-77). If the instability occurs convective models based on a generalization of the Avrett-Krook procedure
deep enough for the diffusion approximation to be valid, then (F,ad/F) = have been used to study F-type main-sequence stars (422), middle-type
(V/VR), and equations (7-77) and (7-69) reduce to supergiants (500), and M-stars (dwarfs through supergiants) (48). A detailed
description of a computer code that treats convection is given in (379). An ex-
A(V - VE)t = VR - V (7-78) tensive grid of nongrey models (4000°K ~ T.rr ~ 50,000°K, 2 ~ log g ~
192 Model Atmospheres 7-4 Results of LTE Model-Atmosphere Calculations 193

5), including convection effects where appropriate and making allowance for we shall merely give a few typical references and invite the reader to examine
line-blanketing, is available (247, 377). More limited grids of blanketed these papers and the references cited therein. A comprehensive list, through
convective models may be found in (512; 513; 514), and extensive computa- 1965 can be found in (506); many of the models in that list use a grey tem-
tions for M giants and supergiants, allowing for molecular line-blanketing, pera~ure distribution on a mean optical-depth scale. Extensive ~rids of
are given in (341; 342). The solar convection zone has been studied with unblanketed, nongrey, radiative-equilibrium models can be found m (421;
both the mixing-length approximation (652) [see also (479)] and more 608)· models including hydrogen-line blanketing by the "direct approach"
detailed hydrodynamical theories (99; 100). Recently, methods for com- hav; been calculated for A- and B-type main-sequence stars and giants
puting convective models using a linearization procedure similar to that (423; 357), and for white dwarfs (620; 412). Models for 0- and B-st~rs,
described in §7-2 have been developed (274; 275; 479). The basic change in allowing for blanketing by hydrogen lines and strong lines of abundant h~ht
the formulation is to use equation (7-77) as the energy balance equation; ions, by the direct approach, are given in (449; 7; 298; 105; 471~. MaJor
introducing a discrete representation of the flux [ cf. equations (6-15) and improvements in the simulation of real atI?~sphere~ have ~een ac?1eved by
(6-26)]. On an angle-frequency mesh {µ,, v1} we can write including the blanketing of thousands to millions oflmes, usmg various types
I of opacity distribution functions. A preliminary model of Procyon (F5IV)
4n L w,µ/(ud+l,I
1•1
- Ud1)/~td+½,I + nFconvld+t = aRT:,, (7-80) allowed for about 30,000 lines (612); extensive grids including hundreds of
thousands of lines semiempirically have been published in (247; 512; 513;
514); and recently these efforts have culminated in the publication of models
The convective flux can be regarded as Fconv(P, p,, T, V) [given these variables,
(331; 516,271) allowing for 1,760,000 lines on the range
VE follows from equation (7-76) and Fconv from equation (7-69)]. The radiative
term may be linearized as before. In linearizing the convective term, the 8000°K =::; T.rr =:;; 50,000°K, 2 ~ log g =:;; 5
total pressure is fixed, and the derivatives appearing in the expression
(as well as a solar model). A few illustrative results from these calculations
F 0 onv = F~onv + (oF,/op,) op, + (oF,/oT) oT + fj)F /0V) oV
0 (7-81) will be described below.
are computed numerically. Several approximations are introduced (274) to
reduce this to an expression in oT only, and practical procedures for handling EMERGENT ENERGY DISTRIBUTION
......
numerical problems have been developed (274; 275). Improvements, in 0
The ultimate goal of stellar atmospheres analyses is the construction of 01
convergc:ince might be obtained by including terms /n oN as well as oT,
mathematical models that describe the physical properties of the outer layers
but, as described earlier, this is inherently more costly. I
of stars. Having computed detailed models on the basis of the theoretical
At the present time, convection theory as applied in stellar atmospheres
principles described in this chapter, one .then compares predicted and ob-
analysis is only heuristic; improvements to the physical theory are being
served values for the distribution of radiation within the spectrum, and
actively pursued and, when more accurate treatments of convection become
available, our understanding of the atmospheres ;,f late-type stars will be attempts to associate a real star with a definite model. In ~his way val~:s of
the parameters that describe the model, (Terr, log g, chemical compo~1tu;m),
improved significantly.
can be assigned to the star. We shall concentrate here on the comparison of
observed and computed values of continuum features, deferring a discussio_n
of lines to the second half of this book. In early-type stars, spectroscopic
7-4 Results of LTE Model-Atmosphere information about gravities comes mainly from profiles of the hydrogen lines
Calculations for Early-Type Stars (for which the broadening mechanisms are density-sensitive) and about
abundances from an analysis of line-strengths; we shall therefore focus
The largest group of reliable model atmospheres available pertains to solar mainly on the determination of T.,r and related parameters-e.g., the bolo-
and earlier spectral types; therefore we shall confine attention primarily metric correction. In fitting the continuum we may follow several approaches.
to these stars. For later types, many difficult problems related to molecular (a) A fit can be made to the entire spectrum. T?is a~sumes t_hat a complete
line-blanketing and the hydrodynamic structure of the atmosphere must be energy distribution (perhaps including spectral regions macce~s1bl~ to groupd-
overcome. There is now a very large literature concerning LTE, plane- based observations) is available. In most cases the comparison 1s based on
parallel, model stellar atmospheres, which cannot be described fully here; the relative distribution of energy in the spectrum-i.e., F./F. 0 , where ~
7.4 Results of LTE Model-Atmosphere Calculations 195
194 Model Atmospheres

denotes some prechosen frequency. In a few cases it is possible to make the All comparisons between models and observations rest, in the end, on the
comparison in absolute energy units using fluxes in ergs cm - 2 sec - 1 hz - 1 fundamental calibration of the energy distribution ofa standard star (or stars)
for both the star and model; here we obtain an enormously important check in the sky, and it is impossible to overemphasize the importance of this basic
on the validity of the whole theory. connection between theory and observation [see also (516, 241)]. Because
(b) More limited information concerning a few outstanding features in the it is, in practice, impossible to make an a priori determination of the absolute
flux distribution may be used. For example, in A- through O-stars the slope efficiency of the telescope-spectrometer-receiver system, one proceeds by
of the Paschen continuum (3650 A ~ A. ~ 8205 A) is useful; the name is derived comparing a star to a standard blackbody source of known emissivity, using
from the fact that the dominant opacity source on this wavelength range in the same observational apparatus. It would take us too far afield to describe
early-type stars is from photoionizations of then = 3 level of hydrogen. Two the details of this procedure here; it is worth the reader's effort to study the
other important features are the Balmer jump, literature on the suhjecl [e.g.. (261. Chap. 2:484:485:486; 285; 487; 286;
287; 288) and references cited therein]! For main-sequence 8-stars, both the
= 2.5 log[F.(,t3650+)/F,(A.3650-)]
DB slope of the Paschen continuum and the Balmer jump depend strongly on
and the Paschenjump, Dp = 2.5 log[F,(A.8205+)/F.(A.8205-)]. These param- T.rr and are insensitive to surface gravity (sec Figure 7-3); hence both may
eters give measures of the effects of the onset of photoionization edges near be used to infer T.rr·
the wavelengths stipulated.
In particular, at the Balmer jump, towards shorter wavelengths the opacity
is large, owing to photoionizations from the n = 2 level of hydrogen, hence 1.8
we receive radiation only from the upper, cooler layers; whereas towards
longer wavelengths, the material is much more transparent and we see deeper, 1.6
hotter, layers from which the flux is larger. The result is a fairly abrupt drop
in the flux across these frequency boundaries (actually the drop is not sharp
1.4
because of the opacity of overlapping lines of the series converging on the
continuum). The continuum slope can be observed and computed unam-
biguously, but one must be able to correct the observed values for interstellar 1.2 3
__.
reddening effects, and must have a reliable absolute energy distribution 0
standard (see below). The "jumps" are not as strongly affected by reddening 1.0 CJ)
or calibration problems because they are defined over a very limited fre- Ds I
quency range. However, although the flux ratio is obtained easily from 0.8
unblanketed models, this abstract quantity is not actually measureable,
owing to the confluence of lines near the series limit; hence one must use 4.44
blanketed models, and apply the same operational process to both observed 0.6
and computed distributions to obtain meaningful comparisons.
(c) Finally, we may employ colors measured with filters that isolate speci- 0.4
fied bands within the spectrum. Colors can be obtained easily and accurately
by standarized observational techniques, and such measures can be extended 0.2
to very faint stars by use of broad-band filters. On the other hand, it is
easier to calibrate colors against theoretical models for narrow bandwidths,
for then one can account more accurately for line-blanketing effects in the
model. In practice a compromise must be struck, and a large number of O,rr
color systems with various properties exist, many of them measuring param-
FIGURE 7-3
eters that are specially designed to characterize the properties of particular Balmer jumps computed from LTE model atmospheres, as a
groups of stars [see, e.g., the systems described in (516)]. A widely used function of effective temperature and gravity. Ordinate:
system that has been well-calibrated in terms of models is the Stromgren Balmer jump in magnitudes; abscissa: 0.,1 = 5040/Terr•
uvby system. Curves arc labeled with log g.
196 197

-0.2

0.0

0.2
"!

Avrett and Strom


i 0.4

Hayes (1968) 0.6


'i.
cc Leonis, B7V
0.8

3.0 2.5 1.5 1.0 1.0 Vega


• Hayes and Latham
1.2 9400,4
FIGURE 7-4
Comparison of the energy distribution of cc Leo (B7V) as observed by
Hayes (285) with the model atmosphere (608) that best fits the Paschen 1.5 2.5 3.0
1.0
continuum, namely, with T, 11 - 13,000°K and log g = 4. Note that the
computed and observed Balmer jumps arc consistent. Ordinate: relative
ftux in magnitudes; abscissa: 1/i.. where). is in microns. From (285). FIGURE 7-5
Comparison of energy distribution of the fundamental standard Vega as measured by
Hayes and Latham (287) (dots) with a line-blanketed model [(381; 561, 271)) that has
T,rr = 9400°K and log g = 4. Ordinate: relative flux in magnitudes; abscissa: 1/).
where ). is in microns.
Until about 1968, a serious discrepancy existed between these two deter-
minations in the sense that, if a fit was made to the Paschen continuum, the
observed Balmer jump was smaller than computed (or if the Balmer jump -0.20 0
was fitted, the slope of the observed Paschen continuum was too shallow); -..J
the discrepancy in T,rr amounted to 3000°K (the Balmer jump temperatures
0.00
.. .-
A•·---- I
being higher). The problem was resolved when a new calibration was made
at Lick Observatory by Hayes (285; 286), who showed that the original
0.20
,,-· ~--- .
0.40 I
I o

calibration had too flat a Paschen continuum slope. With his calibration it I
I
0.60 I
became possible to fit the observed spectrum very well (see Figure 7-4), and llm(l/i.)
I
I
I
effective temperatures deduced from the two parameters were consistent 0.80 I
I
I
[see, e.g., Figure 3 of (682)]; by this procedure, an effective-temperature scale I
1.00 ,•
for the B-stars can be established (682;555). A second recalibration made at / Blanketed model
Palomar (487) by Oke and Schild disagreed with the Hayes calibration (and 1.20
Unblanketed model
the models) below the Balmer jump; recent work by Hayes and Latham 1.40 Observation
(287; 288), however, has shown conclusively that the source of the discrepancy
1.60 '--'--''--__,_--'--'---'--_.__........_..__~
was a faulty correction for the effects of atmospheric extinction in the Palomar 3200 4000 4800 5600 6400 7200
data, and when this is removed, the Lick and Palomar results agree. A ;_ (Al
comparison of the energy distribution of Vega with a line-blanketed model is
FIGURE 7-6
shown in Figure 7-5. In fitting relative energy distributions in spectral regions Unblanketed and blanketed energy distributions for models with
visible from th~ ground, it is important to allow for line-blanketing in spectral T,rr = 6500°K and log g = 4, compared with observations of
types later than A. For example, in Figure 7-6 we see that the blanketed model Procyon. Ordinate: relative ftux in magnitude units; abscissa:
of Procyon (612) mentioned at the beginning of this section fits the observed wavelength). in A. From (612), by permission.
198 Model Atmospheres

energy distribution quite well, whereas the unblanketed model is much too
bright. Blanketing effects are minor in the visible for B- and 0-type stars but
become large in the ultraviolet; fits made to models ignoring u.v. line-
blanketing will be systematically in error (see below).
An entirely different approach to the derivation of effective temperatures
can be made using absolute fluxes. From the fundamental calibration one can
determine the actual energy output of a star at a particular wavelength; N

specifically, for Vega (ex Lyr), which is the standard star, the average of the ~ " ..... >,
'"fS;::i.o
8 'o o,;:
Palomar and Mt. Hopkins work vields (287) a flux. at the earth. off. = u 'il II i
3.50 x 1o- 20 ergs cm - 2 sec- 1 hz- 1 at -l5556 A. For any other star we use the '3 :; '8 1::-
magnitude difference .1m of the star relative to Vega (at this wavelength) to M s .c a q,• s
oi
,l9
s.., - LI.e
scale the flux quoted above by 10- 0 •44"'. As was discussed in §1-4, we can ICU~
u t:]
-
§ .
0

convert fluxes measured at the earth to fluxes at the stellar surface ifwe know 'S:.] lil • 7
the angular diameter of the star. Angular diameters have been measured (113) ..,.
'il
'8-!;11-
e
:o ;:!i ~
for 32 stars on the spectral-type range 05 to F8; these may be used to con- 8 '3 ~ I

struct an effective temperature scale. One could, for example, deduce the 1l -5 ! ~ ~
absolute stellar flux at some particular wavelength, and choose the model ]~~-S7
; J 'li ij
that yields the same flux to assign Terr• By comparing the total energy emis-
sion with that observable in the visible, one can then obtain the bolometric
.,,
~~ ji ~
. _g t a~
correction. ,-: ;;-~~"'8-
Such an approach is vulnerable, however, to serious system~tic errors if ::::- !~:.a]x
.!? 0 .C C ,.
inadequate allowance is made for line-blanketing (190), and will tend to ~ II g~ ~ _.,
"'
assign too-high values of Terr and bolometric corrections. The nature of the e c:,•I: .,8 =
O C ~
"'
ultraviolet line-blanketing is illustrated in Figure 7-7. The blanketed model
Ul - •- C ,!:,
f- ~ I! Oj 1! 0
..J o J! l'.J 0 (X)
(449) allows for the strong lines ofH, He, C, N, 0, Si, Cl, Fe, etc. on the rarige
912 A ~ ·;_ ~ 1600 A by the "direct" approach. The effects of the blanketing .... i~-i"B.; I
~ N = 96
are quite dramatic. The integrated flux of the blanketed model corresponds to lil .cu tf ·-tl
Terr = 21,900°K, but so much flux has been removed from the ultraviolet, :g I,.;IIli~aE
:S 2$ .!: C
and redistributed to longer wavelengths, that the energy distribution there -g g :~ ~·~
most closely resembles an unblanketed model with T.rr = 24,000°K. Had C ~
00 tU bl)
]:ij-5~~
we used unblanketed models to fit the visible energy distribution (whether ~cof.t
c8,i:3~
absolute flux values or the Paschen continuum slope), the derived effective ·- e
:0 ~ 1l 8. ::!. .
oi

temperature would have been systematically too high by 2100°K ! In fact, ~e0:i 6 ~g
"direct-approach" models provide, at best, a lower bound on the total amount °' .., 2 ~ ~ ~ Sl 'i
of blanketing, and only the recent calculations (381) allowing for millions of
D::
:i s s-::s -~-
c.;;;;

~i:;:g]~~8.
~ ~ e
,!:? ·-

lines with opacity distribution functions provide reliable estimates of these


effects. ::
In the face of these difficulties it is preferable to avoid direct reference to
the models, and use known angular diameters, visible energy distributions,
and recent space observations in the ultraviolet to construct complete abso-
lute energy distributions empirically (516, 221; 169). In this procedure there
are nontrivial problems of ultraviolet calibrations and interstellar reddening
effects but, with care, these can be overcome (96). From the integrated flux,
200

8
-
I
<
7
6
-1.8
-1.6 OAO-2 data I
Ground-based data
i -1.4
~ s -1.2
N
4 -1.0
e3
I
~ -0.8
..;:
.£,
Si 2 ---r..::: -0.6
00 -0.4
~
..2
.,, -0.2
N 0.0
2000 3000 4000 5000 6000 I
+0.2
J.(AJ +0.4 ... II Aur, A0p, B - V = -0.08
l'IOURE 7-8
+0.6 - 134 Tau, B9.5 V, B - V = -0.07
Comparison or empirical absolute energy distribution or ex Leo
(B7V) (516, 221) with a line-blanketed model (381) or the same
+0.8 --- y UMa, AO V, B - V = 0.00
effective temperature (12,200°K). The agreement is excellent and +1.0
lends strong support to the model techniques. Ordinate: absolute 1000 1500 2000 2500 3000 3500 4000 4500 5000 5500 6000
flux 109 / 1 in ergs cm- 2 sec-• A-• at the earth; abscissa: wave- Wavelength (A)
length ). in A.
FIGURE 7-9
Comparison or relative energy distribution or the peculiar (Si 3995) A-star OAur with
the actual effective temperature is obtained; this value is esse._ntially indepen- those of 134 Tau (B9.5V) and y U Ma (A0V) (391). Because of the enhanced line
dent of any model atmosphere. A comparison of the empirical absolute energy blanketing arising from the greater heavy-element abundances in the peculiar star, the
distribution with that from a model that has the same (i.e., the empirical) T.rr energy distribution or OAur matches neither or the normal stars, but resembles the cooler
is therefore extremely significant, for it provides a .test of both the absolute.
star in the ultraviolet and the hotter star in the visible. Ordinate. relative flux in .......
magnitude units; abscissa: wavelength in A.. From (391), by permission.
and relative flux predictions of the model. Such a confrontation is shown in 0
Figure 7-8 for the B7V star ex Leo (516,221) and a blanketed model (381) of (0
the same effective temperature. The agreement is excellent, and lends strong -1.8 I
support to the validity of the new models. -1.6 14,000°, Normal opacities
••• 13,D90°, 100 x Normal opacities
As an example of an extreme case ofline-blanketing effects, it is interesting -1.4
-1.2 11,000°, Normal opacities
to consider the ultraviolet flux distribution in the Ap stars as observed by
OA0-2. The Ap stars are objects with anomalous abundances of certain -1.0
elements (e.g., Si, Mn, Cr, Eu, Sr) that are enhanced by factors of 10 2 to 10 3 • 1 -0.8
-0.6
These stars have strong magnetic fields and show spectral variations with i£
,.:.: -0.4
time; the observed variation of the field is well explained by an oblique rotator 00 -0.2 11 , ,___ __

model in which the magnetic axis is inclined to the rotation axis of the star, ..2
.,, 0.0
• • I( • •
while the spectral variations indicate concentrations of the elements into r-i +0.2
I
definite zones or patches on the stellar surface [ see, e.g.• (522; 125; 194 )]. The +0.4
I
greatly enhanced heavy-element abundances produce strong additional +0.6 I
I
+0.8 I
blanketing in the ultraviolet, over and above that found in normal stars. The I
I•
+1.0
effect is nicely illustrated in the peculiar (Si 399S) star 0 Aur (see Figure 7-9),
whose energy distribution in the visible matches a normal star of the same
+ 1.2
+1.4
I
I
color, but in the ultraviolet (391) fits that of a cooler star. The effect of 1000 1500 2000 2500 3000 3500 4000 4500 5000 5500 6000
enhancing the line opacities in models is shown in Figure 7-10, which repro-
Wavelength (A)
duces the behavior seen in Figure 7-9 at least semiquantitatively. Note that
FIGURE 7-10
Line-blanketed models (391) showing effects or 100-rold enhanced heavy-element abundances;
........ , ... ,. ... ,.,,."........ n ...... 1--1,,..,,..~ .,.., th,-. ,,.fr,,.,...t<: ~hnwn in F;pm·~ 7.q Frnm (:\91), hv nermission.
203
202
).
2.85 0
· .• · .
2.87
0

00
0
0
····· ... . ... •······· ·······.................. .........................·
"
3317
y 00 0
2.89 0
0
2.91 0 0 8 •,•,,••····•• ,.. : .................. ,.................: ·••. ,..:..••·•.:" •........,'• 2985
00 0 0

......... ...:..................··.. ····· ...........................:·•.... ........... .


2.93
•• • 2945
'76~
2.78 • •• .....
B
2.80


••• •
• • •
·1 E
~
8.
·····················
.................................................

····.....•...........
2462

...
~

,~~
2.45 "'
"" ••


0
!!l
·a::, ...... ..••,•·.
.• . 2386
XX JC1CM
u
2.47
2.49
• ••
·J e
q
0
....
··........: .
..... ;.:.;.....
-0.02
.,,.----...............
II
0.00
',,,, ____ ..,,.,,,, , ,,,,, ------ M ....·······•""''.. ... " •··•..
.... 1913
3317 A 0.02 -a
!:!
..................... ····•·. •···• . •·. ,,.....
u
0.04
-~
>
1554
0 0.2 0.4 0.6 0.8 1.0 ]
Phase
!!l
C
. ·:•,·.......... ..
::,
7-11 8
FIGURE
Variation or the peculiar (Si-Cr-Eu) A-star oi 2 CVn in UBV and ] ...:•·· ·... .:·:·:· .. ........
·····.•...·•·
at ).3317 Aas measured from OAO-2. From (464), by
0
'ob
........ ........ 1430
0
permission.
I
........
the peculiar star has a lower T.rr than a normal star of the same visible color .:··. :.: ....
(or energy distribution), and a total energy distribution that is distinct from 1332
a normal star of the same Terr. A further effect is shown in the Ap (Si-Cr-Eu) .. ·:
spectrum variable tX 2 CVn. The light variations in the visible are shown in
Figure 7-11 along with a near-ultraviolet band observed from OA0-2 (464);
the far ultraviolet behavior is shown in Figure 7-12 where we see an alltiphase 0.0 0.2 0.4 0.6 0.8 0.0
variation. These results are easily interpreted in terms of much-increased Phase
ultraviolet line-blanketing at phase 0.0, which depresses the ultraviolet flux FIGURE 7-12
and redistributes the energy into longer wavelength bands, thus forcing a Variation of oci CVn in ultraviolet as measured from OAO-2.
brightening in the visible; this interpretation is consistent with the fact that Curves are labeled with wavelengths (A) or filters. 'Note anriphase
variation or rar ultraviolet flux relative to visible! From (464).
the rare-earth lines reach maximum strength at this phase. In contrast, at
by permission.
phase 0.5 we observe regions of the atmosphere where the rare-earth lines are
at a minimum, hence the ultraviolet blanketing is lowest; at these phases flux
emerges more freely in the ultraviolet (leading to a brightening there) and is
204 Model Atmospheres
7-4 Results of LTE Model-Atmosphere Calculations 205

no_t redistributed !nto the visible regions. which. accordingly. decrease in we have
bn~ht_ness. The existence of a "null wavelength" near ).2960, which shows 110
van_at1on, supp?rts th_e differential line-blanketing interpretation and argues
agamst others mvolvmg, e.g., gross geometrical deformations of the stellar (i - j)obs = - 2.5 log [ f
0
te T1(2)F().) d).; f
0
00
TAJ.)F().) di.] + kli (7-82)
surface.
For most ~ta.rs w~ do no~ have detailed energy distributions, but only where the constant kii accounts for the unknown telescope-photometer
much more lu_mted mfo~mat1on such as colors measured in a photometric transmission and photomultiplier response. To determine klJ, the stapd,ird
system. By s_u_1table choices of filter combinations, colors can be obtained approach is to use observed energy distributions F(..l) ofreal stars, with known
that are sens1t1~e t~ effective temperature, gravity. and metal abundance and colors (i - j)0 b•• in equation (7-82), and to require agreement of the computed
:•llow a de~_ermmat1on of the amount of interstellar reddening. For example, and observed colors [see, e.g.• (411; 488; 516, 31; 516, 45)]. The second step
111 is to apply equation (7-82), with known values of klJ, to model-atmosphere
the Stromgren uvby system for, say, A-G stars, the index (b - y) is a
go~d temperature indicator, c 1 = (u - v) - (v - b) is gravity-sensitive flux distributions to obtain the "observed" colors of the models. Compari-
whi~e m, = ~v - b)_ - (b - y) is sensitive to metal abundance. To recove; sons between stellar and model colors then allow the estimation of stellar
the mformat1on available in these data the system must be calibrated against parameters; see Figure 7-13.
model ~tm~spheres. A first step in the procedure is the determination of Once again line-blanketing plays an important role, (a) because of blocking
normahzat1o_ns_between observed colors and those computed from the known effects in individual filter bands, and (b) because the value of Terr of the model
filter transm1ss1ons. If T 1().) denotes the filter transmission in color ;, then depends on the effects of line-blanketing. In later-type stars it is often neces-
sary to perform very detailed spectral synthesis studies [ see, e.g., (80; 81;
1.8,-r-----,---,,---.,----~- 82; 83; 516, 319)] to evaluate the blocking effects, while models with very
complete opacity distribution functions such as in (381) are indispensible
1.6 for estimating T.rr· At present considerable effort is being devoted to the
development of opacity distribution functions for molecular line-blanketing;
when these become available, a reliable analysis of the energy distribution
1.4
of late-type stars should become possible.
~

1.2 ~

TEMPERATURE STRUCTURE
1.0
In addition to emergent fluxes, model atmospheres give the variation of
the physical properties of the atmosphere with depth. In particular, we
obtain an estimate of the temperature structure of the atmosphere, which,
as mentioned in §7-2, is of central importance in LTE models. Let us now
0.6 consider how the temperature structure in a nongrey atmosphere differs
from the grey-body distribution determined in Chapter 3. We shall focus
0.4 attention on two features: (a) the ratio of boundary temperature to effective
temperature, T 0 /Terr, which has a value 0.811 for a grey atmosphere, and
(b) the effects of backwarming. To gain insight into the physics of the situaa_
0.3 0.4 tion, we shall consider two idealized problems: (1) an opacity step in the.
(b - y) continuum, and (2) the "picket-fence" model for lines. Let us first ask quali-
RGURE 7-13 tatively what the effects of an opacity jump or strong lines might be.
Co?1parison of observed (dots) Str6mgren (c,. b - y) indices for If the opacity is grey, x. = K, + <1, and the emissivity is given by ri. =
main-sequence stars (which have log g ~ 4) [(516, 17; 516, 45)) with i<,B. + aJ,,; then the condition of radiative equilibrium is
calculated values from line-blanketetl models [(381; 516, 271)): note
good agreement in log g, which suggests that T., 1 can be estimated
accurately. K,
r B,(T0 ) dv = K, Jor J. 0 dv
Jo
00 00
(7-83),
206 Model Atmospheres 7-4 Results of LTE Model-Atmosphere Calculations 207

Here the scattering term has cancelled out. From the Eddington-Barbier where J, denotes fq,.J. dv for the ith line, and 1:1 1 is the frequency band con-
relation, we expect that at the surface (i.e., Tv « 1), / vO ~ B,(Tv ~ 1), hence taining that line. Again both terms in the braces are positive, so To must be
J .° ~ ½B.(Terr), Substituting this result into equation (7-83) yields the usual less than T 0. There is the additional feature that the effect of the lines depends
grey-body result TO 4 ~ ½T!rr• Suppose now there is a large jump in opacity on their thermal coupling coefficient e,. In LTE, e, = 1, J ~ ½B.(To), and
(e.g., at the Lyman edge) at some critical frequency, so that K = K, for v :.; v0 , for 11 » Kc, a large cooling term results; thus LTE line-blanketing must
K = ')'K, for v > v0 • Then the atmosphere will equilibrate to some new sur- drastically lower the boundary temperature. If the lines merely scatter the
face temperature T 0given by radiation (i.e., e, -+ 0), then (just as is the case for continuum scattering!)

K, s: 0
B.(To) dv + ')'K, I.: B.(To) dv = K, s: J. dv + ')'K, I.: J. dv
0
(7-84)
they have no effect upon the energy balance and the boundary temperature
is not changed markedly. We shall see that this.conclusion is also supported
by the detailed analysis to which we now turn.
Assuming that for v :.; v0 , J. ~ J.° (i.e., neglecting backwarming), and The qualitative results obtained above can be put on a quantitative footing
noting that for v > v0 , the surface value of J. ~ ½B(To), equation (7-84) by consideration of the illuminating treatment of line-blanketing offered by
can be rewritten as · the picket~fimce model proposed by Chandraskhar (150) a~d further d~ve!-
oped by Milnch (474). In this model we assume (a) the contmuum opacity 1s
K, So"' B.(To) dv ~ K, So"' J.° dv frequency-independent, (i.e., Kv = K); (b) the lines have square profiles of
constant width and a constant opacity ratio {J = 1/K relative to the con-
- K, [(')' - 1) f"'
Jvo B.(To) dv f"' J.° dv
+ Jvo Jvo J. dv]
- ')' f"' tinuum; and (c) the lines are distributed at random uniformly through~ut
the spectrum, such that within a given frequency band a fraction w1 contams
~ Kc fo"' J.o dv pure continuum, and a fraction w2 = 1 - w1 contains continuum plus lines.
(Alternatively, the probability offinding line opacity at a specified frequency
- Kc HI.: B,(To) dv + I.: [~ B.(T.~r) - ~.(To)] dv} is w 2 .) A pictorial representation of the problem is given in Figure 7-14,
which shows why the name "picket-fence" is appropriate. (A slightly different
interpretation of w1 and w2 allows treatment of an opacity step; see below.)
(7-~5) Adopting the continuum as the standard optical depth scale, we have for
....I.

....I.
Both of the terms in the braces are positive; hence we conclude that T 0 < T 0, I\)
and that .the amount of cooling is larger, the larger the value of y. This result
is not rigorous because we expect J. must, in fact, rise above J. 0 for v :.; v0 ; I
but we shall see below that a rigorous analysis verifies the correctness of our
conclusion. Suppose we evaluate T'(T.) at some point inside the atmosphere (P + l)K
where Tv » 1 for v > vOwhile Tv < 1 for v < vO• Then the mean intensities
in the square bracket of the second equality of equation (7-85) saturate to
the local Planck function and the whole bracket vanishes, and T'(Tv) equals
T0 for the grey case; i.e., the surface temperature drops only in those layers
where the opacity jump has become transparent.
Suppose now there are spectrum lines at frequencies {vi} that add to the w,
opacity, Xv = Kc + a + Li
11q,., and make both a thermal and scattering
K~--......,
contribution to the emissivity, riv = KcBv + aJ v + r,1,q,.[e,B, + (1 - e1)J .].
Then the condition of radiative equilibrium reduces to

Kc So"' B.(To) dv = Kc So"' J .° dv


- {~ 11e1[B,,(T0) - J,] + Kc~ t (J.° - J.) dv}
FIGURE 7-14
Picket-fence model. Lines are assumed to be a factor of Pmore
opaque than continuum, and to occur with a probability
(7-86) w2 = I - w1 in any frequency band.
208 Model Atmospheres 7-4 Results of LTE Model-Atmosphere Calculations 209

frequencies in the continuum, we find that k satisfies the characteristic equation


2 2 n
(7-87a)
and, in the line
L WmYm = L WmYm L aJl(l - k2µ/!Ym2) (7-95)
m=I m=I J=I

µ(d/~21/dt) = (1 + P)Jt2 > - (1 - r.)PJt2 > - (1 + r.P)Bv (7-87b) This equation yields 2n - 1 nonzero roots for k2 (bounded by poles at
1/µ/, ... , 1/µ/ and y2/µ 1 2 , ••• , y2/µ/) and hence 4n - 2 values fork of
Integrating over all frequencies (radiation quantities without a subscript v the form ± k1• In addition, we see by inspection that k2 = 0 is a root of the
denote integrated quantities) and accounting for the relative probabilities characteristic equation; this root yields a particular solution of the form
that the band is covered by line or continuum, we find
[jll = bw1(t + Q + µ;fy 1) (7-96)
(7-88a)
which may be verified by direct substitution into equation (7-93). The general
and µ(d/ 121/dt) = (1 + P)/< 2 > - (1 - e)pJ< 2 > - (1 + e/J)w 2B (7-88b) solution (7-93) is thus of the form
These equations are to be solved simultaneously with a constraint of radiative 2n-1 L e-k., 2n- I L_.ek•• )
equilibrium, which is obtained by integrating over angle and demanding that
Il0(t) = w,b ( t +
.
Q + µ, +
Y, •= 1
L •
1 + k.µ,/y,
+
•= 1
L _l -___.k.µ,/y,
; .__ •
po> + F(2) = constant, namely
(I = 1, 2)
(7-97)
JU> + (1 = [w 1 + w2 (1 + r.P)]B
+ r.p)J 121 (7-89) (i= ±1, ... ,±n)

Consider now the case of LTE (i.e., r. = 1). Let y 1 = 1 and y2 = 1 + p. Demanding that the solution not diverge exponentially as t -+ oo, we set
Then equations (7-88) become L_ 0 = Ofor all et. Requiring that the total flux
µ(dJ!l>/dt) = y,(J<ll - w1B), (I = 1, 2) (7-90) 2 n
F=2L I a1µi~0 (7-98) __.
lc!.J=-n __.
where, from equation (7-89),

B = L
2

/cl
2 •
y,J!I>/ L w,y,
le!
(7-91) we find b = (~4 F) / l=I ± w1y1-
1 (7-99) w
I

The constant Q and the L.'s are determined from the surface. bou~dary
To solve this system we use the discrete-ordinate approach, and choose conditions /<!! 1(0) = 0, which yield a linear system of 211 equations m 211
{111}, (i = ± l, ... , ±n), such that
unknowns:
=
(7-92) Q - (µ;fy 1) + J, L./(1 -
2n-1
k0 µ;/y 1) = ,
0 (I
(i
l, 2)
= l, ... , n)
(7-100)

Then, substituting equations (7-91) and (7-92) into (7-90), we have Using equations (7-99), (7-97), and (7-92), we find

¾F~
2n-1 • aj )
c1/ (1)
Jl1
i Jjll - ( W1 ) ~2 ~n /Cm) (I = I, 2) J!ll(t) = L WmYm
_1 ( t + Q+ L L.e
-~
L 1_ k 2µ 2/y 2 (7-101)
--= t.., y
Yr dt 2 Lm WmYm m= I m -n al J '
t..,
)= (i=±l, ... ,±11) •=I J=I • l I

(7-93) and, from equation (7-91),


If we now assume a solution of the form
B(t) = -3 F ( t + Q + L
2n- I
L.e-k••
)/
L 2
WmYm -
1 (7-102)
(7-94) 4 o=I m=I
7-4 Results of LTE Model-Atmosphere Calculations 211
2IO Mode/ Atmospheres
polynomial of order 2n - 1 in X; now we know that T(X) has 2n - I
We shall se~ below that, for the picket-fence model, (K/RR) = L wmYm - 1, so
nonzero roots Xm = l/km 2 so the polynomial must be of the form
from equation (7-102) we see that the asymptotic solution for B(t) is ¾FtR,
C(X - X ) • • • (X - X ln- 1). To evaluate C we equate the coefficients of
as would be expected. The Rosseland mean scale tR exceeds t; in the limit of 1

infinitely strong lines (y 2 -+ oo), tR(t) = t/w 1 and we see from equation the terms in x 2 •- 1 to find
(7-102) that the temperatures must be larger at depth. This is the backwarming
effect, and clearly depends mainly upon the bandwidth available for con-
tinuum flux transport.
Thus we have
Exercise 7-11: (a) Verify that equation (7-96) is a particular solution of the transfer
equation. (b) Verify equations (7-99), (7-101), and (7-102).

As was true for the grey problem, we may calculate the value of B(0)
explicitly. Define the function
ln-1 From the middle equality of equation (7-106), we have T(0) = L WmYm; and
S(x) =Q- x + :E L./(1 - k0 x) (7-103)
from equation (7-107), we have
••l
The boundary conditions (7-100) show that S(x) = 0 at the 2n values
x = µ 1/)' 1• But ifwe clear equation (7-103) of fractions by multiplying through
by a function composed of the product of the 2n - 1 denominators [i.e.,
R(x) = n::.
11 (1 - k0 x)], then the product R(x)S(x) is clearly a polynomial Combining these two results we then find, from equation (7-105),
of order 2n in x. But we know 2n zeros of S(x), hence the polynomial must
be of the form R(x)S(x) = C(x - µi) · .. (x - µ.)(x - µify) ~ .. (x - µ.h). S(0) =( L
2
Wm"lm - l
)½ (3 L 2 )-t
Wm"lm
(7-108)
m• 1 me 1
If we equate the coefficients of the two terms in x" on the two sides of
this equation we can evaluate C as C = k 1k 2 • • • k2._ 1, and hence obtain But comparison of equations (7-102) and (7-103) shows that
......
finally ......
. S(x) = k1 · · · k2n-1 [Ii Il
I• I I• 1
(x - µ,/)',)]/ ff (1 -
2

o• 1
k0 x) (7-104) B(0) = ~ FS(0)/(L WmYm - 1 ) (7-109)

which implies that Hence we conclude that


(7-110)
S(0) = k 1k2 • • • k2 ._ 1µ/µ/ · · · µ_2/y" (7-105)
Now consider the characteristic function This result may be restated in a form that reveals its physical content. The
Planck mean opacity is
T(X) = I m= I
WmYm [1 -r, a1/(l -
J• I
µ/ hm 2 X)] Kp = B- 1 fo
00
ICyByclv = B- 1 K(W1B + W2YB) = K(W1 + W2Y) (7-111)

= I
m•l
WmYm [1 - I X
J=I
aJl(X - µ//ym 2)] while the Rosseland mean opacity is
2 n (RR)- 1 = (dB/dT)- 1 f0 00 1
Kv - (dBvfdT) dv
= L WmYm -l L a1µ/ [(µ/hm 2
) - x]- 1
(7-106)
m= 1 J• 1 = (dB/dT)- 1 i-:- 1(w 1 + w 2 /y)(dB/dT) (7-112)
where X = l/k 2 • We clear equation (7-106) of fractions by multiplying (7-113)
through by the product of the 2n denominators. The resulting function is a or
212 213

Grey
I.I
0.8
1.0

0.9

0,8 7;, 0.7


B(tl r,,;
0.7
T
0.6 0.6

0.5

0.4 0.1 0.2 0.3 0.4 0.5 0.6 0.7

0.3
o.,,
FIGURE 7-16
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 Ratio of boundary temperature T 0 to effective temperature T., 1 as
a function of 0crr = 5040/T.rr • The break near 0.,1 = 0.25 results
FIGURE 7-15 from inclusion or the Lyman continuum in the high-temperature
Depth-variation or integrated Planck function in models. Upper line gives value of T 0/T,,1 for a grey atmosphere.
picket-fence models. Solid curve: grey solution,
=
p 1; solid docs: y = 10, w1 = 0.8, w2 = 0.2,
e"' 1 (LTE); rriangles: y = 10, w 1 =0.8, choose the values appropriate at T = T.rr· As pointed out by Munch
=
w1 .. 0.2, e 0 (pure scattering). Note (261, 38), this assumption is crude; but we employ it because it simplifies
backwarming effect in both blanketed models, the
the analysis while retaining the essential physical content. Consider the
large surface-temperature drop in the LTE model,
and the absence of a surface effect in the scattering results shown in Figure 7-16 for the ratio T 0 /T .,, from nongrey LTE model-
model. From (474), by permission. atmospheres calculations. For all 0.rr ~ 0.25, the Lyman continuum has
been omitted. For the coolest models T 0 /T.rr is near its grey value; this is ......
Thus equation (7-110) reduces to
not surprising because the dominant opacity source is H - , which is only ......
weakly frequency-dependent. At higher temperatures the effects of the Balmer
jump become important and T O drops below its grey value. At e.,r = 0.23 01
[B(0)/F] = (Jj/4)('i<.R/'f?.p)½ (7-114) I
the curve shows a sharp break caused by the effects of the Lyman continuum,
or (7-115) which is first included at that temperature. At higher values of T.rr, hydrogen
becomes more strongly ionized, the size of the Lyman jump diminishes, and
Now in the limit as y -+ oo, the Rosseland mean (being a reciprocal mean) the flux maximum moves beyond the jump, so T 0 /T.rr rises toward the grey
simply saturates at a value 1e/w 1 (which, in effect, shows the decrease of value again. At still higher temperatures T 0 /T err drops again as a result of
bandwidth available for flux transport) while the Planck mean increases the He I edge at l504 A and the He II edge at l226 A.
without bound; thus the effect of opaque lines in LT Eis to lower the boundary We can estimate the drop in the boundary temperature caused by the
temperature (in principle to very low values). An example is shown in Figure Lyman jump by applying equation (7-115). Assume that bound-free and
7-15, where B( t)/Fis plotted for the grey case and for one of M iinch 's solutions free-free absorption by hydrogen are the only sources of opacity. Using
withe = 1, w 1 = 0.8, w2 = 0.2, and y = 10; in this case B(0)/F decreases equation (4-124) for the free-free contribution, summing nra. •. 1(b - f) over
from the grey value 0.4330 to 0.286; i.e., T 0/T.rr drops from 0.811 to 0.721. all bound levels with u. = n- 2(x 100 /kT) ~ u = (hv/kT), using equations
The analysis just described can also be applied to an opacity jump at a (4-114) and (5-14), correcting for stimulated emission, and setting all Gaunt
critical frequency, v0 , beyond which the opacity increases by a factor of y. factors to unity, we may write the opacity coefficient in the form
We now apply the two versions of equation (7-90) on the ranges (v ~ v0 )
and (v ~ Vo), and define W1B = =
Jo0 B. dv and W2B I~ B. dv. We must K~ = cu- 3 (1 - e-") [1 + L 2u 1n- 3 exp(ui/n 2 )] (7-116)
then assume that w1 and w2 are constant with depth; for example we may u>u
215
214 Model Atmospheres

where the first term in the square bracket accounts for free-free and the 12
second for bound-free absorption. Because the Rosseland mean is a recip-
rocal mean, it will be essentially unaltered whether the Lyman continuum
is included or not. Thus we need calculate only the Planck mean with and II FIGURE 7-17
without the Lyman continuum, and use these values to estimate the ratio Temperature distribution for LTE and
of T O for these two cases. We take the limits of the integral in equation non-LTE models with T.rr = 15,000°K
and log g = 4. The atmosphere is composed
(7-1 ll)to be0and u0 , where u0 = u 1 when the Lyman continuum is excluded, ! of hydrogen, which is represented by a
and u0 = oo when it is included. Writing B, = C'u 3 e-•(1 - e-•)- 1, I

------- - ---,.- ,, schematic model atom with two bound

9
/' levels and continuum. This model atom
/ - NL TE, no lines accounts for the· Lyman, Balmer, and
Kp(u 0 ) = C"{t - e-•0 + ~2u 1n- 3 [1 - exp(-u 0 + n- 2 u1)]} (7-117)
/ --- LTE. no lines free-free continua, and the Lyman-Gt line.
/ ...... NL TE, Let From (40), by permission.
Now for e.rr = 0.23, u, = 2.3 x 0.23 x 13.6 = 7.2; if Uo = 00, the expo- 8 .•.,.;l -·-·· LTE. I.~
nential term is zero identically, while if u0 = u1 it can be neglected unless
n = 1 because u 1 » 1. Thus -8 -7 -6 -5 -4 -3 -2 -I
log r

Rp(oo)/R,.(u 1) = (1 + 2u 1 f
n•l
3
n- )/(1 + 2u 1 f
n•l
n-
3
)

a further drop to 7800°K. Further lines would produce yet additional cooling;
= (1 + 2.4u 1)/(l + 0.4u 1) (7-118) the non-LTE results will be discussed in §7-5.
If we now consider scattering lines (e '# 1), the results obtained above
For u 1 = 7.2 we thus find R,.(oo)/R,.(u 1 ) = 4.7, hence change radically. Define l = =
(1 + e/J) and <1 (w 1 + lw 2 )- 1 • Then equation
(7-89) becomes B = cr(J(I> + u< 2>) and equations (7-88) become
T 0 (Lyman cont.)/T0 (no Lyman cont.)= (4.7)-t = 0.825 (7-119)
__.
In Figure 7-16, extrapolation of the results without Lyman continuum to µ(di 11l/dt) = [(I) - W1<1(J(I) + u< 2 >) (7-88a') __.
0.rr = 0.23 yields to T 0/T.rr ~ 0.65, while T0 /Terr ~ 0.56 when the Lyman µ(d/12)/dt) = y/12) - (')' - (7-88b') CJ)
and W1<1..1.)Jl2) - W2<1LJl1)
continuum is included, which gives a boundary temperature ratio of 0.865,
I
in good agreement with equation (7-119) (considering all the approximations
Applying the discrete-ordinate method, we obtain the characteristic equation
that have been made). It should be noted that this temperature drop occurs
(474)
only at very shallow depths where the Lyman continuum becomes trans-
parent. At optical depths of even 10- 4 in the visible, the Lyman continuum
is opaque, and temperatures in models with and without the Lyman
continuum are practically the same.
where (7-121a)
A further illustration of the cooling effects of LTE continua and lines is
given in Figure 7-17, which shows the temperature structure in a model
atmosphere with Terr = 15,000°K, log g = 4, consisting of hydrogen sche- 1 n

matized as a two-level atom plus continuum (40). The transitions allowed


and H = 2,f•a,/(1 + kµ,M (7-12lb)
in this atom are La., the Lyman and Balmer continua, and the free-free
continuum. The temperature plateau at T ~ 10,300°K for - 4 ~ log t ~ - 2 Equation (7-120) has 2n - 1 positive roots k«. The solution for B(t) is
occurs where the Balmer continuum is optically thin but the Lyman con-
tinuum is thick; this temperature gives a "T0 "/T,rr ~ 0.68, in fair agreement
(7-122)
with Figure 7-16 where the Lyman continuum is omitted. Including the
Lyman continuum drops T 0 to 9400°K; adding the La. line (alone) produces
216 Model Atmospheres
7-5 Non-LTE Radiative-Equilibrium Models 217
where M« = uLa[Ga(l - G«)- + ()./y)Ha(H« - 1)- 1] and where, in turn,
1

the 2n - 1 constants L« and the constant Qare determined from the boundary problem of constructing models in which the populations of the atomic
conditions 11~1 (0) = 0, which imply levels and the radiation field are computed by self-consistent solutions of
the equations of transfer and of statistical equilibrium. To understand fully
2n- I
some of the difficulties inherent in this problem, the student should, ideally,
Q + W1-I L La(l - Ga)- 1(1- k«µ;)- 1 = µ;, (i = l, ... ,n) already have mastered the material in Chapters 11 and 12; on the other hand,
«=I
some of the material presented here provides background for those chapters.
(7-123a) It is recommended, therefore, that this section be read again after Chapters 11
and and 12 are studied. In this section we shall follow a somewhat "historical"
2n- I approach in developing methods that treat, first, continuum-formation alone,
Q + (}'W2)- 1 L
«=I
L«<H« - W 1(1 - k«µ;/y)- 1 = (11,/y), (i = I, ... , 11) and then, a final method that treats both continuum and lines. We shall
not describe the line spectrum here (cf. §12-4), but will discuss the effects
(7-123b) oflines mainly from the point of view of energy bala.nce.
The fundamental difficulty in the solution of the non-LTE model-
Exercise 7-12: Verify equations (7-120) through (7-123). atmospheres problem is that the occupation numbers in the outer layers
of the atmosphere are determined mainly by radiative rates. Thus the state of
A solution obtained by Munch for w1 = 0.8, w2 = 0.2, y = 10, and the material is only weakly coupled to local conditions (e.g., temperature and
e_ = 0 is sh~wn in Figure 7-15. Here one finds that the boundary temperature_ density) and is dominated by nonlocal information contained in the radiation
hes only shghtly below its grey value, _with B(0)/F = 0.4308 compared to field, which responds to global properties of the atmosphere, including
the grey result 0.4330. Thus lines, when. formed by scattering, have almost boundary conditions. We shall see that the mathematical manifestation of
no influence upon the boundary temperature; that is, the effect of lines upon this physical circumstance is that the source functions implied by the equa-
the boundary temperature depends sensitively upon the mechanism of line- tions of statistical equilibrium contain dominant (noncoherent) scattering
formation. The backwarming effect is, of course, still present because the terms. We have already seen (§6-1) that these terms introduce mathematical
frequency band for free-flowing radiation has been restricted. In fact, the difficulties into the solution of the transfer problem.
backwarming effect is almost identical in the two cases, which shows that The first approaches to the non-LTE model atmospheres problem used _..
backwarming is determined mainly by the frequency bandwidth blocked by the an iteration procedure, which was successful only for continua in which the
lines, and but little by details of the line-formation process. It is important scattering terms were not large. Subsequent approaches attempted to solve --...J
to realize· that LTE line-blanketing cools the surface layers (and produces the transfer equations simultaneously with the rate equations by introducing I
darker lines); but scattering lines are also dark (cf. §10-2) even when there is information from the latter explicitly into analytical expressions for the
no cooling at the boundary. It is not valid to argue for low values of T source functions used in the transfer equations. Scattering terms reduce the
in a stellar atmosphere just because observed lines have dark cores for i~ degree of the coupling of the material to the local thermal pool and hence
general, the lines may be decoupled from the local temperature distributi~n also tend to introduce nonlocal information into the energy-balance criterion.
and their central depths may have nothing to do with T 0 • We shall retur~ Thus it becomes difficult to satisfy the requirement of radiative equilibrium.
to this point again in our work on line-formation. Finally, it is interesting As we noted earlier in our discussion of temperature correction procedures
to note that under certain circumstances the abrupt introduction of an (cf. 117-2), even small errors in energy balance may severely affect the solution
opacity edge can cause local heating in the atmosphere [ cf. (198)]. of the statistical equilibrium equations. It is thus necessary to find methods
that apply the constraint of radiative equilibrium in addition to the simul-
taneous solution of the transfer and rate equations. Initially this was done
7-5 Non-LTE Radiative-Equilibrium Models by a linearization procedure for the temperature alone; this procedure works
for Early-Type Stars when there is fairly direct coupling to the temperature structure (as there
is for continua via radiative recombinations) but fails for lines where neither
The methods and results described thus far in this chapter have been based the emission nor absorption rates depend directly upon temperature. For
on the simplifying assumption of LTE. We now turn to the more general models including lines it becomes necessary to make a sweeping general-
ization to a complete linearization procedure that places all physical variables
218 Model Atmospheres 7-5 Non-LTE Radiative-Equilibrium Models 219

of interest on an equal footing and accounts for the global interactions of of terms of the form [n 1R(J - ni(ni/n1)*R11 ]; we thus eliminate the most
all variables throughout the atmosphere. troublesome terms from the equations at the outset. Physically, the approxi-
mation proposed here recognizes that photons first "see" the surface in the
most transparent regions of the spectrum (i.e., in the continuum), and that
SOLUTION BY ITER·A TION: DETAILED BALANCE
free escape out of the atmosphere in these bands leads to departures from
IN THE LINES
LTE at the greatest geometrical depths in the star. The simplified continuum-
. The first attempts to construct non-LTE model atmospheres employed only problem leads, therefore, to the correct asymptotic behavior at depth,
an iteration procedure in which one (a) starts with estimated occupation and provides a starting point for the solution of problems that include the
numbers (say from LTE, (b) uses these to compute the radiation field, and line terms.
then (c) uses this radiation field to compute radiative rates in the statistical In practice, the iteration procedure treats the departures from LTE as a
equilibrium equations, which are then solved for a new estimate of the level perturbation away from the LTE state. Comparison of equations (7-2) and
populations. In practice it was found that this lambda iteration procedure (7-4) shows that with the line-terms omitted, departures from LTE do not
failed (283, 217) when lines were included. The lines are very weakly coupled affect the expression for the emissivity [if by nr, the LTE population of
to local conditions (see Chapter 12) and are very opaque; therefore the severe level i, we mean the value calculated from the Saha-Boltzmann equation
problem of radiative control of the populations over a very large range of (5-14) using the actual (non-LTE) electron and ion densities]. Comparison
optical depths is encountered. Just as described in §6-1 for the archetype of equations (7-1) and (7-3) shows that (again omitting lines) we can write
scattering problem of the transfer equation, the iterative process then Xv = x: + f>x •. If we define b, = nifnr, then f>x. = Lt
d,nrix,K(v) where
stabilizes to a spurious value without converging, and successive iterations d1 = b1 - 1. Now suppose that at any stage of the calculation we regard
differ but slightly, even though the current estimate is far from the true as given both T(m) and either (a) the values of b1 for all bound levels or
solution. (b) the values of all radiative continuum rates. We may then integrate the
It is therefore of interest to inquire whether it is possible to treat only hydrostatic equation in the usual way, and solve for the electron and ion
the continua and to ignore, or at least defer, treatment of the lines. The densities using either essentially the same formalism as in §5-2, but with
continua are basically simpler because (a) they are strongly coupled to local nr replaced with b,ni throughout, or the linearization method in §S-5 with
thermal conditions (via recombinations), and (b) they are relatively transpar- all terms in f>T and f>Jk set to zero. The latter method yields a consistent
ent down to depths where densities are high enough to assure domination by current value for n, and n100 ; the former does not, for it ignores the non-
collisions (a~d hence recovery of LTE). This means that the self-consistency linearity in n, in the collision rates. In the work cited below where this
problem occurs in regions that are not optically thick, and hence that the iteration method was employed, the former alternative was used, and the co
iteration procedure has a chance of working (these remarks do not apply whole process iterated to convergence. I
in the Lyman continuum, which is as difficult to handle as the Jines). An We next solve the transfer equation
affirmative answer to the question posed above was given by Kalkofen
[(283, 175; 34S; 346); see also (424)), who showed that, for early-type stars, (7-124)
the Lyman and Balmer lines are so opaque that, at depths where the visible
or (7-125)
continuum is formed, they can be expected to be in radiative detailed balance.
In this case the bound-bound radiative rates upward and downward essen-
tially cancel each other. In particular, for T.rr ~ 104 °K, it is found that the where e.
= K:/(x: + f>x.) and ,. = n.a./(x: + f>x.). using any technique
detailed balance criterion is met for continuum optical depths t 5000 ~ 10- 4 , that can handle the electron scattering term correctly. This yields new
which implies that the continuum is already formed (i.e., is optically thin) values for the radiation field, which are used in the rate equations to solve
before the lines go out of detailed balance; thus the continuum-formation for new departure coefficients. For detailed balance in the lines, the complete
problem can be treated essentially independently (except for the Lyman rate equations (S-87) reduce to
continuum, which is about as opaque as the Jines). This result is valuable,
for it offers an opportunity to assess the importance of departures from LTE
from continuum observations alone.
d, [4n
Jvo I
1"' (ix,J ./hv) dv + ,. C,1]
1 1
- I
)"I
d1C1J
Mathematically, radiative detailed balance implies that in the rate equa-
tions (5-87) we may analytically cancel out (or, equivalently, omit) all pairs
= 4n f,"'•o (B, - J ,)(1 - e-hv/kT)(ix./hv) dv (7-126)
221
220 Model Atmospheres

where / denotes the last bound level and c/1 = (b1 - 1). As before, we can
20..-------~-----.-------i------
solve these linear equations for the d1's if we regard ne and n10n as fixed, or
we can iterate ne to consistency.

Exercise 7-13: Verify equation (7-126); note that, unlike equation (5-87), both
upward and downward collision rates appear.

Further, temperature correction may be performed to enforce the requ~re-


ment of relative equilibrium; in the papers cited below this was done usmg 0
the Avrett-Krook procedure [ equations (7-19) and (7-20)], modified trivially
e.
to use the generalized definitions of and C. in equation (7-125) (specifically,
C• .,, l - e.). With the new estimate of t_he temperature structure _and
departure coefficients, the hydrostatic equilibrium and transfer equations
may be solved again and the whole process iterated to convergence.
A number of models of the type described above have been constructed
to study the effects of departures from LTE in the continuum and the
observational implications of these departures [(283, 217; 348; 610; 425; -4
426; 427; 452)]. In the earliest work, very substantial changes (decreases) I and 2
in the Balmer jump were predicted and it was suggested (610) that these _gL__ _ _ _ _.J__ _ _ _ _...J...__ _ _ _ _ _ ~1-----
effects might explain the then-existing discrepancy between observed and -4 -3 -2
computed Balmer jumps (see discussion in §7-4). Subsequent work using
log t,ooo
more refined atomic models and better collision cross-sections snowed that
FIGURE 7-)8
departures from LTE have negligible effects on the Balmer jump in main- Non-LTE departure coefficients in the first 7 states or hydrogen ror a model with
sequence B-stars (important effects remaining for supergiants and 0-stars), T.rr = 10,000°K and log g = 3. Ordinate gives c/1 x 102 ;_ curves are labeled "".ith ~uantum
and ultimately the discrepancy was removed by a change in the fundamental number of the level of the model atom. Note that level 2 1s unclerpopulated while higher levels
calibration of the energy distribution of Vega. Nevertheless, the first papers are overpopulatecl. Levels I and 2 are locked together by the assumption or detailed balance ......
were important, for they stimulated interest, and called attention to the in the Lyman continuum. co
possibility that departures from LTE could have observable consequences in I
the continuum. We defe~ further discussion of the observational implications
On the other hand, for n ~ 3, hv/kT ~ 1, and the dilution factor of½ in
to later in this section.
J,, outweighs the effects of the temperature gradient, so J, < B, a~d the
Results for the departure coefficients d1 of the first six levels of hydrogen
levels are overpopulated. In Figure 7-18 we have d 1 = c/ 2 because 1t was
in a model with T.rr = 10,000°K, log g = 3, are shown in Figure 7-18.
assumed that the Lyman continuum was also in radiative detailed balance;
Here we see that the deviations are, in fact, fairly small at depths representa-
in this case the radiative rates in equation (7-126) for n = I cancel analyti-
tive of continuum formation. The n = 2 level is underpopulated, while levels
with n ~ 3 are overpopulated. The values of dn decrease rapidly for large n
cally, and we are left with dt LJ=i CIJ = IJ=2
diC, 1 which implies d, ::::: d2
because C 12 » C 1 . for j > 2. Thus, collisional coupling of n = 1 ton= 2
because (a) the collisional ionization rate~ for highly excited levels become
allows the upper level to drive the same departure into the ground state
large and force recovery of LTE, and (b) the radiation field at low frequencies
population. Closer to the surface, where the Lyman continuum comes out
is dominated by free-free processes-which are purely thermal, and again
of detailed balance, the n = 1 level becomes overpopulated (see below). At
tend to force recovery of LTE. Then n = 2 level is underpopulated because, higher temperatures, characteristic of the O-stars (i.e., T.rr ~ 35,000°K), the
for that level at the temperatures prevalent in the model, hv 2 /kT » 1, and the situation is different, for now n = 2 becomes overpopulated, and the ground
increase of temperature into the atmosphere implies that B, ~ exp( -hv/kT)
state n = 1 becomes underpopulated at depths where the Lyman continuµip
rises rapidly. Thus at the surface J, :==: ½B.(T.rr) exceeds B,(To), and the is formed. This would be expected on the basis of the scaling of hvo/kT,rr
level is preferentially photoionized as can be seen from equation (7-126).
222 Model Atmospheres
7-5 Non-LTE Radiative-Equilibrium Models 223
mentioned above; these results may also be understood (346) in terms of
the anticipated variation of the flux with depth in the various continua The ratio r is generally much smaller than unity, and the essence of equation
(recall that dH,/dr, = J, - S,). Finally, it should be emphasized that the (7-128) is that the Lyman continuum source function differs from its equi-
departure coefficients obtained from the procedure described here cannot librium value by a factor of 1/b 1 . We now assume that the opacity ratio r
be used to calculate line profiles, for they lead to spurious results, as would is evaluated using the current value of b1 (as is the optical depth scale r 0 ),
be expected because the level-populations will be inconsistent with the values but in equation (7-128) we substitute an analytical expression for (1/bi)
they would have in the presence of lines (43). obtained from the ground-state statistical equilibrium equation:

FORMATION OF THE LYMAN CONTINUUM

The calculations described above assume that the Lyman continuum


(1/bi) = [41t f.: (a.,/hv)J,dv + C]/[41t s.: (a..fhv)B,dv + C] (7-130)

is in radiative detailed balance. But at some point in the outer atmosphere, where C = C 12 + C IK' Substituting equation (7-130) into (7-128) shows
the Lyman continuum must begin to become transparent, and significant that s.
is of the form
transfer effects occur that force n1R 1" to depart from n"R~it and hence lead
to an uncoupling of the n = 1 state from the n = 2 state. This situation
(7-131)
becomes most relevant at high values of Terr, where the high degree of
ionization of hydrogen implies that the Lyman continuum is weakened to
the point of being only somewhat {rather than markedly) more opaque than That is, the source function consists of a noncoherent scattering term (i.e.,
the visible continuum. Application of the iteration method to the Lyman intensities at all frequencies in the continuum are coupled) and a thermal
continuum fails, and we can gain some important physical understanding term (i.e., a term independent of the radiation field). By straightforward
of the problem {and also a preview of the problems of line-formation) by numerical estimates [see, e.g., Table 5-1 and equation (5-44)] it is easy to
analyzing why this is the case. We shall see that one must account for the show that the thermal term is small compared to the scattering term.
information in the statistical equilibrium equations by introducing them To solve equations (7-127) with (7-131) correctly, it is necessary to perform
directly into the transfer equations in such a way as to yield a simultaneous a direct solution (e.g., by the Feautrier or Rybicki methods) allowing fully
solution of the two sets of equations. for the scattering term. If equations (7-127) and {7-131) are solved directly, ........
Consider the following simplified problem. Represent the model hydrogen we have in principle obtained a simultaneous solution of both the transfer I\)
atom by two. bound states and continuum, and assume that departures and the statistical equilibrium problems. Because the scattering term appears
from LTE occur only in the ground state. Let the Lyman continuum threshold explicitly in the transfer equation, the correct solution is obtained (for 0
frequency be v0 ; consider only frequencies v ;;i, v0 , and suppose that a given run of the thermal term) over the entire range of optical depth in a I
hv 0 /kT » 1 so that stimulated emission can be neglected. Ignoring electron single step, and the slow convergence properties of lambda iteration are
scattering and Gaunt factors, the ground-state, upper-state, and free-free avoided. In practice it is still necessary to iterate because the optical scale
opacities all have the same "profile" cf,, = (v 0 /v) 3 • Writing n 1 = b 1nf, we and the overlap parameter r are evaluated at each stage of the calculation
then have x. = Xo<P, = (b1xT + x:)</J., where the superscript • denotes LTE using current estimates of b1, but experience shows that this iteration
values, and the subscript u denotes the contribution of the upper-level and converges immediately. Calculations using an approach of this type (426)
free-free continua. Similarly, tf, = (tff + tf:)</J. = (xf + x:)B,q,,. Let dr 0 = for B-stars show that characteristically the ground state becomes over-
- Xo dz be the optical depth at the continuum head. Then the transfer populated in the outer layers (though the results in the reference cited are
equation to be solved is only schematic because the constraint of energy balance was not adequately
satisfied-we shall return to this point below).
= cf,.(]. - S,)
µ(dl,/dr 0 ) (7-127) We may gain considerable insight into the problem by noting that, for
hv/kT » 1, the frequency variation of J. and B, [roughly as exp( -hv/kT)]
where s. = (xf + x:)B,/(bixT + x:> = {[(1 - r)/b1] + r}B. (7-128)
shows so rapid a drop with increasing v that practically all of the contribution
and r = 1J/(b1xT + x:) (7-129) to the photoionization and recombination rate integrals comes from v ~ v0 •
Therefore, we replace the integrals by 41tw 0 (a. 0 /hv 0 )J O and 41tw 0 (11. 0 /hvo)Bo,
224 Model Atmospheres 7-5 Non-LTE Radiative-Equilibrium Models 225

where w0 is an appropriate weight factor, and consider the transfer prob- where (omitting bound-bound transitions)
lem only at the continuum head, v = v0 [see also (195)]. If we define
e = C/[4nw0(a 0/hv 0)B0 + C] then
Xv = L (n; - nre-hfkT)IX,"(v) + L n,n"IX""(v, T)(l - e-hv/kT) + n,u.
I "

(7-136)
S0 = (1 - r)[(l - e)J0 + eB0] + rB 0 :::::: (1 - e)J0 + 'lB 0 (7-132)

where 'l = e + r, and the transfer equation in the Eddington approximation is


and l'/v = (2hv 3/c 2 )e-hv/kT [ ~ nrrx,,c(v) + ~ n,n,ctX,c,c(v, T)]
= Bv [ ~ ntrx;"(v) + ~ n,n"rx""(v, T)] (1 -:- e-hv/kT) = K~Bv
(7-133)
(7-137)
Assuming 'l is constant, and that the atmosphere is nearly isothermal so that
B 0 is roughly constant, the solution of equation (7-133) can be written We now divide the spectrum into a series of·characteristic frequency ranges
immediately [cf. equation (6-9)] as (v10 ~ v ~ v11 ) upon which we assume the opacity and emissivity of a
particular continuum i -+ " is much larger than all others. For example we
J0 = Bo{V + 1 - exp[ -Jj'i-r 0]}/(1 + V) (7-134) might assume that the Lyman continuum dominates for all frequencies such
that l ~ 912 A, the Balmer continuum dominates for 912 A ~ l ~ 3650 A,
which shows that (a) the solution thermalizes only at depths ~ 1/V, and the free-free continuum dominates for l ;;;i, At, the threshold of the last
(b) the departure coefficient at the surface is b 1(0) :::::: e-t. Because 'l « 1, bound level, etc. We then factor out the dominant terms from both the
the departure from LTE is large, and pe,:sists to great depths. The actual emissivity and the opacity, and write, for (v; 0 ~ v ~ v11 ),
value of 'l is a bit difficult to estimate if collisions dominate (i.e., e > r), for
the density, and hence e, increases exponentially into the atmosphere; on (l'/vlXv) = [nrf(n; - nre-hvjkT)Je,vBv = e,.B./(b, - e-hv/kT) (7-138)
the other hand, the parameter r is almost depth-independent, and often where the three quantities
.....I.

~
dominates. For example, at T 25,000°K, r ~ 10- 2 [see (633, 193)] and I\)
b1 ~ 10, which is in agreement with calculation. In summary we see that e,v = [(n1 - nre-hv/kT)/nt](K~/X,) (7-139) .....I.
it is important to use the information inherent in the statistical equilibrium
equations directly in the transfer equation; further, given the dominant f., and •• are all evaluated using current values of the temperature and level
scattering term, it is obvious that one must use care in satisfying the constraint populations, whereas the term (b 1 - e-h•JkT) in equation (7-138) is to be
of radiative equilibrium. replaced with an analytical expression, involving the radiation field, to be
We now shall consider a partial /inearization method that is fairly general, found from the statistical equilibrium equations. The statistical equilibrium
and provides a satisfactory means of solving the non-LTE continuum- equation for level i is [cf. equation (7-126)]
formation problem, with energy balance; the method fails, however, when
lines are included, in which case one must use the complete linearization
method described in the next subsection. The basic idea in the method
b1 (R," + ±
)~I
C,1) = R" 1 + C1" + t
J~i
b1CIJ (7-140)
[ cf. (40; 41)] is to incorporate the statistical equilibrium equations into the
transfer equations by manipulating them analytically to obtain explicit where R 1" and R" 1 are defined, as usual, by equations (5-66) and (5-67).
expressions for source functions, and to incorporate the constraint of Solving for b; - e-hvJkT we then have
radiative equilibrium into the transfer problem by linearization. Then, in
principle, one solves all three sets of equations: transfer, statistical and
radiative equilibrium, simultaneously.
(b; - eh••JkT)-1 = ( R," + J, C;1)/[ R,cl - e-hv/kTR,"
The transfer equation is

(7-135)
+ ±
1-1
(bj - e-hv/kT)C,1 + c,"(1 - e-hv/kT)] (7-141)
226 Model Atmospheres 7-5 Non-LTE Radiative-Equilibrium Models 227

so that, defining C, = n,<J,/X, (again using current values), the source function where
in equation (7-135) on the range (v10 ,i;; v :i;; vu) can be written as

S, = ~,. (;,,."J::: <I>,.J. dv + e,.B.) + C.J. (7-142)

where <I>,. = 4mx1K(v)/hv E1 = (±i= k=k,o


I
I wkxk,1ke1kBko)/E2 (7-148)

I k11
and E2 =L L wkxk,ikelk(8B 0/8T)k (7-149)
i=lk=km
(7-143)
Thus equation (7-144) becomes
and D1, is the term in square brackets in the denominator of equation (7-141).
In the evaluation of y,. and e1, we again use current values for the b/s of d 2(f,,.Jk)/d,/ = (I - (k)Jk
other levels and for the terms in J. and T; it is thus clear that a large amount
of information is lagged in this technique. Note that e1, contains an overlap
integral from those frequency ranges outside the presumed range of domi-
- ,,k [Yik
lzk,o
Iw,<I>,,J, + e,k(aBo/aT)k ±
k'=I
Wt,t/Jk.Jk,]

nance (v10 , vu) of the transition i -+ ,c; usually this integral is small compared - ,,keik[Bk O - (8B 0 /8T)kE1] (7-150)
to the term displayed explicitly in the scattering integral. Equation (7-142)
accounts for the most important terms directly, but because the information which is completed with boundary conditions given by equations (7-37a)
involving other levels and in the nonlinear terms is lagged, iteration will still and (7-37c) (the latter specifies the flux explicitly).
be required. Finally, discretizing, we have equations of the fo'rm Equation (7-150) has the following important properties. (a) It incorpo-
rates both the equations of statistical equilibrium and radiative equilibrium
in terms of the new (i.e., as yet undetermined) radiation field; thus a solution
of the system satisfies both of these constraints automatically. (b) Cancella-
tion of large scattering terms is done analytically in the coefficients [cf. I\)
on the range· (k 10 :i;; k :i;; ku) delineated for the ith transition. espcially equation (7-147)] of the radiation field before the transfer equation I\)
We must now solve the system (7-144) subject to the constraint of radiative is solved; hence good control of the residual thermal terms is obtained. I
equilibrium which requires that (c) The equations are of the standard Feautrier form (equation (6-31)] with
the frequency coupling spread over the entire spectrum, and can be solved
K
f' x.(J. - S,) dv = L WtXkJk -
I k11
using the usual elimination scheme of equations (6-39) through (6-41). (Why
k•I
L k•koo
l•I
L WtXk is the Feautrier method more efficient than the Rybicki scheme in this
problem, even though the number of frequencies is large?)
x [~,t (i',t l•k,o
~ w,<I>,,J, + e,tBk) + CtJk] = 0 (7-145) The whole solution proceeds iteratively. After solving equation (7-150)
for J., the temperature distribution is revised via equation (7-146), the new
radiative rates are used to solve the statistical equilibrium equations for new
We suppose that B., the value of the Planck function that does yields radiative occupation numbers, new Eddington factors are then calculated from the
equilibrium, can be written in terms of the current value B.° as B, == B, 0 + revised values of 11. and x., and the procedure starting with equation (7-139)
(cB.° /8T) t:. T. Substituting into equation (7-145) we find is repeated.
Methods of the type described above (which is related to the equivalent-
K
two-leve/-atom approach for line-formation problems; see §12-2) have been
t:.T = L wkt/JtJk - E1 (7-146)
developed by Feautrier (211) and Auer and Mihalas (40; 41). Feautrier's
k•l
228 7-5 Non-LTE Radiative-Equilibrium Models 229

the atmosphere, and {b) we obtain reliable level-populations that can be


used to compute line profiles (to be compared with observation) only if line
transitions have been taken into accbunt in the solution of the statistical
equilibrium equations. The importance of these effects is illustrated in
Figures 7-19 and 7-20. In Figure 7-19 we see that the effect of including Lr,,
is to drain the n = 2 level, which goes from being slightly overpopulated to
markedly underpopulated. The drop in b 1 is caused by the drop in tempera-
ture; actual occupation numbers n 1 do not change much [see (40)]. In
Figure 7-20 we see a similar effect for b2 and b3 when Hrt. is taken into account,
If we had computed the Hrt. profile with the continuum-only values (which
are similar to those of Figure 7-18) we would have found a spurious emission
0 core in the line (because n = 3 would have been too overpopulated relative
logb

2.0 ,---,,---,.--,--..,...........-..---.---.-..----,-..---,---,

-1 - La. 1.8
--- No line

1.6

1.4
-2'-~---.....- ~ ~ - ~ - ~ - ~ - ~ -
-9 -8 -7 -6 -5 -4 -3 -2
log t __..
1.2
FIGURE 7-19 I\)
Non-LTE departure coefficients in the lirst twa states of
· hydrogen for a model with T.rr = 15,000°K and log g = 4.
w
Dashed curves: Lyman, Balmer and free-free continua only; I
full curves: uc line included. The optical depth scale in the
abscissa is measured just longward of the Balmer jump. 0.8
From (40), by permission.

0.6
formalism differs significantly in appearance from that displayed above, but
the physical content is about the same. These methods have been used
successfully to construct continuum models for B-stars, including the Lyman
continuum in cases where lambda iteration would be utterly hopeless.
-7 -6 -5 -4 -3 -2 -1 0
Characteristically the ground state of hydrogen becomes strongly overpopu-
lated in the outer layers; see Figure 7-19. This has important consequences log t
for the temperature structure and emergent spectrum of the atmosphere, as FIGURE 7-20
Non-L TE departure coefficients in the second and third level of
will be discussed below.
hydrogen in a model with T,rr = 15,000°K and log g = 4. Dotted
Attempts have been made (40; 41) to apply the techniques described above curves: Lyman. Balmer. Paschen, and free-free continua only; full
to cases including line transitions. It is essential to include the lines because curves: H(X included. Optical depth scale as in Figure 7-19. From
(a) they drastically affect both occupation numbers and energy balance in (41), by permission.
230 Model Atmospheres 7-5 Non-LTE Radiative-Equilibrium Models 231

to n = 2); inclusion of the line from the outset drains n = 3 into n = 2 and these variables is intricate and essentially complete. For example, we have
produces a strong absorption line (we shall compare with observations in already seen in §5-5 that a change in the radiation field at any frequency
§12-5). implies a change in the occupation numbers of all bound states [ cf. equation
The above method does not work well for determining the temperatu;e (5-108)]; further, any change in a radiation field implies a change in the local
structure when lines are included; thus in (40), the solution for the tempera- temperature and density, and so on. Second, the variables interact globally
ture structure including LIX tended to stabilize rather than converge, and in throughout the atmosphere, and a change in any variable at a given point implies
(41) a solution including HIX (which is formed at about the same depths as changes in all other variables at all other points. Thus, if we alter occupation
the Lyman continuum) tended to become strongly unstable. The reasons numbers locally, this changes the emissivity and opacity of the material, and
for the failur can be tra~d to basic inadequacies of the method. (1) Many hence, via the transfer equation, the radiation field throughout the atmo-
7
of the terms m the equations are lagged, and interactions among transitions sphere; when scattering terms dominate the source fµnctions, information
are treated only iteratively rather than collectively from the outset. (2) The propagates over large distances in the material, and the global nature of the
temperature dependence of many variables (e.g., the opacity) is ignored in interaction is accentuated markedly. A truly adequate method must, there-
the linearization procedure. (3) Most important, it is presumed that the fore, put all variables on an equal footing, allow fully for all possible couplings
temperature is, in some sense, "the fundamental variable" of the problem. among the physical variables resulting from the imposed physical constraints,
Such a presumption is adequate for treating continua that are coupled and for the interaction of each variable with all others at every point in the
strongly to the temperature via radiative recombination but, as noted before atmosphere via transfer equations and requirements of energy and momen-
is totally unsatisfactory for lines, where both emission and absorption rate~ tum balance.
are decoupled from the temperature structure. Let us now consider a method To determine the solutions \j,4 we require a total of K + L + 3 equations;
of great generality and power that overcomes these difficulties entirely. we may choose these to be K transfer equations (Jk, k = 1, ... , K); the
equation of hydrostatic equilibrium (N); the equation of radiative equilib-
rium (T); a total particle-number conservation condition (ne); and L statis-
THE COMPLETE LINEARIZA TION METHOD tical equilibrium equations specifying level-populations, total abundances
of chemical species, and total charge conservation (n 1, ••• , ni), exactly as
From a physical point of view, the solution of the non-LTE stellar atmo-
written in §5-4 [see equations (5-91) through (5-93)]. The essential difficulty
spheres problem entails the specification, at each point in the medium, of,
to be faced now is that these equations are. nonlinear and must, in general, I\)
the distribution of the radiation field as a function of frequency, the tempera-
be solved by some kind of iteration procedure. In particular, we suppose that
ture and density of the material, and the distribution of atoms and ions over ~
our desired (but as yet unknown) solution \j, 4 can be written in terms of the
all bound states. These distribution functions are to be determined in such
current (but imperfect) solution \j,/ as \j, 4 = \jt/ + 6\j, 4 • We then choose I
a way that the constraints of energy balance, momentum balance (i.e. 6\j,4 so as to satisfy all constraints more closely; i.e., iff4'\j, 4) = 0 represents
h_ydrostatic equilibrium in the present work), steady-state statistical equilib~
the entire system of constraints, we demand that f4'\j,/ + 6\j, 4) = 0, and
num, and charge and number conservation are all satisfied rigorously. To
solve for 6\j, 4 by linearing the entire system:
ac~ieve the desired result, it is important to recognize two basic physical
pomts, and to construct a generalized method in the light of them. First,
no one variable is more ''fundamentaf' than any other, for they all interact. (7-152)
Thus we must regard the solution at a given depth-point m4 to consist of
the vector If we express the transfer equations as difference equations, and the con-
straints in terms of quadrature sums, the f's are coupled algebraic equations
\j,4 = (J1,,, •, Jx, N, T, ne, n 1, •.• , nL)T, (d = 1, ... , D) (7-151)
and the linearized system reduces to the standard block tridiagonal form of
equation (6-31). In the step-by-step elimination scheme we account (con-
w~ere D is the number of depth-points, K is the number of frequencies, and
sistently to first order) for the interaction of va.riables at one depth-point
L ts the total number of bound levels (of all kinds) considered. (Note that
with all other depth-points, subject, ultimately, to the requirements of the
there is, in principle, redundant information in \j,4 as written; e.g., given all
imposed boundary conditions. The global nature of the problem is thus
the J's, the n,'s follow from the rate equations. However, all of the information
is of interest in practice, and hence will be retained.) The coupling among all taken fully into account.
232 Moc/c,f Atmospheres 7-5 Non-LTE Radia1ive-Eq11ilibrium Models 233

The complete linearization procedure described here is a generalization or (5-104)] is


that described in §7-2, and treats an enlarged set of constraints as well as a
K
larger set of basic physical variables (42). As before, we introduce a depth
discretization {md} and frequency discretization {v.}. We suppose that a
ond - (8n/8n,)don,,d -(on/oT)doTd - L (8n/8Jk)doJdk =0 (7-157)
k=I
starting solution lj,/, (d = 1, ... , D), has already been obtained (assuming,
say, LTE). The transfer equations are again equations (7-37a) through (7-37c)
and the linearizations of these equations have already been written in
where (7-158)
equations (7-39) through (7-47) and in the results of Exercise 7-9. [See also
(437, 22-32)]. The essential difference is that we must now use the general
formulae of equations (7-1) and (7-2) for x. and 17., and write for any variable x [see equations (5-105) through (5-108) and expressions in
(42) and (437, 38-47)]. Here we assume that n is already a solution of the
L
= (ox./oT)d oTd + (ox./on,)d on,, d + ,L (ox./on,)d on,. d current system dn = /?4.
OXdn
... (7-153)
The complete system is of the general form

and a similar expression for Ol'/dn [ detailed expressions for all these derivatives (7-159)
may be found in (437, 51-57)]. Further, we no longer regard the level popula-
tions as functions of (N, T), but rather as independent variables whose which may be solved by the standard Feautrier elimination scheme.
coupling to T, J., etc. is specified by the rate equations.
For the hydrostatic and radiative equilibrium equations we again use equa- Exercise 7-/4: (a) Sketch the form of the A, B, and C matrices, indicating nonzero
tions (7-9) and (7-13) as linearized in equations (7-53) and (7-52), respectively, elements by x's as was done in §5-4 for the rate equations. Show that the A and C
employing equation (7-153) for Ofo and a similar equation for Ol'/dn• The matrices are void below the row specifying hydrostatic equilibrium, and differ for
that row [see (450, 130)]. (b) For typical problems, the number of depth-points
equation of particle number conservation can be written as
D ~ 70, the number of frequencies K ~ 100, and the number of constraints
L L + 3 ~
15. Show that despite the size of K, it is more economical to use the
Nd= n,,d +L n1,d Feautrier elimination scheme than the Rybicki scheme. ......
1•1
I\)
If we assume the same hydrogen-helium mixture as described in §5-4 and let In the system written above, L, is the residual error in the constraints 01
Y denote the (number) abundance of helium relative to hydrogen, we can °
found when current values \j,4 are used; as L., -+ 0, the corrections 6\j,4 -+ 0. I
rewrite equation (7~154) in a simpler form: We have already stressed the physical implications of the complete lineariza-
tion. Mathematically, the equations are internally self-consistent, and are
equivalent to a generalized Newton-Raphson procedure; convergence, if
(7-155) obtained, should therefore be quadratic. This is not achieved in practice
because, after each set of corrections 6\j, have been determined and applied
which, when linearized, yields to the solution, it is necessary to recompute the Eddington factors (which
were assumed fixed under linearization) from a formal solution using current
source functions, as was done in the LTE case as well. Nevertheless, an
order-of-magnitude reduction in the 6\j,'s is often obtained from successive
iterations, and on the whole the procedure is stable and efficient, and it
= N, - n,,d - (1 + Y) [np,d + I
1=1
n,,d(H)] (7-156) easily handles physical problems that defeat the other techniques described
above. It appears that at present the complete linearization method is the
best technique available for solving non-L TE stellar-atmosphere problems.
Finally, the rate equations and charge-conservation equations are of the form The method can handle multilevel, multiline, multispecies problems of great
given in §5-4, and their linearized form [see equations (5-102) through generality, and treats all the physical constraints because the full interaction
234 Model Atmospheres 235

among all variables is allowed, and a priori assumptions about the depen-
dence of quantities upon only a restricted set of variables (e.g., T) are avoided. 0
I
The complete linearization method has been used with good success to I
I
construct models in the temperature range from A-stars (42; 43; 368; 224) 0.10 I
I
I
through O-stars (436;45). The early work, carried out before the introduction
of the variable Eddington factor technique (44) made use of the Eddington
\I
I
approximation (i.e., one angle-quadrature point) and has been supplanted by I
I
I
the later calculations. An extensive grid of models for 0- and B-stars is I
0.20 I
available (430; 432). In this work, departures from LTE are taken into account
for the first five levels of hydrogen, the first two levels of He I and He II,
17.5 \ Q3
and for an "average light ion" (which represents C, N, and 0) consisting of \I
I
five stages of ionization, each with a ground-state only. These models \
\
typically allow for six hydrogen-line transitions: La, LP, Ly, HrJ., HP, and PrJ. 0.30 \
\
for the O-stars, and HrJ., HP, Hy, PrJ., PP, and BrJ. for the B-stars (for which \
(u - b) \
the Lyman lines may be set in detailed balance), and yield results for the \
physical structure of the atmosphere, continuum parameters, and H-line \
\
profiles. \
\
\
0.40 \
\
NON•L TE EFFECTS ON ENERGY DISTRIBUTIONS
15 \
b
Departures from LTE affect both the continuum and the line-spectrum
from a stellar atmosphere; the discussion in this section !ocuses on the
continuum only. Enough results are available for models of early-type stars 0.50
to allow us to delineate the regions of the H-R diagram where departures ......
from LTE have important effects and where they can be ignored. At the' I\)
present time the non-LTE models do not include line-blanketing effects;· 15 °
(J)
thus we compare LTE and non-LTE unblanketed models to determine -0.11 -0.10 -0.09 -0.08 -0.07 -0.06 -0.05 -0.04 -0.03 -0.02
differential effects (which then, presumably, can be applied as corrections to I
(b - y)
values of parameters obtained from blanketed LTE models). Extensive sets
of results for continuum jumps and Stromgren-system colors are given in FIGURE 7-21
Theoretical S1r!lmgren-sys1cm colors for LTE and non-LTE models.
(430; 516,241), and complete energy distributions are given in (432); we shall Ordinate: (11 - /,); abscissa: (b - y), LTE values are represented by open
summarize some of the principal results here, and we suggest the reader circles and dashed curves; non-LTE values by solid dots and curves. Curves
3
examine the literature cited for further detail. are labeled with log g, and individual models are labeled with T,rr/10 ,
For the B-stars (15,000°K " Terr " 30,000°K) the effects of departures
from LTE in the visible spectral regions are generally found to be negligible
for main-sequence stars, but become important for giants and supergiants could have significant consequences in certain applications. The effects of
(low gravities). Non-LTE effects on colors are shown in Figure 7-21. There departures from LTE upon the continuum jumps at the Lyman, Balmer, and
we see that results for both LTE and non-LTE models at a given gravity lie Paschen edges (DL, D 8 , and Dp) are fairly substantial for B-stars. The Lyman
along the same curves, but the position of models with given T.rr is somewhat jump is increased by non-LTE effects for most B-stars because the ground-
different. Thus if we ignore departures from LTE, we introduce a systematic state is overpopulated (b 1 » 1) and hence the flux for 2 < 912 A_is reduced;
error into estimates of T.rr· For main-sequence (log g = 4) stars the errors these changes are significant in making estimates of the far-ultrav1olet energy
can be neglected, whereas for log g = 3 the errors are about 200°K and at output of these stars. Balmer jumps are generally decreased by non-LTE
log g = 2.5 they are about 500°K; these errors are small but systematic and effects in B-stars because b2 < 1 while b3 > 1, so the opacity contrast across
236 Model Atmospheres 237

the Balmer edge is reduced by departures from LTE. The Paschen jump is 0.25

\
only slightly affected because both b3 and b4 > 1, and the opacity ratio stays
about the same. In a plot of D8 vs (b - y) (or some other flux ratio), the LTE 0.24

and non-LTE models at a given gravity lie on the same curve, but slightly
shifted (recall Figure 7-21). If one assigns values of Terr to stars using such a
diagram, the departures from LTE again imply systematic errors in estimates
of r.,,; these errors are negligible for main-sequence stars, are about 350°K
0.23

0.22
\.
at log g = 3, and are about 500°K at log g = 2. \ • , . , . 10'

-~~
From the point of view of obtaining direct observational evidence for 0.21
departures from LTE in the continuum, the results mentioned above are not tf,
helpful because there is not a clear discrimination between the two cases.
Strom and Kalkofen (611) pointed out that the parameter <f, = Dp/D 8
0.20
·~ .
·~·, ·~
___
provides a sensitive observational indicator of non-LTE effects, as may be 0.19
seen in Figures 7-22 and 7-23. The LTE models predict values of <f> of about
4
0.16 to 0.17, independent of gravity, whereas the non-LTE models predict 0.18 ·--. --.._:.:__.
g • 10 •

much larger values of¢, increasing with decreasing gravity. This effect was
found observationally (583); the supergiants have systematically larger values 0.17
of <I> than main-sequence stars with the same value of D 8 • In the observational
system there were some difficulties of calibration, so only a differential com- 0.16
0.50 0.60 0.70 0.80 0.90
0.00 0.10 0.20 0.30 0.40
parison was possible; these difficulties should be surmountable with the
De
FIGURE 7-23
0.24 r-.--,--,--,--,--,--,--.--r-r-r-r-r-r-r-r--- Same as Figure 7-22 for middle B-type stars. Shaded area contains predictions of
2 LTE models at various values or g; curves (labeled with surface-gravity g) show
0.22 non-LTE values.
I\)
. 0.20 -..._J
new Vega calibration. Further, departures from LTE have major effects on
I
the ultraviolet continua of A-stars (587; 588) for which it is found that the
0.18
ground-states of CI and Sil are strongly underpopulated, and the flux ob-
tained from LTE calculations is too low by very large factors. Only when the
0.16
non-LTE effects are taken into account is a satisfactory agreement with
space observations obtained for these stars.
0.14
For O-stars (T.rr ;:;:; 30,000°K) the effects of departures from LTE on
visible continuum parameters and colors are much more important. For
0.12
these stars, as described earlier, the ground state hydrogen becomes under-
0.10 ...._.....__.....__.....__.....__.....__.....__,__i_,..._i_,..._L_._L......JL_JL......JL-..JL-...l
populated, hence the flux below the Lyman limit increases, and the Lyman
0.6 0.7 0.8 0.9 1.0 I.I 1.2 1.3 1.4 1.5 jump decreases. In contrast, at the ).227 A grourid-state edge of He II, the
flux is decreased by non-LTE effects because n = l of He+ is overpopulated
Da
[see (45; 432)]. These changes are of importance in estimating the energy
FIGURE 7-22 .output from O-stars into, say, nebulae or the interstellar medium. The most
Theoretical continuum-jump parameters for LTE and non-LTE models
or late B-type stars. Ordinate: tf, e D,/D 8 ; abscissa: Balmer jump D8 • significant change in the visible region is that the Balmer jump predicted by
Solid dots and curves: non-LTE values; open dots and dashed curves: the non-LTE models is about 0.07 mag larger (for log g = 4) than that from
LTE values. Curves are labeled with log g. LTE models, and shows but little variation with gravity, whereas the LTE
238 7-5 Non-LTE Radiative-Equilibrium Models 239

very large changes in the ultraviolet flux below the hydrogen Lyman edge,
0.30
and in the resonance continua of certain light ions such as CI and Sil. For
later-type stars one must examine departures from LTE in the H- ion. With
D8 (mag)
present estimates of the relevant reaction rates (556; 188), the predicted effects
(607) are negligibly small for G and K main-sequence stars and giants
0.20 [see also (491; 492)]. Virtually nothing is known about the possible impor-
tance of departures from LTE in other types of stars (e.g., M giants or super-
giants), and much work remains to be done to evaluate these effects. Finally,
the problem of accounting for non-LTE effects in line-blanketing remains
0.10 to be attacked.

TEMPERATURE STRUCTURE: THE CAYREL MECHANISM


AND LINE EFFECTS
0.00
As we have seen in Chapter 3, the temperature in a grey atmosphere
3S decreases uniformly outward to a limiting value T 0/T.rr = 0.811. Further,
30 40 45 so for a nongrey atmosphere that has a large opacity jump or strong lines, under
T.,r/10 3 the assumption of LTE the boundary temperature falls below the grey value
FIGURE 7-24
(exceptions: scattering lines leave the boundary temperature unchanged and,
Theoretical Balmer jumps for O-stars. Ordinate: D8 in magnitudes; abscissa: in certain special cases, "new" absorbers can cause small temperature rises).
3
T.,,/10 • Solid symbols and curves: non-LTE models; open symbols and dashed
curves: LTE models. Curves arc labeled with log g. From (45), bt permission. In sum, the basic prediction for LTE radiative-equilibrium atmospheres is
that the temperature distribution is a monotone decreasing function outward.
Often the temperature structure exhibits plateaus where the then-dominant
models predict emission edges at low gravities (see Figure 7-24). The absolu\e transition (e.g., the Balmer continuum) has become optically thin while other __..
sizes of the non-LTE Balmer jumps are in excellent agreement with observa- transitions (e.g., the Lyman continuum and Lyman lines) are completely I\)
tions of O~stars if the Hayes calibration for Vega is adopted (as now seems opague, followed by a series of drops as successively more opaque transitions (X)
correct). Further, t~e differential behavior of low-gravity versus high-gravity become optically thin in turn. I
non-LTE Balmer Jumps agrees with observation while the LTE results do For non-L TE atmospheres the situation is quite different, and typically
not (405), and the same conclusion holds true for the differential behavior of the temperature distribution passes through a minimum and then shows a
O-star versus BO-star D8 's (45); both of these results are independent of the rise outward. The basic reason this occurs was pointed out by Cayrel (142;
calibration and support strongly the non-LTE calculations. The fact that the 283, 169), who called attention to the similarity between the physical con-
Balmer jump remains about constant in strength for the non-L TE models ditions in the outer layers of a stellar atmosphere and those in a nebula
but weakens markedly in the LTE models implies very large differences in the surrounding a star. Suppose the energy balance is completely determined
computed colors [say (u - b) and (b - y)] of the two sets of models (516,241). by absorption and emission from a single state (e.g., the sole bound state
At the highest temperatures, the discrepancy in T.rr for a given value of of H-, or the ground state of hydrogen). Then, in LTE, the condition of
(u - b) is some 15,000°K. Comparisons of Terr derived from colors with those radiative equilibrium can be written
derived from helium-line strengths supports the non-LTE results. The
comparison could be much refined yet by using a large body of data and nf 1"' rxvB,.(T 0 )dv = nf J~
('«> rxJvdv = nfW J~
('«> rx,B,(TR)dv (7-160)
comparing the O-stars differentially with B-stars (thus eliminating calibr~tion
J~
problems) [see also discussion in (516, 241)]. where the last equality introduces a parametric representation of J, in terms
In summary, departures from LTE have significant effects on the visible of a dilution factor W < 1 and a radiation temperature TR· As W < 1, it
cont_in_uum parameters of O-stars and B-type giants and supergiants, but are is clear that T O ·< TR; in particular, if W = ½and TR = Terr, T O has essen-
neghg1ble for B-type main-sequence stars. Deviations from LTE produce tially the grey value. The result just obtained follows because we have forced
240 Model Atmospheres 7-5 Non-LTE Radiative-Equilibrium Models 241

LTE. But now suppose that we allow departures from LTE; then to make an a priori estimate of an appropriate value of TR in the Lyman
continuum, for it, unlike H- which is nearly grey, shows a sharp fall-off in
nT J~
f"' a..B.(T0 )dv = b 1nT J~
f"' a..J.dv = b 1nTW J~
f"' a..B.(TR)dv (7-161) opacity with increasing frequency, which implies that the surface radiation
at high frequencies emerged from deeper, hotter layers and is characterized
In the limit of low density the departure coefficient is determined by the by larger values of TR· The final result is established by some kind of average
photoionization and recombination rates [cf. equation (5-95)], and of TR over frequency. Similar results are obtained at other effective tem-
peratures (40; 211); detailed discussions of how particular transitions are
b1 = Jvo
f"' (a..B.fhv) dv/ Jvo
f"' (a..J .fhv) dv (7-162) affected may be found in the references.
Let us now consider the effects of lines on energy balance in the non-LTE
case; here again it will repay the reader's effort to reread this material after
Hence combining (7-161) and (7-162) we have
Chapter 11 has been studied. A qualitative feeling for the results to be

[L: a..B.(T 0 )dv/ L: a.,B,(T 0 )v- 1 dv]


expected follows immediately from the form of the non-LTE source function
for a collision-dominated line (cf. §11-2), namely
S1 = (1 - &1) ]1 + &1B, (7-164)
=[ L: a.,B,(TR) dv/ L: ex,B.(TR)v- 1
dv] (7-163) where J denotes the average of J, against the line-profile <f, •. By an argument
(625) exactly analogous to that leading to equation (7-86), we find a result
which shows that T O = T th independent of the value of W ! Thus we expect that differs only trivially from that obtained there, even though the line-
typically the temperature will decline from T = Terr near t ~ 1, to a value source function has a noncoherent rather than coherent scattering term.
T min• approximately equal to the grey-body boundary-temperature, and We therefore reach the same conclusion as before: for LTE (& = 1) there is
then rise to a value T 0 , with T min < T O :S Terr· These res~lts can also be a large drop in boundary temperature, but when e « 1, the lines have prac-
viewed in terms of the "quantity" and "quality" of the radiation field (239). tically no effect on the boundary temperature. The argument is verified by a
That is, in LTE, it is the energy density (i.e., quantity) of the radiation that detailed picket-fence calculation (448) that allows the continuum to adjust
fixes T O ; in the non-LTE case, the energy density may be lower than its self-consistently. In this work the line-strength is assumed to remain un-
equilibrium value, but each photon has an energy characteristic of TR and' changed, which would be valid, say, for the·resonance lines of the dominant I\)
hence can still ionize the material and deposit an excess energy, per ionization,' ion of a given species (e.g., Ca II in the solar atmosphere). co
which is again characteristic of TR [see also (240)f A very detailed calculation of non-LTE line-blanketing effects in the solar I
For the sun, T 0 r, = 5900°K, and Cayrel estimated T O = TR ~ 5600°K atmosphere (17) leads to the conclusion that a self-consistent model yields
compared to T min ~ 4800°K (grey value). From a detailed calculation of a T min ~ 4330°K, and that the outward tempe~ature rise dr!ven by the Cayrel
solar radiative equilibrium model, allowing for departures from LTE in the mechanism in the continuum is strongly resisted by the Imes. In early-type
H- ion, Feautrier obtained (211) T min ~ 4700°K and T O ~ 5200°K. Similar stars the effects of lines upon the energy balance can be studied in non-LTE
effects are found for 0- and B-stars in which the main source of surface models constructed with the complete-linearization technique (references
heating is the Lyman continuum. For example, in Figure 7-17 we saw that given above); several interesting results emerge. For example, in Figure 7-17
the LTE temperature structure of an atmosphere with T.rr = 15,000°K, we see the effect of including Lex (only) in addition to the Lyman continuum.
log g = 4, had a plateau at 10,400°K, where the Lyman continuum was In LTE the boundary temperature drops about 1600°K, from 9400°K to
optically thick and all others were transparent, followed by a drop to T O = 7800°K; but in the non-LTE case the boundary temperature had risen to
9400°K when the Lyman continuum became transparent. In the non-LTE T ~ 10,350°K (owing to heating in the Lyman continuum), and inclusion
model, the temperature for 10- 4 ;$ t :S 10- 2 lies below the LTE value or° Lex produces a temperature drop of only about one-third that obtained
(because b2 < 1 in the Balmer continuum and therefore the efficiency of in LTE, to 9800°K. This result is not surprising, for e is « 1. The final tem-
heating is reduced). The non-LTE temperature distribution shows a minimum perature structure is relatively complex in the non-LTE case.
near 10,100°K and then a rise outward to TO ~ 10,350°K, which lies 1000°K An even more interesting example is shown in Figure 7-25, in which we
above the LTE surface value. The agreement of this value of T O with the see the importance of the coupling between lines and continua. Here the
LTE Balmer plateau temperature is probably fortuitous. In fact, it is hard model hydrogen atom had three levels, and the formation of the first three
242 243

--- LTE, no lines 28


12 ····· NLTE. H2 T,rr = 30,000°K
--·- LTE, Hi
FIGURE 7-25
Temperature distribution for LTE and 26 FIGURE 7-26
... ,
non-LTE models with T,rr = 15,000°K T/103 Temperature distributions in models with
II T,rr = 30,000°K .. Ordinate: T/10 3 ;
T and log g = 4. The atmosphere is composed
w or hydrogen, which is represented by a 24 abscissa: logarithm or column mass in
gm cm- 1 . Solid curve: non-LTE model
schematic model atom with three bound
levels and continuum. This model atom including lines; dashed curve: non-LTE
accounts ror the Lyman, Balmer, Paschen, 22 continuum-only model; dotted curve:
and rree-free continua, and the Hcc line. LTE model. Cross-bars on curves mark
From (41), by permission. ----+---
He II - .... "f. H :
optical depth unity in most opaque region
or transition indicated (those with ion
20 He 1' f designations are ground-state continua).
_f H From (436), by permission.
-7 -6 -5 -4 -3 -2 -1 0 He II He 1./
18
log t ............I••············~
-7 -6 -5 -4 -3 -2 -I 0
continua, plus free-free, and the H11. line was treated (41); the Lyman lines logm
were omitted because they are formed in the very outermost layers, whereas
H11. is formed at about the same depth as the Lyman continuum and can
lem of non-LTE line-blanketing in stellar atmospheres has barely been
interact with it. In LTE, H11. decreases the boundary temperature from its
approached, and a large amount of work remains to be done.
continuum-only value of 9400°K to about 8900°K. In the non-LTE case,
however, inclusion of H11. raises the boundary temperature from about w
10,500°K {the continuum-only value) to 11,200°K; this is a decidedly non- 7-6 Extended Atmospheres 0
classical result! The line itself makes a negative (i.e., cooling) contribution I
to the energy balance equation. But at the same time it provides an efficient For all of the models discussed thus far, it has been assumed that the atmo-
channel for atoms to fall into the n = 2 level, where the radiation field sphere is stratified in plane-parallel layers; this is an excellent approximation
produces strong heating; thus the direct cooling effect of the line is out- when the density scale-height in the atmosphere is small compared to the
weighted by indirect effects of the line on continuum energy balance, a radius of the star. However, many stars, in particular the supergiants and
possibility recognized by Cayrel (143). Addition of higher Balmer and Wolf-Rayet stars, have extended atmospheres whose thicknesses are an
Paschen-series lines (42) raises the temperature still further; the effect of appreciable fraction of a stellar radius; to a first approximation we suppose
H11. alone is about equalled by inclusion of H{J and P11., and yet-,higher lines that these atmospheres are spherically symmetric. Atmospheric extension has
lead to only a small additional rise. A final example in Figure 7-26 shows important physical and observational implications. Thus stars with extended
the temperature structure in a model with T.rr = 30,000°K, log g = 4, (436) envelopes show a continuum energy distribution that has an anomalously
including the L11., L{J, Ly, H11., H{J, and P11. lines. The temperature in the LTE low radiation temperature in comparison with the excitation temperature
model decreases uniformly outward; that in the continuum-only non-LTE inferred from spectral lines. Equivalently, the energy distributions are
model shows heating in the H Lyman and the He I and He II ground-state "flatter", as a function of frequency, than those of main-sequence stars (which
continua; the non-LTE model with lines shows enhanced heating from the have compact, planar atmospheres) of the same spectral type, and show excess
Balmer lines followed by a drop produced by the Lyman lines. emission in the infrared and a deficiency in the ultraviolet. Almost always
The variety and complexity of the effects just described emphasizes the there are indications of rapid atmospheric expansion in stars with extended
need for carrying out physically consistent analyses with great care. With envelopes, so that one should consider dynamical models that include the
the exception of the sun, for which a fairly detailed analysis exists, the prob- effects of flow; we shall examine such models in Chapter 15, but for the
7-6 Extended Atmospheres 245
244 Model Atmospheres

solution will be obtained merely by patching together asymptotic results for


moment will deal with the more limited problem of solving the transfer
the two limiting regimes mentioned above; this provides a good starting
equation in static extended envelopes.
point for estimation of extension effects on the emitted energy distribution.
In an extended, tenuous atmosphere, the radiation field at large distances
from the underlying stellar disk becomes very dilute, and is confined primarily
to a narrow solid angle {that subtended by the disk) around the radial di- SPHERICAL GREY ATMOSPHERES
rection. These facts imply that the temperature structure of the atmosphere
must be quite different from that in a planar model, and that mathematical The problem of grey spherical atmospheres in LTE and radiative equi-
complications introduced by the angular peaking of the intensity will be librium was analyzeu approximately by Kosirev (360) and Chandrasekhar
encountered. The equation of transfer to be solved {cf. §2-3) is (148); numerical results of high precision have rece1,1tly been obtained by
Hummer and Rybicki (323). If we assume x. = X, and integrate equations
µ(ol,/or) + r- 1(1 - µ 2 )(fJI,/fJµ) = Y/, - x.J. (7-165) (7-166) and (7-167) over all frequency, omitting the subscript v to denote
integrated quantities, we obtain
with moments (cf. §2-4)
r- 2 [fJ{r 2 H)/fJr] =0 (7-168)
r- 2 [o{r2 H,)/fJr] = r,. - x.J. (7-166) ·
and [o(JJ)/or] + r- 1(3/ - l)J = -xH (7-169)
and (fJK,/or) + r- 1(3K, - J ,) = - x.H. {7-167a)
where in equation (7-168) we have demanded radiative equilibrium
or [o(f.J,)/or] + r- 1(3/. - l)J, = -x,H, {7-167b)
It is obvious that these equations are more complicated and more difficult f0
"' r,. dv = f0"' x.J. dv = xJ (7-170)
to handle than their planar counterparts. For example, equation (7-165) is a
partial differential equation involving explicitly two independent variables. Equation (7-168) yields an integral for the total flux,
Similarly the moment equations do not yield a simple form if, for example, r2 H = H 0 = L/16n2 (7-171)
we eliminate H, between equations (7-166) and (7-167) {though a trans-
w
formation will be described later that does allow a reduction of the two
equations to a single combined moment equation of an attractive form).
where L is the luminosity of the star. Define the optical depth, measured
radially from an arbitrarily large outer radius R, to be
.....
The moment·equations could be dealt with, at least approximately, if we
could relate K, and J • accurately, as is possible in the planar case. But now,
even though f. ➔ ¼at great depth where the radiation field is homogeneous,
t(r) = r x(r') dr' (7-172)

f. ➔ 1 near the surface [recall Exercise (1-12)], and a direct application of Deep within the atmosphere (i.e., r « R, t » 1), we expect the radiation
the Eddington approximation, which gives remarkably good results in the field to become isotropic and f ➔ ½;in this limit equation (7-169) becomes
planar case, will not be even roughly correct near the surface. As was re-
cognized by Mc Crea (413) and Chandrasekhar (148), it is possible to con- (7-173)
struct a consistent approximation scheme in a tenuous atmosphere that
envelopes a parent star of radius r •• by considering averages of the radiation which yields the integral
field computed on the range µ• ~ µ ~ 1 [where µ. = (1 - r• 2/r 2)½]
separately from those computed on the range -1 :Iii µ :Ii; µ. [see also (403)].
But such a method is suitable only if r • can be chosen unambiguously; for
J(t) = H0 (3 f; r- 2 dt' + C) (7-174)

extended photospheres of appreciable density and nonnegligible optical If the usual Eddington-type boundary condition at t = 0 could be applied,
depth this method breaks down [see, however, (635)]. It is thus not surprising
then J(O) = 2H(O) = 2H 0 /R 2 so that
that effective and general methods for treating transfer problems in spherical
geometry have been slow to develop. A general and flexible numerical
technique giving a direct solution of equations (7-165) through (7-167) will (7-175)
be presented later in this section. To treat the grey problem, an approximate
246 Model Atmospheres
7-6 Extended Atmospheres 247
a result obtained by Chandrasekhar (148). If, further, we impose the assump-
tion ofLTE, then = rr. x.B. xB.
= and from equation (7-170) it follows that Equation (7-182) reveals an important characteristic difference between the
J(t) = B('t) = aT4 /n, which assigns a temperature structure to the atmo- planar and spherical cases, namely T - 0 as t - 0 in an extended atmo-
sphere. Equations (7-174) and (7-175) can be valid only at great depth. Near sphere, rather than approaching a finite value. From this result we can see
the surface, the free-flow regime occurs and f - 1, in which case equation that the contribution from the outer cool layers, which occupy a large volume,
(7-169) becomes will enhance the flux observed at longer wavelengths and lead to the dis-
tinctively flatter energy distribution mentioned above.
[8(r 2J)/8r] = -xr 2H = -xHo (7-176)
We may calculate the flux received by an observer at a great distance D
which yields
from the center of the star by using the (p, z) coordinate system shown in
J(t) = r- 2 H 0 (t + C) (7-177)
Figure 7-27. The impact parameter p is the perpendicular distance of a ray
In the limit f = 1, J(0) = H(0) = H 0 /R 2 , so equation (7-177) becomes from a parallel ray passing through the center of the star; z is the distance
along the ra)', measured from the plane through the cente_r. of the star per-
(7-178) pendicular to the central ray. We shall take z to be pos1t1ve towards the
observer and formally place the observer at z = oo for purposes of calculating
a result that is expected to be valid only when t « 1, r ::::: R. integrals. The coordinates (p, z) are related to polar coordinates (r, 0) as
Considerable progress can now be made if we adopt a power-law opacity follows: z = r cos 0, p = r sin 0, and r = (p 2 + z 2 )¼. Ifwe choose a value of
(i.e., x = c.r-•), as was done in the earliest work (360; 148). As noted by p, then the equation of transfer along the ray in the direction of increasing z is
Kosirev, there is strong physical motivation for this choice in an expanding
atmosphere, for the equation of continuity (see §15-1) demands that pvr 2 = (iJIJ8z) = 17v - xJ. (7-183)
constant, where p is the density, and v is the velocity of expansion; in the
which follows from first principles. The formal solution of equation (7-183)
limit of very rapid expansion, when v > v••••P• (actually observed), the ma-
for the intensity emerging at z = oo along the ray with impact parameter p
terial moves at practically constant velocity, so that p r-~ The opacity of ~ may be written immediately as
the material can be expected to vary as some power of the density (e.g.,
linearly for electron scattering. or as the square for free-free) and hence as /v(p, oo) = f~«> Bv[T(p, z)] exp[ -t(p, z)]X(P, z) dz (7-184)
some power of 1/r. Substituting the above expression for x into equation
(7-172) and adopting, for simplicity, R = oo, we have where ,(p, z) is the optical depth measured from z = oo, inward along the ray. w
I\)
t(r) = c.r-<•- 1>/(n - 1) (7-179) I
Using equation (7-179), the limiting form for J(t) (as r - 0, t » 1) is

(7-180)
while equation (7-177), valid fort < 1, remains unchanged. It is therefore
attractive (387) to interpolate between these two extremes with an expression
of the form

J(t) = ( -3H
r2
0
- )(n -
-1)[t + -31 -
n+l
+-1)]
n-1
(n (7-181)

Comparison with precise calculations demonstrates that equation (7-181)


is quite accurate (323). Using the identity of J and B, and writing T 1 for the
temperature at t = 1, we can rewrite equation (7-181) as
I 0
\ - ~ Observer
( (z= 00)
T(t) = T 1t 21
" -
11
{[. + (n + 1)/3(n - 1)]/[l + (n + l)/3(n - l)]}¼
FIUURE 7-27
(7-182) Coordinate systems for solution of transfer equation in spherical
geometry.
248 Model Atmospheres 7-6 Extended Atmospheres 249

Exercise 7-15: Calculate the operator (/J//Jz)p and show that equation (7-183) is one with T 1 = 30,000°K and n = 5. On the basis of these results we can
exactly equivalent to equation (7-165). In the (p, z) coordinate system, the lines p = understand why supergiants and WR stars have lower color temperatures
constant are the characteristic rays of the partial differential equation, which than main-sequence stars of the same spectral type; it is also evident that
reduces to an ordinary differential equation along these particular curves. atmospheric extension effects introduce ambiguities in the choice of a
structural model for a star because of the trade-off between temperature
The total flux received by the observer per unit receiver area is and envelope-size in fitting the data. The frequency-variation of the flux
from an extended grey atmosphere is shown in Figure 7-28, where it is
(7-185) compared to that from a Planck function at T., the color temperature of
the flux at ,l, = 5000 A, and to a Planck function at T(t = ½) (the distribution
Following Kosirev (360) [see also (61, 165)] we change variables in equation that would emerge from a planar atmosphere). It is clear that the flux from
(7-184) to O = cos- 1(z/r) and write t(p, z) = t(p, 8) which, using a power-law an extended atmosphere shows a pronounced ultraviolet deficiency and an
opacity, becomes t(p, 8) = C"p-ln- 0 t{l"(8) where

(7-186)
Then -1.0 ··············
•.......... B,(T213)

I.(P, oo) = C"p-ln-o f 0


" B.[T(p, 8)] exp[ -t(p, 8)] sin"- 2 8 d8 (7-187) -0.5
F• ....·• ··•... ·•....··,.......

Substituting equation (7-187) into equation (7-185), and introducing suc-


-- -----
--
0.0
cessive transformations from p = r sin 8 to t(r), the radial optical depth
given by equation (7-179), we obtain I.)
--
.,...,..,. B,(T,)
+ 0.5
/. = n(Ri/D)2 Jo"' B,[T(t)]t{~\,"(t) dt (7-188) _g .......
.,.,
,.,;
1.0 w
where R 1 is chosen as the radius at which t = 1, and I w
I
<I>n(t) = 2J0" exp[ -(n - l)t csc•- 1 8 t/1~(8)] sin 8 d8 (7-189) 1.5

Exercise 7-16: Derive equations (7-186), (7-188), and (7-189). 2.0

Using the temperature law given by equation (7-182) in equation (7-188), 2.5
we may calculate the flux from a grey spherical atmosphere for a particular
choice of T" a characteristic temperature for the atmosphere, and the index 3.0 ____.......__ ___....__.....__ __._ _...__ __._ __,
n, which determines the degree of extension (n -+ co implies planar models; 4.0 3.0 2.0 1.0
n small yields large extent). Observationally, color temperatures T, are
assigned, at a wavelength ..t,, by measurements of colors in a filter-system, 1/).
or of the spectrophotometric gradient; both of these approaches, in effect, FIGURE 7-28

measure the slope of the continuum. Van Blerkom (61, 165) has calculated Solid curve: emergent flux F, from spherical grey atmosphere with
T 1 = 5 x 104 °Kand n - 2. Dashed curve: black-body curve at
T, at ..t, = 5000 A for various models with T1 = 50,000°K. He finds that, color temperature T, that matches slope or F, at ).S000 A; note
at n = oo, T,/104 = 5, and at n = (10, 5, 3, and 2), T,/10 4 = (4.3, 3.5, 2.2, ultraviolet and infrared excesses or F, relative to B,(T,). Dotted
and 1.2) respectively, which shows that increasing atmospheric extent pro- curve: black-body curve at T = T(t = 2/3), 1he value characteristic
duces effects that simulate lower atmospheric temperatures. Indeed, a model or a planar atmosphere. Abscissa: 1/J., where). is in microns.
with T 1 = 50,000°K and n = 3 has a flux distribution nearly identical to From (61, 165), by permission.
250 Mode/ Atmospheres 7-6 Extended Atmospheres 251

infrared excess relative to that from a planar atmosphere of the same char- where S,. is assumed to have a general form
acteristic temperature. Relative to a Planck function at Tc, the temperature
that would be assigned observationally, the flux distribution shows both an
infrared and ultraviolet excess. These results emphasize the rather inhomo-
S,. = ix, f R(r; v', v)J,. dv' + fl. (7-192)
geneous nature of the radiation field from an extended atmosphere, arising
from the greater variation of the temperature through the envelope. We For a pure continuum problem, we would have only a coherent scattering
shall consider more realistic nongrey and non-L TE models after developing term and s. would be simpler. There are two essential difficulties with
a general method for solving the transfer equation. equations (7-190) and (7-191). (a) Direct elimination of H does not yield
a simple equation, but rather a complicated one involving both first and
s~cond derivatives. (b) The term in equation (7-191) in (x.r)- 1 tends to
SOLUTION OF THE TRANSFER EQUATION
drverge strongly at the surface (recall x. is opacity per unit volume and
IN SPHERICAL GEOMETRY
hence varies with the particle number density over se:veral orders of magni-
A variety of methods have been employed to solve the transfer equation tude). This term destabilizes the system. Both of these difficulties can be
in spherical geometry. Both Kosirev (360) and Chandrasekhar (148) used eliminated entirely by the introduction (32) of a sphericality factor q.,
variants of the Eddington approximation. which, however, is not accurate defined by
near the surface. Subsequent work (154; 679; 680) approximated the angular ln(r 2 q,) = J.: [(3/. - 1)/(rf,,)] dr' + In re 2 (7-193)
behavior of the radiation by an expansion in spherical harmonics, but only
to second order. As demonstrated convincingly by Chapman (163), these where re is a "core radius" corresponding to the deepest point in the atmo-
methods cannot be adequate because the radiation becomes very sharply sphere considered in the solution. It is obvious that q. is known if f. is known.
peaked in the outward direction (/,c -+ 1) at the surface [see (163, Fig. 1) or The factor q, allows equation (7-191) to be rewritten as
(323, Fig. 6)]. Methods that employ direct differencing of both the angular
and spatial derivatives in equation (7-165). [related to the Sn-method (127) (7-194)
of reactor physics] have been suggested (130; 255; 507), but these fail when
the discrete spherical shells become optically thick, and hence must be which, when substituted into equation (7-190), yields a combined moment
supplemented by special techniques to be useful for stellar atmospheres equation
work. (7-195)
The differential-equation technique to be described below, which is general,
stable, and efficient, uses a Feautrier solution along individual impact
parameters (tangent to discrete shells) to generate the angular information or, introducing a new variable dX,. = -q.x. dr = q, d-c.,
required to evaluate variable Eddington factors (323) from an estimated
source function. Then, with known Eddington factors, it employs a Feautrier 82(J,q,r 2 J,)/cX.2 = q,- 1r2(J,. - S,) (7-196)
scheme to solve a combined moment equation obtained by using the elegant
transformation introduced by Auer (32). Alternatively, a direct solution can Exercise 7-17: Veriry that equation (7-193) allows the reduction of equations
be obtained with a Rybicki-type method, if the scattering integral in the (7-190) and (7-191) to (7-195).
source function is frequency-independent (442); an equivalent integral-
equation method has also been developed (558). To obtain an upper boundary condition we define
Consider first the moment equations; regard all variables as functions of r
and v, and introduce the radial optical depth scale dt, = - x. dr. Then the
moment equations are h,. = f I(R,
0
1
Jl, v)µ dµ / f l(R, µ,
0
1
v) dJt (7-197)

a(r 2 H.)/a-c. = r 2(J,. - s.) (7-190) so that from (7-194)


and [a(J;,J,)/a-c.] - (3/, - l)J./(x,r) = H, (7-191) (7-198)
252 Model Atmospheres
253

while at the lower boundary we apply a planar diffusion approximation to


write

H,(r, ) - 1 (x, -11 8B,/8rI),. (7-199)


3
and fix the gradient by demanding that the integral of H ,(rc) over all fre-
quencies equal the correct integrated flux H, = L/(16x 2 r/). Then

2
[o(f,q,r J .)/oX.],.,. I 1=,.
= rc 2 Hc [ x. - 1(8B,/oT) fo"" x. - 1(oB./oT) dv
(7-200)
The diffusion approximation, and hence equations (7-199) and (7-200), will
be valid when the photon mean free path x. - 1 « eR where e is some small
number; this criterion can always be met by choosing r, sufficiently deep
L__
in the atmosphere. For other physical situations (e.g., in a nebula), alternative
inner boundary conditions can be posed (374).
With the introduction of a discrete radius mesh {r4 }, (d = 1, ... , D),
r- A Observer
Y (z- oo)

=
whereR = r 1 > r 2 > ... > rD r.,andafrequencymesh {vn}, (n = l, ... ,N), FIGURE 7-29
we may replace derivatives with finite differences [perhaps using splines Discrete (p, z) mesh used in solution of spherical transfer equation. The impact parameters
{p1} are chosen to be parallel to the central ray, and tangent to spherical shells chosen to
(374; 442) or Hermite formulae (34)] and the frequency integral (if any) in describe the depth-variation of physical properties of the envelope. The intersections of the
the source function with a quadrature formula. Equations ('7-196), (7-198), rays with the radial shells define a z-mesh along each ray.
and (7-200) are then of the standard tridiagonal form of equation (6-31) and
can be solved with the usual Feautrier elimination scheme; the computing
time scales as TM = c DN 3 , which is particularly economical if the source· µ41 = µ(r 4, p1) = (rl - p,2)¼/r4 = z4,lr4 • ijence if we first construct the
......
(JJ
function is purely thermal or has only a coherent scattering term (N = 1). • · solution along all rays {p1}, and choose a particular value r4 , then knowledge
In the calculation just described, the scattering integral in the source function of the variation of 1,(z41 , p1) (i = 1, ... , 14) is equivalent to knowledge of the 01
appears explicitly, hence the correct global thermalization properties of the µ-variation of I, (r4 , µ) on the mesh {µ 41 }, (i = 1, ... , / 4 ), which spans the I
solution are obtained. interval 1 ;l!: µ ;;: 0. Here / 4 = I + l - d. It is therefore clear that the ray-
To carry out the computations described above, we require knowledge by-ray solution described above will allow determination of the requisite
of the Eddington factors /,; these can be found if we know the angular Eddington factors. The astute geometrical trick employed here to synthesize
behavior of the radiation field at each depth. To obtain the required infor- the angular information from ray solutions is actually nontrivial and makes
mation we perform, at each frequency, a ray-by-ray solution along a grid direct use of the symmetry of the problem, which allows us to treat all points
of impact parameters {p1}, chosen to be tangent to each discrete radial on a given shell as equivalent; without strict spherical symmetry, the problem
shell, augmented by an additional set of C impact parameters, which are is much more complex.
chosen to intersect the core, and which include the central ray. The geometry Consider now the ray specified by p1• The transfer equation along the ray is
of the situation is shown in Figure 7-29. The impact parameters are labelled
with an index i, (i = 1, ... , J), where I = D + C; p 1 denotes the central (7-201)
ray; Pc the last ray inside the core (i.e., Pc < r,); .Pc+ 1 = rc; and p1 = R. where the + and - signs refer, respectively, to radiation flowing toward
Each ray p1 intersects all shells with r 4 ;;: p,. and these intersections define and away from the external observer, and we have written r as the space
a mesh of z-points {z41 }, (d = 1, ... , D 1). Here D1 = D + C + 1 - i for variable in 1/ and x with the understanding that r = r(z, p1) = (p,2 + z 2 )¼.
i > C, D1 = D for i ~ C, and z41 = (r/ - p,2)¼. It is seen from the figure
Defining the optical depth along the ray dt(z, p1, v) = - x(r, v) dz, setting
that the ray p1 intersects the radial shell r 4 at an angle whose cosine is S(r, v) = l'/(r, v)/x(r, v) (assumed to be known values), and introducing the
7-6 Extended Atmospheres 255
254 Model Atmospheres
tending the ditTerence equation for equation (7-204) beyond the central plane
mean-intensity-like and flux-like variables Zv;= 0, and using symmetry of udln about this plane.

u(z, p,. v) -·[+


= / (z, p1, v) + I - (z, p,. v)] (7-202) In the case that the source function contains a single scattering integral
2 involving J (or J in a line), rather than a frequency-dependent scattering
term involving partial redistribution, we may avoid the iteration procedure
and (7-203) between ray equations and moment equations, and develop a direct solution
by a Rybicki-type scheme (442). Along each ray, for each frequency, we have
a tridiagonal system of the form
we obtain the second-order system
(7-204) T an. U1 n -- uI J + w. (i = 1, ... ' /) (7-209)
n an (
n
= l ' ... ' N)
with an upper boundary condition
where J describes the depth-variation of J [cf.equation (6-46)] on the range
ou(z, Pit v)/ot(z, P1t v)I:_, = u(zm.., p,, v) (7-205) (d = 1, ... , Di), and ui. the variation of u(z, p1, v.) along the ray. This system
can be solved for u1• = c,.J + Din, and this solution can be substituted into
where Zmax = (R 2 - p/)½. The inner boundary condition depends upon the equation defining J [cf.equation (7-207)] to yield a final system for J; the
whether the ray intersects the core, r0 , = r,, or misses the core and intersects computing time for this solution scales as T O = cN D3 for D » C. Details
the central plane at z = 0. In the former case we apply the diffusion approxi- of this procedure, which is general, stable, and economical, can be found in
mation as was done to obtain equation (7-200); in the latter, symmetry the reference cited.
considerations show that v(O, p1, v) = 0, hence

ou(z, Pt, v)/ot(z, P1, v)l,-o =0 (7-206) EXTENDED MODELS FOR EARL Y·TYPE STARS

Equations (7-204) through (7-206), when written as difference equations, Nongrey spherical model atmospheres in LTE have been constructed
for the central stars of planetary nebulae (130; 131; 376) and for O and B
yield (with known S) a single tridiagonal system of the standard Feautrie'r
supergiants (136; 325; 376; 441; 442; 559; 516, 241), using a variety of
w
form, and .can be solved by the usual algorithm. The computing time CJ)
for N frequencies, C core-rays, and D radial depth-points scales as techniques. Non-L TE models for both classes of objects are given in (376;
441; 442; 516, 241). All of these models assume hydrostatic equilibrium, and
I
TR = cN[D · C + L D1] ~ c'ND 2 for D » C. Having calculated the com-
plete solution u41" = u(z4 , p1, v.) we construct the moments, as described the extension effects are produced by a near-cancellation of gravity by radia-
above, tion forces on the material. In actuality, there is strong evidence that atmo-
,. spheric extension is almost always associated with large-scale expansion, and
J dn L w~?>udln
= ,~1 (7-207) static models can be expected to yield, at best, only qualitative information;
we shall discuss dynamical models, which are more difficult to construct, in
t. §15-4.
and Kdn = IsLI wWudln (7-208) In an extended atmosphere, it is necessary to account for the variation of
the gravitational force with radius, and if .A is the mass of the star, the
and thus the Eddington factor fdn = Kdn/Jdn• Here the w's are appropriate equation of hydrostatic equilibrium becomes
quadrature weights, obtained analytically by integration of moments of a
piecewise polynomial representation of u(rd, µ) on the mesh {11i}, generated (7-210)
by the intersection of the rays {Pd with the radial shell rd. Using the new
Eddington factors, the moment equations are re-solved and the process is Introducing the Rosseland mean opacity XR, the integrated flux H =
iterated to convergence; experience shows the convergence to be very rapid. L/(16n 2 r 2 ), and the parameter

Exercise 7-18: (a) Write difference approximations for equations (7-204) through (7-211)
{7-206). (b) Derive a second-order accurate lower boundary condition by ex-
256 Model Atmospheres 7-6 Extended Atmospheres 257

we have [dp,fd(l/r)] = p[GA - (YLXRf4ncp)] (7-212) obtained using a complete linearization procedure, both assuming LTE and
As was true for planar atmospheres, radiation forces lead to an increase in taking into account departures from LTE. It should be stressed that, for such
the scale-height (and hence the extent) of the atmosphere. The parameter extreme values of r,_hydrostatic equilibrium is very unlikely to be possible,
r = 'l'XRL/(4ncGAp) measures the ratio of radiation to gravity forces. In and dynamical models are necessary; models such as these merely demon-
the limit of a pure electron-scattering opacity we find an upper bound on the strate the effects of a greatly increased scale-height for the envelope. Extension
critical ratio for (L/A) at which radiation forces just balance gravity forces, effects produce major changes in observable parameters such as colors, see
namely (L/A)., 11 = 3.8 x 104 (L0 /Ad. (cf. Exercise 7-1). We shall denote Figure 7-30, which show a strong reddening with increasing atmospheric
the value of r obtained from pure electron scattering as r •. For an extended size. The colors for LTE and non-LTE models differ markedly because in
envelope, the meaning of the stellar "radius" (and hence of the effective LTE a spurious emission edge occurs at the Balmer jump (caused by the
temperature) becomes ambiguous; typically one uses the value rt, at which
'R = i, as a characteristic radius. -0.40
The effects of radiation forces are quite important in the planetary nebula
stars, as estimates of L and A for these objects (279) lead immediately Po-
-0.38 ~2.7
to values of r on the range 0.8 < r. < 0.93. Further, it can be shown 50,00() 3.0 0.75 0.6
(442) that the absolute thickness of the atmosphere, /lr, is proportional -0.36 P o.&0.1· 40,000
g.,i = [(l - r)GA/R 2] - 1, which means that for a given g.rr, the relative %:
1.0
0
1.4 1.6 1.7
thickness of the atmosphere, llr/R, will be larger for stars of small radius, -0.34 p
such as the planetary nebula stars. Thus several of the models in (130; 131) --~I.S
have atmospheric thicknesses comparable to, or even larger than, the radius -0.32
of the "core" star (depths below-r = 10)! The effect of extensioo in these stars p~~
-0.30
greatly reduces the color temperature relative to the effective temperature;
at i., = 5000 A, T,/Terr for some of the models is as small as 0.3. The
-0.28
complete energy distribution shows the characteristic flattening described , (u - b) 60 ..II0
above for grey models [indeed in the visible, the nongrey energy distrjbution . -0.26
can be fitted quite accurately by grey models with n = 3 to 4 (61, 165)]. The
flatter energy distributions predicted by these models strongly resembles -0.24
those of WR stars (though the model parameters L, A, and Rare not really
appropriate for such stars). -0.22
For 0-type supergiants and Of-stars, it is clear, from the failure of planar
models to match the energy distribution [see (376)] and to produce the -0.20
observed emission lines, that the atmosphere must be extended. On the other
-0.18
hand, stellar evolution tracks at plausible masses all yield values of r • < 0.5,
which, taking into account the large radii of these stars, implies insignificant -0.16
atmospheric extension (136). But absorption in spectrum lines can greatly
increase the total force of the radiation on the material (102,404) and indeed -0.14
it is precisely this mechanism that is thought to drive the wfnds from 0-stars
(see §15-4). In several calculations (376; 441; 442; 516, 241), attempts were -0.14-0.13-0.12-0.11-0.I0-0.09-0.08-0.07-0.06-0.05-0.04-0.03-0.02
made to simulate the effects of enhanced radiation forces by adopting ad hoc (b - y)
values for the radiation-force multiplier')' in equation (7-212), chosen so as to FIGURE 7-30
raise the maximum value off to values approaching unity (the most extreme Strtlmgren-system (u - b) and (b - y) colors ror models with .I(= 60 .I( 0 Open symbols:
case considered had rm.. ::==: 0.995). Although such models become very LTE models; filled symbols: non-LTE models. Models with equal values or Terr are joined
on curves. Curves are labeled with parameters describing radiation-force multiplier y.
difficult to compute because of numerical instabilities, which directly reflect
Circles: y = 11 , labels give value of y 1 ; triangles: y = I + Yz exp( - ta), labels give value of
the physical instability of the atmosphere, a number of solutions were y2 • From (376), by permission.
258 Model Atmospheres

Schuster mechanism-see §10-2); detailed discussion of the overall properties


of the energy distributions can be found in the references cited above.

7-7 Semiempirical Solar Models


All of the discussion presented thus far in this book has concerned the
development of methods that allow the analysis of a stellar spectrum, and
the inference of physical properties of the atmosphere, by a model-fitting "'8
technique. Our information about stars, which we perceive as mere points X ~
.,..
N r--
of light, is quite limited in quantity, quality, and scope, primarily because
of the low intensity of the radiation received. But for one star, the sun, we "'8
have unparalleled opportunities to obtain information at very high spatial, .,,:: X
§.
temporal, and spectral resolution, over an enormous range of energies, from E ": 00
N §
.,..
0
u
X-ray, ultraviolet,. visible, infrared, and radio wavelengths. While stars are
unresolved, structures of the order of 1S0 km in size can be distinguished in
ii:
0
"'8
..J
the solar atmosphere, and a wide variety of small-scale features can be seen. X §
Further, measurements of velocity fields and magnetic fields can be made. - 00

~
The sun thus provides a unique testing ground for our theoretical methods, f,.. ..,
and, because so much information can be obtained by semiempirical analyses e
that rely only weakly upon theory, it offers guidance in seeing how to extend ~
the theory in those places where it is oversimplified. It is fair-to say that most
C ,::
~ ·;:;i
(though not all) major advances in stellar atmospheres theory were motivated .,!.
!! 6 .!2 0 :t
e ·= ~
'ob
by attempts to understand solar phenomena; time and again, the refined I- - e .......
confrontations possible between solar observations and the theory have i:t e::,
forced radical changes in our interpretive picture. And many enigmas remain1 -~ ~ w
so we may e.xpect much further development yet. eu ·e 0)
.c
.,,
c:,. e I
There is no hope of summarizing adequately, in a few pages, the vast ~ ::,
amount of information known about the solar atmosphere, and the reader
0
e
E
:2
~
e8. -~
should consult the many excellent references available, particularly (244; .c
u e
694; 11, Chaps. 9 and 10) for general information, (628; 19; 20) for detailed ------y··-------- ~
information about the chromosphere, and (94; 20) for material about the
0 )
' u
corona. The goal of the discussion below will be only to present a brief s ti 0
0 .c
summary of some of the primary structural properties of the solar atmo- .c
c...,,c:,.
sphere, in order to provide background and orientation for the development r-- .,.. ,..,
of more realistic pictures of stars in general. '°
f,..
In terms of basic morphology, the solar atmosphere can be divided into Oil
.S?
four major parts: (a) the photosphere, the opaque disk seen visually; (b) the
chromosphere, a region extending some 2S00 km above the limb, showing the
characteristic emission spectrum of hydrogen; (c) the corona, a tenuous, faint
outer envelope first seen at eclipses; and (d) the wind, a supersonically ex-
panding region that streams past the earth. The overall temperature structure
of the atmosphere is shown in Figure 7-31. In the photosphere the tempera-
ture decreases outward to a minimum of about 4200°K. At the relevant
260 Model Atmospheres

temperature (T ~ 5 x 10 3 °K) and gravity (g~ 3 x 104 ) the pressure scale-


height is about 120 km. This length subtends an angle of 0.15", which lies
well below the typical limit of resolution set by seeing effects in the earth's
atmosphere; as a result, the sun appears to have a crisply defined edge, which
=
occurs at height h 0 km on the scale of Figure 7-31. In the low and 0
middle chromosphere the temperature rises to about 8000°K and then to a
plateau of around 30,000°K. There is then a very thin transition region in
which the temperature suddenly rises to coronal values of about 1.5 x
I0 6 °K. In the corona the scale-height is about 50,000 km, which is a signifi-
t[~[ ~
.,.,
00

°'
cant fraction of a solar radius, so the corona extends to large distances from
the sun; indeed, even if the corona were static, it would envelope the earth
<[·
;; ~
~
--[
_i- <~ =vi
with a substantial density of particles. i5.
The temperature structure described above has been inferred mainly from 'E ...t

~Ie
analysis of the solar spectrum, which is strongly affected by the nature of
the temperature distribution. The features in each spectral region arise from
some characteristic range of height (see Figure 7-32) and, since it has become
l- .,,. :.( °';:;-[vi
:.(

C:
N <- -[
~u ~
·c
~ e
.. tr.
.2
possible to survey a wide range of wavelengths from space vehicles, our
t!
:z: a= :z:
<~[ ti
knowledge of the temperature structure has been greatly improved. In the a §[ -8 5
spectral range from 168S A~ .t ~ 3S0 µ the continuum originates in the
photosphere, and the lines have unit optical depth at heights that extend
from the photosphere (in the line wings) into the chromosphere (at line
center). Observed on the disk, lines with wavelengths .t ~ 1900 A are domi-
nantly in absorption, and a transition to emission lines occurs in the range
'[
l
u ~ ...j
]o..·-
0 ~
-s
.<;: ti
C:

e~
C:
:::> :.::

e . .1!]
1700 A ~ .t :E;; 1900 A. The continuum on the range 1525 A ~ .t ~ 1685 A,
:,
=- o.. B ......
c:<
-~ ~
:.(
I ~ '6
comes from the photosphere-chromosphere transition region, and for .t <
1525 A is dqminantly chromospheric. On the range 504 A :E;; .t ~ 912 A 8 I
-
. ·-.
-~-
:;i e
C: w
co
;~ e
the Lyman continuum is in emission; for .t ~ 504 A the emission is from the
continuum of He I. Chromospheric emission lines extend at least down to ti
~-
,.J
r, -~ ~
:,
=-
c:< u 1!
.c cc:
-C: ..
0
I
288 A, through the resonance series of He II. Coronal emission lines begin ~ 8 'I
~
·- 0,

to appear prominently at about 800 A, and extend down into the X-ray
C:
.. r;::-, ; ~ ,-----, - e
i.--
~ n;;:-7 ·u C
region. In the infrared, one sees most deeply into the atmosphere near 1.6 µ, ..Jr-[7
..J
,.J
~e .....
.c u
0 ·;;
the minimum in the H- opacity. At longer wavelengths, the temperature .c
·- .
C:
0 -
minimum region is encountered near 300 µ; and at radio wavelengths of ~ u
§ .2,
100 cm and beyond, the continuum emission is entirely coronal. (The corona
is completely transparent in the continuum at centimeter wavelengths and ~ -:; l1
shortward.) Off the disk, the chromosphere and corona can be seen in strong §.
~
§.
~
§.
~ ~ ~ ; ~
~ .§ c:!
u C: 0
!:i 8 "ilj
emission lines, by using special instrumentation, or in special circumstances . :. ei ·-
such as eclipses, from which an enorm9us wealth of data has been derived. Q'
e_..
~ 8."'
e i'.l E8.
A variety of analytical techniques have been used to make estimates of (... ~~:.§~
physical properties of the solar atmosphere from the data described above.
A very powerful tool is the analysis of limb-darkening data. As we saw in §3-3,
the Eddington-Barbier relation implies that the dominant contribution to
1,(0, µ) comes from S,('r, ::::: µ); hence by scanning from center to limb,
262 Model Atmospheres 7-7 Semiempirical Solar Models 263

O :;;;; µ :;;;; 1, we sample depths O :5 t, :5 1. Suppose we assume that the occupation numbers of low-lying states of Fe I (7.9 eV below the ionization
source function is the Planck function, and that we can write B.( t.}/I .(O, 1) = limit) of about 6 log n :=:,; ±0.25, or about ±60%; thus even assuming that
b.(t.). Then every other step in the analysis of the spectrum were perfectly accurate,

<I>,(µ) = [/,(0, µ)/1,(0, 1)] = f


0
"' b.{t,) exp(-t./µ) dt,/µ (7-213)
significant errors are introduced into estimates of element abundances from
uncertainties in the model alone.
The range of depths available to analysis using limb-darkening data may
If we now assume that b, can be represented by some analytical formula, be extended by using different wavelengths. One then needs to map the
<I>, can be computed in terms of the expansion coefficients of the formula. By curves of T(t.) from one frequency to another, which can be done if we
fitting this calculated function to the observations of <I>., one finds the coef- know the frequency variation of the absorption coefficient-i.e., x(v)/x(v.,d)-
ficients and hence b,. For example, if we take Alternatively, we can demand that the different T(t.) curves correspond to
a unique T(h) variation, and turn the procedure around to infer the frequency
(7-214) profile of the absorption coefficient. Studies of this kind (e.g., 517) have shown
that the empirically-determined absorption coefficient is consistent with the
then we find hypothesis that H - is the primary opacity source (though additional sources,
(7-215)
mainly lines, are needed in the ultraviolet).
Another method that can be used to determine the temperature structure
Thus a fit of equation (7-215) to the data yields the a/s, hence b. and, ulti- is the analysis of disk-center absolute intensities as a function of frequency.
mately, if we know /,(0, 1) we obtain B,(t,), and hence T(-r.). A very large With the advent of precise absolute intensity measurements in the ultraviolet
body oflimb-darkening data exists, extending from ultraviolet through radio from space vehicles, this method has surpassed limb-darkening analyses in
wavelengths. importance. An extensive compilation and discussion of existing data is
The procedure outlined above encounters many practical gifficulties. Thus given in (646). The basic procedure employed is to construct models by (1)
even large-amplitude fluctuations in the physical properties do pot affect assuming a temperature distribution, (2) integrating the equation of hydro-
the observations if they occur in regions whose line-of-sight optical thickness static equilibrium, and (3) solving the coupled transfer and statistical equi-
is « 1; we thus remain in ignorance of any such inhomogeneities. Similar librium equations for the relevant atoms (e.g., H, He, C, Si) to calculate
remarks apply to horizontal structures that lie below the limit of resolution absolute emergent intensities. The assumed T(h) relation is then adjusted to ~
set by the observing procedure. Moreover, we clearly obtain but little infor~ yield an optimum fit to all available data. A sequence of successively more 0
mation froin depths with t, > 1, while a limit is also set at small depths by refined models have been constructed in this way, starting with the Utrecht I
seeing effects at the limb; a typical resolution of 1" implies a blur of about Reference Photosphere (283, 239), which was soon displaced by the Bilderberg
0.001 solar radius, which corresponds toµ ~ 0.05. Thus at any given wave- Continuum Atmosphere (BCA) (248). Introduction of ultraviolet space obser-
length we are confined to 0.05 :5 t, :5 1. Further, with data oflimited quality vations led to the highly-successful Harvard-Smithsonian Reference Atmo-
the number of expansion coefficients that yield real information in equation sphere (HSRA) (summarized in Table 7-1) (249); subsequent work (645;646)
(7-214) or (7-215) is restricted. For example, for data with an accuracy of has produced a new model ("Model M") that fits a very large body of data
± 1%, at most three coefficients can be obtained (98); attempts to use more extremely well. Improvements in these models are constantly being made,
coefficients yield results that fit the observed values of <I>.(µ) but show wild and important advances will result when eclipse data are taken into account
oscillations in b.(t,.). Numerical methods (such as the Pmny al{Jorithm) exist and spectrum-line synthesis is pushed to its utmost. It should be recognized
that automatically limit the coefficients to only those justified by the accuracy that there are still many fundamental difficulties to be faced by these semi-
of the data (669). One may choose a variety of fitting functions in equation empirical models. One of the most worrisome is that observation reveals
(7-214), each of which fits the observed <I>.(µ) equally well; in general these that the layers under analysis contain a great deal of small-scale structure
yield different values of b.{t,) and hence of T(t,). Among three-term solutions, and are far from homogeneous. Some difficult questions remain essentially
discrepancies in the temperature of the order of ±200°K are found; these unanswered: To what extend does the value assigned to a physical variable
can be important in certain contexts. For example, in the solar atmosphere in any layer of a one-dimensional chromospheric model represent the average
the dominant ionization state of iron atoms is Fe+. Using the Saha equation property of that layer? How large is the fluctuation of any physical property
it is easy to show that an error of ±200°K introduces an uncertainty in the about its mean value? Are fluctuations in various properties correlated? It
264 7-7 Semiempirical Solar Models 26S

TABLE 7-1 is imperative that we face these queries, for one-dimensional models are the
Harvard-Sml1hsonlan Reference A1mosphere• best we can hope to obtain for stellar chromospheres, and the validity of
such models will be supported or undermined, depending on the answers to
h(km) T(°K) p,(dyncs cm- 2) n,/[n, + n(H)]
the questions posed above for the solar chromosphere.
1.00 - 8 1860 8930 1.52 - 1 4.82 - 2 5.11 - 1 In the corona, which is optically thin, the density structure can be inferred
2.00 - 8 1850 8750 1.54 - 1 4.84 - 2 5.05 - I from the brightness of the light scattered by free electrons, after allowing for
3.16 - 8 1840 8630 1.56 - 1 4.92 - 2 5.06 - I the effects of integration along the line of sight [see, e.g., (94)]. Coronal
6.31 - 8 1830 8450 1.62 - 1 5.10 - 2 5.05 - 1
density distributions have been tabulated for the equatorial and polar
LOO - 7 1820 8320 1.69 - I 5.34 - 2 5.09 - I directions, for varying conditions during the solar cycle (477); a typical
2.00 - 7 1790 8090 1.88 - 1 5.88 - 2 5.02 - 1
3.16 - 7 1760 7910 2.10 - 1 6.44 - 2 4.86 - I electron density near the base of the corona is of the order of 10 8 cm - 3 •
6.31 - 7 1690 7630 2.76 - I 7.50 - 2 4.09 - 1 That the temperature of the corona must be very high was first recognized
1.00 - 6 1620 7360 3.76 - I 7.85 - 2 2.89 - I by Grotrian, who suggested that two broad, faint absorption features seen
2.00 - 6 1430 6720 1.00 + 0 6.85 - 2 7.99 - 2 in the light scattered by the corona were the very strong Ca II H- and K-
3.16 - 6 1230 6180 3.79 + 0 6.00 - 2 1.72 - 2 lines of the solar spectrum, "washed out" by electron scattering at very high
6.31 - 6 947 5590 2.92 + I 6.80 - 2 2.26 - 3
temperatures. This idea was further supported by the identification, by Elden,
1.00 - 5 840 5300 6.79 + I 6.77 - 2 9.13 - 4
of coronal emission lines with transitions in highly ionized atoms-e.g.,
2.00 - 5 720 4910 1.87 +2 4.43 - 2 1.41 - 4
3.16 - 5 654 4660 3.37 +2 3.96 - 2 2.46 - 5 Fe X (l6374) and [Fe XIV] (l5303). Modern determinations of the temper-
6.31 - 5 588 4280 6.34 +2 5.11 - 2 7.65 - 7 ature in the corona are in agreement with one another and give a typical
1.00 - 4 557 4170 8.68 + 2 6.12 - 2 2.26 - 7 temperature of about 1.5 x 106 °K, based on (a) thermal line-widths,
2.00 - 4 SIS 4205 1.34 +3 9.09 - 2 2.13 - 7 (b) ionization equilibria (including dielectronic recombination!), and
3.16 - 4 487 4250 1.77 +3 1.20 - I 2.47 - 7 (c) radio-wavelength noise temperatures. X-ray observations reveal intensely
6.31 - 4 447 4330 2.65 + 3 1.81 - I 3.40 - 7
hot regions (several millions of degrees) near active regions. For more de-
1.00 - 3 420 4380 3.46 +3 2.37 - I 4.06 - 7
tailed discussion see (20) and (94).
2.00 - 3 379 4460 5.12 +3 3.53 - I 5.43 - 7
The temperature distributions shown in Figures 7-31 and 7-32 obviously
3.16 - 3 352 4525 6.65 + 3 4.65 - I 7.11 - 7
6.31 - 3 311 4600 9.81 +3 6.86 - I 8.87 - 7 bear little resemblance to the classical predictions of radiative-convective
+4 8.95 - I 1.09 - 6 equilibrium models, and indicate a major breakdown in this approach in
1.00 - 2 283 4660 1.27
1.33 + 0
.......
2.00 - 2 · 241 4750 1.87 +4 1.46 - 6 the outer layers of the atmosphere. The classical models are adequate to
3.16 - 2 212 4840 2.41 +4 1.78 + 0 2.13 - 6
predict the structure of the photospheric layers where the visual continuum
6.31 - 2 168 5010 3.54 +4 2.81 + 0 4.42 - 6
and line wings (and very weak lines) are formed; but once the temperature
1.00 - I 138 5160 4.56 +4 3.95 +0 8.49 - 6 rises outward we must introduce new phenom.ena and an appropriate
2.00 - 1 92.6 5430 6.61 +4 7.04 + 0 2.47 - 5
3.16 - 1 63.1 5650 8.31 +4 1.13 + I 5.29 - 5 heating mechanism (clearly the Cayrel mechanism is not the cause of the
6.31 - I 22.6 6035 1.12 + 5 2.65 + I 1.58 - 4 rise, for Te becomes » T.rr !). It was suggested by Biermann (89) and Sch-
LOO 0.0 6390 1.31 + 5 5.64 + 1 3.65 - 4 warzschild (564) that acoustic waves would be generated in the solar con-
2.00 -25.3 7140 1.54 + 5 2.35 + 2 1.55 - 3 vection zone, propagate outward, steepen into shocks, and deposit energy
3.16 -37.1 7750 1.65 + 5 6.26 +2 4.04 - 3 in the material, heating it to high temperatures. A specific mechanism for
6.31 -51.4 8520 1.78 + 5 1.81 +3 I.II - 2
10.00 -60.8 8880 1.86 +5 2.61 + 3 1.67 - 2 the production of soundwaves was suggested by Lighthill (394). Subsequent
20.00 -76.7 9390 2.00 +6 5.00 + 3 2.79 - 2 work has also called attention to the role of magnetohydrodynamic waves,
for the outer solar atmosphere has a significant magnetic field. In addition,
• The notation a· ua - b denotes a• aa x 10·•. the discovery (393) that large regions on the solar surface oscillate with a
Adapted from (249), by permission.
period of about 300' provided another source of nonradiative energy that
can be tapped to produce heating. The problems of wave generation, pro-
pagation, and dissipation are complex and difficult, and a tremendous
266 Model Atmospheres 7-7 Semiempirical Solar Models 267

amount of work has been done on the subject; an excellent review of the hope to infer, from the behavior of the ensemble, the time-evolution of the
field has been given by Stein and Leibacher (602) [see also (20)]. No fully phenomena for an individual such as the sun, on otherwise inacessible time-
consistent picture of the heating has yet emerged, but current work indicates scales (billions of years). Work at the interface between solar and stellar
that the 5-minute oscillations can indeed heat the upper chromosphere and atmospheres is both active and richly rewarding [see, e.g., (344)], and will
corona, while shorter periods are required to heat the lower chromosphere. unquestionably be synergistic in the development of our concepts of stellar
Estimates of the actual amount of mechanical energy input can be made by atmospheres.
comparing the semiempirical temperature distributions with radiative-
equilibrium models [ e.g., (17) or (380)] and calculating the energy required to
produce the heating. Difficulties here are that at the temperature minimum (I)
T in the empirical models lies below T in the radiative models, and (2) small
errors in T (say ±100°K) imply unacceptable errors in the energy input
estimates. Much work remains to be done to establish both the empirical
and theoretical temperatures to the required accuracy.
The presence of temperature plateaus followed by large jumps in temper-
ature can be understood in terms of thermal stability of the gas. In essence,
the gas is heated by mechanical input and cooled by radiative losses, which
establish a balance [see (628) and (20)]. The losses are largest just as some
atom is becoming nearly completely ionized, and this tends to provide a
thermostatic action that holds the gas near a definite temperature. Thus,
hydrogen provides the dominant cooling in the low chromosphere and keeps
the gas at about 7000°K to 8000°K; after hydrogen becom.es too strongly
ionized to be effective, losses from He I and He II dominate, and ,the tem-
perature jumps to about 20,000°K to 30,000°K; and, finally, after helium is
strongly ionized, losses from ions of C, N, 0, Ne, Mg, and Si dominate,.
yielding strong cooling rates at temperatures above 10 5 °K. In regions of ~
I\)
steep gradients, conductive energy transfer also becomes extremely important
and the final 'temperature profile is a result of all these mechanisms operating I
together.
Given that the sun is a typical G-dwarf, it is clear that we must conclude
that the chromosphere-corona-wind phenomenon must be a basic property of
stars in general. There is ample evidence that this is true. Thus, most stars
in which convection zones exist show chromospheric emission features in
the Ca II Hand Klines (cf.§ 12-2), with many stars that are younger than the
sun showing very active chromospheres. An extreme case is provided by the
T-Tauri stars where most of the prominent spectral features arise in a
"super-chromosphere" and very dense wind. Mass loss in stellar winds,
particularly in the early-type supergiants and WR stars (estimated rates up
to 10- 6 -10- 5 .,11 0 /year) is well established. These winds make the solar
wind (mass-loss rate ~ 10- 14 .,/I 0 /year) seem puny by comparison.
The solar atmosphere provides a kind of Rosetta stone that helps us first
understand a rich "literature" containing many dramatic stellar phenomena;
these phenomena, in turn, extend our knowledge over a broad range of
physical conditions. Further, by studying large numbers of stars, we can
8-1 Observational Quantities 269

Spectral lines are much more opaque in the core than in the wing, and
hence can provide a sampling of a wide range of atmospheric depths, from
8 very high layers (seen in the core) to the deepest points observable (the
depth of continuum formation). Further, lines are narrow in frequency-width,
and hence are sensitive to the effects of velocity fields; they thus provide
the means by which motions of the material in stellar atmospheres can be
studied.
The Line Spectrum: Moreover, it is clear that the strength of a given line must contain infor-
mation about the number of atoms absorbing photons along the line of
An Overview sight, and hence about the abundance of that chemical species in the atmo-
sphere. Thus, by a suitable interpretation, the line spectrum offers the
opportunity to perform a quantitative chemical analysis of the material
of which stars are composed. This information, in turn, provides valuable
clues when we attempt to construct a coherent picture of the structure and
evolution of stars, the Galaxy, and the Universe as a whole.
It is, therefore, of prime importance to develop a theoretical framework
within which line profiles can be predicted and the desired physical infor-
mation can be inferred. A great deal of effort has been devoted by many
astronomers to this end, and considerable progress has been made. The
second part of this book will describe the theoretical techniques that now
exist to treat the problem of line-formation. In this chapter a few of the
basic aspects of the problem, and a summary of the kinds of information
required to solve it will be pointed out to orient our later work.

Superimposed upon the continuum of a star, we observe discrete spectrum 8-1 Observational Quantities
lines, either in absorption or emission. These lines arise from transitions
between bound states of atoms and ions in the star's atmosphere. An ex- A line in a stellar spectrum is most completely characterized by its profile,
tremely wide variety of lines is found, from a wide range of atomic and which is the observed distribution of energy as a function of frequency. For
ionic states, leading to very different-looking spectra for different classes of all stars except the sun, we can observe only the flux integrated over the
stars. A panoramic view of this variety is best obtained by inspection of entire disk of the star. We measure F, (the flux in the line) relative to F,
actual stellar spectra, particularly as shown in (5) and (465) [see also (261, (the flux in the continuum) and describe the profile in terms of its absorption
Chap. 14; 330, Chap. 1)]. One finds that lines in stellar spectra show enormous depth
ranges of strength and striking variations in profile. A close examination A,. = 1 - (F,/F,) (8-1)
shows that the spectra can be arranged into a two-dimensional scheme or residual flux
reflecting primarily the effective temperature and luminosity of the star. It R, = (F,/F,) = 1 - A, (8-2)
would take us too far afield to describe here the details of this procedure-
developed to a high state of refinement by Morgan and his collaborators-or In the case of the sun, the frequency distribution of the radiation can be
the full implications of the results; therefore the references cited above should observed at each point on the disk. We can then describe the profile in
be studied carefully. Suffice it to say, the spectrum lines contain a wealth terms of / ,(0, µ), the emergent specific intensity, in units of the nearby
of information concerning the run of physical variables in the star and continuum intensity 1,(0, µ), and write
therefore provide important diagnostic tools for inferring the state of the
atmosphere. a,.(µ) =I - [I ,(0, µ)/ I,(0, µ)] (8-3)
270 The Line Spectrum: An Overview 8-2 The Physical Ingredients of Line-Formation 271

or r ,(µ) = I ,(0, µ)/1,(0, µ) = 1 - a,(µ) (8-4) 8-2 The Physical Ingredients of Line-Formation
Information about the center-to-limb variation of a profile is extremely As in the case of the continuum, the calculation of the flux in a line requires
valuable because it provides (via the Eddington-Barbier relation) an addi- the solution of a transfer problem, for the observed radiation originates
tional depth-resolution otherwise unavailable, and places important con- from a wide range of depths, over which the physical properties of the material
straints on the theory. Such information is, of course, available only for the may vary more or less strongly. Let us inquire here what information is
solar spectrum, and is one of the reasons why the solar spectrum provides an needed to formulate and solve the transfer equation in a spectrum line.
ideal testing-ground for proposed theories of line formation. Consider a frequency v, and suppose that we know the continuous absorp-
Often, because of the low light levels involved, it is not possible to measure tion and scattering coefficients, K, and u., and the line absorption coefficient
a stellar spectrum with sufficient resolution to determine a line-profile in x,</>. as a function of depth. We could then construct the line optical depth
detail, and one must then substitute the integrated line strength-the scale
equivalent width-in place of this more detailed information. For stars, t, = s:m.. (K, + <J, + x,</>.) dz' (8-9)
where we measure the.flux from the disk, the equivalent width is defined as
and continuum optical depth scale
W. = Saco A. dv (8-5)
t, = J:""• (K, + u.) dz' (8-10)
in frequency units (hz), or, more usually, as Then if, in addition, we knew the run of the source function S., we could
immediately calculate
(8-6)
(8-11)
in wavelength units (A or mA). For the sun, where angular information is
available, we can define, in addition, Fe = 2 Saco S.{t.)E2(t.) dtc (8-12)
and hence A. from equation (8-1), and K'J_ from equation (8-6) if desired.
W.(µ) = Saco a,(µ) dv (8-7) Of course, in practice, it is the source function that must be determined. Only
in the trivial case of LTE is s.
known beforehand (namely S, = B,). As
or K'J_{µ) = Saco a,(µ) d). (8-8) was described in Chapters 2 and 7, in the general (non-LTE) situation, the
source function and optical depth depend explicitly upon the occupation
The equivalent width is clearly the width of a perfectly black line with the numbers of the particular levels involved, but these, in turn, depend upon
same area under the continuum level as the line under study (hence the the radiation field, and hence, ultimately, upon the source function. Thus
name); obviously W gives a direct measure of the total energy in the con- what is required, as was the case for the continuum, is a simultaneous, self-
tinuum removed by the line (assuming that it is in absorption). consistent solution of the coupled transfer and statistical equilibrium
Ideally, one attempts to obtain profiles rather than equivalent widths, for equations. ·
they contain far more information. In particular, it is apparent that there Before we attempt to do this, however, some insight into the kinds of
is an infinite number of radically different profiles (each with distinct impli- information that will be required to attack the problem can be gained from
cations for the structure of the atmosphere) that will produce a given equiv- the following phenomenological arguments. Consider the propagation of
alent width. An interpretation based on an equivalent width alone can photons in a line with overlapping continuous absorption and scattering.
be misleading (the same remark holds even for profiles!). Nevertheless, Some photons will interact with the continuum, others with the line opacity.
approaches exist that use equivalent width information from many lines Some of the photons absorbed in the line will be scattered, and in general
simultaneously (the curve of growth), and these can yield important and will suffer redistribution in frequency and angle according to the redistri-
reasonably unambiguous results. The actual measurement of the data re- bution function R(v', n'; v, n). Others may be destroyed by collisional de-
quires refined instrumental techniques; we shall not discuss these techniques excitations or transitions to other levels. Photons may be introduced into
here as they lie beyond the scope of this book, but excellent discussions exist the line by collisional excitations or by transitions into the upper level fr?m
elsewhere [see, e.g., (300, Chaps. 2, 4, and 13)] other levels, with subsequent cascade to the lower level. A transfer equation
272 The Line Spectrum: An Overview

accounting for these processes will be of the form


;t(al./oz) = -(K,
+ YX,
+ <J, + x,cf,.)I. +
p (dw/41t) f
0
""
K,B. + a,J.
dv'l(v', n')R(v', n'; v, n) + ex,cf,,B, (8-13)
9
Here the coupling coefficients y and l describe, respectively, the fraction of
the photons scattered, and those emitted by other processes at a characteristic
temperature TR (not equal, in general, to the local electron temperature). The Line Absorption Pro.file
At this point we can see four important ingredients needed to compute
line profiles theoretically.
(a) We must be able to calculate the absorption profile cf, •. This will be
treated in Chapter 9.
(b) We must be able to specify the coupling between photons and the
material via the parameters y and l. In general, the expressions for these
quantities may be very complicated and may contain both radiative and
collisional rates from the levels giving rise to the line, as well as terms coupling
to other levels. The specification of y and l follows from the statistical equi-
librium equations for the atom as a whole. We shall examine this aspect of
the problem in Chapters 11 and 12.
(c) We must be able to calculate the redistribution function R(v', n'; v, n),
and to determine the effects of the details of photon scattering upon the
line profile. We shall consider these problems in Chapter 13.'
(d) We must be able to solve the resulting transfer problem. Here we will
simply refer back to the difference-equation methods discussed in Chapter 6.
In all four of the areas listed above, great strides forward have been made
~
recently. The greatest improvement in our understanding has come In The profiles of lines in stellar spectra contain information about both the 0,
regards to points (b), (c), and (d). In early treatments of line-formation, the physical conditions and the abundances of chemical elements in the stellar
question of redistribution was often evaded, and line-scattering was treated atmosphere. Therefore they provide extremely valuable diagnostic tools and I
as coherent. We know today that this is a poor approximation, and that, must be exploited as fully as possible. To carry out an analysis of observed
in fact, a much better approximation is the opposite extreme assumption line profiles one needs to know how the distribution of opacity with frequency
of complete redistribution over the line. In the specification of the parameters in the line-the line absorption profile-depends upon local conditions of
~ and e, the classical approach was sketchy, and led to some serious mis- density, temperature, etc. For an isolated atom with levels having essentially
conceptions. More modern treatments have brought to light the importance infinite lifetimes, the spectral lines would be almost perfectly sharp; but in
of a clear understanding of these coefficients. In the area of actually solving reality there are several different mechanisms that produce an indefiniteness
the transfer equation, important advances have been made possible by in the energy levels of real atoms in a plasma, and thereby lead to line
application of high-speed, large-capacity computers, using the recently- broadening.
developed, powerful, numerical techniques. The first line-broadening mechanism to be considered below is natural
Finally, it, it should be noted that in equation (8-13) it has tacitly been (or radiation) damping, which refers to the line width produced by the finite
assumed that the atmosphere is static. The important effects of velocity lifetime of the atomic levels set by their decay via the radiation process itself.
fields upon line-formation are discussed in Chapter 14 where several different Natural damping occurs even for a solitary, isolated atom. If the atom is
techniques for solving the transfer equation in moving media are described. imbedded in a plasma, then there will be an additional pressure broadening
The role of lines in establishing the dynamical state of the atmosphere is of the line caused by perturbations of the radiated wavetrain through
discussed in Chapter 15 in the sections on radiatively-driven stellar winds. collisions with other atoms, or charged particles, in the gas.
274 The Line Absorption Profile 9-/ The Natural Damping Profile 275

Classically, pressure broadening is described in terms of two limiting transform F(w) is defined to be
approximate theories. The first of these is known as impact theory. Here the
radiating atom is considered to be an oscillator ·that suffers a collision that F(w) = s~C() f(t)e- 1"'' dt (9-1)
occurs essentially instantaneously, and that interrupts the radiation wave- which satisfies the fundamental reciprocity relation
train with a sudden phase shift, or by inducing a transition. These collisions
thus cause the radiator to "start" and "stop" in intervals of finite duration, f(t) = (21t)- 1 s~C() F(w)e 1"'' dw (9-2)
leading thereby to a frequency spread in the radiated wavetrain, and to a
The quantity
shift of the line away from its unperturbed frequency. The alternative E(w) = (21t)- 1 F*(w)F(w) (9-3)
approach is the statistical theory in which we consider the atom to be
radiating in a field produced by an ensemble of perturbers. This field will is called the energy spectrum of the oscillator. This designation derives from
fluctuate statistically about some mean value as a result of motions of the the fact that
perturbers. At a given value of the field, the energy levels of the radiating
atom are shifted slightly, and, correspondingly, the frequency of the line is s~C() E(w) dw = (21t)- s~C() F*(w)F(w) dw = s~C() f*(t)f(t) dt
1
(9-4)
altered. The intensity of the radiation at any specified frequency shift is
taken to be proportional to the probability with which a perturbation of the which may be verified by direct calculation, using equation (9-1). If f(t) were
appropriate field strength occurs. the voltage across a one-ohm resistor, then f *(t)f(t) would be the instan-
The basic limits on the classical theories of pressure broadening are set taneous power delivered to the resistor, and the integral over all time gives
by their inability to account for the actual structure of the radiating atom, the total energy. Thus, E(w) is a direct measure of the energy in the wavetrain
or for transitions produced by the collisions with the perturbing particles. at frequency w.
Both of these defects are overcome in the quantum theory of pressure In many cases, the energy spectrum itself is not used, but rather the energy
broadening, which yields results in good agreement with experimental delivered per unit time-i.e., the power spectrum l(w), which is defined as
determinations.
Finally, we must account for the fact that in a stellar atmosphere we l(w) = lim (27tT)-
T ➔ co
1\r;,l f(t)e- 1"' 1 dt\
2
(9-5)
observe an ensemble of atoms moving with a velocity distribution along the
line of sight. The profile for each atom is Doppler shifted according to its However, for oscillations of finite duration (or with, say, an exponential
line-of-sight velocity, and the profile seen from the entire ensemble is a decay), the power spectrum will be zero because on the average over an ~
superposition of these shifted atomic profiles, calculated by a convolution infinite time interval, the finite total energy emitted yields zero power. In O>
with the velocity distribution. these cases, which are of practical interest, we use the energy spectrum itself, I
The theory of line broadening has progressed enormously in the past assuming that we observe an ensemble of oscillators created at a constant rate
decade, and reliable calculations now exist for many profiles of fundamental with random phases; a finite power then results, with a frequency distribution
astrophysical interest. The quantum theory has become very refined, but proportional to the energy spectrum of an individual oscillator.
also quite complicated. As excellent treatises exist on the general subject In certain situations it is not possible to calculate the power spectrum
of line broadening [ see particularly (264; 268; 629)], only a brief summary directly, using equation (9-5). It is then valuable to make use of the
of the most important results will be presented in this chapter; for further autocorrelation function
details one should consult the books just cited or the research literature, for
which references are compiled in (228; 229 ;232 ). <l>(s) = Thm
.
... co
T _ I fT/2 f*(t)f(t + s) dt
-T/2
(9-6)

from which the power spectrum is obtained by the relation


9-1 The Natural Damping Profile
I(w) = (21t)- 1 s~C() <I>(s)e- 1"'' ds (9-7)
' AND
ENERGY SPECTRA, POWER SPECTR_A,
THE AUTOCORRELATION FUNCTIO~ as may be verified by direct calculation, using a limiting procedure in the
We first derive some basic relations that will be needed in later dis- integration over s. The autocorrelation function provides a powerful tool
cussions. Consider a time-varying oscillation of amplitude f(t). The Fourier for calculating power spectra of a radiating atom perturbed by collisions.
276 The Line Absorption Profile 9-/ The Natural Damping Profile 277

THE DAMPED CLASSICAL OSCILLATOR QUANTUM MECHANICAL CALCULATION

The simplest picture one can construct of the process of emission in a A quantum mechanical analogue of the damped oscillator is obtained
line is to consider the atom to be a classical oscillator. In §4-2 it was shown by assuming the radiation arises in transitions by an atom from an excited
that the profile from a driven damped oscillator is Lorentzian, by a calculation state of finite lifetime to the ground state. Following Wigner and Weisskopf,
of the power emitted by the oscillator. We shall apply the techniques outlined we write the probability of finding an atom in the excited state j as
above to a decaying oscillator to show that the same profile is obtained. From
classical electromagnetic theory the equation of motion [ cf. equation (4-27)] (9-17)
for a radiating oscillator is where r = A 1i, the spontaneous emission rate. Then the time development
3l = -w0 2
x - yx (9-8) of the wave function of the state is

where y is the classical damping constant (9-18)


')' = (2e 2 2
w 0 /3mc 3
) (9-9) Consistent with the uncertainty principle, we consider that the decaying
state j (with characteristic lifetime ~t1) no longer has a perfectly defined
The radiation damping term is numerically quite small, and may be estimated
energy E1, but is, rather, a superposition of states with energies spread
using the undamped solution x = x 0 exp(iw 0 t), which yields
~
about EJ (with a characteristic width ~E1 li/M1). From the fundamental
3l = -(w/ + iyw0 )x (9-10) reciprocity relations of quantum mechanics, the amplitude of the energy
distribution is given by the Fourier transform of the time dependence of the
Neglecting terms in y2 , the solution of equation (9-10) is wave function, and the probability distribution of energy states is given by
(9-11) the square of the amplitude. Thus calculating the Fourier transform of
equation (9-18), it is clear that the result must be of the same form as that
which is an exponentially damped oscillation. Calculating the Fourier derived from equation (9-11), and by analogous arguments we obtain finally
transform, assuming the oscillation starts abruptly at t = 0, we find
/(w) = (r/21t) [(w - wo)2 + (½1}2]- 1 (9-19) ......
F(w) = x 0 f0
(1) e- 11"'-"'0 J1e- 1112 dt = x 0 /[i(w - w 0 ) + h] (9-12)
r
From equation (9-17) we see that is to be interpreted as the reciprocal
of the mean lifetime of the upper state. If several transitions out of the upper
The energy spectrum of the oscillator is then
state U are possible, then
E(w) = (xo2/21t) [(w - wo)2 + (h)2]-1 (9-13)
ru = L Au, (9-20)
l<U
The power spectrum of an ensemble of such oscillators created continuously
with random phases is proportional to E(w), hence the profile, normalized is the appropriate width of the state. Suppose the line under consideration
such that arises from a transition between two excited states, so that the lower state
f_:°CI) /(w) dw = 1 (9-14) L also has a width r L given by an equation analogous to (9-20). The line
profile must, in general, reflect the width of both states. We assume that the
is /(w) = (y/21t) [(w - wo)2 + (h)2] - i (9-15) distribution of substates of each level around its nominal energy is given
by a Lorentz profile with the appropriate r. Let {J = 1/2 for either level,
The damped classical oscillator thus yields a Lorentz profile with a full and let x = (E - E0 )/fi be the frequency displacement of a particular substate
half-intensity width y. In wavelength units this width is from the nominal energy E0 • We assume that the probability of ending in a
~l, = (21tcy/w 2) = (41te 2/3mc 2) = 1.2 x 10- 4 A (9-16) particular substate x' of the lower level is independent of the substate x
of the upper level from which the transition starts; then the joint probability
This width is much smaller than those observed in laboratory or stellar of starting in substate x and ending in substate x' is
spectra. We must therefore develop a more general picture of the radiation
process. p(X, X
1
) = ({JL{JU/1t 2)[(x 2 + fJu 2)(X' 2 + {JL 2 )]- I (9-21)
278 The line Absorption Profile 279

If we restrict attention to transitions producing radiation of a specific fre- 9-2 Effects of Doppler Broadening:
quency w, then x and x' must be related by w 0 + x - x' = w, or writing The Voigt Function
x0 = w - w 0 ,
x' = x - x 0 (9-22) When one observes a line in a stellar atmosphere (or a laboratory plasma),
one sees the combined effects of absorption by all atoms in the ensemble.
The total intensity at w is obtained by summing over all upper substates x, Each atom will have a velocity along the line of sight, measured in the
with x' fixed by equation (9-22), that is observer's frame, and the intrinsic profile of that atom will be Doppler-shifted

l(w) = s~"' p(x, X - Xo) dx


a corresponding amount in frequency. If the damping process producing
the intrinsic profile of each atom is uncorrelated with its velocity, then the
shifted profiles may be superimposed to yield the- total absorption cross-
OuOL f"' dx
(9-23) section by a simple folding procedure.
=7 -oo (x2 + Oul)[(x - Xo)2 + lJL2] Assuming the plasma is characterized by a kinetic temperature T, the
velocity distribution is Maxwellian, so that the probability of finding an
This integral may be evaluated by contour integration using the residue
theorem, taking into account the poles at z = ± iou and z = x 0 ± ilJL.
atom with a line-of-sight velocity e on the range (e, + de) is e
Performing the integration one again finds J(w) is given by equation (9-19) W(e)de = (nte 0 )- 1
exp(-e ;e 0 2 )de
2
(9-27)
but now with the damping width given by
e
where 0 = (2kT/m)t = 12.85 (T/10 A)½ km s- , where A is the atom's
4 1

(9-24) atomic weight. Then, if we observe at frequency v, an atom with velocity


e
component is absorbing at frequency v[l - (e/c)] in its own frame, and
Hence the profile is Lorentzian with a half-intensity width equal to the sum the absorption coefficient for that atom is ex{v - ev/c). The total absorption
of the half-intensity widths of both levels. coefficient at frequency v is thus given by the convolution integral

Exercise 9-1: Evaluate equation (9-23) by contour integration and verify equation (9-28)
(9-24). .....
Equation (9-28) can be applied to any absorption profile to allow for the
The Lorentz profiles calculated above are, strictly speaking, emission effects of Doppler broadening; for the remainder of this section consideration
profiles. If, however, we assume detailed balancing, then the absorption profile will be restricted to the case where the intrinsic profile is Lorentzian.
will have the same form. To convert to absorption cross-sections per atom Substituting equations (9-26) and (9-27) into equation (9-28), and defining
we recall from equations (4-34) and (4-35) that
v = (v - v0 )//!J.v0 (9-29)

(9-25) y = (!:,.v/fJ.vo) = (e/eo) (9-30)

Thus using a profile of the form of equation (9-19) and converting to ordinary and a = (f/4n fJ.v 0 ) (9-31)
frequency units, the absorption cross-section is
where !:,.v 0 is the Doppler width of the line
2 2
ne ) (r/4n ) = (eovo/c) (9-32)
ex, = ( mc f (V - Vo)2 + (r/47t)2
(9-26) /!J.vo
we find the absorption coefficient can be written as
Radiation damping is of primary importance for strong lines in low-density
media, for example Lex in interstellar space. In most cases of interest in stellar ex, = (n½e2f/mc fJ.v 0 )H(a, v) (9-33)
atmospheres, however, the line is formed in regions where the density of
perturbing atoms, ions, and electrons is high enough that pressure broadening where
a f""
H (a, V) -= -7t - ""(v - y)2
e-y> dy
(9-34)
is significant (or dominant). + a2
280 The Line Absorption Profile 9-3 Collision Broadening: Classical Impact Theory 281

is known as the Voigt function. In deriving equation (9-34), the approximation The function F(v) is known as Dawson's integral; efficient techniques for
that ~v/c ~ ~v 0 /c (appropriate in a stellar atmosphere) was made. Extensive computing F(v) are given in (170). The functions H.(v) are tabulated in
tables of H(a, v) are given in (219) and (314). (281) and (11,325) for n ~ 4.
General methods of computing the Voigt function are described in (314;
527; 528). The usual case of astrophysical interest is when a « 1; in this Exercise 9-2: (a) Show that the Voigt function has the normalization
limit one can develop a convenient expression for H(a, v) as a power series
in the parameter a as follows. Using the Laplace transform of the cosine
f.'.". , H(a, v) dv = nt
function (b) Prove the relation H.(v) = -[d H._ 2(v)/dv 2 ]/[n(n - I)] from which higher-
2

order terms can be generated by recursion from H 0 (v) and H 1(v). (c) Write explicit
(9-35)
formulae for Hz(v), H 3 (v), and H4(V), expressing H 3(v) in terms of F(v). (d) Show
that for (a 2 + v2 ) » I, H(a, v) ::::: (n-ta)/(a 2 + v2 ).
and the addition rule for cosines, the Voigt function can be written as

H(a, v) = 7t-l Jojoo dx e-ax cos vx f-oo dy e-yl cos xy


00
(9-36)
In view of the result of Exercise 9-2(d), we see that for v » 1, H(a, v)
a/(1t¼v 2 ). Thus a schematic representation of the Voigt function is
~

But the cosine transformation of the Gaussian is H(a, v) ~ e-• + a/(1ttv2)


2
{9-45)
where the first term applies in the line core, v ~ v•, and the second in the
(9-37)
line wing, v ~ v., the quantity v• being chosen such that the two terms are
equal. The line core is clearly dominated by Doppler broadening, while the
from which we see that
line wings are dominated by the damping profile.
(9-38)

Assuming that a« l, we can replace e-ax by its.power series, and inte- 9-3 Collision Broadening: Classical Impact Theory
grating term by term we find ......
THE WEISSKOPF APPROXIMATION
~
co
01)

H(a, v) =L a"H.(v) (9-39) The simplest classical impact theory has its origins in an analysis by
••0 Lorentz, who considered the atom to be a radiating oscillator that suffers I
= [{- l)"/1t½n!] f0 00
changes in phase during encounters with perturbing particles. It is assumed
where H.(v) e-" 214 x" cos vx dx (9-40) that the collisions occur between the radiating atom and a single perturber,
one at a time. The collisions are assumed to occur essentially instantaneously,
From equation (9-37) it follows immediately that so that the wavetrain suffers an instantaneous phase dislocation that, in
effect, terminates it. During the time between collisions the atom is assumed
Ho(v) = e-v2 (9-41) to be unperturbed. Thus suppose that the time between two successive
Integrating by parts, the first odd-order term can be written as collisions is T, and that in this interval the radiator emits a monochromatic
wavetrain f(t) = exp(iw 0 t). The Fourier transform of this.finite wavetrain is
H 1(v) = {-2/1tt)(1 - v f0
00
e-" 214 sinvxc/x) (9-42)
F(w T) = jT e'<wo-wlr dt = exp[i(w - Wo)T] - 1 (9-46)
and using the sine transform of the Gaussian
' Jo i(w - w 0 )

f = e-•' f:
The energy spectrum E(w, T) of this wavetrain is given by substitution of
0
oc e-yl sin 2vy dy e>" 2 dy = F(v) (9-43) F(w, T) into equation (9-3).
we have In general there is not a unique time interval between collisions; rather,
(9-44) these intervals are distributed probabilistically about some mean value. If
282 The line Absorption Profile 9-3 Collision Broadening: Classical Impact Theory 283

the collisions occur as a result of a random-walk process, and the mean time Following Weisskopf (661) we assume that (a) the perturber is a classical
between collisions is T, then the probability that the interval between two particle; (b) the perturber moves with constant velocity past the atom on a
successive collisions lies on the range (T, T + dT) is straight-line path with impact parameter p; (c) the interaction between atom
and perturber is described approximately by
W(T) dT = e-Tf•(dT/t) (9-47)
(9-53)
Hence averaging over all collision times T, we obtain a mean energy spectrum
where r(t) = (p 2 + v2t 2)½, t = 0 occurring at the point of closest approach;
E(w) = (E(w, T))T = (21t)- 1 J0 CX) F*(w, T)F(w, T)W(T) dT (9-48) and (d) no transitions in the atom are produced by the action of the perturber.
The validity of these assumptions will be considered ·later. The form of the
Computation of the integral, with normalization, yields interaction in equation (9-53) is only approximate but holds over a fairly
(1/1tt) (r/21t) wide range of distances. The value of the exponent p depends upon the
E(w) = (w - Wo)2 + (l/t)2 = (ca - wo)2 + (r/2)2 (9-49) nature of the interaction. Values of astrophysical interest and the interaction
they represent are as follows: p = 2, linear Stark effect (hydrogen + charged
Exercise 9-3: Derive equation (9-49). particle); p = 3, resonance broadening (atom A + atom A); p = 4, quadratic
Stark effect (nonhydrogenic atom + charged particle); p = 6, van der Waals
The collision broadening theory described above again yields a Lorentz force (atom A + atom B). The interaction constant CP must be calculated
profile (a result of assuming that all the collisions are distinct), with a damping from quantum theory or measured by experiment.
parameter r = 2/t. To complete the theory we must obtain an estimate of T. The phase shift induced by the perturbation is
As was done for the radiation-damped oscillator, we take the profile of an
ensemble of randomly phased oscillators, continuously crea'ted, to be pro- 17(1) = f 00 6.w(t') dt' = CP f~<X) (p 2 + v2t' 2)-P12 dt' (9-54)
portional to the energy spectrum of a single oscillator [averaged' over all
times, as given in equation (9-49)]. If both radiation and collision damping The total phase shift 17(p) = 17(t = oo) is found directly to be .....
occur, with widths r R and r c respectively, and are assumed to be completely·
uncorrelated, then the profile is a convolution of the two Lorentz profiles., (9-55) 01
By an analysis similar to that leading to equation (9-24), one may readily 0
show that the combined profile is again Lorentzian with a total width where (9-56) I
r = r R + r c• The effects of Doppler broadening can be taken into account
as in §9-2, by using a Voigt profile with the appropriate total damping width. Here r denotes the usual gamma function; for p = (2, 3, 4, 6) one finds
We must now calculate the mean collision time T. If the radiating atoms 1/J P = (7t, 2, 1t/2, 31t/8}.
and perturbing particles have atomic weights A, and AP respectively, and We now assume that only those collisions that produce a total phase shift
both have a Maxwellian velocity distribution at a temperature T, then their greater than some critical value 17 0 ai:e effective in broadening the line. The
average relative velocity is effective impact parameter for such collisions is thus
(9-50) (9-57)
Assuming that the effective impact parameter of the collisions responsible
and the corresponding value for the damping constant is
for the broadening is p0 , we then have
(9-58)
t- 1 = 1tp/Nv (9-51)
and r = 21tp/Nv (9-52) Weisskopf arbitrarily adopted 17 0 = 1 as the critical phase shift; with this
choice we obtain the Weisskopf radius Pw from equation (9-57) and the
where N is the perturber density. We must now determine p0 • Weisskopf damping parameter r w from equation (9-58).
284 The Line Absorption Profile 9-3 Collision Broadening: Classical Impact Theory 285

If CP is given, the theory described above yields a definite value for r, the averages, i.e.,
and the results are found to be of the right order of magnitude. Yet there
remain serious defects in it. (a) The choice l'/o = 1 is arbitrary, and there is d<f>(s) = (e'q<r,•>) 7 (e'q' - 1) 7 = ef,(s)(e'q' - 1) 7 (9-64)
no means of determining a priori the correct value of11 0 to be used. (b) The lfwe can calculate the average of e1q', we obtain a differential equation for ef,(s).
theory does not account for the collisions that produce small phase shifts
By forming the average over a sufficiently long time interval T, the
even though the number of such collisions increases as p 2 • (c) The theory
randomly-occurring collisions will happen at all values of p with an appro-
fails to predict the existence ofa line shift; as will be shown below this failure
priate statistical frequency. We then invoke the ergodic hypothesis to replace
arises from the omission of weak collisions, as mentioned in (b).
the average over time by the appropriate sum over impact parameters. The
number of impacts that occur on the range (p, p + dp) in time ds is just
THE LINDHOLM APPROXIMATION (21tp dp)Nv ds, hence
A significant improvement in the classical impact theory was made by
Lindholm (397;398) and Foley (221). In this approach we consider the radia-
( e'q' - 1) r -+ ( efq'(p) - I) P = 21tN v ds f 00

0
[ e'q(p) - 1]p dp (9-65)
tor to have an instantaneous frequency w(t) which, because of perturbations, The integral in equation (9-65) has both a real and imaginary part, so we write
differs from the nominal frequency w0 by an amount tiw(t). Then we write
(e 1q'(p) - 1)11 = -Nv ds(uR - io- 1) (9-66)
f(t) = exp [iw 0 t + i f~co tiw(t') dt'] = e 1wor+q(llJ
1
(9-59)
where uR = 21t f 00
[1 - cos 11(p)]p dp = 41t f 00
sin 2 [½1'/(p)] p dp (9-67)
0 0

where l'/(t) is the instantaneous phase of the oscillator. To ,obtain the line
profile, we calculate the autocorrelation function <I>(s) defined by equation
and u1 = 21t f
0
00
sin 11(p)p dp (9-68)

(9-6). Let ef,(s) be defined by ef,(s) = e- 1-<I>(s), which eliminates the unper- Combining equations (9-64), (9-65), and (9-66) and solving the resulting __..
turbed oscillation. Then from equation (9-6), differential equation with the initial condition ef,(0) = 1 we find
U1
rf,(s) = Jim r-• fT/2 e-lwose-l(o,ol+~(!lJellOJo(l+•)+q(r+,)) dt (9-69) __..
T->co -T/2
Finally, calculating the intensity from equation (9-7) and normalizing the
= Jim r- 1 fT/l ellq(r+•)-q(I)) dt (9-60)
profile we obtain
T->co -T/2

Clearly ef,(s) is the time-averaged value of the additional phase shift occurring (9-70)
in the time-interval s. For brevity, write
11(t, s) = 11(t + s) - 11(t) (9-61) Exercise 9-4: Verify equations (9-69) and (9-70).

Then ef,(s) = (exp[i11(t, s)])T (9-62) Thus Lindholm theory yields a Lorentz profile with a damping width

Further, writing drf,(s) = rf,(s + ds) - rf,(s), we have (9-71)


drf,(s) = (e1q<r,,i(e'q' - l))T (9-63) and a line shift (9-72)

where 11' denotes the change in phase occurring in the time-interval ds as a The prediction of a shift is in agreement with experiment, where such shifts
result of collisions that take place in that interval. The phase change cannot are observed. Quantum theory yields a profile of the same form as equation
be correlated with ·the· current value of the phase if the collisions occur at (9-70), and gives explicit expressions for rand 6w 0 in terms of matrix ele-
random. Thus the average of the product can be replaced by the product of ments of the perturbing potential and transitions within the atom. As we
286 The Line Absorption Profile 9-3 Collision Broadening: Classical Impact Theory 281

shall see below, Lindholm theory yields a unique value of r/ll.w 0 for each Resonance broadening, p = 3, is of importance mainly for collisions of
choice of p; quantum theory shows that this ratio actually varies over a hydrogen atoms with one another. As the atmosphere must be hot enough
moderate range as T and N vary, and is different for each line. The effects that hydrogen is excited to then = 2 state (to produce the observable Balmer
of Doppler broadening are taken into account by using a Voigt profile with lines) but cool enough that it is not dominantly ionized, resonance broadening
the appropriate damping parameter and shifted from its rest frequency by effects are of interest for solar-type stars. The interaction constant C3 in
an amount ll.w 0 • equation (9-53) for level n is
The dominant contributions to CTR and c, 1 come from quite different ranges
of impact parameter p. From equation (9-55) we note that r,(p) oc p-<p- 1l, (9-74)
Thus for (p/pw) > 1, the integrand of aR rapidly drops to zero [see (629, 16)
or (638, 305)], and the dominant contribution to the line broadening comes from
[see (662; 112, 231)]. A quantum mechanical calculation gives a r slightly
different from the Lindholm value, namely ·
(strong) collisions inside the Weisskopf radius-i.e., (p/pw) < 1. In contrast,
for a,. the integrand for (PIPw) < 1 fluctuates in sign, and averages to nearly (9-75)
zero. Thus the dominant contribution to the line shift comes from (weak)
collisions outside the Weisskopf radius. It is easy to understand physically how For the hydrogen lines there is no shift ll.w 0 because individual Stark com-
the shift arises. The very weak collisions (r, « 1, p » Pw) are extremely ponents split symmetrically about line center (see §9-4) and the shift is
numerous and occur at an essentially constant rate, yielding an average identically zero. Resonance broadening is most significant for the lowest
phase change per unit time of members of a series where Stark broadening is the smallest. The effects of
resonance broadening have been shown to be important in the solar Hr,. line
'ff = 21tNv f,7 r,(p)p dp (9-73) but negligible for higher series members (146).
Quadratic Stark effect, p = 4, is important for the broadening of lines of
where p* is chosen to assure that r,(p*) « 1. But as can be s~en from equation nonhydrogenic atoms and ions by impacts with charged particles (electrons),
(9-59), the rate of change of phase is by definition a change ll.w ip the oscil- and is the dominant pressure-broadening mechanism for these lines in the
lator's frequency. atmospheres of early-type stars. In applications of the classical Lindholm
theory the interaction constant C4 was typically estimated from experimental ......
measures of line shifts in electric fields or from time-independent perturba-
SPECIFIC CASES 01
tion theory for the quadratic Stark effect [see (11, 319-320) or (638, 326-328) I\)
Lindholm theory has been most widely applied in astrophysical work for examples of this procedure]. The resulting damping widths are usually
much too small, however, because the Lindholm approximation assumes the
I
for the cases p = 3, 4, and 6. For these cases the integrals in equations (9-67)
and (9-68) can be evaluated explicitly to yield the values listed in Table 9-1 collisions are adiabatic (i.e., do not cause transitions in the radiating atom);
[ see (629, 14) for details]. The last line of the table gives the value of r, 0 that, this assumption is frequently poor, and accurate quantum mechanical calcu-
when inserted into the Weisskopf formula [equation (9-58)], gives the . lations including nonadiabatic effects (cf. §9-5) yield much larger line-widths.
Lindholm r. As r, 0 is always less than unity, it can be seen that the Weisskopf Van der Waals interactions (p = 6) of noh-hydrogenic atoms in collisions
formula always leads to too small a value of r. with neutral hydrogen atoms is the major source of pressure broadening in
solar-type stars. The usual classical treatment accounts for the dipole-dipole
interaction term in the potential, and yields [cf. (629, 91-97; 638, 331-334)]
TABLE 9-1
Results of Lindholm Theory (9-76)
p 3 4 6 where ex is the polarizability of hydrogen and R. 2 is the mean square radius
of the two levels. Quantum mechanical results are sometimes available for
r 21t 2 C3N 11.37}
9.85 c. i Vt N 8.08} C tvtN R. 2 ; if not, hydrogenic estimates are used. Using C6 determined in this
2,94 6
6wo
manner, r can be computed from Lindholm theory. When this is done
r/6wo 1.16 2.75 (e.g., for lines of Fe I), it is found that the predicted values are much too small,
'lo 0.64 0.64 0.61 by factors of 5 to 30 (382). Quantum mechanical calculations that again
288 The Line Absorption Profile 9-4 Collision Broadening: Statistical Theory 289

employ only the dipole-dipole term do not lead to large increases in r shifts (i.e., D.cv,, » I) and hence to the strong collisions that occur inside the
[e.g., (264; 98; 86; 599)], which points to the breakdown of the dipole-dipole Weisskopf radius. It is difficult to construct a theory that makes the transition
approximation rather than other theoretical problems [see also (301; 302; from impact to statistical theory. A useful conceptualization is to suppose
541)]. Calculations using the more realistic Lennard-Jones potential (303) that there is a "boundary" frequency 6w 9 inside of which impact theory holds
lead to significantly larger widths. Attempts have been made to include more and beyond which statistical theory is valid. To a fair approximation [see
terms in the expansion of the interaction potential (233); this increases r, (637; 306)] 6.rog :::.: 6ww where llww denotes the frequency shift produced
but still falls short of the observed values. The expansion technique converges by a perturber at the Weisskopf radius Pw; i.e.,
slowly, and an alternate approach, applied to the lines of Fe I, evaluated the
exact expression for the interaction using scaled hydrogenic wave functions (9-78)
(116) and showed reasonable agreement between predicted and measured r's.
Note that 6ww corresponds to a phase shift of unity. We shall see in §9-4 that
Yet another method is based on the proposal that the dominant interaction
equation (9-78) implies that broadening of the hydrogen lines by ions follows
leading to the line broadening is between the perturber and the valence
the statistical theory, while electron broadening is gi\len by impact theory.
electron (539; 540). One can then use a Smirnov potential (578) and obtain
(4) Classical impact theory assumes that the collisions are adiabatic-i.e.,
an expression for the damping parameter. This approach has been used to
transitions are not induced in the atom. A collision occurring in an impact
produce extensive tables (193) giving the parameters a and p in the formula
time '• will have Fourier components of frequencies up to w, ~ 1/t,. To
r = No:T/J, as functions of the effective quantum numbers n• of the lower guarantee that the collision is adiabatic, w, should be much smaller than any
and upper levels, for s-p, p-d, and d-f transitions.
characteristic transition frequency wlJ; i.e.,

VALIDITY CRITERIA (9-79)


(1) An effective impact time t, can be defined such that,t, times the peak For nondegenerate levels, the energy separation is often large enough that
value of 6.w for a collision at the effective impact parameter, namely CPp 0 -p, the condition stated above will be met. But for degenerate levels (e.g., for
yields the total phase shift given by equation (9-55). This gives hydrogen), the energy separation between levels will be proportional to the
(9-77)
IE, -
perturbing field itself; i.e., £11 :::.: hCP/pP = h llw(p). Then the condi-
tion for adiabaticity implies that 6w(p)t, = 17(p) » 1; that is, only collisions 01
For impact theory to be valid, we demand that only one collision at a time inside the Weisskopf radius will be adiabatic. In the· case of hydrogen, llww w
occur, so that t, < t = 1/(Nnp/v). Writing N = 3/(4nr/) where r 0 is the for ions is very small, and for virtually the entire profile the statistical theory I
mean interparticle distance, we find (t,/t) = ¾v, p(p 0 /r0 ) 3 • Thus impact theory is valid, and the collisions causing the broadening occur inside the Weisskopf
will be valid only when the density of particles is so low that the Weisskopf . radius. Thus the ion broadening will be adiabatic. Precisely the opposite is
radius is small compared to the interparticle distance. true for electrons. Here 6ww will be large, and almost the whole profile is
(2) It is clear that, as p -+ oo, the effective impact time t, becomes larger described by impact theory with the relevant collisions lying outside the
and larger, and eventually exceeds t, the mean time between collisions, so Weisskopf radius. The electron broadening is strongly nonadiabatic (and
that the collisions overlap. Thus from the very weak collisions there is an hence must be described by quantum theory). When the adiabatic assumption
essentially continuous perturbation of the atom, and here we expect statistical breaks down, much larger damping parameters than those predicted by
theory to begin to be valid. Indeed, we saw earlier that these weak collisions classical theory are found; for this reason the modern quantum mechanical
produce the line shift, just as would be given by the application of a steady results are often drastically different from earlier classical work.
perturbation. Even though Lindholm theory (an impact theory) treats the
weak collisions, the calculation is not strictly logically consistent.
(3) Impact theory fails for sufficiently large frequency displacements llw 9-4 Collision Broadening: Statistical Theory
from line center, and statistical theory becomes valid. In impact theory, it
follows from the general properties of Fourier transforms that the charac- The basic picture in this theory is that the atom finds itself radiating in a
teristic interruption time t corresponding to a frequency displacement 6.w is statistically fluctuating field produced by randomly distributed perturbers.
t ~ 1/6.w. For sufficiently large 6.w's we will eventually have t « t., and The motion of the perturbers is ignored; this is known as the quasi-static
impact theory breaks down. These values of llw correspond to large phase approximation. (As we shall see later, this approximation is good for the
290 The Line Absorption Profile 9-4 Collision Broadening: Statistical Theory 291

slow-moving ions-e.g., protons-in the plasma.) A specific distribution of


perturbers produces a definite field; the relative probability of fields of dif- (9-84)
ferent strengths is thus determined by the statistical frequency with which
particle distributions producing the appropriate strengths are realized. For a and equation (9-83) can be rewritten as
given value of the field, the oscillation frequency of the radiator is shifted by
a definite Aw. The intensity of the radiation at this Aw is assumed to be W(r) dr = exp[ -(Aw0 /Aw) 3''] d(Aw0 /Aw) 3' ' (9-85)
proportional to the statistical frequency of occurrence of the appropriate
field. Thus the central problem is to determine the probability distribution
For the case of linear Stark effect, the perturbing field is F = (e/r 2
). If we
define the normal field strength to be
of the perturbing fields. Once this is ki1own, line profiles can be computed.
The applications in this section will be.restricted to quasi-static broadening
of hydrogen lines by linear Stark-effect interactions with protons (though the
theory is relevant in other contexts as we!J).
Fo = (e/r/) = e G 1tNr = 2.5985eNt (9-86)

and measure Fin units of FO [i.e., fJ = (F/F 0 )], then nearest-neighbor theory
THE NEAREST-NEIGHBOR APPROXIMATION yields
As a first approximation, we assume that the main effect on the radiator W({J) d{J = ~ p- t exp( - p-~) d{J (9-87)
results from the strongest· perturbation acting at any given instant-namely, 2
that from the nearest neighbor-and that the effects of all other particles are Clearly, as {J--+ oo, W({J)--+ Jp-t; hence statistical theory predicts that, in
neglected. Then if W(r) dr is the probability that the nearest neighbor is the wings of a line broadened by linear Stark effect, the profile falls off as
located on the range (r, r + dr) from the radiator, the frequency spectrum is Aw-t, in contrast with the prediction Aw- 2 given by impact theory.
/(Aw) d(Aw) ex: W(r)[dr/d(Aw)] d(Aw) (9-80) The basic failing of this theory is that the profile is, of course, the result
of perturbations by all particles, not just the nearest neighbor. To obtain
where it is assumed that Aw is given by equation (9-53); i.e., Aw = C,/r'. accurate results a more elaborate theory must be constructed. ......
To find W(r) we calculate the probability that a particle is located on the 01
range (r, r + dr), and that none is at a distance less than r. Then assuming it ~
HOL TSMARK THEORY
uniform particle density N, W(r) is given by I
The effect of an ensemble of particles upon a radiator was determined by
W(r) = [1 - J; W(x) dx]<4nr2 N) dr (9-81) Holtsmark (30S), who calculated the net vector ·field, at the position of the
radiating atom, from the superposition of the field vectors of all perturbers.
An elegant treatment of the problem was given by Chandrasekhar (ISi); this
where the factor (41t r 2 N) dr is the relative probability of a particle lying in the
paper should be consulted for the derivation of the results quoted here:
shell (r, r + dr) while the term in square brackets is the probability that no
For an interaction of the form F = C,/r', the analysis yields
particle lies inside this shell. By differentiation we find

_<!_ [ W(r)] = -(41tr2 N) [ W(r)] (9-82)


, W(/J) = (2{J/1t) f0
00
exp( - y 31 P)y sin {Jy dy (9-88)
2
dr 41tr N 4nr 2 N
Here fJ = F/F 0 , the normal field strength F0 now being defined as
and thus by integration and normalization,
F0 = yC,N'13 (9-89)

W(r) = 4nr 2
N exp ( - ; 1tr 3 N) (9-83) where y = {(2n 2
p)/[3(p + 3)r(3/p) sin(31t/2p)] }P13 (9-90)

In particular, for linear Stark effect, p = 2, c, = e, and y = 2.6031 so


It is customary to adopt the mean interparticle distance r0 = (3/41tN)½ as the that F 0 = 2.6031eNt, which differs only inconsequentially from the normal
reference distance at which a perturber produces the normal frequency shift field strength given by nearest-neighbor theory.
292 The line Absorption Profile 9-4 Collision Broadening: Statistical Theory 293

The integral in equation (9-88) cannot be evaluated exactly for p = 2, In view of equations (9-93) through (9-95), this expression reduces to
and W(P) must be expressed in a series expansion. For small p

4) 00
(41 + 6) ep = -eij, ( n, + ~ Z,2n,) (9-98)
= (37t ,~o (-l)r - 3 -
I p<21+2l
W(P) (21 + l)! (9-91)
Substituting equation (9-98) into (9-96), we may rewrite Poisson's equation
while for p » 1 one finds an asymptotic expansion as 'v 2 <f, = (<f,/D 2 ) where

W(P) = 1.496p-½(l + 5.107p-t + 14.43p- 3 + · · ·) (9-92) D = (kT/4ne 2 )½ [ n, + ~ Z,2n,Tt (9-99)

the leading term of which is essentially the same as given by nearest-neighbor is the Debye length. Solving for <f, we find <f, = ,- 1 [Ae-,tD + Be'10]. De-
theory. Tabulations of W(P) are given in (151) and'(629, 28).
manding that <f, ... Oas r ... oo, we set B = 0. Further, to recover the potential
of the ion itself as r ... 0, we set A = Z;e. Thus the potential produced by an
ion imbedded in a plasma is
DEBYE SHIELDING AND LOWERING
OF THE IONIZATION POTENTIAL <f,(r) = [Z,e exp(-r/D)]/r (9-100)
In deriving the probability distributions described above, interactions As is clear from equation (9-100), beyond the Debye length the field of an
among the perturbers were ignored. In reality, the probability that a particle ion is strongly shielded and rapidly vanishes. Physically this occurs because
is found in a volume dV is not just N dV, but depends also upon the electro- a charged particle tends to polarize the plasma in its vicinity, and the
static potential </) in dV. For example, if at some point <f, > _o, electrons ~ill oppositely charged particles that cluster around it shield the field of the
tend to migrate toward it while ions will tend to migrate away,,and vice original particle at· 1arge distances. Thus the Debye length sets an upper
versa if <f, < 0. Following Ecker (203; 204; 205; 206), we may account _for limit on (a) the distance over which two charged particles can effectively
these effects schematically by introducing a Boltzmann factor depending interact, (b) the size of a region in which appreciable departures ~rom c~a~ge .....L

on if, = (e<f,/kT). Thus for electrons and ions, res~ctively, we write neutrality can occur, and (c) the wavelength of electromagnetic rad1at1on 01
n,W, ,IV = 11, exp(ij,) c/V ~ 11,(1 + if,) dV (9-93)
that can propagate through the plasma without dissipating. 01
In most astrophysical applications we can assume a practically pure I
and n,W, dV = n, exp(-Z,if,) dV ~ n1(1 - Z 1ij,) dV (9-94) hydrogen plasma; then z, = I, and n, = n,,and inserting numerical constants
into equation (9-99) we find ·
where n, and n, are the electron and ion densities, z,
is the ionic charge,
D = 4.8(T/n,)f cm (9-101)
and it is assumed that if, « 1. As the plasma is electrically neutral over
sufficiently large volumes, Exercise 9-5 : Compare Debye lengths in (a) a stellar photosphere with
n, rz,n,
= (9-95) T = 104 °K, n, = 10 14 ; (b) the solar corona T = 106 °K, n, = 108 ; (c) an
H II region, T = 104 °K, n, = 10 2 •
'
We now calculate the potential around a particular ion under the sim-
plifying assumption that all particles can be smeared out into an equivalent To calculate the effect of shielding on W(P) one can make the very simple
charge density. We then may use Poisson's equation assumptions (205; 206) that the field of the perturber is unchanged for
r ~ D, but is identically zero for r > D. One then finds
(9-96)
W(P, b) = (2P o! /n) f0
00
e-d 9(y)y sin(b½py) dy (9-102)
to determine <f,, where p is given by
where g(y) = ~2 yt Jy100 (l - z- 1 sin z)z-½ dz (9-103)
ep = -en,W. + e L Z,n,W, (9-97)
'
9-4 Collision Broadening: Statistical Theory 295
294 The Line Absorption Profile
The treatment of perturber interactions given above is somewhat over-
and o = !nD3 N, the number of particles contained in the Debye sphere. simplified. Very precise calculations of the perturber field-strength distri-
As o_. oo, one expects to recover the Holtsmark distribution; i.e., W(P, oo) = butions have been made using cluster-expansion methods (65; 472; 515),
W8 (P). From equation (9-102) one can show that for o _. oo, W(P, o) = and numerical integrations using Monte Carlo techniques (308; 309; 310;
W8 (P) + o-+F(P) where F(P) is a bounded definite integral. Recovery ~f 482). Most modern treatments of hydrogen-line b~oadening use these refined
the Holtsmark distribution for large ocan also be seen from the asymptotic distribution functions. In practice, the effects of shielding are often quite
expansion important in laboratory plasmas, while in stellar atmospheres the densities
are so low that the number of particles in a Deybe sphere is large (o ~ 100),
and the departures from the Holtsmark distribution are not large.
The presence of nearby charges partly neutralizes the effects of the nuclear
which may be compared to equation (9-92). In principle, for small o the charge upon an orbital electron and thereby weakens the potential well in
theory should merge continuously into the nearest-neighbor approximation, which the electron is bound. The reduction in binding energy can be calcu-
but in practice, when o ,;; S, the assumptions employed (particularly that a lated, using the Debye potential in equation (9-100), as
smeared-out charge distribution may be used) break down. A plot of W(P, o)
for several values of ois shown in Figure 9-1 along with the nearest-neighbor (9-105)
and Holtsmark distributions.
for r « D. An electron in a state that, in the unperturbed atom, lies at an
energy 6.x below the ionization limit, can be considered unbound if 6.x ~ 6.E.
That is, the ionization potential of the atom is decreased by an amount
t'\
I I

0.6
I I
I
I I
I
6.x = Ze 2 /D = (27.2Za 0 /D) eV = 3 x 10-szn,tr-t eV (9-106)
I I
I ~ Nearest-neighbor theory
I I where use has been made of equation (9~101). This calculation of 6.x is only
I I
I I schematic. A discussion from several points of view of the lowering of
o.s I I
I
I
I
I
I
I ,,_ ionization potentials in a plasma is given in (178).

I
I
I
I
I
I
01
I I THE QUASI-STATIC ION BROADENING OF HYDROGEN LINES 0)
0.4 I I
I
I
I
In the absence of a perturbing field, each level of hydrogen is degenerate
W(P,cS)
with 2n 2 sublevels. Analyses by K. Schwarzschild (563) and Epstein (208)
0.3 3 showed that, when an electric field is applied, these sublevels separate and,
because hydrogen has a permanent dipole moment, the energy shift is
Holtsmark distribution directly proportional to the applied field strength F (linear Stark effect). If
no other broadening mechanisms are operative, the line profile will consist
of a number of Stark components, arising from transitions between the
sublevels of the lower and upper states. Each Stark component has a charac-
0.1 teristic relative intensity Ik (561) and will be displaced from line center by
a characteristic shift

(9-107)
0 2 3 4
(1 where Z is the charge on the atom ( = 1 for hydrogen) and nk is an integer.
FIGURE 9-1 The line that we observe will be a superposition of these components,
Probability distribution or field strength at a test point, including weighted by their relative intensities and the probability of being shifted to
shielding effects; t, is the number or charged particles within the the appropriate wave-length position.
Debye sphere. From (205), by permission.
296 The Line Absorption Profile 297

The Stark pattern of a hydrogen line is symmetric about line center with TAIILE 9-2
identical components at ±k with Lk = Ik, c_k = Ck [see, e.g., (638,320; 1'rw1silim1 Waudengt/1 .O.i,w(A) between
S1111i~tical and /111p11c1 Bwatleni11g for
629, 73)]. Assuming that the intensities are normalized such that Lk Ik = 1 Hydrogen Lines
(where the sum extends over all components), then the line profile will be
T
/(.:U) d(A).) = L IkW(F/F0 )(dF/F0 )
k Line Perturber 2.5 x 104 °K 10• °K
= L h W(A)./CkF 0 ) d(A).)/CkF0 (9-108)
Electrons
k H~ 580.0 230.0
Protons 0.63 0.25
It is customary to define the parameter a. as JI/I Electrons 120.0 48.0
Protons 0.13 0.05
(9-109) Hy Electrons 48.0 19.0
Protons 0.05 0,02
where F 0 is the normal field strength F 0 = 2.60eN!, Then the line profile Hb Electrons 32.0 13.0
is given by a function S(a.), Protons O.o3 O.ot

S(cx) dcx =I / W(cx/Ck)(dcx/Ck)


k (9-110)
t
at very large displacements from line center and, as mentioned in §9-3, are
which is normalized on the range ( - oo, oo) for ex. The absorption cross- nonadiabatic. The complete profile consists of the effects of both ions and
section per atom can be written as electrons and, as we shall see in §9-5, the latter increase the line-broadening
markedly. The functions S(cx) for ions alone seriously underestimate the
(9-111) hydrogen-line widths, and a satisfactory theoretical description of stellar
hydrogen-line profiles became possible only after the development of the
Extensive tables of Ct, ltt and S(cx) for numerous hydrogen lines are given
in (634). The largest Ct for a line of upper quantum number n increases as
quantum mechanical line-broadening theory. ......
n2, and recalling that for /J » 1, W(/J) oc p-t, we see from equation (9-108) u,
or (9-110) that the Stark widths of lines rise rapidly up a series, roughly as ---J
113,
9-5 Quantum Theory of Line Broadening I
Let us now consider when the quasi-static profiles are applicable. Let
Quantum mechanical calculations yield precise profiles for pressure-
nk = I Jknk/L It, the sum being taken over positive values only. Then, broadened lines. The development of this theory brought about a major
writing Liw = C2 /r 2 and F = e/r 2, it follows from equation (9-107) that
improvement in one of the most important (and difficult) applications of
(9-112) atomic physics in the analysis of stellar spectra. Excellent discussions of
the general theory can be found in (62; 63; 64; 73, Chap. 13; 179; 264,
with nk ~ ½n(n - 1) for n » 1. From equation (9-78) the wavelength shift Chap. 4; 268; 582); only a brief outline will be presented here. We shall
delimiting the transition between the impact and statistical theories is (for focus attention on the case where the atom suffers impact broadening from
p = 2) electrons and quasi-static broadening by the ions.
(9-113)
THE LINE PROFILE
Notice that A).w ex: v2 , and thus A).w(electron) ~ 10 3 A).w(proton). Using
equations (9-112) and (9-50), we obtain the results listed in Table 9-2. It is As was shown in Chapter 4, the power radiated by an isolated atom in a
obvious that the ion broadening is very well described by the quasi-static transition from an upper state j to a lower state i is [ cf. equation (4-62)]
theory (especially when we note that Doppler motions will dominate in the
core). The electrons, however, are in the impact-broadening regime except (9-114)
298 The Line Absorption Profile 9-5 Quantum Theory of Line Broadening 299

where d is the atomic dipole moment. The total emission, summed over all
possible substates contributing to a line is
Now la, t) satisfies the SchrBdinger equation
Hla, t) = ih(dla, t)/dt) (9-121)
P(w) = (4c.o4 /3c 3 ) I>,o(w - w,1)l(ildli)l 2 (9-115)
,. J Substituting equation (9-120) into (9-121) we can derive a SchrBdinger
Here p1 is the probability that an atom is in the upper state j; in thermo-
la,
equation for T(t, 0). Noting that 0) is fixed in time, we find
dynamic equilibrium HT(t, 0) = ih[dT(t, 0)/dt] (9-122)
(9-116) which has the solution
T(t, 0) = e-lHr/h (9-123)
where U(T) is the partition function.
To calculate the broadening of lines emitted by an atom in a plasma, we where the exponential is to be understood as an operator. We may now
consider the radiating system to consist of atom plus perturbers, and gen- rewrite equation (9-119) in terms of time-development operators, including
eralize the meaning of states It) and Ii) to include perturbers also. The perturbations; we have
profile of the line is then written as
cf,(t) = L P1e'<E,-E,>1/hl(ildli>l2
(9-117) '· J
= L P1<ildli)e-1E,1fh(ildlj)e'EJ1/h
Now p1 refers to the probability of a particular state of atom and perturber. '· J
If the plasma is in thermal equilibrium, p1 is proportional to = L pJ<jldTli)(ildTtli> (9-124)
'· J
(9-118) But the expansion rule, relative to a complete set of states IY), is
where His the total Hamiltonian of the system, HA and Hp are the Hamil- (alfJ> = L (aly)(ylfJ> (9-125) .....
1
tonians of the atom and perturber alone, respectively, and Vis the interaction
Hamiltonian.
01
so we see that equation (9-124) can be rewritten as (X)
It is easiest to account for collisions using the Fourier transform
cf,(t) = L pi(jldTdTtjj) = tr(p dTdTt) (9-126) I
J
(9-119)
The trace includes both atomic and perturber states. This expression is
which is analogous to the classical autocorrelation function [ see (72, 498)]. quite general; the detailed form of cf, depends upon the form of T.
The effects of collisions at a specific impact parameter upon the autocorrela-
tion function (and statistical averages over all possible perturber paths) can THE CLASSICAL PATH APPROXIMATION
be calculated directly. The intensity profile then follows from the inverse
Fourier transform [ cf. equation (9-7)]. Let us now consider the calculation of T in more detail. Let the wave
The problem at hand is to obtain an expression for cf,(t). To do this we must function t{I describing the solution of the complete system of atom plus
find the change with time of the eigenstates Ii) and Ii) under the effects of perturber be the solution of the equation
perturbing collisions in terms of a time-development operator T(t, 0). This
operator is defined such that a state of the system at time t is related to the
ih(dt{l/dt) = (HA + Hp + V)t/1 (9-127)
state at time t = 0 by Note that HA is independent ofperturber coordinates, Hp is independent of
la, t) = T(t, 0)la, 0) (9-120) atomic coordinates, while V depends on both. Further, because the atom
300 The Line Absorption Profile 9-5 Quantum Theory of Line Broadening 301

is in a static ionic field of strength F, HA = HA(F). For brevity, F will make the identification
be suppressed here but included explicitly in the final result.
To make progress, we now assume that the perturber follows its c/assicc.l (9-131)
path-i.e., a straight line past a neutral atom and a hyperbola around an
ion. This assumption will be valid when the deBroglie wavelength is small where V,,1(t) is the classical interaction potential; this is the essence of the
compared to the impact parameter for those collisions that dominate the classical path approximation. The Schrodinger equation for the atom's
broadening. That is, we must have p » J. = (h/mv) or mvp » h. But mvp is time-development operator then becomes
just the angular momentum of the perturber ( = lh); the criterion just stated
is equivalent to the requirement that the quantum number I be much larger (9- I 32)
than unity. Under such circumstances the classical-particle picture would
be expected to be valid on the basis of the correspondence principle. The and the time-development operator for the complete system is
validity of the approximation must always be checked in line-broadening
calculations; in most cases of astrophysical interest it is found to hold.
T(r, 0) = TA(r, 0)Tr(r, 0) = TA(r, 0)e-mP,,h (9-133)
Some perturbers that violate the condition can always be expected, but the Finally, we write the probability-density matrix p as p = PAPP where PA
approximation remains useful if they do not dominate the broadening [see refers to atomic states only, and pp refers to perturber states only and is
also (73, 498 ff.)]. A theory that does not use the classical path approximation diagonal in the perturber coordinates. Again, this may be done if the back-
is given in (64). reaction of the atom on the perturber can be neglected.
The wave function for the system of atom plus perturber is assumed to be When these expressions for p and T are inserted into equation (9-126)
separable; i.e., tf,(t) = cx(t)1t(t), where cx(t) and 1t(t) are the atomic and perturber for </>(r), and the separated form of 1/1 is recalled, we find
wave functions. We further suppose the perturber path to be fixed and to
be independent of the state of the atom with which the inreraction takes (9-134)
place. In this way the effect of the perturber on the atom is taken into account,
The trace over perturber states has reduced merely to a thermal average
but the back-reaction of the atom on the perturber is ignored. This will be ......
valid if the energy gained or lost by the perturber (of order hr, where r is over all perturbers (denoted by braces) and trace is now carried out over
the linewidth) is much less than its kinetic energy (kT), a condition that is atomic states only. 01
(0
almost always satisfied. A few collisions will always occur in which large
energy exchanges take place, but again these will not invalidate the I
THE IMPACT APPROXIMATION
assumption if they do not dominate the broadening.
Under the assumptions made above, 1t(t) is the solution of To calculate </>(t) as given by equation (9-134) we assume that both the
ih[d1t(t)/dt] = Hp1t(t) (9-128) initial state a and the final state b consist of several substates, denoted by
ex and /J, respectively, and that dipole transitions exist only between substates
and the time-development operator for the perturber is of a and b, but that radiative transitions among the substates of a or b can
be ignored. On the other hand, we ignore collision-induced transitions
Tp(t, 0) = e- 111 ••' 1h (9-129)
between states a and b, and assume that collisions can result only in transi-
Now consider the SchrBdinger equation for the atom alone. To obtain it, tions among substates of a orb. That is, we have (cxldlcx') = 0, <fJldlP') = 0,
we multiply equations (9-127) and (9-128) by 1t•, integrate over perturber and (cx!T.,bl/J) = 0. Then writing out the trace in equation (9-134) we find
coordinates, and subtract to find
</>(t) = Pu L (cxldlfJ)(/J'ldlcx')(cxl<Pl{TbT:}lcx')IP') (9-135)
ih[ dcx(t)/clt] = (HA + f n* V1t cltp) cx(r) (9-130)
•• •·,p, P'

Here we have neglected the variation of Pa among the upper substates and
If the perturber wave packets are indeed narrow enough that the perturbers have noted that only the time-development operators can depend upon the
can be considered to be classical particles on classical paths, then we can statistical averages.
302 The Line Absorption Profile
9-5 Quantum Theory of Line Broadening 303
It is convenient to replace the complete time-development operator T by
U, the time-development operator in the interaction representation, defined as Then performing the inverse Fourier transformation, and reintroducing the
quasi-static ion field F, we obtain the intensity profile
(9-136)
This definition factors out the time-behavior of an eigenstate and isolates
the perturbation effects; for an unperturbed eigenstate U(t, 0) = 1. By
I(w, F) =~gee C..~.P' (ixjdj/1)(/1'ldlo:')
substitution into equation (9-135) we have
x (o:l</JI [iw - <I>ab(F) - ~ [Ha(F) - Hb(F)]r 1 lix')I/J')) (9-144)
cp(t) = Pa L (o:ldl/1)(/J'ldlo:')(o:l<Ple- 18" 1'1'{Ubu:}e1H••1h10:')I/J')
a:, cz', /J, fl'
where use has been made of the fact that <I>0 b(F) is found to have a negative
(9-137) real part. If W(F) is the probability of an ion field of strength F, the final
profile, averaged over all ion fields is
Substituting equation (9-136) into (9-132) we obtain a SchrBdinger equation
for U(t, 0),
ih[dU(t, 0)/dt] = e18 A'1hV,,(t)e-mA,,hU(t, 0) = v;,(t)U(t, 0) (9-138)
/(w) =: 0
f
0
00
W(F) fie C.)i,p• (o:jdl/J)(/J'ldlix')
We solve this equation by iteration, starting with U(t, 0) = 1 as a first
approximation, to obtain a solution of the form
x <ixl</31 [iw - <I>ab(F) - f 1
[Ha(F) - Hb(F)]r lix')l/3')) dF

(9-145)
U(t, 0) = 1 + (ih)- 1 J~ v;,(t1) dt1 + (ih)- 2 f~ dt2 v;,(t2) f~· dti v;,(ti) + ... Equation (9-145) is quite general, and has been used in most quantum
(9-139) mechanical calculations of Stark-broadened line profiles. The result is valid
if(l) the interval At in equation (9-140) can be chosen to include a complete
Now consider the calculation of the statistical average {Ubu:}. We pro- collision; (2) when the collisions overlap, they are weak enough that their .
ceed in a way analogous to that used in Lindholm theory. That is, we write contributions to the iterative solution for U are simply additive; and (3) the
the change in { UbU:}, caused by a specific collision (treated as an impact), perturbers can be treated as classical particles. These validity criteria must CJ)
in some time At as A{Ubu:} = {Ub(t + At, t)U:(t + At, t) - l}{Ubu:}, be checked in each case.
where again we argue that the changes in (t, t + At) are statistically indepen- 0
dent of the current values, so that we can replace the average of the product APPLICATION TO HYDROGEN
I
by the product of the averages. Then using equation (9-139) on the interval
(t, t + M) one obtains an explicit expression for {Ub(t + flt, t)U:(t + flt, t) - 1} One of the most important applications or'the theory outlined above has
[see, e.g., (264, 70; 268, 37)] of the form been to the calculation of the effects of electron impacts upon the broadening
of hydrogen lines. The theory has reached a very refined state, and the
{Ub(t + At)U:(t + At) - 1} = el(H.-H.Jr/A(<I>.b At)e- 1<n.-n.ir/A (9-140) theoretically predicted profiles are in excellent agreement with laboratory
measurements (671) and provide satisfying fits to observed stellar profiles
We thus obtain a differential equation for {UbU:}. namely (see Figure 10-4).
The impact theory outlined above has been intensively applied by Griem
A{ ubu:} = ei(H.-H.)l/h<1>.be-l(H.-H.)1/h{ ubu:} At (9-141) and his collaborators (263; 265; 270; 271). Although the evaluation of <I>0 b
whose solution is is straightforward in principle, it is complicated in practice, for it entails
cutoffs both at small and large impact parameters, the former arising from
{Ub(t, 0)U:(t, 0)} = el(H.-H.lr/A exp[ - i(Hb - H0 )t/h + <1>0 bt] (9-142) strong collisions inside the Weisskopf radius (which are no longer correctly
described by the iterative series development for U, the time-development
Now, substituting into equation (9-137), we obtain operator) and the latter to account for Debye shielding effects. Further cutoff
procedures are required to allow for the transition of the electrons from the
cp(t) = Po L (o:ldl/1)(/1'ldlo:')(1Xl(/11e1'<H•-Ho)l/h+tl>••'1jix')l/1') (9-143) impact- to statistical-broadening regimes. This work culminated in the publi-
a:,au,/J,/J'
cation of extensive tables (353; 356) for the first four members of the Lyman
t.i.(A) 9-5 Quantum Theory of line Broadening 305
_ r--0.,...1_ ___,o.,...3__,o.....
s__•,...·o_ _ _J..,..o_s.....0_ _1~0.o_ _ _Jo-.o_s_o~.o
10 0 and Balmer series lines. These tables give S(cx), analogous to the profile
.,::;::.;~--==~~-..
described by equations (9-109) through (9-111), but including the effects of
,·..,•, electron impacts. An alternative approach developed by Cooper, Smith, and
Vidal (581; 647; 648; 649) uses a unified theory that accounts automatically
for the transition of electrons from the impact to quasi-static regimes. Exten-
sive tables of results based on this theory have been published (650) for the
1.0 first four members of the Lyman and Balmer series in temperature-density
ranges appropriate to stellar atmospheres. These tables include the convolu-
tion of the Stark profile with a thermal velocity distribution of the hydrogen
atoms. Figure 9-2 compares the profile of Ho from the unified theory
(including both electron and ion broadening) with the quasi-static theory for
ions only. For higher series members one may use an approximate theory
(262), after corrections to some of the matrix elements are made (265) [see
0.1 also (45)].
In astrophysical applications, the hydrogen lines are significantly affected
by broadening mechanisms other than the Stark effect. The core of the line
is dominated by Doppler broadening. The effects of radiation and resonance
damping may be important in the wings at low electron densities. Assuming
these mechanisms are all uncorrelated, we may account for their combined
S(cx) 0.ot effects by a convolution procedure. Folding the Doppler profile with the
Lorentz profiles from radiation and resonance damping gives a Voigt profile,
H(a, v),wherea = (r,.d + r, •• )/41tl\v 0,andv = (£\v/£\v 0).ThisVoigtprofile
is then convolved (numerically) with the Stark profile S(cx), yielding the
cross-section
0)

0.001
cx,.(L\v) = (nte 2
/mc)f s~ao S*(l\v + V l\vo)H(a, v) dv (9-146)
.......

where S* is S(cx) converted to frequency units..

HYDROGENIC IONS

Hydrogenic ions (of charge Z) have Stark patterns identical to those


of hydrogen, though the energies are, of course, different. The profile from
ion-broadening alone is

(9-147)

where SQS is the quasi-static hydrogen profile given in (634). Note that on
a wavelength scale the ion lines are narrower by a factor of l/Z 5• Indeed,
10-• ::----:-:-:::-:':::---:1:----.L-~--1..----1-..J..._.J
because the Stark widths decrease for ions while the radiative transition
0.ot 0.Q3 0.05 0.1 0.3 0.5 1.0 3.0 s.o 10.0 probabilities increase, there comes a point where Stark broadening can be
"' neglected compared to r,ad•
FIGURE 9-2 The effects of electron broadening for hydrogenic ions are similar to those
1 3
Stark profiles for Ha at n,"" 3.16 x 10 • cm - , T"" 10• °K. Dotted mrve: Holtsmark for hydrogen, though the expression for <1>0 b changes because now the
profile for ion broadening only (634). Daslie,I mrve: Profile for quasi-stalic ion broadening
and impact brondeninii hy electrons (650). Solid c1tr1>e: Electron and ion-broadened profile
306 The Line Absorption Profile 9-5 Quantum Theory of Line Broadening 307

perturber moves on a hyperbola around the positively charged ion instead OTHER LIGHT ELEMENTS
of a straight-line path. Early calculations of the broadening of the lines He II
Electron collisions broaden the lines of other elements observed in
).3203, ).4686 (prominent in O-star spectra) are given in (272), and mucb-
stellar spectra. Widths and shifts for these lines may be calculated using
improved calculations for He II .U.256, 304, 1085, 1216, 1640, 4686, and 3203
techniques similar to those employed for He I [see (268, §§Il.3c, II.3d)].
are given in (354; 355). Unified theory (258; 260) calculations for He II 2304
Extensive results for neutral atoms are given in (264, 454-527; 268, Appendix
are given in (259). Unfortunately, precise calculations for the astrophysically
IV; 84). For charged ions, the Coulomb interaction between radiator and
important Pickering series lines (4 ... n) (e.g., 2210124, 5412, 4542, 4200, etc.)
perturber implies a hyperbolic path for the latter (110; 111), and line-widths
are not yet available, and only an approximate theory exists (262; 265; 45).
calculated allowing for this are substantially larger than those calculated
with a straight-line path. Extensive results of detailed.calculations for ions
NEUTRAL HELIUM LINES are given in (268, Appendix V; 111; 548; 167; 549; 550). Convenient approx-
imate formulae for estimating Stark-widths are given in (267; 183; 549).
The lines of He I are prominent in the spectra of B-stars. Here the
electron impact broadening and quasi-static ion broadening act by quadratic
Stark effect, and for isolated lines yield profiles of the form

/(A ) _ (~) r,., W(F) dF (9-148)


w - n Jo (Aw - d - C4 F 2/e 2 ) 2 + w2
where W(F) is the probability of a field of strength F, w is the electron-
impact width of the line, and d is the line shift. Note that because F enters
as F 2 , the ionic fields always skew the line components in one direction and
thus lead to an asymmetric profile.
Explicit expressions for wand dare given in (269; 264, 81-86; 268, §II.3C<X;
180) for isolated lines; tables allowing the calculation of the profile in terms,
of convenient dimensionless units are given in (269). Detailed numerical CJ)
results for wand d for several lines are given in (269), and improved results I\)
are given in (180) and (67). A much more interesting (and difficult) case is I
presented by the diffuse series lines (2 3 P-n 3D), (2 1 P-n 1 D) for n ~ 4. As was
first recognized by Struve (614;615), the (2 3 P-4 3 D) He I 24471 line shows a
"forbidden" component (2 3 P-4 3 F) at 24470. The other diffuse series lines also
show these components which arise from the mixing of the (3 D, 3 F) or
(1 D, 1 F) states in the presence of the electric fields in the plasma. As the
diffuse-series lines are among the best observed in stellar spectra, a reliable
theory for them is of great interest. The first attempts at constructing such
broadening theories (66; 245; 266) were not too successful, for they gave
"forbidden" components that were too narrow and too intense. Compared
to observed stellar profiles the theoretical predictions showed too much
contrast between the "forbidden" absorption component and the gap between
it and the allowed component; the theory also disagreed with experimental
measurements (122; 123; 124). A comprehensive new theory was then de-
veloped for the He I 24471 (68) and ).4921 (2 1 P-4 1D) lines (69); predictions
based on this theory are in excellent agreement with observed stellar spectra
(438; 439).
JO./ Characterization of the Problem 309

photons absorbed are simply scattered so that the excited electron returns

10 directly to its original lower level, and normally it is assumed that the scat-
tering is isotropic and coherent (actually complete redistribution is a much
better approximation). Then the scattering contribution to the emission
coefficient is
ri: = (1 - e)x,<f,J, (10-1)

Classical Treatments of where x, = (1te 2 /mc).fii n, - (g 1/g1)n1] (10-2)

is the line absorption coefficient in the transition between levels i and j. The
Line Transfer remaining fraction e of the photons is assumed to be destroyed and converted
to thermal energy by various processes (cf. §2-1). One then argues that this
loss into the thermal pool must be balanced exactly by thermal emission,
which contributes to the total emission coefficient an amount

(10-3)

In the limit of strict LTE, e = 1, and all the emission is thermal.


In addition, there are contributions to the opacity and emissivity from
continuum thermal processes and electron scattering. Thus the transfer
equation is

µ(81 Jaz) = -(Kc + u + x,<f,.)l, + KcB, + uJ, + ex,<f,,B, + (1 - e)x,<f,J,


(10-4)
0)

In this chapter we discuss some of the early approaches to the line-formation We have ignored the v-dependence of Kc and a because most lines are so w
problem; these provide background for the more modern treatments to be narrow that these coefficients vary only negligibly over the line in comparison I
presented in following chapters. A fuller discussion of these older methods with the swift variation of <f,,. lfwe write dt, =
-(Kc + u + x,<f,,) dz, and let
and of their application to stellar spectra can be found in (684, Chap. 7; 638, p = u/(Kc + u), and
Chaps. 15-17; 1S, Chaps. 12-16; 11, Chap. 8; 2S6, Chaps. 14-16). It is (10-5)
essential to be familiar with such methods because of the large body of we have
literature based upon them. More important, one must understand the
physical basis of the classical treatments in order to evaluate the reliability µ(ol,/or:,) = I,. - {[(I - p) + i-:/1,.]B, + [p + (l - e)/J,]J.}/(1 + /J,)
of spectroscopic diagnostics derived from them, and to realize the conceptual (10-6)
differences inherent in recent work. Or, defining
J., = [(l - p) + e/J,]/(1 + fl.) (10-7)

l 0-1 Characterization of the Problem the transfer equation becomes

In the usual classical approach one notes, at the outset, the existence of two (10-8)
different line-formation processes: scattering and absorption. We have dis-
cussed these two categories in Chapter 2, and described there the physical E. A. Milne and A. S. Eddington developed equation (10-8) as an approxima-
differences between them. It is usually supposed that a fraction (I - e) of the tion to the line-transfer problem, and customarily it bears their name.
/0-2 The Milne-Eddington Model 311
310 Classical Treatments of Line Transfer
Under these conditions an exact solution may be obtained (158) but this
Excellent discussions of the physical implications of this equation have been differs only slightly from the approximate solution derived below.
given by Milne (416, p. 169ff) and StrBmgren (613). Taking the zero-order moment of equation (10-8) we find
From a physical point of view, equation (10-8) provides a rather severe
idealization of line formation and may be criticized on several counts. (1) (10-11)
Line-scattering is not, in fact, coherent. This inadequacy is overcome if we
write, instead of equation (10-8), Aside from the definition of l., equation (10-11) is the same as equation (6-4),
and fro_m t~e a~alysis in §6-1 we know that the solution, in the Eddington

(of,) = I, - f (, approx1mat1on, 1s
'B pl, (1 - e)/3, )J d, (109)
µ OT, 11.y y - 1 + /3. - (1 + /3,)<J,. R V' V y' V -
J, =a+ p,t, + (p, - .Jja) exp[ -(3l.)tt,]/[J3 + (~.)½] (10-12)
where R(v', v) is an appropriate redistribution function (cf. §2-1 and Chap. 13). and the emergent flux is
(2) To solve equations (10-8) or (10-9), both the parameter & and the occupa-
tion numbers n1 and .n1 must be known. In classical treatments it is often
( 10-13)
assumed that LTE holds, so that & = 1 and n, = nr, n1 = nj. It must be
stressed, however, that this merely an assumption and, as will be seen in
Chapters 11 and 12, the assumption is often unjustified and may yield results Equation (10-12) shows that thermalization (J, .... B,) occurs only at depths
seriously in error. In many treatments, an ad hoc value is chosen for &, but of the order of l, -t. Recalling the definition of l., we see that this depth is
the occupation numbers are still assumed to have their LTE values. Such an (1 - p)-t in the continuum (/3, = 0), and e-t in a strong line (/3, .... co). In
approach is internally inconsistent, for when scattering occurs in a line, the both cases thermalization occurs at a depth p-t, where p is the probability
level-populations depend upon the radiation field via ... the equations of that a photon is destroyed and converted to thermal energy each time it
statistical equilibrium. (3) An equation of the form of(l0-8) was derived by interacts with the material; these results are compatible with the random-
Milne for a strict two-level atom, and his analysis yields a unique (and correct) walk arguments given in Chapter 6. Note that these results apply only for
value for the parameter & (416, pp. 172-178). However, as mentioned abo:ve, coherent scattering (cf. Chapter 11).
coherent scattering is not an accurate approximation. More importa,nt, Equation (10-13) may be used to compute the profile of a line in a stellar
CJ)
analysis @fthe equations of statistical equilibrium for general (i.e., multilevel) atmosphere. In the continuum, {3, = 0; then l, = (l - p), and the continuum
atomic models shows (cf. Chapters 11 and 12) that other kinds of terms appear flux is ~
in the source function; these may depend upon the radiation fields in other 1 I
H,(0) = [b + a ✓3(1 - p)]/[l + (1 - p)¼] (10-14)
transitions (continua and lines), and thus in principle couple all the lines in 3
the spectrum together. In short, both equations (10-8) and (10-9) are seriously
incomplete from a physical point of view, and this should be borne in mind Thus the residua/flux in the line is, by equation (8-2),
during the discussion that follows in this chapt~r.
R,. = [Pv + (3A.-ta] [ 1 + (1 - p)¼ ] (10-15)
l + l, b + a ✓ 3(1 - p)
10-2 The Milne-Eddington Model
There are four important results that can be derived from this classical theory,
DEFINITION
which provided the conceptual orientation of much of the early work on line
formation. Let us now examine these briefly.
Let us now consider the Milne-Eddington equation [equation (10-8)]
under the simplifying assumptions that l., &, and p are all constant with SCATTERING LINES
depth, and that the Planck function B, is a linear function on the continuum
optical depth scale t; i.e., Consider the case where p = 0, so that there is no scattering in the
continuum. Further, suppose that & = 0, so that there is pure scattering in the
B, = a + bt = a + [bt,/(1 + /3,)] = a + p,1:, (10-10)
312 Classical Treatments of Line Transfer 10-2 The Milne-Eddington Model 313

lines. Then},. = (1 + /JS I and the residual flux in the line is Using the Eddington-approximation result for the grey temperature distribu-
tion, namely T 4 = T 0 4 (1 + fr), it is easy to show that

R,=2L :/i. +a(i :/J.Y][(1 + J 1 ~p.)(J3a+b)J


1

(10-16)
(10-20)
If we consider the case of a very strong line and take the limit /J,. -+ oo, we
obtain R, = H ,(0)/H ,(0) = 0, which shows that the core of a very strong line where (10-21)
formed by scattering can be completely dark; this result is in contrast to the
case of a line formed by absorption, as is shown below. and u0 = (hv/kT0 ). Thus the parameters in equation (10-10) are: a = B0 ,
3
Exercise /0-1: Writing p, = p0 H(a, v), where H denotes the usual Voigt function, b = g X 0B0('i?./K) (10-22)
plot residual flux profiles from equation (10-16) for scattering lines with Po = 1, 10,
JOO, 1000, 104. and (b/a) = 1, 2, and 3, assuming a= 10· 3 •
Exercise /0-2: In the Schuster-Schwarzschlld model, the lines are assumed to be 3
and Pv = g X oBo(ic/K)/(1 + /J,) (10-23)
confined to a finite layer (the "reversing layer") of thickness t., illuminated from
below by an incident intensity 10 • In the reversing layer the continuum opacity
is zero, and the lines are pure scatterers. Using the two-stream approximation where ic and K are the mean opacity, and the monochromatic continuum
(I = 1+ for 0 ~ µ ~ I; I = r for -1 ~ µ ~ 0; and µ = ±½ in the transfer opacity at the line wavelength, respectively. Then equation (10-18) becomes
equation) show that (a) H. = ½U.+ - I,-)= constant = ½lo/(1 + t,), and
(b) J,(c.) = H,(1 + 21.), 0 ~ r. ~ i •. (10-24)

Exercise 10-4: Derive equations (10-20) through (10-23).


ABSORPTION LINES

Again assume p = 0, but now set & = 1 (LTE in the line); then J.. = 1, and For the sun, TO ~ 4800°K from a grey model, and if we choose). = 5000 A,
then u0 ~ 6, X 0 ~ 6, and K ~ K, so that equation (10-24) predicts .......
(10-1.7)

In this case, as /J. --+ oo, the flux in the line does not go to zero, but approaches R0 = ( 1 + 43 J3 )-1 = 0.44 (10-25)
CJ)
01
a finite value I
R0 = R,(& = 1, /1.--+ oo) = [1 + (b/J3a)]· 1 (10-18) This value is in fair agreement with the depths of many of the stronger lines
observed in that region of the solar spectrum. Some lines, however, are much
This is to be expected, for as {J, --+ oo, only the surface layers of the star are deeper, particularly resonance lines such as the sodium D-lines; this fact led
seen, and the emergent flux is then determined by the surface val~e of the to the conceptual identification of resonance lines as "scattering" lines and
Planck function, which will be nonzero. In contrast, in the scattermg case, subordinate lines (e.g., Ha.) as "absorption" lines. It was believed that the
photons are constantly diverted out of the pencil of radiation, and in the central intensities of the latter group reflected information about the surface
limit /J. --+ oo, none survive to emerge at the surface. temperature of the atmosphere. Such a division of lines into two groups is
intuitively not unreasonable, for we expect that in a resonance line the most
Exercise /0-3: Repeat exercise 10-1 for an absorption line, using equation (10-17). probable route of exit from the upper state is, in fact, a direct decay to the
lower state. In contrast, for subordinate lines a large number of possibilities
It is convenient to re-express equations (10-17) and (10-18) in terms of the will, in general, exist, and the photons may effectively be removed from the
Planck function and its gradient. Assume that on a mean optical depth scale, line and destroyed. It must be emphasized, however, that this characterization
is only schematic and often does violence to the actual physics of line-
(10-19) formation. For example, we shall find in Chapter 11 that the boundary value
314 Classical Treatments of Line Transfer 315

of the source function of the Hr,, line (a subordinate line) has essentially TAIILE 10-1
nothing to do with the temperature of the outer atmospheric layers. Cenrer-to-Limb Variation in the Milne-Eddington Model

a,(µ)
CENTER-TO-LIMB VARIATION
A,
The specific intensity emergent at frequency v, on the disk at an angle P. µ=I 0.5 0.3 0 (Flux)
0 = cos - 1 µ from disk center, is given by 1.0 0.DI 0.007 0.006 0.005 0.000 0.006
0.1 0.068 0.055 0.043 0.000 0.058
1,(0, µ) =f a:,
0
S,(t,)e-<•.Mµ- 1 dt, 1.0
10.0
0.375
0.682
0.300
0.545
0.237
0.431
0.000
0.000
0.317
0.516
100.0 0.743 0.594 0.469 0.000 0.628
=f a:,
0
[B. + (1 - ,1,,)(J. - B,)] exp(-t,/µ)µ- 1 dt, (10-26) 00 0.750 0.600 0.474 0.000 0.634

0.DI 0.007 0.005 0.004 -0.002 0.006


where the form of s. has been taken from equation (10-8). Substituting from 0.1 0.066 0.051 0.037 -0.019 0.054
0.3 1.0 0.378 0.306 0.246 0.026 0.322
equation (10-12) we find 10.0 0.723 0.633 0.565 0.334 0.653
100.0 0.798 0.713 0.648 0.438 0.731
00 0.808 0.723 0.659 0.452 0.741
1,(0, µ) = (a + p,µ) + [ (p, - Jja)(l - ,l,) ] (10-27)
Jj(l + A)(l + Jr[,µ) 0.DI 0.007 0.005 0.004 -0.003 0.006
0.1 0.066 0.049 0.035 -0.024 0.053
In the continuum, {J. = 0, and again taking p = 0, 0.1 1.0
10.0
0.379
0.751
0.308
0.687
0.250
0.639
0.035
0.483
0.324
0.700
100.0 0.847 0.799 0.765 0.658 0.809
I.(0, µ) = a + bµ (10-28) 00 0.860 0.815 0.783 0.684 0.824

The residual intensity r,(µ) =


/,(0, µ)/1.(0, µ) and the absorption depth 0.01
0.1
0.007
0.066
0.005
0.049
0.003
0.034 ·
-0.004
-0.027
0.006
0.053
.......
a,(µ) = 1 -:- r,(µ) follow immediately from equations (10-27) and (10-28).' 0)
Consider first a pure absorption line, for which s = 1 and ,1,, =
1. Then
0.0 1.0
10.0
100.0
0.379
0.778
0.931
0.310
0.732
0.920
0.252
0.698
0.912
0.039
0.589
0.885
0.325
0.742
0.922
0)
00 1.000 1.000 1.000 1.000 1.000
I
r,(µ) = [1 + (b/a)µ/(1 + {J,)]/[1 + (b/a)µ] (10-29)

Here we see that as the limb is approached (i.e., asµ-+ 0), /,-+ 1., and the In the solar spectrum, some lines do weaken towards the limb while others
line vanishes. This is compatible with the physical picture sketched above for do not, or weaken only slightly. This observed behavior again led to the
as the limb is reached, only the very surface can be seen at any freque~cy. classification oflines into "absorption" or "scattering" categories, though in
:ore = 1, the same source function (the ~urface value of the Planck function) some cases the categorization made on the basis of limb-darkening was in
ts observed throughout the profile, and the contrast between line and con- conflict with that based on central intensities. Furthermore, some studies
tinuum disappears. On the other hand, if s = O (a pure scattering line), [e.g., that by Houtgast, reviewed in (596)] have shown that neither category
;.,, = (1 + /J.)- 1, and taking the limit as /J.-+ oo, equation (10-27) reduces is adequate because in some lines the effects of noncoherent scattering
to I ,(0, µ) = 0 for all µ's; thus the cores of pure scattering lines always dominate. In short, the approach described here is quite schematic, and one
remain dark, even at the limb, and there is a clear distinction between the must recognize that it simply does not contain much of the essential physics.
center-to-limb behavior of absorption and scattering lines. A few representa-
tive values of the variation of line-depth with disk-position are given in SCHUSTER MECHANISM
Table 10-1 for several values of sand /J,. In this table (b/a) = 3; negative
values imply emission, and are an artifact of the approximations made in In the above discussion, the continuum has been assumed to be purely
solving the transfer equation. thermal. When continuum scattering is taken into account, several interesting
316 Classical Treatments of line Transfer /0-3 The Theoretical Curve of Growth 317

effects are found; these were first discussed by Schuster in one of the funda- To begin with, we assume that the line formation occurs in a layer to
mental papers of radiative transfer theory (562). To emphasize the role of which we can assign a unique temperature and electron pressure. This
continuum scattering, set p = 1 (a realistic value for, say, O-stars); then assumption is normally valid in laboratory work, but in the stellar-atmo-
i.• = e{J.f(l + /J.) and equation (10-15) becomes spheres situation there are usually strong gradients in both temperature and
pressure. Thus the choice of an appropriate temperature and pressure is

R.
1
= [ 1 + /J. + b
(a) (13e{J.
+ /J.
)t] [l + (1 e{J,+ /J.)t]-1 (10-30)
difficult and, at best, can refer only to some ill-defined mean value in the line-
forming region. Having made this choice, we can compute (assuming LTE)
the populations of the atomic levels and, hence, the continuous opacity Kc
First, consider the case with e = O; then R, = 1/(1 + fJ.), and the line is a and the line opacity in a transition (i -+ j),
pure absorption feature. This occurs because both the lines and continuum
are formed by scattering, and the scattering length in the line is larger. On (10-31)
the other hand, suppose e = 1; then it is obvious from equation (10-30) that
The line profile is assumed to be given by the Voigt function [ equation (9-34)]
in the line core, as /J. -+ ao, R, -+ (/ja/2b). Thus the line can appear in
absorption or emission depending on the ratio (a/b); the more shallow the
temperature gradient (i.e., the smaller the value of b), the brighter is the line
H(a, v) = (a/rr) J~"' e-)·2[(v - y)2 + a2 ] - 1 dy
compared to the continuum. The reason for this is that because e = 1, the
source function in the line is everywhere equal to the thermal value while,
=
where v /:J.v/1:J.vo, a =
f/(4rr /:J.vo), and l:J.vo = veo/c; here eo
is the most
probable velocity of the atoms in the material. We then write X1J(v) =
in the continuum, scattering causes J. (and thus S,) to drop below B,.
xoH(a, v) where Xo = XiJ/(rrt /:J.v 0 ). We assume that both l:J.v 0 and the param-
If(a/b) is greater than the critical value (a/b) = 2/Jj, the line is in emission
eter a are fixed in the region of line-formation, and employ the Milne--
for all values of fJ. (i.e., at all frequencies in the profile). If (a/b) just equals
Eddington model in which the ratio {J. = X1J(v)/Kc is independent of depth.
the critical value, then both the extreme wing (/J. -+ 0) and -the very core of
(The assumption of depth-independent {J. is actually a fairly good approxi-
the line (/J. -+ ao) lie at the level of the continuum, and all other ,points in
mation for some spectrum lines; e.g., Mg II 24481 and Si II lA.4128, 4131.
the profile are in emission. If (a/b) is less than the critical value the line may
Indeed for these lines both Xo and K, may vary over orders of magnitude in
have weak emission wings and a central reversal into an absorption core, --1.
the atmosphere while their ratio remains nearly constant.) Finally, we assume
lf(a/b) < 1//j, the line is everywhere in absorption. the line is formed by absorption processes in LTE; the flux is then easily 0)
This interplay of scattering and absorption, which gives rise to either -..,J
computed as
absorption or emission lines depending on the temperature gradient, is I
known as the Schuster mechanism; a thorough discussion of the various
possible cases was given by Schuster himself (562). From time to time it has (10-32)
been suggested that this mechanism could be responsible for emission lines
observed in certain early-type spectra, but on the basis of recent critical As before, we adopt B,[T(r)] = B0 + B 1 r, where r is the continuum
discussions (238; 280) this seems unlikely. optical depth. Then, having taken {J,. to be constant with depth, we have

10-3 The Theoretical Curve of Growth


F,. =2 f0
00
(B 0 + B1r)E 2 [(1 + /J,)r](l + /J,.) dr
2 (10-33)
Using equation (10-15) one could, in principle, compute the profile of a = B0 + 3 Bi/(1 + /J..)
spectrum line and, by integration over frequency, determine its equivalent
width. Such a procedure is laborious, however, and requires the use of a From equation (10-33) the continuum flux is clearly F,. = B0 + jBi, so that
computer. It is instructive, therefore, to consider a simple model that allows the line-depth in the flux profile is
the equivalent width to be computed analytically. In this way we can con-
struct what is known as a curve of growth, which gives the equivalent width A,. = [/J./(1 + /J.)] [ 1 + 23 (B 0 /Bi) ]-I (10-34)
directly in terms of the number of absorbing atoms that produce the line.
318 Classical Treatments of Line Transfer
319
It i~ co~venient to wo~k in teT?1s of the parameter A 0 , the central depth of
an mfimtely opaque !me. Takmg the limit of equation (10-34) as P. -+ oo
we find 1.0

(10-35)
0.9
which is esse?tially the result stated in equation (10-18). Then equation (10-34)
may be rewntten 0.8
A, = AoP./(1 + P.) (10-36) R,

The equivalent width may now be computed by integration of the line-depth 0.7
over frequency to obtain
81 3
0.6 80 =2
W. = Jo"' A, dv = 2A 0 Av0 Jo"' P(v)[l + /J(v)J- 1 dv (10-37) a= 10· 3
where it has been assumed that the line is symmetrical about its center. But 0.5
P(v) = (xo/1e,)H(a, v) == PoH(a, v), and if we define the reduced equivalent 0 2 3 4 8 9 10
width w• == W/(2A 0 Av0 ), we have
V

FIGURE 10-1
w•(a, Po) =, Jo"' PoH(a, v)[l + PoH(a, v)]- 1 dv (10-38) Development of a spectrum line with increasing number or atoms along the line of sight. The
line is assumed to be formed in pure absorption. For Po~ I, the line strength is directly
Before we attempt to evaluate this integral, let us consid~r qualitatively proportional to the number of absorbers. For 30 :6 Po :6 103 the line is saturated, but the
how the line develops as more and more atoms absorb. At the start, with wings have not yet begun to develop. For Po ;e 104 the line wings are strong and contribute
most of the equivalent width.
only a few absorbers, each will be able to remove photons from the radiation
field, and the line strength should be proportional to the number of absorbing ,
atoms. Only the Doppler core (where the opacity is highest) will contribute . Consider now the contribution from the core only, and write /J( v) = /J 0 e- •2,
to the line strength; the line wings will be transparent and will not reduce assuming that Po < 1. Then equation (10-38) becomes 0)
th~ emitted flux. As yet more atoms absorb in the line, the core, at some CD
~01~~• becomes. completely ~paque, and the intensity there reaches its W* = Po fo"' e_",[1 + Poe_•,]-• dv I
hm1tmg value given by equation (10-35). Now, essentially all the photons
that can be absorbed in the core already have been, and so long as the line-
wings remain transparent, the addition of more .absorbers does little to = /Jo fo"' e-•lo - /Joe-•2 + .. ·) dv (10-39)
increase the equivalent width of the line, which is said to be saturated.
Finally, when enough absorbers are present, the opacity in the wings be-
or (10-40)
comes appreciable, and the equivalent width again increases as the contri-
bution of the line wings grows. These different behaviors are shown in
Figure 10-1. On the basis of the above discussion, we see that there are three For small values of Po (weak lines) the linear tenn dominates, and the equi-
?as~~lly distinct regimes in the curve of growth; these will be treated valent width of the line is directly proportional to the number of absorbers
md1V1dually. present. This is known as the linear part of the curve of growth. Note that
The Voig~.frofile_ can -~e repre~e~ted schematically [equation (9-45)] as Po varies as Av0 - 1, as does w•, hence the equivalent width Wis independent
~
H(a, ~) e + n tav ; here 1t 1s assumed that the first term applies of Av 0 on the linear part of the curve.
only m the core for v ~ v•, and the second only in the wings for v ~ v• In the saturation part of the curve of growth, Po is so large that the line
where v• is chosen as the transition point where the two tenns are equal'. core has reached its limiting depth, but not yet large enough that the line
wings contribute to the equivalent width. Again P(v) = Poe-•2, and if we
320 Classical Treatments of line Transfer 321

(10-41) log-
w
61•0

where we have set Po = e"'. This integral may be rewritten [see (160, 389)] as 0

2 w• -_ Jo
f« -t d + f"" [ex(l + t)]* dt
+ e"''
f« [cx(l - t)]* dt
1 + e«'
u u ex Jo - ex Jo (10-42) -1
1
Following Sommerfeld, one may replace the upper limit in the third integral -2
-1 0 3 4 5 6
by oo (because ex » 1) and expand [ex(l + t)]* as a power series in t around
t = 0. The results may be written in closed form in terms of the Riemann
zeta function, and one obtains finally the asymptotic expression FIGURE )0-2
Curves or growth for pure absorption lines. Note that the larger the
w• ~ Jin Po{l - [1t 2/24(1n P0 ) 2] - [71t 4 /384(ln P0 ) 4 ] - .. ·} (10-43) value or a, the sooner the square-root part or the curve rises away
from the Hat part.
The above expansion is only semiconvergent and must always be truncated
after a finit~ number of terms; in practice, the series is useful for Po ;;::; 55.
From equation (10-43) we can s~e clearly !hat on the satura{_ion, or fiat, part and hence the sooner the damping part of the curve rises away from the flat
of the curve of growth, the equivalent width grows extremely slowly with part. Numerous curves of growth have been computed by various authors.
increasing numbers of absorbers, namely w• oc Jin p0 • The we~k depen- A particularly useful set has been published by Wrubel for a wide range of
dence of w• on Po implies that W on this part of the curve is essentially the temperature-distribution parameters B0 and Bi, and under different
proportional to Av0 • It is easy to understand why this should be so: the assumptions concerning the atmospheric model (i.e., Milne-Eddington or
Schuster-Schwarzschild) and the transfer problem (i.e., absorption or CJ)
depth of the line profile is fixed at A0 , hence the integrated absorption must'
be proportional to the linewidth; i.e., to Av0 (cf. Figure 10-1). Similarly the scattering lines); see (687; 688; 689). co
deJ?Cn~ence of W o~ Po.can be un?erstood by recognizing that to produce I
a s1gmficant depression m the contmuum, the optical depth in the line must
exceed unity before the continuum reaches unit optical depth. This occurs 10-4 The Empirical Curve of Growth
at_ frequencies v < v0 where p0 e-•02 ~ 1; clearly v0 , which measures the
w1d~h of the dark core (and hence determines W*), varies as Jin p0 • The curve of growth has long been one of the astronomer's favorite tools
Fmally, for very large numbers ofabsorbers, the line wings become opaque for performing an analysis of a stellar atmosphere, and the literature of the
enough to provide the dominant contribution to the equivalent width. Here various applications of this approach is enormous. The reasons for the
we adopt H(a, v) ~
a/(1t*v 2 ), and writing C = p0 a1t-t we find from equation popularity of the curve of growth are that it provides estimates of several
(10-38) key parameters with speed and ease, and that it makes use of equivalent

w• = f0"" (1
widths alone, which can be well-determined observationally even for faint
+ v2/C)- 1 dv = ~ 1t-/C = ~ (1taP0 )* (10-44) stars where profiles would be impossible to measure accurately.
As described in §10-3, the theoretical curve of growth gives log(W,,/Av 0 ) =
Thus W oc Po*, giving rise to what is called the damping or square-root part log(~/A,1. 0 ) = log(W;.c/,1.e 0 ) as a function of log Po, Here e 0 is the total
of the curve of growth. Again, note that both a and Po each contain a factor random velocity of the atoms forming the line, and
of Avo; hence Wis independent of Av 0 on this part of the curve.
The entire curve of growth is shown in Figure 10-2. Notice that the larger p = Xo = ..!:....!!.__ *k(1 - e -hv/kT)
t 2f ) n 11 t 2) (f ,1.) ( niJk
=(~ •) (l0-45)
the value of the damping parameter a, the sooner the wings dominate w, O Kc ( me AVo k,(l - e-hv/kT) me eok,
10-4 The Empirical Curve of Growth 323
322 Classical Treatments of Line Transfer

Here k., the continuum opacity uncorrected for stimulated emissions, has to the theoretical is normally required. From this fitting procedure we can
been introduced for convenience (and continuum scattering has been deduce three essential bits of information. .
(1) The ordinate of the empirical curve is log(W.i/l), while that of the
ignored). The population n1jt, of excitation-state i, of ionization-stage J, of
chemical species k, is assumed to be given by the Boltzmann-Saha equations theoretical curve is log(W.i/~).. 0 ) = Jog(W.i/l) - log(eoM· Thus when the
as described in §5-1. two curves are superimposed, the difference in the ordinates yields log(eo~c),
In a stellar spectrum one can usually observe lines from several multiplets and thus the velocity parameter e0 • The value derived can be compared with
the most probable thermal velocity at the excitation temperature T.,.,
of a given ion, and one or more ionization stages of a given element, for each
ly ,. - (2kT /Am")½ where A is the atomic weight of the element.
of several elements. Let us first consider only lines from a definite ion of a ds
single element. Because of the factor exp( - XtJt! kT) in the Boltzmann ex-
nam C '9therm - ...
It is usually found that eo h f th
as inferred from t e curve o grow excee
citation formula, each multiplet has its own curve of growth. In view of
e sometimes by a large factor. To explain this dilterence, it has been
c~;~~~ary to postulate the existence of additional nont~ermal motions ~f
equation (5-8) we may write
the stellar material, which are usually referred to as m1croturbulence. It 1s
log /Jo = log(g,1tf l) - Bx,1t + log C1t (10-46) assumed that these motions occur on scales that are small c~~pared to a
photon mean-free-path and hence constitute, in e~ect, an add1t1onal_ sour~
where 0 = 5040/T, x is expressed in eV, and of line-broadening. If these small-scale mass motions have a Gaussian dis-
tribution around some most probable speed e~urb, then
(10-47)
eo = [(2kT...IArntt) + e~urbJ½ (10-48)
Clearly the value of Po will be affected by the choice for the temperature (0)
in the atmosphere, and will be different for each multiplet (because of "Turbulent" velocities have been derived for many stars by this method of
differences in x,1t), and for each line (because of differences in .f-values). analysis; the most dramatic results are obtained for supergiants. where
To construct an empirical curve of growth, we plot the v'alue o(log(W,,/l) velocities in excess of the sound speed in the gas have been obtained. It
versus log(g/l) for each line. Now if we assume that there is a unique relation should be realized, however, that such diagnoses are not _on entirely firm
between Po and ~. we attempt to force all points for the different lines to ground because the introduction of a velocity field drast1cal!y affects. t?e
define a single curve as nearly as possible. To do this, we adjust 0, and choose details of line formation (cf. Chapter 14). Indeed, w~th ~he high veloetties
the value that minimizes the scatter around a mean curve. This value is sometimes derived, one must inquire whether the exc1t~t1on state of.the gas --..J
called the ·excitation temperature, 0exc, and is considered to be the charac- is affected by interchange of energy between mass motions and t~e i~temal 0
teristic temperature of the line-forming region. It should be realized that it energy of the material. Very little work has been done on ~his difficult I
may not always be possible to derive a meaningful excitation temperature, problem, and our knowledge about small-scale stella~ ~eloc1t~ fi~lds re-
for different lines will actually be formed in different regions of the atmo- mains rudimentary, almost to the point of merely recog~i~mg their ex1ste~ce.
sphere. For example, one would expect that lines from levels with high (2) The difference between the abscissae of the empmcal and theoretical
excitation potentials will be formed deeper in the atmosphere where tem- curves of growth yields
peratures are higher. Similarly, we would expect the average excitation
temperature to be higher for higher stages of ionization. Such refinements Jog c = log Po - [log(gf )..) - e•••x] (10-49)
can be taken into account by more elaborate calculations that use model To proceed further, we need an estimate of the electron density. This is
atmospheres, but they are normally ignored in curve-of-growth work. derived from a theoretical model of the atmosphere. Once we know n., we
There are other complications that arise in practice: there are errors both
in the values of ~ and /, and these introduce scatter into the curve. Also,
may compute kc, and as we already know eo,
~e find N 1t, the _numbe~ of
atoms of chemical species kin ionization stage j directly from C, via equation
lines arising from only a limited range of excitation potentials may be (10-47). Then by use of the Saha ionizati~n fo~mula_we_ ca~ convert NJk to
observable, and Om may not be very well determined over this short baseline. N the number of atoms of chemical species km all 1omzat1on states._More
Once the empirical curve, corrected for excitation effects, has been estab- pr~cisely, we obtain the abundance CXt = (Nt/Ntt), as kc is usually dom_mated
lished, it may be compared with theoretical curves. To superimpose the two by hydrogen or H-, and thus is proportional to NH. In short, the horizontal
curves, a shift (both in abscissa and ordinate) of the empirical curve relative
324 Classical Treatments of Line Transfer

shift between the two curves of growth yields the abundance of the element.
This method of analysis has been applied to a wide variety of stars, and has
shown that certain stars (e.g., Population II stars and peculiar stars) have
abundances that differ markedly from solar values, which, in turn, are fairly
typical of Population I stars. The uncertainties in a curve of growth abun-
dance analysis should be clear: it has been assumed that the Saha-Boltzmann
relations are valid; depth variations of the relevant parameters (called strati- ....q
fication effects) have been ignored; and a schematic solution of the transfer
problem has been employed. The accuracy of the results is, therefore, limited.
(3) The horizontal and vertical shifts described above make use of the
linear and flat parts of the curve respectively. For a given set of theoretical
curves, a comparison between the observed and computed damping parts
determines the value of a = r/(41t Av 0 ), and hence r; this in turn may be
compared with the value predicted by line-broadening theory.
Let us now turn to a brief discussion of a few typical results. The solar
spectrum has been analyzed extensively with curve-of-growth techniques;
one of the most outstanding early treatments was by H. N. Russell (542),
who derived element abundances (from eye-estimates of line-strengths!) that
are in remarkably good agreement with current estimates. A very interesting
study using curves of growth was carried out by K. 0. W.right (686) who
analyzed the Sun and three other solar-type stars. This study is by no means
the most recent available [see, e.g., (182)], but it is a classic example of the
procedure, so we shall consider it here. From an extensive set of equivalent~ ___._
width measurements and laboratory .f-values, an empirical curve of growth q ---J
was constructed using lines of Fe I and Ti I. In all, some 75 lines of Fe I I
___._
arising from'states with 0 ~ x ~ 1.6 eV, and 137 lines ofTi I with 0 ~ x ~
2.5 eV, were used to obtain the curve shown in Figure 10-3. Slightly different
excitation temperatures were found for the two atoms, namely T ... =
4850°K ± 150°K for Fe I and T ... = 4550°K ± 150°K for Ti I. These
values are about what we would expect for a radiative-equilibrium model
atmosphere. As can be seen, the curve is well defined, though it is true that
the linear part is defined mainly by Ti I lines and the damping part by Fe I
q
lines. It would be more satisfactory if lines of each atom were found in both ....
the linear and damping portions of the curve. I

The vertical shift of the empirical curve, relative to a theoretical curve


e
by Menzel, yields a velocity parameter 0 = 1.6 km s - 1• As the thermal ....q
I
q
.,.,
I
q
'°I
q
r-
1
velocity for these atoms in the solar atmosphere is about 1.2 km s - 1, this
implies a microturbulent velocity of about 1.0 km s- 1 • By a fit to the damping
part of the theoretical curves, log a = - 1.4 was obtained. Adopting an
average wavelength of 4500 A for the lines, this yields r = 1.7 x 109 s - 1,
which is very nearly a factor of 10 larger than the classical damping constant
rr. It is clear that the main source of the line broadening must be collisions
326 Classical Treatments of Line Transfer 10-4 The Empirical Curve of Growth 321

and because hydrogen is mostly un-ionized in the solar photosphere, the Thus, if we plot log(W/,W for lines of a given ion, say Fe I, versus log pg>
most likely process is van der Waals interactions between hydrogen and the instead of log p~, lines of different excitation potentials will scatter around
radiating atom. An approximate result for r 6 was obtained previously [ equa- a mean curve if there is a difference in excitation temperature between the
tion (9-71)], namely r6 ::::: 8.1 c6fv¾NH where c6 ::::: [13.6/Cx,on - x,)] 2 X star and the sun. Therefore, we plot log(W/).)* versus (log /J~ - XiJk ,M) and
10- 32 [equation (9-76)]. For a typical value (x100 - x,) = 6 eV, C6 ::::: choose 60 to minimize the scatter. For the G-subdwarfs it is found (13) that
5.2 x 10- 32 ; typical values for v and NH are 10 6 cm s- 1 and 10 17 cm- 3 , 60 ;$ 0.05, while for Population II K-giants M is about 0.25 to 0.35 (294).
so r 6 ::::: 109 , which is in basic agreement with the empirical value. The After the effect of 60 is eliminated, the empirical curve for the star is super-
damping part of the empirical curve is found to have a slope (on a logarithmic imposed upon the solar curve, in which log( W c/;,e 0 ) 0 is plotted versus log P'l,
scale) closer to 0.6 than 0.5. The likely explanation is that lines of widely The vertical shift gives directly [eo] while the horizontal shift yields the
differing strengths are formed in different layers, so that the parameters average value of o = [N1k] - [e 0k.U] for the ion under consideration. For
describing them (e.g., damping widths) are not identical, as was assumed in the G-subdwarfs e0 is found to correspond closely to a pure thermal value
constructing the curve of growth. (no turbulence), while for the K-giants there is appreciable turbulence. As
Many other analyses of the solar spectrum have been made. An improve- the temperatures of these stars are fairly close to the solar value, the partition-
ment over the basic curve of growth method is obtained by using a saturation function ratio is usually set to unity, and the value of o depends primarily
Junction (505) that accounts for the fact that the line cores and wings are upon abundances, and upon differences in ionization and continuous opacity.
actually formed in different layers of the atmosphere. This is conveniently Because H- is the main contributor to k. in the temperature range under
done in the method of weighting functions, which is described in detail in consideration, the value of [kc] will essentially equal [ n,]. We can determine
(261, Chap. 4), and which has been applied in one of the classic abundance [n.] from an analysis of the ionization equilibrium using informati~n from
analyses of the solar spectrum (252). two stages of ionization of the same element k. If we have observations for
An important application of the solar curve of growth is in differential two stages of ionization (say "O" and "1 "), we may write
abundance analyses of stars with respect to the Sun. Excellent examples of
this approach for G-type subdwarfs (extremely metal-deficient stars) and
Ojk = [N1d - [eokc], (j = 0, 1) (10-52)

for Population II K-giants cari be found in (13) and (294) respectively. An Using these two values of owe can derive
advantage of this approach is that oscillator strengths, which are often poorly· --.J
known, cancel out of the analysis to first order. Also, because the temperatures /l = 0 1k - Ook = [Nu] - [Nok] I\)
of these stars are close to the solar value, one might hope that their atmo- = log(N?JNa't) - log(N!JNtt) (10-53) I
spheric structures are at least roughly similar to the Sun's. If this is the case,
then other effects, such as line-blending, stratification, departures from LTE, As log(N?,JNg'k) is known, the observed value of /l yields log(N!t/Ntt), But
etc. might also cancel out to first order. if we assume that the temperature is specified by 0:.•
(which is known), then
The fundamental assumption made in a differential analysis is that the Saba's equation gives log(N!tn:/N~t); hence we may determine log and n:
stellar curve of growth, log(Wc/..le0)• versus log Pt, is identical to the solar [ n,]. Estimates of [ n,J can be obtained from several different elements,
curve, log(Wc/..le 0 f> versus log PW, In practice we do not know Pt or et, and a mean value taken. Knowledge of [n,] allows us to evaluate [k.] and
and thus we cannot construct the stellar curve directly. What is known for log(NMNt). Hence finally we can calculate the ratio of the abundance of
each line is log(W/,l)• and the value of log P~ for that line from its observed the element in the star to its abundance in the Sun as
solar equivalent width. From the definition of /Jo [see equation (10-45)],
we have log(NP/N:) = log(NP/N~) + log(N1t/Nt) + log(N~/NTt)
= log(N'f/N~) + log(N1t/Nt) + o + [e 0 k,U] (10-54)
log p~ = (N°)
(p0) Ni +
log Xuk(0:•• - 0~.) + log (eik~
e•k•u•)
U0 (10-50)
because all four terms on the righthand side of the second equality are now
If we define [X] = log(X 0 /X*) for any quantity X, then equation (10-50) known.
The results of the two analyses mentioned above lead to the striking
may be rewritten as conclusion that the abundances of the heavy elements in Population II
(10-51) stars are lower than their abundances in the Sun by a factor of the order
328 Classical Treatments of Line Transfer 329

or 10 2 ! This is a fact or tremendous significance for the construction of a 1.oo r-,--.-"'r""lrr----.....--T"7--,----r-:::=:;:::I:::I::::.:::~9


picture of the evolution of the Galaxy, and it implies that the heavy elements 0.90 -~·

--------~
in Population I stars such as the Sun are the result of nucleosynthesis in 0.80
earlier generations of stars. The errors in abundances derived from curve- 0.70
of-growth studies are typically of the order of a factor-of-two uncertainty; 0.60 0.90
although such errors are serious, and motivate the development of more 0.50 0.80
precise methods, they do not change the qualitative results for Population 0.40 -~-- 0.70
JI stars mentioned above. F(Al)/F, ~ 0.60

: : ~ / • ' H,
0.90
I0-5 LTE Spectrum Synthesis 0.30r 0.80
with Model Atmospheres 0.70
0.60 0.60
The curve of growth approach introduces a large number of simplifying 0.50
approximations, which seriously limit the accuracy of the results obtained
0 10 20 30 40
Considerable improvement is achieved if the physical assumption of LTE
is retained, but now a model atmosphere is employed to represent the depth- Al
variation of all physical parameters. Given such a model, one may compute l'IGURE 10-4
Hydrogen Balmer-line profiles in Vega, Solie/ clors: observed profiles from (509).
both ", and x,cfi. as functions of depth, allowing fully for the variation of the Crosses: observed profiles from (71). Abscissa: Ai. in A. Curves are computed
temperature, ionization-excitation equilibrium, Doppler widths, damping profiles from a model with T,rr = 9650°K and log g = 4.05. From (555), by
parameters, etc. It is then possible to compute the optical depth permission.

(10-55)
observed hydrogen line profiles in the standard star Vega (555). Often a
and hence the emergent flux in a line at frequency v plot is made, in a diagram of T.rr versus log g, of the loci of (T.rr, g) values ....J
for which the computed value ofa given parameter [say D8 or W(Hy)] equals w
(10-56) the observed value. The loci for different criteria will intersect, and thus I
define an optimum value of (T.rr, g). Ideally the curves would intersect at
by direct numerical quadrature, with as much mathematical accuracy as a single point; in practice they will intersect within a small area, which
desired. introduces an uncertainty into both 7'.rr and g. Examples of this procedure
The first step in this method is to select the model atmosphere that most for the normal B-star y Peg and the He star HD 184927 are given in (508)
closely resembles the stellar atmosphere to be analyzed. This choice is made and (299), respectively.
by comparing observed and computed values of certain key features in the Having chosen a model atmosphere, it becomes possible to carry out an
spectrum. Typically the comparison is made for (1) continuum features abundance analysis. One now replaces the curve of growth by a computation
such as the overall energy distribution, the Balmer jump D8 , or a color such of the equivalent width, of each line under consideration, as a function of
as (b - y); (2) line profiles of the hydrogen lines (which are density- element-abundance relative to hydrogen. Knowledge of the observed equiv-
sensitive); (3) ratios of line-strengths for lines of two ionization stages of alent width of each line thus leads to an estimate of the element abundance;
a given element. For example, in B-stars, the continuum parameters deter- these estimates may be suitably weighted and averaged over all lines to
mine mainly the effective temperature T.rr; the hydrogen lines determine yield a final abundance estimate.
mainly log g; the ratio of, say, Si III A.4552 to Si II ,l4128, 31 is a function Although the choice of the model atmosphere used in the abundance
of both Terr and log g (the ratio is insensitive to the element abundance). analysis is based on continuum and hydrogen-line criteria, there is usually
Examples of fits of computed and observed energy distributions are shown some ambiguity in ( T ,rr, log g). Often the results of the abundance analyses
in Figures 7-4 through 7-6 and in Figure 7-8. Figure 10-4 shows a fit to the for several elements can be used to narrow the range of uncertainty in
330 Classical Treatments of line Tran4er 331

(T.rr, log g), and to choose a more refined model, much in the same way 1.0 .----.---r--,---r--,-----r--,---n
as atmospheric properties may be inferred from a curve-of-growth analysis.
For example, after the analysis is completed, the derived abundances for
many lines may be examined to see if they show a correlation with excitation 0.8
potential; such a correlation would be expected if the temperatures in the
atmosphere are incorrectly chosen. The elimination of any such correlations
0.6
may, therefore, lead to an improved choice of T.rr for the model. Similarly,
for a given element, identical abundances should be derived from all ioniza- I i0, 0)/1,(0, 0)
tion stages. Differences that are found between ionization stages contain
information about errors in the choice of temperatures and pressures, and
thus about T.,, and Jog g for the model. Finally, one can examine the derived
abundances to see if they show a correlation with equivalent width. If, for
example, stronger lines systematically lead to larger abundances, the velocity
parameter may have been underestimated, and thus information about u
turbulent velocities to be included in the model can be derived. (As we shall 0~.1--L-3-9~84-.0-L-...___.__~3-9-84.5
see in Chapter 11, however, departures from LTE must also be considered
).
when a determination of the velocity parameter from strong lines is at-
tempted.) Naturally, errors in the observed equivalent widths, inf-values, FIGURE 10-5
Spectrum synthesis for solar Fe, Cr, and Ni lines. Curve:
and in other atomic parameters will introduce scatter; hence the procedure observed intensity; open circles: computed intensity. Ordinate:
outlined above may not be completely unambiguous. Nevertheless, exam- emergent specific intensity at disk-center relative to continuum;
ination of the correlations just mentioned can often lead to a significantly abscissa: wavelength in A. From (538) by permission.
better choice for the model. Finally, if the derived abundances _are very
different from those used in construction of the model (as they might be All of the methods described in this chapter make the assumption of LTE.
for, say, Ap stars), it may be necessary to recompute models with appropriate As we shall see in Chapters 11 and 12, however, this assumption is often a --1.
abundances, re-derive T,rr and log g, and perform the analysis over. poor description of the physics of line-formation, and can lead to results
An example of the model-atmosphere method of abundance analysis is ......J
containing serious systematic errors. ~et us no~, _therefore, turn to t~e ~
that for the'two bright A-stars Vega and Sirius in (609). The results of this
problem of solving the combine~ equations of rad1at1ve transfer and atomic I
work show that Vega has very nearly solar abundances of the heavy elements,
statistical equilibrium fully consistently.
while Sirius has abundances a factor of 4 to 10 higher. In many respects
Sirius mildly resembles the group of A-stars called metallic-line A-stars (Am
stars) because of the strength of the metal lines in their spectra. The present
literature of abundance analyses (and estimates of atmospheric parameters
such as T.rr and log g) by both curve-of-growth and model-atmosphere
techniques is vast; extensive lists of references to research papers on the
subject for both normal and peculiar stars can be found in (12; 70, 57-204;
144; 450, 157-237; 493; 523; 552; 560; 658). The work thus far described
makes use primarily of equivalent widths. For some stars, for which very
high-quality spectroscopic data are available, a detailed point-by-point syn-
thesis of the spectrum is possible. An example of such a synthesis for the
Sun (538) is shown in Figure 10-5. Analyses of this kind can lead to accurate
abundances and to the identification of hitherto unnoticed weak blends in
the spectrum; applications of spectrum synthesis to the calibration of photo-
metric systems were described in §7-4.
I/./ Diffusion, Destruction, Escape, and Thermalization 333

is handled. In the LTE approach, it is assumed that the local values of two

11 thermodynamic variables {T and N) are sufficient to determine compl~t~ly


the excitation and ionization state of the gas {and hence transfer quantities
such as x., r,., and s.), independently of the state of the gas at othe~ po_ints.
{This remark is strictly true only when the temperature structure 1s g1~en;
if it is determined by the condition of radiative equilibrium, then there is_ a
Non-LTE Line Transfer: coupling among different points in the atmosphere-see §?-2-:-but n~t m
the sense we are concerned with here.) It has been emphasized m previous
chapters that the excitation and ionization state of the gas is, in fact, strongly
The Two-Level Atom influenced by the radiation field, which, in turn, is determined by the state
of the gas throughout large volumes of the ~t~osphere, via the tr~n~fer
equation. In actuality the two problems of radiative t~ansfer ~nd stat1st1cal
equilibrium are inextricably coupled, and must be considered simultaneously.
Much of the basic physics of line-formation can be understood by con-
sideration of the characteristic lengths for photon diffusion, destruction, and
thermalization, which are related to photon escape and destruction prob-
abilities. For simplicity we shall suppose that the atom consists of a ground
state (denoted /) and a single excited state (denoted u). The distance over
which a line photon will, on the average, move in the atmosphere between
successive interactions with the material (absorptions) can be represented
by the average mean-free-pa~h I. Th~ mean-free-path f?r a p~oton emitted
at a given frequency is the distance mterval whose optical thickness at that
Much of the progress in understanding the physics of line-formation has
come from a study of solutions of the combined transfer and statistical.
x,.
frequency is about unity-i.e., I.:=::: 1/x. = {x!•~• + x.)- 1. where_ and Xe
are the line and overlapping contmuum opac1t1es, and q,. 1s the !me pr~file.
equilibrium equations: the so-called non-LTE approach. In this chapter we,
For complete redistribution, the probability ·of emission at frequency v 1s cf,.,
.....
shall consiµer some schematic line-formation problems that are simple
enou~h . to be solved ~eadily: but which, nevertheless, provide a good
J
so / = (I.) = /,cf,. dv. It is obvious that photons emitted in the core travel --...J
relatively small distances, while those i~ the win~s can_travel_much g~eater 01
~e~crtpt1on ~f the phy~1cally important processes-and yield considerable distances, up to a distance correspondmg to umt optical thickness m the I
ms1ght. It will be obvious that some of the assumptions made are over-
continuum. ·
simplifications, ~nd are not valid in actual stellar atmospheres, for which
When a photon is absorbed in the line and excites an atom to the upper
elaborate numencal calculations are required to yield accurate results. On
state, it is usually re-emitted in a radiative de-excitation and travels another
the other hand, the real goal is to understand the answers, not merely obtain
mean-free-path. This process may occur again and again before the photon
them. 1:'his ~an be done o?IY with a clear grasp of the prototype cases dis-
is ultimately destroyed, and has its energy deposited ~nto_ the thermal p~ol,
cussed m this chapter, which provide a conceptual framework of great use-
either by a collisional de-excitation or by an absorption m the overlappmg
fulness for the interpretation of results from computations with complicated
model atoms and detailed model atmospheres. continuum. Thus there exists a characteristic length L, the destruction length,
over which a photon may travel before it is destroyed. The destructio~ length
has a more basic physical significance than the mean-free-path, for 1t mea-
sures the distance over which a photon emitted at a given point retains its
11-1 Diffusion, Destruction, Escape, identity and hence can "communicate" information about conditions at that
and Thermalization point to another. Thus L determines an interaction r~gion: the volume con-
taining those points that can influence one another via photon exc~ange.
The most important difference between the LTE and non-LTE treatments of
The relative sizes of L and I depend upon the photon destruction prob-
line-formation is the way in which the coupling between the gas and radiation
abi/icy Pd, which gives the average probability that the photon is destroyed
334 Non-LTE Line Transfer: The Two-Level Atom 11-1 Diffusion, Destruction, Escape, and Thermalization 335

when it next interacts with the material. The probability for photon destruc- a departure of the radiation field from its equilibrium value at positions far
tion by collisions following absorption in the line is c.,/(A.1 + c.,) where c., is below the surface.
the rate of collisional de-excitation and A., is the spontaneous emission rate; The depth at which the radiation field (or source function) ultimately
now c., ex: n,, so that it is clear that the contribution by collisions to the approaches closely its equilibrium value is the thermalization depth A; this
total destruction probability becomes large (i.e., approaches unity) in the concept, introduced by Jefferies (333), has proven to be extremely fruitful.
deeper, denser regions of a stellar atmosphere, and can be quite small in the To obtain a quantitative estimate of A we define the escape probability Plr)
uppermost layers. If we assume that all photons absorbed in the continuum as the probability, averaged over the line, that a photon emitted at line optical
are destroyed thermally, the contribution of continuum processes to Pd is depth t escapes from the medium before being absorbed along its path of
the average of x./(x,.,f,. + x,) over the line profile. The continuum sets an flight. The mechanism competing with photon escape is photon destruction;
upper bound on both I and L because a photon cannot travel more than a the former induces a departure from equilibrium while the latter leads to
unit optical depth in the continuum before it is absorbed-and when it is coupling to the local thermal pool. We therefore compare P, with Pd. Deep
it is also destroyed. ' within the atmosphere where P,{t) « Pd, photons are surely thermalized
At great depths in the atmosphere, Pd approaches unity because of large before they can escape, hence S ➔ B; at the surface where P,{t) » Pd,
densities (and hence collision rates) and because of the strength of the photons escape freely before thermalization, hence S will depart from B.
overlapping continuum. Then L -+ I, for the photon is almost surely de- It is therefore reasonable, on physical grounds, to identify the thermalization
stroyed when it is next absorbed. Assuming that the continuum is already depth A as that point at which P,(A) = Pd. Defined in this way, A is essen-
optically thick, it is clear that the radiation field becomes very strongly tially the greatest depth from which photons have a significant chance to
coupled to local conditions and thermalizes to its local thermodynamic escape before being destroyed.
equilibrium value. In contrast, when the destruction probability is very small, Now the escape probability (and hence A) depends sensitively (a) upon
L » I, and the interaction region may become enormous c~mpared to the the nature of photon redistribution over the line profile upon emission,
volume over which a photon can diffuse in a single flight. In this,case, the and (b) upon the amount of.background continuum absorption. The prob-
radiation field is dominated by nonlocal influences, and represents the result ability for photon absorption is highest at line center. If photons are emitted
of physical conditions that may be quite inhomogeneous. For example, coherently, then those absorbed at line center will be re-emitted there, and
within the volume there may be large variations of the kinetic temperature hence will tend to be trapped by the large line-core optical depth up to the
.......
that imply ~trong variations of thermodynamic properties and of the Planck very uppermost layers, where the line finally becomes optically thin; this ---J
function. The radiation field may then depart markedly from its local equilib- will tend to prevent serious departures of the line-core radiation field (which CJ)
rium value, and this departure will extend throughout the entire interaction contributes heavily to the total photoexcitation rate) from its equilibrium I
region. value up to the shallowest layers of the atmospher~. In contrast, if photons
The importance of these notions becomes manifest when we consider a are completely redistributed over the line profile, then there is a significant
sequence of test points approaching the surface of the atmosphere. At chance that after a number of scatterings a photon absorbed at line-center
the deepest points we obtain equilibrium. But as we approach the surface, the will be emitted in the line-wing where the opacity is low, and the probability
distance L grows, and eventually the interaction volume contains the of escape is high. Photons that would have been trapped, if emitted co-
boundary itself (more precisely, extends above the depth at which t = 1 herently, now freely escape from the atmosphere, depressing the line-core
anywhere in the line). Then a new phenomenon enters the picture: photons intensity {and photoexcitation rate) in much deeper layers in the atmosphere.
escape from the medium into space. It follows that the radiation field, at The radiation field in the line as a whole now responds to the fact that the
test points whose interaction regions extend into space beyond the surface, boundary lies within a mean-free-path at some frequencies, even if not at
must be depressed below the value it would have had if there were no bound- others. The role of the continuum is obviously important in the case of
ary, for no radiation is emitted in the region void of material, and photon complete redistribution, for it sets a lower bound on the total opacity, and
escapes are therefore uncompensated. This effect leads to a deficiency in the hence an upper bound on the depth from which escapes become possible
radiation throughout an entire interaction volume, and hence extends to at any frequency, no matter how far out in the line-wing. It is thus clear that
depths at least of the order of L. But of course the radiation field at these we should expect the thermalization depth to be much larger, and the surface
points influences that at points that lie yet another destruction length L departures from equilibrium larger, when the scattering process is non-
deeper, and therefore there is a "compounding" of the effect, which leads to coherent instead of coherent; further, we expect the magnitude of these
336 Non-LTE line Transfer: The Two-level Atom 11-2 The Two-Level Atom without Continuum 337

effects to be larger, the more highly developed the line-wings are relative to The statistical equilibrium equation for level I is
the core. We shall find these expectations are fulfilled by the detailed analvsis
given below. ·
It has been possible to extract practically all of the physical flavor of the
problem using the qualitative arguments given above. Let us now turn to a
mathematical discussion to obtain quantitative results, and to extend the This equation can be used to write an expression for (n 1/n.). Upon substituting
analysis to cases where such heuristic discussion becomes less effective. this expression into equation (11-4), making use of the detailed-balance result
that c,. = (n,,/n 1)*C.,, and using the Einstein relations, we obtain quite
directly
11-2 The Two-Level Atom without Continuum
THE SOURCE FUNCTION
S, = [f cf,.J. dv + e'B.]/o + e') = (1 - e)J. + eB. (11-6)

Consider a schematic atomic model consisting of only two levels, 1and u, where e' = c.,(l - e-hv/kT)/ A., (11- 7)
between which radiative and collisional transitions can occur. This model is
obviously very incomplete, but it nevertheless provides a fairly good de- and e = e'/(1 + e') (11-8)
scription of the real situation for some lines. In particular, resonance lines
Exercise I 1-I: Derive equation (11-6} in detail.
arising from the ground state are well described by this model when the
coupling of the lower and upper levels to the continuum (and of the upper
Each of the terms in equation (11-6) admits a straightforward physical
level to other levels) is weak. For the present, assume that the only sources
interpretation. The source function contains a noncoherent scattering term J
of opacity and emission at the line frequency are from the line itself; then
and a thermal source term e' B•. The thermal source term represents photons
the transfer equation is
that are created by collisional excitation, followed by radiative de-excitation.
µ(dl./dz) = [-n,B,ulv + n.(A.1 + B.,I.)]cf,.(hv/41t) (11-t) The term e' in the denominator is a sink term that represents those photons ......
that are clestroyecl by a collisional de-excitation following a photoexcitation.
Here we h_ave assumed complete redistribution so that the absorption and These two terms describe completely the coupling of the radiation field to
emission profiles are identical. Define the optical depth scale in terms of the the local thermal state of the gas. The scattering term may be viewed as a
frequency-integrated line opacity (which characterizes the average opacity in reservoir term that represents the end result of the cumulative contributions
the line as a whole), dt = - x,.
dz where of the source and sink terms over the entire interaction region.
It is clear that, if densities are made sufficiently large, then the collision
x,. = (n B
1 1• - n.B.1)(hv/41t) = (1te 2/mc)f, 11 [ n1 - (g 1/g.)n,,] (11-2) rate C,,1 may eventually exceed A,, 1, so that e' becomes » 1; then S, -+ B.(T),
and LTE is recovered. However, in virtually all situations of astrophysical
Then µ(dl./dt) = cf,.(!,. - S1) (11-3)
interest, e' « I in regions of line-formation, and, in general, the source
where S, = n.A.1/(n 1B1• - n.B.1) = (2hv 3 /c 2 )[(n1g./n.g1) - l]- 1 (11-4) function cannot be expected to have a value close to the Planck function.
This state of affairs was partially recognized in the classical theory by the
the second equality following from application of the Einstein relations division of lines into the categories of "absorption" and "scattering" lines;
[ cf. equations (4-8) and (4-9)]. As the factor v3 varies only negligibly over the this division, however, was largely ad hoc, and the thermal coupling param-
sharply peaked profile cf,., S1 is called the frequency-independent source eter had to be guessed from heuristic arguments. In the present analysis
function; when the emission profile differs from "the absorption profile, S1 the coupling parameter follows directly, and uniquely, from the statistical
becomes explicitly frequency-dependent (see §2-1 and §13-4). Equation (11-4) equilibrium equation. Further, in the classical treatments of"pure" scattering
is an implicit form for the source function because the level-populations lines, it was sometimes incorrectly argued that the small thermal terms could
depend upon the radiation field; this dependence can be displayed explicitly be discarded. The important point to bear in mind is that, even if the thermal
by incorporating information from the equations of statistical equilibrium term e'B is small compared to the scattering term J locally, when integrated
that determine n1 and n•. over the entire interaction region it accumulates to a value of importance;
I 1-2 The Two-Level Atom without Continuum 339
338 Non-LTE Line Transfer: The Two-Level Atom

moreover, at depths greater than the thermalization depth the intensity must and (l l-13b)
ultimately couple to the thermal pool. Indeed, if the thermal terms are dis-
carded, the transfer equation becomes homogeneous in the radiation fi.::ld
and the scale of the solution is unknown; this scale is, in fact, fixed by the where the points are chosen symmetrically-Le., X-n = -x. and µ_m = - µm
(small) thermal terms at the point of thermalization. (also a_. = a. and b_m = bm). Then the source function can be written in
the same discretized form as was used in Chapter 6 [cf. equation (6-24a)],

SOLUTION OF THE TRANSFER EQUATION


sd = (1 - ed) La.
n
Lm bmUdmn + edBd (11-14)
Having obtained an expression for the source function, let us now con-
sider the solution of the transfer equation. It is convenient to work with the and the transfer equation can be reduced to the standard second-order
dimensionless frequency-variable x, measured from line center in units of
differential equation form:
Doppler widths (or damping widths for Lorentz profiles). In terms of this
variable we shall write Doppler profiles as (11-15)
<f,(x) = ,r-te-xz (11-9a) where u, = ½[I,(+µ) + I,( - µ)]. Equation (11-15) can be discretized on a
Voigt profiles as mesh {td}, and solved numerically using either the Feautrier or Rybicki
schemes discussed in §6-3. All of the relevant parameters (e, B, and <f,) may
<f,(x) = an-t f~"" e_',[(x - y)2 + a2]-1 dy (11-9b) be depth-dependent without causing any difficulty in the calculation.
Alternatively, the transfer equation can be written in integral equation
and Lorentz profiles as form. The formal solution of equation (11-12) is

<f,(x) = (1/n)/(1+ x 2) (11-9c)


J(t, x) = ~ Jo"' S1(t)E 1 \.J:' <f,(t', x) dt'\ <f,(t, x) dt (11-16)
....I.

which are aH normalized such that


-....J
f~ <f,(x) dx = 1 (11-10)
For a depth-dependent profile, the argument of the £ 1 function will depend en
00
upon both t and t; this greatly complicates the analysis, so we shall consider I
only the case of a depth-independent profile, in which case only the dis-
We will absorb a factor of l:,.vD [or (r/4n) for a Lorentz profile] into the
definition of x and write dt =- x,(z) dz, where now x, =x,u/llvD, x,u
placement (t - t) enters:
being given by equation (11-2). Then
J(t, x) = ~ <f,(x) f 0
00
S1(t)E 1 l(t - t)<f,(x)I dt (11-17)
S (t) = [1 - e(t)] f~ <f,(t, x)J(T, x) dx + e(t)B(t)
1 00
(11-11)
Substituting this expression for J into equation (11-11) yields an integral
and the transfer equation becomes equation for s,.
µ[dl(t, x)/dt] = <f,(T, x)[J(t, x) - S,(T)] (11-12) S1(t) = [1 - e(t)] f0
00
S1(t)K 1 It - ti dt + e(t)B(t) (11-18)

We now introduce discrete angle and frequency meshes {µ,,,} and {v.},
and replace integrals with quadrature sums: where the kernel Junction K I is given by

(11-19)
(ll-13a)
340 Non-LTE Line Transfer: The Tll'o-Level Atom 11-2 Tire Two-Level Atom without Continuum 341

Exercise 11-2: Fill in the missing steps in the derivation of equation (11-18). THE THERMALIZATION DEPTH
Exercise 11-3: (a) Show that the normalization of K 1(s) is Jo' Ki(s) ds = ½.
(b) Using the result in (a), show that, fort --> oo, where S1(t) must be a slowly-varying Let us now consider the behavior of the source function in an atmosphere
function and hence can be removed from the integral, S1(t) --> B. in which e and B are constant with depth. Naturally this is a very schematic
case, but it provides useful insight. One of the most important characteristics
Numerical methods for solving equations of the form of (11-18) are dis- of the problem to be determined is the thermalization depth at which the
cussed in detail in (52) and (18, Chap. 8); in most respects these resemble line source function approaches closely the Planck function B. Expressions
the Rybicki method of solving the differential equations, though Rybicki's for the thermalization depth can be derived by a wide variety of methods
method is simpler to apply when the profile is depth-dependent. In essence including (a) analysis of the asymptotic behavior of the solution of the
the solution is obtained by introducing a functional representation of S(t) integral equation (11-18) (53); (b) calculation of the distribution of the
on a discrete grid {td}; these functions are integrated analytically against distances photons can travel from their points of creation to their places
the kernel to construct a final matrix system of the form of destruction (545; 322); and (c) calculations of the probability distri-
butions for photon escape from a given point of origin in the medium
S=KS+L (11-20) (214; 215; 216). One of the simplest and most appealing derivations can be
made on the basis of the physical arguments given in §11-1: that the ther-
where s = (St, ... 'sd, ... 'sor
represents the depth-variation of the malization depth A must be near that point in the atmosphere where the
source-function. The system is then solved by standard numerical techniques. probability of photon escape, P,(A), following a given scattering event is
One of the advantages of the integral-equation formulation is that it equal to Pd, the probability of photon destruction (28).
displays explicitly the intimate dependence of the source function upon the For the strict two-level atom with no overlapping continuum opacity, the
mathematical behavior of the kernel function, and considerable insight can probability of destruction per scattering event is just
be gained by analytical study of the kernel. In particular, the asymptotic
form of K 1(s), for s » 1, shows that a characteristic feature of line.,formation (11-21)
problems with noncoherent scattering is an extremely long-range interaction
of one part of the atmosphere with another. Recall that for coherent scatteririg where e is as defined by equations (11-7) and (11-8) (ignoring stimulated _...
the appropriate kernel function in the formal solution is E 1 It - ti [~f. emission for simplicity). The escape probability can be calculated by summing --..,J
equation (2-57) and Exercise 2-10]. This kernel decays very rapidly, falling exp(- t</Jx/µ), the probability of escape at frequency x along a ray at angle
off as exp( -It - tl)/jt - ti, thereby limiting severely the range of depth- cos- 1 µ to the normal, over all angles and frequencies, with weight if>x (the
co
I
points that are directly coupled together in the scattering process. In contrast, probability for emission at frequency x within the profile):
the asymptotic behavior of K 1(s) is found to be (53, Appendix I) K 1(s) ~
l/{4s 2 (/n(s/,r½)]½} for Doppler profiles, K 1(s) ~ a½/(6si) for Voigt profiles,
and K 1(s) ~ l/(6s½) for Lorentz profiles. Clearly the range of these kernel
functions is very large compared to that for coherent scattering, and this
implies that the radiation fields at widely separated points in the medium (11-22)
become mutually interdependent.
The physical reason for the long range of the noncoherent kernel functions
is, of course, the redistribution of photons into the transparent line-wings. Fort » 1, the exponential integral is approximately zero for !xi ~ x 1, where
In the coherent case, any photon absorbed in the core will be re-emitted q,(xi}t = 1, and unity for !xi > x 1• Physically this states that photons are
there, and hence always encounters high opacity and remains trapped. In trapped in the core, where t</Jx > 1, but in the wing where r</J., ~ 1, they
the noncoherent case it has a probability of being emitted in the wings where, escape freely. We may thus approximate P,(r) as
because of the lower opacity, it can travel over a 1arge geometrical distance,
and a correspondingly large integrated line optical depth, before it is again
P,(t) ~ f"' q,(x) dx (11-23)
absorbed. Jx,
342 Non-LTE Line Transfer: The Two-Level Atom 11-2 The Two-Level Atom without Continuum 343

The above integral is easy to evaluate for x 1 » 1 (which it will be when In Chapter 13 we shall find that in the case of resonance lines, the scattering
t » 1). By direct substitution of equations (11-9) we find in strong radiation-damping wings is nearly coherent, and A is better approx-
imated by equation (11-26a) than (11-26b).
P.(t) = ~ erfc(x 1) :::::: 2
e-x• /(2n¼x 1) (Doppler) (ll-24a) From a mathematical point of view, equations (11-26) show that any
iterative solution of the transfer equation starting from L TE is futile for
small e (much more so than the cases already discussed in §6-1 !), and only
P,(t):::::: a/(1tX1) (Voigt) (11-24b) direct solutions of the transfer equation will yield correct values.
P.(t) :::::: 1/(nx1) (Lorentz) (ll-24c)
BOUNDARY•VALUE AND DEPTH-VARIATION
In deriving equation (ll-24b), use was made of the asymptotic form of the OF THE SOURCE FUNCTION
Voigt function [cf. equation (9-45)]. Now the condition q,(x 1)t = 1 implies
Having shown above that the source function may be decoupled from
x1 = [ln(t/ni)]¼ (Doppler) (11-25a) the Planck function over great depths in the atmosphere, let us now inquire
what value the source function attains near the boundary of a semi-infinite
x 1 = (ar/n)t (Voigt) (ll-25b)
atmosphere. If e and B are taken to be constant with depth, the transfer
X1 = (t/1t)¼ (Lorentz) (11-25c) equation (11-12) is easily solved by the method of discrete ordinates (53;
476). Writing An = l/tp(xn) we have the coupled system of equations
Hence, substituting equations (11-25) into (11-24), setting P,(A) = Pd = e, l N M
and solving for A we find
Anµm(d/mn/dt) = lmn -
2 (1 - 8) n'f:. Nan' m'J;.M bm,lm'n' - BB (11-27)
A:::::: c/s (Doppler) (11-26a)
2
Considering first the homogeneous equation (B = 0) and seeking a solution
A:::::: a/s (Voigt) (ll-26b) of the form Imn = gmn exp( -kt) one finds that
A :::::: 1/s2 (Lorentz) (ll-26c) Imn = c(l + klnµm)- 1e-kt (11-28) .......
co
where c is a number of order unity that depends implicitly upon s; other where the constants k are roots of the characteristic equation 0
factors of order unity have been suppressed. I
N M

Exercise I 1-4: Show that for a profile with asymptotic form q,(x) ~ ex·•, with 2(1 - E) L an L
n=l m= 1
bm(l - k 2 l/µm 2)- 1
·
= 1 (11-29)
ix> 1, the thermalization depth varies as A ~ e· 1, where p = ix/(ix - 1).
Exercise 11-5: (a) Verify equation (11-28). (b) Show that equation (11-29) is the
The striking feature of the results given by equations (11-26) is that the appropriate characteristic equation for the system (11-27).
thermalization depth for a line with small s is enormous. Recall that for
coherent scattering the thermalization depth ~s-t (cf. §6-1 and 10-2). It is For e = 0, the root k2 = 0 is a solution. For e > 0, k2 must be greater
clear that the effects ofnoncoherence, with their attendant increase of photon than zero. As in our earlier use of the discrete-ordinate method (§3-3, §7-4)
diffusion in the line wings, greatly increase the depth in the atmosphere over we may delimit the roots by the poles at k2 = 1/l/µm 2, and ifwe label these
which the source function can depart from the local Planck function. In fact, poles in the order of decreasing values of (lnµm) we may write
intercomparison of equations (ll-26a) through (ll-26c) shows immediately
that an increase in the relative importance of the line-wing within the profile (11-30)
leads directly to an increase in the thermalization depth. It is worth stressing,
however, that the above results do explicitly depend upon the assumptions of which shows that the roots may be determined rapidly by a systematic search
(a) no background opacity and (b) complete redistribution. We shall see later on appropriate finite intervals. As can be seen, it is c?nvenient to label_ t~e
in this section that the effect of a background opacity can greatly reduce A. roots with a subscript IX, 1 ~ IX ~ MN. By an analysis of the characteristic
344 Non-LTE Line Transfer: The Two-Level Atom 11-2 The Two-Level Atom without Continuum 345

equation similar to that used in deriving equation (3-65) one may readily By substitution of equation (11-37) into (11-35), we obtain finally the
show that extremely important result
MN
S1(0) = etB (11-38)
TI
«~1
(kaJ..µ,,,) 2 =& (11-31)
This result is independent of the order of the quadrature sum and the form
We shall use this result below. For B = constant, a particular solution of of the profile, and hence is general. The basic physical content of the results
equation (11-27) is Im• = B. Thus the general solution must have the form obtained thus far is contained in equations (11-38) and (11-26). It may be
summarized in words as follows: when the coupling of the line-emission
lmn =B [1 + I a•l
Lae-k•t(l + k«J.nµm)- 1] (11-32) process to the termal pool is weak (e « 1), the source function can depart
drastically from the thermal values, and this departure can extend to great
depth in the atmosphere. Note that the surface value of the source function
where the ascending exponentials have been discarded to keep the solution exceeds the local creation rate by a factor of 1/et. This implies that S1(0)
bounded at infinity. The surface boundary condition is, as usual, Im•(- Jtm) = is controlled primarily by photons fed in from the line wings. These photons
0. If we define originated deep in the atmosphere where S1(t) » S1(0); it is therefore clear
MN that the surface value of S1 has little to do with the local thermal source term
9'(x) =1+ L La(l -. kax)- 1 (11-33) but is dominated by nonlocal effects.
«•1 Numerical solutions (53) for the full depth-variation of S1(t) are shown in
then the surface boundary condition may be expressed as 9'(J..µm) = 0 for Figures 11-1 and 11-2. The cases in Figure 11-1 are for a semi-infinite atmo-
(m = 1, ... , M) and (n = 1, ... , N). This set of linear equations may be sphere, constant Planck function (B = 1), a Doppler profile (a = 0), and
solved numerically for the L«'s. Substituting equation (11~32) into the ex- various values of e. It can be seen that in each case S1(0) = et, and that
pression for S, on the right-hand side of equation (11-27), and making use S1 --+ B at t ~ 1/e. The results shown in Figure 11-2 are for lines with e =
of the characteristic equation (11-29) one finds 10- 4 , and for Voigt profiles ranging from a pure Doppler profile to a pure
......
S,(t) =B (1 + I L«e-k•r)
«•1
0 (X)
......
In particular, at the surface, -1

S,(O) =B (1 + I L«) =
a•I
B9'(0) (11-35) -2
logS
Now, by an analysis similar to that used to derive equations (3-70) and -3
(7-104) one may show that

9'(x) = [Ji .6 Jt

1
(),.µ,,, - x)]/Ji (1 - k«x) (11-36)
-4

-5
from which it follows, using equation (11-31) that
-1 0 I 2 3 4 5 6 7 8 9 10
MN
9'(0) = TI (k«Anµm) = e+ (11-37) log t
a=I FIGURE 11-1
Line source functions in a semi-infinite atmosphere with B = I,
Exercise JJ-6: (a) Derive equation (11-31). (b) Derive equation (11-34). (c) De- for a line with a pure Doppler profile (a = 0), and various values
rive equation (11-36). of e. From (53), by permission.
Jl-2 The Two-Level Atom without Continuum 347
346

atmosphere to which a particular ion is confined owing to changes in the


0
11 =0 ionization balance (e.g., chromospheric tines of, say, He II are limited to
u = 10··1 layers bounded above by the corona and below by the photosphere)._ In
Cl= JO-l finite atmospheres two physically distinct behaviors are found, depending
on whether the atmosphere is effectively thick or effectively thin. If T » /\,
then photons from the stab center will not escape before they thermalize;
logS -1
in this case S1(0) will attain its semi-infinite atmosphere value for the corre-
sponding value of e, and will approach B at depths ~ /\ from the surfaces.
If, however, T « /\, then the solution never thermalizes, and S1(t) becomes
proportional to the local creation rate-i.e., S1(t) = 1$B/(t) where /(t) is
independent of e for a given T.
An estimate of S1 at stab-center can be obtained as follows [see also (28)].
-2 -1 0 2 3 4 s 6 1 8 The ratio of the total number of emissions along a column through the slab
log r to those thermally created must be equal to the mean number of times,
FIGURE I 1-2 (N), a photon is scattered before it escapes or is destroyed, i.e.,
Line source functions in a semi-infinite atmosphere with B = I, for lines with & = 10-• and
Voigt profiles ranging from a pure Doppler profile (a = 0) to a pure Lorentz profile (a :, oo)
From (53), by permission. · (N) = s~CX) dv(4nhvcf>v) s: S,(t) dt Is~CX) dv(4nhvcf>v) s: e(t)B(t) dt
Lorentz profile (a = oo). The increase in thermalization depth from e- 1
= J: S1(t) dt If: e(t)B(t) dt (11-39)
to e- 2 is shown plainly. The source functions in both Figures 11-1 and 11-2
Here we have used the definition r,, = x,S., and noted that S, is frequency-
will yield absorption lines with dark cores; in contrast, the LTE solution
with S, = B, would yield no line whatever. This difference can be.attributed independent. For a finite slab the dominant photon loss-mechanism is
to th_e effects. of scattering, and indeed the classical theory would also have escape, hence (N) ~ [P,(T)]- 1 ; at stab center P,(T) ~ [~P.(½T)]oo, w~ere
the subscript denotes the escape probability. from the md1cated depth m a
predicted a hne of the same central depth for the same e. However there re-
semi-infinite slab, and the factor of two accounts for losses through both
ro
main two important differences. (a) Both the upper- and lower-state occu- I\)
faces. To obtain an order-of-magnitude estimate from equation (11-39), we
pation numbers differ from their LTE values; in many earlier "scattering"
replace S (t) with Smax = S(½T) and assume eB is constant so that (N) ~
I
calc~lations LTE populations were (incorrectly) assumed. (b) The dark 1
portion of the non-LTE profiles, where noncoherent scattering is assumed SmaJ(eB). Then, using (11-24) and (11-25) to cal~ulate P,(T), and agam
are wider. The Eddington-Barbier relation implies that/ (0) ~ S (t = 1), ignoring numerical factors of order unity, we find
hence the lines will be dark for !xi ~ x 1 where Acf,(x 1) ,,; 1. Bec~u;e A i~ (11-40a)
so much larger for noncoherent scattering (e- 1 to e- 2 instead of e-t) the Smax ~ eT(lri T)½ B (Doppler)
~orrespon~ing val_ues of x 1 will also be larger; we shall return to this ~oint (Voigt) (11-40b)
m §11-5. Fmally, 1t must be remarked that the solutions obtained here are
fully ~onsistent solutions of both the transfer and statistical equilibrium (11-40c)
(Lorentz)
equations; any approach that falls short of this level of consistency is and
unsatisfactory.
The behavior described above is seen very clearly in the numerical results
(53) shown in Figure 11-3, whi~h gives S1(t) for lin~s with Doppler profiles
FINITE SLABS
(a = 0), for various values of e, ma~ ~tm~sphere with T = _10 . The d~!hed
. The finite_ slab atmosphere of total thickness T is a case of astrophysical curve gives the solution for a sem1-mfi~1te atmosphere with e = _1~ . _It
importance; 1t can be used to represent nebulae, or limited zones in an can be seen that for e ;;:: 10- 4 the soluttons closely resemble sem1-mfimte
348

-1 e = 10-2 .,..
8
II
f-,
'<I'
_ 2 _e_.=:. 10_-_4..

..... ,s
>< 'ii
logs -3 fl
N <C
...0
Q,
e .. 10- 6
.i
-4 ~
0 ....
'ti
C:

-s .,, u
-a
Q,
0
0
e • 10- 1 '<I' .5
-6
-2 -1 0 2 3 4 E
....
><
-a"'
log t 0

FIGURE 11-3 N .
!l
Linc source functions in finite atmospheres
with (total thickness) T • 104 and B • 1, for
-~
c:l
ci::
C: __.,
a line with a pure Doppler profile (a • 0).
Dashed curve corresponds to the semi-infinite

I;! cl (X)
atmosphere solution withe • 10-•. Note that 0

S1(t) is symmetric about T/2 (not shown owing


0 'ti·-
I; .!a w
to use of a logarithmic scale for abscissa. 8 "' u ~ I
From (53~ by permission. ·a
ci::
8..
>,
'<I' e ..o
0 ,;f-
J:: Ill
»-
atmosphere solutions at the corresponding e, while fore < 10- the atmo- 4
..... .1::
"' 0
e
>< C: ...
sphere becomes effectively thin and S1 falls below the corresponding semi- B t...
infinite-case curve, and, in fact, scales linearly with e. Emergent intensities
(53) are shown in Figure 11-4 for lines with e = 10- 4 , and for various
N
-
-
,. .5
-1:11,ho
C: I
~

values of a and T. For T = oo, an absorption line is obtained in every case, g i;


I.: IJ.l ..
with central intensity independent of a. In finite atmospheres, emission lines
are obtained, for the line wing becomes completely transparent for suffi- L.10_ _ _ _ _...J__ _ _.ll.-:'N".""""_ _. . _ _ _ - ; . . . . , ; - - - - - - ~ - - - ~ O
ciently large x, and the intensity must go to zero. At smaller x the intensity I I
rises rapidly, and, for effectively thick atmospheres, saturates to the semi-
infinite atmosphere value. Finally, in the line core, scattering leads to a
self-reversal. The profiles shown in Figure 11-4 strongly resemble those
from laboratory emission sources with saturated lines, and from hot
chromospheric layers above a relatively cool photosphere.
//-2 The Two-Level Atom without Continuum 351
350 Non-LTE Line Transfer: The Two-Level Atom

THE EFFECTS OF AN OVERLAPPING CONTINUUM destruction. The formal solution of the transfer equation (11-42) yields

All of the discussion presented thus far in this section has been based
on t?e as_sumption that th~ on~y opacity source is the line itself. In virtually
J,,(,:) =~ S S,,(t)E
0
00
1 j(t - ,:)(<Px + r)j (</Jx + r) dt
all s1tuat1ons of astrophysical interest, however, the line will be superposed
on a background continuum arising from bound-free absorption by other
levels of the atom, or by other atomic species. As was noted in §11-1, the
= 1S
cf," "' S (t)E 1 !(t - ,:)(<Px
0 1
+ r)! dt + 1S rB 0"' E1 j(t - r)(cp" + r)I dt

presence of a background continuum affects the destruction probability (11-45)


Pd, and sets an upper bound on the photon mean-free-path and destruction
length. We therefore expect the thermalization depth to have an upper The solution given by equation (11-45) could be substituted into (11-43)
bound set by continuum processes. Furthermore, the continuum provides to produce an integral equation for the total source .function S,,(r), but
an additional source of emission into the_ line. A detailed study of the line- the resulting equation would be two-dimensional (,: and x), and would be
formation problem with a background continuum has been presented in impractical to solve. Alternatively we can write an equation for S,(r); if the
(315) for idealized problems similar to those considered above in the no- line source function is known, we can compute S,,(r) = (cj,,,S,(r) + rB]/{cp,, + r)
continuum case; we shall summarize some·of the results of this work in the as desired. T~btain the equation for S1(r), the integral Jcp,,J" dx is calcula~ed
following paragraphs. from equation (11-45) and the result substituted into equation (11-11) to give
For simplicity we shall consider a semi-infinite atmosphere, and assume
depth-independence for the line and continuum opacities x, and x , the line S,(1:) = (1 - e)(l - c5) S S(t)K
00

0
1 1 ,, jt - rl dt + (1 - e) c5B SL
00

0 1,, It - rl dt + eB
profile <J,"' the line thermalization parameter e, and the Planck-f~nction B (11-46)
(all of these assumptions may be dropped if numerical techniques such as
the Feautrier or Rybicki methods are employed). The transfer equation is
where c5 =r J~ 00
cp,,(cp" + r)- 1 dx (11-47)

µ(dl,,/dz) = -(x. + x,<J,,,)I,, + x,<J,,,S, + x.B (11-41)

Defining the optical depth in terms of the line opacity, dr = - x, dz, which
K
1,r
(s) = !(1
2 - c5)- 1 f 00

-oo
E 1[(cp" + r)s]cp/ dx (11-48)
___.
implies that th~ continuum is regarded as a perturbation, and writing
r = x,/x,, equation(ll-41) becomes (11-49)
and
µ(d!,Jdr) = (<J,,, + r)(l,, - S") (11-42) The particular choice of coefficients in front of the integrals defining the two
kernel functions was made to assure normalization of the kernels.
where s" = (1 - e)<J,,,
(<J,,, + r)
f- <J, dx' + (<J," ++ r)r) B
00

00 "
.J .
"
(B</Jx
Exercise/ l-7: (a) Derive equation (11-46). (b) Show that lim,-o (6/r) = oo,
lim,-o (I - «5) = 1, and hence lim,_ 0 L 1,,(t) = 0 and lim,-o K1,,(t) = K1(t)
= (1 - e,,) f.~ 'Px·J,,. dx' + e,,B (11-43) where K (t) is defined by equation (11-19). (c) Show that K1,, and L,,, are
00
normali~d such that Jo K 1 , ,(t) dt = ½and Jo L1, ,(t) dt = ½, (d) In view of (c),
show that, for t .... oo, where S1(t) is slowly varying and hence can be removed
and e,, = (e<J,,, + r)/(<J,,, + r) (11-44)
from the integral, S1(t) --+ B.

It is easy to see that e,, is the total destruction probability for a photon at The physical significance of equation (11-46) becomes clear if we introduce
frequency x, for it is the sum of (a) the probability that a photon is absorbed the average destruction probability
in the line, <J,,,/(<J,,, + r), multiplied by B, the probability that a line photon
is destroyed by collisions, and (b) the probability that the photon is (11-50)
absorbed in the continuum, r/(<J,,, + r), multiplied by unit probability of
/ l-2 The Two-Level Atom without Continuum 353
352 Non-LTE Line Transfer: The Two-Level Atom

for then (1 - ?) = (1 - e)(l - o) and we can rewrite equation (11-46) as in the continuum will occur; equations (11-26) with e replaced by ? are
thus consistent. .
S1(r) = (1 - ~ J S (t)K
0
"' 1 1 ,, It - rl dt + (? - e)B JL 0
"' 1 ,, It - rl dt + eB To estimate the source function at the surface we can agam use the
similarity of equation (11-51) to (11-18) to infer that there exists a relation
= (1 - ~ J S (t)K
0
"' 1 1 ,, It - rl dt + ?B.rr(r) (11-51) of the form
(11-52)
In the first form of equation (11-51), the first term obviously represents where B.rr is an appropriate average of Berr(r) over a thermalization length.
scattering in the line, with a total photon destruction probability given by It is apparent from the discussion given above that whenever ? exceeds e
?. The second term represents photons fed into the line from the continuum, (either because r > e or because exten~ive line wing~ assure tha~? > e even
and the third term represents photons created by collisional processes. As though r < e), then both the destruction and creation· terms (1.e., B.rd are
in the pure-line case, the depth-variation of the source function will be dominated by the continuum, and the line is said to be continuum controlled.
determined by the properties of the kernel functions, and thermalization This implies that S1(0), in particular, is set by the continuum even thou_gh the
properties of S, can be derived from a mathematical analysis of the asymptotic line opacity and line source function in the line core exceed the_ contmuum
forms of K 1,, and L 1,, (31S). On the other hand, the second form of equation terms. This is yet another manifestation of the fact, noted earher, that the
(11-51) shows that the problem with a continuum is of the same form as the surface value of the source function is controlled by the line wings.
pure-line case, except that the coupling constant between the radiation and Numerical solutions for S1(r), in a semi-infinite atmosphere with B = 1
the thermal pool is ?, not e. Indeed, exploiting the same line of reasoning and e = 10- 6 , obtained using the method of discrete oroinates (31S), are
as before, but using? for P4 , one finds that the thermalization depth is again shown in Figures 11-5 and 11-6. It is easily seen that even modest values of
given by equations (11-26), but with e replaced by?. Values of? and A, for
e = 10- 6 and for various values of rand a, are presented in Table 11-1.
The striking result seen there is that ? can greatly exceed e, even when r ~ e,-
with a consequent dramatic decrease in A. These results- show that by
inhibiting photon transfer in the line wings, the continuum can dominate
line thermalization. Note also that whenever r ;;i: e, then ? is already suffi- ........
ciently large to assure that A < r- 1, as it must be, because we know from· CD
physical considerations than over distances greater than r- 1 thermalization 01
I
TABLE 11-1
Average Photon Destruct/on Probability? and Thermalizat/011 Dep1h A fur,, Line log S1(T)
ll'ith Overlapping Continuum(£ - 10" 6 )

0 10· 3 10-1

' ' '


r A A A

0 10·• IX 106 10-• 109 to·• 1010 -4L-......I.-....L..-L-...J..--'-_..-~----::


-2 -I 0 2 3 4 5 6 7
10·' J.79 X 10- 6 5.6 X 106 J.87 X tO·l 2,9 X 106 5.7J X to·' 3.1 X 10 7
6 log T
10·• 8.26 X J0- J.2 X 10' 5.72 X to·' 2.1 X JO' J.78 X to•• 3.2 X 10'
10-, 6.69 X JO-, 1.5 x 104 J.85 X 10- 4 2,9 X 104 5.63 X to· 4 3.2 X 104 FIGURE 11-5
10·• Line source functions in a semi-infinite atmosphere for a
5.85 X to· 4 1.,7 x 103 7.84 X to·• J.6 X 10 3 1.83 X to·l 3.0 X 10 3
two-level atom withe = 10·• and a pure Doppler profile
10-J 4.98 X 10- 3 2.0 X 102 5.23 X JO· l 3.7 X 10 1 7.25 X 10· J 1.9 X 102 (a = O), with an overlapping continuum having various values
of r. From (315), by permission.
SOURCE: (315), by permission.
354 11-2 The Two-Level Atom without Continuum 355

only the strongest lines have values of r as small as 10- 4 or Io- 3 , and
continuum terms therefore are generally dominant. On the other hand, the
H and K lines of Ca II in the solar spectrum have r :::::; 10- 9 , and hence
their thermalization is essentially independent of the overlapping continuum.

EFFECTS OF DEPTH-VARIABLE THERMALIZATION


PARAMETERS AND LINE PROFILES

In the idealized problems considered thus far in this section it has been
assumed, for the sake of simplicity, that all of the quantities B, r, 8, and cf,,
logS 1(r) are independent of depth, whereas in real stellar atmospheres they may all
be strong functions of depth. These variations can, of course, markedly
affect the line-formation process, the behavior of the source function, and
computed line profiles. To account for these complications in attempts to
model accurately real stellar spectra, recourse must be had to direct numerical
methods such as those described earlier in this chapter. But physical insight
can be gained from a study of simple parameterizations of the variations of
some of the quantities mentioned above. A discussion of the extremely
important effects of Planck function variations will be deferred to §11-3,
_3,___.__......_._....,__.,_-J._..J..._,___, where line-formation in the presence of a chromospheric temperature rise
-2 -1 0 2 3 4 5 6 ,7 will be described. Variations in the parameter r can be important, but are
log t not easy to summarize in a few words; see (18, Chap. 3) for further discussion.
FIOURE 11-6 We shall concentrate here on the effects of depth-variations of 8 and cf,,. ......
Line source functions in a semi-infinite atmosphere for a two-level The collisional de-excitation parameter e is proportional to the electron
atom withe = 10- 6 and an overlapping continuum with various density, and hence must reflect the rise of the density with depth in the
CD
CJ)
values of r, showing effects of different line profiles. From (315), atmosphere. If the atmosphere is supposed to be essentially isothermal, then
by permission. hydrostatic equilibrium implies that the total density increases linearly with I
m, the column-mass; if further we suppose that (x/p) is essentially constant,
r produce major changes in the source function; the effects are even more then tis also proportional tom. Then in an early-type star where the material
striking in finite slabs (315). The results in Figure 11-5 pertain to a pure is appreciably ionized, we expect n, to be proportional to t; in a later-type
Doppler profile; the vertical arrows indicate the predicted thermalization star the material may be neutral in the outer layers and may then abruptly
depths listed in Table 11-1, and the horizontal arrows show ?t, the value ionize at some depth, in which case a much more sudden rise in n, (and 8)
of S1(0) if B.rr = I. For Voigt profiles (Figure 11-6) a fairly good estimate may occur. Even though an actual numerical solution is required to deter-
for S1(0) is obtained if one adopts mine S1(t) with precision when 6 is variable, it is reasonable on physical
grounds that we should still expect thermalization to occur at a depth A
r (,\/«) Berr(t) dt
Berr = (Alo:)- , Jo (11-53) where P,(A) :::::; PiA). To illustrate the usefulness of this idea, suppose a
rapid rise in 6 occurs, from some value 6 1 to higher value 6 2 , at a particular
with o: chosen to be a little larger than 3; this choice allows for the strong depth t 0 in an atmosphere with r, B, and cf, all constant. Then if to > 1/e 1
decrease in the true weighting function with depth. For finite slabs, a first (for a Doppler profile) it is clear that the line will already have thermalized,
approximation to Sm.. at slab center is given by equations (11-40) with e and hence the rise in 6 has no effect. In this case S1(0) will equal 8 1½B. Con-
replaced by?; this accounts for photons emitted by the continuum. versely, suppose that to < 1/62 • Then the line source function thermalization
In real stellar atmospheres the effects of a background continuum on will have been delayed by the low value of 8 at the surface to a predicted
line-formation and thermalization are usually major. For early-type stars depth of 1/6 1 ; but at that depth 6 already equals 8 2 > 8 1 • Hence the line will
356 357

log t -4
-0.5 3

-1 0 2 3 4 5 6 7 8
(a) log r
2

-1.5

(u)

log(S,/B) -2
logS1

___ ,.. _____ 3


-1
/' / /
I I I
I I I
I I I

-2 I
,'Jo, I /to•I /10 3
2
I I I
-2 -1 0 2 3 4 5 6 7 8 I I I __..
I I I (b)
(b) logr 1061 / /
l.2.~.....::l_Q:J•:.;,,..::_::..:t:~:..,,.:.::"'i.:,,.:...,,._"'.1--1.--'--'--..J 1
co
FIGURE 11-7
0
--..J
(a) Variation of s with depth: curves are labeled with log To, and -7 -6 -5 -4 -3 -2 -1
I
t(t) .. 10- 3 (1 - 0.99 exp(-r/t0 )]. (b) Line source function in a semi- log r,
infinite atmosphere with B .. I, for a line with a pure Doppler profile and
t(T) as shown in (ai Curves labeled "a" and "b" correspond to constant FIGURE 11-8
values oft of 10- 5 and 10- 3 respectively. From (284, IOI). Line source functions in a semi-infinite atmosphere with B "". I, for a
line with t = 10-•, r = 10- 6 , a= 10- 3, and a depth-variable
Doppler width. (a) Doppler width rising toward surface,
ti.v 0 (t) = I + 2 exp( - ext,). (b) Doppler width decreasing toward
behave as if &were constant at &= &2 , and S1(0) will equal &2 ½B. These surface, ti.volt) = 3 - 2 exp(-at,). Curves are labeled wit~ parameter
qualitative expectations are confirmed by detailed calculations (284, 101) a. Curves labeled "0" have constant Doppler width. Ordmates:
such as those shown in Figure 11-7 where &is assumed to have the form Left-hand scale and solid curves give log(S1/B); right-hand scale and
&(t) = 10- 3 [1 - 0.99 exp(-t/t0 )]. Notice that, when to > 105, S1 behaves dashed curves give ti.v 0 • From (18, 52), by permission.
as if&= 10- 5 , whereas ift0 ~ 10 2, S1 behaves as if&= 10- 3 ,
A change in the form of the line profile alters the way photons get redis- constant (see Figure 11-Sa). Then we fi?d th~t the surfac7 value S,(O)_ is
tributed into the line wings, and their probability of escape. Numerical increased markedly because the broader hne-wings can now intercept rad1~-
studies (320; 18, 51) have revealed a number of interesting effects of profile tion from deeper layers. As the position of the rise in !:J.v0 moves deeper in
depth-variations upon the line source function. Suppose, for example, the the atmosphere, the surface effect diminishes, basically beca:1se the surf~ce
Doppler width rises sharply near the surface in an atmosphere with B = layer becomes more and more opaque in the wings, and the bright underlying
I 1-3 The Two-Level Atom with Continuum 359
358 Non-LTE Line Transfer: The Two-Level Atom
As before, assume complete redistribution, so that S, is given by equation
continuum radiation becomes more attenuated. At the same time the source (11-4). The statistical equilibrium equations now are
~unctio~ at g~eat ~epth also increases, basically because the higher opacity
m the hn~-~mgs m. the surface layer impedes the escape of radiation from
?elow; this 1s sOI_net1mes referred to as the reflector effect. The effect at depth n1 (B,u f cp,.J,. dv + C,u + R,. + C,.)
1s ~reater, the thicker the upper layer (until, of course, the rise lies below the
pomt ?fthe~malizati_on). Notice in Figure 11-Sa that when ex = 10 6 the upper = nu (Au, + Bu, f cp,J, dv + Cu,) + nr(R., + C,.) (11-54)
layer 1s optically thm, and thus there is essentially no reflector effect even
though the surface rise is largest.
When the Doppler width decreases sharply at the surface, the value of
~,(?) dr~ps becau~e ~he narrower profile no longer intercepts bright radiation
and nu ( Au, f
+ Bu, cp.J. dv + c., + R•• + c••)
m its wmgs. Radiation trapping is also reduced in lower layers, and hence
s, dec~eases there as well. At very great depth, however, the effect is the = 11, (B,u f cp.J, dv + C,.) + n!(R•• + c••) (11-55)
opposite, and s, actually rises a little above its value for constant 6.v
because now the decreased bandwidth available to the lines implies a d!~ for the lower and upper levels, respectively. The photoionization and re-
creased escape probability. combination rates are given by equations (5-66) and (5-67). Solving equations
It m°;St also be noted that, in addition to changing S,(t), variations in the (11-54) and (11-55) for (11,/11.), substituting into equation (11-4), and making
abso~t1on profile c?ange the run ~f optical depth, at a given frequency, with use of the Einstein relations we obtain
physical depth. In view of the Eddington-Barbier relation it is obvious that
s,
this implies a change in the "mapping" of into I (O).'and hence in the
emergent intensity distribution in the line profile, :nd its center-to-limb S1 = [f cp.J.dv + eB.(T) + e]/(1 + e + 17) (11-56)
behavior [see, e.g., (26)].
where e is given by equation (11-7),

11-3 The Two-Level Atom with Continuum (R,,. + C,,.)nr(R., + C,.) - g1(R,. + C,.)n!(R •• + c•• )/g. (l l- 5?) en
'1 = A.,[nr(R.1 + C,.) + n!(R •• + c •• )] en
THE SOURCE FUNCTION I
and
O = (2'1v
3
)(_J}_!.,_) . (R,. + c,.)11!(R •• + c •• ) (ll- 58)
The discussion given in the preceding sections of this chapter is based on - c2 g.A,,1 [nr(R., + c,.) + 11:(R •• + c •• )]
a ve~ sc~ematic and admitted~y restrictive atomic model. Naturally the
Equation (11-56) was first derived by Thomas (622) as an extension of earlier
true s1tuat1on for any real atom ts more complicated. To introduce some of
work by Milne (416, 159-164) and StrBmgren (613), and was studied exten-
the physi~lly importa?t effects whil~ retaining analytical simplicity, let us
sively in an important series of papers by Jefferies and Thomas (335; 336;
now consider an atomic model consisting of two bound states and a con-
tinuum. Th!s model provides at lea~t a rough representation of reality for 623; 337).
~eson~n~ Imes ~nd also for subordinate lines when the resonance lines are Exercise //-8: Derive equations (11-56) through (11-58) in detail.
m rad1at1ve detaded balance. The addition of the continuum greatly increases
the number of processes that may take place. In the strict two-level case Despite the apparent complexity of these expressions, each term admits
t?e only processes that can occur are photoexcitations or collisional excita~ a simple interpretation. Consider the numerator of equation (11-56). The
t1ons from the lower to upper state, and their inverses. Now there are in first term again represents the scattering reservoir. The second term is the
addition, p~otoionizations and collisional ionizations from the bound le~els thermal source, giving the rate at which photons are created by collisional
to the continuum, and radiative and collisional three-body recombinations excitation; note that this term depends upon the local value of T,, the
to each bound level. It is clear that this is a much more general model, and electron kinetic temperature. The third term is proportional to the total rate
we shall find that the additional physics has major implications for the line- at which electrons are ionized from the ground state to· the continuum,
formation process.
360 Non-LT/:,' Line Transfer: The Two-Level Arom //-J The Two-Level Atom with Continuum 361

times the fraction that recombines to the upper state and thus becomes new terms do not couple S1 to the local thermal pool, characteriz.ed by T,,
available for emission into the line by radiative decay to the ground state. as do collisions, but rather to a radiation temperature that may be markedly
Similarly, in the denominator, the second term accounts for photons de- hotter or cooler than the local kinetic temperature.
stroyed by collisional de-excitation of the upper state. The third term is a
new sink term that is proportional to the total ionization rate from the upper CLASSIFICATION OF LINES
state times the fractional recombination rate to the lower state; this term
clearly accounts for the destruction of potential line photons via the con- In view of the fundamentally differing natures of the various source and
tinuum processes. We thus find all of the basic physics of the situation sink terms that appear, we may expect that line-transfer problems may show
represented clearly in the source function. substantially different characteristics depending on which terms provide the
Further insight follows from rewriting 8 as rJB*, where B* is found to be dominant contributions to S1• Thomas has suggested (611; 616, 174) a
classification of lines into broad categories by consideration of which terms
B* = (2h:3){(nruu)[<R.K + c.K)(RKI + c,K)]- 1}-I (11-59) are the largest for different atoms in typical stellar-atmosphere regimes.
C n:g, (R,K + c,K)(RKU + c.K) For example, if e > '1 and eB, > rJB*, the line is called collision dominated;
here S couples to T,. On the other hand, if '1 > e and rJB* > eB, the line is
Exercise I 1-9: Verify equation (11-59). photoi~nization dominated, and the line couples to a char_acteristic T, '# T •.
Intermediate cases where, say, e > rJ but rJB* > eB, or vice versa, are called
It is clear that B* bears a formal resemblance to the Planck function, and, mixed domination lines. The recognition of these classes of lines represents a
in fact, we can think ofit as B• = B,(T,), where T, is a characteristic radiation considerable advance over the rather ill-defined classical division of lines
temperature whose value is set by the photoionization and recombination into "absorption" and "scattering" categories, and has led to impor~ant
rates in the two bound-free continua. This radiation temperature can, in insights about line-formation. In particular, the emerge~t profiles for h~es
general, be quite different from T •. At great depth in the atmosphere where in these classes are quite different when a chromosphertc temperature nse
the continua are optically thick, J, .... B,(T,); then R,K ....
RrK =;. R:,, and occurs in the outer layers of the atmosphere.
R.K .... R:K = R:u,so that B* -+ B,(T.), as expected, and S1.... B,(T,). But The category to which any particular line belongs depends (via the atomic
near the surface, the continua may become transparent (while the !in~ cross-sections involved) upon the structure of the ion from which it arises,
remains optically thick) and the continuum radiative rates become ess1:ntially and upon the structure of the atmosphere (because of the dependence of 0)
fixed. Depending on the relation of the radiation temperature (which char- the relevant rates upon atmospheric parameters such as temperature, density, co
acterizes J, emerging from t, ~
1) to the local electron temperature, quite and incident radiation fields). Different lines of the same ion will, in general, I
distinct situations emerge. For example, suppose that collisions are negligible, fall into different classes, and a comprehensive a priori classification is not
and that R,K RKI
> while RKu> Rn (i.e., suppose that a relatively cool possible; rather, each case must be examined in turn. Broad groupings of
medium is irradiated by a "hot" radiation field in the ground-state con- lines in a solar-type atmosphere have been suggested by Thomas; these are
tinuum). Then it is clear that B* > B,(T,), and that S1 will be larger than the displayed in Table 11-2, and can be understood qualitatively as follows. The
value it would have had by coupling to the thermal pool; the excess emission resonance lines of the singly ioniz.ed metals are collision dominated because
comes from a preferential depopulation of the lower state into the upper via the excitation energy is only a few electron volts (compared to a thermal
the continuum, followed by radiative de-excitation in the line. If the line is an
absorption line, an increase in S1tends to weaken it; taken to the extreme limit, TABLE 11-2
Ca1egories of Line Source Funclions in a Solar-Type Atmosphere
this is the mechanism that produces the photoionization-recombination-
cascade emission spectrum in a nebula. If the inequalities posed above are Collision dominated Photoionization dominated
reversed, then the upper state is selectively depopulated into the lower, and
S1 decreases; if the line is an absorption line, it will strengthen. Resonance lines or singly ionized Resonance lines or neutral
metals (Mg•, Ca•, Sr•, etc.) metals
The two essential points that have emerged from the above analysis are
Resonance lines or H and other Hydrogen Balmer lines
the following. (a) The source function for a given line contains terms in
nonmetals (C, N, 0, etc.)
other transitions (the two continua for the present model). This result is
quite generally true and carries over to the multilevel case (see §12-1). (b)The SoURCE: Adapted from (623; 626, 174).
362 Non-LTE Line Transfer: The Two-Level Atom 11-3 Tire Tll'o-Level Atom with Co11tin1111m 363

energy of about 0.5-1 eV), while the ionization energies are 11-15 eV; recall- which provides an approximate representation of a temperature distribution
ing that both the collision and photoionization rates scale as exp( - E0 /kT), that has: a uniform gradient in the photosphere; a plateau at a minimum
it is plausible that the collisions should prevail. Even for hydrogen, where temperature at some characteristic depth (they used Tmin ~ 4000°K at tr ~
E 0 in I.A is 10 eV, the collision rates dominate because the solar u.v. radiation 10- 2 ); and then a steep outward gradient mimicking the sharp chromospheric
is weak and the photoionization rate is small. In contrast, for the hydrogen rise. The transfer equation was solved in the Eddington approximation:
Balmer lines and the neutral metals, the relevant photoionization edges lie
in regions where the solar radiation field is intense, and the photoionization
rates prevail over collision rates. The sodium D-lines are an exception
because of an unusually large collision cross-section (284, 333; 284, 347);
this example serves as a warning that the broad classes listed are illustrative, using the method of discrete ordinates, for a Doppler profile and for typical
and detailed analysis is needed in each case. In a higher temperature regime values of r, e, I'/, and n•.
(e.g., in an 0-star) some of the remarks just made are no longer valid, and Results for a collision-dominated line (I'/ = 0) withe = 10- 4 , r = 10- 4 ,
lines switch categories. For example, the radiation field in the Lyman con- and various choices for the continuum source function are shown in Figure
tinuum becomes extremely intense (the hydrogen is virtually completely 11-9. Note that at great depth S1 thermalizes to Sc, but as the wings begin to
ionized), and the hydrogen Lyman lines become photoionization dominated. become transparent (tc :$ 1), S1 drops below Sc. Proceeding outward, Sc
The higher subordinate lines now have energies E0 ::5 kT(~3 eV at 0-star rises very steeply at tr ~ 10- 2 • The line source function tries to follow this
temperatures), and are relatively weak, hence are formed in deeper, denser rise through collisional coupling, and actually does increase outward, but
layers; this tends to lead to collision domination of the subordinate series. ultimately the effects of scattering dominate, and, at the surface, S1 lies about
three orders of magnitude below B,(T.). It follows from the Eddington-
LINE-FORMATION IN THE PRESENCE OF A CHROMOSPHEII F. Barbier relation that the depth-variation of S1(t) should be reflected in the
frequency-variation of the emergent intensity. The computed emergent
The great physical importance of the division oflines into the two broad
categories described above becomes strikingly apparent when we consider
the nature of the source function variation, and emergent line profile, in an, 5 .......
atmosphere with a chromospheric temperature rise outward. The basic 4 s, co
point is that collision-dominated lines are linked to the local electron kinetic ............. 0
temperature, while photoionization-dominated lines are not. The latter couple
3
- \
\
I
instead to a radiation temperature characteristic of the energy distribution 2
In S/S 1
(emitted at some other point in the atmosphere) in the continua of the upper
and lower levels. We therefore expect collision-dominated lines to be at /,..
,,..✓
least partly responsive to the local temperature, and to exhibit, in their 0
profiles. features attributable to the outward temperature rise. On the other -I
---- s,
_.,

hand, we expect photoionization-dominated line-profiles to be insensitive


to variations in local parameters (in particular, to variations in T,). The -2
strong dichotomy of behavior just described was clearly demonstrated in -5 -4 -3 -2 -I 0
the pioneering work of Jefferies and Thomas (336). They showed that these
log t,
considerations explained the observed presence of emission cores in lines
FIGURE 11-9
such as the Hand Klines of Ca+ (and the analogous lines of Mg+), and their Line source functions for a collision-dominated line
simultaneous absence in the hydrogen Balmer lines. in a semi-infinite atmosphere with a chromospheric
In their work Jefferies and Thomas adopted a schematic continuum source temperature rise. Upper curves show continuum
function of the form source function S,. and lower curves show
corresponding line source functions S1• In all cases
(11-60) £ = 10·• and r = 10·•. From (336) by permission.
364 365

0.0
4
o.s 3 FIGURE 11-11

,,,-"..,.--.....,,
lnS/S 1 2 Source function for a photoionization-
1.0 A dominatcd line in a semi-infinite
(/, - 1,)/S1 .,,,,,,·.,./ ~ ____.,.,.,, .,,/ ., atmosphere with a chromospheric
I.S .,.,,,
..,,,. ,'
,' 0 temperature rise. From (336), by
permission .
........ - - • • - ' ' I
_,.____,__.__.,_....,_--L........._
~/ -5-4-3-2-1 0
2.0

---- -------
2.S c..:.==-::.=.:._ __,___ _ _ __,___ _ _ _.,___ _
-----------
-- _.,. /
log r,

0 3
rapidly to its chromospheric value). The line profile is a pure absorption
feature with no hint of central emission reversal. This behavior is precisely
MUUllll ll-10
what is observed for the Balmer lines and, in fact, the run of S1 correlates
Linc profiles for a collision-dominated line in a semi-infinite atmosphere with a
chromosphcric temperature rise. Curves arc coded to correspond to those in well with the empirical source function deduced by Athay and Thomas (29).
Figure 11-9. From (336), by permission. The recognition that the Ca II H- and K-lines have source functions that
couple partially to the local Planck function variation led to the realization
that the emission features often observed in the cores of these lines in stellar
intensity is shown in Figure 11-10; only half of the profile (which is symmetric spectra contain valuable information about the temperature distribution in
about line-center) is shown. Note that the lines show an emission peak near stellar chromospheres. If suitably analyzed, these lines provide unique diag-
± 1.5 Doppler widths, with a central absorption reversal; lines of this form nostic tools for the determination of the physical structure of the outer layers
are called doubly reversed. The overall variation of intensity within the line of solar-type stars. It is known that the intensity of the emission correlates
is in good qualitative agreement with the observed behavior in the solar with stellar age (67S; 677), and this fact offers the possibility that by combining
_...
Ca II H- and K-lines and the corresponding lines of Mg II. Numerous other. accurate chromospheric diagnostics with a reliable theory of stellar evolution,
calculations, with a chromospheric temperature rise, for other choices we may be able to infer the time-evolution of the solar chromosphere. One co
_...
(sometimes including depth-variations) of e, a, r, </>., and different para- of the most fascinating empirical results relating to chromospheric Ca II H
meterizations of B,(t) can be found in the literature [see, e.g., (284, 101; and K-line emission is known as the Wilson-Bappu effect (676), which shows
26; 27; 18, 45-48)]. All show the same kind of behavior as that shown in that the half-intensity-width of the emission components correlates closely
Figure 11-10, and some yield semiquantitative agreement with the solar with stellar luminosity, over a range of 106 in L. Several theoretical sug-
data; excellent quantitative fits to the observations are obtained when multi- gestions have been made to explain this observation, but, as was early
level calculations, using realistic atmospheric models and atomic models for emphasized by Jefferies and Thomas (337), an understanding of the phe-
Ca+, are employed (see §12-1). nomenon must be based on an accurate physical picture of line-formation.
The behavior of photoionization-dominated lines, for the same assumed Their work, and more recent efforts, have elucidated the dependence of the
Planck function, contrasts strongly with that of collision dominated lines. emission intensity and width upon the form of Te{t) (amplitude and depth
Results for a line withe = 0, '1 = 10- 2 , and r = 10- 4 are shown in Figure of rise) and upon the functions e, r, and <J> ••
11-11. At very great depth, S1 -+ s. when the line thermalizes in the con- Some useful insight concerning the basic properties of chromospheric
tinuum. Proceeding outward, S1 remains fixed at the value set by B•, and at emission lines can be gained from approximate scaling rules [(24;, 18, 46)]
a line optical depth of order 1/'1, shows a strong drop outward, characteristic such as those summarized in Table 11-3. These give estimates of the surface
of scattering, to a value equal to '1½ B•. The source function shows no response value, S1(0), and peak value, S,(max), of the source function in a chromo-
whatever to the variation of B,(T.). Note that proceeding outward, S1 first spheric slab, with a Planck-function variation given by equation (11-60)
lies aboves. (because B• has a radiation temperature set in the continua in with S 1 set to unity. The rules are based on the assumption that the Doppler
deeper, hotter layers), and then, near the surface, lies below S, (which rises width is independent of depth; Doppler width variations alter the results
366 I J-4 Static Extended Atmospheres 367

TABLE 11-3
Scali1111 Rules for L/11e Source Fun,·rion in Fin/le
0.3 and thus the coupling of the Ca II lines to the temperature rise is relatively
Chromospherfc Slabs inefficient. In contrast, for the Mg II resonance lines, r is about a factor of
10 smaller (because Mg is about 10 times more abundant than Ca), and the
Doppler profile Voigt profile other parameters are about the same; in this case we obtain a much stronger
(a<£) e< a< I coupling to the rise and much brighter emission, as observed. It ~ho_uld be
Chromospheric
thickness S1(0) S1(max) S1(0) S1(max) noted that the arguments given about are meant only to be qualitative, for
the depth-variation of the parameters is actually fairlr com~licated, and ?ot
Optically thin well-represented everywhere by the values chosen; m particular, equat,1on
(yr)" 1 < I ~fl ~fl ~p w (11-60) provides a relatively poor fit to the actual run of B.(t). Detailed
Effectively thin analyses of Ca II line-formation in stellar chromospheres have been per-
l<(yr)- 1 </\ W!yrJ-t W!yri- • ~p(yra)- 1 W(yra)-t
formed using' realistic atomic and atmospheric models (56; 58), and a
Effectively thick
I\ < (yrr I ~'P p ~'P p physically plausible explanation of the Wilson-Bappu effect has begun to
emerge (57).
SOURCE: Adapted from (18, 46), by permission.

somewhat (24) and, more importantly, strongly affect the emergent line- 11-4 Static Extended Atmospheres
profile for a given S1(t). The parameters of interest for describing the run
of B,(Te) are {J, the amplitude of the rise (assumed to be » 1), and y- t, the As was described in §7-6, many stars have extended atmospheres whose thick-
continuum depth at which the rise occurs [the corresponding line depth is nesses are comparable to the radius of the star. We shall suppose that, to a
(yr)- 1 ]. If (yr)- 1 « A (the thermalization depth), then S1 will-respond only first approximation, these atmospheres are sph~rically symmetric.. For the
weakly to the chromospheric temperature rise; but if the inequality is re- purposes of the present discussion we shall ~ons1_der the st~ll~,r radius r • ~o
versed, a strong effect will occur. If the slab is effectively thick [i.e., (yr)- 1 > A, be that of the surface on which t R ~ ½. We 1magme that this photosphenc
......
where A = ?- 1 or a?- 2 for Doppler and Voigt profiles, respectively, and ? core" is surrounded by an envelope of large size, within which unit optical co
is the total destruction probability given by equation (11-50)], then s, depth in the most opaque spectral regions (i.e., line centers) is encountered I\)
saturates to {J at depth and falls by a factor of ?t at the surface. If the slab at radii R » r •· In reality, virtually all stars with very extended _atmosph~res I
is optically thin, then S1 just equals the local creation term ?/J. If the slab is also have large-scale velocity fields (usually overall atmospheric expansion)
effectively thin, the maximum value of S1 is given by equations (11-40) with that strongly affect-indeed dominate-transfer in the lines, so that. the
T = (yr)- 1 • Note that the quoted results for the Voigt profile presume that assumption of a static atmosphere is physically less useful in an analysis of
a(yr)- t > 1; if a is so small that this is not true, the line wings are negligible line-formation than it was for the continuum. Nevertheless, there are some
and the results listed for a Doppler profile apply instead. The surface value extremely important effects of a fundamentally geomet~ic or!gin that en~er;
of S1 can be derived by recognizing that if Sm•• oc ?/J(N) (where (N) gives it is worthwhile to examine these here, and to defer a d1scuss1on of velocity-
the mean number of scatterings needed to escape), then from random-walk field effects until Chapter 14.
arguments we expect S1(0) oc ?/J(N)t. Note that the effectively-thin results As seen by an outside observer, the size of the emitting surface where
merge smoothly with· the effectively-thick results when (yr)- 1 = A, and with r = I at the more opaque frequencies, in particular the cores of spectral
the optically-thin results when (yr)- 1 = I. 1{nes, can be much larger than that of the continuum. Then the line has a
From the results of Table 11-3 we see that, with a given chromospheric larger effective emitting area, and if_ w~ assume LTE an~ sup~ose that t~e
structure, some collision dominated lines will respond strongly to the envelope is essentially isothermal, 1t ts clear that the )me _will appear m
temperature rise, while others may not, depending on the values of e, r, and a emission relative to the continuum. This behavior contrasts with the result
appropriate to them. For example, for the Ca II lines in the solar chromo- for an isothermal planar atmosphere where the line is neither in absorption
sphere we have (to order of magnitude only!): a ~ 10- 3 , r ~ 10- 9 , or emission. In fact, the basic geometric effect just described is actually the
~
e ~ 3 x 10- 4 , and y 106 ; therefore A~ 104, while (yr)- 1 ~
103, so the primary mechanism that produces the extremely intense emission in v~ry
lines are effectively thin. Furthermore a(yr)- 1 ~ 1, so (SmaJ/J) .~ e(yra)-t ~ opaque spectral lines observed-e.g., in Wolf-Rayet spectra. The assumption
'
369
368 Non-LTE Line Transfer: The Two-Level Atom

that the atmosphere is essentially isothermal implies a nonradiative source


of energy input, for we have seen in §7-6 that in radiative equilibrium (in
LTE), J ~ B ~ r- 2 in an extended atmosphere; thus the situation just
described may be a bit extreme (though we have no strong physical reason -0.S
to prefer radiative equilibrium-recall the solar corona!). Also, if the line
source function has a scattering term, we expect S1(t) to decrease outward
(in fact the drop is enhanced by extension effects; see below). However, it is ,
clear that, for any given S1(t), the change in effective emitting area from line
core to wing will always tend to increase the core emission, relative to the
continuum, compared to the value it would have had in a planar atmosphere.
Absorption lines will tend to weaken, and emission lines will tend to grow -1.5
brighter, as atmospheric size increases.
A second effect of atmospheric extension is a systematic increase of the
escape probability at a given (radial) optical depth. It is obvious from ele- logS1 -2,0
mentary geometric considerations that along all rays, except the single ray
at µ = + 1, the optical path-length from the test point to the surface is
smaller in a spherically symmetric atmosphere than in a plane-parallel semi-
infinite atmosphere of identical properties (i.e., same run of physical variables
along the radial direction in the two cases). Thus, in view of equation (11-22),
P.(t) increases, and we expect S1(t) to decrease accordingly.
A third effect, which also increases the effective probabifity of escape, is
the systematic bias toward larger radii in the scattering. Suppose that the
scattering process is isotropic and that the material is homogeneous, so that
the photon mean-free-path is the same in all directions, and that l = 1 de.!
fines a spherical volume around the test point. Then, from the basic geometry co
of the situation, it is easy to see that, in a spherical atmosphere, more photons w
end their flights at larger radii (hence closer to the surface); in contrast, in a -1 0 2 3 I
-2
planar atmosphere the probabilities that the photon ends its flight at a logt
greater or smaller depth (by a given amount) are equal. If the material has
an opacity that decreases strongly outward, the bias is enormously enhanced FIGURE 11-12 r d' R
Line source runctions in a spherical atmosphere o outer ra iu~3 d
[see Table II and Fig. 2 of (374)]. This bias implies that the net chance of , . s or r _ I) B = l £ = constant, T, = 0, T, = 1 • an
escape is enhanced even further. (m,unit oc' r-:.2, Curves ar~ labeled with R, and results_ror tw~ values
The transfer equation for the two-level-atom line-formation problem ~~:c~~r!~sponding to ctTcctivcly thick and ctTcctively thin media, are
[i.e., S1 given by an equation of the form of(ll-56)] in spherical geometry is sho~n. from (374), by permission.
easily solved by the methods described in §7-6 [see equations (7-190) through
(7-208) and related discussion]. Calculations for idealized models (374) values of & chosen yield effectively thick and effectively thin.media. The m;j~r
similar to those employed earlier in this chapter provide illustrations of the
concepts outlined above. We characterize the atmosphere by its outer radius effects of increased photon escape in decreasing s,l~re ~ead~ly ~~~~r~\02) i:
R (in units of the core radius r, = 1), a total line optical depth T,, a continuum
tha for the effectively thick medium, the therma 1zat1on ep ~
sca;cel affected by sphericity. Emergent flux profiles for these source fu?c-
optical depth T., and opacities x, cc r- 2 and x. cc r- 2. We set & = constant, . y h . F' ure 11-13 There we see that the central reversal, which
and B = 1. Source functions for an envelope with no background continuum, ttons ares own m ig · · h ~ very
is a prominent feature in the planar limit {note log scale), vams es or
surrounding an empty core {i.e., a nebula) are shown in Figure 11-12. The two
370 371

0
1.2
-1

-2 1.0

-3
0.8
-4
F,
-S 0.6

-6
0.4
-7
log F,
-1
0.2
-2

-3 0
-3 -2 -I 0 2 3
-4 X

11-14
-s flGURE
Emergent flux profiles from spherical atmospheres with B = I,
e = 2 x 10· 3 , T, = 2, T, = 10 3, and x oc ,- 2 • The source
-6 functions in these atmospheres are shown in Figure 14-1 I.
-7

-8
__.,
11-5 Comments on LTE Diagnostics
-9
From the developments presented in the preceeding sections of this chapter
-10
0 2 3 4
we have obtained a very different, and physically a far more satisfying,
X
picture of line-formation from that based on the assumption of LTE. The
FIGURE 11-13 full multilevel problem must be solved before close quantitative agreement
Emergent flux profiles from spherical atmospheres with same properties with observation can be obtained. It is, nevertheless, worthwhile to con-
as used in Figure 11-12. From (374), by permission. solidate some of the changes in perspective inherent in the new conceptual
framework fashioned above, by summarizing here a few of the important
differences between the LTE and non-LTE methods of analysis, and by
extended atmospheres; this is a manifestation of the greater effective emitting stressing the implications of these differences for the reliability of diagnoses
area in the line core. Analogous results for atmospheres with Tc = 2 and of physical conditions in a stellar atmosphere. An extended discussion of
c = 2 x 10- 3, and otherwise identical to those mentioned above, are shown many of these points can be found in (626).
in Figures 14-11 (source functions; note only curves with Vmax = 0) and 11-14 In LTE line-formation theory, the source function is linked uniquely to
(line profiles). Again, the source functions are dramatically reduced by in- the electron temperature T., and the line profile reflects the depth-variation
creased escapes, and the line profile transforms from a pure absorption line of the Planck function, to within the limits of resolution set by photon
to an emission line with a central reversal (the reversal results only because diffusion over a single mean-free-path. In contrast, in the non-L TE theory
S1 has decreased so much at t = 1; if S1 remained unchanged, the whole S1(,) is no longer tied to T,(,); instead, it is fixed by the scattering t~rm,
line would come into emission). which results from the interplay of the rates of photon escape, destruction,
372 Non-LTE line Transfer: The Two-level Atom //-5 Comments on LTE Diagnostics 373

and creation, by all mechanisms. Of these, only collisions couple the creation S,(t) is often markedly different from B.(T); thus the emergent intensity
and destruction events to the local value of T,. The other photon sources and within the line (and hence its equivalent width) must differ from its LTE value.
sinks (which may far overshadow the collisional terms) couple to radiation In a general way, the central depths of collision-dominated lines are increased
fields in spectral regions outside of the line. Thus far, only continuum fields by departures from LTE, and the equivalent width is increased (which im-
have been considered; but, as will be seen in Chapter 12, al/ lines in the entire plies that the abundance require~ to fit an observed width ~ill d~crease).
transition array of the atom are, in principle, involved. To be sure, in the For photoionization-dominated Imes, non-L TE effects may either increase
classical theory there was some flexibility implicit in the division of lines into or decrease the linestrength; a useful summary of expected consequences of
"absorption" and "scattering" lines. But this division was ad hoc, and actually departures from LTE on abundances is given in (225).
misleading, for we have seen that all line source functions have scattering' Until recently the question "Do departures from LTE actually lead to
terms, and that the important question is, "What are the sources and sinks significant errors in stellar abundance analyses?" remained open, and was
that dominate in fixing the level of the scattering term?" The classical theory the subject of vigorous and lengthy debate in the literature. As we shall see
did not clearly recognize the photoionization category, and was essentially in §12-4, the question can now be answered in the affirmative in many
unprepared to cope with the mutual interaction of several lines. important cases [ see also (435)]. This is not to say that LTE abundance
One of the important implications of the above-mentioned changes in the estimates are always in error; in many cases they are not. It is, however,
theoretical structure is that, ifs, is no longer uniquely specified by the run apparent that it must be shown, and not ~ere!y assumed, _that the applicati~n
of T,(t) (and, in fact, it may be totally unconnected with T,), then it is clear of LTE provides an adequate approx1mat1on to reahty for any case m
that we cannot hope to infer T,(t) from the emergent intensity l(t = 0, v). question.
In particular, we have repeatedly found that S1(0), and hence the central Another parameter often inferred from curve-of-growth analyses is the
intensity of the line, are almost completely unrelated to the "boundary characteristic "microturbulent" velocity in the atmosphere. Leaving aside
temperature" T 0, but are determined by transfer over an entire thermalization the questions raised about whether _a simple :urvc~ of growth adequa~ely
length, by photons fed into the line wings. Thus the literature containing describes the complex problem of !me-formation m a turbulent medmm
estimates of stellar boundary temperatures, obtained from line-core intensi- (discussed in §14-1), it is important to note that the diagnosis of this particular
ties using an LTE theory, is little more than a collection of elaborate "maps" parameter is especially vulnerable to error from non-LTE effects. The
of nonexistent "territory." . position of the flat part of the curve of growth depends sensitively upon the ......
Extending this conclusion, we notice that with an LTE theory, to explain way, and the distance over which, the line saturates to its thermal value, for co
the observed emission cores in the Ca II H- and K-lines we would have to these characteristics of the solution determine the depth and width of the 01
propose a nonmonotone temperature distribution r:(z), which first de- line. For example, as remarked earlier in §11-2, even though coherent scat- I
creases outward, then rises, then drops again. Not only would this be in- tering may produce a line of the same depth as noncoherent scattering (with
consistent with the distribution T,(z) inferred from (LTE !) methods using the same value of e), the line will be wider in the latter case, and therefore
infrared and continuum data [ which show a photospheric drop to a minimum the equivalent width larger. Calculations, for idealized model atmospheres
plateau, followed by the chromospheric-coronal rise-cf. Figs. 7-31 and and two-level atoms, have shown important effects of departures from LTE
7-32], but also the particular distribution inferred from the Ca II lines would upon the flat part of the curve of growth (25), and recently (cf. §14-1) non-
not yield a fit to other lines (e.g., of Mg II) showing similar reversals. Worse, LTE theories that include stochastic velocity fields in the line transfer have
the use of such a temperature profile would produce emission reversals in been developed. Although there is little doubt that mass motions do exist
all lines of sufficiently great strength, even those where none are observed in stellar atmospheres, the accuracy of the actual values assigned will remain
(e.g., Ha.). The entire problem vanishes when we discriminate between the in doubt until internally-consistent analytical methods are employed.
collision-dominated and photoionization-dominated classes; and further, We now turn attention to the problem of attempting to match observed
the new approach allows us to understand the varying degree of the effective- stellar spectra, using realistic multilevel model atoms and detailed model
ness of the coupling of S, to T, for different collision-dominated lines. In atmospheres, in a physically self-consistent solution of the full transfer and
short, the non-LTE analysis leads to an enormous improvement in our statistical equilibrium equations.
conceptualization of the physical situation.
One of the primary applications of LTE line-formation theory has been
the estimation of stellar abundances. As we have seen, the depth-variation of
Non-LT£ Line Transfer: The Multilevel Atom 375

Moreover, it is essential to realize that this increased complexity is not

12 merely one of having to treat larger sy!\tems of equations, but that there are
extremely important new physical effects involved. In particular, we shall
now find that the radiation field in any one transition may affect that in every
other transition in the atom, and that in many cases these effects are of over-
whelming importance. Furthermore, the way in which the effects manifest
Non-LTE Line Transfer: themselves is often very subtle, being the result of extremely complicated
chains of interactions. This mutual interaction of line radiation fields (and
source functions) is referred to as interlocking, and the successful treatment
The Multilevel Atom of interlocking effects poses the central obstacle to the solution of multilevel
transfer problems.
The most straightforward approach to the multilevel problem proceeds
from a rather direct extension of the techniques employed in the previous
chapter for a two-level atom. Here one writes analytical expressions for the
source function in each line in such a way as to isolate explicitly the radiation
field in that line, and solves the corresponding transfer equation holding all
other terms fixed. This is the equivalent-two-level atom approach, in which
only one line at a time is considered in the solution of the transfer equation,
and the interactions among lines are treated by iteration. In the method just
described, one has tacitly assumed that the coupling between lines is fairly
weak in some sense; but often this is not the case, and the equivalent-two-
level-atom approach then becomes unsatisfactory. For example, the indi-
vidual line components of multiplets usually have, as their initial and/or final .....
states, closely-spaced levels that can interact ·physically, very strongly. In co
such situations, photons may switch from one line to another in the multiplet, (J)
and the radiation fields in the lines become strongly interlocked. More I
generally, in chains of transitions within a complicated transition array,
The discussion presented in the previous chapter, based on the use of a conditions often arise under which photons are rather freely converted from
highly simplified atomic model in idealized atmospheres, affords deep insight one transition, with its corresponding spectral region, to another. In effect,
into much of the physics of spectrum line-formation. But to analyze real the photon no longer belongs uniquely to a specific line, but, to a certain
stellar spectra, it is now necessary (1) to consider much more realistic atomic extent, belongs to the ensemble of radiation fields of the entire set of transitions
models, having many levels (perhaps spread over a sequence of ionization of the atom. The profound importance of this point was recognized and
stages), giving rise to a multiline transition array; and (2) to solve the com- emphasized by Jefferies (334, Chap. 8; 284, 177), who advanced the appealing
bined transfer and statistical equilibrium equations for such atomic models, picture that the photons should be considered to be interchangeable members
in fairly elaborate atmospheric models that attempt to describe the physical of a collective pool. Viewed in this light, it is apparent that it is essential to
structure of the star with a high degree of realism. Throughout the discussion treat all of the lines and their interactions simultaneously, to a high degree
we shall assume that we are dealing with an "impurity" species (i.e., other of consistency. This may be done quite directly in special cases (e.g., in mul-
than hydrogen) that has no significant influence on the structure of the tiplets with a common lower state). In the general case, strict consistency
atmosphere, and shall therefore regard the atmospheric model as given and can be achieved by the complete /inearization method; this scheme allows
fixed. It is obvious that with a larger number of levels, the number of transi- fully for all interlocking effects from the outset, and may be considered as the
tions and interactions that are possible increases greatly, and both the mathematical realization of Jefferies's physical conception of the collective
physical and mathematical nature of the problem become more complex. photon pool.
376 Non-LTE Line Transfer: The Multilevel Atom 12-1 The Equivalent-Two-Level-Atom Approach 377

12-1 The Equivalent-Two-Level-Atom Approach and for the upper level we have

FORMULATION

In our study of the two-level atom, we made use of the statistical equilib-
rium equations to eliminate analytically the population ratio appearing in
the line source function, and thus to obtain an expression of the general form
- n, (B,. f <f,.J. dv + C,.)
L n1A1.Z1u + u>l¢1
= n~(RKU + c.K) + u<J ·L n,C,.Y,. (12-3)
(12-1)
In equations (12-2) and (12-3) we have used the net radiative bracket Z 11
where oc and /J describe the possible ways of creating or destroying a photon. defined by equation (5-58), the net collisional bracket Y, 1 defined by equation
In writing equation (12-1), explicit use is made of the fact that for a strict (5• 71), and the radiative ionization and recombination rates defined by
two-level atom, there is only one line, hence only one radiation field ·or equations (5-66) and (5-67); further nr
denotes the LTE population of level I,
relevance. Indeed, we saw that if the atom also has a continuum, then the given by equation (5-14), at the actual value of the ion density.
radiation fields in the two bound-free transitions appear in the terms oc and fl. If we solve equations (12-2) and (12-3) for the ratio (n1g./n.g1), substitute
Because the line is usually much more opaque than the continua, the latter the result into the relations,.= (2hv 3/c 2)[(n,g./n.g1) - 1]- 1, and make use
will often be transparent at the depth of line-formation, and the continuum of the Einstein relations among the transition probabilities, we find
rates then will be.fixed (though this is not always the case). If other levels and
other lines are included, it is to be expected that the radiatfon fields in these
transitions will also appear in the terms oc and p, and, in addition; that these
terms will vary (i.e., cannot be fixed a priori) in the region of line formation. __.
Nevertheless, it is obvious that, however strong the coupling of a given lirte where e' is defined as in equation (11-7), while (0
to other lines may be, the line must always at least respond to the collisional
rate coupling the two levels in question, and to the continuum rates from these (12-5) --..J
levels. This suggests that, as a computational strategy, one might attempt to I
write a mathematical expression for the source function in which the line- and (12-6)
scattering, collisional interaction, and continuum radiative and collisional
rates from the two levels forming the line appear as direct rates, while all where, in turn,
other rates are grouped analytically into net rates; one hopes in this way to
minimize the effect of interlocking at each stage of the calculation. a, = R,K + c,K + L AuZu + l<J,i,u
L C,1Yr1 (12- 7)
Consider a line formed between levels I and u. The rate equation for the l<I

lower level is
a2 = nt(R,c1 + c,K) + L
l<),f,u
n1A1,Z1, + L n,C,, Y,,
l<I
(12-8)

a3 = RUK + c.K + L
u>l,f,I
A.,z., + L c.1Yu1
u<)
(12-9)

- n. ( A., + B., f <f,.J. dv + c.,) (12-10)

= nt(R,c1 + c,K) + L n1A1,Z1, + L n,c,, Y,, (12-2)


I<}¢• l<I Exercise 12-1: Verify equations (12-4)-(12-10).
3711 Non-LT£ Line Transfer: The Multilevel Atom 12-I The Equivalent-Two-Level-Atom Approach 379

It is clear that the terms o 1 and o2 represent, respectively, the loss rate In a calculation designed to simulate conditions in a real atmosphere, it is,
from level / to all levels other than u, and the number of electrons fed into in general, necessary to account for the variation in the ionization state of
level / from all levels other than u; a3 and a 4 denote similar quantities for the particular atomic species under consideration (perhaps over a sequence
level u. of several different ionization stages). Although continua arising from excited
The resemblance of equation (12-4) to the two-level form is obvious. Note states of "impurity" species will generally be transparent (compared to the
that the populations of all other levels appear in the terms a 2 and 04 • This dominant absorption and emission terms from H and He), so that the radia-
occurs becau_se _we have used only two of the entire se_t of rate equations, and tion field may be considered to be fixed, this situation will not usually be
hence c~n ehmmate only n1 and n•. It is possible, in principle, to obtain an true for ground state continua, which normally will be opaque enough to
expression for the source functions,. that does not involve any of the occupa- determine their own radiation fields. Thus it is necessary to supplement the
tion numbers explicitly; here the full set of rate equations .riln = f!JJ is line transfer equations (12-12) with corresponding continuum equations, as
manipulated to give analytical expressions for the quantities ex and p in described in the discussion of the formation of the Lyman continuum in
equation (12-1), in terms of the cofactors of .ril [see (333; 284, 187)]. Howe~er, §7-5 [see equations (7-127) through (7-131) and (7-135) through (7-144)].
the algebra_ in such a ~r~cedure quickly becomes hopelessly unmanageable; The complete set of rate equations for a particular impurity species will be
and, ~ore important, _it 1s not at all obvious that anything is gained, for the composed of as many equations as there are bound levels (over all stages
result mg source function then is nonlinear in the radiation fields of all other of ionization taken into account), plus a final equation setting the sum of all
transitio~s. Moreover, ~hese radiation fields depend upon the occupation occupation numbers equal to n110m, the total number density of the species.
numbers m the appropnate levels, hence
the n1's, (i ,:/: 1, i ,:/: u).
s,.
will still depend implicitly upon For a given abundance of the element, relative to hydrogen, n110m is a pre-
specified function of depth. In overall form these equations are similar to
If we write l., for the continuum opacity and the first MH• lines of the matrix displayed, in §5-4, between equations (5-91)
and (5-92).
l.r. = (B,.h11/41t)[n; - (g 1/g.)_n.] The iteration procedure required to obtain the solution (which consists
of the run of the occupation numbers of all the levels with depth) is fairly
for the lin_e opacity, then the transfer equation at each frequency in the line straightforward in principle, though often complicated in practice. If one
can be written as
starts with an initial estimate of n1(z) for all atomic levels i, say from LTE
relations, then provisional optical depth scales can be constructed. By re-
µ(dlJdz) = -(x, + x,</>.)1, + x.s. + x,c/>,S, (12-11) garding x. and 17, as known [ via equations (7-1) and (7-2)], an estimate of the co
where S, den~tes the c??tinuum source function (which is not necessarily the
radiation field in each transition may be obtained by a formal solution of en
I
Planck fun~t1on). Wntmg 1t,
= -(1., + x,</>,) dz, and substituting for s, the transfer equation (i.e., S1 is regarded as given). One may use this estimate
to compute continuum photoionization and recombination rates. A transfer
from equation (12-4), equation (12-11) can be rewritten as
equation of the form (12-12) for each line (and opaque continuum) is then
µ(dl,/dt,) = I, - a, f cf>.J. dv - b,. (12-12)
solved in sequence, supposing, initially, that all the net brackets Z 11 and
Yj 1 appearing in equations (12-7) through (12-10), and their continuum
analogues, may be set identically to zero. This sequence of calculations pro-
which ~an be rec~st. into a second-order form, and solved by the standard duces radiative rates R 11 in all lines and opaque continua. The rate equations
Feautner or Ryb1ck1 methods described in Chapter 6. A calculation of this dn = f!lJ are then re-solved, at each depth, to obtain an improved estimate
type is carried out for each line of the entire transition array considered. Note of all the occupation numbers n. All Z 11's and Yj/s, and hence If and 0 for each
that to compute l.i, and hence t,, a., and b, in equation (12-12), it is necessary line can now be evaluated via equations (12-5) through (12-10) because all
to hav~ actual values for n1 and 1111 ; as these quantities are known only of the required rates and level-populations are known. Equations (12-12)
approximately at any stage of the calculation, it is clear that an iteration are then re-solved, using the new values for t., '1, and 0 and one obtains,
procedure is required, in which successively more accurate values of the
occupation numbers are employed. This iteration may be carried out simul-
thereby, new estimates of J,. and s,. in every line (and continuum). The rate
equations are again re-solved, and the process is iterated to convergence.
taneously with the more basic iteration required to obtain estimates of the: Once a converged solution is obtained, all the line source functions are
radiative rates in all lines. known, and line profiles may be calculated for each line.
381
380 Non-LTE Line Transfer: The Multilevel Atom

As mentioned above, the practical implementation of a successful iter-


ation procedure using the equivalent-two-level-atom formalism is often quite
complicated, for the rate of convergence (or lack thereof!) may be strongly
affected by technical details, such as the way the net rates are computed,
source functions are evaluated, and many others. Careful discussions of these
points can be found in (18, Chap. 4; 23, 27-63; 23, 113-132; 187), as well as
in the references cited therein. A more fundamental difficulty with this overall
approach is that the iteration procedure may fail to converge but, instead,
stabilizes on a solution that is inconsistent; see the discussion in (18, §4.2)
and (23, 27-63). This failure, which is not too common, and which can some-
times be overcome with special procedures, is, nevertheless, not surprising;
7.22 eV
as we shall see in the next section, there are many physical situations in which
the radiation fields in different lines are very strongly interdependent, in
contradiction to the basic assumption of the equivalent-two-level-atom
1420
approach. The inconsistency problem is overcome wholly by the complete-
linearization method presented in §12.3.
1219
APPLICATION

The equivalent-two-level-atom approach has been extensively applied 3.14 eV


to calculations of a wide variety of spectra, particularly for lhe solar atmo-
sphere and for solar-type stars. For example, for the sun, analyses have been
made of the Ca II H- and K-lines (400; 401; 569), 0 I and C II lines (168), __..
the Mg I b-lines and Na I D-lines (21), and the spectrum of Fe I (22), using' 1.70 eV
(0
sophisticated multilevel atoms and elaborate atmospheric models. Very
(0
complete spectrum syntheses, using this general approach, of both continuum
and line intensities have led to refined photospheric-chromospheric models I
(645; 646). Further, similar analyses have been made of the Ca II H- and
OcV
K-lines in solar-type stars (56; 58).
Space does not permit a discussion of all of these results, and the reader
should examine the references cited. It is, nevertheless, of interest to quote a 12·1 I 4 25 4 2 p. and
few results (401) for the solar Ca II lines, which play a central role in studies
FIGURE +
Energy-level diagram for lower states or c_a . The fivetvehs ~
3 12 D suffice to describe the line-formation process or t e a
1r k'
Hpand" lines at
h I I
of the chromosphere. The H and K lines arise from the transitions 4s 2 St -+ . '. 39Ji! land 3933 and the infrared-triplet i.i.8498, 8542. 8662._ Note t~at t e_ nexth eves
1 1
n be i nored though interactions with t c
4p 2 Pt. i (see Figure 12-1); the two upper states are coupled by collisions, • · • '
tie fairly high •~ energyd, anfd hcn(~Ot) by pe~missi~n. (Courtesy of the Publications of
and may also decay to the metastable 3d 2 Di-t states in the infrared triplet continuum are include • rom •
i.i.8498, 8542, 8662. As the next levels lie well above the 4p level (recall the Astronomical Society of the Pacific.)
kT ~ 0.5 eV in the solar photosphere), the five levels mentioned above, plus
the continuum, suffice to provide an accurate description of the physics of
line-formation. A solution of the transfer and statistical equilibrium equa-
tions for this 5-level ca+ atom yields source functions in all 5 lines; the
frequency-independent K-line source function (obtained assuming complete
redistribution) is shown in Figure 12-2, along with the Planck function
0.12 -------......---..---,----,,---,,-,,,
382

10-,,........,......,.....,.......,....,.....,....._,....,.........,--,....,.........,,....,.-,-.,.....,,.....,.-,-.,.....,
0.10

0.08

1(6).)//,

0.06

0.04

0.02 L - - l . . - - L - - . J . . - - ' - - - - ' - - - ' - - - '


10- 1 10 102 103 10' 105 106 0 0.2 0.4 0.6
t
10-7
2500 2000 1500 1000 500 0 FIGURE 12-3
h(km) Computed ff. and K-line profiles atµ • 1 (dashed lines)
compared with observed profiles (solid lines). From (401 ),
FIGURE 12-2 by permission. (Courtesy of the Publlcarlons of the
Ca II K-line source function S, and corresponding Planck-function Astronomical Society of the Pacific.) I
8, in a model solar atmosphere. Source function is in absolute I\)
units: ergs cm- 2 sec-• hZ- 1 sr- 1• Optical depth scale is at line
center and dashed lines indicate computed intensities at K., K 2 , 0
and K 3 (see Fig. 12-3). From (401), by permission. (Courtesy of ~~
0
the Publications of the Astronomical Society of the Pacific.) 0.6 ..&~ I
Calculated -.___ .~
8498 Observed ~~
corresponding to the adopted run of temperature Te(h). The emergent inten-
sity profiles, at µ = 1, for the H- and K-lines and the infrared triplet, are
compared with observation in Figures 12-3 and 12-4, respectively. The agree- 0.4
-
----.:tl!!=~--
8542
~

Calcula
Observ
ment between theory and observation is quite satisfying. However, when the
1(6).)//,
computed variation of the H- and K-line profiles from center to limb is
compared with observation, significant disagreements are found; as we shall
see in §13-4, these are removed when the partial coherency of the scattering
0.2 Observed
process, and the resulting frequency-dependence of the source function, are 8662 Calculated
taken into account. Further, the calculations shown in Figures 12-3 and 12-4
represent only the average quiet chromosphere. To match plage profiles (569),
a different atmospheric model is required; and to match the detailed variation
of the K-line profile, as observed from point to point on the disk, will require 0.2 0.4 0.6 0.8
0
a full treatment of the three-dimensional fine-structure and the velocity fields
6).
in the chromosphere. Such calculations have not yet._been undertaken in a
completely satisfactory fashion, but are needed, both to diagnose details FIGURE 12-4
Computed Cu II infrared-triplet profiles at 11 = I compared with observe~ profil~s.
of chromospheric structure, and to answer the question "To what extent F'rom (401 ), by permission. (Courtesy or the Pub/ic111io11s or the Astronomical Society
r ,1.. n ... :1~, \
384 Non-lTE line Transfer: The Multilevel Atom /2-2 Effects of level Coupling: Source Function Equality in Multiplets 385

(if any) is it possible to replace the complicated multidimensional struc- not be discussed here, and the reader is encouraged to refer to the complete
ture of the chromosphere with a meaningful horizontally-averaged 'mean discussion by Jefferies (334, Chap. 9).
chromosphere'?"
PHOTON DEGRADATION AND CONVERSION

12-2 Effects of Level Coupling: Spurce Function Consider, first, a three-level atom with two resonance lines 1 +-+ 2 and
Equality in Multiplets 1 +-+ 3, and a subordinate line 2 +-+ 3. As an example, in hydrogen we would
have the Lrx, L{J, and Hrx lines; in fact, for expository convenience, the three
As was emphasized in the preceding section, the equivalent-two-level-atom lines will be referred to with these names, even though we may not be dealing
approach works best when the coupling terms from two particular levels with hydrogen. For the present purposes, interactions· with the continuum
to all other levels are small, compared to those between the levels themselves will be ignored, and attention focussed entirely upon the lines. Generally,
and from these levels to the continuum. There are, however, many important we can expect the ground-state population 11 1 to be vastly larger than the
physical situations for which this will not be the case, and it is important to upper-state populations 11 2 and 11 3 ; hence the resonance lines will be much
understand the effects of strong coupling among several levels. These effects more opaque, and will have much larger optical depths, at a given physical
have important implications that affect both our perception ofline-formation, depth, than the subordinate lines. Suppose now that all three lines were
and our choice of numerical methods used to obtain solutions of multilevel somehow uncoupled, and were formed in distinct two-level atoms; each
transfer problems. Most of the fundamental physical concepts were outlined would then thermalize at a characteristic depth AIJ ~ 1/e,i, where elJ is the
by Jefferies (333; 334, Chap. 8), and much of the subsequent theoretical collisional destruction probability in transition i +-+ j. Because hv 23 « hv 12
work [e.g., (567)] has been based on his ideas. Numerous instructive cal- (or /iv 13 ) we will generally find e23 » e12 (or e 13 ); hence the subordinate
culations, using idealized multilevel atomic models and parameterized model line will thermalize in a smaller number of optical depths, measured on its
atmospheres, have been carried out; see, e.g., (54; 18, Chap. 4; 187; 213; own scale, than the resonance lines.
217; 218. Rather than attempt to summarize these diverse investigations But this will be overshadowed by the much greater opacity in the resonance
here, we shall, instead, confine attention to two concrete examples which lines, and the subordinate-line thermalization will occur at much greater I
illustrate the nature of interlocking effects in a particularly illuminati~g way. geometric depth in the atmosphere. Put another way, the subordinate line I\)
First we shall consider, briefly, the case of three levels coupled together is so much more transparent than the resonance lines, that photon-escapes 0
by two lines of a resonance series, and one subordinate line. The primary (and hence departures of S from B) first occur in this transition, as one
result of interest here is photon degradation from the higher resonance line proceeds outward from great depth toward the surface of the atmosphere.
into the lower resonance and subordinate lines; this phenomenon is related There will thus be a certain range of depths over which the resonance lines,
to fluorescence in nebulae and to the Rosseland cycles discussed in §5-5. were they not coupled to the subordinate line, would by themselves be
Next, we shall devote the remainder of the section to the interlocking of thermalized, but within which the subordinate line has become transparent
lines within a multiplet; here photon conversion may occur as collisions enough to permit escapes. In this critical range, some of the electrons
shuffle electrons among the fine-structure levels of the spectroscopic terms photoexcited by L/J from level I to level 3 will decay into Hrx, and the Hrx
involved. As a result, the source functions in the different lines become photons will escape; there will thus be a systematic c/egrac/c,tio11 of L/J photons
dependent upon one another, and in a certain limit become equal at each into Hrx and Lrx photons, at a rate determined by the branching ratio
depth in the atmosphere. Source-function equality in multiplets is of great A 32 /(A 31 + A 32 ). Note that the inverse process of conversion of Hrx photons
theoretical importance, for it implies that we may replace several lines within into L{J is essentially ineffectual. If an H'.1. photon is absorbed, leading to
a multiplet by a single representative line, and thereby reduce greatly the an excitation 2 -+ 3, most of the subsequent emissions from state 3 will be
amount of computation required. Further, source function equality, when it in the 3 -+ I transition; because, however, r 13 is so large and e 13 is so small,
occurs, allows application of a non-L TE method of analysis of line-profile these photons are trapped, and merely scatter until a 3 -+ 2 transition is
data, by means of which the physical properties of an atmosphere may be finally made, and the photon escapes. In this atmospheric region, Lrx will
inferred directly with a minimum of theoretical interposition. In keeping adjust in such a way as to remain almost in detailed balance (with its source
with the remainder of the book, in which theoretical prediction rather than function near to the Planck function), while L/J excitation is drained, and
analysis of observations has been emphasized, this analytical method will its source function is depressed.
386 Non-LTE Line Transfer: The Multilevel Atom 12-2 Effects of Level Coupling: Source Function Equality in Multiplets 387

Ignoring stimulated emissions, we can express "reduced" source functions approach equality, at a value well below that which the L{J source function
as (S 2i/B 21 ) = (b2/bi), {S3i/B31) = {b3/bi), and (S 32 /B 32 ) = (b 3/b 2), where would have attained if degradation did not occur. Thermalization in Ha.
BIJ denotes the Planck function B(vlJ, T), and b, = n,/nr. When La. adjusts (and, along with it, L{J) occurs at about t 13 ~ 10 5 , which implies t 23 ~ 10,
to detailed balance, b2 = b1 (recall the discussion pertaining to the ground- as would be expected from 8 23 . Further discussion and interpretation of the
state population of Hin B-stars given in§7-5), and thus(S 32 /B 32 ) = (S 3i/B 31 ), results may be found in the paper cited, but it is clear from what has been
so that the "reduced" L{J and Hr,. source functions become equal below the said here that there are major effects of interlocking on the source function
point of La. thermalization. As we proceed to the outermost layers of the variations in this problem. See also (18, §4.8).
atmosphere, the L{J line becomes more and more transparent and, ultimately, Let us now turn to the case of a three-level atom consisting of a ground
the probability of direct escape exceeds the branching ratio into Hr,., with state, and two close-lying upper states. We assume that both radiative and
its consequent escape; at this point the L{J and Ha. source functions become collisional transitions can take place between states 1 and 2, and between
uncoupled. I and 3, but that only collisional transitions occur between states 2 and 3.
Results from a calculation (218) for a three-level atom with the parameters This model simulates the actual physical situation for a resonance doublet,
A 31 /A 32 = 10, 8 12 = 8 13 = 10- 3, 8 23 = 9 x 10- 2, (r 12/r 13 ) = 6.2, and where the upper states are separated by fine-structure splitting. For example,
(r 23 /r 13 ) = 1.6 x 10- 4 [{S 3i/B 31 )/(S32 /B 32 )] are shown in Figure 12-5. the sodium D-lines arise from transitions from the 3s 2S½ ground state
There we see that Lr,. thermalizes at r 13 ~ 2 x 10 2, which impliesr 12 ~ 103, (state 1) to the upper states 3p 2P½and 3p 2P1 (states 2 and 3).
as expected from 8 12 . At t 13 ~ 10, the probability of L{J escape is comparable Consider first the limit in which the 2 +-+ 3 collision rate is zero (i.e., the
to the probability of branching into Hr,.; therefore for t 13 < 10 the "reduced" uncoupled case). The transitions 1 +-+ 2 and 1 +-+ 3 can then take place without
L{J and Hr,. source functions diverge. But below this depth, they rapidly reference to one another, and the two lines are formed independently. Each
line will have a source function that falls below the local Planck function
near the surface, and equilibrates to the Planck function at thermalization
depths given by
(12-13a)
I
and (12-13b) I\)
0
10· 1 Here we have made the simplifying assumptions of Doppler profiles and an I\)
absence of strong gradients. In the limit considered here, the run through I
the atmosphere of the source functions for the two lines will, generally, be
different, and at any given depth they will not usually have the same value.
On the other hand, suppose that collisions occur very rapidly between
levels 2 and 3. In this case, excited electrons are shuffled from one level to
the other; we then say that photons have been converted back and forth
between the two lines. In the transfer problem the two upper states become
effectively a single state, and in the extreme limit, the occupation numbers
of the two upper states will be proportional to their statistical weights. The
10-3 ..__ __,__ __.__ _.....__ __._ __.__ _......__ ___..___ __,
source functions for the two lines (which depend on the ratio n,g./n.g1) then
become equal at each point in the atmosphere. (Here we ignore the inconse-
10" 2 10" I 10 102 10 3 JO' 10 5 106
quential differences that may arise because v12 is not exactly equal to v13 ).
In general, the actual situation will be intermediate between the two extremes
FIGURE 12-5 described above, and we expect that the source functions will be equal only
Source runctions ror 3-level atom in which both resonance lines, (I ..... 3) and (I ..... 2). and the
subordinate line (2 ..... 3) are radiatively permitted. Abscissa: optical depth in the I -, 3 over a limited range in the atmosphere.
line; ordinate: "reduced" source runctions [S1,(r)/S1,(oo)]; S1,(oo) is equal to the Planck Proceeding outward from the deepest layers of the atmosphere, we will
runction at the appropriate frequency. From (218), by permission. find that the source functions in the lines begin to drop below the Planck
388 Non-LT£ Line Transfer: The Multilevel Atom 12-2 Effects of Level Coupling: Source Function Eq11ality in Multiplets 389

function at depths smaller than their thermalization depths; this occurs disk-position is given by
(cf. §11-2) because the escape probability begins to exceed the destruction
probability. But if the collision rates C 23 and C 32 are nonzero, there is now 1,(0, 11) = µ- • So"' [S,(r,) + /3,S,(r,)] exp [- µ- 1 s:· (1 + /3,.) cir,] dr,
a finite chance that the photon will be switched from one line to another,
rather than be destroyed or escape, and there is now a conversion length (12-16)
(333; 284, 177) over which the source functions in the two lines are still where r, is the continuum optical depth scale, and {3. = x,<P,IX,· In the core
coupled together. The lines become uncoupled only when the escape prob- of the line /3. » 1, so that, with good accuracy, we may write
ability exceeds both the destruction and the conversion probabilities-Le.,
when /,.(0,11) = µ-• S S (r)exp(-r<f,Jµ)<f,,.dr
0
"' 1 (12-17)
P,(t12);,;:; (C21 + C23)/(A21 + C21 + C23) (12-14a)
where r is now the line optical depth scale. For the D-lines one has
and Ph13};::; (C31 + Cn)/(A31 + C31 + C32} (12-14b) (/13 //12 ) = 2; thus we may write
From equations (12-14), we see that the two lines could thermalize indepen- ld0,µ,v) = Sow Su(t 12 )exp(-</>.rufµ)(</>./µ)dr 12 (12-18a)
dently over the entire range predicted by equations (12-13) only if, everywhere
on that range, C 31 > C 32 and C 21 > C 23 (by a fair margin); if either of
these conditions is violated, photon conversion plays an essential role. Again
and / 13 (0,µ,v) = J S 13(t
0
"' 12 )exp(-2<f,.rufµ)(2<J,,/µ)dr 12 (12-18b)
assuming Doppler profiles, we can estimate the optical depths from which Thus clearly / 13(0, 11, v) = / 12(0, ½µ, v) if S 12 and S 13 have a common depth
photons in each line can retain their strict identity, and can emerge, without clepenclence. In essence, one may compensate for the higher opacity in one
being either collisionally destroyed or converted, as line by increasing the path length in the other. When this comparison is
carried out, very good agreement between the cores of the two D-lines is
AT2 :::::; (A21 + C21 + C23)/(C21 + C23) (12-15a) found, as shown in Figures 12-6a through t 2-6c. By way of contrast,
and + + C32)/(C3 1 + C32) (12-15b) Figure t 2-6d shows the inverse comparison of/ 13(0, ½µ, v) with / 12(0, µ, v),
Af3 :::::; (A31 C31
and demonstrates the genuine significance of the agreement shown in the
I
F\.)
in the 1 +-+ 2 and 1 +-+ 3 transitions respectively. Let zf 2 and zf 3 be the other figures. The disagreement in the wings arises from the increasingly
0
geometric depths (from the surface) corresponding to Af 2 and Afl, Then large contributions from the continuum, which invalidate the assumptions
the two groups of photons may propagate independently only over a depth required to write equation (12-17). w
z* = min(z! 2, z! 3). These .striking observational results provide an impetus to examine in I
From a mathematical point of view, whenever conversion is competetive detail the conditions under which source function equality can occur (aside
with thermalization, the usefulness of the iteration procedure employed by from the trivial case of LTE), and to develop methods well suited to handle the
the equivalent-two-level-atom approach becomes less clear. It then appears transfer problem for multiplets. The strong agreement shown in Figure 12-6
more attractive to consider the two (or more) lines, and their effects upon is so impressive that, in some of the early analyses (6S5; 656), it was concluded
one another, simultaneously, and to develop a different computational that there must be strict source function equality in the two lines all the
technique. way to the surface. To obtain such equality, one would have to impose the
exacting requirement that C 23 » A 21 and C 32 » A31 ; these conditions
cannot actually be met by the collisional rates for the D-lines in the solar
OBSERVATIONAL INDICATIONS
atmosphere, and an apparent contradiction arises. Subsequent work (16; SI)
OF SOURCE FUNCTION EQUALITY
has shown, however, that the requirements stated above are far too stringent,
To motivat~ further the theoretical development, let us consider some and that one needs to have only (C 32 /A 31 ) > (C 3i/A 31 ) and (C 23/A21) >
of the observational evidence that source function equality in multiplets (C 21 /A 2i) (conditions that are met for the solar D-lines) to obtain a very
actually occurs. An excellent example is provided by an extensive set of near (though not exact) equality of the source functions. These conditions
precise observations of the sodium D-lines at various positions on the disk are sufficient to produce profiles that look identical to within the accuracy
of the Sun (655). The emergent intensity at a specified frequency and of the observations.
390
12-2 Effects of Leue/ Coup/in(/: Source Fllnction Equality in Multiplets 391

.-.--.--r-,-....,...-.--.--,----,, a ,......,,--,__,.""'T"....,...--.--.---..-,--, a SOLUTION OF THE TRANSFER EQUATION IN MULTIPLETS

Let us now consider how the coupled transfer problems for lines in a
multiplet, arising from a single lower level to two (or more) upper levels
00 U that interact via collisions, are formulated and solved. For simplicity,
ou stimulated emissions and coupling to the continuum will be neglected, and
II c.
:~
.g u
the differences in line frequencies ignored; these may, of course, be taken
into account, but the algebra becomes much more complicated without a
-o
ec compensating enhancement of the physical content. The source function
in the (1 -+ j) transition is
511
(12-19)
The required population ratios are obtained from rate equations; fer the
case of only two upper levels these are:
0...............00~-'°~~..,~-,..-~o~
d o o o ol n2(A 21 + C 21 + C 23 ) = n1(B 12 112 + C 12 ) + n3C 32 (12-20a)
and (12-20b)
where

,......,.-T--.,,C""T'"-,--,-,-,----r-, d00 .--.~--.---.--.-.. . . -.-..---,--, a J 11(t) = _ooao J u(t, v)cpi(t, v) dv = 21 J""_ao dv </>1(t, v) f_I 1 dµ I 11(t, µ, v)
J
..
-·.... I
..
··•.:•, \}\ ~ Making use of the Einstein relations, and detailed-balancing relations of
the form CI1 = (nifn I )*C1,, and defining &12 = C21 /A 2 i, 'ltl = C 23 /A 21 ,
(12-21) I\)
0
~

• ··----) 0 1 '.§: ·--- ••) 0 1 &13 =C3ifA3i, and '113 =


C32/A31, we may rewrite equations (12-20), in
terms of source functions of the form of (12-19), as
I

.._. ... ,,,

......•'{
,
....
/

~
../•''
•'•
/=. ~
and
S12
S13
= <112 + &12B, + '112S13)/(l + &12 + '112)
= (113 + &13B, + '113S12)/(l + &13 + '113)
(12-22a)
(12-22b)
. . I
.f ..I, I These expressions display transparently the linear dependence between
,•' S 12 and S 13 • We see that if17 12 = 17 13 = 0, the source functions are uncoupled.
But if either 17 12 » 1 or 17 13 » 1, then S 12 -+ S 13 . We see further, that even
o'---'~oo~-'°~~..,.--0"!--::0a if the strong inequalities just mentioned are not satisfied, merely having
..: d d d d I 17 11 » elJ and 11uS,h » euB,. will imply a near-equality of S 12 and S 13 , for
then the physical source-sink terms in each source function will be dominated
by the other line. The 'lu terms are thus expected to influence the solution
in a way reminiscent of the collisional versus photoionization domination
found in the two-level case (i.e., the lines in a multiplet may be "conversion-
dominated").
In view of the linear dependence discussed above, it is evident that a
simultaneous solution for both source functions is mandatory. Let us
392 Non-LTE Line Transfer: The Multilevel Atom 12-2 Effects of Level Coupling: Source Function Equality in Multiplets 393

consider a single ground state and an arbitrary number of upper levels. which describe the depth-variations of these quantities on the range r 1 ~
Equations (12-22) can be replaced by the general form r ~ rv. Then the system (12-28) can be written in the general form
L+I
(12-23)
TJ.mnUj.mn + L Ul,mnJl = KJ,mn (12-31)
l=l

where the sum extends over all upper levels. This system can be solved (at for each (j = 2, ... , L + 1), (m = 1, ... , M), (n = 1, ... , N). Here T
each depth-point in the medium) to obtain an expression of the form is a (D x D) tridiagonal matrix, the U's are (D x D) diagonal matrices, and
K is a vector of dimension D. These systems are solved in succession to
(12-24) yield, in effect,
L+I
Exercise 12-2: Write explicit analytical expressions, of the form of equation UJ,m• = ksl
L cjk,mnJk + DJ.mn (12-32)
( 12-24), for the case of only two upper levels, starting from equations ( 12-22).
where C is a full (D x D) matrix. Equations (12-32) are substituted into the
To write the transfer equation we now adopt the optical depth r 11 in some matrix representations of equation ( 12-27). namely JJ = Lmn w), mnUJ, mn, to
particular line (1 ➔ I) as standard, and write y1 = (dr1J/dr 11 ) = (JIJ/f11 ). generate a final system of the form
Then for each line (1 ➔ j) we have
L+I

(12-25) L p}kJJ = QJ, (j = 2, ... , L + I) (12-33)


k=l

We introduce depth, angle, and frequency discretizations {r 4 }, {µm }, and Here each PJk is a (D x D) matrix, so the whole system is of order LD x LD.
{v.}, and appropriate quadrature sums required to perform the integrations The system is then solved by standard numerical methods, and yields the
indicated in equation (12-21). Then defining full depth-variation in all lines simultaneously. When stimulated emission
terms are included, the system becomes nonlinear and requires an iterative
solution. A linearization method for the problem is described in (23, l); we I
shall not discuss this particular procedure, but instead will describe, in I\)

and rewriting equation (12-21) as §12-3, a more general method that handles this problem and many others 0
from the outset. tn
JJ.d = JIJ(r,) =LL Wmn<Pi(t,, v.)uJ.dmn (12-27) The sodium D-Jines have been studied (51), using schematic atomic and I
n m atmospheric models, with an integral-equation method essentially equivalent
to the system described above. For these particular lines the following
the transfer equation can be written in the discrete form
relations among atomic parameters exist: g1 = 2, g2 = 2, g 3 = 4,
A31 = A 21 , B 13 = 2B 12 , C 21 = C 31 ; hence e12 = e13 and 11 12 = 211 13 •
(12-28) Adopt the 1 .... 3 transition as standard so that y2 = ½and y3 = I. Solutions
were obtained for a range of typical values of the parameters e13 and 11 13 ,
Here cf,1, 4• =<Pit 4, v.), a1k.d =a k(t
1 4 ), b1• 4 =bi(r 4 ),
and deoth- variations of the Planck function. Results for cases with B. = 1,
e = 10- 4 , and '1 = 11 13 = 0, 10- 4 , 10- 3 , 10- 2 , 10- 1, and 1, are shown in
and Y1., = yj(r 4 ) Figure 12-7. We see that, while 11 13 would have to be larger than 1 to guarantee
strict source function equality to the very surface, much smaller values of
the sum extends over all L lines (or upper levels) considered. An equation '1 13 yield equality for r <! l, which is the significant range for determining
of the form of (12-28) can be written for each line. the emergent intensity. The emergent specific intensity forµ = 1.0, 0.8, and
To solve the system we may use the Rybicki method. Define 0.6 is shown in Figure 12-8. Clearly the profiles agree very closely when
'1 .r= 10- 3 , and are indistinguishable when '1 = 10- 2 • These calculations
(12-29)
show that, for all practical purposes, source function equality can occur for
and (12-30) '1 « 1; indeed this result is not surprising, for we would, in fact, expect
394 395

1.0 1.0

0.8 0.8

0.6

0.4

0.2
µ C 1.0
o"="""'=~-~
0.5 1.0
0.4

0.2 0.2
µ • 0.8
o"====~--
o.5 1.0
o'--.....
0.5
- •1.0- - ~
0.4 0.4

0.2 0.2
µ • 0.6 µ • 0.6

0 0.5 1.0 t.5 2.0 2.5 3.0 3.5 4.0 0 0.5 1.0 1.5 2.0 25 3.0 3.5 4.0
X

1.0

0.8

0.6
I
0.4 I\)
0.2 0
Jl • 1.0 CJ)
o'====~---
o.5 1.0 I
0.4

0.2

0.2
Jl • 0.6

0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
X X

flGURE 12-8
Line profiles comparing/ 1 2(0, ½µ, v) with / 1 3(0, µ, v) for various values of II, atµ = I, 0.8, and
0.6. Solid curves give/ 13 , dashed curves give/ 12 • Abscissa: displacement from line center in
units of Doppler widths. From (SI), by permission.
396 Non-LTE Line Transfer: The M11/rilevel Atom /2-3 The Complete Linearization Method 397

conversion to dominate individual line-thermalization as soon as 11 » e. ionization stages. The rate equations are formulated to describe all i'nter-
The effects of a schematic temperature gradient were also studied, and it actions among all levels, and to fix the total number density at the specified
was found that, when B,(t) increases inward, the value of 11 required for value na,om• These equations will be of the general form of the_ first MH•
source function equality decreases (compared to the B. = 1 case), and lines of the matrix displayed in §5-4, supplemented by an equation of the
increases when B,(t) decreases inward. These results, however, may depend form ~ 11 = n
/..,II, J) I} atom
. For ease of exposition we make
,
the simplifying
, ,
upon the specific forms chosen for B.{t), and may not be general. assumption that, at any chosen frequency v., there ts only one trans1t1?n
(say / +-> u) of the atom that can interact (bound-bound or ~ound-free~ "'.1th
the radiation field, and that all overlapping sources of opacity and em1ss1on
12-3 The Complete Linearization Method are fixed.
The transfer equation to be solved at each frequency vis
In the light of the discussion presented thus far in this chapter, it can be
appreciated that the effects of interlocking in the multilevel line-formation (12-34)
problem are important, complex, and lead to systems of equations that may
be ill-conditioned. Neither of the methods discussed thus far is entirely From equations (7-1) and (7-2) we can write the opacity and emissivity in a
satisfactory for multilevel problems. The equivalent-two-level-atom formu- bound-bound transition as
lation is ill-posed when there is strong coupling among lines. The method (12-35a)
developed to treat multiplets is too specialized. What is needed is a method
x. = ix,.(v)[ n, - (g 1/g.)n.] + X.
that is general, flexible, and computationally robust, and that handles the 17,. = (2hv 3/c 2 )ix1.(v)(g1/g.)n. + E, (12-35b)
physical complexities of the problem. These requirements are met by the
complete linearization method (42). In this method, the full set of rate equations and that in a bound-free transition as
is incorporated into the transfer equation, from the outset, by linearizing in x,. = ix 1.,(v)[ n1 - n.(n1/n.)•e-h•tkT] + X. (l 2-36a)
terms of all occupation numbers and their dependences upon the radiation
field. The rate equations themselves constitute the physical prescription of 11. = (2'1v 3 2
/c )ix,.(v)n.(n,/n.)•e-h•/kT + E, (12-36b)
how the photons interact with the material, are created, destroyed, converted, I
where X,. and E,. are fixed. To unify the notation, write I\)
degraded, and interlock within the collective photon pool envisioned by
Jefferies. The transfer equations determine how this information is propa, Xv= ix1.(v)[n 1 - G1.(v)n.] + X. (12-37a) 0
gated from one depth to another. By means of the linearization procedure, -...J
a set of equations is developed that describes fully consistently (to first 11 • = (2hv 3/c 2
)ix,.(v)G1.(v)n. + E, (12-37b) I
order) the response of the material, at every point in the atmosphere, to the
radiation field at any frequency and any depth-point and, reciprocally, the
where G,.(v) = (g 1/g.) for bound-bound transitions, and G,.(v) n,<l>,(T) =
exp(-hv/kT) for bound-free transitions; here <l>,(T) i_s th~ Sah_a-B~ltzmann
response of the radiation field at all frequencies and all points in the medium factor for level / [cf. equation (5-14)]. The system 1s d1scret1zed m de~th
to a change in material properties at any point. When this system is solved {m4 }, (d = 1, ... , D), and frequency {vk}, (k = 1, ... , K). We then can wnte
(iteratively), one has a simultaneous solution of the full set of rate equations a transfer equation of the form of equation (7-37) at each frequency vk; these
and transfer equations that reflects both the global nature of the transfer equations are linearized to obtain an equation of the form of (7-39), where
problem (inherently large destruction/conversion lengths interplaying with now c5xdk and c517 4k can be expressed simply as
an open boundary surface) and the intricate infrastructure of the statistical
equilibrium equations. In practice, this method has provided a means of DXdk = ix1.(vk)[c5n1, 4 - G,.(vk) c5n•• 4] (12-38a)
treating extremely complex model atoms with a very high degree of realism. c517dk = (2'1v//c 2 )ix,.(vk)G 1.(vk) c5n.,d (12-38b)
We shall examine here the statistical equilibrium problem for a multi-
level "impurity" atom that has no effect upon the structure of the atmosphere. The c5n's can be expressed in terms of changes in the radiation field only, as
We regard the model atmosphere as given-i.e., the temperature (T), electron T and n, are assumed fixed, and from equation (5-102) or (7-157) we can write
density (n,), mass density (p), and the total number density of the species
under consideration (n 010 m) are all specified, fixed, functions of depth. The (12-39)
atom is assumed to have L discrete levels, distributed, perhaps, over several
/2-J The Complete linearization Method 399
398 Non-LTE line Transfer: The Multilevel Atom
To initialize the problem, the model atmosphe~e _is used ~o determine
where analytical expressions for (on/oJ.) are given by equation (5-108). By background opacities and emissivities, and the rad1at1on fi~ld 1s _co!11puted,
means of equations (12-38) and (12-39), it is now easy to eliminate all the assuming LTE populations for the atom under study. T_h1~ rad1at1~_n ~eld
on's in the linearized transfer equation analytically, in terms of the M's only is used to calculate photoionization rates, and the stat1st1cal _eqmhbr1um
(33), and to obtain equations in the standard Feautrier form equations are solved assuming radiati~e detailed balance in the _Imes, b~t not
in the continua; in this way one obtams the correct asymptotic s_olut1on at
-Ad6Jd-l + Bd6Jd - Cd6Jd+I = Ld (12-40)
depth. It is then possible to do an equivalent-two-level-atom solution to find
where 6Jd = (oJdt, ... , oJdk; ... , oJdKf (12-41)
J,. in each line. . . . .
and A 1 = C0 = 0. The resulting J's allow the radiative rates m the Imes to be computed, a~d
the complete statistical equilibrium equations can !hen be.solved to obtam
Exercise I 2-J: Write explicit expressions ror the elements of the matrices A,, improved level-populations. This process may be iterated, and may be ~x-
B,, and c,, and the vector L, in equation (12-40). tended to include the analogue of the equivalent-two-level-c1.tom solution
for the opaque continua as described in §7-~ [cf.. eq_uations _(7-135) through
In the formulation of the linearization procedure described in §7-5, both (7-144)]. The process outlined here could, m prmc1p_le, be iterated t~ con-
the M's and 011's appeared explicitly. By eliminating the bn's the system size sistency, and one would then have the method of §12-1; however, m the
has been reduced from [(K + L) x (K + L)] to (K x K), which saves present context, only one or two iterations are perfori:ned, and the results_ are
computer storage; but now A and Care full, whereas in the earlier formulation adopted as initial estimates of the n's, J ;_s, an~ Eddington factors reqmr~d
[cf. equation (7-159)] both A and C were diagonal (omitting the constraint to start the linearization procedure. At this pomt, and henceforth, the radia-
of hydrostatic equilibrium which is not relevant in the present context). All tive rates in the "secondary" lines are held fixed at their equivale~t~two~level-
of the operations in the elimination scheme given by equations (6-40) and atom values, and only the "primary" lines are treated exphc1tly m .the
(6-41) now involve full matrices, and the computing time therefore increases; linearization. This will provide an adequate treatment of the secondary Imes
the overall scaling is T = c DK 3 , compared to the older scaling T' = c' if they have been chosen astutely, ~nd ifth:Y ~a) are very weak, and th7refore
D(K + L) 3, but c is much larger than c'. dominated by overlapping absorpt1on-em1ss1on processes, or (b) are, m fact,
The solution of equations (12-40) is carried out iteratively, as described "isolated" and therefore accurately described by a two-level atom, or (c) are I
I\)
below. The great power of the method is that equations (12-40) account for only very 'weakly coupled to the lines of primary interest, even th?ug~ th~y
the full depth-to-depth and frequency-to-frequency coupling in the problem. are members of, say, a multiplet, and the two-level-atom approx1mat1on 1s 0
That is, the effect of Mdk at (td, v.) upon Jd'k' at all other (td,, v•• ) is given poor. . . • d ·
co
consistently, and thus photon propagation within the collective pool is taken Starting from the initial solution just described, the lmeanze equations I
fully into account. (12-40) are now solved for 6J at all dept~s. _The resulting 6~'s _are used to
In actual application of the method, we must now consider how to obtain obtain a more accurate estimate of the rad1at1on field; the rad1at1ve rates are
a starting solution, and how to treat a large number of lines, as required by a then updated with this revised field, and th~ statistical equilibri~m equations
realistic model atom. These questions are related, and the latter is of partic- are re-solved for new level populations. With the new occupation n~mbers,
ular importance. For example, suppose the model atom has, say, 20 levels. opacities and emissivities may be computed, and a formal solut1~n (/\-
On combinatorial grounds there could be of the order of 200 transitions; iteration) is then performed to obtain. the radiation. field and Eddington
if we require, say, 10 frequencies per line profile, the problem becomes factors (this step also provides a smoothing of the solution). The con_vergence
unmanageable. But of course spectroscopic selection rules severely limit the of the method is ordinarily very swift, yielding lloJ./J.11 ~ 10-s m 4 or 5
number of transitions that are actually possible in the spectrum, and among iterations, typically with order-of-magnitude improvement between succes-
20 levels one would typically have about 30 permitted lines. Of these, only a sive iterations. This scheme is the one used to carry out most o~ the work
particular subset will directly influence those levels of primary interest in the discussed in §12-4. A detailed description of a p~rticu_lar version of the
treatment of definite limited set of lines in the spectrum, and usually it is method, including a program adapted for calculat1ons m the solar atmo-
possible to divide the transition array into a set of"primary" lines for which sphere, is given in (36). . .
a strictly self-consistent treatment is necessary and a set of"secondary" lines A difficulty with the Feautrier-type approach developed above 1s tha! 1t
for which a less precise treatment is adequate. This division is exploited in becomes costly for computations involving large numbers of frequencies.
the initialization procedure.
400 Non-LTE Line Transfer: The Multilevel Atom 12-4 Light-Element Spectra in Early-Type Stars 401

The problem can be overcome, however, with a new method (37) that.uses a The overall size of the system (12-48), which contains the full transition-to-
Rybicki-type elimination. Again we assume that at frequency vk there ts only transition coupling (interlocking) over all depth-points, is (DT x DT). A
one transition (t), arising from levels I and u. At this particular frequency, the direct solution of these equations would thus require a time that scales as
linearized transfer equation (7-39) can be written, in view of equations T O = c D3 T3, which for, say, T ~ 20 and D ~ 50 becomes impractically
(12-38), as large. The system is therefore solved by iteration, using a successive over-
(12-42) relaxation (SOR) method (526, 438). In this approach there are two basic
iteration cycles: (a) the SOR iteration, to obtain a definite set of 6Z,'s
where each of the 6-vectors contains the depth variation of the quantity-e.g., • (t = I, ... , T) within a given stage of linearization, and (b) the overall
6Jk = (l,J U• •,,, l,Jdk• •,,, {,J Dkf (12-43) linearization procedure, where successive sets of 6Z,'s are used to update
rates, and the full statistical equilibrium equations are then re-solved. The
and the matrices Tk, Lk, and Uk are of dimension (D x D) and are tridiagonal.
SOR procedure is started by computing the solutions of the systems
Equation (12-42) may be solved to obtain an equation of the form E, 6Zl 0 > = C, (t = I, ... , T); this initial solution requires c D3 Toperations,
(12-44) and the resolved systems (equivalent to E, - 1) are saved. Then with any set
where !t' and ~ are now full matrices. Now the basic radiation-dependent of current estimates of the 6Z's, the righthand side of equation (12-48) can
quantities entering the rate equations are radiative rates integrated over the be evaluated for each transition in turn (note that only vector multiplications
are involved, so the procedure is very fast); this yields a single vector of
transitions in question. We therefore introduce the variations in the net rates,
known value on the righthand side, and, using the previously resolved E,,
defined as a new value of 6Z, is obtained. Each cycle in the SOR procedure requires
(6Z,), = n,., oR, •. , - "•·' l,R.r,d c D2 T 2 operations, so if J iterations are necessary, the overall computing
= L [4nwka 1.(vk)/hvk][nr,d - G,.(vk)n •. ,] M,k (12-45) time scales as TsoR = c D3 T + c'l D2 T 2 , which is clearly favorable com-
k pared to T 0 if J < DT.
where the sum extends only over those frequencies contained within transi- By actual tests it is found (37) that this method works well, even though
tion t. If we substitute equations of the form of (12-44)' into (12-45), and the SOR iterations resemble the equivalent-two-level-atom approach, in
perform the indicated summations, we obtain finally · that only one transition at a time is treated. The reason is that this part of
the calculation is required only to determine the 6Z's, which are merely
I
6Z, + A, 6n 1 + B, 6n. = C, (1216) I\)
one step of the overall linearization procedure. (in itself designed to handle
the interlocking problem self-consistently). Inasmuch as further steps in the 0
where A, and B, are full matrices. (0
linearization are presumed, the 6Z's at any given stage need not be known
Exercise I 2-4: (a) Write explicit expressions ror the elements or the matrices perfectly, but only with sufficient accuracy that the error in the current I
Tt, Lt, and Ut, and the vector Rt in equation (12-42). (b) Assuming that the estimate of 6Z, is smaller than the full size of the 6Z,'s of the next linearization
matrices !i't, d/1 t, and the vector /:Jet are known, write explicit expressions for the step. In practice, the requirement loZl 0 - oZl;- 1)1 < eloZl 0 1(where i denotes
elements or A,, B,, and C, in equation (12-46). the SOR iteration number) works well withe set at about 10- 2 • Both of the
From equations (5-108) and (12-39) it follows that one can write methods described in this section have proven to be very effective in a wide
variety of physical problems, for different atoms, in stellar atmospheres of
(12-47)
various types; we now turn to a discussion of some of the results obtained
for early-type stars.
where Dm, is a diagonal matrix with elements (Dm,)d = (8n,,,/8Z,)d =
(dd);;/ - (d,),;;/, where d, is the unperturbed rate matrix at depth-point
d, and i andj are the lower and upper states in transition t. Using equation 12-4 Light-Element Spectra in Early-Type Stars
(12-47) in equation (12-46), we obtain the system
From our study of the combined equations of transfer and statistical equili-
E, 6Z, = (I + A, D1, + B, D.,) 6Z, brium, we have gained deep insight into the physics of spectral line-formation
L Dr,· 6Z,, - B, ,.,,.,
= C, - A, ,.,,., L D.,. 6Z,, (12-48) in stellar atmospheres. But it has also emerged that, when departures from
LTE are taken into account, the equations to be solved become extremely
with one such equation for each transition t. intricate, and require special methods (which are computationally expensive).
402 Non-LTE line Transfer: The M11ltilevel Atom 12-4 Lig/,r.£/eme111 Spectra i11 Early-Type Stars 403

It is therefore of considerable interest to answer the questions: "Do depar- from stellar-structure computations. Similarly, when detailed comparisons
tures from LTE actually have a sensible effect in real stellar spectra?" "Are are made with line profiles (510), it is found that, if a fit is made to the line-
serious errors made in diagnostics of element abundances, or physical prop- win~s, the line-core predicted by LTE is much too weak. These discrepancies
erties of the atmosphere, when non-LTE effects are ignored?" "Are there ~amsh when departures from LTE are taken into account (436; 45), and it
regions of the spectrum, portions of the H-R diagram, or classes of atoms 1s found that the non-LTE profiles are invariably much stronger than the
and ions, for which we can conclude categorically that the use of an LTE LTE profiles, as shown in Figure 12-9 for a typical model. A comparison
approximation is safe, or is unsound?" From the practical point of view, it between the observed equivalent widths of H/J and both LTE and non-L TE
is important to know when it is adequate to assume LTE in the analysis calculations is shown in Figure 12-10. It can easily be seen there that the
of a stellar spectrum, for then the amount of computational work required non-LTE results are in much better agreement with observation, and that
is greatly reduced. Equally important is knowledge of when the results the LTE equivalent widths are systematically too small by factors of 3 to 5
obtained from such analyses are not to be trusted. Answers to the questions in the extreme cases [see also (174)]. Further, an excellent fit can be obtained
posed above have come forth only quite recently as it became possible to to observed profiles by using the non-LTE computations; this can be seen
obtain accurate numerical solutions of the non-LTE transfer equations, for by comparing the fits displayed in (510) and (45), both of which use the same
elaborate multilevel atomic models in realistic model atmospheres, by apply- observational data. In sum, by taking into account departures from LTE
ing on high-speed computers the techniques discussed in this chapter. It is major improvements are obtained in matching observed H-line profiles fo;
now known that non-LTE effects are indeed of great consequence in several the O-stars. F~r _the B-stars, the hydrogen lines are much less strongly
situations of basic interest and importance, and that a number of serious affec!ed by dev1at1ons from LTE, and for a line such as Hy, LTE actually
discrepancies between theory and observation are removed when departures provides a very good approximation. Some important effects are found
from LTE are taken into account. A few of the results obtained thus far will for Ho: (SI I; 430), _where de~artures from LTE produce a deeper line-core,
be described below; further details can be found in the papers cited as well and a shallower wmg than given by LTE. The latter effect arises because at
as reviews given in (433; 434; 435). It is not an exaggeration to -state that the depths where the Ho:-wing is formed, b3 > b2 (cf. §7-5) and therefore
work in this essential area has just begun, and much is yet to be learned S23 > B(v2 3 , T); the predicted changes have been confirmed by observation. I
from further efforts. After hydrogen, helium is the next most important element in stellar F\)
In many respects, the most fundamental spectrum in early-type stellar atmospheres, and is represented by the He I spectrum in B-stars, and by
atmospheres is that of hydrogen. Not only do the hydrogen-line profiles both the ·He I and He II spectra in O-stars. These lines may be used to 0
serve as effective tools to measure basic stellar parameters, such as the I
effective temperature and surface gravity, but also, when one recognizes 1.0 r~:::i:=::=;e=-=====:!l:=!:liiP'
,~,..,
-.,...,
that hydrogen dominates the transfer of radiation through the major part
of the spectrum in early-type stars, one realizes that a serious discrepancy
0.9 ,,
,,./
between theory and observation for these lines has grave implications for 0.8
the overall properties of the radiation field. Although very good agreement F(!J.i.)/F,
is usually obtained between predicted and observed H-line profiles for B- 0.7
and A-type stars (cf. §10-5), strong disagreement has been found for the
O-stars (T.rr > 30,000°K). The basic problem is that the H-lines are observed 0.6
to have nearly-constant strength in the O-stars over the spectral range 09-05
[as can be seen by inspection of the plates in (465) or (S)], while the LTE 0.5 '--'--1....L.....L__,_--L_._..J.......,__.J..._.,__.L_..____Jc_J
equivalent widths decrease markedly over the corresponding temperature 0 2 4 6 g 10 12 14
range (30,000°K ~ Terr ~ 50,000°K), owing to increased hydrogen ioniza- /J.,i
tion. This decrease can be offset somewhat by assuming a higher surface FIGURE 12-9
gravity in the models, but then inconsistencies arise. Typically it is found Hr1. profiles ror O-star models with T., 1 = 45,000°K; curves are
(420) that, if a match is made between the observed H-line equivalent widths labeled with surface gravity g. Absci.,sa gives displacement
and those calculated from LTE models, then the surface gravities required from line-center in A: ordi11crte gives residual Hux. Solid
l'l/rl'1•s: non-L TE profiles; daslied rnri·e.,: LTE profiles. Note
to obtain the fit are too large (by about a factor of 3) compared to those marked strengthening of line by non-LTE effects. From (45),
deduced from fundamental measurements of stellar masses and radii, or by permission.
405
404
He 111s s 1
BO 09.5 090807 06 05
200------------------------
4 4/
190
\

3
\
.,
\
\
\
180
\
\
W(HP) \\ 170
2
'' :(cm· 1 )/10 3
' ',,, 160
...... 2s
----- LTE
g = 10•---------

QL-L---.L....--...J...---'-----"----'
---- .... 150

30 35 40 45 50
3
T.,,/10
FIGURE 12-10 lpO
HP equivalent widths in O-stars. Soll,/ c·urves: non-LTE 'D
calculations; dasl1ed curves: LTE calculations. Dors: observed FIGUII~ 12-11
values (693). Ordinate gives equivalent width in A; absciss11e give Model He I atom used in non-L TE calculation or He I spectrum. Ordinate gives wavenumber
(T,rr/10 3 ) and spectral types. in cm· 1 times 10· 1 . All the transitions shown were included in the calculation explicitly as
..primary.. transitions in the sense defined in §12-3; all other dipole-permiued lines up through
11 = IO were treated as .. secondary" transitions. with rates fixed by an equivalent-two-level-

estimate the helium abundance N(He)/N(H), and in the 0-stars; the ratio atom computation. The singlets and triplets are treated separately. From (46), by permission.
I
of strengths of the He II lines to He I lines provides a sensitive temperature- I\)
indicator. LTE calculations of He I lines in B-stars have produced excellent .......
agreement with the observations for the blue-violet lines [see, e.g. (390; which shows that a given fractional departure /J of the level-population ratio .......
480; 481)], especially when the most accurate line-broadening theories avail- is amplified by a factor o- 1 when b < 1. The occupation numbers, and hence
able are used (438; 439). However. lines in the yellow-red regions of the the ratio (brf b.) will, in general, be determined by processes other than those
spectrum [ e.g., }.5876 (2p 3 P-3d 3 D) and i.6678 (2p 1 P-3c/ 1D)] are always in the line. In the 0- and B-stars the level populations are set by photo-
found to be stronger in B-star spectra than given by LTE computations, using ionizations and recombinations, and thus /J will generally be nonzero despite
the same model and abundance that fit the blue-violet lines. Formally these a large collisional rate in a line. Once /J #- 0, large departures of S, from
lines indicate abundances almost three times larger than that given by the B,. become possible; the same argument, by the way, explains why even tiny
blue-violet lines. When a detailed non-L TE calculation is made (46). using a departures from LTE have very large effects in the radio-frequency recom-
realistic atomic model (see Figure 12-11), it is found that the lines mentioned bination lines observed in nebulae ( 199). In the 0-stars, LTE predicts too-
above are strengthened dramatically. and that good agreement with obser- weak lines for both He I and He II. If one attempts to match the observed
vation is obtained, as shown in Figure 12-12 for }.6678. It is interesting that equivalent widths, without regard to line-profiles (420), the derived helium
the largest non-L TE effects are found for the lines with the smallest values abundances are about a factor of two higher than the value obtained from
of /11•, for, as mentioned in §5-3, these lines, with relatively small values of B-stars, or from the nebulae in the interstellar medium from which the
hv/kT, have collisional rates that are comparable to the radiative rates, and 0-stars have just formed. If a fit is made to the wings of the line profiles
the classical argument is that these lines should therefore be in LTE. The (510). it is again found that the line cores are much too weak. On the other
reason for the large effects can be seen by examining the source function in hand, a non-LTE calculation (45) yields excellent agreement with observed
the limit b = hv/kT < 1. If we write (b 1/b.) = (1 + /J), then line strengths and profiles at the "standard" abundance N(He)/N(H) = 0.1
[see also (174; 175)].
(Sr/ B,) = (ehv/kT - I )/[(b,lb.)ehv/kT - I] Departures from LTE on occasion introduce large errors into abundance
= (e 6 - l)/[e 6(1 + fJ) - l] ~ o/(/J + b) = (1 + {J/0)- 1 estimates based on LTE calculations. For example, LTE determinations of
406
407

HD41692
1.0
y Peg
0.9

0.8 W().4481)

0.7
0.1
1.0 0.6 '
·, -.; ;: o••,-:,,.
0.9 0.5
""'···················
0.8 I 5 17.5 20 22.5 25 27.5 30 32.5 35 37.5 40
3
He 1). 6678 T.rr/10
0.7
FIGURE 12-13
0.6
Strength or Mg II ).4481 in B- and O-stars. Ordinate: equivalent width
0.5 in A; abscissa: (T, 1,!10 3 ). Squares: observed values at the indicated
• spectral types. Dotted curves: LTE calculations at a solar or ten-times
0.4 solar Mg abundance, with zero microturbulence. Dashed curve: solar
L,l__._,.L_a......L_._..L.&....J...........1.--L-.L..........1.--L.....1._,J Mg abundance, and 4 km s • 1 microturbulence. Solid curves: non-LTE
-4 -3 -2 -1 0 2 3 4 calculation at solar Mg abundance, and zero (open circles) or 4 km s • •
,1). (filled circles) microturbulence. Note that ror the O-stars the choice of
microturbulent velocity docs not influence the result, and that the LTE
FIGURE 12-12
predictions yield much too small a line-strength, while the non-L TE I
Comparison or observed (390) He I i.6678 profiles (docs) I\)
results agree well with the observations.
with non-LTE (solid cim•es) and LTE (clashed curves)
.......
calculations ror three B-stars. Ordinate: residual flux;
abscissa: displacement from line-center in A. From (46), produce lines of similar strength in LTE, an abundance 5 to 7 times the solar I\)
by permission. value would be required. In short, there is ample evidence that, at least for I
early-type stars, deviations from LTE may produce large and important
changes in abundance estimates.
the abundance of Mg in O-stars are in error by a factor of 10 or more (431),
Yet another context in which non-LTE effects enter preeminently in early-
and only when a non-LTE calculation is made, is a fit obtain:d to _t~;
type stellar spectra is in the production of emission lines. For example, in
observations, with an abundance near the solar (and accepted cosmic )
value as shown in Figure 12-13 [see also (384; 586)]. Similarly, LTE analyses the Of stars, the lines of He II J4686, C III A.5696, and N III U4634-41 are
of th~ spectrum of Ne I in B-stars have routinely given a neon abundance seen in emission, while other lines of the same ionic spectra are in absorption.
of 5 x 10- 4 relative to hydrogen, in disagreement with the value 10- 4 In the case of He II A.4686, studies have been made of selective excitation
mechanisms (45) and the effects of atmospheric extension in static models
obtained from nebulae, the solar corona, the solar wind, and cosmic rays
(376); while weak emission lines have been obtained in such work, it is clear
(both galactic and solar-produced). Again the discrepancy is removed when
that a quantitative fit to the observed line-strengths will be obtained only
a non-LTE calculation is done (47) and it is found that the Ne I spectrum,
when large-scale atmospheric expansion is taken into account (see Chapters
consisting primarily of lines in the red, is affect_ed by essentially the s~me
14 and 15). The C III line has not yet been subjected to analysis. For the
mechanism as that described above for He I. Finally, the spectra of St III
N III lines there exists a subgroup of the Of stars, designated 0( (f) ), in
and Si IV in B- and O-stars show significant non-LTE effects (349). Depar-
which the N III lines are weakly in emission and the He II line is in absorption
tures from LTE increase the computed line-strengths by 50 to 70 percent,
(657); these appear to be near-main-sequence objects in which the effects
and line-core intensities decrease by a factor of 0.6 relative to LTE. To
of atmospheric extension and expansion, if present at all, may be ignored.
No11-LTE Line Transfer: The Multilevel Atom /2-4 Light-Element Spectra in Early-Type Stars 409
408

A detailed non-LTE study of the N III doublet spectrum (429; 440; 443; 41 emission lines can be produced in planar, static atmospheres, solely as a
433) gives a plausible explanation of how the J.i.4634-41 lines fro~ the result of the atomic structure of the ion responsible, they must be regarded
3p 2 p .... 3d 2 D transition can be in emission, while the next lo_we~ multiplet, as intrinsic emission lines. In expanding atmospheres, direct pumping of the
i.i.4097, 4103 (3s 2 S .... 3p 2 P) is in absorption. The two essential mgredtents 3s and 3d states will occur in the 2p --+ 3s and 2p --+ 3d transitions, because
required to produce this result are (a) a mechanism for populating the 3d the resonance lines will be Doppler shifted into the bright adjacent continuum
state with enough electrons to induce the 3d --+ 3p emission, and (b) a way and will not be in detailed balance. This greatly enhances the 3d .... 3p
of depopulating the 3p state so that emission does not appear in the 3p --+ 3s emission (thus explaining the very bright emission seen in the Of stars, which
transition. Both the 3s and 3d states can be populated directly by transitions are known to have expanding envelopes and stellar winds), while the in-
from the ground state (see Figure 12-14) and, when the ultraviolet radiat!on creased 3s population, coupled with the drain from 3p; assures that the
field at ).374 and ).452 is intense, these states will have large occupation 3s--+ 3p lines remain in absorption. The calculations show (see Figure 12-15)
numbers. In static atmospheres, however, the resonance lines come into that the ).).4634-40 lines make the transition from absorption to emission
detailed balance, and therefore do not lead to a very large overpopulation
of the 3d state. Rather, the overpopulation results from die/ectronic recom- 0.3 r--rc3...,._3---,----,----,---..,..--...,....--~--~
bination from the 2s2p( 1P)3d state-which, for the N + + ion, just happens 3.3
to lie barely above the ionization limit. Stabilization of this state occurs 3.5
when the 2p electron drops to the 2s level, and proceeds at a very rapid rate, 0.2
feeding electrons directly into the 3d state. These electrons then decay
3d .... 3p, producing the emission. Further, it turns out that, because of the
particular structure of the N+ + ion, the most probable route of exit from
2
3p is not 3p --+ 3s, but rather via "two-electron jumps" of tl)e form 2s 3p --+
2s2p 2 (see Figure 12-14); the latter process occurs so efficiently tha_t it drains
the 3p state, leaving the 3s --+ 3p line in absorption. Inasmuch as the V.4634- I
!\)
.......
w
I

-0.3
3.5
__.
-0.4 i.4640-41

3.5

-0.S
32.5 35 37.5 40 42.5 45 47.5 so
3
T, 11/10
FIGURE 12-15
N Ill line-strengths for the 3s -+ 3p (i.4097) and 3p -+ 3c/ ().).4634-40) lines in planar,
static, model atmospheres. Orclinate: equivalent width (negative values denote
FIGURE 12-14 emission); abscissa: (T, 11 /10 3 ). Note that the i.4097 line remains in absorption while
Simplified term diagram for the lower states or N • •, as well as the autoionizing state
2
2s2p( 1 PJ3d from which diclcctronic recombinations can occur directly into the 2s 3c/ state, i.i.4634-40 make a trnnsilion from absorption to emission 111 a spectral type near 06.
From (443) by permission.
leading to emission at ).).4634-40.
410 Non-LTE line Transfer: The Multilevel Atom

13
at about spectral type 06 near the main sequence, in agreement with obser-
vation (443).
Many more analyses, for a wide variety of atoms and stellar atmospheres,
need yet to be done (particularly for later spectral types) before the questions
posed at the beginning of this section can be answered fully, and efforts in
this direction will be richly rewarded with interesting results.

Line Formation with Partial


Frequency Redistribution

I
1\)

In the process of scattering in spectral line-formation, an atom is excited from


one bound level to another by the absorption of a photon, and then decays
radiatively back to the original state, with the emission of a photon. In our
work on line-transfer we have, thus far, assumed that scattering is either
strictly coherent, or that the photons are completely redistributed over the
line profile. Neither of these limits is achieved exactly in stellar atmospheres,
and it is necessary to consider the redistribution of photons, in angle and
frequency, in some detail, and to calculate redistribution functions, which
describe the scattering process precisely. This calculation proceeds in two
steps. We first consider a single atom in its own frame of reference, and com-
pute the form of any redistribution that occurs within the substructure of
the bound states. Then, recognizing that what is actually observed in a
stellar atmosphere is an entire ensemble of atoms moving with a thermal
velocity distribution, we take into account the Doppler redistribution in
frequency produced by the atoms' motion. Doppler redistribution arises
because the incident and emergent photons travel, in general, in different
directions; in this event the projection of the atom's velocity vector along the
propagation vectors will be different for the two photons, and a differential
/3-1 Redistribution in the Atom's Frame 413
412 Line Formation with Partial Frequency Redistribution

Doppler shift must occur. The final redistribution function is obtained by (a) Case I. Here we consider an idealized atom with two perfectly sharp
averaging over all possible velocities. Once the redistribution function is e
states. Then f(e') de' = «5(f - 0 ) de', and p(e', e) = «5(e' - e) where «5
known, a transfer calculation can be carried out, taking into full account denotes the Dirac function, and , 0 is the line-center frequency; clearly there
any correlations (or lack thereof) between incoming and outgoing photon is no redistribution in the atom's frame in this case. It is obvious that the
frequencies. conditions described here do not apply to any real line, for normally one
In this chapter the notation and methods of Hummer (313) will be adopted; (or both) levels will be broadened. It is useful, nevertheless, to study this
the reader is referred to this paper for further details and other results. An limiting case, for it demonstrates the effects of Doppler redistribution alone,
interesting analysis of the problem from a rather different conceptual point as seen by an observer in the laboratory frame examining the ensemble of
of view can be found in (295) and (296). moving atoms.
(b) Case II. Here we envision an atom with a perfectly sharp lower state,
and an upper state whose finite lifetime against radiative decay (back to the
13-1 Redistribution in the Atom's Frame lower state) leads to a Lorentz profile

Let us first consider the nature of redistribution in the rest frame of the atom.
f(f) = («5/n)/[(e' - eo) 2 + «5 2 ] (13-3)
Let the frequency displacement from line center of the incoming photori, where «5 = r RI4n, and r R is the radiative damping width of the upper state.
measured in the atom's frame, be denoted by e', its direction by n', and the We assume that there are no additional perturbations of the atom while it is
frequency displacement and direction of the outgoing photon by and n. e in its upper state; then there will be no reshuffling of electrons among sub-
Assume that the material has no preferred directions on an atomic scale so states of the upper state, and the decay to the lower state will produce a
that the atomic absorption profile f(e') is isotropic; f is normalized such that photon of exactly the same frequency as was absorbed. Thus we again have
f~ oo f( e') de' = 1. Further, suppose that the frequency redistribution function p(e', e) = «5(e - e'). This case applies to resonance lines in media of such
p(e', e) gives the probability that a photon absorbed in the frequency range low densities that collisonal broadening of the upper state is completely
(e', e' + de') is emitted into the range (e, e+ de), while the ang~lar phase negligible-for example, the Lyman a line of hydrogen in the interstellar I
function g(n', n) describes the probability that the photon is scattered from medium. N
solid angle dw' in direction n', into solid angle dw in direction n. These funC: (c) Case III. The basic physical picture here is of an atom with a perfectly .....
lions are normalized such that sharp lower state, and a broadened upper state, in a medium where collisions 01
f~ P(e', e) de' = f~ p(e', e) de =
00 00
1 (13-1)
are so frequent that all excited electrons are randomly reshuffled over the
substates of the upper state before emission occurs. The absorption profile
I

and f g(n', dw' = f g(n', dw =


(41t)- 1 n) (4n)- 1 n) 1 (13-2)
is again the Lorentz profile given by equation (13-3), where «5 now represents
the full width (radiative plus collisional) of the upper state. In this extreme
limit, the frequency of the emitted photon will have no correlation with the
The phase functions most useful in describing atomic scattering are those for frequency of that absorbed; the probability for emission at any particular
isotropic and dipole scattering, given by equations (2-17) and (2-18). frequency is then proportional to the number of substates present at that
In terms of the functions just defined, we may now write the probability frequency, and hence to the absorption profile itself. When complete redis-
that a photon (e', n') is absorbed as f(e') de' dw'/4n, and the probability that,
tribution in the atom's frame occurs we thus have
if a photon (e', n') is absorbed, then a photon (e, n) is emitted as p(e', e)g(n', n)
de dw/4n. Thus the Joint probability that a photon (e', n') is absorbed and a p(e', e) de = f(e) de = («5/rt) de/[(e - eo>2 + «5 2] (13-4)
photon (e, n) is emitted is f(e')p(e', e) de' de g(n', n)(dw'/4n)(dw/4n). which shows clearly that p(f, e) is independent of e', and that the joint
We must now specify the functions f(e') and p(e', e). Following Hummer, probability of absorption at f and emission ate is proportional to f(e')J(e).
we shall consider the following four categories: (a) Case I, zero line width;
(b) Case 11, radiation damping in the upper state, and coherence in the atom's (d) Case IV. Here we suppose that the line is formed by an absorption
rest frame; (c) Case III, complete redistribution in atom's frame; (d) Case from a broadened state i, to a broadened upper state}, followed by a radiative
IV, subordinate-line redistribution between two broadened states. Let us decay to state i. This picture is appropriate to scattering in subordinate lines.
examine these in turn. Because the electron returns to the same level as that from which it was
414 line Formation with Partial Frequency Redistribution 13-2 Doppler-Shift Redistribution 415

excited, the entire circuit is treated as a single quantum-mechanical process the typical situation of interest for resonance-line formation is one in which
and .a? expressio_n, is obtained for the product J(e')p(e', e). An expression fo; the upper state is broadened by both radiation damping and elastic collisions,
the Jomt probab1hty of absorption and emission was derived by Weisskopf with damping-widths oR and Oc respectively. In this case, one would expect
(66~) and by Woolley (683). A subsequent analysis by Heitler (293, 198) yielded the line profile to be given by equation (13-3) with c'i = OR + c'ic, Of the atoms
~ different result th~t was widely quoted in the astrophysical literature, but in the upper state, a fraction y = c'iR/(c'iR + Oc) would be expected to decay
1s now known to be m error (489, 195); the earlier formula is in fact correct radiatively, and hence to emit coherently in the atom's frame (recall the lower
T~e derivation is straightforward but lengthy, so only the final ex,pressio~ state is perfectly sharp). The remaining fraction (1 - y) = c'icMR + c'ic)
will be quoted here, namely would have suffered collisions, and would be expected to have been com-
pletely redistributed. Thus we could now write
f(f)p(f. e)
p(e', e) = 'I c'i(e' - e) + (1 - y)(0/7t)/[(e - eol2 + c'i 2 ] (13-6)
= (o,2 01/1t2)4(o, + 6)/{[(e' - eo)2 + 62][(e - eo)2 + 62][(e' - e)2 + 4o,2]}
2 This result was derived by Zanstra (690; 691), who treated the radiating atom
+ (0101/1t )/{[(e - eo) + 6 2][(e - e') 2 + 4ci,2]}
2
as a classical oscillator. A detailed quantum-mechanical calculation (489)
+ (010i/1t 2)/{[(e - e') 2 + 4o,2]({e' - eo) 2 + 6 2]} recovers equation (13-6) when the lower state is presumed to be perfectly
sharp. If inelastic collisions occur with sufficient frequency, ci,, to contribute
+ (c'i,2/1t 2 )/{[(e' - eo) 2 + Li 2]({e - eo) 2 + 6 2 ]} (13-5)
to the total width of the state, then c'i = oR + c'ic + c'i 1; we now write y =
whe:e Li = ci, + 01 [see (684, 164-168) for a detailed derivation and dis- (c'iR + 01 )/(c'iR + c'ic + 01 ), for it is only the elastic collisions that reshuffle
cuss10n]. atoms among upper-level substates (489). In this case it is also necessary to
It is easy t? show that f(e')p(e', e) = J(e)p(e, e'). as would be expected in introduce an additional emission source accounting for collisional excitations
the symmetrical process i -+ j -+ I. Further, by inspectio11 of the denomi- of the upper level (see §13-4). In sum, we see that the redistribution function
!s
nators in_equati~n (13-5), it easy to see that for a given value of~. there are in this more general situation can be expressed as a linear combination of
the results for Cases II and III.
two relative maxim~ one with e• = e and the other with e• = eo, These can
be understood physically as follows. (1) Most absorptions will arise froI)l I
the center of the lower level; when these excite a particular substate of the I\)
u~per state, the d~cay occurs most often back to the center of the lower level. 13-2 Doppler-Shift Redistribution ......
It_ 1s clear tha~ ~ will then equal~•• and that t~is process occurs with a relatively in the Laboratory Frame 0)
high pr~bab1hty.. (2) Alternatively, there 1s a very high probability that an I
atom will be excited from the center of the lower state to the center of the GENERAL FORMULAE
e
up~r by photons with e' = 0 ; the upper state may then decay to any
~rb1trary substate of the l~~er, in particula~, yielding the frequency e. The Let us now consider the effects of the Doppler shifts introduced by the
mver~e also occurs: a trans1t1on from an arbitrary substate of the tower level motion of the scattering atoms relative to the laboratory frame. In this section
has high probability of exciting the center substate of the upper level and we shall derive expressions that describe the full angular andfrequency depen-
th_e most prob~ble ~ec~y is back to the center of the lower level. Thus ~here dence of redistribution in the scattering process. In practice, relatively few
will be a p~ak m em1ss10n at line center (i.e., at e = eo), and indeed one sees radiative transfer calculations of relevance to line formation in stellar atmo-
that there 1s such a peak in equation (13-5). spheres have been done with the scattering treated in this much detail (the
The laboratory-fram~ redist_ribution function corresponding to equation dimensionality of the problem is large!), and the angle-averaged functions to
(13-5) _has not appea~ed m the hterature [ owing to use of the erroneous result be derived in §13-3 are generally much more useful in application. Following
of He1!ler], though 1t could be derived straightaway (the calculation would Hummer's treatment, we first deduce general formulae for the observer's-
~e tedious, how~ver). ~oreo~er,. the _for~ulat!on of a correct transfer equa- frame redistribution function, and then calculate explicit results for the
tion for subor~mate-hne red1stnbut1on 1s quite complicated; we shall not, specific cases defined above. As was discussed in §2-1, the redistribution
therefore, consider Case IV further in what follows. function R(v', n'; v, n) dv' dv(dw'/4n)(dw/4n) gives the joint probability of
0~ the _four cases defined above, Cases II and III are of the greatest astro- scattering a photon from a laboratory frame frequency (v', v' + dv'), and
physical importance. In fact, neither of these extreme limits is attained, and solid angle dw' in direction n', into frequency (v, v + dv), and solid angle dw
416 line Formation with Partial Frequency Redistribution 13-2 Doppler-Shift Redistribution 417

in direction n. The function is normalized such that [ equation (2-7)] Resolve v along these axes, and write v1 = v · n1; for convenience express
velocities in dimensionless thermal units
(4n)- 2 pdw' pdw f dv' f 0
"'
0
"' dv R(v', n'; v, n) =I u = v/V1hermal = (mA/2kT)¼v (13-11)
Suppose an atom moving with velocity v, which remains fixed during the where mA is the mass of an atom, and introduce the Doppler width
scattering process, absorbs a photon (v', n') and emits a photon (v, n), as
measured in the laboratory frame. Neglecting the aberration of directions in w = (v 0 /c)(2kT/mA)t = Vo(V1hermai/c) (13-12)
transforming from the atom's frame to the laboratory frame, the correspond-
ing atom's frame frequencies for the absorption and emission are Then the Maxwellian velocity distribution [equation (5~1)] is
2
e' := v' - v0 (v · n')/c (13-7a) P(u 1, u2 , u3 ) du 1 du 2 du 3 = n-t exp[ -(u 12 + u/ + U3 )] du 1 du2 du3
and e= V - Vo(V ' n)/C (13-7b) (13-13)

As was shown in §13-1, the joint probability of absorption ofa photon (e', n1 We now average equation (13-8) against the velocity distribution of equation
with subsequent emission of a photon (e, n) measured in the atom's frame is (13-13) to obtain
f(e')p(e', e)g(n', n) de' de(dw'/4n)(dw/4n). Transforming this expression to
the laboratory frame via equations (13-7) we can write R(v', n'; v, n) = f_:'°.., du 1 f_:'°.., du 2 J:°.., du 3 P(u,, u2 , U3)Ru(v', n'; v, n)
R.(v', n'; v, n) = f(v' - v0 v · n'/c)p(v' - v0 v · n'/c, v - v0 v · n/c)g(n', n) (13-8)
= n- 1g(n', n) f~.., du 1 e-u,i f~.., du 2
where the subscript v implies that the redistribution is produced by an atom
of velocity v. To find the net result for the entire ensemble o( atoms, we must x e-• 22.f[v' - w(ixu 1 + flu2)]
average over the velocity distribution, which is assumed to be Maxwellian. x p[ v' - w(ixu 1 + {lu 2), v - w(ixu 1 - Pu2)] (13-14) I
To perform this average, introduce an orthogonal triad of reference axes I\)
(ni, n2 , n3 ), with n 1 and n2 chosen to be coplanar with n' and n, and with n 1' where the integration over u3 has been carried out explicitly. An alternative
form for R, which will prove useful, can be derived by choosing n 1 to tie along
.....
bisecting the angle 0 between them (see Figure 13-1). Then we may write· -...,J
n'. Then v · n' = v1, while
I
(13-15)
and (13-10)
We then find

R(v', n'; V, n) = n- 1g(n', n) s~<X> du, e-••2.f(v' - wu,)

x f_:'° du 2 e-u22 p[v' - wu 1, v - w(ix'u 1 + P'u 2 )] (13-16)


00

FIGURE 13-1 With the aid of these general formulae, we can now compute the redistribu-
Coordinate axes used in calculation or tion functions for the various cases defined in §13-1. Note first, however, that
redistribution function. The vectors n., n 2 , n, in both Cases I and II the scattering is coherent in the atom's frame, so that
and n' are coplanar. The vector n 1 bisects the
angle 6, (0 ,;; 6 ,;; n), between n' and n. (13-17)

In these cases u 1 no longer appears explicitly in the expressio~ for p, and_ the
integrations written above are simplified. Substituting eq_u~t1on (13-17) 1~to
(13-14), the integration over u2 can now be performed ex~hc1~ly (transfor~mg
to the variable z = 2wflu 2 in order to preserve normahzation of the Dirac
418 Line Formation with Partial Frequency Redistribution 13-2 Doppler-Shift Redistribution 419

a-function). We then obtain, for p :/= 0, From this result, we immediately notice that, even though the scattering in ·
the atom's frame was presumed to be strictly coherent, there is significant
2 redistribution in the laboratory frame. This is a point of very great physical
R (1•' n'· v n) = g(n', n) exp [ (v - v') ]
2 2
• ' ' ' 21r.Pw 4{J w significance, as was stressed by Thomas (622). He examined the frequency
dependence of scattered radiation [cf. equation (2-9)] using several simple
(13-18) analytical expressions for the incident intensity, and showed that in the
Doppler core (x :5 3), over which the absorption profile varies by a factor
while if f3 = 0 we find of 104, the emission profile departs from the absorption profile by less than
a factor of four! Thus, within the Doppler core, the assumption of complete
R,(v', n'; v, n) = 1r.-¼g(n', n) o(v' - v) f~
00
e-"1(v' - wu) du (13-19) redistribution of the scattered radiation provides an excellent approximation
(it will be easy to see this using the angle-averaged redistribution functions
Here the subscript c denotes coherence in the atom's frame. derived in §13-2; see also Figure 13-2). This result provided much of the
motivation for assuming complete noncoherence in the treatment of line
Exercise 13-1: Verify equations (13-18) and (13-19). formation, as was done in Chapters 11 and 12. Outside the line core, complete
redistribution is not achieved as closely, and one must use the correct redis-
Finally, we note that the results can be written in a concise and convenient tribution function. For example, in the case considered here, a simple physical
form by expressing frequency displacements from line center in Doppler argument (597) shows that the radiation in the far line-wings is about two-
units: · ' thirds noncoherent, and one-third coherent. On the other hand, outside the
x = (v - v0 )/w. (13-20) Doppler core the line-wings will often be swamped by the continuum·(unless
the profile has a large damping parameter), in which event the details of
and x' = (v' - v0 )/w (13-21) redistribution become less significant.

and writing (b) Case II. Here we substitute equation (13-3) into (13-18) to obtain
2
I
R(x', n'; x, n) = R(v', n'; v, n)(dv'/dx')(dv/dx) = w2 R(v', n'; v, n) (13-22) Rn(v,' n, ; v, n) g(n', n) [-(v - v') ] ( o) I\)
= 21r.cx/3w2 exp 4p2w2 ncxw
to yield normalization of R(x', n'; x, n) when integrated over x' and x. co
xs~ao e-• [(v' +;C(: 2v uy + (:wYTI du
2 0
- (13-25) I
RES UL TS FOR SPECIFIC CASES
Transforming to dimensionless frequency units via equations (13-20) through
(a) Case I. If we substitute f(f) = o(e' - v0 ) into equation (13-18), we
(13-22), and recalling the definition of the Voigt function H(a, v) [see equation
obtain immediately
(9-34)], we obtain, finally
2
, ,, _ g(n', n) [-(v - v') ] [-(v + v' - 2v0 ) 2 ] g(n', n)
R,(v, n 'v, n) - 21r.cx/3w2 exp 4p2w2 exp 4cx2w2 (13-23) Ru(x', n'; x, n) =~ exp[ -½(x - x') 2 csc 2 ½0]
1r. sm 0•
Now noting that 2cx/J = 2 sin(½0) cos(½0) = sin 0, that cx 2 + {3 2 = 1, and x H[a sec½0, ½(x + x') sec ½0] (13-26)
transforming to dimensionless units via equations (13-20) through (13-22),
we have · where a = (o/w). Although this result is relatively complicated, efficient
methods to evaluate H(a, v) exist, and Ru can be calculated fairly easily.
R1(x', n'; x, n) = [g(n', n)/1r. sin 0] exp[ -x 2 - (x' - x cos 0) 2 csc 2 0] As will emerge in §13-3, Ru simulates nearly complete redistribution in the
(13-24) line core (x $ 2.5), but becomes nearly coherent in the line wings.

Exercise 13-2: Fill in the steps required to obtain equation (13-24). Exercise l 3-3 : Fill in the missing steps leading to equation (13-26).
420 Line Formation with Partial Frequency Redistribution 13-2 Doppler-Shift Redistribution 421

(c) Case Ill. Here, as was noted in equation (13-4), p(e', is independent e) Now for all four cases defined in §I 3-1, the scattering is symmetric in the
off. In this case equation (13-16) is particularly useful, and substituting atom's frame-i.e.,
from equations (13-3) and (13-4) we find
(13-30)
R111 (v', n'; v, n) = g(n~ n) f:' 00
du 1 e-u,z (~) [(v' - v0 - wu 1) 2
+ c5 2
]-
1
Suppose we replace x' ... -x' and x ... -x in equation (13-29); ifwe simul-
taneously change the signs on u 1 and u2 [which we may do because the
X s~"' du2 e-uzZ G) {[v - w(u, cos 0 + U2 sin 0) - Vo]
2
+ 02 }- 1 (13-27) integrals span the entire range ( - oo, oo)], we immediately see from equation
(13-30) that
R(-x', n'; -x, n) = R(x', n'; x, n) . (13-31)
Converting to dimensionless units, and again using the definition of the
Voigt function, we find Next, noting that replacing n' ... - n' and n ... - n changes the sign of ex
and /J, and recalling that g(n', n) depends only on n' · n, it is clear from
, ,. ) _ g(n', n)
R111 (x , n , x, n - a
f"' 2
e-" H(e1, x csc 0 - u ctn 0) d
(X , ) u
( _ )
13 28 equations (13-29) and (13-30) that
1[
2 -00 - U 2 + a2
R(-x', -n'; -x, -n) = R(x', n'; x, n) (13-32)
where a ;s a csc 0 and a ;s (c5/w). This result is no longer expressible in
terms of simple functions, and must be evaluated by numerical integrations and from (13-31) and (13-32) it follows that
(529).
R(x', -n'; x, -n) = R(x', n'; x, n) (13-33)
Exercise J3-4 : Fill in the steps required to obtain equation (13-28).
Exercise I3-5 : In the limit that Compton scattering effects are negligible (i.e., Equations (13-31) through (13-33) depend only on the validity of (13-30)
hv/mc 2 « I where m is the mass of an electron), the scattering of radiation by and hence are rather generally true. Further, because R depends only on
electrons is coherent, p(f, ~) = c5(f - ~). and grey, /({) = u, = (81tt'/3m c ).
2 4 the angle between n' and n, we may interchange them without changing R;
(a) Show that the redistribution function for electron scattering is i.e., I
mc 2
]¼ [ -mc (v - v') ]
2 2 R(x', n; x, n') = R(x', n'; x, n) (13-34) I\)
R,(v', n'; v, n) = g(n', n) [ 41tkT(l - cos 0)v 2 exp 4kT(l - cos 0)v 2
Next we notice that, for coherent scattering in the atom's frame (Cases I (0
[See (196; 475; 161; §86)). (b) Show that for T = 10'' °K, ). = 4000 A, 0 = and II), the redistribution functions depend only upon (x + x') and Ix - x'I
n/2, the characteristic width of the above redistribution is about 10 A. This result [cf. equations (13-23) and (13-26)]; hence x and x' may be interchanged
I
suggests that spectral lines in early-type stars can be significantly broadened by without altering R; i.e.,
electron scattering (159; 475; 161, §86; 318; 39).
R1(x, n'; x', n) = R1(x', n'; x, n), (i = I, II) (13-35)
SYMMETRY PROPERTIES
Finally, a transformation from n' ... - n' changes the sign of ex and {J in the
The functions obtained above have certain symmetry properties which expressions for incoming photons, hence if g(n', n) is an even function of
often may be exploited to simplify the form of the transfer equation. To n' · n,
obtain these, it is convenient to rewrite the general result (13-14) in terms
of x and x'; let /(v - v0 ) ;s f(v) and p(v - v0 , v' - v0 ) ;s p(v, v'). Then R(-x', -n'; x, n) = w2 x- 1g(n', n)f"' du 1 e-u,z f""-er:> du 2
-ex>

from equations (13-14), and (13-20) through (13-22) x e-v,'J[w(-x' + exu 1 + /Ju 2 )]p[w(-x' + exu 1 + {Ju 2 ), w(x - exu 1 + /Ju 2 )]
(13-36)

x e-•,'l[w(x' - exu 1 - /Ju 2 )]p[w(x' - exu 1 - /Ju 2 ), w(x - exu 1 + /Ju 2 )] which will reduce to (13-29) in certain cases. In particular, the integrals will
be equal if!(~") = c5(~') or if J(f>p(e', ~) = /(- ~')p(-e', ~). The former is
(13-29) true for Case I, while the latter is true for Case III where !(O = J(- ~')
422 Line Formation with Partial Frequency Redistribution
13-3 Angle-Averaged Redistribution Functions 423

and p(e', e) = l(e). Hence we conclude that emissivity from scattering processes can be written in the form

= o-(r) f_:'°
R1(-x', -n';x,n) = R1(-x',n';x, -n) = R 1(x',n';x,n), (i = I,III) (13-37) 5
17 (r, v) 00
R(v', v)J ,· dv' (13-38)
where the second identity follows from equation (13-33).
where J,. is the mean intensity and R(v', v) is the angle-averaged redistribution
function
APPLICATIONS

The solution oft he transfer equation allowing for full angle and frequency
R(v', v) = (41t)- 1 pR(v', n'; v, n) dw' (13-39)

redistribution is fairly difficult, and requires techniques more powerful than The rationale for this approach is essentially as follows. The radiation
those discussed in this book; both Monte Carlo methods (31; 50) and field will depart significantly from isotropy at any given frequency only at
difference-equation methods (460) have been developed to handle the prob- · points whose optical depths t, from the surface are of order unity or less;
lem. There are a number of situations where both angular and frequency at depths t, ~ 1, the radiation field is essentially isotropic. Now a basic
effects may become important. For example, the escape of In. photons from characteristic of non-LTE line-formation that emerged clearly in Chapter 11
a very thick nebula depends sensitively upon the details of the scattering is that the surface value of the source function is determined by photons
process, and a complete, careful treatment is required (31). In addition, contributed from over an entire destruction length, and over virtually all
when the material has a macroscopic velocity, both the absorption and of this region the radiation field will, in fact, be isotropic. We may expect,
emission coefficients become angle-dependent in the observer's rest frame therefore, that even at the surface where I(µ, v) shows departures from
(cf. §2-1). There then results an inextricable coupling between angles and isotropy, S, will still have a value already fixed by processes occurring at
frequencies in the transfer problem (see §14-1), and the detalls of the redis- depths where the anisotropy is negligible; hence the value of l(t, = 0, µ, v)
tribution process may be of consequence; in this book, however, Yle shall computed from this S, should be quite accurate. On the other hand, use
consider only the case of complete redistribution in moving media, and will of the angle-averaged functions accounts completely for the frequency-
not pursue this point further. Finally, when observable angle-dependen
I
reshuffling of scattered photons, the action of which (as we saw in Chapter r\)
information is available (e.g., center-to-limb variations as measured in the 11) affects crucially the photon escape-probability, and hence the thermal- r\)
solar atmosphere), the possible importance of angular effects in the redis- ization process. In short, in this approach we account for the critical aspects
tribution process must be examined. The investigations of this question 0
of the redistribution process, and sacrifice information only in an area of I
conducted thus far (640; 641; 460) show that, at least for semi-infinite plane- secondary importance.
parallel atmospheres with homogeneous layers, the differences between re- The functions defined by equation (13-39) are normalized such that
sults obtained using angle-averaged redistribution functions and those
obtained using the full angle-dependent functions are negligible. On the other
hand, for optically thin media or small-scale structures (e.g., chromospheric
f~ dv f~
00 00
dv' R(v', v) = 1 (13-40)
fine-structure, spicules, etc.) the angular effects arc often quite important Integration over all emitted photons yields the absorption profile
(186), and should be considered in detail. However, these questions lie beyond
the scope of this book, and will not be discussed here; rather, we now turn
to the derivation of angle-averaged redistribution functions, and then to
clv' f~ R(v', v) dv = ip(v') dv'
00
(13-41)

their use in radiative transfer calculations. while integration over all absorptions yields the natural-excitation emission
profile [cf. equation (2-14)]

13-3 Angle-Averaged Redistribution Functions l/t*(v) dv = dv f~ R(v', v) dv'


00
(13-42)

As was noted in §2-1, ifwe assume that the radiation at any particular point
in the atmosphere is essentially isotropic, then the contribution to the
If R(v', v) = R(v, v') (we shall find this to be true for cases I-III), then clearly
l/t*(v) = ip(v). It should be emphasized, however, that l/t*(v) is not, in general,
424 Line Formation with Partial Frequency Redistribution 13-3 Angle-Averaged Redistribution Functions 425

the actual emission profile in the line when J. varies over the line profile equation (13-45) we have
(sec the discussion in §13-4); that profile follows, ultimately, from the equa-
tions of statistical equilibrium and knowledge of the frequency-variation of
J •.
R.(v', v) = (l61t 2 )- 1 ft dq, f~ 1
dµ' f(v' - wµ'u)
Two hypothetical limiting cases sometimes considered are strict coherence
and complete redistribution in the laboratory frame. In the former circum-
x f~ 1
dµ p(v' - wµ'u, v - wµu) ft d<f,' g(µ', µ, <f,') (13-46)

stance we would have Define g(µ', µ) = (47t)- I ft g(µ', µ, <f,') d<f,' (13-47)

R(v', v) = <f,(v') c5(v - v') (13-43) Then, noting that the integral .over <f, in equation (13-46) is trivial, we have
and in the latter
R(v', v) = <f,(v')<f,(v) (13-44)
R.(v', v) f
= ½ ~ 1 dµ' f(v' - wµ'u) f I
dµ g(µ', µ)p(v' - wµ'u, v - wµu)
Neither of these limits can actually ever be achieved. As we have seen already, (13-48)
even if the scattering is strictly coherent in the atom's frame it is not in the·
laboratory frame unless the atoms have zero velocities (impossible). Further, For simplicity, in what follows only isotropic scattering in the atom's frame
as we shall see below, complete redistribution in the atom's frame does not will be considered; in the case g(µ', µ) = ½[formulae for a dipole phase
produce exactly complete redistribution in the laboratory frame. However, function are given in (313)]. Applying this restriction we have
in the latter situation it is nonetheless found that equation (13-44) does, in
fact, provide a very good approximation to reality in most cases, and hence
may normally be used to describe that limit. Note in passing that, because
RA, .(v', v) f
= ¼ ~ 1 dµ' f
f(v' - wµ'u) ~ 1 dµ p(v' - wµ'u, v - wµu) (13-49)

the dependences of R on v' and on v are separated in equation (13-44), the Although equation (13-49) is general, it is not convenient in the case of
true emission profile t/1. = t/1: = <f,. regardless of the behavior of J..,, which coherent scattering in the atom's frame because of complications that arise
justifies our earlier use of this relation [but only if equation (13-44) is actually ih setting the limits of integration; Jet us, therefore, derive a more refined I
accurate!]. I\)
formula for this case.
I\)
If the scattering is coherent in the atomic frame, ......
GENERAL FORMULAE
p(v' - wµ'u, v - wµu) = c5[v' - v - wu(µ' - µ)] (13-50)
The integration indicated in equation (13-39) could, in principle, be
Because the range of integration forµ' andµ is only ( -1, 1), it is clear that
carried out directly, using each of the redistribution functions derived in
for a given value of u, the singularity of the c5-function will be outside the
§ I 3-2. This however, turns out to be rather complicated. It is simpler to
range of integration for sufficiently large values of Iv' - vi, and RA, .(v', v)
derive first a general formula, by performing the angle-average for arbitrary
will, accordingly, be zero. Physically this corresponds to the fact that an
f(f) and p(e', e), and then to obtain specific forms for the particular cases atom moving with velocity u can change a photon's frequency by no more
of interest. We begin by rewriting equation (13-8), using the Doppler units
than 2uw, this maximum shift occurring if the propagation vectors of the
defined in equations (13-11) and (13-12):
incoming and outgoing photons lie along the velocity vector and are op-
Ru(v', o'; v, o) = f(v' - wu · o')p(v' - wu · o', v - wu · o)g(o', o) (13-45) positely directed. We substitute equation (13-50) into (13-49), and consider
first the integration over µ. Let y = wµu, and write
We now wish to fix v' and v and integrate over all angle. Choose an ortho-
normal triad (o 1, o 2 , o 3 ) such that u = uo 3 • Then u · o = µu and u · o' = µ'u,
whereµ = o · o3 andµ' = o' · o3 • An element of solid angle may be written
I= (wu)- 1 r:. c5[y - (v - v' + wuµ')] dy (13-51)

dw = dµ d<f, where <f, is the azimuthal angle around o3 • The phase function The integral will equal (1/wu) if -wu ~ v - v' + wuµ' ~ wu, and will be
g(o', o) can be expressed in general as g(µ', µ, <f,). Thus, angle-averaging zero otherwise. Define A(x) such that A = 1 if -1 ~ x ~ 1, and A = 0
426 Line Formation with Partial Frequency Redistribution 13-3 Angle-Averaged Redistribution Functions 427

otherwise. Then equation (13-49) can be rewritten, using equation (13-51), as From equation (13-49) we have, for noncoherence in the atom's frame (e.g.,
Case III),
RA,u(v', v) = (4wu)- 1 f~ 1
f(v' - wuµ')A[µ' + (wu)- 1(v - v')] dµ' (13-52)
R.,.(v', v) = 7t-¼ f du u e-" f~ dµ' f(v' - wµ'u)
00
2 2
0 1
If u is sufficiently small, then l(v - v')/wul > 1 and A will vanish for all
values of µ'. Thus there is a minimum speed Umin, whose value we must x f dµ p(v' - wµ'u, wµu)
1
-1
v - (13-58)
determine, for which scattering from v' to v can actually occur. Define
ii= =
max(v', v) and.! min(v', v). First suppose that v > v'; then the require- Finally it is convenient to use Doppler units [cf. equations (13-12), (13-20),
ment that the argument of the A-function fall in the range ( -1, 1) implies and (13-21)] and write · ·
that [(v - v')/wu] - 1 = [(V - 1)/wu] - 1 ~- 1; this inequality yields
R.,.(x', x) = R.,.(v', v)(dv'/dx')(dv/dx) = w2 R.,.(v', v) (13-59)
Umin = (V - .l!)/2w = Iv -. v'l/2w (13-53)
But the same result is obtained by similar reasoning if we assume v' > v; RES UL TS FOR SPECIFIC CASES
therefore equation (13-53) is general. For u ~ Umin, Ru will be zero. For (a) Case I. Here f(y) = 8(y - v0 ); hence the integral over yin equation
u > Umin, a contribution to R. will come from part of the range of integration
(13-57) is nonzero only if.! + wu ;.i: v0 ;.i: V - wu. This implies that Umin now
overµ'. To determine this range, we suppose, for definiteness, that v > v'. becomes effectively u;,,1n = max(lx'I, !xi), which clearly satisfies the inequality
Then a contribution is obtained when u;,, 1n ;.i: Umin as given by equation (13-53). Then from equations (13-57) and
-1 ~ µ' ~ 1 - [(v - v')/wu] = 1 - [(V - 1)/wu] (13-59)
' (13-60)
which implies that V - wu ~ 1 - wuµ' ~ .! + wu; recalling that f = v' by
hypothesis, the result just stated can be rewritten as I
where the complimentary error function is defined as I\.)
V - WU ~ v' - wuµ' ~ !. + WU (13-54)
(13-61)
I\.)
I\.)
Ifwe assume instead that v' > v, we again obtain equation (13-54), which is, I
therefore, general. Substituting for u;,,;n,
Now introducing the Heaviside function cI>(x, x0), defined such that cI> = 1
when x > x 0 and cI> = 0 when x < x 0 , making the substitution y = v' - wuµ', R1, ,4(x', x) = ½erfc[max(lxl, Ix'!)] (13-62)
and using equations (13-53) and (13-54), equation (13-52) may finally be
written as This redistribution function is easy to compute from well-known approxi-
mation formulae for erfc(x) (4, 299); asymptotic formulae and results for
R.,., .(v', v) = (4w 2 u2 )- 1cI>(u - Iv - v'l/2w, 0) J.r_+:,: f(y) dy (13-55) dipole scattering are given in (313).
A plot of R1,.,.(x', x)/<f>(x') is shown in Figure 13-2; the curves are labeled
Finally, the results expressed in equations (13-49) and (13-55) must be with the incoming photon frequency x', and give the probability for emission
averaged over the Maxwellian velocity distribution at frequency x, per absorption, as a function of x. We see that a photon
absorbed at frequency x' will be emitted with equal probability at all x such
(13-56) that - lx'I ~ x ~ lx'I, and with exponentially decreasing probability beyond
From equation (13-55) we then have for coherence in the atom's frame this range. It is easy to understand this result. The absorption can occur only
(Cases I and II) when e• = O in the atom's frame. Therefore atoms absorbing at frequency;,.'
in the laboratory frame have a velocity of at least x' Doppler units. As the
R.,.(v', v) = (7t¼w2) -1 I.a, 2 rx+wu
du e-u J,-wu f(y) dy (13-57)
emissions occur with equal probability in all directions, the photons can be
redistributed with equal probability O\'er the entire range ± x·.
Umoo
428 429

1.0 1.0

x' =0
,:--
.!:l. ~
-&
::::. o.s -&
::::.
>< >: 0.5
:!£
~ :ic,:

X X
FIOURE 13-2 FIOURE 13-3
Probability or emission at frequency x, per absorption at frequency x', for pure Probability or emission at frequency x, per absorption at frequency
Doppler redistribution. Ordinate: R1(x', x)/q>(x'): abscissa: emission frequency x. x', for Doppler redistribution from a coherently scattering (in the
Curves are labeled with frequency x' of absorbed photon. atom's frame) Lorentz profile with a• 10- 3• Ordinate:
R11 (.~', -~)/cf,(x'); c,bscissa: emission frequency x. Curves are labeled
with frequency x' of absorbed photon.
(b) Case II. Substituting equation (13-3) into (13-57), we have

R11 • ..c(v', v) = (w2n•)- 1 l, J.:,. due-•• f;_+:: dy [(y - vO) 2 + l, 2)- ! (13-63) emissions are from atoms absorbing at line center and moving with velocities
near x'. For large x' there will be very few atoms with high-enough velocity
Converting to Doppler units, writing z = (y - v0 )/o = x/a, and writing to absorb at line center, and most emissions then come from atoms moving I
I\)
um1n = ½Ix - x'j from equation (13-53) we have with low velocities and absorbing in the line-wing. As the scattering process
I\)
is coherent in the atom's frame, and the appropriate atoms are nearly at
( , X)
R11, AX, I"'
= 7t -i Jix'-xl/l d -• 2 f(l!i+u)/ad ( 2 -1 rest in the laboratory frame, the scattering will be nearly coherent in the w

e: ")- e
U e J(ii-u)/a Z 1 + Z )
laboratory frame as well. Thus in the line-core there is Doppler redistribution I
= 7t-i fi;_,. 112 e-•'[tan-
1
tan- 1
~ ")]du and strong noncoherence, while in the wing the scattering is more nearly
coherent; this dichotomy has important implications for the line-transfer
problem, as we shall see in §13-4.
(13-64) In early work by Jefferies and White (338), it was suggested that a simple
=
where x max(jxj, jx'j) and ~ =
min(jxj, jx'j). The asymptotic behavior of approximation for R11 could be written in the form
R11 , and results for dipole scattering, are given in (313); an accurate method
for the evaluation of R11 is given in (6). R11 , ..c(x', x) ~ [1 - a(x)]</>(x')</>(x) + a(x)</>(x) o(x - x') (13-65)
The redistribution function R11 is of great interest, for it describes the where a(x) is nearly zero for x ~ 3 and approximately unity for x ~ 3. The
important case of scattering by a resonance line that is broadened by original suggestion for the form of a(x) is not adequate, however, and fails
radiation damping, and it has been extensively studied. A plot of to meet requirements of normalization, symmetry, and wing-coherence.
R11 ,..i(x', x)/</>(x') for a= 10- 3 is shown in Figure 13-3; again, the curves Nevertheless it is easy to define an appropriate function (358), and an
are labeled with the incoming photon frequency x', and give the probability, approach of this type does simplify the calculation somewhat, although in
per absorption, of a subsequent emission at frequency x. Here we see that the most precise work the correct form for R11 [equation (13-64)] should
for small x' (x' ~ 3), the curves resemble those for R., because most of the be used.
430 Line Formation with Partial Frequency Redistribution

(c) Case III. In this case we no longer have coherence in the atom's frame, :t
and equation (10-55) may be used without complication to give
~
R v' v - ! fl (l>/1t) dµ' fl (l>/1t) dµ
.5 .•
A,u( ' ) - 4 -1 (v' - wµ'u - v0 ) 2 + c5 2 -1 (v - wµu - v0 ) 2 + c5 2 ~~-
0 )<
~ ;s.~~
[ tan - t (x'+u) (x'-u)] ~~~
= 41t2w2u2
1 --a- - tan -1 --a- I:! - [
0 a: -
00
-; \i C
.. -I:,·-
0

[tan _ (x+u) _ (x-u)] )< ?t ,5 .fa


x 1
- -
- tan 1
- - (13-66)
,c
0 1!
cc 5
0 0 0 ..
'§..,· ~
Averaging over a Maxwellian velocity distribution and converting to Doppler
units we have st
- - ..,
,&> I
·co •-
::I
.!!! II"'=
il cl
~ ~~-2

= TC -½ Jof"" e-u2[tan - t (x'-a-


+ u) - _ (x'-a-
- u)] -~-1
.!!-~·
R 111 • ix,, x) tan 1 N
s~ -5. cgs.
o. II u

~~
.g 'li g ...i:
(13-67)
.
I
0 -..; ,.; 'ilj t:l
~ ~ ·e !j,
~ = ..= "e
:!: C e
!j i:°'
8" g .gt.i..
Asymptotic formulae, results for dipole scattering, and computational I!: tU !! I
methods for evaluation of R111 are given in (313,212; 529). ,
i,j _g 'tl g I\)
~ g~ ~ 0
A plot of R 111 (x', x)/<f,(x')
for a = 10- 3 is shown in Figure 13-4. For small e- ,s 'ij
·- u .cl
0. I\)
x', most absorptions are at line center by atoms moving with velocities near
x', hence redistribution occurs with equal probability over the range ~
] ·5
tU C: u
! ij ~
... .9 .c I
- x' :!6; x :!6; x', as was true for R1 and R11 • Again, for large x' most absorptions 8:] ,; :!
)'(·cu o
occur in the line wings of nearly stationary atoms, but now the emitted 00 ~-~ ~ '><
photons are completely redistributed over the absorption profile in the C -0 ~
I:! al
atom's frame, hence R111 (x', x)/<f>(x') .... <f>(x)
for x' » 1. It has, on occasion, ><
u
&1:;_6
~c.85-
been argued on intuitive grounds that if the redistribution process is com- ,c
~ 8 . :;
- 0. "
pletely noncoherent in the atom's frame, and if this is combined with random
Doppler motions, then the redistribution should be completely noncoherent
st
-~ -g
·e ~ u]
f'i
in the observer's frame as well; this conclusion is false, however, as is clearly
shown by Figure 13-5. In fact, Doppler motions introduce a correlation ~ ~ .§
,. ~ ,s -
·e
» ·c I:!
between incoming and outgoing frequencies near the line core, and the N - :::.':: ·"'e l1li .,
...,
deviations from complete frequency redistribution in the laboratory frame ~ .g !LS! rJ
=> -g ., u e:
can be large. Despite these deviations, it turns out (see §13-4) that the as- ~~-sga
sumption R(x', x) = <f>(x')<f>(x) produces line profiles quite similar to those
I
...I .
I
0

obtained from the exact R 111 (x', x), and in practice, the case of complete =! =! =!
noncoherence in the atom's frame i;nay be treated as complete noncoherence
(.x)4>/(x ',x) 111 ~
in the laboratory frame also, without serious errors.
432 13-4 Radiative Transfer with Partial Redistribution 433

1000--------------~
r,x' • I I
3
which is general so long as equation (13-30) holds. Further, by a direct
extension of the arguments leading to equation (13-35) one finds
I I
I I
I I R 1(x', x) = R 1(x, x'), (i = I, II, III) (13-69)
I I
I I
I I The validity of this result can also be seen by inspection of equations (13-62),
100 I I
(13-64), and (13-67). From equation (13-69) it follows that the natural-
I I
I I excitation emission profile [see equation (13-42)]
,::::, / I
~
~
x' I I 4 t/lf(x) = cp 1(x), (i = I, II, III) (13-70)
E
....,
~
10 Exercise 13-7: Verify the result stated in equation (13-70) for Case I by direct
;::,,
~ integration of R1, ix', x) over x'.

~
:I
r:,:
x' 13-4 Radiative Transfer with Partial Redistribution
' , ... -----
Consider now the problem of accounting for the effects of angle-averaged
partial redistribution in spectral line formation. For simplicity we shall con-
fine attention to a two-level atom with continuum, and develop a general
method that gives a rigorous solution of the problem. (This method can be
0.1 .___ ___..___ _.___ _.____ _.____ __, generalized easily to more complicated atomic models as well.) Next, a
0 2 4 6 8 10 somewhat simpler, but less general, approach will be developed, which still
X treats the troublesome stimulated emission term correctly. Then a still
FIGURE 13-5 simpler method, which treats the stimulated emission terms only approxi- I
Ratio or actual redistribution function, for complete redistribution in the atom's mately, will be described, and results from the application of this method in I\)
frame and Doppler redistribution in the laboratory frame, R111(x', x), to the highly idealized models will be discussed. Finally, results from calculations I\)
limiting case or complete noncoherence in the laboratory frame, ,t,(x'),f,(x). The ofresonance-line profiles, using the general method, for realistic model atoms 01
absorption profile ,t, is a Voigt profile with a = 10- 3, From (212), by
permission. and stellar atmospheres, will be presented. I

FORMULATION FOR A TWO-LEVEL ATOM

Exercise 13-6: Show that the angle-averaged redistribution function for scattering Let us examine the formation of a resonance line connecting a perfectly
by electrons (under the same assumptions as employed in Exercise 13-5) is sharp lower level to an upper state that is broadened into a distribution of
substates by both radiation damping and collisions. Denote the population
R,, ,i(v', v) = w- 1 ierfcl(v' - v)/2wl
of the lower level by ni, the total population of the upper level (summed over
where ierfc(x) = J:"" crfc(z) dz = 1r·+e-~• - x crfc(x) all substates) by n2 , and the number of ions by nK. The distribution of the
and w denotes the electron Doppler width, w .. v0(2kT/m,)¼/c, which is about atoms over the upper state is specified in terms of the observer's-frame
43A¼ times as large as the Doppler width of an atom of atomic weight A. Sec also emission profile t/J., defined as the fraction of all atoms in the upper state
(318) and (39). that, if they decay radiatively, emit photons of frequency v as seen in the
laboratory frame. This method of counting atoms in the upper state is well-
SYMMETRY PROPERTIES posed physically, for it describes the distribution in terms of observables;
mathematically, t/J. is a complicated, but unique, one-to-one mapping from
From equation (13-31) we see that, after angle-averaging, we must have
the distribution of atoms over their own rest-frame frequencies ~. and
R(-x', -x) = R(x', x) (13-68) velocities v, the transformation being specified by the redistribution function
434 line Formation with Partial Frequency Redistribution 13-4 Radiative Transfer with Partial Redistribution 435

(456). Clearly t/,, is normalized, and n2 f


= n2 t/1, dv. We shall find it con- but in practice it is adequate (and much easier) to use
venient to define an auxilliary variable ff 2 such that n2(v) = n2tJ,, = n2(v)cf,,
gives the number of upper-state atoms in substates that can emit photons of R(v', v) = yR 11 (v', v) + (1 - y)</>.,</>. (13-73)
frequency v; the dominant variation of the upper-state distribution is
factored out this way, and ff2 reflects only the departure from natural ex- which assumes that complete redistribution in the atom's frame leads to
citation. complete redistribution in the observer's frame.
The substate occupation number n2(v) (or, equivalently, t/,.) is specified The rate equation for the ground state is
by a rate equation of the form
n1(B12f<1>.J,,dv + C12 +RIK+ C1K)
n2(v)(A21 + B21J. + C21 + R2K + C2K)
= n2 (A21 + B21 f 1/!Jv dv + C21) + nT(RKI + c.K) (13-74)
= n, [ B12 f R(v', v)J•. dv' + C t/l:] + n!t/l:(RK
12 2 + C2K) (13-71)
which has an interpretation entirely analogous to that of equation (13-71).
This equation has a simple physical interpretation [see also (456)]. The term Finally, the total number of atoms is presumed known, hence
on the lefthand side is the population of the substate, times the total rate of
(13-75)
exit (a) to the ground state by spontaneous emissions, stimulated emissions
and colJisions, and (b) to the continuum by photoionizations and collisions.
The required photoionization and recombination rates may be considered
The only noteworthy term is that for stimulated emission which is strictly
as given, in which event only the radiation field in the line needs to be cal-
coherent in both the laboratory and the atom's frame. [This follows from the
culated, or these rates may follow from a solution of the transfer equation
fundamental quantum-mechanical characteristics of the process (197, §62;
in the continuum (assumed necessary for the ground state only).
293, §17), in which a photon, incident upon the atom, induces an emission
The transfer equation to be solved may be written as
in such a way as to create yet another photon of exactly the same properties:
energy (hence frequency), momentum (hence direction of propagation), and iJ2(f.J .)/ih.2 = J. - S, (13-76) I
polarization. Thus a photon (v', n') in the laboratory frame undergoes a I\)
definite transformation (depending on n' and v) to (e', n') in the atom's frame, where, as usual, drv =- Xv dz and Sv = ti.Ix •. In the line I\)
and creates another identical photon; both photons undergo exactly th'e 0)
same inverse transformation back into the laboratory frame, returning two Xv= (Xdn1 - (gifg2)ii2(v)]</>. + X, (13-77)
photons (v', n') into the radiation field.] The terms on the righthand side
I
describe processes of excitation to the substate n2(v), from the ground state and tlv = (2hv 3/c 2)cx12(gi/g2)ii2M</>v + Ev (13-78)
and by recombinations from the continuum. The latter produce atoms in and in the ground-state continuum
level 2 distributed according to the natural-excitation profile t/1:; n! denotes
the LTE particle density n! = nKn.ct> 2 (T), where ct>(T) is the appropriate Xv= (XIK(v)(n 1 - nTe-h•/kT) + Xv (13-79)
Saha-Boltzmann factor. For all three physically interesting cases of redis-
tribution considered in §13-3, t/1: = cf,, [see equation (13-70)]. From the and tlv = (2hv 3 /c 2 )e-hv/kT (X 1K(v)nT + E, (13-80)
ground state, atoms excited by collisions are also distributed according to
Here Xv and E, represent (fixed) background opacity and emissivity sources,
1/J~. The radiative excitations are given by the number of absorptions of
photons of frequency v', namely n1 B 12 J •. , times the joint probability R(v', v) while (X 12 = (B 12 hv)/41t.
The solution of the transfer equation is now quite complicated because
of absorbing at v' and emitting· at v, summed over all v'. In the scattering
the emission pro.file 1/1, is not known a priori, but follows from the statistical
process we assume that a fraction y of all excited atoms emit coherently,
and the remainder are completely redistributed in the atom's frame by equilibrium equations (13-71) through (13-75). Unlike the case of complete
redistribution, where only the ratio (n 2 /n.) is required to specify the source
elastic colJisions [so that p(f, e) is given by equation (13-6)]. The correct
function [ cf. equation ( 11-4)] and hence only one statistical equilibrium
laboratory frame redistribution function in this case would be
equation is needed, we must now compute i/t., and this introduces as many
R(v', v) = yR11 (v', v) + (1 - y)R 111 (v', v) (13-72) equations of the form of (13-71) as are required to define this function to
436 Line Formation with Partial Frequency Redistribution 13-4 Radiative Transfer with Partial Redistribution 437

the desired precision. Indeed, the situation at hand strikingly resembles that When the system is solved, the oJ ;s are applied to the current estimates of
of a multiline "multiplet" problem with very strong interlocking effects, J., and the rate equations (13-81) through (13-83) are re-solved for new n's.
each frequency within the line playing the role of a separate transition, and These values are used to compute x. and "" and a formal solution of the
the line as a whole acting as the collective photon pool. This analogy suggests transfer equation updates the Eddington factors / •. The whole process is
that the solution can be obtained effectively by means of a complete linear- iterated to convergence. The formalism can be extended (459) to include the
ization technique. case of several lines from different sharp lower levels to common broadened
upper levels [e.g., the Ca II H- and K-lines and infrared triplet, for which
METHODS OF SOLUTION the 4s ground state and metastable 3d levels are sharp, while 4p is broad
(see Figure 12-1)]. The convergence properties of this method are good,
A general and powerful method of solving equations (13-71) through typically yielding a factor of 5 to 10 reduction in errors in the solution per
(13-75), simultaneously with the transfer equations (13-76) through (13-80), iteration.
is to use the complete linearization technique. We introduce a discrete set The method described above is effective, but is relatively expensive compu-
of upper-state substates specified by frequencies {v1}, measured relative to tationally, and it is worthwhile to explore less costly approaches (292). One
line center. The substate populations can then be written as n2(v1) = ff2/PJ, of the basic problems encountered in treating partial redistribution is that
and the rate equations discretized as the unknown emission profile appears explicitly in the stimulated emission
term in x.. and hence in the denominator of S, = '1,IX,· However, in many
n1 (B 12 o/ W/J, 1J 1 + C12 + R 1,. +. C1,c)- o/ w1n214>1(A 21 + B 21 J1 + C21) astrophysical applications, particularly for ultraviolet resonance lines in
solar-type atmospheres, this stimulated emission correction, which is of
- n,.[n,<1> 1(T)(R,. 1 + C1,c)] = 0 (13-81) order exp( - hv/kT), is extremely small. In this event we may proceed by
iterating the ratio w. = (1/1./4>.) in the stimulated emission term ofan analyt-
ical expression for the source function obtained from manipulation of the
-n 1 {B12 [y L Bt}\JJ' + (1 - y)iJ,1 L wJ'iJ,J'JJ'] + C124>1} statistical equilibrium equations.
J' J' .
Consider the source function for a strict two-level atom, omitting the I
+ n21iJ,j(A 21 + B21 J1 + C21 + R2,. + C2,.) - n,.[n,<l>2(T)(R,. 2 + C2,.)] = ,0 continuum for simplicity. Using the Einstein relations we may write the line
I\)
I\)
(13-82) source function as
....J
and n1 + L W1ff21'P1 + n,. = "■tom (13-83)
(13-84) I
J

Here a,l} is a discrete representation of R 11 (v,, v1) in an appropriate quadra-


Here, and in what follows, we assume that a current estimate of w, is known.
ture. Writing n e (ni, ff2i, ff22 , ••• , ff21, ... , ff21 , n,.t,
where J denotes the From equation (13-71), omitting the continuum terms,
total number of substates, equations (13-81) through (13-83) are of the form
dn = !M. 3/ 2) .,. / 1
,i, ) - A21'P,- f R(v',v)J,,dv' + C 21 (1 - e-hv/kT)B,(T)
Suppose we have an estimate of the occupation numbers n for all depths; (2h\I C (g1n2'l'v g2nl'l'v - A B J C
21 + 21 • + 21
then current values of x.. '1 .. and J, may be computed at all frequencies in
the line and continuum. We may then linearize the transfer equation in (13-85)
terms of lJJ,. ox,. and 0'1,, and express the latter two quantities in terms of and from equation (13-74)
Oni, the 0n2/s, and on,.. In turn, the on's can be written as 6n = L (on/oJk) oJb
where the sum extends over all frequencies in the line and continuum. The _ A21 + B21(1, - w,Ja) + C21(l - w,e-h•fkT) ( _ )
[ 1 - (g1n2 / 92n1 )w,] - A +B J +C 13 86
derivatives (EJn/EJJt) can be written as (EJn/EJJt) = -d- 1 [(0.91/EJJt) · n], and 21 21 e 21
explicit analytical expressions can be obtained for the derivatives (o.91 /oJk)
[ see (456; 459)]. The final system of equations to be solved is of the standard
where Ja = f 4>.J. dv and J,
C21 (1 - e-h•lkT)/A 21 we have
=J 1/1.J. dv = JiJ,,w,J, dv. Defining e' =
Feautrier form -A, 6J 4 _ 1 + B4 6J4 - C4 6J4 + 1 = L4 where
S,(v) = e. ["'·-I f R(v', v)J •. clv' + e'B.] (13-87)
438 Line Formation with Partial Frequency Redistriblltion
13-4 Radiative Transfer with Partial Redistribution 439
where
obtained from complete redistribution, for R1, R11 , and R111 in constant-
e. = [(A21 + B211, + C2i)/(A 21 + B 21 J, + C2i)] property atmospheres (53; 212; 316 ). The source functions S1(v) and s,CR
1 + (B21/A21)(], - w.Ja) + (C2i/A2i)(l - w.e-h•/kT) (13-88) may also be used to compute line-profiles; the errors made in calculated
profiles if complete redistribution is assumed in place of an accurate re-
Using equation (13-87) in an expression for the total source function of the distribution function can then be assessed. An intermediate approximation
form
is the iterated source function
s. = [x,(v)S,(v) + x.s.]/[x,M + x.J (13-89)
where now
(13-94)
x,(v) = oc12[n1 - (gifg2)n2w.]cf,. + x. (13-90)
the transfer equation now reduces to the general form which is evaluated using the mean intensity obtained from the complete
redistribution solution based on equation (13-93). It has been found that
(13-91) sj1>(v) in isothermal media is nearly equal to S1(v) (316), and obviously it is
much simpler to compute.
which may be solved straightaway by the standard Feautrier technique. In Let us first consider the case of redistribution by Doppler shifts only-i.e.,
this method we use current estimates of w. and J, to calculate in equation e. R(v', v) = R1(v', v). A number of solutions for S1(v) have been obtained (316)
for both finite and semi-infinite isothermal media, assuming zero continuum
(13-88), and of the level populations and w. to determine x,(v) in equations
(13-89) and (13-90). The transfer equation (13-91) is then solved for improved opacity, withe = 10- 4 and 10- 6 • Results for variation of the source function
values of J •. These J.'s are then used in equations (13-81) through (13-83) with frequency and depth are shown in Figure 13-6 for e = 10- 4 , The
to update ni, n2(v), and n,., and from these, 1/1. and thus w •. The process is vertical arrows designate the frequency at which the monochromatic optical
then iterated to convergence. One would expect this meth_od to work well depth t_. == 1. It is clear that, at all frequencies where t_. ;:: 1, S1(x) essentially
whenever (B 21 1/A 21 ) « 1; when stimulated emissions are very important equals s,cR; large deviations occur when t_. < 1, but are oflittle consequence
one must use the full linearization technique described above. because these optically thin regions do not contribute significantly to the
intensity in a line profile. In fact, the line profiles computed from S1CR and I
from S1(x) are virtually identical; hence we conclude that for Case I the I\)
RES UL TS FROM IDEALIZED MODELS
scattering process is, for all practical purposes, adequately described by the I\)
A great deal of insight into the nature of partial redistribution effects simpler assumption of complete redistribution. A similar conclusion, is en
can be obtained from studies similar in spirit to those described in Chapter reached (212) for the case of complete redistribution in the atom's frame-Le., I
11, using idealized model atmospheres and atoms. To simplify the problem R(v', v) = R 111 (v', v). Here the iterated source function, equation (13-94), was
as much as possible, it is customary to make the additional physical assump- evaluated in an isothermal semi-infinite atmosphere, for lines with a = 10- 3,
tion that the stimulated emission profile is given by cf,., not 1/,., and that the and e = 10- 6 and 6 x 10- 3 , Again it is found that S! 0 (x) departs from
stimulated emission rate in equation (13-71) can be written as ni(v)B 21 J. s,cR only fort,, < I, and that emergent line profiles computed from S1CR are
rather than 11 2(v)B 21 J •. Then the parameter c.o,. defined above is identically almost identical to those computed from the frequency-dependent function.
unity, J, = J., and ~. = (1 + er
1
, and equation (13-87) reduces to Thus the assumption of complete redistribution provides a very useful and
accurate approximation for Case III as well.
S1(v) = (1 - e)cf,. - 1 f R(v', v)J •. dv' + eB, (13-92) The situation for Case II (coherent scattering in a broadened profile in
the atom's frame with Doppler redistribution in the laboratory frame) is
where e = e'/(1 + e'). With this·source function, the transfer equation can quite different. Results from an isothermal atmosphere of total thickness
be solved in a single step without iteration. This approach has been used to T = 10 6 , for a line with a= 10- 3 and e = 10- 4 , given in (316), are shown
estimate the differences between the frequency-dependent S1(v), obtained in Figure 13- 7. Here we see that at line center, S1(v) is near s,cR for shallow
~hen partial redistribution effects are taken into account, and the frequency- optical depths, but rises above s,cR at great depth, and rhermalizes to the
independent Planck function sooner than the complete-redistribution source function.
s,cR = (1 - e) f cf,.J. dv + eB. (13-93)
This result is obtained because the coherent nature of the scattering process
in the line-wings inhibits photon escape from the line-core, and forces more
440
t ..

t • 104

t. 103 t

• = 10•

t = 10 2

-1
log(S/B) -1

log(S/B)

t = 10

I
I\)
I\)
co
I

-2 -2

t =0

0 2 3 4
0 2 3 4
X
X
FIGURE 13-7
FIGURE 13-6 Source functions in an isothermal atmosphere of total thickness T = 106, for a
Source functions in an isothermal, semi-infinite atmosphere for a line with£ "" 10-•, line with£= 10-• and a = 10- 3, assuming coherent scattering in a radiation-
assuming pure Doppler redistribution (case I). The vertical arrows show the frequency damping profile in the atom's frame, and Doppler redistribution in the laboratory
at which the monochromatic optical depth t, • I. Das/red curves: frequency- frame. Das/reel c11rves: frequency independent s,0 obtained assuming complete
independent s,0 obtained assuming complete redistribution. Solid curves: redistribution. Solie/ curves: frequency-dependent S1(x) obtained using correct
frequency-dependent S1(x) obtained using correct redistribution function. Abscissa: redistribution function. Abscissa: frequency displacement from line-center in
displacement from line center in Doppler units. From (316), by permission. Doppler units. From (316), by permission.
442 U,w F11r111111i1111 wi1/r 1'11rtial Frc·q111•111·y R1•tli,1·1rih111i1111 13-4 Radiative Transfer with Par1ial Redislribution 443

rapid thermalization in the core by reducing the net escape probability for 4.5 km s - 1 ; (2) the HSRA with the microturbulent velocity distribution
such photons. In contrast, in the line-wings, the frequency-dependent source given by (401); and (3) the same as model 1, but with the temperature
function lies substantially below s, 0 \ because photons are no longer being structure modified to be T 3 = max {TttsRA• 4450°K), which raises the tem-
fed into the wings from the core as efficiently. These effects are even more perature minimum by about 300°K. In all cases the full depth-variations of
pronounced for finite atmospheres where s,cR may exceed S1(x) by orders of the line profile {allowing for radiation, van der Waals, and Stark broadening),
magnitude in the line wing! The emergent line profiles faithfully reflect the the redistribution function, and background sources were taken into account.
discrepancies between S,(x) and s,CR. Profiles for the correct source function The results from complete redistribution (CR) and partial redistribution
have intensities that lie well below those for complete redistribution in the (PR) computations of the double-reversal near line center {which provides
wings, and that nearly agree with the intensities predicted by coherent vital information for diagnostics of the temperature-minimum region and
scattering. Although these effects will be diminished when there is an over- the chromosphere) are strikingly different. As may be seen in Figure 13-8,
lying continuum, we may, nevertheless, expect partial redistribution effects
to be important for resonance lines formed in the outer layers of stars where
densities, and thus collision rates, are low. I
~ 3 X 10- 6
I
APPLICATION TO SOLAR AND STELLAR RESONANCE LINES
i 2 X 10- 6
The resonance lines of many- ions [ e.g., those of hydrogen (particularly ~
L(X), Ca+, and Mg+] are formed in chromospheric layers of low density, N
I
and hence are rather well characterized by the redistribution function of B 1 X 10- 6
equation (13-73) with y almost unity. Several calculations have now been
made for solar lines, using the HSRA (249) or related models, and·for solar- I
..: 0 ..........................._._.._._........................_.............................,
type giants (where densities are even lower than in the sun) using realistic 0.0 0.2 0,4 0,6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
model atmospheres. I
(a) ~). I\)
One of the first examples of the importance of partial redistribution
effects in resonance-line formation arose from attempts to fit the observed w
solar chromospheric La profile (645). It was found that when the line profile 0
I
was computed under the assumption of complete redistribution, using models ~ 3 X 10- 6 I
that provided accurate fits to the continuum data formed in the same I
atmospheric layers as the La line-wing, the intensity in the calculated profile
wing was much larger {by a factor of 5 to 6) than observed. From the first
i 2 X 10- 6
study (645) it emerged that a much better fit to the profile is obtained if the
scattering is assumed to be about 93 percent coherent, and only 7 percent
completely redistributed. Subsequent work showed (456) that this param-
eterization is equivalent to using equation (13-73) with realistic values of
y (determined from the known collisional and radiative rates), and finally, 0 ........_._......_._...._._._............._._..._.............................,
that when the full depth-dependence of the profile and redistribution func- 0.0 0,2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
tions, atomic rates, level populations, and background opacity are taken (bl
into consideration, an excellent fit to the observations is achieved (457). FIGURE 13-8
An even more interesting example is provided by the solar Ca II H- and Ca II K-line profiles computed using the HSRA (249) model solar
K-lines, for which the earlier work {e.g., 401) assuming complete redistribu- atmosphere with the distribution of microturbulent velocities
given in (401 ). Orilinare: specific intensity I(µ,~).) in absolute
tion gives a good fit to the disk-center profiles, but fails to fit the observed units; 11bscissc,: displacement ~.l. from line-center in A. (a)
center-to-limb variation. Calculations using a five-level atom similar to that Partial redistribution results; (b) complete redistribution results.
shown in Figure 12-1 were made (570) [see also (642)] for three model Curves are labeled with µ, the cosine of the angle from disk
atmospheres: (1) the HSRA with a depth-independent microturbulence of center. From (570). by permission.
444 Line Formation with Partial Frequency Recb:1·trib11tion 445

the PR profiles show uniform limb-darkening throughout the entire profile. 1.2 r--r---r----r---r---r---r---r--...---.---,
In contrast, the CR results show limb-brightening at K 2 (the emission 0
maximum) and no center-to-limb variation at K 1 (the minimum outside K 2 );
both of these results are contrary to observation. Furthermore, the wave- 1.0
length position of the K I minimum shows a rapid center-to-limb increase
for CR (again contrary to observation), while for PR the increase is much 0

slower. The problem of limb-brightening of K 2 when CR is assumed can be 0.8


overcome if a very special distribution of microturbulent velocities (27) is ~
,-:
employed. However, neither of the problems just described for K I is elimi- <I 0

__ ____
nated in this way; moreover, the need for this special assumption is obviated 0.6
when PR is employed. 0
The quantitative improvement in the comparison with the observed
behavior of the K 1 feature is shown in Figure 13-9. There we see that CR 0.4
y
0
, , . --, ~

produces much too rapid a rise in L\),(K 1) asµ -+ 0, while both models 1 and
3 yield an excellent fit to the data. Model 2 does not do as well, though it • • • • •
provides a better fit to certain data for K 2 (not discussed here). In part (b) of 0.2
the figure we see that CR fails badly to fit the limb-darkening of K I intensity, 1.0 0.8 0.6 0.4 0.2 o.o
while PR (with any of the three models) yields the correct center-to-limb (a) µ
variation (notice the logarithmic scale). Note that the absolute intensity of the
K I feature is reduced by PR, relative to CR, for a given model (compare the 1
2 IQ" 6 r----r----r-,----r---r--.----r--r----.---.
open and filled circles at µ = 1). This results from the essentially coherent
~ X

nature of the scattering process in the K I feature, and is consistent with the ..
I

~
results for R11 redistribution in the idealized models discussed above. One I
~ay to obtain a fit to the absolute intensity is to raise the temperature I\)
minimum by 300°K to 400°K over the HSRA value, as was done for model 3. w
While this change seemingly is small, it should be recalled that a change of ......
about 100°K near Tmin changes the energy content of that' region by ari 6 X 10" 7 .___.___._...__.___,'--..L.......1.-'---•L....J
amount comparable to the energy content of the entire corona (cf. §7-7.). One ..: 1.0 0.8 0.6 0.4 0.2 0.0
hesitates to conclude on the basis of one line that T min must be higher than (b) µ
the HSRA model value; however, similar results are also obtained from an FIGURE 13-9
analysis of the Mg II h- and k-lines (59), and it may, in fact, be necessary to (a) Wavelength position of minimum intensity outside emission
raise the empirical value of T min to about 4400°K. Such an adjustment would core, K 1, as a function ofµ, cosine of angle from disk center.
(b) Limb-darkening at K 1 ; note logarithmic units of intensity,
bring the semiempirical result into harmony with the estimates of T min from
Open circles: complete redistribution results, model 2 (see text).
radiative equilibrium models, and would eliminate the difficulty described Filled circles: partial redistribution, model 2. Open tri1111g/es:
in §7-7. In any case it must be emphasized that the drop in intensity of PR partial redistribution, model I (see text). Crosses: partial
relative to CR is a differential effect for a given model, and therefore will occur redistribution, model 3 (see text). From (S70), by permission.
no matter what model is used. Thus if T min is estimated from a fit to the K 1
intensity (as is sometimes done iri stellar work), one must use a PR description frequency-independent, and the frequency-dependent PR source functions
of the scattering process, especially for giants (571), or systematic errors will S1(v) shown in Figure 13-10. There we see that CR yields a unique source
be made. Finally, it is found that the PR calculation accurately reproduces function that has a single absolute minimum. As one observes from center to
the relative behavior of the H- and K-lines while CR does not. limb, the slant-length optical depth at a particular frequency increases; if
The marked differences in the CR and PR predictions of limb-darkening s,cR is used, this implies that the intensity at L\),K, (µ = 1) must rise as µ
and wavelength-position of K I can be easily understood in terms of the decreases. The minimum stR will manifest itself only at some larger £\l
differences in the depth-variation of the source function s,cR, which is where the material is more transparent and thus reflects the source function
446

14
I

.j! Radiative Transfer in


I

I9 10-•
L---;;.i::...-=:.!---..:~- Moving Atmospheres
c:,{

10- 7 ""'"-::--"--'---'---'---..i"---....J.-J.._J
10- 7 10-• 10-s 10- 4 10- 3 10-z 10- 1 I 10
m(gm cm- 2)
FIGURE 13-10
Depth-variation of K-line source function (for a five-level atom)
and Planck function in model I. Dashed curve: Planck
function B,. Dotted curve: complete-redistribution source-
function S1 (see also Figure 12-2). The other curves show'the
partial-redistribution source-functions at line-center (K 3 ), the
_emission peak (Kz ), and the profile minimum outside the peak
(K 1). Abscissa: column density gm/cm 2 • From (570), by
permission.

deeper in, near the minimum. This explains both the rapid increase in The existence of macroscopic motions (i.e., nonthermal velocities that are
6J.(K 1) as a function ofµ, and_ the ne~r-constancy of /K 1 predicted by CR. In coherent over distances much larger than a particle mean-free-path) in
contrast, the PR source function at hoe-center (K 3 ) lies above the CR value stellar atmospheres is well-documented by a wealth of observational evi-
(as a result of photon trapping as discussed previously). But at K the source dence. These motions appear to be present on all scales, from "eddies" whose
function has an essentially coherent-scattering character, is dec~upled from sizes are small compared to a photon mean-free-path, up to expansion of the
other frequencies, and shows a monotonic decrease outward in the atmo- atmosphere as a whole. Although velocity fields have but little effect on
spher~ (similar _results were ?btained in (641)]. For this behavior of Si, radiative transfer in the continuum, they strongly influence line-formation
changmg the point of observation from center to limb merely samples S (K ) because even a small (Doppler) frequency shift of a line produces a major
higher in the atmosphere, where it has (for PR) a lower value, and hen~ change in its absorptivity as seen by a stationary observer.
pro~uces a_de~rea~e in, IK,• as desired: at about the same.:\).. In short, taking In the analysis of spectra ofsupergiants, Struve and Elvey (617) discovered
partial red1stribut1on mto account yields a substantial improvement in the that the Doppler widths inferred from the position of the fiat part of the curve of
degree of agreement between theory and observation for the solar Ca II growth (cf. §§10-3 and 10-4) were far in excess of the thermal value. They attri-
H- and K-lines. Similar calculations, yielding similar results, have also been buted this broadening to non thermal "turbulent" velocities, presumed to have
made for the Mg II hand k (3s-3p) lines for the solar atmosphere (458; 59) a Gaussian distribution. The geometric scale of these motions (which are
and for solar-type stars (4SS). It now appears that it is mandatory to account called "microturbulence") is supposed to be so small that they act as additional ·
for partial ~distri~ution effects i!1 t~e interpr~tation of strong chromospheri• line-broadening agents, and thus enhance the line-strength. The inferred ve-
resonance Imes with strong rad1at1on-dampmg wings and further efforts it locities often approach or exceed the speed of sound in the material, and it is
this direction will undoubtedly be amply rewarded. ' clear that astrophysical "microturbulent" velocities are not to be identified
448 Radiative Transfer in Moving Atmospheres
.14-1 The Transfer Equation in the Observer's Frame 449

with turbulence in the strict fluid-dynamical sense, but rather with unresolved rapid flow with large velocity gradients. Comoving-frame methods span the
motions. two extremes and provide general solutions applicable in both limits. Special
The picture is clarified when we examine the solar spectrum. From curves of techniques have been developed to treat random or stochastic velocity fields.
growth, or from spectra of low spatial, spectral, and temporal resolution, one
again infers a significant microturbulent velocity. But high-resolution spectra
have a characteristic "wiggly-line" appearance [see Figure III-2 of (20) for 14-1 The Transfer Equation
an excellent example], showing distinct Doppler shifts and asymmetries that in the Observer's Frame
fluctuate rapidly along the length of the slit, and in time. Probably a large
part, or even all, of the velocity field is composed of wave motions of various FORMULATION AND SOLUTION OF THE TRANSFER EQUATION
types, scales, and periods, which (when superposed) give rise to a pattern
When the material in the atmosphere moves with velocity v(r) relative
that appears chaotic and "turbulent". However, the diagnostics are still in a
to an external observer at rest, there is a Doppler sliift of photon frequencies
primitive state, and the precise nature of the velocity field is not at all well
between the observer's frame, and the frame of the atoms of which the material
known.
is composed. If the frequency in the observer's frame is v, then in the atom's
Evidence for velocity patterns on a large scale ("macroturbulence") was
frame the frequency at which a photon traveling in direction n was emitted,
provided by Struve's observation (616) that the widths of line profiles in
or can be absorbed, is
certain stars exceeded the Doppler widths obtained from the curve of growth
v' = v - v0 (n · v/c) (14-1)
of their spectra; h.ere the line strength is unchanged, and one envisions areas
on the stellar surface so large as to be practically independent "atmospheres" Thus, the opacity and emissivity of the material, as seen by a stationary
moving systematically along the line of sight. A discussion of the observations, observer, become angle-dependent. The transfer equations for a time-
with numerous ref~rences, can be found in (261, Chap. 8). Further, periodic independent moving medium in planar geometry is then
Doppler shifts of the lines in some stellar spectra reveal that they are from
pulsating stars. Beyond this, objects such as the WR stars, P-Cygni stars, and tt[ol(z, tt, v)joz] = 17(z, tt, v) - x(z, tt, v)/(z, µ, v) (14-2)
early-type supergiants all show characteristic iine profiles, with blue-shifted
absorption components and red-shifted emission components, indicative of It is convenient to measure frequency displacements from line center in
large-scale expansion. units of a fiducial Doppler width 6.vt = v0 v~/c, where v~ is a thermal
It is clear that a characterization of stellar velocity fields by the two velocity parameter, and to measure velocities in the same units, V = vfvrh·
extremes of "micro"- and "macro"-turbulence is an oversimplification and, Then the transformation between observer's frame and atom's frame fre-
in the end, we wish to know distribution functions describing the amplitudes quencies is
and scales of the velocity patterns. Further, one would like to relate observed x' = x - µV (14-3)
parameters to more fundamental quantities such as the velocities of convec-
where x = (v - v0 )/6.vt and x' is defined similarly. The effects of Doppler
tive motions. Finally, a truly consistent theory of stellar atmospheres will
shifts are inconsequential for continuum terms, which do not vary much over
require a dynamical theory of the interaction of material velocities, the the frequency range implied by velocity shifts, so we account only for changes
thermodynamic state of the matter, and the radiation field; only then will we
in line terms and write
be able fully to understand stellar chromospheres and coronae.
At the present time, however, we are far from having such a complete x(z, µ, x) = Xr(z) + x,(z)cf,(z, µ, x) (14-4)
theoretical structure. In this chapter we shall focus almost entirely on the
"kinematics" of radiative transfer in moving media; i.e., given the velocity and 17(z, µ, x) = 17c(z) + 11,(z)cf>(z, µ, x) (14-5)
field and the model atmosphere, compute the emergent spectrum. A variety
of techniques exist to attack the problemjust posed. Observer's-frame methods where the normalized line-profile is defined by
can handle complicated velocity fields and multidimensional structures
(although we shall confine attention to one-dimensional problems only), but
cf>(z, µ, x) = <f,(z; x - ttV) (14-6)
are generally restricted to velocities of the order ofa few Doppler widths, and For example, for a Doppler profile,
hence are not well suited to handle rapid atmospheric expansion. Sobolev's
method, on the other hand, provides an approximate solution in the case of cf>(z,tt,x) = 1r-ti5- 1(z)exp{-[x - 11V(z)]2/<5 2(z)} (14-7)
450 Radiative Transfer in Moving Atmospheres 14-1 The Transfer Equation in the Observer's Frame 451

where omeasures the local Doppler width in units of the fiducial value; i.e., where e is the usual thermalization parameter. Note that in the scattering
o(z) = Av0 (z)/Avt. We now define line and continuum source functions integral we can no longer replace / with J because <p is angle-dependent; note
S,(z) = ,r1(z)/x,(z), and S,(z) = ,r,(z)/x.(z); a total source function also that the intensity can no longer be assumed to be symmetric around the
line center, hence the full profile must be considered. The approximation of
S(z, µ, x) = [cf,(z, µ, x)S1(z) + r(z)S,(z)]/(1/>(z, µ, x) + r(z)] (14-8) complete redistribution becomes questionable for moving media, as the
conditions that help validate it in static media no longer occur; a good
where r(z) = x.(z)/x,(z); and an optical depth scale measure along a ray discussion of this point is contained in (273, 87) (this is a superb paper that is
specified by µ, highly recommended to the reader). Recently some work has been devoted
t{z, µ, x) = µ- 1 f:'""• x(z, µ, x) dz (14-9) to ·the problem of partial redistribution in moving atmospheres; it has been
shown that to treat the problem in the observer's frame, the full angle-
where Zmu denotes the upper surface of the atmosphere. Then the transfer frequency dependent redistribution function must be employed. In contrast,
equation becomes in a comoving-frame method (see §14-3), one can employ static redistribution
functions in the fluid frame, and angle-averaging again yields accurate
[ol(z, µ, x)/ot(z, µ, x)] = /(z, µ, x) - S(z, µ, x) (14-10) results.
Accurate calculation of the scattering integral in equation (14-12) with a
The formal solution of equation (14-10) can be written immediately as quadrature sum poses a fundamental difficulty in an observer-frame solution
for two reasons. (1) The line-profile rj,(x - µ V) is clearly shifted by an
l(zm.., µ, x) = 1(0, µ, x)e-r(O,µ,x) + f;(O,µ,x) S(z, µ, x)e-r(z,µ,x)d,(z, µ, x) amount 2V in frequency asµ varies from -1 to 1. Thus, in the frequency
quadrature, an amount equal to twice the maximum macroscopic flow
= /(0, µ, X)e-r(0,µ,x) velocity must be added to the bandwidth required to describe the static
+ f:'"" µ- 1 (1/>(z, µ, x}S,(z) + r(z)Sc(z)]e-r<•, 11 ·"1x1(:z) dz
line-profile. This requirement is not severe in studies of, say, wave motions in
the solar atmosphere, but becomes prohibitive for atmospheres in supersonic
(14-11) expansion where v/c ~ 0.01, or 2(v0 v/c)/Avt ~ 200. (2) The angle-quadra-
ture scheme must employ a large number of angles. Because the argument of
from which we can compute the emergent intensity for given source functions. the profile function is (x - µ V), there is an inextricable coupling between the
(For example in LTE, S, = S, = B; or we might use the values of S1 found for angular and frequency variations of the intensity. Thus if some maximum
the line in a static atmosphere, an approximation that often is surprisingly frequency increment Axm 1 .( ~t) is required to obtain sufficient precision in
accurate as we shall see below.) In equation (14-11) we have taken the atmo- the frequency quadrature, the maximum tolerable angle increment will be
sphere to be a finite slab with intensity incident at z = 0: for a semi-infinite !:,.µ,,,,x ~ llxm,JV. which is quite stringent! These difficulties are ameliorated
mmosphere, we set ttO, µ, x) = ex, and omit the term in /(0, µ, x). The formal by transforming to the comoving frame.
solution allows a direct evaluation of the effects of velocity fields on profiles Equation (14-10) may be cast into second-order form. If the line profile is
by accounting for the velocity-induced shifts in the opacity and emissivity symmetric about line center, then I/>( - x + µ V) = cp(x - µ V), which sug-
of the material. gests that we group the two pencils /(z, µ, x) and /(z, - µ, -x) together, for
The line source function will, in general, contain a scattering term, and d,(z, µ, x) = d,(z, - µ, -x) and S(z, µ, x) = S(z, - µ, - x). Thus, defining
therefore depend on the radiation field; thus the source function can be
strongly affected by material motions. For example, an expansion at the u(z, µ, x) = 21 (/(z, µ, x) + /(z, - µ, -x)] (14-13)
upper surface ofan atmosphere can displace a line away from a dark absorp-
tion feature, at the rest position, into the bright nearby continuum, thus
raising] (and S1) dramatically. If we assume that photons scattered by the and v(z, µ, x) = 21 (/(z, µ, x) - l(z, -µ, -x)] (14-14)
line are completely redistributed, then the source function for a two-level
atom becomes we obtain [o 2 u(z, µ, x)/iJr(z, µ, x) 2] = u(z, µ, x) - S(z, µ, x) (14-15)
At the upper boundary there is no incoming radiation, hence
SJz) = ½(l - e) f:'.., f~
dx 1
dµ l(z, µ, x)cf,(z, µ, x) + eB(z) (14-12)
[ iJu(z, µ, x)/ot(z, µ, x)] ..... = u(zmm µ, x) (14-16)
452 Radiative Transfer in Moving Atmospheres 14-1 The Transfer Equation in the Observer's Frame 453

At the lower boundary we assume either !hut the incident radiation is LINE-FORMATION WITH SYSTEMATIC MACROSCOPIC
specified, in which case VELOCITIES IN PLANAR ATMOSPHERES

The effects of velocity fields on line-formation in planar atmospheres


[ou(z, µ, x)/ot(z, µ, x)] •• 0 = 1(0, µ, x) - u(0, µ, x) (14-17)
have been studied by a number of authors; we shall review some typical
results here. Basic insight into effects of motions can be gained by using just
or that, in a semi-infinite atmosphere, the lower boundary is chosen to be so the formal solution. Consider, for example, a semi-infinite atmosphere with
deep that the diffusion approximation is valid [ which demands that the a line formed in LTE, with r = constant, and S1 = S, = B, = B0 (1 + at1),
velocity gradient be small enough that 1
x-
(dV/dz) « 1; i.e., there is a neg- wpere t 1 denotes the static line optical depth. Choose a velocity field of the
ligible change in the velocity over a photon mean-free-path], in which case form v(t1) = v0 /[ I + (t1/t 0 )], taken to be positive toward the observer (i.e.,
toward increasing z). Then it is easy to calculate the emergent intensity
ou(z, µ, x)I
ot(z, µ, x) r•O =
[x(z,µµ, x) (oB,)
af dz
ldTI] r•C
(14-18) /(µ, x) over the surface of the star, and to construct the flux F(x) by integrating
overµ; results for models with a = 3 x 10- 2 , r = 10- 2 , v0 = (0, 1, 3,
5, .10), and , 0 = (1, 10, 100) are displayed in Figure 14-1. There we see that
As in the static case, we introduce a discrete depth-mesh {z,}, angle-mesh the line flux profile shows an asymmetry toward the blue. Similar asymme-
{µm}, and frequency-mesh {xn}, and combine angles and frequencies into a tries result even if there is no velocity gradient and the atmosphere is assumed
single quadrature set {µ,. x 1} =
(µm, x.) where / = m + (n - 1) M; the to expand with constant velocity, because of the way velocities and intensities
angle-points are distributed on the interval [O, 1], while frequencies must now are weighted in the flux integral (Cf. Exercises 14-1 and 14-2).
span a range [xmin, xm.. ], Xmin < 0 and xm .. > 0, large enough to contain
hoth halves of the line profile and to allow for Doppler shifts ± 2 Vm••· We Exercise 14-1: (a) For a linear limb-darkening law tf,(µ) = /(µ)//(0) = I + pµ,
then replace equations (14-15) through (14-18) with difference equations and show that the function normalized to give unit flux is tf,* (µ) = ½(I + pµ)/
write (½ + ½P). (b) Assume that a weak line is formed on the surface of an atmosphere
(14-19) expanding with velocity v0 , and that the line depth, as a fraction of the con-
tinuum, does not vary with µ. Derive an expression for the radial velocity mea-
where ix and fJ are the appropriate combinations of r,, ¢,,, and e,, and sured from observations of the flux in a spectrogram. In particular, show that
for the grey-body limb-darkening law, in the Eddington approximation, v0 bt =
L
(17/24)v 0 • What is the ratio of v0 bJv0 for p = O; for p = oo?
J, =L w1tf>41 u41 (14-20) Exercise 14-2: Calculate the flux profile from a line, idealized as a delta-function
of constant depth, on a stellar surface expanding with constant velocity v0 ; i.e., take
I• 1
/(µ, x) = ½(I + pµ)[I - a0 o(x - µV0 )]/(½ + ½P), Derive an explicit expression
where tf,41= tf>(z4 ; x 1 - µ 1V.,). The resulting system is then of the standard for F(x; a0 , P, V0), and plot the profile in the limiting cases p = 0, p = oo.
Rybicki form [see equation (6-47)] and is solved for J as described in §6-3. (b) Expand the analysis to a line with a Gaussian profile, and compute numeri-
An analogous integral-equation solution can also be constructed (273, 120), cally a typical flux profile.
but in application the differential-equation method is easier to use. The whole
procedure is stable and general, and quite efficient, as the computing time The formal solution can be used to evaluate proposed velocity-diagnostic
TR= cl D2 + c' D is only linear in L, the number of angles and frequencies.
3 techniques by computing profiles for given velocity fields, subjecting these
The depth-mesh must be chosen sufficiently fine to assure that only modest profiles to the diagnostic analysis, and comparing the inferred with the
changes in V(z,), say ,:5 ½, occur between successive depth-points; otherwise originally-assumed fields. For example, the "bisector shift" technique has
the profile function tf,41 may change radically with depth and lead to inac- been examined (373) for a variety of cases. This method supposes that the
curacies in the optical depth increments. Except for supersonic winds, this is displacement c5x, from the static line-center, of the position of the point
not a stringent requirement. Note also that the same methods can be used to midway between two points of equal intensity in the line-profile, gives the
construct the formal solution, when S is given, by solving a single tridiagonal Doppler shift caused by velocities in a layer at unit optical depth for a (static)
system (at each angle-frequency point desired) of the form T 1u1 = S1; here line frequency ~x. where 2~x is the full distance between the two points on
the computing time required is only Ts = cL D, which is minimal. the profile. It is found that the inferred velocities are in fair agreement with
454
14-1 The Transfer Equation in the Observer's Frame 455
1.0
the input velocities for measurements in the line-core, but that spurious
velocities at great depths are inferred from the wings. It is easy to see why this
0.8
is so. Suppose that the atmosphere moves with velocity v0 on the range
0 ~ tr ~ ti, and is at rest for tr > t 1 • It is clear that the shift of the line in
F(x)/F,
0.6 the upper layer forces the line-wings to be asymmetric, because opacity from
the upper layer intrudes into the wing and absorbs radiation from below.
0.4 By assuming that radiation at frequency displacement 6.x from line-center
arises from depths .tr ~ 1/¢(6.x), velocities will automatically be ascribed to
these depths even ift1 > t 1 • This example shows that care must be taken in
-4 -2 0 2 4 6 8 10 12 inferring velocity fields!
(a) X A further example of such problems is shown in (321) where a calculation
is made [using the Riccati method (544)] of the non-LTE source function of
1.0 a line with e = 10- 3 and r = 0, in a differentially expanding finite slab of
total (static) optical depth tmax = 50. The expansion is taken to be symmetric
0.8 about the midpoint of the slab (assumed to be at rest) with a linear velocity
law of the form V(t) = V0 + Vitr; the medium is effectively thin, and simu-
lates an expanding nebula. Line profiles are shown in Figure 14-2. There we
F(x)/F, 0.6
see that the usual central reversal arising in the material nearest the observer
is blue-shifted, and thus obliterates the blue emission peak, while the red
0.4 peak is enhanced because photons more easily emerge from below; the line
as a whole appears red-shifted even though the average velocity of the
material is zero! One cannot, therefore, immediately conclude that a small
-4 -2 0 2 4 6 8 10 12
observed red shift implies a receding emitter.
(b) X
To evaluate the effect of velocity fields on the source function one must
solve the transfer equation self-consistently. Some of the early work on this
1.0

0.8 0.03

F(x)/F, _
06 0.02
I,(µ= I)

0.4 0.01

-4 -2 0 2 4 8 10 12
(c) X -3 -2 -1 0 2 3
FIGURE 14-1 X
Flux profiles from an expanding atmosphere, for a line formed in LTE, with FIGURE 14-2
B,(t1) "' B0 (1 + at1) and v(t1) • v0 /(l + (t1/t 0 )). Ordinate: flux in units of Normally-emergent intensity from differentially
continuum; Abscissa: frequency displacement from line center in Doppler units. expanding slab with total (static) thickness t ..,. "' SO, for
Parameters arc B 0 .. I, a• 3 >< 10- 2 , r .. 10- 2 , for all curves. Each curve is a line with£ • 10- 3 and r - O. Abscissa: x a l:,.v/l:,.v0 •
labeled with v0 • Panels (a), (b), and (c)·havc t 0 • I, 10, 100, respectively. Curves arc labeled with v0 , velocity of expansion at
surface: velocity law is linear in t and gives zero velocity
at slab center. From (23,215).
456 Radiative Transfer in Moving Atmospheres 0.15

B
problem (370; 371) treated the case of an isothermal atmosphere with a 0.12
velocity jump A at a depth t" with constant velocities above and below. The
transfer equation was solved in the Eddington approximation, for a two-
level atom, using a discrete-ordinate method. When A = 0, the static solution 0.09
for the appropriate values of s and r is obtained. As A becomes greater than S,B
about 4, the lines (which have Doppler profiles) in the upper and lower regions 0.06
are strongly shifted with respect to one another and no longer interact. The
atmosphere then acts as if it consists of two independent parts: (a) a finite
O.Q3
upper layer of optical thickness t1, and (b) an underlying semi-infinite atmo-
sphere in which t = 0 at the depth, ti, of the velocity jump. In this limit, the
source functions in the two layers both achieve their respective static limits. 0.00I.....--L..-.....1---'--......L--.J..---'-- _ ......._ ~
-2 0 2 4 6
Thus for 6 = 0 and 6 -+ oo, one recovers profiles identical to those computed
from static source functions, the major effect just being the Doppler shift of (a) log t
the line-center in the formal solution. This result is strengthened when there
is an appreciable background continuum [see also (372)]. For A on the range· 1.50
of 2 to 3, the two layers interact strongly and a full solution must be found. 103
A more realistic problem (273, 120) is presented by an atmosphere with a 1.25
"chromospheric" Planck-function rise at the surface (see Figure 14-3a) and
with velocity laws of the form V(t1) = 10/[1 + (tr/t 0 )], where tr is the static
optical depth in the line. The resulting source functions for a line with l(µ = l,x)
r = 10- 4 ands= 10-·2. and various values oft 0 are shown ln Figure 14-3a,
0.75
while emergent intensity profiles are shown in Figure 14-3b, and flux profiles
in Figure l.4-3c. The striking result seen in Figure 14-3a is that the line source
function is only weakly affected by the velocity field, even though the pro.files' 0.50
show drastic changes. The basic reason for this result is that photon-escape'
o.2:4L-_1--_.J..o_...J._.....14_ _1..-_ _._s_ __,__--:1:-2-......._-:16
through the outer layers is increased in the red wing, but decreased in the
violet wing-and, to a large measure, these effects nearly cancel [ see also X
(b)
(18, 53)]. On the whole, the photon-escape probability is slightly enhanced
by the atmospheric expansion, which explains why Sr tends to lie below its
static value for 1 ~ tr ~ 10 2 • In the case to = 10, the value of Sr increases 1.25 10
for r 1 :S 10 because the line intercepts underlying continuum radiation while 104
it is optically fairly thin, which leads to an increase in J; for larger values of
t 0 , the line becomes optically thick above the velocity rise, and the effect
vanishes. When t 0 ~ 103, the point of the velocity rise already lies below the
thermalization depth of the line; line-formation in the upper (effectively F(x) 0.7S
thick) layer then proceeds as if the atmosphere were static, and the static
value of Sr is recovered very closely. The flux profiles for t 0 = 10 2 to 10 3 0.50
show "P-Cygni" features with red emission components and violet-shifted
absorption. Here the emission, however, arises from the assumed tempera- o.2~4L---'---_.:ro:..__.,____4..__ __.__ _s.____-"-_ _.12-~
ture rise, and not from the geometrical effects that occur in extended atmo-
spheres. To a high degree of approximation, one would find the same (c) X
profiles from a velocity-dependent formal solution using static source FIGURE 14-3 . . -2
functions. (a) Planck runction and source function in an expanding atmosphere, for a h~e wit~ e = 10
and,. = 10·•. Abscissa: static line optical depth. (b) Normally-emergent intensity.
Abscis.,e1: x = /l1•/ll1•,,. (c) Flux profiles. Curves are labeled with value of to in velocity law
,./,\ - I n,r I + (T,/T., n. From (273. 120).
458 Radiative Transfer in Moving Atmospheres 14-1 The Transfer Equation in the Observer's Frame 459

The results described above apply to expansion, where an increased s,


and increase the local coupling of to B,; this, in turn, leads to substantial
escape probability in one wing can be compensated by a decrease in the increases first in the violet, and then the red, emission peaks of the doubly-
other. In.fluctuating velocity fields, however, the coherence of shifts from one reversed profile. These results again point to the need for a dynamical theory
point in the atmosphere to another is lost, and effects similar to a marked for the velocity fields.
depth-variation of the line profile are produced. The effects of fluctuating
mesoscale velocity fields on line-formation have been studied (568) for SPHERICAL ATMOSPHERES: LOW-VELOCITY REGIME
sinusoidalwaveswith V(t,t) = ,Bsin[2n(r 1 log 10 t + t)],andfor"sawtooth"
waves; the latter simulate steepened shock-like structures. Using the Rybicki- Observer's frame calculations in the low-velocity regime have been
type method described earlier, the transfer equation was solved at times carried out for radially expanding spherical atmospheres (375). Such an
spaced equally over a period, and time-averaged profiles were found for approach can be useful in studying line formation in the deeper layers of
various values of /J and ,t The limit l ➔ 0 corresponds to a "microturbulent" expanding atmospheres, but for the large-velocity regime a comoving-frame
regime, while l ➔ co yields a "macroturbulent" limit. For a given source formulation is preferable. The method is similar to that described in §7-6 for
function (e.g., S1 = B,), the line profiles for finite l invariably lie between static atmospheres, and carries out a ray-by-ray solution in the same (p, z)
those given by the two extremes l = 0 and l = co. The lines computed with coordinate system. The transfer equation along the ray is
a given {J and l = 0 are always stronger than those for l = co, as expected.
When non-LTE cases are considered, the source functions are modified by ±[aJ±(z, p, x)/oz] = YJ(Z, p, x) - x(z, p, x)P(z, p, x) (14-21)
the velocity field; characteristically, S1 shows ripples as a function oft, and
where x(z, p, x) = xc(r) + x,(r)<f,(z, p, x), and a similar expression defines
the departure of the results for finite l from the microscopic limit (l = 0)
rJ(z, p, x); we use the relations r(z, p) = (z 2 + p 2 )½ and µ(z, p) = z/(z 2 + p 2 )½.
are larger at smaller values of e. The primary result found for non-LTE
The profile is defined as <f,(z, p, x) = tf>[r(z, p); x - µ(z, p)V(r)]. The velocity
isothermal atmospheres is a significant rise in the core intensity of the time- V(r) is positive in the direction of increasing r. Introducing the optical depth
averaged profile for finite l; the profile lies between the limiting (l = 0,
along the ray
). = co) profiles only in the wings, and is much brighter than both in the
core (by a factor of 2.5 for fJ = 2.5 and l = 4). The same behavior of the ,(z, p, x) = f.'m"
x(z', p, x) dz' (14-22)
line core is also found in computations using the HSRA, and substantial
changes in the source function occur. In particular, even though a collision- and defining u(z, p, x) = ~ [J+(z, p, x) + r(z, p, -x)] (14-23)
dominated source function is used, S1 ,::ises above B, (as it can for photo-
ionization-dominated Jines), because velocity shifts allow the line to intercept
bright continuum radiation. The brightening of the line core appears to and v(z, p, x) = ~ [J+(z, p, x) - r(z, p, -x)] (14-24)
improve greatly the agreement between the observed and computed solar
Na I D-lines at disk center, without recourse to unusually high densities as
we may rewrite equation (14.21) in second-order form:
hitherto required; this result may also offer an explanation for similar dis-
crepancies observed in the solar Ca I and Fe I lines. Further work in this [a 2 u(z, p, x)/a,(z, p, x) 2] = u(z, p, x) - S(z, p, x) (14-25)
area will be quite rewarding.
The results discussed above take no account o_f possible effects of the where S(z, p, x) =
YJ(Z, p, x)/x(z, p, x) has the general form S(z, p, x) =
velocity field on the state of the gas. Recently a study has been made (291) of a(z, p, x)J[r(z, p)] + {J(z, p, x). Here ix and fJ contain combinations of the
effects of acoustic pulses on formation of the solar Ca II lines, allowing for line parameters e and r 0 = x,/x,. and the profile function <f,(z, p, x), while
the temperature and density changes the pulses produce in the gas. These
computations show that, if changes in the physical variables are ignored,
the velocity field alone produces little change in the source function, and the
J(r) = f::. f dx 0
1
dµ <f,[r; x - µV(r)]u[z(r, µ), p(r, µ), x] (14-26)

correct profile is predicted using the static source function in a velocity- In formulating the boundary conditions a difficulty arises. On the axis z = 0
dependent formal solution; a similar conclusion was reached in (185). In we can no longer write v(O, p, x) = 0 beca.use two frequencies ( ± x) of radia-
contrast, the changes in Tand N produced by the pulses have major effects on tion are involved. We may circumvent this problem by following the ray for
s, and hence on the profiles. In particular, both T and ne increase together, its entire length-Le., we consider the whole interval [ - zmm ZmaxJ. The lower
460 Radiative Tmn.vfer i11 Moving Atmo.rplrere.v 14-1 Th,, Tran.i:/l'r Equation i11 ti,,• Ob.rer1•a'.,· Fram,• 461

and upper boundary conditions for rays that do not intersect the core then computing time is prohibitive; in this case one may use a comoving-frame
become method. An advantage of the observer-frame method is that it can be used
[ aucz, P, x>Jot(z, P, x>J ... ±•m.. = ± u(i, P, x>I ... ±•m.. (14-27) for arbitrary variations in the velocity field (e.g., nonmonotone flows), which
is not true for comoving-frame methods, as currently formulated.
For rays that intersect the core, p· :i;; r 0 , we either (a) apply a diffusion
approximation at an opaque core (stellar surface), which yields v(Zm1,., P, x) Exercise 14-3: (a) Verify the symmetry relations quoted above for rt., u, and v,
directly, or (b) for a ltollow core (nebular case), apply eq~ation (14-25) at and the reduction of (14-29) to (14-30). (b) Sketch the form of the rectangular
: (forcing the points at ±zmin to be identical), and equation (14-27) at the chevron matrices U1• and V1" in the Rybicki scheme, and show that the dimensions
111111 of the matrices mesh in the correct way to allow a solution.
ends of the ray.
To solve the system we introduce the same discrete meshes {r.,} and {Pi}
used in §7-6 to solve the static problem. The frequency mesh now includes The method described above has been applied (375) to highly idealized
thewholeprofile{xn}, 11 = ±1, ... , ±N,withx_" = -x";weshall,how- spherical atmospheres with power-law opacities and linear velocity laws of
ever, be able to eliminate half of these (see below). We again obtain equations the form V(r) = V(R)(r - r 0 )/(R - r 0 ). Calculations were made in extended
of the form of(6-27) and (6-48), and hence can apply the Rybicki method to isothermal models with R/r. = 30, for a line with e = 10- 2 and 10- 4 and
obtain J. Because J(r.,) need be defined only for {r.,}, 1 :i;; d :!E; D, while zero background continuum, with V(R) = 0, 1, and 2. The results include
udln = u(z.,, p,. Xn) is defined on a mesh {z.,,}, d1 = 1,.:., D,. w~i h runs the following. (a) The source function is more strongly affected by ex-
7 tension than by small"velocity fields; to a first approximation, one may use
the whole length of the ray, it now turns out that, while the tridiagonal
T-matrix is square, the U-matrix .is rectan(Jular, and is a chevron matrix. the static spherical source-function in a velocity-dependent profile com-
Solution of these systems for each choice of (i, n) yields an expression of the putation. (b) The effects of extension and velocities are more significant
for smaller values of e. (c) The dominant effect of the velocity field is to
form
Din = A1nJ + B,n (14-28) reduce photon trapping, and hence to increase the photon escape probability.
(d) The line profiles become skewed to the red as the central absorption

N I•
'
Equation (14-26) defining J can be written in the discrete form feature obliterates the violet emission peak [true also in planar geometry,
cf. Figure (14-2)]. In a second sequence of models, the maximum velocity was
](r4 ) = L Wn L a.,,q,[r,; Xn - µ(r.,, p,)V(r,)]u,1n (14-29) held fixed, V(R) = 2, and R/r. was chosen to be 3, 10, and 30; B = r- 2 ,
n• -N I• I e = 10- 4 , x.,lx, = 10- 4 , and Xi = cr-
2
• The effects of velocities on the
source function are small compared to the changes produced by spherical
But, from the spherical symmetry of the problem, /± (z, p, x) = (+ ( - z, p, x), geometry, and the relative departure from static results increases as R de-
and thus u(z,p, -x) =
u(-z,p,x), and v(z,p, -x) -v(-z,p,~~; the~e = creases, probably because the velocity gradient increases. The emergent
relations allow elimination of the values of u at negative x and pos1t1ve z, m profiles all show a strong P-Cygni character. A final calculation worthy of
equation (14-29), in terms ofu at positive x and negative z. Thus mention is a case in which the velocity is constant throughout the atmosphere;
=
i.e., V(r) V0 • For a planar medium the source function would, of course,
]= LN
Wn L adl{<J,[r.,; x. -
'•
µ41 Y,i]u41 • + f/>[r.,; x. + µ41 ¥,i]u.,, 1.} (14-30) be unchanged (though the flux profile computed by averaging over the surface
n• l /a 1 ofa star changes; recall Exercise 14-1). For a spherical medium however, the
radii diverge from the center, leading to a transverse velocity gradient that
where d' = D, + 1 - d. Equation (14-30), when used in the Rybicki method, decreases photon trapping; as a result, the escape probability increases, and
yields V-matrices that are rectangular chevron matrices. Using equations the source function decreases. Even a modest constant velocity of expansion
( 14-28) for all values of i and n in equation (14-30), we obtain a final system has quite dramatic effects on emergent flux profiles.
for J, which is then solved. The computing time required for the solution
scales as TR ~ cN D 3 + c' D3 ; this is less favorable than the result for the EFFECTS OF LINES ON ENERGY BALANCE IN MOVING MEDIA
planar case, because now there are about as many angles (i.e., impact param-
eters) as there are depths. The method is stable and easy to use for small The energy balance in the outer layers ofan atmosphere can be dominated
velocities (i.e., a few times thermal). For larger velocities, the number of by spectral-line contributions. Hence Doppler shifts, which may move a line
depth-points required to resolve the velocity field becomes large, and the away from its rest position and allow it to interact with the continuum (in
462 Radiative Transfer in Moving Atmospheres 14-1 The Transfer Equation In the Observer's Frame 463

which the intensity can be markedly different), can significantly alter the Even though the models upon which the results quoted above are based
temperature distribution. Qualitatively, we can expect three effects to be are very schematic, it is clear that velocity-field effects on line-absorption can
present, over and above the usual boundary-temperature change and back- lead to very substantial changes in the energy balance of the outer layers of
warming that occur in static media. (1) Lines Doppler-shifted away from stellar ?tmospheres. These c~anges could, in principle, influence the hydro-
their rest frequencies can intercept continuum photons from deeper layers; dyna~1cs of the flow. Thus, m a pulsating atmosphere, energy deposition in
we may call this the irradiation effect. The absorbed continuum flux provides the Imes. produces a kind of radiative precursor that could affect shock
additional energy input to the gas and, because the color temperature of the propa~atlon; for expanding atmospheres, significant energy deposition could
flux exceeds the ambient temperature locally, irradiation will lead to a net occur m the transsonic flow region, which might alter the nature of the stellar
heating of the outer layer. The effectiveness of the input is determined by the wind. It is also possible that velocity-dependent line-absorption contributions
strength of the coupling of the lines to the thermal pool; thus heating will to energy balan~e c~uld. affect the flow-dynamics of novae, supernovae, and
be greatest when e = 1, and should be negligible for e -. O. (2) As the lines ma~s-exchange m bmanes. Much further work remains to be done on this
shift away from their rest positions, photons that would have been trapped subJect.
in the deeper layers by overlying line absorption now encounter only con-
tinuum opacity, and hence may diffuse freely to the surface and escape; we LINE-FORMATION IN TURBULENT ATMOSPHERES
may refer to this effect as escape-enhancement. In general, an increase in
photon escapes will lead to a cooling of deeper layers. (3) A velocity gradient . As mentioned earlier, the dichotomy of velocity-field effects on spectrum
in the atmosphere smears the lines over a larger bandwidth, thus impeding !me-formation into the "microturbulent" and "macroturbulent" limits is
the free flow of photons; we may call this the bandwidth-constriction effect. obviously o~ersimplified. In these two extreme limits, the effects of the velocity
At depth in the atmosphere, where the diffusion approximation is valid, a field u~on !me strengths and profiles may be predicted from simple phenom-
velocity gradient causes little, if any, change in the temperature structure enolog1~l a~guD?ents. To stud~ the.influence of velocity fields that have a
when the scale of the velocity variation is large comphred to a photon scale which 1s neither zero, nor mfimte, with respect to a photon mean-free-
mean-free-path. But if a velocity shift of the order of a Doppler width occurs path, ~ detailed computation must be made. One could, in principle, specify
in a mean-free-path, then bandwidth constriction leads to a decrease in. the a particular run o~ th: velocity, solve the equation of transfer, and average
effective radiation-diffusion coefficient, and hence to increased backwarming. over as many reahzatlons of the velocity field as are necessary to embrace
In the extreme limit of an abrupt velocity step near the surface, photons the possible ranges of its inherent degrees of freedom. Such an approach,
emerging from lower layers in fonnerly-open continuum bands encounter however, v:'ould be co~tly, and ~ould not yield direct insight into the problem.
opaque material, and have their escape impeded; this might more appro- An attractive alternative (motivated by the expectation that in stellar atmo-
priately be tenned a "backscattering" or "reflector" effect. sphere~ the velo~ity field is chaotic, and possibly even turbulent in the hydro-
Studies of the influence ofline-shifts on energy balance have been made for dynamic sense) 1s to assume that the velocity is a random variable, described
highly schematic picket-fence models, in planar geometry with an abrupt locally b~ a. probability distribution for the amplitude, and nonlocally by a
velocity jump (428), and in spherical expanding atmospheres (445). The characteristic correlation length.
velocity-step can be regarded as a caricature of a shock front. The escape- Considerable progress in solving the transfer equation in turbulent media
enhancement and irradiation effects show very clearly for the velocity-jump, has recently been made with two distinct approaches. One formulation,
where the layers above the jump heat markedly (one case gives !J. T ~ 1100°K dev~l~ped by the Heidelberg group (70, 325; 234; 235; 236; 557), employs
for T.rr = 10,000°K), and a cooling of a few hundred degrees occurs imme- the Joint probability ~(z; v~ /) ~hat, at point z, the velocity lies in the range
diately below. For the expanding spherical atmospheres one obtains large (v, ~ ~ dv), and the mtens1ty m the range (/, / + d/), for Markov-process
irradiation-effect rises at the surface, and substantial backwarming below. variations of v and /. P, or some other suitable distribution function derived
The irradiation-effect temperature rise is larger for extended models than from P, is found by solving a Fokker-Planck equation. This method is
for nearly-planar models; this is because the discrepancy between the color pow~rful and ge~eral, and allows the treatment of velocity fields that are
temperature of the flux, and the ambient temperature characteristic of the cont1~uous func~1ons of depth. However, the resulting partial differential
local energy density, becomes larger as atmospheric size increases. For one equat10ns are difficult to solve, and a succinct description of the method
extreme case, the velocity field produces a change !J.B/B ~ 3, which implies would presume considerable familiarity of the reader with the mathematical
llT/T ~ 0.33, or !J.T ~ 10,000°K for an O-star. methods for treating Markov processes. A rather different formulation has
464 Radiative Transfer in Moving Atmospheres /4-1 The Transfer Equation in the Observer's Frame 465

been developed by the Nice group (49; 226; 227). The flow is conceived as We may obtain an analytic~! solution of the problem if we introduce the
consisting of turbulent eddies or cells. The velocity is taken to be uniform following additional assumptions: (I) set B,(t) = B 0(1 + ext); (2) use a
within each cell, and to jump discontinuously at shar-p boundaries that separate Milne-Eddingtonmodel-i.e.,p = constant,a = constant,6.v0 = constant;
the cell from neighboring cells with uncorrelated velocities; this description (3) adopt a constant eddy density n(t) = n. All of these assumptions may be
is called a Kubo-Anderson process. Although the discontinuity of the velocity relaxed if the problem is solved numerically (49: 226). To simplify the treat-
structure is unphysical and introduces some artifical high-order correlations ment still further, we consider the intensity emergent at disk-center, and set
in the field (49; 70, 325), this approach nevertheless has the advantages of µ = I. The transfer equation thus becomes
yielding exact analytical results in certain limits, and of expository simplicity.
We shall therefore describe the Nice method here, but quote results from both [ol(t, x)/ot] = [I + .fJH...(t)][/(t, x) - B0 (1 + ext)] (14-36)
bodies of work.
We assume that the cell boundaries, at which the velocity changes, are where H...{t) denotes H[a, x - uh(t)] and is constant between successive
located at random continuum optical depths {t.}, distributed according to jump-points. If we define
a Poisson law characterized by an eddy density n(t), which gives the reciprocal
of the correlation length I (in continuum opticat' depth unit_s) of the velocity q,.(t) = exp {- J~ [l + /lH,.{t')] dt'} (14-37)
field. The probability that no jump has occurred on the interval (t', t) is given
by exp[ - J:. n(t") di"]. Let vh(t) denote the hydrodynamic velocity of the cell
then the emergent intensities in the continuum and the line are
at -r; these velocities are independently distributed according to a probability
distribution function P(vh), which we shall take to be Gaussian. Let v,h(t)
denote the thermal velocity of the line~forming atoms at t, and measure all (14-38)
velocities in thermal units, uh(t) = vh(t)/v,h(t), and frequencies from line
center in units of the corresponding Doppler widths, x = llv/llv 0 (t). Charac- and 1(0, x) = f 0
"' B0 (1 + ext)q...(t)[I + f?H.,(t)] cit (14-39)
e
terize the turbulent field by a dispersion (in -thermal units); then
It is easy to show that /, = B0 (1 + ex), and that the intensity in a line of in-
(14-3J) finite strength (/J -+ =o) is 8 0 ; the absorption depth of such a line will be
A 0 = ex/(1 + ex). Then, in general, the line absorption depth a,. =
The transfer equation to be solved in [I, - 1(0, x)]/1, can be written ·
µ[ol(z, µ, x)/oz] = -(Xe + x,<f,,.)J(z, µ, x) + x,S, + x,S1 (14-32)
(14-40)
where <J,,.(t) =
<J,[x - µuh(t)]. Equation (14-32) is a stochastic equation-i.e.,
the coefficients in the equation are random variables. We now assume that The ensemble average of the line profile over all possible realizations of the
the turbulent velocity field influences only the line absorption coefficient via velocity field is thus
Doppler shifts; fluctuations in the continuous opacity, source functions, and
occupation numbers are expected to be of secondary importance, and are (14-41)
ignored here. We.adopt LTE and ignore scattering, so that S1 = S, = B,.(t),
and employ a Voigt line-profile so that while the average of the reduced equivalent width is

[x,(-r)<J,,.(-r)]/x,(i) = fl(r)H[a(r), x - Jwh(t)] ( 14-33) (14-42)


where fl(t)= (1tte /mc)[J,1n1(t)(l
2
- e-h•/kT)]/[x,(t) llv 0 (t)] (14-34)
Thus the crux of the problem is the calculation of (q...(t)).
and a(t) = r/41t llv 0 (,) (14-35) · The function (q_.(,)) can be found from the solution of an integral equa-
tion, which is obtained by the following arguments. First, at a given point t,
Here r is the damping width, and 111 denotes the occupation number of the the probability that no jump has occurred on (0, t) is exp( - nt); the cor-
lower level of the transition. responding contribution to (q...(t)) is then exp( -11t)(q,.(t))s, where the
466 Radiative Transfer In Moving Atmospheres 14-1 The Transfer Equation in the Observer's Frame 467

static average is We do not require the inverse transform of Q,. because, as can be seen by
comparison of equations (14-48) and (14-41), the residual intensity can be
(q,h))s = (exp[ -(1 + PH,.)t]) expressed entirely in terms of Q,c(O). Thus we have the general result that

= f_:°.., exp{-[1 + PH(a, x - uh)]t}P(uh) duh (14-43) (a,.)/A 0 = 1 - ((n + 1 + PH_.)- 1 )[1 - n((n + 1 + pH_.)- 1 ) ] - 1
(14-51)
Here uh (and hence H,.) is constant over the entire interval (0, t). On the other
hand, suppose that one or more jumps have occurred on the interval (0, t), · We can recover the macroturbulent limit by taking the correlation length
and let t'{ < t) denote the last jump point. On the interval {t', t), H,. will be I = ex>, or n = 0, so that the entire atmosphere along th~ ray moves with
constant, hence constant velocity. From equation (14-51) we find
q,.(t) = exp[ -(t - t')(l + PH,.)]qx(t') (14-44)
(a,c(/J))m■cro = Ao(/JH,./(1 + {JH,.)) (14-52)
The probability that the last jump, at t', occurs between t' and t' + dt' is
This result agrees with our intuitive expectations when we recognize that
exp[ -n(t - t')]n dt'; thus averaging (14-44), and summing over all t', we
the argument in the bracket is just the emergent residual intensity of a line
obtain the contribution to (q.(t}) from jumps on (0, t), namely
in the Milne-Eddington model (cf. §10-3), shifted bodily by a velocity uh,
(q..(t))Jump = f: (exp[-(t - t')(l + PH,.)]q:,c(t')) exp[ -n(t - t')]n dt' and then averaged over the probability distribution of uh. We may express
the general result in terms of the macroturbulent limit as
(14-45)
(a,.(P)) = (n + l)(ax[P/(n + l)])m■cro[l + nAo -l(a,.[(1/(n + l)])macro]- 1
Now all steps in H,.(t") on the interval O < t" < t' are independent of the (14-53)
value of H,. on the interval (t', t), hence
To obtain the microturbulent limit, we let n -+ oo. First, note that for P-+ 0,
(exp[ -{t - -r')(l + pH,.)]q"(t')) = (exp[ -(t - t')(l + /JHx)])(q..(-r')) equation (14-52) can be expanded as (ax(P))macro ~ A 0 ({1(H) - P 2 (H 2 )),
= (q,.(t - t'})s(q,.(t')) (14-46) which yields

Adding the static and jump contributions we have, finally, (14-54)

(qx{t)) = e-"'(q,c(t))s + f; (qx(t'))(q,.(t - t'))se-n<r-r'>n dt' (14-47)


Because the Voigt function is a convolution of a Lorentz profile with a
Because we have taken n to be depth-independent, the integral in equation
Gaussian [ cf. equation (9-34)], we obtain, from an interchange of the order
(14-47) is a convolution integral, and we ma~;a:pply..a ~aplace transformation
of integration, .!:,,,., ..
to obtain the:.~olution. Thus let J•i· in ,'- • · -..--
= (1'-'+'•~'lttH[a(l + e2)-t, x(l + e2)-t]
J e-"(q,.(t)) d/ (H(a; ;))
1-..,'I-.
(14-56)
Q"(s) e "" (14-48)
0
and That is, in the microturbulent limit the Doppler width ~v 0 is increased by a
factor of (1 + 2 )t. e .
S,c(s) ~Jo~,d:~•(qh))s dt = f:'.., duh P(uh) Jti:,~x.~l_~(l + PH")t]e-" dt In the limit of zero turbµle~e the standard Milne-Eddington curve.,9.(
growth is given by [cfJ:~~a;~i~J]p-38)] ., .. 1'..;. '"\··
,: lj.,., = ([s + (I + PHx)]- 1).;;.\, (14-49)
Thenftomequati6n"(_1~!!f1!)wefind
; ) ,.. J;
.:,-,::·.:•;::, ~ , .;mi: f.}}<~: ;f;;,, f~~lft(a, x)[l + PH(a, x)]- 1
, dx ,,: :" (14-57)

; "'· . . ,, Q~s)b=S,c(s + n)/:{1 - nS,c(s 4' n)] (14-50) N"ol;;,\tub~t'ft~'ti~i~qtililioiR fl'#iS1j•iftto (14-42)-and interchanging the order
of integration, we find w:, •• ,0 (a, P) = ( W~(a, P» = W~(a, P); i.e., in the
Exercise /4-4: Derive equation (14-50). macroturbulent limit the curve of growth is unchanged-which, of course,
468 R,11/iative Tmn.yfi•r ;,, M11vin11 Atmospheres /4-/ The Tran.tfer Equation in tire Observer'.f Frame 469

is the expected result. In the microturbulent limit, by substitution of equa- Results for the average depth, and its dispersion, of a strong line in an
tions (14-54) and (14-S6) into (14-42), we obtain immediately e
atmosphere with the turbulent-velocity parameter equal to the thermal
velocity, and a density of 10 "eddies" per unit continuum optical depth, are
w~lcro(a, /J) = (1 + e2)½W~[a/(1° + e2)½, {J/(1 + e2)½J shown in Figure 14-4. Curves of growth fore = 1 are shown in Figure 14-5.
There we see that the dependence of the theoretical curve upon the eddy
This shows that the linear and damping parts of the curve are unaffected, density, n, implies that a comparison of an observed curve to a theoretical
while the flat part rises by a factor of (1 + 2 )½. Between these limits the e curve with e = 0 cannot lead to a unique value fore. In general, the value of e
curve of growth is found by numerical integration of equation (14-42), using deduced in the microturbulent limit will be a lower bound on the actual value.
equation (14-53). In addition to expressions.for (a.,,) and W*, it is possible Note that this effect is opposite to that produced by departures from LTE-
to derive an expression for u,,, the rms fluctuation in the residual intensity which, typically, raise the flat part of the curve of growth even when no
that would be seen along the slit in a spectrogram of perfect resolution; velocities are present (cf. §11-4).
i.e., a,. = A 0 - 1(<a,,2) - (a,,) 2 )½ [see (49)]. Detailed calculations, using a realistic solar atmosphere, have been made
for the O I )..).7771, 7774, 7775 lines, and the Fe I J..).5576, 5934, 6200 lines
(235); an excellent fit to observed profiles is achieved. In the fitting procedure,
the loci of points in the (e, /) plane [I = l/11 (km)] that match the observed

1.0 -----
....... ....
'
'' '' \
' \ \
\
0.8 '' \
\
'
'' \
\
\
I
I
\
0.6 \
I
(a~)/Ao log w•
0.4
0

0.2

0.0 L____-1.._-=:~~=:t::::...---1
0 2 4 6
0 2 3 4
X

FIGURE 14-4 log/J


Absorption depth in line with fJ = 100, in a turbulent atmosphere
with turbulent velocity~ • I, and eddy-density n .. 10. Abscissa: FIGURE 14-5
x "" Av/Av0 • Solid curve: average profile from all realizations of Curves of growth in turbulent atmospheres with~ = I. Orclln111e: Logarithm
velocity distribution. Dashed curves: average profile ± the rms of reduced equivalent width w• a ( W,./A 0 Av0 ). Ab.~clssa: Logari1hm of line-
fluctuation seen by spectrograph of perfect resolution. From (49), strength P = 1.,/1.,, Curves are labeled with depth-independent eddy density.
by permission. · From (49). by permission.
470 Radiative Transfer in Moving Atmospheres 14-2 Sobolev Theory 471
intensities at several values of L1J.. from line center, as well as the equivalent spheres will undoubtedly emerge from further detailed application of the
width, intersect at almost a unique point. Thus it proves possible to deter- methods described above to the analysis of the solar spectrum, and to stellar
e e
mine both and I= l/n (km) uniquely; numerically ~ 2.2 km s- 1, and spectra where possible.
/ ~ 150 km. The velocity is larger than that usually adopted as a micro-
turbulent velocity; the correlation length is of the same order as the photo-
spheric scale height. It would be of interest to ascertain whether the apparent
14-2 Sobolev Theory
near-equality of the correlation length· with_ the scale height has hydro•
dynamic significance, or is mere coinciden~. • . .The existence of large-scale, rapid (sometimes violent) expansion in stellar
Allowance for a finite scale in the velocity structure also has important atmospheres is well established observationally. Probably the first objects
implications for the interpretation of the center-to-limb variation of so~ar in which such motions were unequivocally recognized were the novae (and
line profiles. A long-standing problem has been that the profiles at disk later the supernovae). In their spectra, after the explosive increase in the star's
center have a characteristic V-shaped appearance, suggestive of macro- luminosity, one observes absorption lines strongly violet-shifted from their
turbulence, while those at the limb have a U-shaped appearance, suggestive· rest positions, indicating material flowing rapidly toward the observer. These
of microturbulence. Moreover, the microturbulent velocities derived from lines are accompanied by extensive red-shifted emission features, resulting
limb profiles are typically larger than those derived from the same lines at in characteristic P-Cygni profiles resembling those in Figure 14-6. In nova
disk center. spectra these features are transients, and indicate episodes of violent ejection
This result has been advanced as evidence for anisotropic turbulence (a of the outer layers of the star. In other objects [the classical P-Cygni stars
situation difficult to understand hydrodynamically). Both of these effects can (79; 366)], these lines, though variable, are more-or-less permanently present
be understood, at least qualitatively, in terms of a finite eddy density for the in the spectrum, and indicate persistent outflow of material. Beals first
velocity field (226). If n0 (t) gives the eddy density at µ = l, !hen the appro• recognized (75; 77) that the great breadths of lines in WR spectra (indicating
priate density at other values ofµ is n0 (t)/µ; this follows from th~ require- velocities of the order of 3000 km s - 1 ) could be interpreted in terms of rapid
ment that the turbulence be isotropic, so that we encounter the same eddy outflow of material. He suggested that the flow was driven by radiation
density per unit path-length along the ray. If the density is depth-independent,, pressure, a conclusion supported by current dynamical models. Similar
equation (14-53) is unaltered except that we replace n with n0 , and now conclusions can be reached for the Of stars. We know today that in the WR
Ao = cxµ(l + cxµ)- 1 • In general, n0 (t) must vary with t. Suppose we demand and Of stars, and in many early-type supergiants, there are transsonic stellar
that the eddies have constant geometrical size L; then n0 (t) = [x,(t)L]- 1•
For the HSRA x,(t) oc t, hence n0(t) oc t- 1• Thus the eddy density pertinent
to observations near the limb must be higher than at disk center, and this
result provides at least a qualitative explanation of the center-to-limb effects
mentioned above. An analysis of the Mg I J..4571 line has been performed
(226) using the HSRA. A good fit to the profiles is obtained with isotropic
turbulence characterized by e= 1.2 km s- 1 and I ::::: 70 km. The smaller FIGURE 14-6
value for I quoted here, compared to the results for O I and Fe I mentioned P-Cygni profiles of the hydrogen lines in
earlier, may reflect the fact that this analysis is based on a Kubo-Anderson the spectrum of HD 190603 as observed by
discontinuous velocity field while the other uses a-continuous field. It turns Beals (79). Ordinate: observed flux in
e,
out that for a given a given line-strength is always achieved at a smaller
units of the cominuum (successive profiles
arc displaced for clarity). Abscissa:
value of I in the discontinuous case (237). displacement from line center in velocity
Non-LTE effects produce line profiles that are deeper in the core, and units-i.c., v .. c !;.)./l
approach the LTE profile in the wing (227; 236). The deviations of the non-
LTE profiles from their LTE counterparts become larger as the turbulent
e
velocity parameter increases, and as the correlation length I decreases. -200 0 +200
Much useful information about the nature of velocity fields in stellar atmo- v (km scc- 1)
472 Radiative Transfer in Moving Atmospherc•,1· 473

winds (cf. §15-4), that have vanishingly-small outward velocities in the deeper Expanding envelope
layers, and a large outward acceleration producing very large velocities
(v/c ::::: 0.01) at great distances from the star.
The solution of the transfer equation in a spherical expanding medium is
difficult, and in the early work by Beals (75; 76; 77; 78), Chandrasekhar (149),
Gerasimovic (243), and Wilson (674) it was assumed that the material was
optically thin so that transfer effects could be ignored. This approach, while
obviously oversimplified, has nevertheless contributed a good deal to our
basic picture of the physical situation. A major advance occurred with the
brilliant realization by Sobolev (590; 591; 15, Chap. 28) that the presence
of the velocity gradient in an expanding medium actually simplifies line-
transfer problems, for it dominates the photon escape and thermalization
process, and implies a geometric localization of the source function not present
in static problems. In Sobolev's theory the solution of the transfer problem is,
in effect, replaced by the calculation of escape probabilities; the basic theory
has been refined and extended by Castor (134), and has been applied to
fairly realistic calculations of spectra from multilevel atoms in WR envelopes
(139; 140).
bsorption-reature 1
region
SURFACES OF CONSTANT RADIAL VELOCITY

Consider a spherically-symmetric, radially-expanding envelope sur-


rounding a star with a fairly well-defined photospheric surface, as sketched
in Figure 14-7. With this basic model we can explain qualitatively the maih
features of P-Cygni profiles such as those in Figure 14-6. For the discussion
in this subsection, we shall assume that the envelope is essentially transparent,
so that every photon emitted towards an external observer can be received.
This approach yields insight, and results that will be useful later. Throughout ~
Observer
the discussion we make. use of the fact that most of the line emission (or
FIGURE 14-7
absorption) occurs at line center; thus radiation from a given region, as Schematic diagram of expanding envelope surrounding a stellar surface. The
received by an external observer, appears mainly at the line-center frequency, material in the occ11/tecl region is blocked from view by the stellar disk, and cannot
Doppler-shifted by an amount corresponding to the velocity of the material be seen by an external observer.
along the line of sight.
The material behind the stellar disk is in an occulted region, and cannot
be seen by an external observer. The matter projected on the stellar disk lobes to the sides of the disk, we receive photons either emitted thermally
can either (a) simply emit radiation without significant reabsorption as or scattered from both the stellar and diffuse (from the envelope itself)
occurs in, e.g., a forbidden line Jo a nebula or in a thermally excited medium radiation fields. The velocities along the line of sight in the emission-Jobe
where T, » T, (the color temperature of the radiation from the underlying material range from positive through negative values, and produce a sym-
photosphere), or (b) absorb the incident photospheric radiation and scatter metric emission feature extending from a wavelength to the violet of the
it out of the line of sight. From this material, in case (a) we would obtain a rest-wavelength, to redder wavelengths. Because material is occulted by the
violet-shifted emission feature, while in (b) we obtain a violet-shifted absorp- star, the maximum redshifts that could be produced will not be observed,
tion dip characteristic of P-Cygni profiles. From the matter in the emission and, in general, we expect to derive information about the maximum flow
474 Radiative Transfer in Moving Atmospheres 14-2 Sobolev Theory 475

velocities from the position of the blueward edge of the absorption feature thin zone, centered on a surface of constant radial velocity such that v, =
(or emission feature if no absorption is present). The volume of the emitting µv, = x. (Norn: here the term "radial velocity" has the usual astronomical
region can be enormous compared to that of the star, and the integrated meaning of velocity along the line of sight, not the more fundamental meaning
contributions to the observed emission line from this volume may far out- of the velocity v, measured from the center of the star.) In the idealized limit
weigh the amount of energy received from the stellar disk. As a result, the that the width of the line profile is negligible (because V1hermal « Vnow), the
peak intensities in strong emission lines may be several times the background zones degenerate to the radial velocity surfaces themselves, which therefore
continuum value (see Figures 14-6 and 14-9). Also, when the size of the stellar play a basic role in the theory.
disk is much smaller than that of the emission ·region, occultation effects The shape of these surfaces depends upon the nature of the velocity
become unimportant. Finally, some lines are much more opaque than others field-which, ultimately, must be obtained from dynamical calculations. We
and may, therefore, have a larger effective volume for emission; thus in can gain insight, however, by consideration of some simple velocity laws of
Figure 14-6 we see a transition from quite strong emission at Ha, to practically the form v = r" (in units v0 = r 0 = 1). (a) Suppose v, = constant. This
no emission at Hy, where we see essentially the photospheric Hy absorption- law could apply to a thin spherical shell (e.g., a planetary nebula at large
line. distance from the star), or in high-velocity flows nearing terminal velocity
To make these notions more quantitative, we can compute the energy [see (c) below]. (b) Suppose v = r. This law could apply in the case of an
received at frequency v by an external observer as explosive ejection that started at some time t 0 such that (t - t 0 ) = r/v; here
the faster-moving particles outrun the slower ones, giving the linear relation
(14-58) of v with r. (c) Suppose the gas leaves the star with velocities greater than
escape velocity. We then can write v = v"'(1 - r,/r)i, which provides a crude
where the integration is carried out over the entire unocculted volume. In simulation of a transsonic wind; the flow accelerates everywhere on the range
performing the integration we may use, as was done in o..ur earlier work in r, ~ r ~ oo. (d) If the material is ejected with just the escape velocity and
spherical geometry, either (r, 0) coordinates with the axis of symmetry from is decelerated by gravity, we may take v = r-½. Each of these laws has a
the center of the star through the·observer, or (p, z) coordinates (see Figure distinctive set of constant radial-velocity surfaces.
7-27). Equation (14-58) can be made more explicit if we write 17(r, v) ,=
ij(r)q,(v - v0 (1 + µv,/c)], which accounts for the shift ofline center, as seen Exercise 14-5: Show that the surfaces v, = constant for cases (a) and (b) are the
by an external observer, to v0(1 + v./c), where v, = µv, is the velocity along cones O = cos- 1 µ = constant, and the planes z = constant, respectively, in the
the line o'r sight resulting from the expansion velocity v,. If we suppose that usual (r, 0) and (p, z) coordinate systems.
rf(r) = 170 (p/p 0 )«, (where reasonable values for a lie in the range O ~ a ~ 2)
=
and, further, assume v v0 (r/r 0 )", then from the requirement of continuity, The constant-radial-velocity surfaces for laws of -the form (c) and (d) are
pvr 2 = p 0 v0 r0 2 , we obtain finally ij(r) = 170(r/r0 )-<n+l'/a., Choose units such shown in Figure 14-8 (a and b).
that r0 = v0 = l (these quantities referring to values at the photospheric From what has been said above, some far-reaching conclusions can be
surface), and measure frequency displacements from line center in units drawn. The fact that the surfaces of constant radial velocity extend over
x = (v - v0 )/llv 0 , where llv0 = v0 v0 /c. Then large regions (infinite if the flow is not decelerating) implies a complete
breakdown of the Eclclington-Barbier relation for expanding atmospheres.
(14-59) We can no longer associate a given frequency in the line profile with a
specific position r in the envelope, but only with a wide range of values
In principle V refers to the entire unocculted volume; in actuality the volume (r, r + llr). From the viewpoint ofan outside observer, geometric localization
of integration can be defined more precisely. occurs only if variations in total particle density and ionization-excitation
Most of the emission observed at frequency x will arise from regions where equilibria confine the region of high emissivity. What is worse, this conclusion
the line center frequency, after Doppler shifting, is at the observed value. is independent of the intl'insic line-strength (317). So we can no longer, in
The observed flow velocities (up to 3000 km s- 1) in WR and Of atmospheres principle, obtain a depth-anctlysis of atmospheric structure, with a precision
vastly exceed the thermal velocity ( ~30 km s- 1). Therefore the geometrical better than the charctcteristic llr defined above, by examining weak and strong
region from which the emission at any one frequency arises must be a very lines. Clearly these considerations imply severe modeling and diagnostic
14-2 Sobolev Theory 477

problems for expanding atmospheres. The problems are· even more' severe
in the case of decelerating flows where, as can be seen in Figure 14-8b, a
particular line of sight may intersect a surface of constant radial velocity at
two distinct points; hence two regions, which may have vastly different
physical properties, contribute to the information received by the observer.
Furthermore, in this case these two distinct regions can also interact radia-
tively, and the SoboJev method to be described below requires reformulation.
Taking into account the geometry of the surfaces of constant radial
velocity, equation (14-59) can be applied to calculate observable line profiles.
For instance, as Beals first showed (75; 76; 77; 78), if we can ignore occulta-
tion, then the profile from an optically thin shell expanding with v =
constant is flat-topped; an example of such a profile appears in Figure 14-9,
where the )5696 line of C III and the ,l58O8 line of C IV in the spectrum of
the hot WR star HD 165763 are shown. The rounded profile for the C IV
line indicates it is optically thick (see below). Beals's result can be seen by
inspection of equation (14-59) using an (r, 0) coordinate system. We note
that radiation at each value of v, (and hence of x in the line profile) is con-
tributed by a conical volume element centered on µ = cos 0 = x, in a range
(a) dµ; the volumes of all such elements are manifestly identical. Results for
other velocity laws may easily be derived so long as the line is optically thin;
this approximation will not be true in general, and the need for it is overcome
by Sobolev's method.

HD165763
WC5

F(i.)

5700 5800

FIGURE 14-9
Observed profiles of C 111 ).5696 and C IV ).5808 in the
WC5 slar HD 165763. Notice the flat-topped profile for
(b) Observer the transparent ).5696 line, and the rounded profile for
FIOURE 14-8 1he optically-thick ).5808 line. From (369), by permission.
Surfaces of constant radial velocity, v, "' constant.
(a) v(r) = 11,:(I - 1/r)t. Curves are labeled with 1,.Jv,.
(b) v(r) = r-t. From {366), by permission.
14-2 Sobo/ev Theory 479
478 Radiative Transfer in Moving Atmospheres
As was done earlier, we measure velocities in units of a thermal velocity
Exercise /4-6: (a) Suppose that the intrinsic line profile is cp(x) = o(x). Consider
an envelope for which v = v0 = I. Ignoring occultation, show that £, = £ 0 = [i.e., V(I') = v(r)/v,h] and frequency displacements from line center in Doppler
constant for - I ..; x =E; I, and£. = 0 for !xi > I. Derive expressions for £ 0 from units, [i.e., x = (v - v0)/6v 0 , where 6v 0 = v0v1h/c]. The optical depth along
equation (14-59), and show that different results are obtained for IX < 1.5, IX = 1.5, a ray to an observer at infinity can then be written [cf. equation (14-22)]
and IX > 1.5. Show that for oc ..; 1.5 the envelope must be bounded, r ..; R, but that
no restriction is required for oc > 1.5. Accounting for occultation, show that£. = 0 -r(z, p, x) = J,"'' x(z', p, x) dz' = J,"'' x,(r')</>(x') dz' (14-61)
for x < Xmtn• where Xmin = -(1 - R- 2 )t, and write an expression for Ex on the
range Xmin ..; x ..; 0. (b) Perform a similar analysis for v = r; determine £ 0 for
appropriate ranges ofoc, and derive analytical expressions for the profiles of£./£ 0 ,
where r' =(z' 2
+ p 2 )t, µ' =(z'/r'), and
i11c/11ding occultation effects.
x' = x'(z', p, x) = x - V,(z') =x - µ' V(r') (14-62)

ESCAPE AND THERMALIZATION IN AN EXPANDING MEDIUM The main contribution to the integral in equation (14-61) must come from
the region where x' = 0-i.e., from z' = z0 (p, x), where z0 is chosen such that
Consider now the formation of a line in an optically-thick expanding z0 r 0 - 1 V(r0 ) = x; here r 0 = (z 0 2 + p 2 )t. The surface z0 (p, x) is, of course,
envelope surrounding an opaque core of radius re; assume that the effects of just a surface of constant radial velocity. To a good approximation, we can
a background continuum can be ignored. Ifwe now transform to a coordinate then replace x,(r') with x,(r0 ), and remove this factor from the integral. We
system at rest with respect to a particular fluid element, and inquire what now change the variable of integration from z' to x'; in view of equation
happens as we look along a given ray, it is clear that (as a result of the velocity (14-62) the transformation is
gradient) there will be a differential Doppler shift of each successive sample-
point along the ray relative to the test-point. Eventually this shift becomes -(iJx'/iJz)p = (iJV,jiJz)p = (iJ{µ(z, p)V[r(z, p)]}/iJz)p
so large that no line photon emitted within the effective Bounds of the line = 2
µ (iJV/iJr) + (1 - µ 2 )(V/r) = Q(r, µ) (14-63)
profile (assumed to be limited by some ± Xma,) can interact with the line
profile at the test point. The velocity field has introduced an intrinsic escape where µ and r are again understood to be functions of z and p. If the interac-
mechanism for photons; beyond the interaction limit (measured from the tion region is small, the transformation coefficient written above may be
point of emission) they no longer can be absorbed by the material (even if. it assumed to be essentially constant and may be evaluated at the resonance
is of infinite extent!) but escape freely to infinity. Thus there is a definite limit point z = z0 (p, x). Then if we define
to the size of the region within which photons emitted, or scattered, can have
any effect upon the intensity within the line at the test-point. In the limit of (14-64)
large velocity gradients the interaction region will be small, and may, there-
fore, be presumed to be nearly homogeneous in its physical properties
where clearly <I>( - co) = 0, and <I>( co) = 1, we can rewrite equation (14-61)
(temperature, density, ionization state, etc.). The theory then can be for-
as
mulated in terms of local quantities, and a parameter p that gives the
-r(z, p, x) = -r( - co, p, x)<I>[x'(z, p, x)] (14-65)
probability of photon escape summed over all directions and line-frequencies.
In the limit of negligible transfer effects, we can therefore write
where x'(z, p, x) is defined by equation (14-62) and

J(r) = (1 - P)S(r) + Pelc (14-60) -r(- co, p, x) = x,(r0 )/Q(r0 , µo) = -r0 (r0 )/{1 + µ 2 [(d In V/d In r) - 1]} 0

The first term is derived from the value that J would have in the limit of no (14-66)
escapes, namely J = S, corrected for velocity-induced escapes. The param-
eter Pc measures the probability of penetration (summed over frequency Here x,(r0 ) = (ne 2 /mc)JiJn,(ro) - (gifg1)nj(r0 )]/6v0 (14-67)
and angle) of the specific intensity le, emitted from the core, to the test-point.
We must now calculate p and Pc·
and -ro(ro) = x,(ro)/(V/r)0 (14-68)
480 Radiative Transfer in Moving Atmospheres 14-2 Sobolev Theory 481

In equations (14-66) through (14-68) it should be borne in mind that r 0 and Note that both fJ and Pc are defined essentially in terms of local quantities:
µ0 are functions of p and x, i.e., r 0 = r 0 [z0 (p, x), p] and µo = µo[zo(P, x), P]. the local opacity and velocity gradient. Given these values, one can compute
Let us now choose a fixed value ofr and calculate /3(r); because of spherical J from equation (14-60) without actually solving a transfer equation; thus
symmetry, integration over µ can be effected by using the above results for we see the enormous simplification that has been achieved. For the particular
various values of p. The escape probability along any ray is just exp(-.:\, co) case of a two-level atom, where the source function (assuming complete
where .:\tco denotes the optical path length from the test-point to infinity. redistribution) is given by S = (1 - e)J + eB, we may use equation (14-60)
Thus, summing over angle and frequency, we have to write
S = [(1 - e)fl,lc + eB]/[(l - e)P + e] (14-73)
P(r) = ½f~ 1 dµ f_:°"' dx t/>[x'(z, p, x)] exp{ - ,[z(r, µ), p(r, µ), x]} (14-69)
which shows that knowledge of p and p, is sufficient to determine S. Further,
if we ignore the continuum contribution (fl,[c = 0), equation (14-60) allows
Here we have assumed that photons that hit the opaque core are absorbed
us to write the net radiative bracket for each line of a multilevel atom im-
and hence lost. To evaluate equation (14-69), we use equations (14-64)
mediately, namely Z 11 = {J 11 ; we shall exploit this result in the discussion of
through (14-67), and assume that the material in the interaction region is
multilevel atoms. .
sufficiently homogeneous that the distinction between r0 and r may be
It is very instructive to consider a uniformly expanding plane-parallel
ignored. Then
at.mosphere, for then we can obtain expressions that show the effects of the

/J(r) = !2 f -1
I
dµ i' d<I> exp[ - x,(r)<I>/Q(r, µ)]
Jo
velocity gradient on the thermalization of the source function in a particu-
larly transparent way [see (273, 87) and (406)]. Lett denote the integrated
line optical depth defined for a medium at rest; assume the velocity gradient
y = oV/o, is everywhere constant. The specific intensity at a test point t
in direction µ is

For the special case that V = kr, Q(r, µ) = k, and equation (14-70) reduces
considerably to l(,,µ,x)= Jrjco S(t')exp [ -µ· I j(r'-r)
Jo q,(x+yµt)dt ] q,[x+yµ(,'-t)]dt'/µ
/J(r) = {1 - exp[ -t0 (r)]}/t 0 (r) (14-71)
(14-74)
where now t 0(r) = k- 1 x,(r). The same result is obtained if the angle-depen-
dent terms in equation (14-63) are merely ignored. Thus the source function for a two-level atom is given by the integral equation
To calculate p,, assume that the test point is relatively far from the core
(i.e., that the surface of the core is at - oo). Then, from its physical meaning,
Pc can be written
S(t) = (I - e)](t) + eB(t) = (1 - e) f_:°"' Kp It' - •I S(t') dt' + eB(t)
(14-75)
{Jc(r) =~ f-;' dµ s: d<I> exp[ - x,(r)<I>/Q(r, µ)] where the kernel function

= x, - 1{r) 21 Ji,,
jl
{1 - exp[ - x,(r)/Q(r, µ)]}Q(r, µ) dµ (14-72) · K 1i(s) = 1f_:" ' f
dx 0
1
dµ µ- 1q,(x)q,(x + yµs) exp [ - µ- 1 f: rp(x + yµt) dt]
(14-76)
whereµ, = [1 - (r,/r}2Jt. Again, for the special case ofa linear velocity law
we obtain a considerable reduction, namely /Jc(r) = Wf3(r), where W is the It is easy to show that, unlike the static case where the kernel is normalized
usual dilution factor given by equation (5-36). The result just quoted is what to unity, in the present case the effects of escapes lead to
would be expected physically, because W is the fraction of the full sphere
contained in the solid angle subtended by the disk, while P measures the
probability of penetration from the disk to the test point. (14-77)
482 Radiative Transfer in Moving Atmospheres /4-2 Sobo/ev Theory 483

where {1 is the planar-atmosphere escape probability that follows from gives the emission from the part of the envelope superposed on the core,
equations {14-70) and {14-63) in the limit that 1/r-> 0, namely the factor <I>{x,) correcting for occultation of material by the core. Note
that, for an expanding atmosphere, <I>(x,) equals zero for x < 0 and will
(14-78) be essentially unity for x > 0, showing immediately the effect of core occul-
tation on the red wing of the profile. The last term gives the continuum
contribution from the core; in view of the properties of<I>(x,)just mentioned,
Exercise /4-7: Verify equations (14-77) and (14-78).
we see that it is unattenuated in the red wing, and_ more or less heavily
attenuated in the blue wing of the line. The flux in the continuum outside
Equation (14-75) may be cast into the standard form for a two-level the line is proportional to
atom by renormalizing the kernel to K*(t) = Kp(t)/(1 - P), and defining
1 - e* = (1 - /J)(l - e) and B*(t) = eB(t)/e* .. Then (14-82)
S(t) = (1 - e*)f:' K*lt' - tlS(.t.')dt' + e*B*(t)
00
{14-79) Transforming the variable of integration from p to ron surfaces(z/r)V(r) = x,
equations (14-81) and (14-82) can be combined to yield an expression for
When thermalization is achieved, S varies slowly and may be removed from
the line profile Rx = <Fx - F,)/F,, namely
under the integral to yield S{t) = B*(t) = eB(t)/(e + /1 - e/1). For e » {1,
S(t)-> B(t), as expected. But for P» e, escapes dominate and S(t) -> eB(t)//3,
showing that S decreases to the local creation rate eB as {J-> 1, which is
Rx = 2(r, 2 /,)- 1 J:. <x> S(r)[t (r)/t(- co,
0 p, x)]{ I - exp[ - r( - co, p, x)] }r dr
reasonable on physical grounds. If the medium has a boundary surface
and B is constant, then [ cf. (406)]
- 2(r,2 I,)- 1 f;• S(r,){ exp[ - t( - oo, p, x)<I>(x,)]

- exp( -t(- oo, p, x)]}p dp


S(O) = (e*}½ B* = eB/(e*)¼ (14-80)

Thus when e » {J we recover the usual static result S{O) = e¼B, while for
- 2r, - 2 f;• {1 - exp[ - t( - oo, p, x)<I>(x,)] }p dp (14-83)
f3 » e, we find S(O) = eBJfJ¼ = fJ¼S 00 , where S 00 denotes the asymptotic where r min(x) is the radius at which V(r) == x, and p is regarded as p(r, x).
value for S at depth. Note the change in sign convention [relative to equation {8-2)] that has been
made to give positive numbers for emission lines. Each term in equation
LINE PROFILES
(14-83) can be interpreted in parallel with terms in equation (14-81).
If we ignore the last two terms, from the core, in equation (14-83), we
Let us now derive expressions for the line profiles seen by an external consider a two-level atom for which the envelope is so thick that the source
observer. The flux emergent at frequency x is proportional to function achieves its asymptotic value S = eB//1, and we replace t( - oo, p, x)
with t 0 , then, in view of equation {14,71), we may write
Fx = 2n f0
"' /(oo, p, x)p dp
R ~ (eBt 0 ) f,"' eBt 0 2r dr _ A(eBt 0 )
{14-84)
= 2n f,~S(r 0 ){ 1 - exp[ - t(- oo, p, x)] }P dp x I, '• (eBt 0 ) r,2 - I,

+ 2nf~• S(r 1 - exp[ -t(- oo_, p, x)<l>(x,)] }p dp


0 ){
where (eB, 0 ) is a typical value of eBt 0 ; here the quantity A denotes the
effective emitting area measured in core units. For sufficiently large effective
+ 2nl, f~• exp[ -t( - oo, p, -~)<I>(x,)]p dp {14-81) emitting areas, the line can become quite bright relative to the continuum.
In fact, most strong emission lines result largely from this geometrical effect.
where, as above, r0 denotes the value of r at the surface of constant radial Another interesting result follows easily from equation (14-81) [see, e.g.,
velocity specified by x, and x, is the value of x' given by equation (14-62) (15, Chap. 28)]. Consider an envelope with constant velocity of expansion V;
at r' = r, and µ' = (1 - (p/r,)2]¼. The first term gives the emission from then Q(r, µ) = (V/r) sin 2 0, and the surfaces of constant radial velocity are
the part of the envelope seen outside the disk (i.e., p > r,). The second term given by cos 0 = (x/V) = constant. The transformation from p to r is
14-2 Sobo/ev Theory 485

p = r sin 0, and in the limit that we can neglect the contribution from the
core,
3

= 2n sin 2 0 f0
00
F" S(r){l - exp[ - x,(r)r/(V sin 2 0)]} r dr (14-85)

F/F, 2 If the envelope is opaque, the exponential term vanishes, and the integral
becomes a constant, so that F" = C sin 2 0 = C[l - (x/V) 2 ]; the line pro-
file in this case is rounded (specifically, it is parabolic). This conclusion is of
importance because it shows that rounded profiles ·occur naturally, as a
result of optical-depth effects, even if the velocity is constant. In contrast,
an interpretation based upon an analysis that assumes the lines are optically
thin would necessarily have involved an accelerating, or decelerating,
0 ....___.._ _.__....__,.____._ _,___ _.___,.____._ __J
velocity field (which would, of course, have quite different dynamical im-
1.0 0.8 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 -0.8 - 1.0
plications). We thus see that the spectroscopic diagnostic procedure must
(a) (c /J.v/v 0 v.,,) be carried out with care, and to a high degree of consistency, if physically
1.5...--.--r---r--.------.--....--r--.------.-- meaningful results are to be derived.
Detailed calculations of flux profiles, using equation (14-83), have been
made (134) using the two-level-atom source function of equation (14-73),
along with assumed distributions of V(r), -r 0 (r), and the constants e and
B/1, (the latter chosen always to have· the numerical value 5). The velocity
F/F, law was taken to be of the form V(r) = V00 (l - r,/r)+. The adopted distri-
butions of -r 0 (r) all are characterized by a maximum on the range 1.1 ~
0.5 (r/r,) ~ 4. If monotone decreasing distributions are used, very asymmetric
profiles (not observed) result. Presumably this indicates that the lines
observed in real stars arise in shell-like zones, produced by variations in the
0 ,_____,__ _.__.....____.__---J.._ _,__...1.-_L..----J..-.....J ionization equilibrium that yield a dominance of a particular ion in a definite
1.0 0.8 0.6 0.4 01 0 -0.2 -0.4 -0.6 -0.8 -1.0 range of radii. A wide variety of profiles can be produced by suitable choices
(b) (c /J.v/v 0 v.,,)
of the parameters. Three characteristic types of profiles, similar to those
10.---.--r---r--.---.--.---r--..----.--
observed in WR stars, are shown in Figure 14-10: (a) rounded emission
with violet absorption, such as observed in the C III l4650 and N III l6438
lines; (b) flat-topped emission with violet absorption, such as seen in the
He I lines; (c) very intense rounded emission with no absorption, such as
5 observed in the He II lines. In each case the intensity of the emission is
F/F,.
proportional to A(eB-r 0 )/J, as expected from equation (14-84); note that the
flat-topped profile results from an optically thin line.

0 .____.__ _.__..._____._____.__ _,___ _.___,.____._ __J MULTILEVEL ATOMS: APPLICATION TO WOLF-RAYET STARS
1.0 0.8 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 -0.8 -1.0
In the spectra of Wolf-Rayet stars, extensive series of extremely strong
(c) (c /J.v/v 0 v.,,)
emission lines can be observed. The spectra fall into two broad classes:
FIOURE 14-10
WC, in which lines of C and O are prominent while those of N seem to be
Calculated line profiles in expanding spherical atmospheres.
(a) t • 0.0092 and t 0 (max) ::::: 15. (b) t z 0.002 and t 0(max) ::::: 0.5. practically absent; and WN, which have prominent Jines ofN and essentially
(c) t =- 0.021 and t 0 (max) "" 2. From (134), by permission. no lines of C. The He II Pickering series (n = 4 -+ n') is very strong and, by
486 Radiative Tra11sfer i11 Moving Atmospheres
/4-2 Sobolev Theory 487
comparison of lines with odd n' (not overlapped by a hydrogen line) and
even n' (blended with a hydrogen Balmer line), it is found that the hydrogen required. To circumvent the need for a detailed solution write J = J • + J d
emission is weak, and hence we conclude that the hydrogen to helium ratio where J: represents the radiation emitted by the stell~r core ;nd /d
is th~
must be significantly less than unity. To derive quantitative information diffuse field from the envelope. Allowing for absorption, we ;dopt •
about these interesting abundance anomalies, as well as about the physical
structure of the envelope, it is necessary to carry out a complete multilevel
(14-89)
analysis of the spectrum. At present it is not possible to specify the atmo- where a representative optical depth between the test-point (R) and the core
spheric structure in detail, and studies (139; 140) have been carried out in (r,) is taken to be
the spirit of a coarse analysis (making a fair number of approximations),
with the goal of obtaining estimates of the physical properties at a single (14-90)
typical point in the envelope.
The statistical equilibrium equations are of the form ~, + f{l 1 = 0 where Fur_ther, if we assume that (a) the envelope is homogeneous, and (b) the
311 and f{l1 are, respectively, the net rates at which level i is populated by optical depth from the test point to the boundary in any direction is t.,
radiative and collisional processes. We have one such equation for each we then may adopt
level of the ion under consideration, plus one additional equation specifying (14-91)
a total abundance for the chemical species. The net collision rate can be
where S,~(v) = (2hv 3/c 2 )[b, exp(hv/kTe) - I]- 1 • With these approximations
written (cf. §5-4)
we have rate equations of the form Jiln = fJJJ where .rd and fJJJ contain rate
((J, =I [n1 - (n1/n I )*n,]C1, + I [(n I/n1)*n1 - ni]C11 + (nt - n1)C1" coefficients, line escape probabilities, and analogous continuum quantities.
J<I }>I To solve the system, the parameters that must be specified are T ,, r r, R ,
(14-86) v~R), _T,, n,, and the total number density "••om of the species under con-
while the net radiative rate is s1derat1on. Most of these quantities can be specified from independent con-
siderations, and typically only Te, n., and notom are free parameters to be
~, = L [n1'A11 + B11111) - n,B,iJIJ] + L [n1B1,J1, - n,(AIJ + B1JJ1,)] ~etern:iined f~om model-fitting. The system of statistical equilibrium equa-
J>I J<I tions •.s nonh?ear beca~se the optical depths t I1 , ,., and the departure
+ nt4n J,.,00
cx1h){hv)- 1 B,(T.)[1 - exp( - hv/kTe)] dv
coefficients b1 m the contmuum terms, depend upon the solution; it is neces-
sa~y, therefore, to solve the system iteratively. The iteration may be effected
(14-87) usmg a Newton-Raphson procedure, which yields swift convergence.
To compute line-strengths we use equation (14-83) and, in the spirit of
coar~e analysis, we assume that SIJ and , 11 are constant for ,. ~ r ~ R,
Here Te denotes the envelope temperature, and nr = nocne<I>,"(Te) [ cf. equa-
and ignore angular factors. Then the three contributions to the line profile
tion (5-14)] where"" is the actual ion density. The expression for~, is useful
are con~t~nt, and are equal to (a) Re .= (R 2/r.2)[S 11 /B,(T.)](1 - e-••J) from
as written, for a given value of Te and ne, but that for ~, must be rewritten.
the em1ss1on component; (b) R 0 = -[SiJ/B.(T.)](1 - e-'1J) from the oc-
To simplify the bound-bound rates we use equation (14-60) to obtain
culted material; and (c) R0 = -(1 - e-••J) from the absorption component.
nj(A 1I + B1,JIJ) - n,BIJJIJ = [n1A 11 - (n IBI1 - n1B1I )WB,(T.)]/JIJ (14-88) Th~ blue half and red half of the line profile each has a width, in wavelength
units, of LU = J.v(R)/c. Thus the equivalent width (taken positive for emis-
where /JIJ is given by equation (14-71), and t I1 (called to there) is given by sion) is
equations (14-67) and (14-68); in equation (14-88) we have used the approx-
imate result /J.::::: W{J, and have parameterized I. in terms of a radiation
temperature T •. The bound-free terms are somewhat harder to reduce,
because the Doppler shifts produced by expansion scarcely affect continuum If occultation and absorption effects are ignored, the two terms involving
formation, and no essential simplification is introduced by intrinsic escapes, "-1" in the above formula are suppressed. Equation (14-92) is only approxi-
as it is for lines; in fact, a solution of the static transfer problem is, in principle, mate, because of the assumption of homogeneity of the envelope, and could
be in error by as much as a factor of two. The above results can be used to
488 Radiative Transfer in Moving Atmospheres 14-2 Sobolev Theory 489

compute intensities and line strengths if SIJ and ti/ (i.e., the level-populations) for u ;;: 10. Because x./kT, « 1 for u » 1, (n./g.) should approach a con-
are known, or, alternatively, may be used in diagnostics to determine these stant value. This, however, is not found for the empirical values just obtained;
quantities. but rather (n./g.) ex: (f A) 4., as expected for optically thin lines because of the
The methodology developed in this section has ·been applied (140) in a factor [l - exp( -t4u)] ~ t 4u oc (f A)4• for t 4u « 1. We thus conclude that
thorough analysis of the He II spectrum of two WN6 stars: HD 192163 and the upper Pickering-series lines are optically thin; this implies an upper bound
HD 191765. As a first step, the total line intensities (which show neither on the parameter A defined in equation (14-97). By imposing the physically
absorption components nor occultation effects) are used to establish the reasonable requirements that (n./g.) (a) be a monotone decreasing function
level populations empirically. For the line (u ➔ 1) the total intensity is of u (i.e., no population inversions), and (b) become constant for u > 10,
we may set limits on A; the results are 3 ;5 A ;5 6 for HD 192163, and
I., oc f A.,hv.,n.(r)P.,(r) dV oc f l'l.,(r)P.,(r) dV 2 ~ A ;5 8 for HD 191765.
As a next step, o~e may use the known absolute magnitude M • to obtain,
oc Js.,(r)[v v(r)/c]{l - exp[ -r,.(r)]} dV (14-93) from standard relat10ns, the absolute continuum flux at )..5500 A; ifwe adopt
0
T, ~ 40,000°K, as indicated observationally, we can then deducer, ~ 13R 0 •
Using the observed ratio F,()..4686)/F,()..5500), we find F,(A4686); then using
where equation (14-71) was used for P.,. and equations (14-67) and (14-68) · the observed value of the emission line strength R.()..4686) we can find R 2 S43 .
for r, •. Therefore, assuming homogeneity, This in turn yields R ~ 70 R 0 because S43 is already known. The ratio
(14-94) (R/r,) ~ 5-6 is in agreement with direct interferometer measures for the
WC star )1 2 Ve!, and also explains the absence of any occultation effects in the
where K is the same.for all lines. For any line that is optically thick, we can profiles. If we adopt v(R) = 1000 km s- 1 from the observed Jinewidths and
set the exponential term to zero; this is expected to be true for )..4686 use the known values of A and R in equation (14-97), we deduce a num~rical
(n = 3 ➔ n = 4). From the observed intensities, the "reduced intensities" value for (n 4 /g.), and hence for all (n./g.), from the empirically determined
(relative to )..4686)J., = (l.,1/v.,4)/(143/v43 4) can be formed. From equation ratios (g 4 n./g.n 4 ). If these values are used in the Saha equation (14-98) for
(14-94), u ~ 10, and if we adopt T, ~ 10 5 °K and set n, = 2n1 (i.e., a completely
ionized atmosphere of helium), we find n, ~ 5 x 10 11 cm 3, which implies
that the

optical depth of the envelope in electron scattering is t I ~ nI aC R =
1.5. Finally, knowing that the upper lines are optically thin, the observed
where we have set exp(-t 34) = 0. Further, we can write excess emission in Pi 14 (which is blended with H 7) relative to Pi 15 can be
used to estimate n(H+)/n(He++) ~ n(H)/n(Her~ 0.5; in these stars the
= A[(n1/g 1) - (n./g.)](g1f,.}..)/(n 4/g4 )
r 1• (14-96) helium-to-hydrogen ratio is thus enhanced by a factor of 20!
Having fixed r., T., R, v(R), and n., the rate equations can be solved
where A = (1te 2 /mc)(n 4/g 4)[R/v(R)] (14-97)
for various assumed values of T, and n(He); the results yield theoretical values
If we now assume that both the )..3203 (5 ➔ 3) and the )..10124 (5 -+ 4) lines of A and (n 1/g1). Calculations (140) were done for a 30-level atom. Good agree-
are also opaque, we may solve for the numerical value of (g 4n3/g 3n4) = ment with the empirical results are obtained for n(He) ~ 2.5 x 10 11 cm - 3
5 54(1 - J 53 )/J 53 , and S43 in each star. The three lines considered thus far and T, ~ 105 °K (the fit is not unique). The calculations show that the upper
are, in fact, opaque, but this need not be true for higher series members which levels are indeed collision dominated, and that b. -+ I for u » l. Further,
have much smaller .f-values. Suppose, however, we presume that all Pickering- because T, > T., it appears that a nonradiative source of energy input is
series line are opaque; then from equation (14-95) we can obtain empirical required to maintain the excitation of the envelope.
values for (g4n./g.n 4). At reasonable values of n, and T,, the upper levels of A second application of the methods described above has been made in an
the He+ ion should become dominated by collisions, and their occupation analysis of the C III line strengths (in the ultraviolet, visible, and infrared)
numbers should have the LTE values observed in the WC 8 star y2 Vet (139). In this work, independent estimates
could be made of the parameters r., T., R, and v(R). The values of T., n.,
(14-98) and n(C+ +) were determined by fitting the observed equivalent widths for
490 Radiative Transfer in Moving Atmospheres /4-3 The Transfer Eq11ation in the Fluid Frame 491

IO lines to theoretical widths obtained from equation (14-92), using occupa- calculations in spherical flows (e.g., pulsation, expansion) can be handled
tion numbers from a model atom consisting of the lowest 14 terms of c+ +, accurately in a Lagrangian coordinate system (i.e., the comoving frame). The
Transitions to higher levels and to the continuum were ignored; this com- Lagrangian equations of gas dynamics are easy to formulate, and offer many
promises somewhat the accuracy of the upper-state populations. Good physical and computational advantages. Obviously it is desirable to be able
agreement with observations was obtained for to treat the radiation field in a closely parallel way. On the other hand, a
disadvantage of the comoving-frame formulation is that present methods
of solution work only for relatively simple velocity fields; otherwise it be-
comes too difficult to pose boundary conditions on the problem (cf. Exercise
[having adopted T, = 30,000°K, (R/r,} = 3.6, and v(R} = 900 km s - 1]. 14-12 below).
The lowest four excited states are found to have a Boltzmannian distribution In the comoving frame we shall develop both (1) _the monochromatic
relative to the ground state, while the higher levels are underpopulated. equation of trw1.~fer, used for calculating, e.g., line profiles, and (2) fre-
Assuming that all the carbon is in the form c+ +, and that all the electrons quency-integrated moment equations, which specify the total radiative contri-
come from hydrogen, a lower bound is found for the ratio n(C)/n(H}-namely, butions to the energy, momentum, and pressure of the gas-plus-radiation
2.5 x 10- 3 • This value exceeds the normal cosmic abundance by a factor of fluid.
8, and suggests that the carbon abundance in WC stars is indeed enhanced, To obtain expressions describing how the relevant physical variables
and that the prominence of the carbon lines in the spectrum is a result of this• change between the rest and the comoving frames, Lorentz transformations
enhancement. Allowing for other ions of carbon and for electrons from other are applied. Here we encounter a problem: strictly speaking, a Lorentz
sources (e.g;, He) would raise the lower limit mentioned above. For y2 Vel, transformation applies only when the velocity v of one frame relative to the
Te < T,, which does not present a compelling argument for nonradiative other is uniform and constant. But in stellar atmospheres we are concerned
energy deposition, and even offers the possibility that the envelope may be with situations where v = v(r, t), and hence the fluid frame is not an inertial
in radiative equilibrium. frame. We must then imagine transfor~a.tions.taking pla~ fro~ uniformly-
moving frames that instantaneously coincide with the moving fluid. Actually,
the difficulty just mentioned introduces considerable complication into the
14-3 The Transfer Equation in the Fluid Frame analysis. It is easy to show that the form of the transfer equation is the same
in two uniformly moving frames {i.e., the transfer equation is covariant),
The formulations of the transfer equation discussed in the preceeding sections providing we account for the effects of Doppler shifts and aberration of
of this chapter are all based in the stationary frame of the observer, who views photons when we calculate atomic properties. Further, it is straightforward to
the stellar material as moving. As we have seen, the complication in this derive the behavior under transformation of the atomic properties them-
approach is that the opacity and emissivity of the material become angle- selves. But for unsteady or steady differential flows, new terms appear in the
dependent, owing to the effects of Doppler shifts and aberration of light. equations that account, in effect, for changes in the 'Lorentz transfor.mation
There results an inextricable coupling between angle and frequency that from one point in the medium to another; these terms can be derived by
presents severe difficulties in the calculation of scattering terms with a discrete application of the different!al operat~r (c- 1 o/ot +. n · ~) in .the transfer
quadrature sum. It then becomes attractive to treat the transfer problem in equation to the transformation coefficient of the specific mtens1ty.
a frame comoving with the fluid.
There are two strong motivations for working in the moving frame of the THE LOCAL FREQUENCY TRANSFORMATION
material. (1) From the point of view of the transfer equation itself, there is
the advantage that both the opacity and the emiss_ivity are isotropic in the Before discussing the details of transformation of the physical variables
comoving frame. Further, in problems involving partial redistribution effects, in the transfer equation, it is worthwhile to extract the essential physical
we may use the standard static redistribution functions. In addition, in the flavor of the problem in simplest possible terms. A velocity field produces a
calculation of scattering integrals we need consider only a frequency band- Doppler shift and aberration of photons, and gives rise to advection terms
width broad enough to contain fully the line profile; this bandwidth is describing the "sweeping up" of radiation by the moving fluid. Formally, these
independent of the fluid velocity. Finally, the angle quadrature may be chosen terms are all of order (v/c). However, in the case of line-profiles, the effect
on the basis of the angular distribution of the radiation alone. (2) Dynamical ofa frequency shift 61' becomes important not when 6v/v = v/c is significant,
492 Radiative Transfer i11 Moving Atmospheres 14-3 The Transfer Equation In the Fluid Frame 493

but rather when 6v/6v 0 = (v/v1h) is significant (i.e., of order unity); in essen~e, solution. We shall discuss a numerical method for solving these equations
velocity-field Doppler effects are amplified a factor of (c/v1h~ by _the swift below.
variation of the line profile with frequency. To a first approx1matton, then, Equations (14-99) and (14-100) contain all the essential physics of the
it is sufficient to consider only Doppler shifts and _to ignore aberration and transfer problem, but there are additional angle-dependent terms that have
advection. been omitted by our neglect of aberration and advection; to obtain these
If v is a frequency seen in the observer's frame, then v0 = v(l - µv/c) is terms (which are important for the fluid equations), we must now develop
the corresponding comoving-frame frequency. The differential operator the transformation properties of the relevant variables.
µ(a/az) in the observer's frame (for a time-independent, planar atmosphere)
is evaluated at constant frequency v; but ifwe move a distance 6z, holding v
fixed, v0 = v0(v, z) will change· because v changes. Thus (a/oz), -+ (0/oz).0 ' . LORENTZ TRANSFORMATION OF THE TRANSFER EQUATION
(fJvo/{)z).(i)/i)v0).0 • Clearly (iJv 0/i)z), = -(v0µ0/c)(ov/az), to order (v/c). Substi- We consider here transformations between the rest system, specified by
tuting into the transfer equation we thus obtain the four coordinates (x 1, x 2, x 3, x 4 ) = (x, y, z, ict), and the fluid system
µ 0[aI 0(z, µ0, v0)/i)z] - [(µ/v 0/c)(iJv/iJz)][aI 0(z, µo, vo)/ovo] (Xo, Yo, Zo, icto), moving relative to the rest system with a constant velocity V
in the z-direction. This choice for v is the one most physically important to
= 11°(z, v0) - x0 (z, v0 )1°(z, µo, Vo) (14-99) our work, and simplifies the calculation; generalizations for an arbitrary
orientation of v are given in (621). Changes from one system to the other
The corresponding result for spherical geometry is are effected by means of a Lorentz transformation, which corresponds to a

1
l
o
ar 0(r, µo, Vo)+ (1 - µ/) oI 0(r, µo, Vo)
or r iJµ 0
-(~oV) [c1 -
er
µo2) + µ02 (d Inv)]
d In r
proper rotation in four-dimensional space-time. Physically the Lorentz trans-
formation is chosen in such a way that the equation for the wavefront of a
light wave is of the same form {covariant) in both systems (i.e., such that the
velocity of light is always =c in both reference frames).
x olo(r, µo, vo)
. . ovo .
= .,,.o(r, vo) - xo(r, vo)lo(r, µo, Vo)· (14-100)
The mathematical form of the transformation is x 0a = L/xP (ix= 1, ... , 4)
where the Einstein convention of summing over repeated indices is employed.
Exercise 14-8: Derive equation (14-100). The transformation can be represented [cf., e.g., (392, 29; 253, 191)] by
the matrix
Several points should be noticed about equations (14-99) and (1~100).
(1) The variables I,µ, and v are no"': all in the comoving fra~e, and m that L= 0
1 0
l
0
O
ol0 (14-101)
frame X and If are isotropic. (2) It 1s clear that any scattering terms n ed 7 0 0 y if3y
be evaluated only on a (small) definite range ofv 0 • (3) The transfer eq uat1on (
0 0 -if3y y
is now a partial differential equation. The frequency derivative accounts
for the change, with position, of a given photon frequency v0 in the comoving =
where y (1 - v2 /c 2 )-t and f3 =
(v/c). Note that L is Hermitian; i.e.,
frame, as seen by an external observer-or, equivalentl~, for the frequency L = Lt where "t" denotes the adjoint (i.e., conjugate transpose). Note also
shift of photons as seen in tlte comoving frame. In_ parllcular, suppose the that L - 1 = LT where "T" denotes the ordinary transpose. In matrix notation,
atmosphere is expanding so that ov/az (or av/or) 1s greater than zero; we x0 = Lx, where x0 and x are column vectors. Clearly x = L - 1 x0 = LTx 0 ;
then see that photons are always systematically redshifted _as t_h~y transfer equivalently, x"- = (L - •)/x/.
from one point in the atmosphere to another, as expected mtutt1vely. Note The Lorentz transformation can be applied to arbitrary four-vectors and
that in planar geometry only velocity gradients matter; in spherical geometry to four-tensors of rank two. The transformation rules of tensor analysis
a net effect occurs even when v(r) = constant, because divergence of the assure that these quantities are covariant under the Lorentz transformation
rays still implies a transverse vel_ocity gradient in this ca~e. (4) From ~ (because it is a proper rotation in four space); hence physical laws written
mathematical point of view, equations (14-99) and (14-100) yield a hyperbolic in terms of four-vectors and four-tensors are automatically covariant. The
system of equations [cf. (181, Chap. 5; 530, Chaps .. ~ and 12; 462, Cha~. _4)], transformation of an arbitrary contravariant four-vector A [ e.g., a space-time
and pose an initial-boundary-value problem requmng boundary cond1tw~s increment (6x, 6y, 6z, ic6t?] is defined such that A 0 a = (ax 0 a/oxfl)Afl =
in the spatial coordinate and initial conditions in frequency to effect their L/ AP, or, in matrix notation, A0 = LA, A = L- 1A0 . The transformation
/4-3 The Transfer Equation in the Fluid Frame 495
494 Radiative Transfer in Moving Atmospheres
Here n is the direction of photon propagation, and (n", ny, n,) = (sin 0 cos <P,
of a covariant four-vector B (e.g., the gradient operator [8/8x, 8/oy, 8/oz, sin 0 sin</), cos 0). Applying a Lorentz transformation to equation (14-105)
~ic)- 1 a~at]T)is ~efin~d sue~ that (B 0 ). = (oxP /ox 0 ")Bp = (L - 1) / Bp-which, we find
m matrix notation, 1s equivalent to B0 T = BTL - 1 or, transposing B =
(L - i )TB = LB; also, B = L- 1B0 • Finally, the transformation of a :on- C"P: (v 0 nx°, v0 n/, v0 n,°, iv 0 ) = [vnx, vny, vy(n, - P), ivy(l - n,.8)] (14-106)
travariant tensorofrank two, is defined such that c0•P = L "L/Od, which(in
matrix notation) is equivalent to C0 = LCLT = LCL - 1 Y(i.e., the transfor- which is equivalent to
mation is a similarity transformation); also we have C = LTC 0 L = L - 1C0 L.
(NoTE: The word covariant describing the four-vector B is used in a [<J>o; (1 - µ/)½; µo; Vo]
different sense from that used two sentences earlier in relation to physical = [qi;y-1(1 - µ 2 )½/(1 - µP);(µ - .8)/(1 - µ/J);vy(l - µfJ)] (14-107)
laws, where it meant "of the same form"; this double meaning, however
deplorable, is standard usage.) The inverse transformation gives
Equipped with the transformation rules given above, we can now derive [(1 - µ 2 )½; µ; v] = [y- 1(1 - µ/)½ /(1 + µ 0 /J);(µo + /J)/(1 + µo/J); voy(l + µo.B)]
a number of important results. First, applying the transformation to the
coordinates themselves it is easy to show that the measurement of intervals (14-108)
perpendicular to the z-axis is unaffected by the relative motion of the two
frames-i.e., .6.x0 = /J,.x and .6.y0 = .6.y. But an object of length .6.z at rest Equations (14-106) through (14-108) contain the familiar results for Doppler-
in the fixed frame will be measured by an observer in the moving frame to shift and aberration; the classical results can be derived by keeping only
have a length terms ofO(v/c), (i.e., by setting y = 1). From equations (14-107) and (14-108),
2
.6.zo = y-1 .6.z (14-102a) it is easily seen that dv 0 = (v 0 /v) dv and dµ 0 = (v/v 0 ) dµ; hence, using
dw = sin 8 dB d</) = dµ d</J, we find
This result is the celebrated Lorentz-Fitzgerald contraction effect. Likewise a v dv dw = v0 dv 0 dw 0 (14-109)
time-interval .6.t measured in the fixe'd frame will be measured by an observer
in the moving frame to be To establish the transformation properties of the specific intensity we
(14-102b) calculate (621) the number of photons passing through an area dS oriented I
perpendicular to the z-axis, in frequency interval dv, into solid angle dw, (
~his is the time-dilation effect. From these results we conclude that a space- propagating at angle 8 = cos- 1 µ to the z-axis, in a time dt. We suppose (
time volume element is invariant; i.e., that dS is stationary in the rest frame; then N = [/(11, v)/hv] dw dv dS cos 0 dt.
This must be the same number that would be counted passing through the
dV dt = dV0 dt 0 (14-103)
same surface element by an observer in the comoving frame, namely
a result we shall use repeatedly below. Next, applying the Lorentz trans- N = [1°(µ 0 , v0 )/hv 0 ] dw 0 dv 0 x (dS cos 00 dt 0 + c- 1 dS v dt 0 ); here the first
O
formation to the four-gradient (a covariant vector), we obtain term gives the number that would have been counted if dS had been stationary
0
in the comoving frame, and the second gives the density of photons (I /hv 0 c)

~ a~J. i (a~
times the volume (dS v dt 0 ) swept out by dS in a time dt 0 • Using equations
(ct. :y' :z' l :t) = [a! 0• a~0 ' Y (a! 0 - 0 - c/J a!J] (14-102), (14-108), and (14-109), we find
(14-110)
(14-104)

Let us now turn to the transformation of the radiation field and transfer- Next consider the emissivity. The number of photons emitted from a given
volum;, into a specified solid angle and frequency-interval, in a definite
related v~riables [further discussi~n can be found in (521; 551; 621)]. For
~ny particle, the four-moi:nentum 1s P" = (p,,, Py, p,, iE/cf, where p1 is the time interval, must be the same in both frames, hence
1th component of the ordinary momentum, and E is the total energy of the (l'/v dw dv dV dt)/hv = (11.0 dw 0 dvo dVo dt 0 )/hvo,
particle. If the particle has rest mass m0 , then p 2 c2 + (m 0 c2 ) 2 = E2 where
P2_ - Px 2 + P, 2 + p, 2ph
. otons have m0 = 0 and E = hv, hence p = hv/c, ' and which implies that (14-111)
P" = (hv/c)(n,,, ny, n., W = (hv/c)(n, W (14-105)
496 Radiative Transfer in Moving Atmospheres 14-3 The Transfer Equation in the Fluid Frame 491

(where we have noted explicitly that I/ is isotropic in the fluid frame). To be where the changes in '7, and X,, are so swift that a first-order expansion is not
able to establish energy balance in any frame, we must be able to balance sufficient. A complication in application of equation (14-115) is that both
emission losses with absorptions; hence, from equations (14-110) and even and odd terms in µ appear on the righthand side, so that the usual
(14-111), we conclude that reduction to a second-order transfer equation is not possible; a satisfactory
numerical scheme can, nevertheless, be devised using a staggered mesh
(14-112) [see, e.g., (460)].
Finally, between two frames moving uniformly with respect to one another,
the differential operator in the transfer equation transforms, using equation TRANSFORMATION OF MOMENTS OF THE RADIATION FIELD
(14-104) and (14-106), to The total energy density, energy flux, and radiation pressure in the field
(14-113) are specified by the frequency-integrated quantities [cf.equations (1-7), (1-19),
(1-28)]
We can now show that the transfer equation is, in fact, covariant, for we see
from equations (14-110) through (14-113) that
ER(r, t) = c- 1
00
f p
dv dw /(r, n, v, t)
0
(14-116a)

(14-114a) .'!F(r, t) = f0 00
pdw n, v, t)n
dv /(r, (14-116b)
transforms to

(v 0/v)[c- 1(o/ot 0) + (n° · V0)][(v/vo)3 I0(µo, Vo)]


and P(r, t) s c- 1 J dv pdw
0
00
/(r, n, v, t)nn (14-116c)

= (v/vo) 2[11°(v0) - x 0(vo)I 0(µo, vo)] These are related by the frequency-integrated moment equations [ cf. equa-
tions (2-75) and (2-68)]
which, if (v/v 0) is a constant-as it will be if the two frames are in uniform
motion with respect to one another (and only then)-can be written as c- 2(o.'!F/ot) + V · P = c- 1 S0"" dv9'dw[11(r, v, t) - x(r, v, t)/(r, v, t)]n
1
[c- (0/0to) +(n°' V 0
)]/
0
(µo, Vo) = 1/0(Vo) - x0(vo)/ 0(µo, Vo) (14-114b) (14-117a)
Obviously equations (14-114a) and (14-114b) are of the same form. Two and
points must be stressed. (a) Despite the similar form of the two equations,
equation (14-114b) (at rest relative to the fluid) is actually much simpler
(fJER/ot) + V · fF = S0"" dv pdw[17(r, v, t) - x(r, v, t)/(r, v, t)] (14-l 17b)
because of the isotropy of ,, 0 (v 0 ) and x0 (v 0 ). (b) The reduction of equation As written, equations (14-117) are already covariant. One way this may
(14-114a) to equation (14-114b) is not valid if the two frames do not move easily be seen is to exploit the results of electromagnetic theory in which one
uniformly with respect to each other; i.e., this equation does not apply in, finds [see, e.g., (331, Chap. 12; 386, Chap. 4; or 494, Chap. 21)] that the
say, an expanding or pulsating atmosphere. .. momentum-energy conservation laws of electrodynamics can be written in
One approach to coping with the shortcoming mentioned above is to leave the covariant form
the streaming terms described by the differential operator, and the radiation (14-118)
field itself, in the observer's frame, but to use first-order expansions to write
the source-sink terms in the comoving frame (S21, Chap. 6). Thus, using Here T"fl is the stress-energy tensor of the electromagnetic field, and f" is
equations (14-111), (14-112), and (14-107) we can write a four-vector giving, per unit volume, the three components of the Lorentz
forces of the field on charged matter, and (i/c) times the rate of work done
[c- 1(O/ot) + (n' V)]/(µ, v) = 11°(v) - x0(v)I(µ, v) by the field on charges and currents. The stress-energy tensor is a contra-
+ (n · v/c){217°(v) - v(o11°/ov) + [x0 (v) + v(ox 0 /ov)]I(µ, v)} (14-115) variant four-tensor of rank two, namely

This approach may be adequate for continua, where the frequency derivatives -TM
icG) (14-119)
are well defined and nearly constant, but will not be accurate enough for lines T = ( icG -W
498 Radiative Transfer in Moving Atmospheres
14-3 The Transfer Equation in the Fluid Frame 499

where TM is the Maxwell stress tensor (cf. §1-5), G is the momentum density, it follows that
and W is the energy density of the field. H0 = H - (J(J + K)
We have seen in Chapter 1 that a one-to-one correspondence between (14-123b)
electromagnetic field quantities and moments of the radiation field can be and K 0
= K - 2{3H (14-123c)
made; specifically, TM = -P, G = c- 2:g;, and W = ER. We thus conclude
that the stress-energy tensor of the radiation field must be The inverse transformations yield

R =(
p
1
ic-1/F)
=c
_1 (f dv pdw Inn if dv pdw In)
(14-120)
(J, H, K) = [1° + 2(JH 0, H 0 + (J(J 0 + K 0 ), K 0 + 2(JH0 ]
Notice that these transformations are valid only for the. frequency-integrated
(14-124)

ic- :F . -ER if dvpdw In - f dvpdw I moments.

We can see by inspection that R is actually a four-tensor merely by noting Exercise /4-?: Derive equations (14-123) and (14-124) by applying a Lorentz
that it is formed from the outer product of the four-vector (vn, iv) [recall transformation to the stress-energy tensor [equation (14-120)) and expanding the
equation (14-105)] with itself, times the invariant [see equations (14-109) and result to O(v/c).
(14-110)] (J dv dw/v 2 ), integrated over all angles and frequencies. Similarly,
the Lorentz force can be identified with the rate at which momentum is THE COMOVING•FRAME EQUATION OF TRANSFER
transferred from radiation to matter, and the rate of work with the rate of
energy input of the radiation into the matter. Thus we conclude that we can . The full transformation of the equation of transfer (including the differ-
write ential_opera~or) for_a ?onuniform velocity field can be done rigorously using

gR = c- 1 [fdv p dw (xI - ,i)n, if dv p dw (xI, - ,i)] (14-121)


cova~1ant d1ffere~t.1at_1on (399; 135; 278); however this approach requires
considerable fam1hanty with tensor calculus. We shall instead use a simple
~rst-order expansion method that yields results correct to O(v/c). The equa-
as a four-vector. This can again be seen to be the case, for gR is composed of t10ns o_f transfer that we consider are (in planar and spherical geometry
respectively) ' I
the four-vector (vn, iv) times the invariants (xI dv dw)/v or (,i dv dw)/v. Tbe I\)
four-diver-gence of a four-tensor is automatically a four-vector, hence 1 01
[c- (8/8t) + µ(8/8r)][(v/v 0 ) 3 J 0 (r0 , µ 0 , v0 , r0 )]
0)
(14-122) = (v/vo) 2 [,i 0 (vo) - x0 (vo)I 0(ro, µ 0 , v0 , t 0 )] (14-125a) I
1
is, in fact, a covariant representation of the conservation relations for the and [c- (8/8t) + Jt(8/8r) + ,- 1(1 - µ 2 )(8/8µ)] [(v/v0 ) 3 J 0(r0 , µ 0 , v , to)] 0
radiation field. One can see immediately that equation (14-122) produces
equations (14-ll 7a) and (14-117b), when the latter is multiplied by ( - ic), and = (v/vo) 2 [,i 0 (vo) - x0 (vo)I 0 (r 0 , µ 0 , v0 , t 0 )] (14-125b)
that these equations are, therefore, covariant between frames moving uni-
'V4!e consider one-dimensional flows, and. apply a local Lorentz transforma-
formly with respect to one another; for nonuniform motion, additional terms !1on to a frame that instantaneously coincides with the moving fluid. We shall
will appear in the comoving-frame moment equations. ignore terms of O(v 2/c 2 ), and set y = 1. We then have
In working with the equations of radiation hydrodynamics (§15-3), it will
be useful to have transformations accurate to O(v/c) of ER, :F, and P (or, r0 =r (14-126a)
equivalently, J, H, and K) between the fixed and comoving frames. These
are. most easily obtained by using equations (14-109), (14-110), and (14-107),
and expanding to first order in (v/c). Thus, setting')' = I, we find readily that
cto(r, t) = ct - c- 1 J: v(r', t) dr' (14-126b)

I. 0 dv 0 dw 0 = (v 0 /v) 2 Iv dv dw :=:: (1 - 2/Jµ)l v dv dw, from which it follows Equation (14-126a) states that space increments will be judged to be the same
that by observers in both _frames (i.e., ~o Lorent contraction). Equation (14-126b)
J 0 = J - 2/JH (14-123a)
7
acco~nts for the fimte propagation velocity of light and accounts for the
classical retardation effect. [The significance of this term can be understood
Similarly, J. 0 µ 0 dv 0 dw 0 = (µ - /J)(l - µfJ)I,.dvdw :=:: [µ - /1(1 + µ 2 )]1.dvdw, more fully by the following thought-experiment. Suppose clocks at a number
and I.,°µ/ clv 0 dw 0 = (µ - /1) 2 1, dv dw ~ (µ 2 - 2µ/J)l, dv dw, from which of stations around some point r 1 in the rest frame are initially synchronized.
500 Radiative Transfer in Moving Atmospheres 14-3 The Transfer Equation in the Fluid Frame 501

Then suppose that at some time t I all these stations simultaneously emit St~rting with equation (14-125a), rewritten with the approximation
pulses oflight received by (a) an observer at rest at r 1 , and (b) an observer at mentioned above as
r I moving with velocity v. Observer (a) will receive the pulses synchronously,
while observer (b) receives pulses from directions r > r 1 sooner than from 1
{(v/vo)[c- (o/ot) + µ(o/or)] + 3µ[o(v/Vo)/or]}JO(r 0 , µ0 , Vo, to)
r < r 1• He thus concludes that.the emission times (t 0 ) were earlier for r > r 1
than for r < r 11 as is shown by equation (14-126b) (for simplicity consider = '1°(vo) - x0(vo)I 0 (ro, µo, v0 , t 0 ),
the case of v = constant}.] The retardation effect is, in one sense, not a
relativistic tenn, but merely acknowledges a finite light-propagation speed; subst~tuting from equations (14-127) and (14-128) and the first-order ex-
its significance from a relativistic viewpoint is, of course, profound. pressions_ for (v/v 0), etc., and retaining only tenns of first-order in (v/c},
one obtains:
To evaluate the derivatives in equations (14-125) the chain rule is applied.

(ara) = (a)ar ,,., = (aro)


Tr ,,., ora + (aµo)
Tr ,,., aµa
0 0
[!~
C Oto
+ (µo + ~) ~ + µo(µo
C Oro
2

C
- 1) (!!:) ~
Oro oµo
2

-avo) a
+ ( or ,,., -+
OVo
(oto) o
-ar ,,., -Oto (14-127a) - Vo:o (:~)
0~0 + :a2 (:r:)] I (ro, µo, v
3 0
t
0, 0)

= 11°(vo) - x0(vo)I 0(ro, µ0 , v0 , t0 ) (14-129)

For spherical geometry we have the additional term

+ ( -ova) a
-+
oµ ,., OVo
(ot o
-oµ0 ) ,~, -Oto (l4-127b) (vo/v) 2 r- 1(1 - µ 2 )(o/oµ)[(v/vo) 3 JO]
= (v/vo)r- 1(1 - µ 2 )(o/ 0 /oµ) + 3r- 1(1 - µ 2 )[o(v/v0 )/oµ]JO
(ato) = (a) (ar 0 )
a o
at ,,,. = Tt ,,,. aro + Tt ,,,. oµo
(oµ 0 )
Again expandi~g to fir:t o~der in (v/c), and using the result from equation
I
I\)
(14-1_07) that v (1 -. ~ ) = v0 (1 - µ 0 2 ), and also [o(v/vo)/oµ] = (v/c), we 01
2

+ ( -ova) -+ a
at ,,,. avo
(at o
-ot0 ) ,,,. --
ato
(14-127c) obtain for these add1t1onal terms co
I
2
We use the first-order expressions (v/v 0 } = 1 + f1µ 0 , (v 0 /v} = 1 - flµ, ro -l(l - µo ){[(1 + /Jµo)(o/oµo) - /Jvo(a/ovo)] + 3/J}JO
µo = (µ - /1)(1 - fJµr 1, and µ = (µo + /J}(l + /Jµo)- 1, and make the
additional approximation that fluid accelerations (which are identically Thus the _comoving-frame transfer equation to order O(v/c) in spherical
geometry 1s •
zero for steady flow} are so small that the change in any velocity during the
time of a photon flight, over a mean-free-path, is negligible compared to
the velocity itself. We then neglect (av/at} and derivatives of the fonn (ax 0 /ot}
for x 0 = r 0 ,µ 0 ,orv 0 ,andretainonly(ot0 /ot) = 1. Theremainingcoefficients
in equations (14-127), to O(v/c}, are easily derived from equations (14-126)
and the first-order relations written above. One finds

(aO\ r },..,
(ro, µo, Vo, to} = [1, C- 1(µo 2 - 1)(av/Oro), - C- I µO Vo(OV/Oro), - /J/c]

(14-128a)

( :µ),., (ro, µo, Vo, to} = [O, (1 + 2µ 0 /1), -v0 {J, O] (14-128b)
Exercise 14-10: Verify equations (14-129) and (14-130) in detail.
502 Radiative Transfer in Moving Atmospheres
14-3 The Tran~/i'r Equation in the Fluid Frame 503
Equations (14-129) and (14-130) are the comoving-frame equations in-
cluding all terms ofO(v/c), and were first derived consistently by Castor (135). flow, it is sufficient to replace equations (14-13la) and (14-13lb) with
The time-derivative written in these equations is, in essence, still in the fixed 2 2
frame (though it allows for retardation); the Lagrangian time derivative,
,- [fJ(r H/)/fJr] - a[o(J. 0 - K.°)/ov 0 + b(oK.°/ov 0 )] = 11°(v 0 ) - x0 (v0 )J. 0
which follows the motion of a fluid element (cf. §15-1) consists of the two (14-133a)
terms (D/Dt) = (o/ot) + (v/c)(fJ/fJr), the second term being the advection 0 1
term. For steady flows, the terms in (fJ/fJt) are identically zero. The equations and (iJK,. /or) + ,.- (3K,.° - J, 0 ) - a[o(H.° - N,. 0 )/ov0 + b(fJN.°/ov0 )]
just derived are obviously fairly complicated. As we shall see, it is important = - x0 (v0 )H.° (14-133b)
to retain all of the terms in the moment equations. However, for a solution
of the transfer equation itself, it is sufficient for most astrophysical flows where a = (v 0 v/cr) and b = (d In v/d In r) [see (447)]. ·
[as verified by a detailed calculation (446)] to retain only the frequency- For problems of radiation hydrodynamics we require the frequency-
derivative terms because of the effective amplification of these terms by integrated moment equations, which follow immediately from equations
(c/v,h), as discussed earlier. In this limit, equation (14-129) for steady flow (14-131):
reduces to equation (14-99), while equation (14-130) reduces to (14-100). We
shall describe a numerical method for solving equation (14-100) in the next (fJER 0 /ot) + t•(fJER 0 /or) + ,·- 2 [fJ(r 2 ff 0 )/ar]
subsection. . + (v/r)(3ER 0 - PR 0 ) + (ov/or)(ER 0 + PR 0 )
For problems of radiative transfer involving partial redistribution, it is
useful to derive frequency-dependent moment equations from equations
(14-129) and (14-130). For spherical geometry we obtain
= 4n f0
00
[17°(v 0 ) - x0 (v0 )J,.°] dv 0 (14-134a)
2 0
C- (fJff /fJt) + (v/c 2 )(fJ§ 0 /or) -f- (fJpR 0 /fJr) + (3pRO - £R 0 )/r

! fJJ.° + ~ fJJ.°
C Ot C or
+ ..!_2 fJ(r'-H~) + (~)(3J o - K.°) +
r or er •
! (J:o +
C
K.°) ~
' or
+ (2v/c 2 r)[l + (d In v/d In r)]§ 0 = -c-• f 0
00
x0 (v0 )ff,.° dv 0 (14-134b)
I
+ (~) J_ [v (JK O _ j 0 )] _ ! (2v + ov) o(v 0 K.°)
0
J
where ER = (4n/c) J 0 (v 0 ) dv 0 =
(4n/c)J 0 and PRo (41t/c)K 0 • In planar = I\)
er OVo O v v C r or OVo , geometry, equations completely analogous to (14-131) through (14-134) can CJ)
also be written down. Equations (14-134) contain additional velocity- 0
= 11°(v0 ) - x (v 0 )J.°
0
(14-131a) dependent terms on the lefthand side compared to their counterparts [cf. I
equations (14-117)] in the fixed frame. But this is more than adequately
for the zero-order moment, and for the first-order moment compensated by the tremendous simplification of the righthand side, where
the isotropy of the opacity and emissivity in the comoving frame allows the
integrals to be expressed in terms of the moments themselves, instead of as
double integrals over the specific inte11sity. The only question remaining to
be faced is: "How do we actually solve the comoving-frame transfer equation
for J0(r, Jlo, 1•0 ) and its moments?"

(14-13lb) SOLUTION FOR SPHERICALLY SYMMETRIC FLOWS

Let us now consider methods for solving the comoving-frame transfer


where (14-132) equations in the limit that only Doppler shifts are taken into account-i.e.,
equations of the form of (14-99) and (14-100). As was mentioned earlier,
and, for brevity, we have written J .° = J 0 (r, v0 , t), etc., and suppressed the the transfer equation in the comoving frame is a partial differential equation,
suffix "O" on r and t. Again, for practical transfer calculations with steady which was first obtained by McCrea and Mitra (414). Chandrasekhar (156;
157) obtained solutions of these equations for planar geometry, with a linear
504 Radiative Transfer in Moving Atmospheres 14-3 The Transfer Equation in the Fluid Frame 505

velocity law, a two-stream description of the radiation field, and strictly where y(z, p, v) = y(z, p)/x(z, p, v), and the source function is assumed to have
coherent scattering; generalizations of this approach are given in (l; 2; 3; the form for an equivalent-two-level-atom with complete redistribution-Le.,
276, 199). Lucy (403) solved the equations in planar geometry, with coherent S(z, p, v) = S[r(z, p), v] = cx(r, v)J(r) + f3(r). The coefficients ex and f3 contain
scattering, in the high-velocity limit, by ignoring the spatial derivative and the thermalization parameter e, the opacity ratio x,/x,. and the profile
treating the equation as an ordinary differential equation in frequency alone. function, while
An integral-equation method for planar geometries has also been devised
(574); this method is restricted to linear velocity laws and hence is not useful (14-141)
for realistic models. A general and flexible n\Jmerical method developed by
Noerdlinger and Rybicki (478) solves the equations in planar geometry using In equation (14-141), vmin and Vmax are chosen to contain the whole line
a Feautrier-type elimination scheme; this method can treat problems in- profile as seen in the comoving frame. Note particularly in equations (14-137)
volving partial redistribution. In a ray-by-ray solution for spherical geometry, and (14-138) that, because we are working in the comoving frame, we can now
then umber of angles must be of the same order as the number of depth-points, average r and r at a given value of v, in contrast to the situation in an
and a Feautrier-type solution becomes costly. If complete redistribution is observer's-frame formulation [cf. equations (14-23) and (14-24)].
assumed, we can construct an efficient Rybicki-type solution (444), as will Spatial boundary conditions are now required. At the outer radius. r = R
be described here; for partial redistribution we use moment equations [e.g,, r = O; therefore u = v, hence
equations (14-133)], thereby eliminating the angle-variable, in which case
a Feautrier-type solution again becomes practical (447). [ou(z, P, v)/Ot(z, p, v)]Zmo, + y(zmax• p, v)[OU(Zmax, P, V)/0\1] = U(Zmaxt p, V)
In spherical geometry we adopt the (p, z) coordinate system introduced in (14-142)
§7-6. Then, along a ray specified by constant p, equation (14-100) becomes
At the plane of symmetry z = 0, we can now write v(O, p, v) = 0, hence for
± [aJ±(z, p, v)/oz] - y(z, p)[of±(z, p, v)/ov] = 17(r, v) - x(r, v)[±(z, p, v) rays that do not intersect the core,
(14-135) [ou(z, p, v)/ot(z, p, v)], .. 0 =0 I
(14-143)
N
where y(z, p) = [vv(r)/cr][(1 - µ 2) +µ 2
(d In v/d In r)] (14-136) For rays that intersect the core (i.e., p ~ r,) we (a) apply the diffusion 0)
and r = (p.2 + z2 )½,
µ =
(z/r). In equations (14-135) and (14-136) we have approximation for an opaque core (stellar surface), which specifies v, or
(b) set v = 0 (by symmetry) for a hollow core (nebular case).
suppressed the suffix "0" for ,notational simplicity (and continue to do so
In addition, an initial condition in frequei1cy is required. For an expanding
henceforth in this chapter), but it is to be stressed that all quantities are
evaluated in the comoving frame. Now introducing the optical depth along atmosphere [i.e., one in which v > 0 and (dv/dr) > OJ, it is obvious that
the high-frequency edge of the line profile (in the comoving frame) cannot
the ray, dt(z, p, v) = - x(z, p, v) dz, and the variables
intercept line photons from any other point in the atmosphere, because
they will all be systematically red-shifted; any photon incident at the high-
_1[+
u(z, p, v) = / (z, p, v) + I - (z, p, v)] (14-137)
2 frequency edge must be a continuum photon. To specify the required initial
condition we may therefore either (a) solve equations (14-139) and (14-140)
in the continuum, omitting the frequency-derivative terms (which yields the
and v(z, p, v) = 21 [ / + (z, p, v) - I - (z, p, v) ] (14-138) standard second-order system) to obtain u(z, p, Ymax) = Ucontlnuum• or (b) fix
the derivative (ou/ov),m .. to any prespecified value given by the slope of the
we can obtain from equation (14-135) the system continuum; in particular, the choice (ou/ov) = 0 leads to equations identical
to option (a) just mentioned.
[cu(z, p, v)/ot(z, p, v)] + y(z, p, v)[ov(z, p, v)/ov] = v(z, p, v) (14-139) The system is now discretized using the same grids {r4 }, {p1}, {z4,} as
were employed in §7-6 and §14-1. We now choose the frequency grid {vn}
and [cv(z, p, v)/cr(z, p, v)] + y(z, p, v)[cu(z, p, v)/ov] = u(z, p, v) - S(z, p, v)
(n = 1, ... , N) in order of decreasing values (v 1 > v2 > · · · > vN) because
(14-140) the initial condition is posed at the highest frequency. We replace equation
506 Radiative Transfer in Moving A1mospheres 14-3 The Transfer Equalion in /he Fluid Frame 507

(14-141) with a quadrature sum where G is bidiagonal and His diagonal. Equations (14-151) can be used to
eliminate vd±t, in from (14-146); we then obtain a set of second-order
N 14
equations for udin• namely
J(r4) =L Wn L a4,cp(r4, vn)u[z(rd, P1), p,, vn] (14-144)
n=I 1•'1
{(ud+l,ln - Ud1n)/[Md+½,in(1 + c5d+½,l,n-t)J
Equations (14-139) and (14-140) are replaced with difference approximations
- (udin - Ud-1,in)/[£\td-t,ln(l + c5d-t,l,n-t)J}/£\td1n
= (1 + c5dl,n-t)Udtn - Sdln - c5dl,n-t Udl,n-1
(14-145) + [c5d-t,l,n-t (1 + c5d-½,l,n-~)-!Vd-t,l,n-1
- c5d+½,l,n-t (1 + c5d+t,l,n-t)- 1Vd+½,l,n-1]/£\t41.
(14-146) (14-154)

where u is presumed to be defined on the mesh-points zd = z(r4, p1), and Adding the boundary conditions to equation (14-154) we obtain the system
udin = u(zd, p1, vn), while v is presumed to be defined on the interstices
zd± ¼ = ½(zd + z4 ± i) and vd±t, In = v(zd± t, p1, v.). Further, we have defined (14-155)

(14-147) where T 1n is tridiagonal, V,n and w,. are diagonal, v,. is bidiagonal, and
X 1• is a vector.
(14-148)
Exercise 14-11: Verify equations (14-151) and (14-154), and sketch the forms of
(14-149) the G, U, and V matrices.
I
N
and (14-150) CJ)
To solve the complete system, we choose a definite ray, specified by a N
Similar difference equations may be written to represent the boundary given p,. and carry out a frequency-by-frequency integration procedure,
I
conditions (444). In equations (14-145) and (14-146), an implicit frequency with n ranging from 1 to N. This is effected by noting that the initial condition
differencing is used to assure stability (444; 462; 530). in frequency implies that U 11 , Vu and H11 are all exactly zero; thus we can
Equation (14-145) can be solved analytically for vd+t, In to yield obtain expressions of the form u11 = A11 - B11 J and v11 = Cu - D 11 J,
where A, 1 = Tjj 1X 11 is a vector, 8 11 = Tjj 1W 11 is a matrix, C11 = G11 A11 ,
and D 11 = G11 B11 . Similar substitutions are carried out for successive values
of n, to yield
(14-151)
(14-156)
Organizing the solution into vectors that specify the depth-variation along
a particular ray at a given frequency-Le., and (14-157)
(14-152a) where A,n = T,;'(X1. - v,.A1,n-l - v,.c1,n-d (14-158)

and (14-152b) B,. = T,; 1(W,n - u,.B1, n-1 - V1.D,, n- i) (14-159)

equation (14-151) can be written in the form (14-160)

(14-153) and (14-161)


508 Radiative Transfer in Moving Atmospheres 509

Each result, of the form of equation (14-156), for every frequency v., along 0
every ray pi, is substituted into equation (14-144) to obtain a final system
for J of the form
(14-162) -0.4

where the F's contain the quadrature weights. The solution of this final
-0.8
system yields J, and hence S(r, v), and u(z, p, v) and v(z, p, v) from equations
(14-156) and (14-157). Knowledge ofu(z, p, v) implies knowledge ofu(r, µ, v),
so it is clear that we can calculate J 0 (r, v) and K 0 (r, v) in the comoving frame;
similarly we can calculate the flux H 0 (r, v) from v(r, µ, v). Thus we obtain a -1.2
complete solution for the radiation field and its moments in the comoving
frame.
The number of operations required to obtain A,. and 8 1• in equation -1.6
(14-156) is proportional to D12, so summing over all frequencies on all rays
logS
one obtains Ts = cN D3 + c' D3 as an estimate of the computing time
required. Note that this time is linear in the number of frequencies. A similar -2.0
formulation can be written down for planar geometry; in this case we dispense
with the rays and use M fixed angles {µ 1}. In planar geometry the computing
time required for the solution is T,. = cNM D2 + c' D3 • For partial redis-
tribution in spherical geometry one would use a Feautrier solution of the
moment equations (447), obtaining the Eddington factors from a· ray-by-ray
formal solution with a given estimate of the source function. For partial I
redistribution in planar geometry the method of Noerdlinger and Rybicki N
(478) is applicable. 0)
w
Exercise 14-12: .Consider flows with monotonic velocity fields [i.e., (dv/dr) every- I
where ;a,O or everywhere :!6;0). (a) In planar geometry show that the choice of
the initial condition in frequency is unique and depends only on the sign of (dv/dr)
[cf. (451) for a discussion of non-monotonic velocity fields). (b) In spherical -2.0 -1.0 0 1.0 2.0 3.0
geometry show that unique conditions can be found only if [v > 0, (dv/dr) > OJ or log t
[v < O, (dv/clr) < O], and that, owing to projection effects along a ray at the plane
FIGURE 14-11
of symmetry (z = 0), velocity distributions of the form [v > 0, (c/v/clr) < O] or
Linc source functions in expanding atmospheres for various values of outer radius R
[v < O, (dv/dr) > O], though monotone in the radial direction, produce non- (in units of r,), and terminal velocity vm., (in thermal velocity units). For all models
monotonic fields along tangent rays. T, = 2, T, = 103, B = I, and e = 2 x 10" 3 • Abscissa: log of static line optical
depth. The dashed line gives the mean intensity in the continuum. The curves are
The method outlined above has been used to calculate source functions labeled with vm., = v(R). From (444), by permission.
and line profiles in idealized model atmospheres (444). Each atmosphere is
characterized by an outer radius R (in units of r, = 1), a continuum optical velocity at larger. Form (b) is a caricature of stellar-wind solutions. Results
depth T, (at r = r,), a static line optical depth T,. opacities Xi r:t:. ,- 2 and for the source function in several models are shown in Figure 14-11; for
Xe r:t:. ,- 2, and constant Planck function B = 1. The two-level-atom thermal- these models T 1 = 103, T, = 2, rv = (R + 1)/2, B = 1,ande =2 x 10- 3 ,
ization parameter was set to & = 2/T,. The velocity field was chosen to be We see that the basic effect of the velocity gradient at depth is to increase
either (a) dv/dt = constant, or (b) v(r) = v0 [tan - 1(ar + b) - tan - i (a + b)], the escape probability; hence the source function drops below its static
which gives a sharp rise at the point rv = -(b/a), and has constant terminal value and, for large values of vm .. , S 1 approaches J,, the mean intensity
510

1.3 ,----,---r---,----r---,----,.--,--,----,---,

1.2

1.1
15
1.0

F 0.9 Stellar Winds


0.8

0.7

0.6

-60 -20 +20 +60 +100


X
FIGURE 14-12
Linc profiles (emergent fluxes integrated over the disk) from
expanding spherical atmospheres. Ordinate: flux relative to
continuum flux. Abscissa: llv/llv 0 • For all models, R = 300,
T, • 2, B • 1, em 2/T£,and(dv/dt 1) = -100/TL. From
(444), by permission.

I
in the continuum (dashed line). Near the surface, the enhanced esc~pe I\)
probability competes with interception of continuum radiation by ,he 0)
Doppler-shifted line. In planar geometries the latter effect dominates, and The outermost atmospheric layers of many stars are in a state of continuous ~
produces an increase in S1; in very extended atmospheres the former effect rapid expansion, and the material lost from a star in such a flow is called a I
dominates, and S1 decreases. Flux profiles, as seen by an external observer, stellar wind. These winds have a wide range of properties. At one extreme
are shown in Figure 14-12 for models with parameters specified in the are very massive flows (mass-loss rate ~10- 5 A 0 /year) that are optically
figure caption. Characteristic P-Cygni profiles are obtained with both the thick in spectral lines (and even in some continua) and produce emission
emission intensity and the absorption depth increasing with increasing lines and P-Cygni profiles; at the other are relatively tenuous flows such as
optical depth. that of the Sun, which is optically thin and inconsequential in terms ofmass-
The solution of the comoving-frame transfer equation by the method loss (10- 14 A 0 /year), but still of great importance to the solar angular
described above is convenient, efficient, and easily generalized to realistic momentum balance. In Chapter 14 we discussed the problem of spectrum
stellar models. The method can also be extended to apply to multilevel formation in a given flow; here we shall examine the dynamics of the wind and
atoms, and should permit the computation ofline spectra for realistic model analyze questions of momentum and energy balance. We shall find that there
atoms in expanding atmospheres. Accurate theoretical spectra, when com- are two primary mechanisms for producing stellar winds. (1) In stars with
pared with observation, should assist in the determination of the physical hydrogen convection zones (such as the Sun), the outer atmosphere is a
structure of the atmospheres of stars with expanding envelopes and stellar mechanically-heated corona of very high temperature. Here we find that the
winds. corona cannot establish a static pressure balance with the interstellar medium,
but must inevitably expand supersonically, driving the flow by tapping the
thermal energy of the gas. (2) In early-type high-luminosity stars, the radia-
tion field is so intense that momentum imparted to the gas by photons drives
512 Stellar Winds 15-1 The Equations of Hydrodynamics 513

the material in a transsonic flow. (There is also some observational evidence dx 2 dx 3 dVi dV2 dV3 gives the number of particles of type k in the volume
that there may be a corona near the surface of these stars, and our theoretical element (r, r + dr) = [(x 1, x 1 + dx 1 ), (x 2 , x 2 + dx 2 ), (x 3 , x 3 + dx 3 )], with
models for their winds are only preliminary.) velocities on the range
Stellar winds have important implications for many astrophysical prob-
lems. In some cases, the mass-loss rate is so large as to produce a significant
change in the star's mass on a thermonuclear-evolution time-scale, and hence
directly to affect the star's evolution track. In other cases, the possibility of at time t. The plasma is assumed to be chemically homogeneous, so that the
noncatastrophic mass-loss over its entire lifetime may permit a star to relative numbers of particles of different species are the same throughout the
evolve to a white-dwarf configuration without becoming a supernova. gas. The velocity distribution is characterized physically in terms of macro-
Stellar winds act as brakes on stellar rotation, and hence strongly influence scopic flow velocities and a microscopic thermal distribution; the former
the angular momentum content of stars. Further, stellar winds represent describe the average motions of particles (and hence the bulk fluid motion)
important sources of mass- and energy-input into the interstellar medium, as seen in a fixed laboratory frame, while the latter gives the random individual
and thus help to determine its composition and thermodynamic state. In this particle motions relative to the average. We assume that the Coulomb colli-
book we shall consider only winds from single, isolated, stars. Stellar winds sion rate in the plasma is so large that (a) there is no drift of any species
can also occur in binaries, where they may induce rapid mass-exchange that relative to any other, and (b) on the microscopic level there is perfect
radically alters the course of stellar evolution, or, in some cases, produce equipartition of energy at each point, so that all species have the same thermal
exotic objects such as X-ray sources; although space does not permit a dis- distrubtion specified by a unique temperature T(r). Further, the microscopic
cussion of these phenomena, the material presented here is basic to a study velocity distribution is assumed to be isotropic.
of the more complex cases just mentioned, and provides a background for an The number density (cm- 3 ) of particles of species k is given by
approach to the literature.

I
15-1 The Equations of Hydrodynamics The mass density of species k is mknk(r, t), and the total density (gm cm - 3 ) is I\)
for an Ideal Compressible Fluid CJ)
(15-2) 01
In this section we shall develop briefly the equations of hydrodynamics for I
an ideal {nonviscous) compressible fluid, which is taken to be a perfect gas. The average velocity (i.e., thefiuid-:flow velocity)_ in the ith direction is
We shall ignore ionization effects, and assume that the material is already
essentially completely ionized. No attempt will be made to discuss the equa-
tions in great depth, as numerous excellent texts and monographs on the
nt<V,)k = s~CX) dVi s~CX) dV2 s~cx, dV3fk(r, V, t)V, (15-3)
subject of hydrodynamics are readily available (385; 692; 104; 490). For
expository convenience, the equations will be derived in Cartesian coordi- As mentioned above, ( V,)k is taken to be the same for all species, hence the
nates when explicit reference to a coordinate system is required, then restated subscript may be omitted. The full velocity component V, of any particular
in vector-tensor notation, and finally rewritten in spherical coordinates particle may now be written as V, = ( V,) + v;, where v; is the random
(assuming spherical symmetry) for application to the stellar wind problem. thermal velocity in the ith direction. Clearly ( v;) = 0. The complete fluid
velocity is
v(r, t) = (Vi)i + (V2 )j + (V3 )k
KINEMATICS
= v1(r, t)i + vi(r, t)j + v3(r, t)k ( 15-4)
Let us first consider some of the basic kinematic properties of the fluid.
The motion of the fluid results in mass transport and the massfiux is given
We consider the gas to consist of a mixture of particles of different species
(e.g., protons, electrons, heavy ions). Each species k has a mass mk, and a
space and velocity distribution function J,.(r, V, t), defined such that Ji. dx 1
by tmknk((V1 )i + (V2 )j + (V3 )k) = (tmknk)v = pv (15-5)
514 Stellar Winds 15-1 The Equations of Hydrodynamics SIS

Similarly, the particles carry momentum, and the rate of transport of the ith position of the fluid element we have
component of momentum, across a surface oriented perpendicular to the
jth direction, by particles of species k, is
cx(t + At) = cx(r, t) + At [(ocx/ot). + ~ (oCL/ox ),v,]
1

n,/(r, t) = mk f~"' dVi f~"' dVi f~"' dV3 Ji,,(r, V, t)Vi~ = cx(r, t) + At[(ocx/ot). + (v · V)cx]
= mk f~"' dVi f~"' dV f~"' dV3 Ji,,(r, V, t)(v + Vj)(v
2 1 1 + Vj) It thus follows that for any ex (scalar or vector)
= mknk(r, t)(v,v1 + v,( Vj)k + vj( Vi)k + ( VjVj)k)
(Dcx/Dt) = (ocx/ot) + (v · V)cx (15-10)
(15-6)
The Lagrangian system is particularly advantageous for time-dependent
The quantity ( ViVi)t is the average of the ith ahd jth components of the one-dimensional (planar or spherically symmetric) flows, while the Eulerian
random thermal velocity, and because the thermal distribution is isotropic system has advantages in complicated geometries. For steady flows [i.e., all
and the individual components are uncorrelated, properties constant in time as viewed by an external observer, so that
(o/ot) = OJ, one usually employs the Eulerian system.
(15-7)
Hence THE EQUATION OF CONTINUITY
(15-8)
Consider a fixed volume element d-r in the fluid. Ifwe demand overall mass
where Pk is the partial pressure from species k. Summing o'(er all species, the conservation, then the rate of decrease of the mass in dt is given by the net
total momentum flux tensor is given by mass flux out of the element through the surface S enclosing the element.
That is,
I
(15-11) I\)
CJ)
CJ)
where p is the total gas pressure. where the second equality follows from the divergence theorem. The equation
Finally, there are two useful (and conceptually rather different) schemes written above is true for an arbitrary element d-r, and hence the integrands I
for describing the changes that occur in the fluid as a result of material must be equal; therefore we have the equation .of continuity
motions. As an external observer views the fluid, the natural description of
its properties will be to write CL(Xi, x 2 , x 3 , t) for any property CL. The variation (op/ot) + V · (pv) = 0 (15-12)
of CL, as a function of time and position, is described in terms of a time deriva-
or, in view of equation (15-10),
tive (o/ot) computed at fixed (x 1, x 2 , x 3 ), and space derivatives (o/ox,)
evaluated at a fixed t. This system is called the Eulerian description. Alter-
natively, imagine that we follow the motion of a fluid element, consisting of
(Dp/Dt) + pV · v =0 (15-13)
a definite sample of material; this system is called the Lagrangian description. Equation (15-12) was also derived in §5-4 from the non-LTE rate equations
The time variation of the properties of a Lagrangian fluid element is described ( cf. equation (5-50)]. For steady flow, (o/ot) = 0, hence
in terms o(thefluid-frame time derivative (D/Dt) (also known as the "Lagran-
gian", "total", or "substantial" derivative). The relationship of the Lagrangian V · (pv) =0 (15-14
derivative to derivatives in the Eulerian frame may be obtained as follows. The
derivative (DCL/Dt) is defined as the limit, as M-+ 0, of[CL(t + At) - CL(t}]/At, For a one-dimensional spherical flow, equation (15-12) becomes
where CL is measured in the fluid frame at times t and t + At, at (in general) two
different positions: r, and r + Ar = r + v At. Allowing for the change in (15-15)
516 S1e/lar Winds 15-1 Th<' Eq11ario11s of Hydrodynamics 511

which, for steady flow implies ENERGY EQUATION

The statement of energy conservation for an element of gas is conveniently


41tr 2 pv = constant = .ti (15-16) given by the first law of thermodynamics, which equates the change in the
Here .,1/ is the mass-loss rate through an entire spherical surface and v denotes specific (i.e., per unit mass) internal energy, de, plus the work done by the gas
the velocity in the radial direction. pressure p when the specific volume changes by an amount d(l/p), to the
amount of heat per unit mass added to the element, dq. That is,
Exercise 15-/: Show that for any physical variable a (scalar or vector) the
equation of continuity implies that de + p d(l/p) = dq (15-23)

p(Da/Dt) = iJ(pa)/iJt + V · (pav) (15-17) Imagine that these changes occur as we follow the motion of a Lagrangian
fluid element for a time 6.t, and assume that heat exchange with the sur-
MOMENTUM EQUATIONS
roundings occurs by means of conduction only (other.sources and sinks-e.g.,
radiation and mechanical energy dissipation-being neglected) with a con-
The momentum density (i.e., momentum per unit volume) in the flow is ductive flux qc, which yields a rate of energy loss per unit mass of (1/p) V • q<.
pv. If we again consider a fixed volume element dt, we may equate the time Then equation (15-23) yields the gas energy equation
rate of change of the momentum in dt to the sum of the losses from the
momentum flux through the surface S enclosing dt, and the gains from the p{(De/Dt) + p[D(l/p)/Dt]} = -V · qc (15-24)
force per unit volume f acting on the material in dt. That is, Normally the conductive flux is written as qc = - K(VT) where K is the
thermal conductivity of the material.
:t (f pv dt) = -Ps n. dS + Jr dt From the equation of continuity, one sees that p[D(l/p)/Dt]
hence equation (15-24) can be written in the alternative form
= V · v;

= f (f - V · Il) dt (15-18) p(De/Dt) + p(V · v) =- V · qc (15-25)


I
where the divergence theorem has again been employed. As the element dtr Now, from the momentum equation, a statement of mechanical energy con- I\)
is arbitrary, we conclude that servation can be derived by taking the dot product of equation (15-21) with CJ)
v, to find
iJ(pv)/iJt = -V · Il + f (15-19) p[D(½v 2 )/Dt] + (v · V}p = v · f (15-26) "'--1
I
Substituting for Il from equation (15-9) we have Adding equations (15-25) and (15-26), and noting that V · (ab) = (b · V)a +
a(V · b), we find a total energy equation
[e(pv)/ot] + V · (pvv) = - Vp + f (15-20)
p[D(½v 2 + e}/Dt] + V · (pv) = v·f - V · qc (15-27)
In view of equation (15-17), we can rewrite equation (15-20) to obtain the
equations of motion Applying equation (15-17) we may write instead

p(Dv/Dt) = p[(iJv/iJt) + (v · V)v] = -Vp + f (15-21) [iJ(½pv 2 + pe)/iJt] + V · [(½pv 2 + pe + p)v + qc] = v·f (15-28)

For a spherically symmetric one-dimensional flow with a gravity force For one-dimensional spherically-symmetric.flow equation (15-28) becomes
f, = -(G.,I{ p/r 2 ), equation (15-21) becomes

p(iJv/iJt) + pv(iJv/iJr) = -(op/or) - (G.,/lp/r 2 ) (15-22)


a
ot r or
o[
- (½pv 2 + pe) + -12 - r2 pv (tv + e + -p)p - r (oT)]
2 2
or
K - pv
= - -r - G.,I{
2

For steady flow the first term of equation (15-22) vanishes. (15-29)
518 Stellar Winds
15-1 The Equations of Hydrodynamics 519
For steady spherically symmetric flow
Equation (15-34) is the wave equation, in which the wave velocity is a. We

:r [r pv(½v
2 2
+ h) - r 2 ,c (~~)] + (r 2 pv)(GA/r2 ) = 0 (15-30)
thus identify the adiabatic speed of sound as

a, = [(ap/ap).]t (15-35)
where h = e + (p/p) is the specific enthalpy. Recalling that (r2 pv) = constant, For a perfect gas undergoing adiabatic changes, (p/p 0 ) = (p/p 0)Y, where y
we may integrate equation (15-30) to obtain the total energy integral is the ratio of specific heats at constant pressure and constant volume
(y = i for an ideal monatomic gas); thus
(41tr 2pv) [½v 2 + h - (G.J(/r)] - (41tr 2),c(aT/ar) = C (15-31)
(15-36)
Equation (15-31) states that the total energy flux through a spherical surface
(consisting of the flux of kinetic energy, gas enthalpy, potential energy, and where µ is the number of atomic units per free particle. In certain situations
heat conduction) is a constant in a steady flow. there can be a very free exchange of energy between the wave and its sur-
roundings via conduction or radiation; in this case the temperature fluctua-
SOUND WAVES tions in the sound wave are destroyed (the wave is thereby damped), and the
wave propagation proceeds isothermally. In this event p = p(kT0 /µm") and
When a compressible fluid suffers a small impulsive disturbance, an the isothermal sound speed is
oscillatory motion of small amplitude (consisting of alternating compressions
and rarefactions) is generated and propagates through the medium. These (15-37)
oscillations are sound waves, and their characteristic speed is the speed of
sound. Suppose the medium is planar and homogeneous, and originally at
THE RANKINE·HUGONIOT RELATIONS
rest with an ambient density p0 , and pressure p0 • Write the ~rturbed values
FOR STATIONARY SHOCKS
of density and pressure as p = Po + p 1 and p = Po + p 1 , respectively,
where it is assumed that p 1 « p0 and p 1 « Po; further, v, the velocity of the The speed of sound determines the rate at which a disturbance propa-• I
disturbance, is to be regarded as a small quantity. Then substituting these gates naturally in a compressible fluid. If the velocity of the flow exceeds the I\)
expressions into the equation of continuity (15-12) and the momentum speed of sound, then the adjustments of the fluid state by sound waves can- 0)
equation (15-21) (with the external force f = 0), expanding, and retaining not proceed quickly enough, and situations occur where the properties of CD
only terms cifthe first order in the perturbations, we find the flow may change markedly over a rather small distance. This gives rise I
to almost discontinuous changes at a narrow interface called a shock front.
(ap 1/at) + p0 (av/ax) =0 (15-32) Shocks may occur in a wide variety of circumstances. For example, in a
pulsating star, a wave initiated at great depth will propagate with larger and
and p0 (av/at) + (apifax) =0 (15-33)
larger velocities as it progresses upward. in the atmosphere, because of the
decrease in density; eventually it runs through the material at supersonic
For an ideal fluid (i.e., no viscosity or conductivity), there is no thermal
energy exchange from one fluid element to another, and the material behaves velocities, producing a shock front that travels outward into the upper layers
of the atmosphere. Such shocks produce interesting spectroscopic phenom-
adiabatically. We may therefore write
ena in RR Lyrae and Cepheid variables, but we shall not be able to pursue
this subject here. Alternatively, in a steady flow, a stationary shock front may
be produced where the gas, in effect, encounters an obstacle (e.g., a constric-
where (ap/ap), is the isentropic derivative of the pressure with respect to tion in a nozzle or flow channel or at the point where a stellar wind runs
density. Now, differentiating equation (15-32) with respect to t, and (15-33) into the interstellar medium). For the purposes of this chapter, only stationary
with respect to x, and eliminating the cross derivative (a 2 v/at ax), we obtain shocks will be considered.
The shock front itself has a structure whose properties are determined by
(15-34) dissipative processes involving viscosity, heat conduction, and radiation;
furthermore, the downstream material may be forced out of equilibrium in
520 St<!llar Winds 15-2 Coronal Winds 521

one or more degrees of freedom, depending on how long are the characteristic Exercise 15-2: (a) Verify equations (15-42) and (15-43): (b) Show that equation
rela~ation times that restore the equilibrium [see (692) for a detailed dis- (15-43) can be rewritten in the customary forms
cussion of these phenomena]. These effects will be ignored here, and it will
be assumed that the fluid is a perfect monatomic gas in equilibrium through- or
e2- - e1 = !(P1 -I - P2 -t)(p1 + P2)
out, and t_hat the shock is a discontinuity at an infinitely sharp interface. h2 - h1 = ½(Pi- 1 + P2- 1)(P2 - P1)
If we consider the flow of particles across an interface, it is clear that the flux Exercise 15-3: (a) Verify equations (15-44) through (15-46). (b) Show that the
of mass, momentum, and energy must be conserved. Using subscript 1 to downstream Mach number is
de~ote upstream quantities (i.e., material flowing into the shock), and sub-
script 2 t_o denot~ downst~eam quantities (i.e., post-shock material), the M/ = [(y - l)M 1 2 + 2]/[2yM/ - (y - l)]
conservation requirements Just mentioned become (in planar geometry) and that this implies the downstream flow is always subsonic.

PtVt = P2V2 (15-38)


2 15-2 Coronal Winds
Pt+ P1Vt = P2 + P2Vl (15-39)
and P1v,[e1 +(pi/pi)+ ½v, 2] = P2V2[e2 + (P2IP2) + ½v22] (15-40) The outermost layers of the solar atmosphere comprise the corona, a tenuous
(characteristic electron density ~4 x 108 cm - 3 ), high-temperature (T ~
The appropriate fluxes have been identified from equations (15-5) (15-9) 1.5 x 10 6 °K) envelope which can be observed, at eclipses, to extend to
and (15-28). In view of equation (15-38), and the results for a perfect ga~ several solar radii. As was described in §7-7, the high coronal temperature
that e = (p/p)/(y - 1) so that h = e + (p/p) = y(p/p)/(y - 1) equation is a result of energy input from wave dissipation. The ultimate source of the
(15-40) can be rewritten as '
waves providing this mechanical heating is the hydrogen convection zone,
and we believe that all stars that have extensive convection zones should
(15-41)
also have coronae.
If we eliminate V2 by means -of equation (15-38), use the definition of the For many years the corona was regarded as an essentially static envelope
I
sound spe~d in equation (15-36), and define the Mach number M 1 = vi/a., bound to the Sun, and although it was realized that particle bombardment
N
then equations (15-39) and (15-41) may be rewritten as of the Earth from energetic events (e.g., flares) on the Sun produced auroral CJ)
and geomagnetic effects (164; 165), it was Biermann (91; 92; 93) who first
(P2/P1) = 1 + yM 12[1 - (pif p2)] (15-42) advanced the idea of continuous particle emission ("corpuscular radiation")
co
I
and from the Sun. Realizing the importance, at coronal temperatures, of energy
2[(P1P2/P2Pt) - 1] = (y - 1)M/[1 - (pifp 2)2] (15-43) transport by conduction, Chapman (166) showed that the corona extended
which are kn~wn as the Ran~ine-Hugoniot relations. Equation (15-42) gives far out into interplanetary space and, in fact, enveloped the Earth in a low-
P2 as a function of P2 for given (pi, p 1). In principle, two solutions exist: density high-temperature medium. Subsequently Parker showed (496; 498).
(a) P2 > Pt and P2 > P1 and (b) P2 < p 1 and p2 < p 1• However, it is that any reasonable hydrostatic model of the corona, starting from known
found [see (385, §§82-84) or (692, §17)] that case (b) is excluded on the basis conditions near the Sun, led to such high pressures at large distances, as to
of the second law of thermodynamics and stability considerations, and that preclude the possibility of containment by the pressure of the interstellar
only compression shocks are physically possible. medium. Thus static models are internally inconsistent, and large-scale coronal
Equations (15-42) and (15-43) can be solved simultaneously for (p 2/p 1 ) expansion must occur. This expansion provides the source of the particle
and (P2/p 1), yielding emission proposed by Biermann.
Parker developed a theoretical model indicating a flow with low velocities
(P2/P1) = (vi/V2) = (y + l)M 1 2/[(y - l)M / + 2] (15-44) near the Sun, rising to very large supersonic values at large distances; he
called this transsonic flow the solar wind, and made predictions of typical
and (P2IP1) = [2yM I 2 - (y - 1)]/(y + 1) (15-45) flow velocities, densities, and particle fluxes at the Earth's orbit. These
Applying the perfect gas law we also may write predictions were supported by many bits of evidence [see the summaries
of the fascinating historical growth of our conceptions of the solar wind in
(T2/T1) = [2yM/ - (y - l)][(y - 1)Mt 2 + 2]/M/(y + 1)2 (15-46) (107, Chap. 1) and (324, Chap. 1)], but an alternative model was proposed
522 Stellar Winds J5-2 Coronal Winds 523

by Chamberlain (147), which was everywhere subsonic and produced only Exercise 15-4: Using the particle flux given in Table 15-1 show that the rate of
~ low-velocity solar breeze at the Earth's orbit. The ensuing debate in the mass-loss, ,,If = 4nr2 nvm, where n is the particle density an'd m is a mean particle
literature, based almost entirely on theoretical considerations, led to a mass, is ~2 x 10- 14 A 0 /year. Compare this with the thermonuclear mass-loss
sharpening of the formulation of the theory (and makes interesting reading rate required to yield the solar luminosity.
today because of the high standard of scientific argumentation it contains!).
As mentioned previously, we believe there should be coronae for all stars
But the question of which picture really applied remained undecided until
that have hydrogen convection zones (i.e., spectral types F and later); these
direct measurements from space vehicles provided conclusive evidence in
stars should also have stellar winds. There is, in fact, direct observational
favor of a high-speed wind.
evidence (e.g., P_-Cygni Iine:Profiles) for. massive coronal winds in many
Detailed measurements of solar-wind properties show that there is con-
late-type superg1ants and giants. Thus estimates of ·mass-loss rates [see,
siderable fluctuation of all the physical variables; many of the variations
e.g., (624, 238; 70, 246)] range from 2 x 10- 10 .,If 0 /year for K-giants to
correlate with a solar rotation period and hence reflect changes in the initial
10-s A 0 /year for M-giants; and from 10- 7 .,If 0 /year for G- and K-
conditions of the flow at the corona, while others are the result of violent
supergiants to 10- 6 or 10- 5 A 0 /year for M-supergiants. All of these are
events such as flare-generated blast waves. DesP.ite these variations it is
coronal winds resulting mainly from the high temperature of the stellar
useful to consider the background low-speed "q~iet" wind, whose charac-
corona (~xcept possibly for the M-stars, where radiation pressure on grains
teristics must result from the conditions prevalent in the corona as a whole.
~ay be important). Another class ofm~ssive winds (10- 6 -10- 5 .,H 0 /year)
A summary of quiet solar-wind properties measured at the Earth's orbit is
1s found for early-type stars (OB supergiants and WR stars); these flows are
given in Table 15-1. Although it must be emphasized that these data define
driven by radiation pressure acting on the material (cf. §15.4).
only a high-order abstraction of the real solar wind, they nevertheless provide
In this section we shall examine the basic physics of coronal winds, using
typical values against which theoretical results can be compared. From the
simplified and idealized descriptions, both of the material properties and
data given, we see that the wind is a highly supersonic (M!l_ch number ~ 8),
of the nature of the flow. Our goal will be to gain insight into the overall
nearly-radial flow. It is easy to show (Exercise 15-4) that the effects of the
nature of these winds, rather than to attempt detailed modeling. We shall
mass-loss in the wind on the evolution of the Sun as a star are negligible.
repeatedly refer to the solar wind as an example, and as an indicator of what I
Moreover, the corona-wind ensemble is optically thin. Thus at first si~t I\)
other processes need to be considered in an "ultimate" stellar-wind theory.
the flow seems to be of little significance from a stellar-astrophysical poi~t -..J
of view. We shall see, however, that even this rather weak (see below) flow 0
EXPANSION OF THE SOLAR CORONA
has implications of great consequence to the question of angular-momentum I
loss by the Sun. Suppose the corona is taken to be a fully-ionized pure-hydrogen plasma
with_ To~ 1.5 x 106 °K and n0 ~ 4 x 108 cm- 3 , where n0 denotes the
TABLE IS-I
Properties of lhe Quiet Solar Wind al the Earth's Orbit
density of protons or electrons (n, = ne for charge neutrality). At these high
tem~eratures the ~lectrons have large velocities, and can conduct energy
Radial component orHow velocity 300-325 km s-• efficiently, producmg an energy flux q, = -IC VT, where the conductivity
Nonradial component or now velocity 8 km s- • ICis given (598, 86) by IC = Ko yt with ICo ~ 8 x 10- 7 ergs cm-• s- 1 deg-I.
Proton (or electron) density 9 cm- 3 At 10 6 °K, the conductivity of the plasma far exceeds ordinary laboratory
Average electron temperature 1.5 x 10 5 °K conductors! If we assume that conduction is the dominant energy transport
Average proton temperature 4 x 104 °K mechanism in the corona, then for equilibrium we must have V , q = O
Magnetic field intensity S x 10-s gauss which implies that • '
Proton flux 2.4 x 101 cm· 2 s· 1 (15-47)
Kinetic energy flux 2.2 x 10- 1 ergcm- 2 s·•
Enthalpy flux 8 >< 10- 3 crgcm· 2 s- 1 This equation can be integrated twice, and demanding that T-+ 0 as r -+ oo,
Gravitational flux 4 x 10- 3 erg cm- 2 s- • we find· T(r) = T (r /r)'+ (15-48)
0 0
Electron heat-conduction flux 7 x 10- 3 erg cm· 2 s· 1
where r 0 denotes a suitable reference-level in the corona (which is of the
Souaa: Adapted from (324, 44-45), by permission. order of R 0 = 7 x 1010 cm). The temperature fall-off implied by equation
524 Stellar Winds
15-2 Coronal Winds 525

(15-48) is extremely slow, and at the earth's orbit (re = 1.5 x 10 13 cm) we magnetic fields in the interstellar medium. The average density of the medium
find T ~ 3.3 x 10 5 °K for T 0 = 1.5 x 106 °K; thus the earth is enveloped is of the order of 1 cm - 3 ; for an H I region, T ~ 10 2 °K, while in a typical H II
by high-temperature coronal material. region T ~ 10 4 °K, so that we expect pressures to range from 1.4 x 10- 14
If the corona were in hydrostatic equilibrium, we would have to 3 x 10- 12 dynes cm- 2 • The magnetic field in the medium is of the order of
(15-49) 10-s gauss; hence the magnetic pressure (B 2 /81t) is of the order of 4 x 10- 12
dynes cm - 2 • The pressure attributable to cosmic rays is estimated to be
The Debye length in the corona at the temperature and density quoted about 2 x 10- 1 2 dynes cm - 2 ; hence a reasonable upper bound of 10- 11
above is only about 0.3 cm [cf. equation (9-101)]. Hence the plasma is dynes cm - 2 can be assigned as the total interstellar pre~sure. This value is a
neutral except on the smallest scales, and we may write n, = np = n and factor of 10 6 smaller than that estimated for a hydrostatic corona, and we
p = n(mp + m.) = nm, where m denotes the mass of a hydrogen atom, and conclude that the corona cannot be contained, but must undergo a steady
p = 2nkT. Then the hydrostatic equation becomes expansion. We therefore abandon the assumption of hydrostatic equilibrium
and inquire into the hydrodynamics of coronal expansion.
d(nT)/dr = -(GA 0 m/2k)(n/r2 ) (15-50)
ONE-FLUID MODELS OF STEADY,
If T were constant, and we were to consider only distances r ~ r 0 , we could
SPHERICALLY SYMMETRIC CORONAL WINDS
approximate equation (15-50) with the usual planar limiting form d(ln n)/dr =
-(1/H), where the scale-height H = (GA 0 m/2kT0 r 0 2 )- 1 ~ 10 5 km. As To simplify the treatment of the hydrodynamics of a coronal wind, the
shown in Exercise 15-5, this scale-height is so large that the predicted particle following assumptions are made: (1) the flow is taken to be steady; (2) the
density near the earth's orbit is quite large. We would then conclude that the wind is spherically symmetric; (3) the gas is an ideal compressible fluid.
Earth is immersed in a hot, fairly dense, extension of the-corona. Actually Clearly all of these assumptions are idealizations, and they restrict the appli-
the assumption that T = T 0 is too crude, and produces somewhat of an cability of the results for, say, the real solar wind; nevertheless they provide
overestimate of n(re)- If, instead, we make use of equation (15-48), we may a good basic physical model for general study. I
rewrite equation (15-50) as I\.)
The equations describing a steady flow are the equation of continuity
(mass conservation) ---..J
which yiel~s the solution
(15-54)
d[(r0 /r)+n]/d(r/r 0 ) = -(r0 /r) 211/H (15-51) the momentum equation

n(r) = n0 (r/r0 )+ exp{ - 7r0 [1 - (r0 /r)']/5H} (15-52) pv(dv/dr) = -(dp/dr) - p(GA 0 /r 2 ) (15-55)

from which we deduce n ~9 x 104 near the earth for n0 =4 x 108 • and the energy equation (including conductivity)

Exercise J5-5: Assume the corona is isothermal, T =


T 0 , but allow for the varia- _!_2 !!.. [,2 pv (
r dr
½v2 + e + !!.)]
p
= _ pv (G"':' 0) + ..;. !!._ (,2,c dT)
r r dr dr
(15-56)
tion of the gravitational force to obtain an expression for n(r). With n0 ~ 4 x
108 cm 3,show that the particledensityneartheearth's orbit isabout4 x 10 5 cm· 3• For a fully ionized gas of pure hydrogen, p = 2nkT and p = nm, where
Exercise 15-6; Show that equation (15-52) yields a minimum for n at (r/r0 ) = n = nP = n, and m is the mass of a hydrogen atom. The ·specific internal
(7/4)(r0 / H), and that n(r) increases outward beyond that point (formally this energy of the gas is 2(fnkT)/(nm) = (3kT/m), and (p/p) = (2nkT)/(nm) =
configuration would be unstable against a Rayleigh-Taylor instability).
(2kT/m), so the specific enthalpy is h = e + (p/p) = (5kT/m). Note that
equation (15-56) does not include any terms representing radiative losses or
Combining equations (15-48) and (15-52), we find that the pressure is mechanical energy deposition from dissipation of waves. The assumption
p(r) = Po exp{ - 7r0 [1 - (r 0 /r)i]/5H} (15-53) made is that these processes occur in a relatively thin layer at the base of the
corona, and that above some reference level only the kinetic, potential,
Instead of vanishing at (r/r0 ) = ex:,, the pressure approaches a finite value. thermal (enthalpy), and conductive terms need be taken into account.
Adopting p0 ~ 0.2 dynes cm- 2, we find p.., ~ 10- 5 dynes cm- 2 • This pres- The continuity equation can be integrated to give
sure is to be compared with the pressure expected from the material and
4nr 2 nv = F = constant (15-57)
15-2 Coronal Winds 521
526 Stellar Winds
The critical velocity v, defined by equation (15-64) is equal to the isothermal
where F denotes the particle flux, or sound· speed.
We now restrict attention to solutions for which both v and (dv/dr) are
.Ii = mF = 41tr 2 nmv = constant (15-58) single-valued and continuous. First, suppose that equation (15-63) is satisfied.
We may then construct solutions for which 1 - (2~T 0 /mv~) has the sa~e
where A denotes the rate of mass-loss. The energy equation can be integrated sign for all r. If v(r,) < v, then v(~) wi!I have a relative max1I?um at r,; this
to give yields solutions of type 1 shown m Figure 15-1. These sol~t1ons are eve~-
where subsonic. On the other hand, if v(r,) > v,, then v(r) will have a relative
(4,rr 2 nv)[½mv 2 + 5kT - (G~m/r)] - (4,rr 2 ),c(dT/dr) = E = constant
(15-59)
which states that the total energy flux through a spherical surface is a constant 3 ,--..-,-.-.----r---r----,----,----,
of the flow. Equations (15-57), (15-55), and (15-59) comprise a nonlinear
system of two first-order differential equations, c;:ontaining two constants of
integration; two more constants will result from the integration of the system;
hence a total of four conditions (boundary conditions or specifications of
the nature of the solution) must, in general, be imposed. Before discussing
this general problem, however, it is extremely instructive to consider the
simple case of an isothermal corona. 2
If we demand that the flow be strictly isothermal, we need to specify only
n and v, and hence we may omit the energy equation ancl employ only the
r
momentum and continuity equations. [The ad hoc assumption = constant
I
is equivalent to invoking some unspecified mechanism to heat or cool the gas (v/v,)
r\.)
in just the right way as to yield the desired temperature.] Equation (15-55)
becomes -...,J
r\.)
nmv(dv/dr) = -2kT0 (dn/dr) - nm(G~/r2 ) (15-60)
I
and using equation (15-57) to eliminate n we find .
½[1 - (2kT 0/mv 2 )](dv 2/dr) = -(G~/r2)[1 - (4kT 0r/G~m)] (15-61)
This equation yields several f ami/ies of solutions that have substantially
different mathematical behaviors and physical significance. For the solar
corona, (4kT 0r0/G.,lf 0 m) ~ 0.3. Thus it is clear that the righthand side of
equation (15-61) is negative for r < r, and positive for r, < r < co, where
0 2
r, is the critical radius
(r/r,)
(15-62) FIGURE 15-1
Schematic variation or velocity in units or the critical velocity v., as a
at which the righthand side is exactly zero. For the solar corona, (r,/r0 ) ~ 3.5. runction or radial distance in units or the critical radius r., for stellar
At r = r., the lefthand side of equation (15-61) must also be zero; this may winds and breezes. Solutions or type I are the subsonic breezes. Solutions
occur in one of two ways. We may either have or type 2 arc everywhere supersonic. Solu!i~ns 3 ~nd 4 are the transsonic
critical solutions that pas~ through the cnucal point (heavy dot)
continuously. Solutions or types S and 6 are double-valued, but are
(dv/dr),. =0 (15-63)
important for fitting shock transitions.
or v(r,) = (2kT 0/m)t = v, (15-64)
528 Stellar Winds
529
minimum at r., and will be everywhere supersonic; these are the type-2 1200,------,--......--.---,--,---,--,-,-.--,
solutions shown in Figure 15-1.
Alternatively, if equation (15-64) is satisfied, we obtain a critical solution 1000
(i.e., v = v, at r = r,) that has a finite _slope at_r = r,. S~ppose that (dv(dr),. >. 0; 3 X106 °K I
then we obtain a unique transsomc solution that 1s monotone mcreasmg 800 2 X106 °KI
from subsonic speeds (v < v,) for r < r, to supersonic velocities (v > v,) for 1.5 x 106 °KI
V 600
r > r,. This unique solution is shown as type 3 in Figure 15-1. If(dv/dr),. "<: 0,
then we obtain a unique solution in which v(r) is monotonically decreasing 400
from supersonic speeds for r < r, to subsonic speeds for r > r,; this is shown
200
as type 4 in Figure 15-1. .. .
Finally, there exist two fam1hes of solutions defined by the restrictions
. . Orbit of Earth - I
r;;,, '• > r, or r ~ r• < r,, both of which have v(r.)·_= v, and (dv/dr),. = <:'· 0 20 40 60 80 100 120 140 160
These give rise to the double-valued solutions of types 5 an~ 6 shown m
Figure 15-1. Initially we exclude consideration of these solutions because FIGURE 15-2
they are double-valued; we shall find later, however, that they may be used Isothermal solar wind solutions; curves arc
to provide part of a complete solution that includes a shock transition to labeled with coronal temperature. Ordinate: velocity
match boundary conditions as r -+ oo. in km s • 1 ; abscissa: heliocentric distance in 106 km.
From (496), by permission.
To decide which solution to choose, we must now invoke physical bound-
ary conditions. First, both the entire family oftype-2 solutions and the unique
type-4 solution can be excluded because they predict velocities v > v, :::::
the orbit of the Earth the solar wind has velocities of the order of a few
170 km s - 1 in the low corona, which is not observed. This- leaves solutions
hundred km s- 1 • Several wind solutions for isothermal coronae at different
of type 1 or the unique transsonic solution of type 3. We may-make the
temperatures are shown in Figure 15-2. The wind solution given above is
choice on the basis of the behavior of the solution as r -+ oo. Equation (15-61) I
not entirely satisfactory because (v/v,) increases without limit as (r/r,) -+ oo. I\)
may be integrated straightaway to yield
This behavior is an artifact of the assumption that the corona is strictly
isothermal. The corona is, in fact, nearly isothermal near the Sun because
--..J
(v/v,) 2 - ln{v/v,) 2 = 4 ln(r/r,) + 4(r,/r) + C (15-65)
conduction is effective and distributes the heat deposited by wave dissipation w
very efficiently, thus obliterating large temperature gradients. At large dis- I
Exercise /5-7: (a) Verify equation (15-65). (b) Evaluate C for the critical solu- tances, however, expansion of the gas must ultimately force it to cool; if we
tions. (c) Evaluate C for solutions of types 1 and 2.
insist that Tremain fixed, we have, in effect, introduced a (spurious) energy
source to maintain the temperature. This effectively results in a continuous
Consider first solutions of type 1 as r -+ oo. Here (v/v,) is < 1 and decreasing deposit of energy into the enthalpy of the gas, and this energy is available
as r -+ oo hence on the lefthand side the dominant term will be - 2 ln(v/v,), to do work· to continue to accelerate the gas without limit. The problem is
and on th~ righthand side it will be 4 ln(r/r,). Thus as r-+ oo, then v oc r- 2 ,
easily overcome, however, and Parker was able to produce satisfactory wind
which implies (from the equation of continuity) that n remains finite. These
solutions by using either of two assumptions. (a) The.corona is isothermal
solutions thus yield a pressure at r = oo which greatly exceeds the ambient
for r 0 ~ r ~ '•• and expands adiabatically for r > '•· In this case, for r > '•
interstellar pressure, and therefore they can be rejected on the same grounds
the pressure and density are related by a polytropic law of the form
as the hydrostatic solution was. On the other hand, the critical solu,tion h~s
p = p0 (p/p 0 )1 with y = ¾(ideal gas). (b)The corona is everywhere polytrop_ic
(v/v) > 1 and increasing as r -+ oo, hence we find v :::.: 2v,[ln(r/r,)Jr. In this with a polytropic exponent y < f. Both of these classes of solutions avoid
cas~ n oc r- 2 v- 1 -+ Oas r-+ oo, and we conclude that this critical solution
an unphysical increase in v at large r, and one finds that v approaches a
can indeed match the boundary condition at infinity. finite value v00 as r -+ oo.
Following a line of reasoning similar to that outlined above, Parker The isothermal solutions described above reveal the broad outlines of the
concluded (correctly!) that the solar corona must be undergoing a transsonic nature of coronal expansion into a supersonic wind, but clearly are inade-
expansion into interplanetary and interstellar space (496), and that near quate for a detailed description of the flow. In particular, in solving the
530 Stellar Winds 15-2 Coronal Winds 531

problem, we wish to determine the temperature distribution T(r). We must, smoothly through the critical point, and that T(r)-+ Oas ,,-'_,. oo. The'last
there~ore, turn to the full set of equations including energy balance, namely of these conditions, however, is actually more complex than it seems, and
equations (15-55), (15-57), and (15-59). As before, there are two basic classes it is now known that the nature of the variation of T with r depends upon
of solutions to be considered: (1) subsonic solutions, called stellar breezes, the mechanism of heat transport at r = oo [see (200; 201; 202; 532;324, 47}].
which resemble those of type 1 shown in Figure 15-1; (2) transsonic critical Suppose first that the heat-conduction flux remains finite at r = oo, with a
s~lutions, stellar winds, resembling the unique solution of type 3 shown in value E0 (00). From equation (15-59) it is clear that we will then find v"' =
Figure 15-1. These two classes are distinguished from one another by the {2[E - Ec(oo)]/..ll}t, and further, because [r 2 T¼(dT/dr)]"' = constant, the
value of the total energy flux E in equation (15-59}. Tht;l breeze solutions all temperature obeys the asymptotic law T oc r-l. This is the same law found
have E = 0, while the wind solutions have E > O. by Chapman, and used by Parker (499) in his wind solutions. Next we might
Although we are primarily interested in wind solutions in the solar context suppose that both a conduction flux and an enthalpy flux persist to r = oo,
the.breeze solution_s, which have been studied extensively (147; 532}, played and that the ratio (5kTv}/[r 2 T¼(dT/dr)] ➔ constant.In this case both fluxes
an ~portant role tn the development of the theory, and it is worthwhile to approach zero at r = oo, and v"' = (2E/Ji)+. Further, the special condition
consider them briefly here. In the breeze solutions, as r ➔ oo, then u ➔ O, imposed on the ratio of the fluxes implies that T oc r-t, a solution first
T-+ 0, and p -+ 0. There is a whole family of such solutions that are obtained by Whang and Chang (667). Finally, we might suppose that the
differentiated from one another' by the limiting value of (mv 2 /kT) as r ➔ 00 conductive flux vanishes, as r -+ oo, more rapidly than the enthalpy flux.
(532}. Because a match to the interstellar pressure can now be achieved we Again v"' = (2E/A}½. In this case there is no energy exchange mechanism in
cannot exclude these solutions from the outset, as we could for an isothe;mal the flow as r ➔ oo, and the gas merely expands adiabatically. If the poly-
wind; indeed, Chamberlain (147} advocated this type of solution for the tropic exponent is y, then T oc p<Y-o. But p cc r- 2; hence, as r-+ oo, then
solar corona, in which case the velocity of the material near the Earth's T oc (r- 2)<t> or T cc r-t, for i' equal to¼ (ideal gas).
orbit would be only 20 km s- 1 • He argued that the correct hydrodynamic The relationship of these solutions to one another was explained lucidly by
solution should be consistent with an evaporative model of the corona in Durney (200), who considered wind solutions that were all chosen to have the
which individual particle motions are calculated assuming a critical l~vel same value for T 0 , but different values for n0 • For small values of n0 , the
above which the density is so low that fast-moving particles suffer no further critical solutions yield large values for e(oo) = (E/F} which is the residual
col!isions, but escape. He found that in the evaporative model the mean speed energy per particle at r = oo, and they have T oc ,-*.
As n0 is increased, I
of ions near re would be about 10 km s- 1, and concluded that the breeze the conductive part of e(oo), say e0 (00), becomes smaller, and at a particular I\)
(rather than wind) solution was correct. The question was laid aside when value (nt), e0 (00) = 0 and T oc r-t (i.e,, this particular value gives the --.,J
direct observation proved the validity of the wind solution. But subsequent Whang and Chang solution). As n0 is increased further, ec(oo) remains equal 01
work (108; 339) has shown that the basic premise that the hydrodynamic to zero, and now e(oo} decreases; here T oc r- 1 . Finally, when n0 increases I
and evaporative pictures should agree is, indeed, correct, and that the diffi- still further, a limiting value is reached at which e(oo) = 0, and a stellar
culty lay in the calculation of the original evaporative model. In particular, breeze solution is obtained (see Figure 15-3}. Phy.sically, these results imply
when account is taken of the fact that particles moving at different speeds that, as n0 increases, more and more of the conductive flux at infinity is
escape from different levels, not from a single "critical" level (the faster consumed by the expansion (the flow is more and more massive), until at
particles escape from deeper, denser layers}, and when a correct dynamic some point ec(oo} vanishes. Ifwe now add more material, the flow continues
calculation is made of the electric field in the plasma that couples the electron to be supersonic, but more and more thermal energy is consumed in the
and proton flows together, then the evaporative solutions yield n ~ 1o cm - 3, adiabatic expansion, and e(oo} decreases. Eventually e(oo} vanishes and the
v ~ .300 km s- 1, and T(proton) ~ 5 x 104 °K, in good agreement with the flow becomes subsonic.
observed values. We would now conclude that the correct evaporative model In a systematic study of stellar wind solutions, it is very convenient to
supports the solar wind solution. work with dimensionless variables. Chamberlain (147} suggested the trans-
Let us now examine the wind solutions in greater detail. In all of these formations
solutions, as r ➔ oo, then v -+ v"' (a nonzero value}, n ➔ O, and T ➔ o. As t = (T/T0 } (15-66a)
mentioned above, four distinct conditions are required to specify the solution i/1 = (v 2 µm/kT 0 } (15-66b}
uniquely. Typically, these may be chosen to be values for the coronal density
and temperature, and to meet the requirements that the solution pass and l = (G.A 0 µm/kT 0 r) (15-66c)
532 15-2 Coronal Winds 533

is the residual energy per particle at infinity (in units of kT0 ), and

A = 4nK0 Gµ 2 m.,I( 0 T 0 f/(k 2 F) (15-71)


In these equations we need specify only Bao and A to perform the integration
(imposing, of course, the requirements that a critical solution be obtained,
and that t -+ 0 as A-+ 0); by redimensionalization, one can obtain several
solutions from a single dimensionless one. A further transformation (531)
reduces the number of arbitrary constants to one (for winds only). Thus, wtite
t• = t/e 00 , " ' • = t/l/B 00 , and -'-• = A/B 00 • Then equation (15-68) has the
same form in terms of the new variables, while equation (15-69) becomes
(15-72)
where K = B00 iA. A large number of solutions for a wide range of the param-
eter K are given in (201), along with an example of how to recover a unique
dimensional solution having specified, say, T O and F. A detailed discussion
of breeze solutions is given in (532).

Exercise I5-8: Carry out the transformation to dimensionless variables, and


To verify equations (15-67) through (15-72).
FIGURE 15-3
Regions of stellar wind and breeze solutions as a function of coronal
TRANSITION TO THE INTERSTELLAR MEDIUM
temperature T0 (in units.ofG.l(eµm/kr 0 ) and density N 0 (in units of I
2,c0 k- 1(G.I( er0 )-i]..Shaded area contains breeze solutions. The I\)
dotted curve marked S delimits the region (to the right) where the The solutions derived above all have pressures that vanish at r = oc.
Actually the interstellar medium does have a finite (if small!) pressure ---.J
flow is already supersonic at the base of the corona (r .. r0 ). The u,
region P contains wind solutions with a Parker temperature law, p1 :5 10- 11 dynes cm - 2 • When the high-speed gas in the wind encounters
T ex: ,- 1, as r ... co; these solutions have a nonzero conductive flux the interstellar medium, a stationary shock front is formed (624, 306; 665) I
at infinity. The region D contains Durney temperature law (Tex: ,-1) in which the flow speed suddenly diminishes to a small value, and the density
winds, which become adiabatic as r ... co. The curve marked WC
gives the locus of solutions with a Whang-and-Chang temperature
and temperature rise. The solution jumps discontinuously from the critical
law, Tex: ,-¼ as r ... co; these particular solutions occur just when solution (curve 3 in Figure 15-1) to one of the subsonic solutions of type 6
the conductive flux at infinity first drops to zero. From (532). shown in Figure 15-1. Beyond the shock the material finally cools and
recombines, and the velocity drops to zero [see, e.g., (106)].
At large distances from the star, the wind is highly supersonic, and the gas
where µ is the number of atomic mass units per particle ( = ½for ionized pressure is negligible compared to the kinetic energy density. Thus, balance
hydrogen). In these units, equations (15-57), (15-55), and (15-59) become, with the interstellar pressure is achieved through the impact pressure of the
respectively, material in the wind. We then have

n-'-- 2 t/Jt = (kT 0 /µm)t F/(4nG 2 ✓lt&) =C (15-67) (15-73)


From the equation of continuity and the assumption that vis approximately
½(1 - (t/t/l}](dt/J/dA) = 1 - 2(-r/..l) - (dt/d..l) (15-68)
constant, we know that n(r) = n 9 (r 9 /r) 2 , and using this relation we may
and At½(dt/d-'-) = B -½t/1 + A- fr (15-69) estimate the radius of the shock front in the solar wind as
00

(15-74)
where Bao = µE/kT0 F (15-70)
534 Stellar Winds
15-2 Coronal Winds 535
Adopting n 61 ~ 10 cm- 3, 11 ~ 300 km s- 1, and the upper limit on p1 given
above, we find r, ~ 30 a.u. {i.e., outside the orbit of Neptune). The physical does rotate, however, and the field lines must be considered to be anchored
conditions in the flow beyond the shock can be estimated by applying the at the solar surface. Field lines from a particular point on the surface will be
Rankine-Hugoniot relations [equations (15-44) through {15-46)]. On the drawn out along the streamlines of fluid elements as seen by an observer in
conservative assumption that the temperature falls as r-+, the temperature the rotating frame of the Sun; this yields a spiral pattern.for the interplanetary
in the preshock flow is of the order of 4 x 104 °K, which implies a sound speed field (496;498, 137; 107, 67;324, 11).
in ionized hydrogen of about 25 km s- 1, so the Mach number is about 12. When the magnetic field is taken into account, the physics of the fluid flow
In the large Mach-number limit, the Rankine-Hugoniot relations reduce to in the wind becomes much more complicated, and considerable mathematical
complexity results; we shall not pursue this problem here but merely refer
112 = 11i{y - 1)/(y + 1) (15-75) the interested reader to the literature [e.g., (463; 659)j. It is of considerable
interest, however, to calculate the angular momentum carried away in the
n2 = n1(y + 1)/(y - 1) (15-76) wind, for this loss has important implications for stellar rotation. The mag-
netic field must satisfy Maxwell's equation V · B = 0, from which it follows,
and P2 = 2n 1mv 2/(y + 1) (15-77) in spherical coordinates, that B,{r) = (r0 /r) 2B,{r0 ). The azimuthal component
where in deriving equation (15-77) we have written in the spiral pattern, B~, can be computed in terms of B,. and hence the total
field can be calculated. We can then estimate the ratio of the energy density
p,M/ = (pifa,2)11/ = {p 1/y)11/ in the flow to the magnetic energy density-i.e., IX = {½mn11 2 )/(B 2/81t); one
finds that IX » 1 in the vicinity of the Earth's orbit, but IX is small at the base
Here the subscript 1 denotes preshock and 2 denotes postshock conditions. of the corona. This implies that-although the fluid motions will dominate
Combining equations (15-76) and (15-77) into a perfect gas law, we estimate the field at large distances, and will drag it along with the material-deeper
the postshock temperature to be in the flow, the magnetic field will dominate and will be able to drag the
material along with the solar rotation as seen in a fixed frame. Thus there is a
(15-78) region with r ~ r" in which the material is forced into corotation with the I
Sun, and a region r > r" where the mater.ial streams essentially radially. f\)
where we have assumed that the material remains fully ionized, and that the
Inuitively we would expect the transition to occur at the radius r,c where --.J
proton an,d electron temperatures equalize. Assuming y = i, we see that
IX = 1, and detailed analysis [(463; 659; 324, §111.15; 107, §3.7)] shows this O>
112 = ¼111 ~ 75 km s- 1 while T 2 = 3mv2/(32k) ~ 106 °K [which shows that
the postshock material is indeed subsonic, as expected from Exercise 15-3(b) conjecture to be correct. The speed of the hydromagnetic A/fven waves is I
for M 1 » 1]. 11..t = {B2/4np )¼, hence ex is nothing more than the square of the A/fvenic Mach-
The injection of hot, high-velocity material from stellar winds has signifi- number ex= (11/11.. )2 = M/; thus the flow corotates with the sun inside the
cant implications for the energy balance in the interstellar medium. The Alfvenic critical point, and streams radially outside it. The Alfven speed can
picture developed here is purposely quite simplified; more elaborate calcula- be directly measured in the vicinity of the earth's orbit; using the equation
tions have been made allowing for the effect ofan "interstellar wind" (arising of continuity and the inverse-square radial dependence of B, and demanding
that MA = 1 at the Alfvenic critical point, we find
from the Sun's peculiar velocity with respect to the interstellar medium),
magnetic fields, and thermal conduction [see, e.g., (498, Chap. IX)]. (15-79)

THE MAGNETIC FIELD AND BRAKING OF STELLAR ROTATION We may obtain a lower bound for r A by assuming 11(r,4) ~ 11(r 61 ), and sub-
stituting numerical values we find'" ~ 20 R 0 • This region is large, and thus
From direct observation it is known that the Sun has a magnetic field.
the effects of magnetic fields greatly increase the angular momentum con-
Because coronal material is completely ionized, it has an extremely high
tained in the material flowing in the wind.
electrical {as well as thermal) conductivity, and thus magnetic fields.are
"frozen" into the material (i.e., the charged particles cannot diffuse readily
Exercise 15-9: Verify equation (15-79) and the numerical estimate of rA•
across field lines), The large-scale expansion of the corona thus implies that
there will be a transport of the solar magnetic field into interplanetary space.
If the Sun did not rotate, the field lines would be drawn out radially. The Sun The angular momentum of a particle of unit mass at the equator of a star
is 10 = wrA 2 ; where w is the angular frequency of stellar rotation. Hence the
536 Stellar Winds I 5-2 Coronal Winds 537

total angular momentum lost from the stellar surface as a whole is mention of the many physical problems that have been addressed; for further
details see, e.g., (324) or (107) and the references cited therein.
(dL/dt) = -(4nr 2 nvm)/0 f;'2 cos = -t Jt/ 0
3
8 dB (15-80) First we might ask how well a one-fluid conductive model fits the data;
taking the model by Whang and Chang(667)as typical,one finds n = S cm- 3 ,
where we have accounted for the fact that (at latitude 8) /(8) = 10 cos 8. This
2
v = 260 km s - 1, and T = 1.6 x 10 5 °K at the radius of the earth's orbit.
angular-momentum Joss into the wind implies a braking of the star's rotation. Comparing with the observed values listed in Table 15-1, we see that the
We may estimate the rotational-braking decay-time t by writing (dL/dt) :::: model gives good agreement for the density, a~d the one-fluid temperature
- L/t, and expressing Las L = Iw, where I is the moment of inertia, from agrees well with the electron temperature, but 1s a factor of four larger than
which we find t = fl/(Jtr ,/). For the Sun, I = 6 x 10 53 gm cm 2, and using the proton temperature; the velocity is about 20 per~nt too low. The.fact
.ti= 2 x 10- 14 .,1(0 /yearandrA:::: 20R 0 ,wefindt :::: 10 10 years,which that the protons and electrons can have di~erent te~peratures is e~plamed
is comparable to the thermonuclear-evolution time-scale for the Sun. That by the very low densities, and hence collision rates, m the solar wmd near
is, if the solar wind maintains its present properties, we expect the solar
angular momentum to be significantly diminished during the main-sequence
lifetime of the Sun.
i:
r · these rates are so low that they cannot equilibrate the thermal energy
;he two separate components (moreover, the thermal_ velocitJ distributi~n
for both electrons and protons is observed to be amsotrop1c). Two-fluid
One of the striking features of the statistics of stellar rotation is the marked models have been constructed (618; 282) in which electrons and protons are
drop in the observed average rotation speed ( v sin i), as a function of spectral allowed to have distinct temperatures; an energy equation is written for each
type, for stars of spectral types F and later. Although it is possible that this component, including an energy exchange term of the form ¾~k_(Tp - T,),
decrease is associated with the formation of planetary systems (in the solar where v is a collision frequency calculated for Coulomb colhs1ons of the
system the Sun has only 2 percent of the angular momentum), it is also charged particles. Typical models of this type yield n ~ 15 cm - 3 , v ~
extremely suggestive that this is precisely the point where stars develop deep 250 km s- 1 , T, :::: 3.4 x 105 °K, and TP :::: 4.4 x 10 3 °K at r = re, The
hydrogen convection zones, and hence presumably have coronae and winds. densities are about a factor of two too large, the velocity is 20 percent too
It is thus very attractive to hypothesize (554) that all later-type ·stars lose low, the electron temperature is a factor of two too hi~h, and the prot~n
their angular momentum via magnetic braking in stellar winds. Time-scales I
temperature is a factor of 10 too low. Althoug~ the bas1~ ~oncepts _used m I\)
for such braking can be estimated by an examination of mean rotation speeds the two-fluid model do admit more physical reahsm than 1s mherent m a on_e
of stars in clusters whose ages can be determined from stellar-evolution con- ~
fluid description, the quantitative results are not impressive; in particular it
siderations,; when this is done, it is found that there is indeed a strong decrease ~
appears that the energy exchange between protons and ele~~rons must pro-
with age (362), and that the braking times may be as short as 5 x 10 8 years. ceed more efficiently than is predicted from Coulomb colh~1ons alo~e. T~e I
Although the stellar time-scale appears much shorter than that derived from exchange has been treated on the assumptions tha~ the pa_rt1cle-veloc1ty dis-
the present solar wind, it is known that chromospheric activity (and pre- tributions are isotropic and Maxwellian, and that magnetic field effects may
sumably also the activity of the corona-wind complex) decreases with age be ignored; possibly all of these assumptions are inadequate. Ho-:veve~, most
(675; 677); hence it is possible that in the early history of the Sun the wind of the discrepancy seems to be removed when the effects of v1scos1ty are
was stronger, and provided more efficient braking. Actually this discussion taken into account.
provides a good example of the fruitful exchange of ideas that occurs when Magnetic fields may play an important role in determining the nature of
solar properties are used to develop detailed physical models of a particular the solar-wind flow, through magnetic forces, by modification of transp~rt
phenomenon, which are then applied in a stellar context, where results of coefficients such as the conductivity, and through energy transpor~ a~d dis-
evolutionary significance on long time-s~ales can be inferred. sipation by hydromagnetic waves. Each of these effect~ leads to a s1gmfi~ant
increase in the complexity of the theory, and can appreciably alter t~e det~ded
DETAILED PHYSICS OF THE SOLAR WIND results. Worse, allowing major sources of momentu?1 of energ~ mput mto
The study of the solar wind by means of direct measurements from the wind throughout the flow (from any source, not JUS~ magnetH~), changes
satellites has greatly deepened our understanding of the physics of coronal the whole topology of permitted solutions of the equations, and mtroduces
flows. In an effort to fit the observations, numerous theoretical refinements many possibilities beyond those shown in Figure 15-1 _(307~. .
have been introduced to extend the basic model described above, and to The models described above all omit the effects ofv1scos1ty. When viscous
achieve greater realism. Space does not permit anything more than mere forces are included, the equations become of higher order, and the momentum
538 Stellar Wind, · 15-2 Coronal Winds 539

equation no longer possesses a singular or critical point (668). The earliest 1030 ergs s- 1 if T O varies from 106 °K to 4 x 106 °K. Thu~-a coronai ,wind
one-fluid viscous models (553; 668), using classical viscosity coefficients for acts as a very effective thermostat for controlling the coronal temperature.
a proton-electron plasma, were disappointing, for they yielded flow speeds Ifa wind is present, we demand that the temperature be compatible with the
and temperatures about a factor of two too small. A resolution of the diffi~ flow-i.e., TO $ (G.,U .m/4kR.). For the Sun, the numerical factor of 4 is
culty is achieved when magnetic effects upon the viscous stress tensor are replaced by 10, so one might expect
taken into account (660), and modern models including viscous terms (681)
have the right velocities, and raise the proton temperature (from the spuriously TO ::::; 0.l(GA .m/kR) (15-81)
low values found in the inviscid two-fluid models) to values close to those
observed. In main-sequence stars, (A .IR.) varies only by about a factor of two above
Even the most refined and accurate spherically symmetric models still or below the solar value, which suggests a similar restricted range for T 0
represent only very high-order abstractions of the real solar wind, for it varies, (less than the variation of photospheric temperatures!). In contrast, the energy
on time-scales of days, over wide ranges of all the physical variables. Prom- in the wind may vary widely (perhaps more so than the stellar luminosity).
inent individual features that appear are high-speed plasma streams; these For a giant or supergiant, (A .IR.) is much smaller than in main-sequence
often are observed to recur with a synodic solar-rotation period, and are, stars and, accordingly, T O should also be much smaller. For example, in an
therefore, the result of specific conditions in localized regions of t?e c~rona. M-supergiant, equation (15-81) suggests T 0 ::::; 4 x 104 °K. At such low
Further, there are energetic flare-produced shock waves that give nse to -temperatures conduction becomes ineffective, and one must invoke some
geomagnetic storms. These structures often interact in a complex way and, mechanism to heat the corona as a whole (665 ;666); quite possibly, significant
more and more, it seems that it is an oversimplification to view the solar wave dissipation occurs throughout the corona. Similarly, the velocity of
wind as a smooth flow upon which "atypical" structural features are super- the winds ultimately approach some significant fraction of the escape velocity,
posed; rather, these complex structures are, in some sense, the ":ind itself. • and thus we can expect
Similarly we are coming to understand that the wind does not anse from a
Va, ;;::: (2G.,({ .1R.) (15-82)
single smooth coronal condition, but that it may be produced in hi~hly
specific regions of markedly differing properties, and can_ be substant~ally Again this implies little variation along the main sequence, and implies that
modified by interaction with magnetic fields and by rapid (nonspher1cal) supergiants have quite low wind velocities. The picture developed above I
divergence from a limited initial volume. At some point the problem of the depends on the assumption that the wind is subsonic at the coronal tempera- N
time-depend~nce of a full three-dimensional model must be solved; only for --.,_j
ture maximum. If, however, the critical point in the wind lies below this
the solar wind do we have data that require (and, reciprocally, permit I) such maximum, the run of temperature with distance is no longer monotonic and CD
solutions, but it is clear that the results will have important implications for the properties of the flow may change substantially. I
stellar winds as well. To construct detailed models or apply existing computations [e.g., (201)],
we need to specify the relevant coronal conditions, such as n0 and T 0 , from
STELLAR CORONAE AND WINDS theoretical calculations-or to specify one of these quantities and~a flow
parameter such as the mass-loss rate, which can sometimes be determined
As mentioned above, it is reasonable to suppose that all stars having observationally. The computation of coronal conditions is exceedingly
hydrogen convection zones should also have coronae and winds. Further, difficult. The basic technique is to calculate the acoustic energy flux created
many stars clearly have much more massive flows than the Sun, for the winds -in the convection zone [ see, e.g., (394; 395; 525; 601}], compute the dissipation
are sufficiently optically thick in some spectral lines to give rise to displaced of this flux in the coronal material, and (taking into account losses by radia-
lines or P-Cygni features. Needless to say, we know less about stellar winds tion and the energy transported by conduction) obtain a temperature-density
than about the solar wind, .and much work remains to be done in this rapidly structure in a model corona [see, e.g., (192; 377; 378; 630; 631; 632)]. Each
developing field; nevertheless, a number of interesting results emerge even ·of these steps requires use of an uncertain theory, and of necessity many
from very simplified calculations. approximations must be made, so a high degree of reliability cannot be
Parker emphasized (498, Chap. XV) that the amount of energy consumed assigned to the final results. In particular, a good fit is not obtained to observed
by coronal expansion rises very rapidly with increasing coronal temperature; solar coronal properties [see Fig. 21 of (192)] in all cases. But if we simply
for example, for the Sun, the energy consumption ranges from 10 27 to 3 x accept the calculations, however uncertain, at face value, it appears that for
540 Stellar Winds 15-3 Radiation Hydrodynamics 541

main-sequence stars the maximum acoustic flux is found for stars near tion. In this way we obtain the equations of radiation hydrodynamics, which
spectraltypesF0;forthesestarsTO i::-s 4 x 106 °KandnO i::-s 3 x 10 1O cm- 3 describe the coupled flow of the gas and radiation. In the applications of
(192). One would expect substantial coronal winds in these stars. If T 0 and interest here, it will be assumed that the flow velocities v « c, so that the
n0 are regarded as given, then the calculations cited previously can be used material particles can be treated nonrelativistically; in certain other contexts
to derive wind models. (e.g., in supernovae or thermonuclear explosions), this assumption may not
Consideration of the energetics of stellar coronae and winds is also quite be valid. Despite the restriction to v « c, the fact that the photons have
instructive
• (290). A corona is heated by the mechanical flux Fm delivered to velocity c leads to subtleties in the way the radiation and material terms
1t, and loses energy via radiation and conduction, and in the kinetic energy interact, and it is a nontrivial task to obtain equations that describe the
of the wind. Analysis shows that, for a given coronal pressure, the rate of interaction fully consistently. In the end it is simplest and most reliable to
energy loss by mass-ejection and by conduction increases with rising tem- develop them in a relativistically covariant form from the outset (621; 524;
perature, while that by radiation decreases. Therefore, for a given pressure, a 664; 135). These equations may subsequently be simplified to retain only
temperature exists at which the total energy loss from the corona is a min- terms of order (v/c), and to omit terms of O(v 2/c 2 ) and higher. It must again
imum. Further, losses from all three processes increase with increasing be noted that, as was true in our earlier treatment of the comoving-frame
pressure, so for minimum-flux coronae there is a monotone relation between transfer equation, it is not sufficient to apply a unique Lorentz transforma-
coronal pressure and the energy flux required to maintain the corona and tion, because the fluid velocity is, in general, a function of position and time.
wind. This suggests that a given Fm determines a unique coronal temperature As before, we consider transformations from a set of uniformly-moving
and pressure, and hence wind. (Actually it remains to be demonstrated that, frames that instantaneously coincide with the moving fluid.
if one has a given minimum-flux corona, and arbitrarily changes Fm• then The radiative contributions to the equations can, in principle, be written
the corona necessarily adjusts to a new minimum-flux configuration.) The in terms of either laboratory-frame or fluid-frame quantities. Although in
competition among these processes establishes three basic classes of coronae. some ways it is easier to write down the equations of radiation hydrodynamics
(1) For low values of Fm the main losses are by conduction and radiation. in a stationary frame [see, e.g., (521; 551; 692)], they are actually ofa simpler
An example is the solar corona-wind complex where only about 10 percent form when written in the comoving frame of the fluid, for the radiative terms
of the mechanical flux delivered to the corona goes into the wind. (If the wind are most easily evaluated in that frame (621; 135). In the past this approach I
arises in geometrically localized regions of the corona-e.g., coronal holes..._ was not widely employed, because one must be able to solve the comoving- I\)
the local conversion efficiency may be much higher.) (2) At intermediate frame transfer equation to calculate the necessary radiation-field quantities. --..J
values of Fm the losses are mainly from radiation and mass ejection. An ex- As was shown in §14-3, the comoving-frame transfer equation is easily solved (()
ample is the wind in the A2 supergiant oc Cyg where losses into the wind are with present-day techniques (and is, in some ways, simpler than the fixed- I
dominant, while those by r~diation are significant, and those by conduction frame equation, which is complicated by anisotropies in the absorption and
are negligible. (3) At very large values of Fm, mass loss totally dominates. emission coefficients arising from Doppler shifts). We shall, therefore, con-
For early-type stars the rate of mass ejection is as high as 10- 6 to 10- 5 .,11 / centrate primarily on a comoving-frame formulation, but it will also be
0
year, and exceedingly high coronal temperatures (which appear to be excluded shown that these equations are consistent (to order v/c) with the fixed-frame
by the observations) would be required to drive the winds; here the winds equations. Only one-dimensional spherically-symmetric flows will be treated
are driven instead by radiation forces (see §15-4). in this book; an extensive collection of formulae for three-dimensional flows
in various coordinate systems can be found in (521, Chap. 9).

15-3 Radiation Hydrodynamics


THE MATERIAL STRESS-ENERGY TENSOR AND
In both the atmospheres and interiors of stars, there exist intense radiation THE RADIATING-FLUID EQUATIONS OF MOTION
fields that can influence heavily the momentum and energy gains and losses In the Cartesian coordinate system (x1, x2, x 3, x 4 ) = (x, y, z, ict), a co-
of the material, and hence its motions. To study the dynamics of a flow variant formulation of the equations of motion and of energy conservation
occurring in such a situation, it is fruitful to consider the fluid as consisting for the material alone yields a system of equations of the form
of both material particles and photons, and to calculate the contributions
of both types of particles to the equations of motion and of energy conserva- (15-83)
542 Stellar Winds 15-3 Radiation Hydrodynamics 543

where P is a four-vector force (the Minkowski force) and T"fl is a tensor Here h is the specific enthalpy of the gas, and (p 0 h/c 2 ) gives the mass equiv-
describing the momentum flux (stress), momentum density, and energy alent (per unit volume) of the total energy contained in the microscopic
density of the material. Similar equations were written in §14-3 for the radia- motions of the gas.
tion [cf.equations (14-118) through (14-122)], and by analogy with equation Noting that in the nonrelativistic limit n 11 = pv1v1 + p /;11, while G1 =
(14-120) we might expect p/J to be of the form pv1, we may hypothesize, by analogy, that
(15-90)
T =( n icG) (15-84)
icG -E where J«/J denotes the standard Kronecker symbol. It is obvious that equation
(15-90) satisfies requirement (1) stated above. The individual components
where Il, G, and E are suitable covariant generalizations of the momentum may be written more specifically as
flux tensor, the momentum density vector, and the total energy density of
the material, respectively. We place three demands upon T: (1) that it be yiJ = PoooV 1V1 + p oiJ = p 1v1v1 + p l,lJ (15-91a)
written in terms of scalar invariants and four-vectors, so that it is covariant; T'4 = y41 = iC')'PoooV1 = icpiv' (15-91b)
(2) that it reduce to the correct limit in the frame at rest with respect to the 44 2 2 2
fluid; and (3) that it yield the correct nonrelativistic limit in the laboratory
and T = -C 'l' Pooo + P = -C P1 + P (15-91c)
frame. where i = 1, 2, 3, j = 1, 2, 3, and p 1 = y2p 000 . Notice, now, that in the
First, we may define the proper time, a four-scalar (invariant) as frame of the fluid itself, where v = 0, T becomes diagonal, with T 11 = p =
(Tiu) 0 , and T
44 = -(p c 2 + p e); this gives the correct static pressure and
0 0
the correct energy per unit volume of tpe fluid (including the material rest-
Clearly dt reduces to dt as v --. 0. Then we define the cbntravariant four- energy). Thus requirement (2) stated above is satisfied. Next, let us examine
velocity as the nonrelativistic limit (v/c « 1) of T. Note first that, if p 0 is the mass
V" = (dx'/dt) (15-86) density in the fluid frame, then [remembering that a volume element trans- I
forms as dV = y- 1 dV O (Lorentz contraction)], the density seen in the I\)
Noting from equation (15-85) that (dt/dt) = (1 - v2 /c 2 )-t = y, where, v laboratory frame will be p = yp 0. Thus, expanding each element ofT, and ro
is the ordinary velocity v2 = [(dx/dt) 2 + (dy/dt) 2 + (dz/dt) 2 ], we have retaining only the leading terms, we see by inspection that T 11 and T14 0
(or T 41) reduce to the correct momentum-flux tensor and momentum density, I
V 1 = (dx 1/dt)(dt/dt) = (1 - v2 /c 2 )-f(dx1/dt) while
= ')'V 1, = 1, 2, 3)
(i (15-87a)
T44 = -y2(poc2 +Poe+ pv2;c2) ... -y(pc2 + pe)
V4 = ic(dt/dt) = icy (15-87b) ~ -(pc 2 + ½pv 2 + pe). (15-92)
Note in passing that V«V" = -c , a world scalar (which is invariant under
2
which is the nonrelativistic energy density (including the rest-mass energy).
any arbitrary coordinate transformation, as can be readily shown from the
Thus T satisfies requirement (3) stated above.
transformation properties of covariant and contravariant vectors). Next,
The action of external forces upon the fluid are expressed in terms of a
if p0 , e, and pare the invariant mass density (i.e., the particle number density
four-force (per unit volume) F" whose spatial components give the rate of
times rest-mass per particle), the specific internal energy, and the pressure
increase of the momentum of the matter per unit volume and whose time
measured in the rest-frame of the fluid, then the equivalent total mass
component specifies the rate of increase of energy per unit volume. The
density is
expression V,,F" is a scalar invariant, and hence may be evaluated in any
Poo = PoO + e/c )
2
(15-88)
frame. In particular, we may evaluate it in the frame comoving with the
Also, following Thomas (621) we define fluid. In this frame

Pooo = Poo + p/c 2 = PoO + h/c2 ) (15-89) (15-93)


544 Stellar Winds 15-3 Radiation Hydrodynamics 545

In the absence of general-relativistic effects, the derivative on the righthand Now, subtracting (c 2 + e + p/p 0 ) times equation (15-96) from equation
side will be zero unless there is a change in the proper mass density arising (15-99) and writing (D/Dt) = V•(a/ax"), we obtain
from a chemical energy release or from a conversion of matter to energy by,
say, a thermonuclear reaction. Thus V,.F" = 0; note that this relation implies Po(De/Dt) - (p/p 0 )(Dp 0 /Dt) = - V.if (15-100)
that in the comoving frame (F4 ) 0 = 0, a result that will be used below.
Exercise IS-JO: Fill in the missing steps between equations (15-99) and (15-100).
The equations of motion for the material alone are given by equation
(15-83), where y«fl is defined by equation (15-90). If R•P denotes the radiation
As the lefthand side of equation (15-100) is evaluated following the fluid
stress-energy tensor [cf. equation (14-120)], then the equations of motion
flow, the righthand side can be evaluated in the comoving frame. In this
for the complete fluid (matter plus radiation) are
frame V 1 = 0 (i = 1, 2, 3), and V 4 = ic instantaneously, and from equation
a(r«P + R"')/axP = F" (15-94) (14-121)

or (aT•fl /axP) = F" + g• (15-95)


g
4
f
= (41ti/c) (x.°J.° - ri.°) dv (15-101)

where g" is given by equation (14-121). In addition, we may write an equation where advantage has been taken of the isotropy of the absorption and
of continuity for the material particles alone (the photons are massless); emission coefficients in the comoving frame. It is thus clear that equation
demanding number conservation, we have (15-100) reduces to equation (15-97), which is then, in fact, the correct gas-
energy equation for the material as it interacts with the radiation field.
(15-96)
THE FLUID-FRAME MOMENTUM EQUATIONS

THE FLUID-FRAME ENERGY EQUATION To obtain an equation of motion in the comoving fluid frame, one
might argue that the forces exerted by the radiation field on the material I
Ifwe were to assume that the first law of thermodynamics remains valid should be added to the usual body forces exerted by, say, gravity. The net
N
for the combined material and radiation fluid in the comoving frame of the force exerted by the radiation, when calculated in the comoving frame, is (X)
material, then the gas-energy equation [cf. equation (15-24)] would need to given by [cf.equation (14-121)] __.
be modified only to account for gains and losses of energy (per unit volume
per unit time) by the material from interaction with the radiation field. Thus (15-102)
we could write
where again x.° and .'F,0 are evaluated in the comoving frame, and advantage
p0 (De/Dt) - (p/p 0 )(Dpo/Dt) = 41t Jo"' CJ..°J.° - ri.°) dv (15-97) has been taken of the isotropy of x.° and ri.°, Thus from equation (15-21)
we expect that
where (D/Dt) denotes the Lagrangian time-derivative, following the motion
of the material, and x.°, ri.°, and J.° are evaluated in the comoving frame,
which is at rest with respect to the gas. It will now be shown that this a priori
p(Dv/Dt) = -Vp f
+ F + c- 1 x.°.'F.°dv (15-103)

expectation is right, and that equation (15-97) is rigorously correct (621; To verify equation (15-103) we examine the ith component of equation
135). Forming the dot product of the equations of motion (15-95) with - V. (15-95), which, using equations (15-91), may be written
[and using the explicit form for T•P given by equation (15-90)] one finds
[a(p 1viv1)/ax1] + [a(p 1v1)/at] + (ap/ax 1) = F 1 + g1 (15-104)
• a(Pooo Vfl) fl ( av•) • ap .
-(V.V) axfl - PoooV V. axfl - V ax•= - V.F" - V.if (15-98) av1 J av' I 8(p1 vi) I apl ap I I
or Pi at + Pt axl + V 8T + V at + ax' = F + g (15-105)
But v,.v• = -c , so V.(av•;axP) = a(tV.V")/axP
2
= 0 and, recalling that
v.F' = 0, equation (15-98) becomes But the time-component of equation (15-95), multiplied by v1 is

(15-99) v'[a(p 1v1)/ox1] + v1[o(p 1 - c- 2 p)/ot] = v1(F4 + g4 )/(ic) (15-106)


546 Sre/lar Winds 15-3 Radiation Hydrodynamics 547

Hence by subtraction Consider first the momentum equation (15-109), which for a spherically
1 1 1 symmetric flow with a gravity force becomes
av' ov op v op v 4
P1-;--
ut
+ P1vl-;;-J
uX
+!Ii+
uX
2 ;;-
C ut
= F 1 + g' - -:--(F
IC
+ g4 ) (15-107)

Writing (D/Dt) = v«(o/ox«) and p = yp000 , equation (15-107) becomes


p(Dv/Dt) = - Vp - (v/c 2 )(op/ot) + F. + g - v(F 4 + g4 )/ic (15-108) (15-111)

Again, the lefthand side of the equation is evaluated following the fluid flow, where use has been made of equation (1-43b) for V · P. To obtain an equation
so the righthand side can be evaluated in the comoving frame of the fluid correct to O(v/c) the full transformations of equations (15-1 lOa) and (15-110c)
(in which v is instantaneously zero). We then find equation (15-103), as must ~e _used. However, even in the_ free-flow limit, SF 0 ~
cER 0 or cpR 0 ;
expected, when g is evaluated using equation (14-121). hence 1t 1s clear that all of the terms m SF in equation (15-111) are already
It is convenient to rewrite equation (15-108) with g« replaced by at most only O(v/c) (remembering that we consider M 6.x/v). Thus the ~
-(oR«PjoxP) [cf. equation (14-120)] to obtain additional terms in v in equation (15-llOb) will introduce terms ofO(v 2 /c 2 )
and can be omitted, so that it is sufficient to write SF = fF 0 • Substituting
p DvDt + Vp + (v. p + .!_c oF)
2
ot
_ !__ (oER + V. F)·
c ot 2
for the inertial-frame radiation field as indicated, we thus obtain
'Dv op [ 1 as; 0 v oSF 0 opR 0
p-+-+ ---+---+--
= F-; (!~) + F]
2 [ v· (15-109)
Dt 2 2
or c ot
0
c or
0
or
0
+ (3pR - ER ) + 2§2 (av+£)]= G.,Hp (15-112)
where the rate of increase of material energy per unit xolume has been r c Or r -----;:i-
written as F4 = -(v · F)/ic. In equation (15-109) the radiation-field quan-
tities are now expressed in an arbitrary (laboratory) frame. Notice now that, if Exercise /5-11: Verify equation (15~112).
I
we are concerned with describing only a nonrelativistic fluid flow the term f\)
from F 4 is O(v 2/c 2 ) with respect to F and hence may be omitted. 'Likewis~. In view of the comoving-frame transfer equation (14-134b), all of the terms 0)
ifwe compute the.fluid flow (as opposed to the time-evolution of the photons), in the square brackets collapse to yield f\)
the characteristic times involved will be !lt ~!lx/v where v is the fluid
~- p(Dv/Dt) + (op/or) = -(G.,Hp/r2) + (41t/c) f x.°Hv° dv (15-113) I
velocity, and the term in (op/ot) is clearly O(v 2/c 2 ) compared to Vp and may
be omitted. Finally, ER ~ PR; hence (v/c 2 )(oER/ot) is O(v 2/c 2 ) compared to It is thus apparent that the comoving{rame tra·nsfer equation derived in
V · P and may also be omitted. All of these terms will be dropped henc('.{orth.
Chapter 14 renders the inertial-frame and fluid-frame equations of motion
Note that the term -v[(oERfot) + V · F]/c 2 in equation (15-109) accounts
fully consistent to order (v/c). · .
for the rate of change of the proper mass density of the matter resulting from
Let us now consider energy conservation (ignoring thermal conduction)
its interaction with the radiation field.
as expressed by equation (15-28). The time derivative operates on the energy
per unit volume, and to this term we should add ER; the divergence operates
THE INERTIAL-FRAME EQUATIONS
on the energy flux, and to this term we should add F. Hence
Most treatments of radiation hydrodynamics formulate the equations [o(pe + ½pv 2 + ER)/ot] + V . [(pe + ½pv 2 + p)v + ~] = V • F
in an inertial laboratory·frame. Let us now show [working to O(v/c) only]
that the inertial-frame and comoving-frame equations are consistent. This (15-114)
may be done by expressing the inertial-frame radiation-field quantities in
To obtain a gas-energy equation, we subtract out the mechanical work terms
terms of their comoving-frame counterparts via equation (14-124), i.e.
obtained by taking the dot product of v with the momentum equations
ER= ER 0 + 2c- 2 v · F 0 (15-ll0a) (15-109). Note that the two terms in (v/c 2 ) will then become O(v 2/c 2 ), as will
F = F 0 + ER 0 v + v · P0 (15-ll0b) c- 2 v · (o~ /ot) for time-intervals .1.t ~ .1.x/v; hence
P = po + c- 2 (vF 0 + F 0 v) (15-ll0c) p[D(½v 2 )/Dt] + (v · V)p + v · (V · P) = v · F (15-115)
548 Stellar Winds 15-4 Radiatively Driven Winds 549

and by subtraction from equation (15-114) we find Finally, it is useful to combine the momentum and gas-energy equations
into a total energy equation that displays the effects of radiative momentum
. [a(pe + ER)/at] + V · (pev + F) + p(V · v) - v · (V · P) =0 (15-116) and energy inputs in a transparent way. Let qR be the rate of n;;t gain of
energy density by the material from the radiation, and let fR be the force
In view of equations (15-17) and (15-13), equation (15-116) reduces to per unit volume exerted by radiation. Then, from equations (15-97) and
(15-13),
p(De/Dt) - (p/p)(Dp/Dt) + [(aER/at) + V· F - v · (V · P)] =0 (15-117) p(De/Dt) + p(V · v) = qR (15-120)

It is clear that, in transforming to comoving-frame quantities, we must now and from equation (15-103), by forming the scalar prpduct with v,
carry all three terms of equation (15-110b) for F; but manifestly the second
terms of equations (15-ll0a) and (15-ll0c) will produce terms of O(v 2 /c 2 ) (15-121)
in time-intervals characteristic of the fluid flow; hence these terms may be
By addition of equations (15-120) and (15-121), and use of equation (15-17),
omitted, Thus we calculate V · F = V · {F 0 + ER 0 v + v · P0 ), and note
that, for spherically symmetric flow, [o(½pv 2 + pe)/ot] + V '[(½pv 2 + pe + p)v] = qR + V' (f + fR)
(15-122)
0
For a steady (o/at = 0) spherically symmetric flow, integration of equation
=V -(a
ER-)
or
+ ERO (av
-+-
ar
2v)
r
(15-l 18a) (15-122) with the usual gravity force gives

(41tr 2 pv)[½v 2 + e + (p/p) - (G..H/r)]


V ' (v ' P0 ) = r- 2 o(r2 vpR 0 )/or
= v(apR 0;ar) + PR 0 [(av/ar) + 2(v/r)] (15-118b) + 47t J,«> (qR + v · fR)r 2 dr = E = Constant (15-123)

which is analogous to equation (15-59) with added radiative terms. For later
Further, (15-118c)
work it will be convenient to denote the first set of terms by E0 and to write
so, expanding the term in square brackets in equation (15-117), we have (15-124)

15-4 Radiatively Driven Winds


+ ER (av
0
ar + r2v) + PR 0 av
ar
O
+ (ER - PR o>(v)r Observational evidence gathered over the last decade has demonstrated
compellingly that rapid mass loss via transsonic winds is ubiquitous through-
= (aER
ot
0
) + v (aER
or
0
) + (a.Far 0
+ 2?
r
0
)
out the high-temperature, high-luminosity portions of the Hertzsprung-
Russell diagram. The basi9 theoretical framework within which such flows
are explained involves winds driven by momentum input into the gas from
+ (ER O + PR O) OV
or + (3ER O - PR O) (V)
r the intense radiation fields of these luminous stars. Although the basic out-
lines of the theory seem fairly well established at present, this is a very active
and rapidly developing field, and a number of important questions remain
(15-119)
open, awaiting future study. We shall, therefore, confine attention to ad-
mittedly idealized models that illustrate the basic physics; detailed fitting
by virtue of equation (14-134a). Equation (15-117) is thus identical to the of computed models to observed data is only beginning, and the reader
comoving-frame equation (15-97). should turn to the current research literature to follow these developments.
550 Stellar Winds 15-4 Radiatively Driven Winds 551

OBSERVATIONAL EVIDENCE FOR TRANSSONIC WINDS envelope having a density profile consistent with steady outflow of the ma-
IN EARLY-TYPE STARS terial (171; 241; 277; 685).
The ground-based data are of great importance for they, in principle,
A large number of characteristic features (e.g., P-Cygni profiles, emission provide information in the region where the flow passes through the sonic
lines, line asymmetries) in the visible spectra of many O and early B stars velocity. The hydrogen Ha. line and the He II A.4686 line are among the
(particularly supergiants, Of, and WR stars) have long provided evidence strongest lines in the visible spectrum of OB stars, and hence usually provide
that these objects have extensive envelopes, and that the material being the first indications of atmospheric extension and expansion by coming into
observed in the lines is flowing outward from the stellar photosphere [see, emission. Extensive observational surveys have been made of both lines. Ha.
e.g., (261, Chap. 10)]. The case for actual mass loss was not unequivocal, is found (537) to show emission in luminous B-stars of all types from BO
however, because the observed velocities measured from the short-wave- to A3. There are well-defined lower limits to the luminosity at which Ha.
length edge of the absorption in P-Cygni features (typically 200-400 km s - 1 ) emission first becomes conspicuous, namely Mu ~ - 5.8 (M bol ~ - 8.8)
did not exceed the surface escape velocity near spectral type BO, and M" ~ - 6.8 (M bol ~ - 7.3) near spectral type AO,
(15-125) and there is a definite relation between the net emission strength of Ha. and
the stellar luminosity. There is clear evidence for differential expansion of
which is of the order of 1000-1500 km s- 1 for a main-sequence O-star, and the atmosphere for stars about 0.5 mag less luminous than those that show
600-900 km s- 1 for an OB-supergiant. The data obtained from ground- conspicuous Ha. emission. For the O-stars (177) there is generally a close
based observations suffer from a fundamental limitation: all of the lines correlation between the emission strength of He II ).4686 and that of Hrx,
observed are subordinate transitions that arise from levels with high excitation and emission at either line is an indicator of an extensive envelope around
potentials, and hence with small populations outside the regions of high the star. Further, it is found that the luminosity at which Ha. weakens, or
temperature and density. The column density of absorbers in these lines is comes into emission, is M" ~ -6 (Mbo 1 ~ -9). The Hrx profiles typically
therefore quite small, and one observes only the innermost layers of the exhibit P-Cygni characteristics, or show an extended blueward absorption I
envelope, just outside the photosphere. wing indicating expansion; typical velocity widths are of the order of I\)
In contrast, complementary information is obtained from the ultraviolet· ± 600 km s - 1 • As the expanding envelope of an OB star becomes more ex- ro
region of the spectrum accessible to observation from space, which contains tended and dense, a definite series of other spectroscopic effects are mani- ~
resonance lines arising from the ground states of the dominant ionization fested, and a useful qualitative assessment of the state of the atmosphere can
I
stages of abundant light elements. The column densities in these lines are so be obtained by an examination of the particular effects that do appear (328).
large that one may sample the outermost parts of the envelope. Decisive Measurements of the radial velocities (in the usual spectroscopic sense)
direct evidence of mass loss was thus first provided when Morton (467) indicated by different lines of various ions in Of spectra (326;327), show some
discovered displacements corresponding to outflow velocities in the range interesting features. (1) There is a velocity progression with line-series
1500-3000 km s- 1 in the P-Cygni profiles of the ultraviolet resonance lines (e.g., the Balmer lines), with the strongest lines showing the largest velocities
of Si IV ,U402.8 A and C IV ).1549.5 A in spectra obtained from rockets of approach; this indicates an acceleraiing outward flow, for the outermost
[see also (468; 469; 470; 129; 600; 579; 580)]. Combining the ultraviolet layers are seen in the strongest lines. By theoretical fitting of the observed
and ground-based data, we infer a transsonic flow in the expanding envelope. profiles it is possible, in priniciple, to infer the variation of velocity with
The measured resonance-line velocities are, in fact, only lower limits on the radial distance from the star. If such information can be obtained reliably,
actual terminal velocity of the flow, because variations in the ionization it is of great value, for it can provide important constraints on possible
equilibrium may cause the absorbing ion to vanish at some level in the en- theoretical models in the crucial transsonic flow regions. For one star
velope (or the material may just become optically thin and produce features (HD 152236), it was found (327) that the velocity rises sharply outward, from
below the detection threshold). Recently, results of great accuracy and ~ ~
a value near the sound speed ( 25 km s c. 1 ) to v 300 km s- 1 in about 0.5
sensitivity have been obtained for ultraviolet OB spectra from the orbiting stellar radii above the photosphere; for two other stars this rise is much more
observatory Copernicus (589). gradual, with velocities of a few hundred km s - 1 being attained at two or
Further evidence for mass loss is provided by infrared and radio continuum three stellar radii above the photosphere. We shall note the significance of
observations (of several OB and WR stars), which are most readily inter- these results later in this section. (2) There is a clear correlation of larger
preted in terms of free-free emission from an extended, optically-thick veJocities with smaller lower-level excitation potentials· for the Jines. If the
552 Stellar Winds 15-4 Radiatively Driven Winds 553

excitation were presumed to be in LTE, this correlation would imply that are all clearly larger than the surface escape velocity (and hence much larger
the temperature decreases outward in the envelope (as would be expected than v••• at great distance from the star), and range from 300 km s - 1 to
if the envelope is in radiative equilibrium, or is expanding adiabatically). Of 3500 km s - 1 • There does not appear to be a significant correlation of v.,,
course, LTE is not likely, and probably the correlation reflects increasing with stellar temperature, luminosity, gravity, or rotation velocity.
dilution of the radiation field (hence lower radiative excitation rates) or A wide variety of ions are observed, ranging from Mg II and C II in the
decreasing densities (hence lower collisional excitation rates). Both sets of coolest stars, through C IV, Si IV, and N V in the hotter stars. Lines arising
results mentioned here urgently need to be refined with trustworthy diag- from excited states (e.g., N IV Ji.1718.5 at 16.1 eV, or He Ihl.1640 at 40.6 eV)
nostics, based on careful solutions of the coupled transfer and statistical show lower velocities (470), again indicating decreasing eicit1;1tion with in-
equilibrium equations, for various flow models. Finally, it should be men- creasing distance from the star; as noted above, this· probably reflects de-
tioned that broad, faint emission features have been observed (678) under- creasing densities and radiation fields. In several stars the O VI Ji.1032 and
lying the relatively narrow, bright, emission lines of He II Ji.4686, C III Ji.5696, Ji.1038 lines are observed (533; 589). These lines are unexpectedly strong, and
and N III Ji.4634-40 in some Of spectra. These features have total velocity indicate a higher density of 0 5 + than would be estimated from the color
widths up to 4000 km s- 1, and if real (they are not seen by all observers) temperature of the stellar radiation field. If it is assumed that the ionization
could possibly result from emission originating in the extensive, rapidly- equilibrium is established by collisions (coronal case), the temperature de-
expanding outer envelope of the star. rived is about 2 x 10 5 °K, which is taken as evidence for coronal heating in
A large body of ultraviolet data has now been accumulated, thanks to these stars. On the other hand, an upper bound of about 3 x 10 5 °K on T
the possibility of making long-term observations from the satellite Copernicus follows from observations of lower ionization stages (e.g., C III and N III)
(589). Because the observed resonance lines are intrinsically strong and easily in the flow, and from comparisons of upper limits of X-ray fluxes to Ha
detected, they provide extremely sensitive indicators of stellar winds. From emission strengths (145). We shall see below that temperatures in this range
an examination of ultraviolet spectra of 47 0-, B-, and A-stars (see Figure are too low to drive the flow by a coronal-wind mechanism and produce
15-4 for an example), it is found that mass loss occurs over a much wider terminal speeds in the observed range.
range of temperature and luminosity than indicated by effects seen in the Mass-loss rates have been estimated from the observed line-strengths (468)
spectrum visible from the ground. Essentially all stars with luminosities o,
and profiles (329) for the Orion supergiants e, and COri, yielding values of
greater than 3 x 104 L 0 (a reduction of a factor of 15 relative to the ground- about 1 to 2 x 10- 6 .,((0 /year. Detailed fits to line profiles for CPup (O5f) r
based limit) are found to show mass loss. The observed terminal velocities yield a mass-loss rate of 7 (± 3) x 10- 6 .,I/ 0 /year (383). (
(
BASIC DYNAMICS OF RADIA TION·DRIVEN WINDS

The first question that arises is whether the flows observed in early-type
stars can result from coronal expansion of the kind considered in §15-2. As
recognized by Lucy and Solomon (404), the answer to this question is prob-
ably negative. To begin, the OB stars are thought not to have extensive con-
vection zones, and are not expected a priori to have coronae. But even if they
1180 1200 1220 1240 1260 did have coronae for some reason, then equations (15-59), (15-62), and (15-64)
;.(A)
imply that the critical-point temperature T. required to drive the flow must
FIGURE l5-4
satisfy the relation 2kT. ~ ½mv.,, 2 • Adopting v.,, ~ 3 x 10 3 km s- 1, we
Far-ultraviolet scan of the spectrum of CPup (05f) obtained from find T ~ 3 x 10 7 °K, a value that is completely excluded by (a) the absence
the satellite Copernicus. Ordinate gives flux (in arbitrary units), and of soft•X-ray emission from the stars, and (b) the presence of lines from ions
abscissa gives wavelength in A. The positions of the C III ).II 75.7, such as C IV, N V, and Si IV (observed to exist in the flow at velocities from
Si Ill .1.1206.S, HI ).1216, and NV .1.1238.8, 1242.8 lines are marked
with vertical arrows. The smooth upper curve is an estimated
~½v.,, to v.,,), which would be destroyed by collisional ionization at tem-
"continuum" level, drawn for illustrative purposes only; note the
peratures greater than about 3 x 105 °K.
marked P-Cygni character of the C Ill and N V lines. From (589), We must therefore seek an alternate mechanism to drive the wind. Lucy
by permission. and Solomon (404) suggested that this mechanism is direct momentum input
554 Stellar Winds 15-4 Radiatively Driven W:nds 555

into the gas through the absorption of radiation by the strong resonance lines the maximum force that can result from a single line, we assume that the line
observed in the ultraviolet spectrum. The momentum input results when intercepts unattenuated continuum radiation-Le., F. = F. = B.(Terr), Then
photons are absorbed by certain ions from the stellar radiation field, which an upper limit to the acceleration of the material by a single line of an atom
is strongly outward-directed, and then scattered isotropically. Because the of chemical species k, in excitation state i, of ionization stage j, is
emission process is isotropic, it produces zero net change in the momentum
of the material; hence there is a net gain of outward momentum from the YR
0
= (1t 2 e2 /mc 2 )f11B.(T.rr)(n11k/NikHNJ1.INk)(r1.kX/mH) (15-129)
incident radiation field. The absorbing ions are thus accelerated radially, and
where n1Jk is the population of the particular level, Nik is the total number of
they then suffer collisions with all other particles in the medium; in this way
atoms in all excitation states of ionization stagej, Nk is the total number of
the momentum gained by the particular ions absorbing the radiation is
all atoms and ions of species k, r,.k is the abundance ·or species k relative to
shared with the other atoms in the gas, and accelerates the material as a
hydrogen, and X is the mass fraction of the stellar material that is hydrogen.
whole. The outward acceleration experienced by the gas is then
Lucy and Solomon considered the C IV line at ).1548 A, and adopting f =
gR = (41t/cp) f0 x.°H.° dv
a:i (15-126) 0.2, Terr = 25,000°K (to maximize 8 1.), r1.e = 3 ~ 10- 4 , X = I, and
(n1Je/N1c) = 1, they found
where x. 0 denotes the absorption coefficient per unit volume from all sources
(15-130)
(continua, electron scattering, and lines); H. 0 is the incident flux.
The outward radiative acceleration is to be compared with the inward For a typical O-supergiant, log g :::::: 3; hence the upper limit on the force
acceleration of gravity, g = (G.,H/r 2 ); if g is everywhere greater than gR, then (obtained when Nie/Ne = 1) from even this one line exceeds the force of
the atmosphere remains in hydrostatic equilibrium and does not expand. We gravity by a factor of 300.
therefore must investigate the circumstances under which gR will exceed g; Of course,. the estimate just derived is (purposely) a gross upper limit,
for convenience we shall define because the carbon atoms in the photosphere of the star produce a dark
(15-127) absorption line in which F. « F•. To account for this, Lucy and Solomon
solved the transfer equation approximately, and found that above a certain
I
In O-stars the continuous opacity is dominated by electron scattering in critical level in the atmosphere the radiation force, computed using equation
I\
those spectral regions where most of the flux emerges. Pure Thomson (15-126) for the C IV ).1548 line, still exceeded gravity. [Interestingly, similar 0
scattering is ,frequency independent, and the resulting r can be written results had also been found in standard plane-parallel, static, model-atmo-
C
immediately as sphere calculations (7; 298). For early-type stars, the radiation force on a I
(15-128) realistic line spectrum exceeded gravity at the surface of the model; but this
was regarded as "nonphysical" for the purposes of model construction and
where se = (neue/P) is the electron-scattering coefficient per gram. Recalling was suppressed in the calculation!] In summary, one finds that, for O-stars,
the results of Exercise 7-1, re:::::: 2.5 x 10- 5 (L/L 0 )(.,H 0 /.,/!); for an O-star the forces obtained when the atmosphere is assumed to be static are incom-
(L/L 0 ) :::::: 106 , (.,H/.,/! 0 ) :::::: 60, and re :::::: 0.4. It is thus clear that continuum patible with that assumption; hence hydrostatic equilibrium in the outermost
absorption alone cannot produce a force that exceeds gravity. (We shall see layers is not possible, and an outflow of material must occur.
below that it is essential for a transsonic wind that it does not!). We must Once the uppermost layer begins to move, the lines will be Doppler-shifted
therefore look to the spectral lines to produce the required force. away from their rest positions and will begin to intercept the intense flux in
At great depth in the atmosphere the diffusion approximation is valid, and the adjacent continuum; this enhances the momentum input to the material
Hv ex: Xv - i [cf. equation (2-91)]; in this limit the product XvHv appearing in and hence increases the acceleration. The underlying layers must expand to
equation (I 5-126) is independent of the value of Xv; i.e., in the diffusion limit fill the rarefaction left by acceleration of the upper layers. Furthermore, the
the lines are no more effective than the continuum in delivering momentum to lines in these lower layers become unsaturated (because the absorption lines
the gas. Thus at depth, r remains essentially equal tor., as given by equation in the upper layer have been Doppler-shifted); hence these underlying layers
(15-128). On the other hand, at the surface of the atmosphere Hv may rise also begin to experience a radiative force that exceeds gravity. In this manner,
far above its diffusion-approximation value, because intense radiation a flow can be initiated; it remains to be shown that (a) the amount of mass
emerges from the material below, and none is incident from above. To estimate loss produced is significant, and (b) the· variation of the radiation force
556 Stellar Winds 15-4 Radiatively Driven Winds 557

with depth is consistent with transsonic flow. Let us consider the latter reaches very large values when the lines shift into the continuum. We shall
point first. find below that, to a good approximation, the force depends upon a power
To obtain a transsonic flow, certain conditions at the sonic point must be of the velocity gradient; this dependence allows the force, and the flow it
met (408; 132), as is the case for coronal winds. For steady flow, mass con- produces, to accomodate to one another, so that a steady transsonic wind can
servation is expressed by equation (15-16), while the momentum equation be obtained.
( 15-103) can be written, using equation (15-127), as Before leaving the question of how the flow is achieved, it is worthwhile
to comment on some other points that have been made in the literature.
v(dv/dr) + p- 1(dp/dr) = -G.,1((1 - r)/r1 (15-131) Cassinelli and Castor (132) studied the problem of energy input into an
Here r is assumed to be a given function of the radius r. The pressure may optically thin flow with continuum opacity only. They reached several
be expressed in terms of the density and isothermal sound speed a [cf. important conclusions. (a) A flow can be driven by thermal energy-input
equation (15-37)] asp = a1 p; a is assumed to be a function of the radius r. from radiation into the gas via true absorption processes near the critical
Then using the equation of continuity (15-14), we find point. Such a mechanism deposits energy in an extended region, analogous
to the solar corona, with radiation now playing the role conduction has at
p- 1(dp/dr) = (da 1/dr) - (2a 1/r) - (a 1/v)(dv/dr) (15-132) coronal temperatures. It was noted that the term E0 in equation (15-124)
is negative near the star, but must become positive at large distances for
Substituting equation (15-132) into equation (15-131), we obtain finite v00 • It was argued that this change must be brought about by the inte-
½[1 - (a 2/v 2 )](dv2 /dr) = (2a 2 /r) - (da 2/dr) - G.,{((1 - r)/r 2 (15-133) grated absorptive energy input (qR), and was concluded that true absorption
is essential to effect the transition from subsonic to supersonic flow. This
For simplicity, assume the envelope is isothermal (a good approximation), conclusion, however, is too restrictive (307), for it is clear from equation
so that the term in (da 2/dr) may be omitted. Then it is clear that, ifwe are to (15-124) that, although appropriate thermal energy input can drive the flow,
obtain a smooth transition from subsonic flow at small r to_ supersonic flow the work done by radiative.forces can serve equally well. In fact, it is precisely
at large r, the righthand side of equation (15-133) must (1) vanish at some this work that drives the winds calculated by Lucy and Solomon (who
critical radius r = re, where v = a; (2) be negative for r < r,; and (3) be assumed that the lines scatter the radiation conservatively, so that qR = 0), I
positive for r > re. The condition for r < r, can be met only if r < 1 in th3.t as well as those described later in this section. (b) In the event that the wind I\)
region; i.e., in the subsonic-flow region the radiation force must be less tha!1 is driven by true absorption, then virtually every early.type star has a wind, (X)
r
that of gravity. If is greater than unity everywhere (which implies that the but in most cases the sonic point is so far from the stellar surface that the '"'-J
whole star is unstable), transsonic flow becomes impossible, and one has either mass-flux is miniscule (the density continues to fall with an essentially hydro- I
an initially subsonic flow that decelerates, or a supersonic flow that accelerates. static scale-height inside the sonic point). Only if unrealistically large values
For r > r, in transsonic flows, r may become arbitrarily large; indeed the for the absorption coefficient and the parameter r (assumed constant) are
larger it is, the greater is the momentum input into the gas, and the larger adopted do significant winds result. These difficulties are overcome entirely
(dv/dr) will be. This is, of course, of great interest for early-type stellar winds, when a realistic representation of the radiative force on lines is used (see
for we have seen that r can become very large in the supersonic-flow region, below). (c) The time required by the material to gain and lose energy
where the lines are sufficiently displaced from their rest frequencies that they radiatively is short compared to the time a fluid element requires to move
absorb continuum radiation; indeed it is just these large values ofr that lead through a density scale-height. Therefore, to a high degree of approximation,
to the high values of v00 that are observed. Because the flow is already super- the energy balance is given by radiative equilibrium; this result is exploited
sonic where r exceeds unity, information about the momentum input cannot in the models described later.
propagate back upstream, and the flow at the sonic point itself is essentially The first successful radiatively-driven wind models for 0-stars were con-
unaffected. structed by Lucy and Solomon (404). They assumed (1) planar geometry
It is important to recognize that the radiation force on the continuum and (adequate for the flow inside the sonic point); (2) constant temperature;
on spectrum lines has precisely the right properties, as delineated above, to (3) and ionization equilibrium fixed by equation (5-46) with W = ½, and
produce the desired transsonic wind. That is, r is less than unity inside the T, and T, equal to 0.7T.rr (the effective temperature of the model photo-
star where the diffusion approximation is valid, approaches unity as the lines sphere); and (4) that the radiation force is dominated by absorption in
desaturate, becomes greater than one when the lines are optically thin, and resonance lines of a few ions (C III, C IV, N III, NV, Si IV, S III, S IV, and
558 Stellar Winds 15-4 Radiatively Driven Winds 559

S VI). The momentum equation was written as radiative momentum absorbed in the lines (assuming perfect efficiency); i.e.,
(15-134) (15-137)
where g.rr = g• - gR, 1; g• =g - (nFs,/c), which allows for the radiation
where the sum extends over all lines. Lucy and Solomon argued that the sum
force from electron scattering; and llR, 1 is the radiation force on the lines. To
calculate gR," the lines were treated as pure scatterers, illuminated from would be dominated by a single line located near the maximum of the flux
below by a photospheric intensity distribution, and wrote .,Hv<X) = (4n 2r 2/c) · Fm ..(vm ..vcx,/c), or

(15-135) .,I(= (4n 2 r 2/c 2 )Fmu"mu::::: (4n 2 r 2/c 2 )F = (L/c 2),


where u = n,u,, the electron scattering coefficient per unit volume. Equation where F denotes the integrated flux, and L is the stellar luminosity. This
(15-135) gives a rough representation of the emergent intensity in the profile yields A= 7 x 10- 14(L/L 0 ).J( 0 /year, or about 7 x 10- 8 .,({ 0 /year for
6
of a pure scattering line formed in the photosphere. In calculating the intensity L = l0 L 0 . Lucy and Solomon considered this result to be an upper limit
higher in the envelope, re-emissions were ignored because, on the average, to the mass loss; actually it is more nearly a lower limit (627); for in essence
they contribute nothing to the force exerted by radiation on the material. it is obtained by accelerating a single line across the entire spectrum, to drive
In this case lhv) decays exponentially, and we can write J.{t.) = I.{O) the material (formally) to the speed oflight. A better estimate (132) is obtained
exp( -t./µ), where 'l'v is the optical depth, at frequency v, from the base of the by replacing (4n 2 r 2 ) IF(v,) .1v1 in equation (15-137) with L; then
envelope to the test point, allowing for Doppler shifting of the line profile
.,I( ~ (L/vcx,c) =7 x 10- 12 (L/L 0 )(3000/vcx,).,({ 0 /year (15-138)
along the path. Then
where v<X) is expressed in km s- 1• For L = 106 L 0 , vex, = 3000 km s- 1, .,H =
gR, 1 = (2n/cp) L Jo' dµ Jo<X) dv x,(v)I .(0)µe-"' 1" (15-136) 6
I 7 x 10- .,({ 0 /year, which is the observed value for CPup. This result pre-
sumes that all the momentum carried by stellar photons is converted into I
where the sum extends over all lines. mass-loss in a single scattering; in actuality only some fraction e will be so N
The density p 0 at the base of the envelope is taken from a model atmo- converted, but observations indicate that e may be substantial, perhaps as en
sphere. Then a trial value is chosen for the velocity v0 ; this fixes the mass-fhix
J = p0 v0 • The mass-flux is an eigenvalue of the problem: if too large a value
large as 0.5. Thus it appears that adequate mass-loss rates can be obtained, en
if (and only if) enough lines are included in the calculation of gR (see also I
is chosen, then at the sonic point, where v = a, the radiation force will be the discussion in (145)]. A still larger bound on .,H may be obtained when
too small [essentially because t. in equation (15-136) will be too large], and account is taken of the possibility that some photons may scatter several
g.rr will be >0, making a continuous transition to supersonic flow impossible. times in the envelope, pas~ing back and forth to regions on opposite sides
Likewise, if J is chosen too small, llerr at the sonic point will be <0. For of the central star between successive scatterings. Quantitative estimates of
precisely the right value of J, the condition Oerr = 0 will be met, and transsonic the importance of this effect have yet to. be made.
flow is possible. A large number of solutions, for a range .of stellar parameters,
were obtained in this way, and it was found that they gave mass-loss rates of
LINE-DRIVEN WINDS IN Of STARS
the order of 10-s .,({ 0 /year (or Jess), which is about a factor of 100 smaller
than the observed values, despite the fact that reasonable terminal velocities, The most complete and internally consistent theory for radiatively driven
vex, ~ 3300 km s- 1, were obtained. stellar winds at present is that of Castor, Abott, and Klein (138; 145). The
Let us now inquire: "How large a mass-flux can be driven by radiation physical parameters employed make this theory applicable to Of stars. The
from a star?" Suppose that the spectrum contains a large number of lines, flow is assumed to be time-independent and spherically symmetric; the gas
each of which totally removes the momentum from the radiation field at is taken to be a single fluid; and conduction and viscosity are neglected. The
each frequency where it absorbs. Assume that the material is accelerated gas is assumed to absorb momentum from the radiation field in spectral
from v = 0 to v = vex, in the process, so that a line at frequency v, is spread lines according to a particular force law discussed below.
over a frequency range L1v1 = (v,v/c). Then the maximum mass-loss that can The single-fluid approximation can be justified (145) by a comparison of
be driven is obtained by putting the material momentum-flux equal to the the fluid-flow velocity to the drift velocity of the ions absorbing the radiative
560 Stellar Winds 15-4 Radiatively Driven Winds 561

momentum (relative to the remainder of the material with which they suffer where R is the photospheric radius, or (b) by the velocity gradient (which
Coulomb collisions). For a representative electron density of ne ~ 10 11 and Doppler-shifts the line from its rest frequency) in a moving medium, in which
temperature T ~ 40,000°K, the drift velocity of c+ 3 ions is found to be case
0.7 km s- 1, which is clearly negligible in a medium with a flow velocity of (15-141b)
1000 km s- 1 • Also the lines of all chemical species are observed to span the
where v,h is the thermal velocity of the absorbing atoms. Equation ( 15-141 b)
same velocity range, indicating that there is no systematic separation of the
is derived from considerations similar to those employed in development of
material, and supporting the single-fluid picture. To justify the neglect of
the Sobolev approximation, and is the planar equivalent of equations (14-61)
viscosity, one computes the Reynolds number !Jt = (vl/v), where l is a char- and (14-62). .
acteristic length-scale in the flow at density p and velocity v, for a fluid with
It is more convenient to have a depth-scale that is independent of line-
kinematic viscosity v. The Reynolds number essentially gives the ratio of
strength, so we introduce /31 = (x,/u), and calculate an equivalent electron
inertial to viscous forces (490, 19; 385, 62); at very large values of !It, the
optical depth scale t = t 1/{3,. which for an expanding atmosphere is defined
fluid may be regarded as inviscid. In an Of wind the minimum calculated to be
value of !It is found (145) to be of the order of 10 10, which assures that viscosity
(15-142)
can be neglected. Finally, because the temperature of the material is not
extremely high, the thermal conductivity is low; and further, because the We shall use equation (15-142) throughout the wind, even though it becomes
mass flux is large (10 8 times the solar wind), the conductive flux is found (145) invalid in the stellar photosphere, for the radiation force on the lines becomes
to be about eight orders of magnitude smaller than heat transport by bulk negligible there anyway. The total line force is obtained by summing equation
motions, and hence it can be ignored. (15-140) over all lines, and can be written
Let us now consider how to calculate the force exerted by the radiation
on the material; the essential point is to account for saturation of the lines,
gR, 1 = (1tFu/cp)M(t) = (seL/4ncr2 )M(t) (15-143)
so that the transition between the optically thick and thin limits-is handled where M(t) = F- 1 IF,(v1)Avn,,min(p,,r- 1 ) (15-144)
correctly. This problem has been treated in detail by Castor (137); his analysis I I
leads to a simple result, for which a heuristic argument that contains the I\)
is the radiation-force multiplier. The calculation of the radiation force is thus
essence of the physics will be presented here. Suppose that the absorbing reduced to the evaluation of M(t), which is a function of only one param- (X)
lines are confined to a discrete layer overlying the photosphere. Let the eter (t). co
continuum flux incident from below be 1tF0 and approximate the momentum Castor, Abott, and Klein evaluated M(t) on the assumption that the line I
absorbed (per unit mass) from the unattenuated continuum, by a line of spectrum is the same as that ofC III (for which an extensive set of /-values
opacity x, and width Av0 , as g11 , 1(0) = (nF,x, Av0 /cp). To account for was available), and by assigning a total abundance ore++ relative to hydro-
attenuation we note that (as only the net momentum input is needed) re- gen of 10- 3 (which is the total abundance of C, N, and O taken together).
emissions, which are presumed to be isotropic, can be ignored. In this case The occupation numbers were computed using LTE. Although these as-
the incident flux decays as e-•• where t 1 is the line optical depth computed sumptions are somewhat rough, the results are certainly qualitatively correct.
allowing for Doppler shifts. Thus the average rate of momentum input into The numerical values of M(t) are found to be well fitted by the formula
the layer is · M(t) = kt-",withk ~ -fo-andcx = 0.7.Inmorerecentwork(l45),allelements
J;•
t 1(g11 , 1) = g11 , 1(0) e-•· dt' (15-139) from H through Ni are included, and the coefficients k and ex are allowed to
depend on the relevant physical variables (but these refinements will not be
or (15-140) considered further here). Using equations (15-142) and (15-143), and the
formula for M(t), we find finally
In their work, Castor, Abott, and Klein replace the term t 1- 1(1 - e-'•) with
min( I, t 1 - 1), which provides an adequate approximation.
At a given observer's-frame frequency, the effective optical thickness t 1
will be limited by (a) the amount of material in the line-layer if the medium
YR,, = (:;~:2) C~,h ~~)" = ~ (r v~~)"
2
(15-145)

is static, in which case where the equation of continuity has been invoked, and the constant C is
(15-141a) (15-146)
562 Stellar Winds 15-4 Radiatively Driven Winds 563

Using equation (15-145) for the line contribution to the radiation force, where r, is the sonic radius (essentially equal to R, the photospheric radius),
the equation of motion [cf. equation (15-133)] becomes and r, is the critical radius. Equation (15-154) is based on the assumption
that T oc r-"; likely values for .n lie between O (isothermal) and ½(radiative
2 2 2 2
! (l _ a ) dv = 2a _ da _ G..i(l - r.) + C (r 2 v dv)" (15 _147) equilibrium), which implies that 1.5 ~ (r,/r,) :S 1.74.
2 v2 dr r dr r2 r2 dr The stellar model is parameterized by a choice of L, ..i, and R, and by an
assumed relation for T(r) [ or a 2 (r)]. The mass-loss rate is fixed almost entirely
Or, defining the new variables w = ½v 2 and u = -r- 1, equation (15-147) is by .,If and L (via r,); the characteristic temperature T.rr also enters into v,h•
equivalent to In practice, the model is determined by guessing a value for r, from equation
(15-154) with r, = R; equation (15-147) is then solved numerically, and the
F(u, w, w') = (1 - ½a 2 w- 1)w' - h(u) - C(w')' = 0 ( 15-148) exact relation between radius and optical depth is computed. The value of
r, is then adjusted until a reasonable photospheric optical depth (::::: f) is
where w' = (dw/d11), and attained at r = R. Having constructed a dynamical model, one may use the
h(u) = -G..i(l - r,) - 2a 2 u- 1 - (da 2 /du) (15-149) resulting density structure in a spherical model-atmosphere code to adjust
the temperature structure in such a way as to satisfy the requirement of
Equation (15-147) [or equation (15-148)] has a singular point at which radiative equilibrium. The new temperature structure is then employed to
solutions terminate, have cusps, or show other discontinuities; this point is reconstruct a new dynamical model, and the process is iterated. In practice
=
not the sonic point. At the sonic point, where r a, it is easily seen that the the dynamics-and hence the velocity, density structure, and mass-loss rate-
lefthand side of equation (15-147) vanishes, and that the righthand side can is insensitive to the temperature structure, and the iteration process con-
be made to vanish as well with a suitable choice of(dv/dr), which need not be verges rapidly.
infinite or discontinuous. This difference from standard corenal-wind theory Castor, Abbott, and Klein have published (138) a solution for parameters
results from the force depending on (dv/dr) instead of just r itself. The critical- appropriate to an 05 star: .,If = 60..i 0 , L = 9.7 x 10 5L 0 , R = 9.6 x
point analysis is relatively complicated, compared to the theory described in 10 11 cm = 13.8R0 , T.rr = 49,300°K, log g = 3.94, re = 0.4. The re- I
§15-2, because the equation is nonlinear in (dv/dr). Detailed study (138) of sulting mass-loss rate is .,I/ = 6.6 x 10- 6 .,If 0 /year, a value appropriate to I\)
a star like CPup. The terminal velocity is v00 = 1500 km s- 1, so .,{j ::::: ½(L/v 00 c),
the situation shows that the locus of singular points is specified by
which shows that about one half of the momentum originally carried by
co
(15-150) radiatio!] has been transferred to the matter. Furthermore, about one-half
0
of the continuum radiation field is blocked by the lines; this value appears
I
But not every point on this locus yields an acceptable solution; if w' is to be to be in good agreement with observation (304). Standard stellar evolution
continuous the additional condition theory (606) gives main-sequence lifetimes at this mass of about 3 x 106
(oF/ou) + w'(oF/ow) = 0 ( 15-151) t
years, which implies a total mass-loss of about the original mass; it thus
appears that the stellar winds of these stars will have very significant effects
must also be satisfied. Equations (15-148), (15-150), and (15~151) determine on their evolution.
(11, w, and w') if C i's given, or (w, w', and C) if u is given. Analytical expressions The density and velocity structures of the model described above are
can then be obtained for the rate of mass-loss, the velocity law, and the shown in Figures 15-5 and 15-6; the letters P, S, and C designate the photo-
acceleration (dv/dr). In the limit v » a these are quite simple: sphere, sonic point, and critical point respectively. The striking characteristic

.,I/ = (41tG..i)
SeVih
(X (~)(1-o)/•
1 - re
(kre)' If•) (15-152)
of the solution is the "core-halo" nature of the density structure: inside the
sonic point the model has a nearly hydrostatic density gradient, while outside
the critical point, p oc ,-- 2 • The model insider, is essentially planar. The
velocity rise outside r, is abrupt and, in fact, the predicted gradient is much
v2 = [2G..i(l - r,)a/(1 - a)][(l/r,) - (1/r)] (15-153) steeper than than those inferred observationally for some Of stars (326; 327);
however, these empirical analyses are rather rough, and a much more accurate
and (r,/r,) = 1 + { -½n + [¼n 2 + 4 - 2n(n + l)Jt}- 1 (15-154) study is urgently needed. The velocity variation is, of course, fundamental to
564 565

104
6
10>
C
C
101

s 4
JO'
V M
10°
p
10- 1 2
s
10- 2
p
10->
0
10 11 10 12 10 13
10-• 10- 1 101
r
FIGURE 15-S FIGURE 15-7
Variation of velocity (in km s- 1) with radius (in cm) in an Variation of radiation force multiplier M with
Of stellar-wind model. From (138), by permission. continuum optical depth in an Of stellar-wind model.
From (138), by permission.

the dynamics, and its empirical determination strongly merits any efforts
I
necessary to yield accurate results. I\)
The variation of the radiation-force multiplier is shown in Figure 15-7. (0
There one sees that in the outer envelope M :::::: 5, which implies that the ......
radiation force on the lines is about twice the force of gravity; adding the
10- 7
acceleration from electron scattering (and subtracting the force of gravity),
we see that the material experiences a net outward acceleration of about 1.4
10- 9 p times gravity.
The emergent relative flux distribution from the wind model is almost
s identical to that from a planar static model with T.rr :::::: 50,000°K. However,
10-11
p the outer envelope has an electron-scattering optical depth of about 0.16,
and this layer (which scatters conservatively and hence does not affect the
10- 13 relative energy distribution) increases somewhat the observed radius of the
star (measured interferometrically), and thus reduces the absolute flux. A
critical discussion (304) of both the visible and ultraviolet data for C Pup
10- 15
(allowing for interstellar reddening effects and line-blanketing) shows that,
if both the observational data and the theoretical models are pushed to the
10-11
10 11 10 12 10 13 10 14
allowable limits, agreement is obtained with the Castor-Abbott-Klein model.
Earlier work had suggested the need to consider sphericity effects in an ex-
tended atmosphere to fit the flux distribution.
FIGURE 15-6
Variation of density (in gm cm -J) with radius (in cm) in an Finally, a calculation of the Ha. and He II ,1,4686 line profiles from the wind
Of stellar-wind model. From (138), by permission. model yields profiles of shape similar to those observed, and with emission
566 Stellar Winds 15-4 Radiatively Driven Winds 567

strengths and redshifts in qualitative agreement with typical values measured (304) of the energy distribution of HD 50896 (WN5) seems absolutely to
for Of stars; these need to be refined with a more accurate calculation. On require an extended subsonic-flow region; some models of this type have
the whole, the Castor-Abbott-Klein model appears to give a fairly satis- been constructed (133), but only with ad hoc force laws, and much further
factory picture of the basic dynamics of Of atmospheres; but many questions work remains to be done before they can be regarded as satisfactory.
remain open. Although all theoretical models of radiatively driven winds assume that
the flow is steady, there is ample evidence that the spectrum (and hence the
FRONTIERS wind) of Of stars is time-variable. A wide range of time-scales (176; 536) has
been noted. It appears, in fact, that essentially all early-type supergiants that
The theory of radiatively driven winds is at an early stage of its develop- show emission lines are intrinsic spectrum variables (537). In a few cases,
ment, and many interesting (and challenging) problems remain to be attacked. pathological spectra with transient inverse P-Cygni profiles (presumably
Within the framework of the Castor-Abbott-Klein theory, refinements of indicating temporary inflow of the material) have been observed (173). If
the force law to include the complete spectrum, a more realistic ionization:- the time-scale of the variations is long compared to the time a fluid parcel
excitation equilibrium, and more accurate treatment of the transfer (account- requires to move from the photosphere to the critical point, then one can
ing for the fact that a photon scattered at one point in the envelope can argue that the flow can be considered as a sequence of quasi-stationary
subsequently interact with the material at some ~ther point), should yield states, each of which is well approximated by steady flow. The only problem
more precise and reliable results. then remaining would be to understand the mechanism leading to the
A more difficult problem is posed by the question of how to treat properly variations. On the other hand, if very short time-scales are ever observed,
the energy equation and thereby determine the temperature structure of the then a fully time-dependent treatment might be required, and this would
flow. The observations (533) of O VI lines, described earlier in this section, introduce staggering difficulties into the problem.
have been interpreted (assuming collisional ionization onl.v-::-an assumption Although the assumption of spherical symmetry of the wind is a reasonable
that requires further consideration) as requiring temperatures of the order starting point, it may not adequately describe the flow for some stars. In
of 2 x 10 5 °K, which is much higher than can be produced radiatively. This particular, if the star is rapidly rotating, then centrifugal forces can appre- I
suggests that there may be mechanical energy deposition that produces ~ ciably lower the effective surface gravity, to the point where it barely exceeds I\)
(relatively cool) corona. Mechanisms for producing a mechanical flux (which
presumably dissipates and heats the outer layers) have been proposed in
even continuum radiation-forces. This may lead to enhanced mass loss from co
the equatorial regions of the star (409), in which case the flow becomes I\)
(289; 290; 404), but these are not entirely satisfactory in their present form. axisymmetric instead of spherically symmetric. The divergence of stream- I
If the flow were to become turbulent (see below), energy dissipation and lines away from the equatorial plane again introduces the possibility of a
heating could occur and, because pv 2 » kT, even a rather low efficiency in radical alteration of the topology of the solution (307). Furthermore, rotation
the conversion of flow energy to heat could have a large effect on the tem- implies that the flow has nonzero vorticity and, in the presence of rotational
perature. Nevertheless, it must be stressed that, although current models do shear, the flow may disintegrate and become turbulent, as suggested by the
not convincingly determine the temperature structure, the dynamics of the very large Reynolds numbers mentioned earlier in this section. In this event,
flow will remain essentially unaltered for T ~ 3 x 10 7 °K (a value that gross inhomogeneities may develop in the wind, and again the theoretical
seems completely excluded observationally), unless the energy deposition complexity becomes overwhelming.
changes the topology of the solution [ e.g., by the introduction of additional Finally, there is the question of whether magnetic fields play a significant
critical points (307)] and, thereby, even the qualitative nature of the solution. role in winds from early-type stars. The O-stars are very young, and have
These possibilities all await further exploration. only recently formed from the interstellar medium. Presumably any fields
One of the problems that will ultimately have to be faced by the "core-halo" present in the medium could persist as weak fields in the atmospheres of
models is that evidence has been presented for atmospheric extension effects these stars. There would then be the possibility of a region of forced corotation
in the continuous energy distribution of the most extreme Of stars (367; 466). out to an Alfvenic point, with subsequent radial expansion. Could this give
Unless it can be shown that there are errors in the observations or in their rise to the profiles with broad emission wings extending beyond the short-
reduction (e.g., in the allowance for interstellar reddening), it will be necessary wavelength edge of the P-Cygni absorption feature, as observed in some
to find ways to construct models showing a slower outward rise of the stars? Could such fields produce structural inhomogeneities with flow-tube
velocity. Such models seem essential for WR stars, where a critical analysis divergences such as are seen in the solar corona and wind (307)? Could they
568 Stellar Winds

provide structures against which shear (as a result of rotation) with con-
sequent turbulence can develop? The present observational detection
threshold for stellar magnetic fields is several hundred gauss; much weaker
fields (a few tens of gauss) in the atmosphere could have major effects on
the flow.
It should be clear from the points raised above that there remains much to
be learned about the physics of stellar winds for early-type stars. Without
doubt it is unrealistic to suppose that these issues can be decided on the basis References
of theoretical considerations alone. It is obvious that a thorough analysis of
the spectroscopic data, at a high level of internal consistency, with the goal
of diagnosing the physical conditions in the flow semiempirica/ly, is required,
and that such efforts will be immensely rewarding.

I. Abhyankar, K. 1964. Astrophys. J. 140: 1353. I


2. __ , 1964. Astrophys. J. 140: 1368. I\)
3.
4.
- - · 1965. Astrophys. J. 141: 1056.
Abramowitz, M., and I. Stegun. 1964. Handbook of Mathematical Func-
co
tions. Washington, D.C.: U.S. Dept. of Commerce. w
5. Abt, H., A. Meinel, W. Morgan, and J. Tapscott. 1969. An Atlas of Low- I
Dispersion Grating Stellar Spectra. Kitt Peak National Observatory, Steward
Observatory, and Yerkes Observatory.
6. Adams, T., D. Hummer, and G. Rybicki. 1971. J.Q.S.R.T. 11: 1365.
7. Adams, T., and D. Morton. 1968. Astrophys. J. 152: 195.
8. Alder, B., S. Fernbach, and M. Rotenberg (eds.). 1967. Methods in Computa-
tional Physics, Vol. 7. New York: Academic Press.
9. Allen, C. 1973. Astrophysical Quantities, 3rd ed. London: Athlone Press.
10. Aller, L. 1956. Gaseous Nebulae. New York: Wiley.
11. - - · 1963. The Atmospheres of the Sun and Stars, 2nd ed. New York:
Ronald Press.
12. - - · 1965. Advances in Astron. and Astrophys. 3: I.
13. Aller, L., and J. Greenstein. 1960. Astrophys. J. Supp. No. 46 5: 139.
14. Aller, L., and D. McLaughlin (eds.). 1965. Stellar Structure. Chicago: Univ.
of Chicago Press.
15. Ambartsumyan, V., (ed.). 1958. Theoretical Astrophysics. London: Pergamon
Press.
16. Athay, R. 1964. Astrophys. J. 140: 1579.
570 References
References 571
17. Athay, R. 1970. Astrophys. J. 161 :713.
18. - - · 1972. Radiation Transport in Spectral Lines. Dordrecht: Reidel. 60. Baker, J., and D. Menzel. 1938. Astrophys. J. 88: 52.
19. _ _ (ed.). 1974. Chromospheric Fine Structure. Dordrecht: Reidel. 61. Bappu, M., and J. Sahade (eds.). 1973. Wolf-Rayet and High-Temperature
20. - - · 1976. The Solar Chromosphere and Corona: Quiet Sun. Dordrecht: Stars. Dordrecht: Reidel.
Reidel. 62. Baranger, M. 1958. Phys. Rev. 111 :481.
21. Athay, R., and R. Canfield. 1969. Astrophys. J. 156: 695. 63. - - · 1958. Phys. Rev. 111 :494.
22. Athay, R., and B. Lites. 1972. Astrophys. J. 176:809. 64. - - · 1958. Phys. Rev. 112:855.
23. Athay, R., J. Mathis, and A. Skumanich (eds.). 1968. Resonance Lines in 65. Baranger, M., and B. Mozer. 1959. Phys. Rev. 115:521.
Astrophysics. Boulder: National Center for Atmospheric Research. 66. Barnard, A., J. Cooper, and L. Shamey. 1969. Astron. and Astrophys. 1 :28.
24. Athay, R., and A. Skumanich. 1968. Astrophys. J. 152: 141. 67. Barnard, A., and J. Cooper. 1970. J.Q.S.R.T. 10:695.
25. 1968. Astrophys. J. 152:211. 68. Barnard, A., J. Cooper, and E. Smith. 1974. J.Q.S.R.T. 14: 1025.
26. - - · 1968. Solar Phys. 3: 181. 69. - - · 1975. J.Q.S.R.T. 15:429.
27. - - · 1968. Solar Phys. 4: 176. 70. Baschek, B., W. Kegel, and G. Traving (eds.). 1975. Problems in Stellar
28. - - · 1971. Astrophys. J. 170:605. Atmospheres and Envelopes. Berlin: Springer-Verlag.
29. Athay, R., and R. Thomas. 1958. Astrophys. J. 127:96. 71. Baschek, B., and J. Oke. 1965. Astrophys. J. 141: 1404.
30. Auer, L. 1967. Astrophys. J. Letters 150:L53. 72. Bates, D. 1952. M.N.R.A.S. 112:40.
3 I. 1968. Astrophys. J. 153: 783. 73. - - (ed.). 1962. Atomic and Molecular Processes. New York: Academic
32. 1971. J.Q.S.R.T. 11 :573. Press.
33. 1973. Astrophys. J. 180:469. 74. Bates, D., and A. Damgaard. 1949. Phil. Trans. Roy. Soc. (London) 242A:
34. 1976. J.Q.S.R.T. 16:931. 101.
35. Auer, L., and J. Heasley. 1971. Unpublished Yale Univ. Observatory Report. 75. Beals, C. 1929. M.N.R.A.S. 90:202.
36. Auer, L., J. Heasley, and R. Milkey. 1972. Kitt Peak Natl. Obs. Contr. 76. 1930. P. Dominion Astrophys. Obs. Victoria 4:271.
No. 555. Tucson: Kitt Peak National Observatory. 77. 1931. M.N.R.A.S. 91:966.
37. Auer, L., and J. Heasley. 1976. Astrophys. J. 205: 165. 78. 1934. P. Dominion Astrophys. Obs. Victoria 6:95.
38. Auer, L., and D. Mihalas. 1968. Astrophys. J. 151 :311. 79. 1950. P. Dominion Astrophys. Obs. Victoria 9: I.
39. 1968. Astrophys. J.153:245. 80. Bell, R. 1970. M.N.R.A.S. 148:25.
40. 1969. Astrophys. J. 156: 157. 81. 1971. M.N.R.A.S. 154:343. I
82. Bell, R., and D. Gottlieb. 1971. M.N.R.A.S. 151:449. I\)
41. 1969. Astrophys. J. 156:681.
42. -1969, Astrophys. J. 158:641. 83. Bell, R., and S. Parsons. 1974. M.N.R.A.S. 169:71. co
43. 1970. Astrophys. J. 160:233. 84. Benett, S., and H. Griem. 1971. Univ. of Maryland Technical Report No. .t::.
44. 1970. M.N.R.A.S. 149:60. 71-097. College Park: Univ. of Maryland. I
45. 1972. Astrophys. J. Supp. No. 205 24: 193. 85. Berger, J. 1956. Astrophys. J. 124:550.
46. 1973. Astrophys. J. Supp. No. 223 25:433. 86. Berman, P., and W. Lamb. 1969. Phys. Re~. 187:221.
47. 1973. Astrophys. J. 184: 151. 87. Bethe, H., a.nd E. Salpeter. 1957. Quantum Mechanics of One- and Two-
48. Auman, J. 1969. Astrophys. J. 157:799. Electron A toms. Berlin: Springer-Verlag.
49. Auvergne, M., H. Frisch, U. Frisch, C. Froeschle, and A. Pouquet. 1973. 88. Bhatnagar, P., M. Krook, D. Menzel, and R. Thomas. 1955. Vistas in Astron.
Astron. and Astrophys. 29:93. 1:296.
50. Avery, L., and L. House. 1968. Astrophys. J. 152:493. 89. Biermann, L. 1946. Naturwiss. 33: 118.
51. Avrett, E. 1966. Astrophys. J. 144:59. 90. 1948. Z. fiir Astrophys. 25: 135.
52. - - · 1971. J.Q.S.R.T.11:519. 91. - - · 1951. Z.furAstrophys.29:214.
53. Avrett, E., and D. Hummer. 1965. M.N.R.A.S. 130:295. 92. - - · 1953. Mem. Roy. Soc. Sci, Liege 13:291.
54. Avrett, E., and W. Kalkofen. 1968. J.Q.S.R.T. 8:219. 93. - - · 1957. Observatory 107: 109.
55. Avrett, E., and M. Krook. 1963. Astrophys. J. 137: 874. 94. Billings, D. 1966. A Guide to the Solar Corona. New York: Academic Press.
56. Ayres, T., J. Linsky, and R. Shine. 1974. Astrophys. J. 192:93. 95. Blaha, M. 1969. Astrophys. J. 157:473.
57. 1975. Astrophys. J. Letters 195:Ll21. 96. Bless, R., A. Code, and E. Fairchild. 1976. Astrophys. J. 203:410.
58. Ayres, T., and J. Linsky. 1975. Astrophys. J. 200:660. 97. Bode, G. 1965. Die Kontinuierliche Absorption von Sternatmosphliren. Kiel:
59. - - · 1976. Astrophys. J. 205:874. Institut fiir Theoretische Physik.
98. Bohm, K.-H. 1961. Astrophys. J. 134:264.
572 References References 573

99. Bohm, K.-H. 1963. Astrophys. J. 137:881. 141. Cayrel, R. 1961. Ann. d'Astroph,1·s. 23:235.
100. _ _ , 1963. Astrophys. J. 138:297. 142. _ _ , 1963. C.R. Acad. Sci. Paris251:3309.
IOI. _ _ , 1969. Astron. and Astrophys. 1: 180. 143. _ _ . 1966. J.Q.S.R.T. 6:621.
102. Bohm-Vitense, E. 1973. Astrophys. J. 181: 379. 144. Cayrel, R., and G. Cayrel de Strobel. 1966. Ann. Rev. Astron. and Astrophys.
103. Bolton, C. 1970. Astrophys. J. 161: 1187. 4:1.
104. Bond, J., K. Watson, and J. Welch. 1965. Atomic Theory of Gas Dynamics. 145. Cayrel, R., and M. Steinberg (eds.). 1976. Physique des Mm11•c•ments clans /es
Reading, Mass.: Addison-Wesley. A rmospheres Srellaires. Paris: Centre National de la Recherche Scientifique. p. 363.
105. Bradley, P., and D. Morton, 1969. Astrophys. J. 156:687. 146. Cayrel, R., and G. Traving. 1960. Z.fiir Astrophys. 50:239.
106. Brandt, J. 1964. Icarus 3:253. 147. Chamberlain, J. 1961. Astrophys. J. 133:675.
107. _ _ . 1970. Introduction to the Solar Wind. San Francisco: W. H. Freeman 148. Chandrasekhar, S. 1934. M.N.R.A.S. 94:443.
and Company. 149. 1934. M.N.R.A.S. 94:522.
108. Brandt, J., and J. Cassinelli. 1966. Icarus 5:47. 150. 1936. M.N.R.A.S. 96:21.
109. Branscomb, L., and B. Pagel. 1958. M.N.R.A.S. 118:258. 151. 1943. Rev. Mod. Phys. 15: I.
110. Brechot, S., and H. Van Regemorter. 1964. Ann. d'Astrophys. 27:432. 152. 1944. Astrophys. J. 99: 180.
111. _ _ , 1964. Ann. d'Astrophys. 27:739. 153. 1944. Astrophys. J. 100:76.
112. Breene, R. 1961. The Shift and Shape of Spectral Lines. Oxford: Pergamon I 54. 1945. Astrophys. J. 101: 95.
Press. 155. 1945. Astrophys. J. 101: 328.
113. Brown, R., J. Davis, and L. Allen. 1974. M.N.R.A.S. 167: 121. 156. 1945. Astrophys. J. 102:402.
114. Brucato, R. 1971. M.N.R.A.S. 153:435. 157. 1945. Rev. Mod. Phys. 17: 138.
115. Brucato, R., and D. Mihalas. 1971. M.N.R.A.S. 154:491. 158. 1947. Astrophys. J. 106: 145.
116. Brueckner, K. 1971. Astrophys. J. 169:621. 159. 1948. Proc. Roy. Soc. (London) A192:508.
117. Buckingham, R., S. Reid, and R. Spence. 1952.· M.N.R.A.S. 112:382. 160. 1957. An Introduction to the S111dy of Stellar Structure. New York:
118. Burgess, A. I964. Astrophys. J. 139: 776. Dover.
119. _ _ , 1965. Astrophys, J. 141: 1588. 161. 1960. Radiative Transfer. New York: Dover. I
120. Burgess, A., and M. Seaton. 1960. M.N.R.A.S. 120: 121. 162. Chandrasekhar, S., and F. Breene. 1946. Astrophys. J. 104:430. I\)
121. Burgess, A., and H. Summers. 1969. Astrophys. J, 157: 1007. 163. Chapman, R. 1966. Astrophys. J. 143:61. (0
122. Burgess, D. 1970. J. Phys. B. 3:L70. 164. Chapman, S., and V. Ferraro. 1931. Terr. Magn. and Atm. Elec. 36:77. 01
123. Burgess, D., and C. Cairns. 1970. J. Phys. B. 3:L67. 165. _ _ . 1940. Terr. Magn. and Atm. Elec. 45:245. I
124. - - · 1971. J. Phys. B. 4: 1364. 166. Chapman, S. 1959. Proc. Roy. Soc. (London) A253:450.
125. Cameron, R., (ed.). 1967. The Magnetic and Related Stars. Baltimore: Mono 167. Chapelle, J., and S. Sahal-Brechot. 1970. Astron. and Astrophys. 6:415.
Book Corp. 168. Chipman, E. 1971. S.A.O. Special Report No. 338. Cambridge, Mass.:
126. Carbon, D. 1974. Astrophys. J. 187: 135. Smithsonian Astrophysical Observatory.
127. carlson, B., and K. Lathrop. 1968. In Computing Methods in Reactor Physics, 169. Code, A., J. Davis, R. Bless, and R. Brown. 1976. Astrophys. J. 203:417.
ed. H. Greenspan, C. Kelber, and D. Okrent. New York: Gordon and Breach. 170. Cody, W., K. Paciorek, and H. Thacher. 1970. Math. Comp. 24: 171.
128. Carrier, G., and E. Avrett. 1961. Astrophys. J. 134:469. 171. Cohen, M., M. Barlow, and L. Kuhi. 1975. Astron. and Astrophys. 40:291.
129. Carruthers, G. 1968. Astrophys. J. 151:269. 172. Condon, E., and G. Shortley. 1963. Theory of Atomic Spectra. Cambridge:
130. Cassinelli, J. 1971. Astrophys. J. 165:265. Cambridge Univ. Press.
131. _ _ , 1971. Astrophys. Letters 8: 105. 173. Conti, P. 1972. Astrophys. J. Letters 174:L79.
132. Cassinelli, J., and J. castor. 1973. Astrophys. J. 179: 189. 174. ___ , 1973. Astrophys. J. 179: 161.
133. Cassinelli, J., and L. Hartmann. 1975. Astrophys. J. 202: 718. 175. _ _ , 1974. Astrophys. J. 187:539.
134. Castor,J. 1970. M.N.R.A.S.149:111. 176. Conti, P., and S. Frost. 1974. Astrophys. J. Le11ers 190: LI 37.
135. 1972. Astrophys. J. 178:779. 177. Conti, P., and E. Leep. 1974. Astrophys. J. 193: 113.
136. - - · 1974. Astrophys. J. 189:273. 178. Cooper, J. 1966. Proceedings of Workshop Conference on the Lowering of
137. - - · 1974. M.N.R.A.S. 169:279. the Ionization Potential [J.I.L.A. Report No. 79]. Boulder: Joint Institute for
138. Castor, J., D. Abbott, and R. Klein. 1975. Astrophys. J. 195: 157. Laboratory Astrophysics.
139. Castor, J., and H. Nussbaumer. 1972. M.N.R.A.S. 155:293. 179. _ _ , 1967. Rev. Mod. Phys. 39: 167.
140. Castor, J., and D. Van Blerkom. 1970. Astrophys. J. 161 :485. 180. Cooper, J., and G. Oertel. 1969. Phys. Rev. 180:286.
574 References References 575

181. 225. Friere, R., and F. Praderie. 1974. Astron. and Astrophys. 37: 117.
Courant, R., and D. Hilbert. 1962. MethodsofMathematical Physics, Volume
226. Frisch, H. 1975. Astron. and Astrophys. 40:267.
II: Partial Differential Equations. New York: Interscience.
182. 227. Frisch, H., and U. Frisch. 1976. M.N.R.A.S. 175: 157.
Cowley, C., and A. Cowley. 1964. Astrophys. J. 140:713.
183. Cowley, C. 1971. Observatory 91: 139. 228. Fuhr, J., G. Martin, and B. Specht. 1975. Bibliography on Atomic Line Shapes
184. and Shifts ( July 1973 through May 1975) [N.B.S. Special Pub. No. 366, Supp. 2].
Cox, A., J. Stewart, and D. Eilers. 1965. Astrophys. J. Supp. No. 94 ll: I.
185. Washington, D.C.: U.S. Dept. of Commerce.
Cram, L. 1972. Solar Phys. 22:375.
229. Fuhr, J., L. Roszman, and W. Wiese. 1974. Bibliography on Atomic Line
186. Cram, L., and· I. Vardavas. 1977. In press.
Shapes and Shifts ( April 1972 through June 1973) [N.B.S. Special Pub. No. 366,
187. Cuny, Y. 1967. Ann. d'Astrophys. 30: 143.
Supp. I]. Washington, D.C.: U.S. Dept. of Commerce.
188. Dalgarno, A., and J. Browne. 1967. Astrophys. J. 149:231.
230. Fuhr, J., and W. Wiese. 1971. Bibliography on Atomic Transition Proba-
189. Dalgarno, A., and D. Williams. 1962. Astrophys. J. 136:690.
bilities: July /969 through June 1971 [N.B.S. Special Pub. No. 320, Supp. I].
190. Davis, J., and J. Webb. 1970. Astrophys. J. 159:551.
Washington, D. C. : U.S. Dept. of Commerce.
191. Dekker, E. 1969. Astron. and Astrophys. 1: 72.
231. _ _ , 1973. Bibliography on Atomic Transition Probabilities: July 1971
192. De Loore, C. 1970. Astrophys. and Space Sci. 6:60.
through June 1973 [N.B.S. Special Pub. No. 320, Supp. 2]. Washington, D.C.:
193. Deridder, G., and W. van Rensbergen. 1976. Astron. and Astrophys. Supp.
U.S. Dept. of Commerce.
23: 147.
232. Fuhr, J., W. Wiese, and L. Roszman. 1972. Bibliography on Atomic Line
194. Deutsch, A. 1970. Astrophys. J. 159:985.
Shapes and Shifts ( 1889 through March /972) [N.B.S. Special Pub. No. 366].
195. Dietz, R., and L. House. 1965. Astrophys. J: 141: 1393.
Washington, D.C.: U.S. Dept. of Commerce.
196. Dirac, P. 1925. M.N.R.A.S. 86:825.
233. Fullerton, W., and C. Cowley. 1'970. Astrophys. J. 162:327.
197. _ _ . 1958. The Principles of Quantum Mechanics. Oxford: Clarendon
234. Gail, H., E. Hundt, W. Kegel,J. Schmid-Burgk, and G. Traving. 1974. Astron.
Press.
and Astrophys. 32: 65.
198. Dumont, S., and N. Heidmann. 1973. Astron. and Astrophys. 27:273.
235. Gail, H., and E. Sedlmayr. 1974. Astron. and Astrophys. 36: 17.
199. Dupree, A., and L. Goldberg. 1970. Ann. Rev. Astron. anil Astrophys. 8:231.
236. Gail, H., E. Sedlmayr, and G. Traving. 1975. Astron. and Astrophys. 44:421.
200. Durney, B. 1971. Astrophys. J. 166:669.
237. _ _ , 1976. Astron. and Astrophys. 46:441. I
201. Durney, B., and P. Roberts. 1971. Astrophys. J. 170:319.
238. Gebbie, K., and R. Thomas. 1968. Astrophys. J. 154:285. I\)
202. Durney, B., and N. Werner. 1972. Astrophys. J. 171 :609.
239. _ _ , 1970. Astrophys.J.161:229.
203. Ecker, G. 1955. Z.fur Phys. 140:274.
240. _ _ , 1971. Astrophys. J. 168:461. co
204. 1955. Z.fur Phys. 140:292. CJ)
241. Gehrz, R., and J. Hackwell. 1974. Astrophys. J. 194:619.
205. _ _·. 1957. Z.furPhys.148:593.
242. Gehman, S. 1962. Astrophys. J. 136:935. I
206. _ _ , 1957.. Z.fur Phys. 149:254.
243. Gerasimovit!, B. 1934. Z.fur Astrophys. 1: 335.
207. Einstein, A. 1917. Phys. Z. 18: 121.
244. Gibson, E. 1973. The Quiet Sun. Washington, D.C.: National Aeronautics
208. Epstein, P. 1916. Ann. d. Phys. 50:489.
and Space Administration.
'209. Feautrier, P. 1964. C.R. Acad. Sci. Paris2S8:3189.
245. Gieske, H., and H. Griem. 1969. Astrophys. J. 157:963.
210. 1967. Ann. d'Astrophys. 30: 125.
246. Gingerich, 0. 1963. Astrophys. J. 138:576.
211. 1968. Ann. d'Astrophys. 31 :257.
247. _ _ (ed.). 1969. Theory and Observation of Normal Stellar Atmospheres.
212. Finn, G. 1967. Astrophys. J. 147: 1085.
Cambridge, Mass.: M.I.T. Press.
213. 1971. J.Q.S.R.T. 11 :477.
248. Gingerich, 0., and C. de Jager. 1968. Solar Phys. 3: 5.
214. 1972. J.Q.S.R.T. 12:35.
249. Gingerich, 0., R. Noyes, W. Kalkofen, and Y. Cuny. 1971. Solar Phys.
215. 1972. J.Q.S.R.T. 12: 149.
18: 347.
216. 1972. J.Q.S.R.T. 12: 1217.
250. Goldberg, L. 1935. Astrophys. J. 82: I.
217. Finn, G., and J. Jefferies. 1968. J.Q.S.R.T. 8: 1705.
_ _ , 1969. J.Q.S.R.T. 9:469. 251. - - · 1936. Astrophys. J. 84: 11.
218.
252. Goldberg, L., E. Millier, and L. Aller. 1960. Astrophys. J. Supp. No. 45 5: I.
219. Finn, G., and D. Mugglestone. 1965. M.N.R.A.S. 129:221.
253. Goldstein, H. 1969. Classical Mechanics. Reading, Mass.: Addison-Wesley.
220. Fischel, D., and W. Sparks. 1971. Astrophys. J. 164: 355.
254. Gordon, W. 1929. Ann. Phys. 2: 1031.
221. Foley, H. 1946. Phys. Rev. 69:616.
255. Grant, I., and A. Peraiah. 1972. M.N.R.A.S. 160:239.
222. Fowler, R., and E. Milne. 1923. Phil. Mag. 45: I.
256. Gray, D. 1976. The Observation and Analysis of Stellar Photospheres. New
223. _ _ , 1924. M.N.R.A.S. 84:499.
York: Wiley.
224. Fransden, S. 1974. Astron. and Astrophys. 37: 139.
576 References References 577

257. Green, L., P. Rush, and C. Chandler. 1957. Astrophys. J. Supp. No. 26 3:37. 296. Henyey, L., and W. Grassberger. 1955. Astrophys. J. 122:498.
258. Greene, R., and J. Cooper. 1975. J.Q.S.R.T. IS: 1037. 297. Herzberg, G. 1944. Atomic Spectra and Atomic Structure. New York: Dover.
259. - - · 1975. J.Q.S.R.T. IS: 1045. 298. Hicock, F., and D. Morton. 1968. Astrophys. J. 152:203.
260. Greene, R., J. Cooper, and E. Smith. 1975. J.Q.S.R.T. IS: 1025. 299. Higginbotham, N., and P. Lee. 1974. Astron. and Astrophys. 33:277.
261. Greenstein, J., (ed.). 1960. Stellar Atmospheres. Chicago: Univ. of Chicago 300. Hiltner, W., (ed.). 1962. Astronomical Techniques. Chicago: Univ. of Chicago
Press. Press.
262. Griem, H. 1960. Astrophys. J. 132: 883. 301. Hindmarsh, W. 1959. M.N.R.A.S. 119: 11.
263. 1962. Astrophys. J. 136:422. 302. _ _ , 1960. M.N.R.A.S. 121:48.
264. 1964. Plasma Spectroscopy. New York: McGraw-Hill. 303. Hindmarsh, W., A. Petford, and G. Smith. 1967. Proc. Roy. Soc. (London)
265. 1967. Asirophys. J. 147: 1092. A297:296.
266. 1968. Astrophys. J. 154: 1111. 304. Holm, A., and J. Cassinelli. 1977. Astrophys. J. 211 :432.
267. 1968. Phys. Rev. 165:258. 305. Holtsmark, J. 1919. An11. cl. Phys. 58:577.
268. 1974. Spectral Line Broadening by Plasmas. New York: Academic 306. Holstein, T. 1950. Phys. Rev. 79:744.
Press. 307. Holzer, T. 1977. J. Geophys. Res. 82:23.
269. Griem, H., M. Baranger, A. Kolb, and G. Oertel. 1962. Phys. Rev. 125: 177. 308. Hooper, C. 1966. Phys. Rev. 149:77.
270. Griem, H., A. Kolb, and K. Shen. 1959. Phys. Rev. 116:4. 309. _ _ , 1968. Phys. Rev. 165:215.
271. - - · 1962. Astrophys. J. 135:272. 310. - - · 1968. Phys. Rev. 169: 193.
272. Griem, H., and K. Shen. 1961. Phys. Rev. 122: 1490. 311. Hudson, R., and L. Kieffer. 1971. Atomic Data 2:205.
273. Groth, H., and P. Wellmann (eds.). 1970. Spectrum Formation in Stars with 312. Hulst, H. van de. 1957. Light Scattering by Small Particles. New York: Wiley.
Steady-State Extended Atmospheres. Washington, D.C.: U.S. Dept. of Com- 313. Hummer, D. 1962. M.N.R.A.S. 125:21.
merce. 314. 1965. Mem. R.A.S. 70: I.
274. Gustafsson, B. 1971. Astron. and Astrophys. 10: 187. 315. 1968. M.N.R.A.S. 138:73.
275. Gustafsson, B., and P. Nissen. 1972. Astron. and Astrophys. 19:261. 316. 1969. M.N.R.A.S. 145:95.
276. Hack, M., (ed.). 1967. Modern Astrophysics, A Memorial to Otto Struve. 317. 1976. In Be and Shell Stars [I.A.U. Symposium No. 70], p. 281. I
Paris: Gauthier-Villars. Dordrecht: Reidel. I\)
277.
278.
Hackwell, J., R. Gehrz, and J. Smith. 1975. Astrophys. J. 192:383.
Haisch, B. 1976. Astrophys. J. 205:520.
318.
319.
Hummer, 0., and D. Mihalas. 1967. Astrophys. J. Letters J50:L57.
_ _ , 1970. M.N.R.A.S. 147:339.
co
279. Harman, R., and M. Seaton. 1964. Astrophys. J. 149:824. 320. Hummer, 0., and G. Rybicki. 1966. J.Q.S.R.T. 6:661.
--..J
280. Harrington, J. 1970. Astrophys. J. 162:913. 321. 1968. Astrophys. J. Le11ers 153: LI07. I
281. Harris, D. 1948. Astrophys. J. 108: 112. 322. _ _ , 1970. M.N.R.A.S. 150:419.
282. Hartle, R., and P. Sturrock. 1968. Astrophys. J. 151: 1155. 323. _ _ , 1971. M.N.R.A.S. 152: I.
283. Harvard-Smithsonian Conference on Stellar Atmospheres. 1964. Proceedings 324. Hundhausen, A. 1972. Coronal Expansion and Solar Wine/. New York:
of the First Conference [S.A.O. Special Report No. 167]. Cambridge, Mass.: Springer-Verlag.
Smithsonian Astrophysical Observatory. 325. Hundt, E., K. Kodaira, J. Schmid-Burgk, and M. Scholz. 1975. Astron. and
284. - - · 1965. Proceedings of the Second Co11fere11ce [S.A.0. Special Report Astrophys. 41: 37.
No. 174]. Cambridge, Mass.: Smithsonian Astrophysical Observatory. 326. Hutchings, J. 1968. M.N.R.A.S. 141 :219.
285. Hayes, D. 1968. Unpublished Ph.D. thesis, Univ. of California, Los Angeles. 327. 1968. M.N.R.A.S. 141 :329.
286. - - · 1970. Astrophys, J. 159: 165, 328. - - · 1970. M.N.R.A.S. 147: 161.
287. Hayes, D., and D. Latham. 1975. Astrophys. J. 197: 593. 329. - - · 1970. M.N.R.A.S. 147:367.
288. Hayes, D., D. Latham, and S. Hayes. 1975. Astrophys. J. 197:587. 330. Hynek, J., (ed.). 1951. Astrophysics: A Topical Symposium. New York:
289. Hearn, A. 1973 .. Astron. and Astrophys. 23:97. McGraw-Hill.
290. - -. 1975. Astron. and Astrophys. 40: 35S. 331. Jackson, J. 1962. Classical Electrodynamics. New York: Wiley.
291. Heasley, J. 1975. Solar Phys. 44:275. 332. Jacobs, V. 1973. Photolonizatlon from Excited States of Helium [NASA
292. Heasley, J., and F. Kneer. 1976. Astrophys. J. 203: 660. Report X-641-73-317). Greenbelt, Md.: National Aeronautics and Space
293. Heitler, W. 1954. Quantum Theory of Radiation. Oxford: Clarendon Press. Administration.
294. Helfer, L., G. Wallerstein, and J. Greenstein. 1959. Astrophys. J. 129: 700. 333. Jefferies, J. 1960. Astrophys. J. 132:775.
295. Henyey, L. 1946. Astrophys. J. 103:332. 334. _ _ . 1968. Spectral Line Formation. Waltham, Mass.: Blaisdel.
578 References References 579

335. Jefferies, J., and R. Thomas. 1958. Astrophys. J. 127:667. 373. Kulander, J., and J. Jefferies. 1966. Astrophys. J. 146: 194.
336. - - · 1959. ·Astrophys. J. 129:401. 374. Kunasz, P., and D. Hummer. 1974. M.N.R.A.S. 166: 19.
337. - - · 1960. Astrophys. J. 131 :695. 375. - - · 1974. M.N.R,A.S. 166:57.
338. Jefferies, J., and 0. White. 1960. Astrophys. J. 132:767. 376. Kunasz, P., D. Hummer, and D. Mihalas. 1975. Astrophys. J. 202:92.
339. Jockers, K. 1970. Astron. and Astrophys. 6:219. 377. Kuperus, M. 1965. The Transfer of Mechanical Energy in the Sun and the
340. John, T. 1967. Astrophys. J. 149:449. Heating of the Corona. Dordrecht: Reidel.
341. Johnson, H. 1974. NCAR Technical Note No. NCAR-TN/STR-95. Boulder: 378. _ _ , 1969. Space Sci. Rev. 9:713.
National Center for Atmospheric Research. 379. Kurucz, R. 1971. S.A.O. Special Report No. 309. Cambridge, Mass.: Smith-
342. Johnson, H., R. Beebe, and C. Snedden. 1975. Astrophys. J. Supp. No. 280 sonian Astrophysical Observatory.
29: 123. 380. _ _ , 1974. Solar Phys. 34: 17.
343. Joos, G. 1959. Theoretical Physics, 3rd ed. New York: Hafner. 381. Kurucz, R., E. Peytremann, and E. Avrett. 1974. Blanketed Model Atmo-
344. Jordan, S., and E. Avrett. 1973. Stellar Chromospheres. Washington, D.C.: spheres for Early-Type Stars. Washington, D.C.: Smithsonian Institution.
National Aeronautics and Space Administration. 382. Kusch, H. 1958. Z.fur Astrophys. 4S: I.
345. Kalkofen, W. 1966. J.Q.S.R.T. 6:633. 383. Lamers, H., and D. Morton. 1976. Astrophys. J. Supp. 32:715.
346. _ _ , 1968. Astrophys. J. 151:317. 384. Lamers, H., and M. Snijders. 1975. Astron. and Astrophys. 41 :259.
347. _ _ , 1974. Astrophys. J. 188: 105. 385. Landau, L., and E. Lifshitz. 1959. Fluid Mechanics. Reading, Mass.:
348. Kalkofen, W., and S. Strom. 1966. J.Q.S.R.T. 6:653. Addison-Wesley.
349. Kamp, L. 1973. Astrophys. J. 180:447. 386. - - · 1962. The Classical Theory of Fields, 2nd ed. Oxford: Pergamon
350. Kaplan, S., and S. Pikelner. 1970. The Interstellar Medium. Cambridge, Press.
Mass.: Harvard Univ. Press. 387. Larson, R. 1969. M.N.R.A.S. 14S:297.
351. Karp, A. 1972. Astrophys. J. 173:649. 388. Lathem, E. 1969. The Poetry of Robert Frost. New York: Holt, Rinehart
352. Karzas, W., and R. Latter. 1961. Astrophys. J. Supp. No.'55 6: 167. and Winston.
353. Kepple, P. 1968. Improved Stark Profile Calculations for the, First Four 389. Lawrence, G. 1967. Astrophys. J. 147:293.
Members of the Hydrogen Lyman and Balmer Series [Univ. of Maryland Report 390. Leckrone, D. 1971. Astron. and Astrophys. 11: 387. I
# 831). College Park: Univ. of Maryland. 391. Leckrone, D., J. Fowler, and S. Adelman. 1974. Astron. and Astrophys. r\)
354. _ _ , 1972. Phys. Rev. A6: I. 32:237. co
355. - - · 1972. Stark Profile Calculations for Ionized Helium Lines [Univ. of 392. Leighton, R. 1959. Principles of Modern Physics. New York: McGraw-Hill. OJ
Maryland Report No. 72-018). College Park: Univ. of Maryland. 393. Leighton, R., R. Noyes, and G. Simon. 1962. Astrophys. J. 13S:474.
394. Lighthill, M. 1952. Proc. Roy. Soc. (London) A211: 564.
I
356. Kepple, P., and H. Griem. 1968. Phys. Rev. 173:317.
357. Klinglesmith, D. 1971. Hydrogen Line Blanketed Model Stellar Atmospheres. 395. _ _ , 1954. Proc. Roy. Soc. (London) A222: I.
Washington, D.C.: National Aeronautics and Space Administration. 396. Limber, D. 1958. Astrophys. J. 127:363.
358. Kneer, F. 1975. Astrophys. J. 200:367. 397. Lindholm, E. 1941. Arkivf. Math. Astron. och Fysik 28B(no. 3).
359. Kondratyev, K. 1969. Radiation in the Atmosphere. New York: Academic 398. - - · 1945. Arkivf. Math. Astron. och Fysik 32A(no. 17).
Press. 399. Lindquist, R. 1966. Ann. Phys. 37:341.
360. Kosirev, N. 1934. M.N.R.A.S. 94:430. 400. Linsky, J. 1970. Solar Phys. 11: 355.
361. Kourganoff, V. 1963. Basic Methods in Transfer Problems. New York: Dover. 401. Linsky, J., and E. Avrett. 1970. P.A.S.P. 82: 169.
362. Kraft, R. 1967. Astrophys. J. 150:551. 402. Lotz, W. 1968. Z.fur Physik 216:241.
363. Kramers, H. 1923. Phil. Mag. 46:836. 403. Lucy, L. 1971. Astrophys. J. 163:95.
364. Krishna-Swamy, K. 1961. Astrophys. J. 134: 1017. 404. Lucy, L., and P. Solomon. 1970. Astrophys. J. 1S9:879.
365. Krook, M. 1955. Astrophys. J. 122:488. 405. Maeder, A. 1971. Astron. and Astrophys. 13:444.
366. Kuan, P., and L. Kuhi. 1975. Astrophys. J. 199: 148. 406. Magnan, C. 1974. J.Q.S.R.T. 14: 123.
367. _ _ , 1976. P.A.S.P. 88: 128. 407. Mark, C. 1947. Phys. Rev. 72:558.
368. Kudritzki, R. 1973. Astron. and Astrophys. 28: 103. 408. Marlborough, J., and J.-R. Roy. 1970. Astrophys. J. 160:221.
369. Kuhi, L. 1973. Astrophys. J. 180:783. 409. Marlborough, J., and M. Zamir. 1975. Astrophys. J. 195: 145.
370. Kulander, J. 1967. Astrophys. J. 147: 1063. 410. Massey, H., and E. Burhop. 1969. Electronic and Ionic Impact Phenomena,
371. 1968. J.Q.S.R.T. 8:273. 2nd ed. (4 vols.). Oxford: Clarendon Press.
372. _ _ , 1971. Astrophys. J. 16S:543. 411. Matsushima, S. 1969. Astrophys. J. 1S8: 1137.
580 References References 581

412. Matsushima, S., and Y. Terashita. 1969. Astrophys. J. 156: 203. 452. Mihalas, D., and M. Stone. 1968. Astrophys. J. 151: 293.
413. McCrea, W. 1928. M.N.R.A.S. 88:729. 453. Miles, B., and W. Wiese. 1969. Critically Evaluated Transition Probabilities
414. McCrea, W., and K. Mitra. 1936. Z.filr Astrophys. 11 :359. for Ba I and Ba II [N.B.S. Technical Note No. 474]. Washington, D.C.: U.S.
415. Menzel, D., (ed.). 1962. Selected Papers on Physical Processes in Ionized Dept. of Commerce.
Plasmas. New York: Dover. 454. ___ 1970. Bibliography on Atomic Transition Probabilities: Januarl' 19/6
416. _ _ (ed.). 1966. Selected Papers on the Transfer of Radiation. New York: through June 1969 [N.B.S. Special Pub. No. 320). Washington, D.C.: U.S.' Dept.
Dover. of Commerce.
417. Menzel, D., and C. Pekeris. 1935. M.N.R.A.S. 96:77. 455. Milkey, R., T. Ayres, and R. Shine. 1975. Astrophys. J. 197: 143.
418. Merzbacher, E. 1970. Quantum Mechanics, 2nd ed. New York: Wiley. 456. Milkey, R., and D. Mihalas. 1973. Astrophys. J. 185:709.
419. Michard, R. 1949. Ann. d'Astrophys. 12:291. 457. _ _ , 1973. Solar Phys. 32:361.
420. Mihalas, D. 1964. Astrophys. J. 140:885. 458. _ _ , 1974. Astrophys. J. 192:769.
421. 1965. Astrophys. J. Supp. No. 92 9: 321. 459. Milkey, R., R. Shine, and D. Mihalas. 1975. Astrophys. J. 199:718.
422. 1965. Astrophys. J. 141: 564. 460. _ _ , 1975. Astrophys. J. 202:250.
423. 1966. Astrophys. J. Supp. No. 11413: I. 461. Milne, E. 1924. Phil. Mag. 47:209.
424. 1966. J.Q.S.R.T. 6:581. 462. Mitchell, A. 1969. Computarional Methods in Partial Differential Equations.
425. 1967. Astrophys. J. 149: 169. London: Wiley.
426. 1967. Astrophys. J. 150:909. 463. Modisette, J. 1967. J. Geophys. Res. 72: 1521.
427. 1968. Astrophys. J. 153:3i7. 464. Molnar, M. 1973. Astrophys. J. 179:527.
428. 1969. Astrophys. J. 157: 1363. 465. Morgan, W., P. Keenan, and E. Kellman. 1943. An Atlas of Stellar Spectra.
429. 1971. Astrophys. J. 170:541. Chicago: Univ. of Chicago Press.
430. 1972. Astrophys. J. 176: 139. 466. Morrison, N. 1976. Astrophys. J. 200: 113.
431. 1972. Astrophys. J. 177: ! I 5. 467. Morton, D. 1967. Astrophys. J. 147: 1017.
432. 1972. NCAR Technical Note NCAR-TN/STR-76. Boulder: National 468. _ _ , 1967. Astrophys. J. 150:535.
Center for Atmospheric Research. 469. Morton, D., E. Jenkins, and R. Bohlin. 1968. Astrophys. J. 154:661.
470. Morton, D., E. Jenkins, and N. Brooks. 1969. Astrophys. J. 155:875.
I
433. _ _ , 1973. P.A.S.P. 85:593. I\)
434. _ _ . 1974. Astron. J. 19: 111 I. 471. Morton, D., and G. Van Citters. 1970. Astrophys. J. 161 :695.
(0
435. Mihal~s. D., and R. Athay. 1973. Ann. Rev. Astron. and Astrophys. 11: 187. 472. Mozer, B., and M. Baranger. 1960. Phys. Rev. i18:626.
473. Miinch, G. 1945. Astrophys. J. 102:385. (0
436. Mihalas, D., and L. Auer. 1970. Astrophys. J. 160: 1161.
437. Mihalas, D., L. Auer, and J. Heasley. 1975. NCAR Technical Note NCAR- 474. 1946. Astrophys. J. 104:87. I
TN/STR-104. Boulder: National Center for Atmospheric Research. 475. - - · 1948. Astrophys. J. 108: 116.
438. Mihalas, D., A. Barnard, J. Cooper, and E. Smith. 1974. Astrophys. J. 476. ___ 1949. Astrophys. J. 109:275.
190:315. 477. Newkirk, G. 1967. Ann. Rev. Astro11. and Astrophys. 5:213.
439. 1975. Astrophys. J. 197: 139. 478. Noerdlinger, P., and G. Rybicki. 1974. Astrophys. J. 193:651.
440. Mihalas, D., and D. Hummer. 1973. Astrophys. J. 179:827. 479. Nordlund, A. 1974. Astron. and Astrophys. 32:407.
441. _ _ , 1974. Astrophys. J. Lel/ers 189:L39. 480. Norris, J. 1970. Astrophys. J. Supp. No. 176 19:305.
442. - - · 1974. Astrophys. J. Supp. No. 265 28:343. 481. _ _ , 1971. Astrophys. J. Supp. No. 197 23: 193.
443. Mihalas, D., D. Hummer, and P. Conti. 1972. Astrophys. J. Lellers 175: L99. 482. O'Brien, J., and C. Hooper. 1972. Phys. Rev. A5:867.
444. Mihalas, D., P. Kunasz, and D. Hummer. 1975. Astrophys. J. 202:465. 483. Oertel, G., and L. Shomo. 1968. Astrophys. J. Supp. No. 145 16: 175.
445. 1976. Astrophys. J. 203:647. 484. Oke. J. 1960. Astrophys. J. 131: 358.
446. _ _ , 1976. Astrophys. J. 206:515. 485. _ _ , 1964. Astrophys. J. 140: 189.
447. _ _ , 1976. Astrophys. J. 210:419. 486. _ _ , 1965. A11n. Rev. Astron. and Astrophys. 3:23.
448. Mihalas, D., a·nd W. Luebke. 1971. M.N.R.A.S. 153:229. 487. Oke, J., and R. Schild. 1970. Astrophys. J. 161: 1015.
449. Mihalas, D., and D. Morton. 1965. Astrophys. J. 142:253. 488. Olson, E. 1974. P.A.S.P. 86:80.
450. Mihalas, D., B. Pagel, and P. Souffrin. 1971. Theorie des Atmospheres 489. Omont, A., E. Smith, and J. Cooper. 1972. Astrophys. J. 175: 185.
Ste/laires. Geneva: Observatoire de Geneve. 490. Owczarek, J. 1964. Fundamentals of Gas Dynamics. Scranton: International.
451. Mihalas, D., R. Shine, P. Kunasz, and D. Hummer. 1976. Astrophys. J. Textbook.
205:492. 491. Pagel, B. 1959. M.N.R.A.S. 119:609.
582 References
References 583

492. Pagel, B. 1968. Proc. Roy. Soc. (London) A306:91.


_ _ , 1973. Space Sci. Rev. 15: I. 533. Rogerson, J., and H. Lamers. 1975. Nature 256: 190.
493.
534. Rohrlich, F. 1959. Astrophys. J. 129:441.
494. Panofsky, W., and M. Phillips. 1962. Classical Electricity and Magnetism, 2nd
535. - - · 1959. Astrophys. J. 129:449.
ed. Reading, Mass.: Addison-Wesley.
536. Rosendhal, J. 1973. Astrophys. J. 182: 523.
495. Pannekoek, A. 1922. Bull. Astr. Inst. Netherlands 1: 107.
537. - - · 1973. Astrophys. J. 186:909.
496. Parker, E. 1958. Astrophys. J. 128:664.
538. Ross, J., and L. Aller. 1968. Astrophys. J. 153:235.
497. 1960. Astrophys. J. 132:821.
539. Roueff, E. 1970. Astron. and Astrophys. 7:4. ·
498. - - · 1963. Interplanetary Dynamical Processes. New York: Interscience.
540. - - · 1975. Astron. and Astrophys. 38:41.
499. - - · 1965. Astrophys. J. 141: 1463.
541. Roueff, E., and H. Van Regemorter. 1969. Astron .. and Astrophys. 1: 69.
500. Parsons, S. 1969. Astrophys. J. Supp. No. 15918; 127.
542. Russell, H. 1929. Astrophys. J. 70: 11.
501. Payne, C. 1925. Stellar Atmospheres. Cambridge, Mass.: Harvard Univ.
543. Rybicki, G. 1971. J.Q.S.R.T. 11:589.
Press.
544. Rybicki, G., and D. Hummer. 1967. Astrophys. J. 150:607.
502. Peach, G. 1962. M.N.R.A.S. 124:371.
545. 1969. M.N.R.A.S. 144:313.
503. - - · 1967. Mem. R.A.S. 71:13.
546. Saha, M. 1920. Phil. Mag. 40:472.
504. - - · 1970. Mem. R.A.S. 73: I.
547. - - · 1921. Proc. Roy. Soc. (London) A99: 135.
505. Pecker, J.-C. 1951. Ann. d'Astrophys. 14:383.
548. Sahal-Brechot, S. 1969. Astron. and Astrophys. 2: 322.
506. - - · 1965. Ann. Rev. Asfron. and Astrophys. 3: 135.
549. Sahal-Brechot, S., and E. Segre. 1971. Astron. and Astrophys. 13: 161.
507. Peraiah, A., and I. Grant. 1973. J. Inst. Math. Applies. 12:75.
550. Sahal-Brechot, S. 1974. Astron. and Astrophys. 35: 319.
508. Peters, G. 1976. Astrophys. J. Supp. 30:551.
551. Sampson, D. 1965. Radiative Contributions to Energy and Momentum Trans-
509. Peterson, D. 1968. S.A.O. Special Report No. 293. Cambridge, Mass.:
port in a Gas. New York: Interscience.
Smithsonian Astrophysical Observatory.
552. Sargent, W. 1964. Ann. Rev. Astron. and Astrophys. 2:297.
510. Peterson, D., and M. Scholz. 1971. Astrophys. J. 163: 51.
553. Scarf, F., and L. Noble. 1965. Astrophys. J. 141: 1479.
511. Peterson, 0., and S. Strom. 1969. Astrophys. J. 157: 1346.
554. Schatzman, E. 1962. Ann. d'Astrophys. 2S: 18.
512. Peytremann, E. 1974. Astron. and Astrophys. 33:203.
555. Schild, R., D. Peterson, and J. Oke. 1971. Astrophys. J. 166:95. I
513. - - · 1974. Astron. and Astrophys. Supp. 18:81.
514. - - · 1975. Astron. and Astrophys. 38:417.
556. Schmeltekopf, A., F. Fehsenfeld, and E. Ferguson. 1967. Astrophys. J. w
Letters 148: LISS. 0
515. Pfennig, H., and E. Trefftz. 1966. Z.fur Phys. 190:253.
557. Schmid-Burgk, J. 1974. Astron. and Astrophys. 32:73. 0
516. Philip; A., and D. Hayes. 1975. Multicolor Photometry and the Theoretical
558. - - · 1975. Astron. and Astrophys. 40:249.
HR Diagram [Dudley Obs. Report No. 9]. Albany, N.Y.: Dudley Observatory.
Schmid-Burgk, J., and M. Scholz. 1975. Astron. and Astrophys. 41:41.
I
559.
517. Pierce, A., and J: Waddell. 1961. Mem. R.A.S. 58: 89.
560. Scholz, M. 1972. Vistas in Astron. 14:53.
518. Placzek, G. 1947. Phys. Rev. 72:556.
561. Schrodinger, E. 1926. Ann. d. Phys. 80:437.
519. Placzek, G., and W. Seidel. 1947. Phys. Rev. 72:550.
562. Schuster, A. 1905. Astrophys. J. 21: I.
520. Planck, M. 1959. The Theory of Heat Radiation. New York: Dover.
563. Schwarzschild, K. 1916. Sitzber. Deutsch. Akad. Wiss. (Berlin), p. 584.
521. Pomraning, G. 1973. Radiation Hydrodynamics. Oxford: Pergamon Press.
564. Schwarzschild, M. 1948. Astrophys. J. 10S: I.
522. Preston, G. 1971. P.A.S.P. 83:571.
565. Sears, F. 1953. Thermodynamics, The Kinetic Theory of Gases, and Statistical
523. - - · 1973. Ann. Rev. Astron. and Astrophys. 11: 115.
Mechanics. Reading, Mass.: Addison-Wesley.
524. Prokof'ev, V. 1962. Soviet Phys.-Doklady 6: 861.
566. Seaton, M. 1958. M.N.R.A.S. 118:504.
525. Proudman, I. 1952. Proc. Roy. Soc. (London) A214: I 19.
567. Shimooda, H. 1973. P.A.S. Japan 25:547.
526. Ralston, A. 1965. A First Course in Numerical Analysis. New York: McGraw-
568. Shine, R. I975. Astrophys. J. 202: 543.
Hill.
569. Shine, R., and J. Linsky. 1974. Solar Phys. 39:49.
527. Reichel, A. 1968. J.Q.S.R.T. 8: 1601.
570. Shine, R., R. Milkey, and D. Mihalas. 1975. Astrophys. J. 199:724.
528. - - · 1969. Math. Comp. 23:645.
571. - - · 1975. Astrophys. J. 201 :222.
529. Reichel, A., and I. Vardavas. 1975. J.Q.S.R.T. 1S:929.
572. Shore, B., and D. Menzel. 1968. Principles of Atomic Spectra. New York:
530. Richtmyer, R., and K. Morton. 1967. Difference Methods for Initial-Value
Wiley.
Problems, 2nd ed. New York: Interscience. · Shaub, E. 1977. Astrophys. J. Supp. 34:259.
573.
531. Roberts, P. 1971. Astrophys. letters 9:19.
574. Simonneau, E. 1973. Astron. and Astrophys. 29: 357.
532. Roberts, P., and A. Soward. 1972. Proc. Roy. Soc. (London) A328: 185. Skumanich, A., and B. Domenico. 1971. J.Q.S.R.T. 11: 547.
575.
584 References References 585

576. Slater, J. 1960. Quantum Theory of Atomic Structure, Vol. 1. New York: 619. Sykes,J. 1951. M.N.R.A.S.111:377.
McGraw-Hill. 620. Terashita, Y., and S. Matsushima. 1966. Astrophys. J. Supp. No. 12113:461.
577. - - · 1960. Quantum Theory of Atomic Structure, Vol. 2. New York: 621. Thomas, L. 1930. Quart. J. Math. 1: 239.
McGraw-Hill. 622. Thomas, R. 1957. Astrophys. J. 125:260.
578. Smirnov, B. 1967. Soviet Phys. J.E.T.P. 24:314. 623. _ _ , 1960. Astrophys. J. 131 :429.
579. Smith, A. 1969. Astrophys. J. 156:93. 624. _ _ (ed.). 1960. Aerodynamic Phenomena in Stellar Atmospheres [I.A.U.
580. _ _ , 1972. Astrophys. J. 176:405. Symposium No. 12]. Bologna: N. Zanichelli.
581. Smith, E., J. Cooper, and C. Vidal. 1969. Phys. Rev. 185: 140. 625. - - · 1965. Astrophys. J. 144:333.
582. Smith, E., C. Vidal, and J. Cooper. 1969. J. Res. Nat. Bur. Standards 73A: 626. - - . 1965. Some Aspects of Non-Equilibrium Thermodynamics in the Pres-
389. ence of a Radiation Field. Boulder: Univ. of Colorado Press.
583. Smith, M., and S. Strom. 1969. Astrophys. J. 158: 1161. 627. - - · 1973. Astron. and Astrophys. 29:297.
584. Smith, M., and W. Wiese. 1973. J. Phys. Chem. Ref. Data 2:85. 628. Thomas, R., and R. Athay. 1961. Physics of the Solar Chromosphere. New
585. Sneden, C., H. Johnson, and B. Krupp. 1976. Astrophys. J. 204: 281. York: 1nterscience.
586. Snijders, M., and H. Lamers. 1975. A.\'/ro11. a11d Astrophys. 41 :245. 629. Truving, G. 1960. Uhc•r clic• Theorie da Druckverbreit,mmy vo11 Spektrallinie11.
587. Snijders, M. 1977. Astron. and Astrophys. 60:377. Karlsruhe: Verlag G. Braun.
588. _ _ , 1977. In press. 630. Ulmschneider, P. 1967. Z.fiir Astrophys. 67: 193.
589. Snow, T., and D. Morton. 1976. Astrophys. J. Supp. 32:429. 631. - - - · 1971. Astron. and Astrophys. 12:297.
590. Sobolev, V. 1957. Soviet Astron. 1 :678. 632. - - - · 1971. Astron. and Astrophys. 14:275.
591. - - · 1960. Moving Envelopes of Stars. Cambridge, Mass.: Harvard Univ. 633. Underhill, A. 1949. M.N.R.A.S. 109:563.
Press. [Russian edition, 1947.J 634. Underhill, A., and J. Waddell. 1959. Stark Broadening Functions for the
592. Somerville, W. 1964. Astrophys. J. 139: 192. Hydrogen Lines [N.B.S. Circular No. 603). Washington, D.C.: U.S. Dept. of
593. - - · 1965. Astrophys.J.141:811. Commerce.
594. Spiegel, E. 1971. Ann. Rev. Astron. and Astrophys. 9: 323. 635. Unno, W., and M. Kondo. 1976. P.A.S. Japan 28:347.
595. _ _ , 1972. Ann. Rev. Astron. and Astrophys. 10:261. 636. Unsold, A. 1931. Z.fiir Astrophys. 1: 138.
637. - - · 1943. Vierteljahresschrift der Astron. Gess. 78:213.
I
596.
597.
Spitzer, L. 1943. Astrophys. J. 98: 107.
_ _ , 1944. Astrophys. J. 99: I. 638. _ _ . 1955. Physik der Sternatmospharen, 2nd ed. Berlin: Springer-Verlag. w
598. - - · 1956. Physics of Fully Ionized Gases. New York: Wiley. 639. Van Regemorter, H. 1962. Astrophys. J. 136:906. 0
640. Vardavas, I. 1976. J.Q.S.R.T. 16: I. .......
599. Stacey, D., and J. Cooper. 1971. J.Q.S.R.T. 11: 1271.
600. Stecher, T. 1970. Astrophys. J.159:543. 641. - - · 1976. J.Q.S.R.T. 16:715.
601. Stein, R. 1968. Astrophys. J. 154:297. 642. Vardavas, I., and L. Cram. 1974. Solar Phys. 38: 367.
602. Stein, R., and J. Leibacher. 1974. Ann. Rev. Astron. and Astrophys. 12:407. 643. Vardya, M. 1965. M.N.R.A.S. 129:205.
603. Stewart, A., and T. Webb. 1963. Proc. Phys. Soc. (London) 82:532. 644. - - · 1967. Mem. R.A.S. 71:249.
604. Stilley, J., and J. Callaway. 1970. Astrophys. J. 160:245. 645. Vernazza, J., E. Avrett, and R. Loeser. 1973. Astrophys. J. 184:605.
605. Stone, P., and J. Gaustad. 1961. Astrophys. J. 134:456. 646. - - · 1976. Astrophys. J. Supp. 30: I.
606. Stothers, R. 1966. Astrophys. J. 144:959. 647. Vidal, C., J. Cooper, and E. Smith. 1970. J.Q.S.R.T. 10: 1011.
607. Strom, S. 1967. Astrophys. J. 150:637. 648. - - · 1971. J.Q.S.R.T. 11:263.
608. Strom, S., and E. Avrett. 1965. Astrophys. J. Supp. No. 10312: I. 649. - - · 1971. Unified Theory Calculations of Stark Broadened Hydrogen
609. Strom, S., 0. Gingerich, and K. Strom. 1966. Astrophys. J. 146:880. Lines Including Lower State Interactions [N. B.S. Monograph No. 120]. Washing-
610. Strom, S., and W. Kalkofen. 1966. Astrophys. J. 144:76. ton, D.C.: U.S. Dept. of Commerce.
611. _ _ , 1967. Astrophys. J. 149: 191. 650. _ _ . 1973. Astrophys. J. Supp. No. 214 25: 37.
612. Strom, S., and R. Kurucz. 1966. J.Q.S.R.T. 6:591. 651. Vitense, E. 1951. Z.fur Astrophys. 28:81.
613. Stromgren, B. 1935. Z.filr Astrophys. 10:237. 652. - - · 1953. Z.fiir Astrophys. 32: 135.
614. Struve, 0. 1929. Astrophys. J. 69: 173. 653. - - - · 1958. Z.fur Astrophys. 46: 108.
615. _ _ , 1929. Astrophys. J. 70:85. 654. Voigt, H., (ed.). 1965. Landolt-Bornstein Numerical Data and Functional
616. - - · 1946. Astrophys. J. 104: 138. Relationships in Science and Technology, new series, group VI, vol. I. Berlin:
617. Struve, 0., and C. Elvey. 1934. Astrophys. J. 79:409. Springer-Verlag.
618. Sturrock, P., and R. Hartle. 1966. Phys. Rev. Letters 16:628. 655. Waddell, J. 1962. Astrophys. J. 136:231.
586 References

656. Waddell, J. 1963. Astrophys. J. 138: 1147.


657. Walborn, N. 1971. Astrophys. J. Supp. No. 198 23: 257.
658. Wallerstein, G., and P. Conti. 1969. Ann. Rev. Astron. and Astrophys. 7:99.
659. Weber, E., and L. Davis. 1967. Astrophys. J. 148:217.
660. - - · 1970. J. Geophys. Res. 75:2419.
661. Weisskopf, V. 1932. Z.filr Phys. 75:287.
662. _ _ , 1933 ... Physik. Z. 34: I.
663. - - · 1933. Observatory 56:291.
664. Wendroff, B. 1963. Los Alamos Scientific Loboratory Report No. LAMS-2795. Glossary of Physical Symbols
665. Weymann, R. 1960. Astrophys. J. 132:380.
666. _ _ , 1962. Astrophys. J. 136:476.
667. Whang, Y., and C. Chang. 1965. J. Geophys. Res. 70:4175.
668. Whang, Y., C. Liu, and C. Chang. 1966. Astrophys. J. 145:255,
Physical symbols used in the text are listed below, along wit.h a brief description of
669. White, O. 1968. Astrophys. J. 152:217.
their meaning and the page number on which each first appears. Standard mathematical
670. Wiese, W., and J. Fuhr. 1975. J. Phys. Chem. Ref. Data 4:263.
symbols, dummy variables and indices, and notations used only in one location are
671. Wiese, W., D. Kelleher, and D. Paquette. 1972. Phys. Rev. A6: 1132.
672. Wiese, W., M. Smith, and B. Glennon. 1966. Atomic Transition Probabilities, not included.
Vol. I: Hydrogen through Neon [NSRDS-NBS-4). Washington, D.C.: U.S. Dept.
of Commerce.
Ratio of damping width to Doppler width, r/4n Av0 279
673. Wiese, W., M. Smith, and B. Miles. 1969. Atomic Transition Probabilities, a
Vol. 2: Sodium through Calcium [NSRDS-NBS-22). Washington, D.C.: U.S. Isothermal sound speed 519
a
Dept. of Commerce. Quadrature weight 65
674. Wilson, 0. 1934. Astrophys. J. 80:259. a,
Branching ratio j --+ i in radiative decay of state j 142
675. - - · 1963. Astrophys. J.138:832. 011
Radiation constant in Stefan's law 7
676. Wilson, 0., and V. Bappu. 1957. Astrophys. J. 125:661. QR
519 I
Wilson, 0., and A. Skumanich. 1964. Astrophys. J. 140: 1401. Adiabatic sound speed
677.
678. Wilson, R. 1958. P. Roy. Obs. Edinburgh 2:61.
a,
Macroscopic absorption coefficient uncorrected for
w
a, 78 0
679. Wilson, S., and K. Sen. 1965. Ann. d'Astrophys. 28:348. stimulated emission
148 t\.)
680. - - · 1965. Ann. d' Astrophys. 28:855. a, Coefficient in linear expansion of Planck function
681. Wolff, C., J. Brandt, and R. Southwick. 1971. Astrophys. J. 165: 181. 89 I
ao Bohr radius
682. Wolff, S., L. Kuhi, and D. Hayes. 1968. Astrophys. J. 152:871.
ai(t) Coefficient of eigenstate iJ,1 in expansion of a general
683. Woolley, R. 1938. M.N.R.A.S. 98:624. 85
state v,(t)
684. Woolley, R., and D. Stibbs. 1953. The Outer Loyers of a Star. Oxford:
a,(µ) Absorption-depth, relative to continuum, in spectrum-line
Clarendon Press. intensity-profile at angle cos-• µ from disk-center,
685. Wright, A., and M. Barlow. 1975. M.N.R.A.S. 170:41. 269
1 - r,(µ)
686. Wright, K. 1948. P. Dominion Astrophys. Obs. Victoria 8: I. 7
687. Wrubel, M. 1949. Astrophys. J. 109:66. A Vector potential
Rate matrix of statistical equilibrium equations 138
688. - - · 19S0. Astrophys. J. 111: 157. d
689. _ _ . 1954. Astrophys. J.119:51. Autoionization transition probability 134
A.
690. Zanstra, H. 1941. M.N.R.A.S. 101 :273. Atomic weight of chemical species i 110
A,
691. - - · 1946. M.N.R.A.S. 106:22S. Einstein spontaneous-emission probability for transition
692. Zeldovich, Ya., and Yu. Raizer. 1966. Physics of Shock Waves and High- A11 78
j ... i
Temperature Hydrodynamic Phenomena (2 vols.). New York: Academic Press. Stabilization transition probability in dielectronic
693. Zinn, R. 1970. Astrophys. J. 162:909. A, 135
recombination process
694. Zirin, H. 1966. The Solar Atmosphere. Waltham, Mass.: Blaisdell. Absorption-depth, relative to continuum, in spectrum line
A, 269
flux-profile, 1 - R,
588 Glossary of Physical Symbols Glossary of Physical Symbols 589

Ao Central absorption depth of an infinitely opaque line 318 elm• Matrix element of dipole moment (</>!Jdl</>.) 86
Ad Matrix coupling depth-points d - I and cl in Feautrier ,/3r Volume element 8
difference-equation solution of transfer equation 155 D Distance from star to observer II
b, Non-LTE departure coefficient (n 1/nt) for level i 219 D Debye length, Debye radius 122
b. Coefficient in linear expansion of Planck function 148 D Electric displacement 7
b0 (t) Planck function normalized to value at surface of Balmer jump (in magnitudes) 195
Ds
atmosphere B,[T(t)]/B,(T0 ) 76
Dp Paschen jump (in magnitudes) 195
b.(,.) Planck function normalized to disk-center intensity
B,[T(t,)]/1,(0, I) ' 262 DJ Auxiliary matrix used in developing Fea:utrier difference-
equation solution of transfer equation 156
B Magnetic induction 7
(D/Dt) Fluid-frame or Lagrangian derivative 514
f1A Righthand-side vector of statistical equilibrium equations 138
(D/Dt),., 11 Time rate of change of distribution function caused by
Bil Einstein absorption probability for transit ion ; .... j 77 collisions 33
BJ, Einstein induced-emission (or stimulated-emission) Electron charge 81
e
probability for transition J .... i 78
e Specific internal energy of a fluid 517
Bo Planck function at T = T0 312
e(oo) Residual energy per particle at infinite distance in stellar
81 (oB.!il'f) evaluated at l = 0 312 wind, E/F 531
B* Equivalent Planck function for recombination emission in
E Total particle energy 494
non-LTE source function 360
E Total energy flux in stellar wind 526
B, Coupling matrix at depth-point cl in Feautrier difference-
0

equation solution of transfer equa tion 155 E Electric field 7


B(T) Frequency-integrated Planck function 52 E Matrix giving definition of J in terms of u,'s in Rybicki
difference-equation solution of transfer equation 159 I
B.rr{t) Effective thermal source in line with overlapping continuum 352
I tf Energy in radiation field 3 w
B,(T) Planck function 7 0
E, Energy of atomic state i relative to ground state 22
B:[T(t)] Planck function for temperature distribution that gives
radiative equilibrium 179 E, Contribution to emissivity from (fixed) overlapping
w
transitions 397 I
C Velocity oflight 4
E. Energy received at frequency v from. an expanding
c,J Collisional transition rate from level ; to level J 127 atmosphere by an external observer 474
c,K Collisional ionization rate oflevel i to continuum 123 Eo Magnitude of electric field 8
Ct Interaction coefficient of kth component of hydrogen Stark Threshold energy of collision process 132
pattern 295 Eo
E; Thermodynamic equilibrium value of radiation energy-
cp Specific heat at constant pressure 186 density 7
Cp, C3, C,, C~ Coefficien! of power-law expression for perturber-radiator
E(w) Energy spectrum of oscillator 275
interactton 283
E,(oo) Heat-conduction flux at infinite distance in stellar wind 531
c,. Specific heat at constant volume 186
40
E.(x) Exponential integral of order n
Co Numerical constant in collision rate formula, 1t11/(8k/m1t)I 133
c, Matrix coupling depth-points cl and cl + I in Feautrier E,(,, ,, ''}
Monochromatic energy density in radiation field 6
. difference equation solution of transfer equation 155 ER(r, v),
'C, Net collisional rate into level i 486 ER(v)
d Electron-impact line shift 306 E,(,,
ER(r),
•h} Total (frequency-integrated) energy density in radiation field 6
d Electric dipole moment 86
d, Departure of ratio of non-LTE to LTE population of level ; ER
from unity, (b, - I) 219 f Force-per-unit-volume acting on a fluid 516
590 Glossary of Physical Symbols Glossary of Physical Symbols 591

fi,
J;
Buoyancy force
Continuum oscillator strength
188
123
... ,. ' }
(,.
:!F(r,v), Monochromatic vector flux or radiation field 9
fl} Oscillator strength of transition i .... J g;;.
84
I, Monochromatic flux received by observer S-ne(v) Flux from blackbody 12
11
f(n',n) Oscillator strength for transition between states with
g Surface gravity in planar atmosphere 170
principal quantum numbers n' and n 88 gcrit Surface gravity at which radiation force exceeds
f(n', /'; n, /) Oscillator strength for transition from substate /' of state n'
gravitational force in atmosphere 170
to substate / of state n 88 g.,r Net acceleration, gravitational minus radiative, of stellar
f(t) Amplitude of time-variation of oscillator
atmospheric material 256
274
f(v), f(v) Maxwellian velocity distribution g, Statistical weight of atomic state i 77
110
gl}k Statistical weight of excitation state i of ionization state J of
f(r, p, t),}
Particle distribution function 32 chemical species k 110
/(r, V, 1),
g., Statistical weight of substate / of state n 88
f(,. ,. ,, } gR Acceleration produced by radiation force S54
f(z, v, t), Variable Eddington factor K./J. 18 gR,I Radiative acceleration produced by all spectrum lines S58
J. gRO Acceleration produced by radiation force in a single
/ik(n., T) Fraction of chemical species k in ionization stage J, NJk/Nk 114 spectrum line 5S5
fK(n', n) Kramers'-formula oscillator strength for transition n' .... n gR Four-force of radiation on matter 498

fR(r, n, v, t)
in hydrogen
Photon distribution function
90
4
g(,',
g(µ', µ,
•l'}
<//), Angular phase functions in scattering process 29
F Perturber field-strength 291 g(Jt', Jt)
F Particle flux in stellar wind 525 I
g1(n', n) Gaunt factor, for bound-bound transition n' .... n in
F Imposed external force 32 hydrogen 90 w
:F, §' Frequency-integrated flux 44 9u(n, k),} Gaunt factor for bound-free transition n -+ k in hydrogen 99
0

t Flux in continuum near spectrum line 269 g11 (n, v) ~
Fcunw Energy flux transported by convection 188 g,,(k, /)'} I
F,.d Energy flux transported by radiation 190 g111 (v, v), Gaunt factor for free-free transitions in hydrogen IOI
F" Four-force S41 g111 (v, T)
Fo Normal field strength 291 G Newtonian gravitation constant 255
F,ad Radiative damping force on an oscillator 82 G.m Momentum density in electromagnetic field 14
3-oa Frequency-integrated flux from blackbody 12 GR Momentum density in radiation field 10
F(v) Spontaneous recapture probability for electrons of velocity v 94 G(v) Induced recapture probability for electrons of velocity v 94
F(v) Dawson's function 280 G(v),} Generalized statistical weight ratio in stimulated emission
G,.(v) correction 165
F(,, ,, ''} §' .In Monochromatic "astrophysical" flux or radiation
F(z, v, t),
field h Planck's constant 4
10
F, h Reduced Planck's constant h/21t 85
F(w) Fourier transfrom 275 h Specific enthalpy of a fluid 518
F(w, T) Fourier transfrom of wavetrain of duration T 281 Ii, Eddington factor giving ratio H,(0)/J, at surface of
.F(,, ,. ,)'} atmosphere 157
§'(z, v,t), Monochromatic flux or radiation field in direction normal H Frequency-integrated Eddington flux 54
to atmospheric layers 10 188
s-. H Pressure scale height
592 Glossary of Physical Symbols Glossary of Physical Symbols 593

H Magnetic field 7 J(r, 1), J Frequency-integrated mean intensity, f0· J, dv 6


.f( Nominal Eddington flux, 11R r:rr/41t 174
HA Hamiltonian for atom 85 J(r, v),
J(,, '·''·}
Hp Perturber Hamiltonian 298 J(z, v), Mean intensity 5
H,o, Ho Current value of Eddington flux in Avrett-Krook procedure 174 J,(t),
Ho Magnitude of magnetic field 8 ),,
Ho Flux-constant in extended atmosphere, r 2 H = L/16n 2 245 k Boltzmann's constant 7
H(a,v) Voigt function 279 k Wavenumber 8
H(q,,p;) Hamiltonian operator 85 k Continuum-state quantum number for hydrogen 98
k Unit vector in z-direction 3
H(,, ,, ''}
H(z, v, r), Monochromatic Eddington flux, ¼F, = ff./41t IO k Continuum opacity uncorrected for stimulated emission 321
H,.
' Root or characteristic equation in discrete-ordinate method 66
k.
H(11) Limb-darkening function 70 K Frequency-integrated second moment of radiation field 55
H.(v) Expansion function in power-series expression for Voigt .% Numerical coefficient in hydrogen cross-section 99
function 280 K; Discrete representation of depth-variation of thermal source
Unit vector in x-direction 3 terms at angle-frequency point i in Rybicki difference-
I Frequency-integrated specific intensity 54 equation solution of transfer equation 159
I Moment of inertia of a star 536 K(,, ,, ''}
I, Specific intensity emitted from core.in expanding, extended K(z, v, 1), Second angular moment of monochromatic radiation field 16
atmosphere 478 K,
Jk Fractional intensity of kth component of hydrogen Stark K,(,) Line-formation kernel function for an expanding I
pattern 296 atmosphere 481 (.,J
I

1.1 Total line intensity in transition u -+ I 488 K1(,) Line-formation kernel function 339 0
Ionization energy of hydrogen 133 Kernel for line-formation with overlapping continuum 351 01
'}
11;1 K 1.,(,)
89 I
/(,, ~ ,, I Azimuthal quantum number
l,(T,µ), I Continuum-state quantum number for hydrogen 100
Specific intensity of radiation 2
/,(µ), Convective mixing length 188
I, Correlation length in turbulent velocity field 464
/(w) Power spectrum of oscillator 275 I,. Photon mean-free-path 51
/,.(p, oo) Emergent specific intensity along ray with impact parameter L Stellar luminosity 49
p in extended atmosphere 247 L Photon destruction length 333
r+(µ, v), r+ Specific intensity traveling in + µ direction 36 L Rotational angular momentum or a star 536
r(11, v), r Specific intensity traveling in - 11 direction 36 L Total orbital angular momentum of atom 92
j Unit vector in y-direction 3 L, L p• Lorentz transformation 493
Current density 7 Critical luminosity at which radiation force exceeds
Lcr11
J Jacobian of transformation of coordinates 32 gravitational force in atmosphere 171
J Total angular momentum of atom 92 L. Integration constant in discrete-ordinate method 67
),°, jO Current value of mean intensity in Avrett-Krook procedure 174 Ld Source-term in Feautrier difference-equation solution of
155
J, J,1 Mean intensity averaged over line profile, J<J,,J, dv 129 transfer equation
L1,,(t) Continuum kernel function 351
J Discrete representation of depth-variation of J(z) in
Rybicki difference-equation solution of transfer equation 159 m Mass of electron 83
594 Glossary of Physical Symbols Glossary of Physical Symbols 595

m Magnetic quantum number 89 p Impact parameter of ray in extended atmosphere 247


m Column mass in atmosphere 170 p Exponent in power-law expression for perturber-radiator
m, Mass of electron 89 interaction 283
m, Mass of proton 89 p Momentum of a particle 32
mH ·Mass of hydrogen atom 170 P, Electron pressure 103
mo Rest (proper) mass of a particle 494 Pg Total gas pressure 115
fll Average mass per nucleus (atoms + ions) 170 Pt Generalized momentum coordinate 85
M Mach number 520 Pt Pressure in interstellar medium 533
.I( Stellar mass 171 P11 Cascade probability of state j to state i 142
M. Integration constant in discrete ordinate method 67 Pt Partial pressure of particle species k in a gas 514
.Ii Mass-loss rate 516 Pv Probability ofphotoionization at frequency v 94
M(t) Radiation force multiplier 561 p. Coefficient in linear expansion of Planck function on t, scale 310
n Index of refraction 4 p(x,x') Joint probability of absorption from substate x and return
to substate x' in a line transition 277
n Principal quantum number 89
p(f.~) Redistribution probability in atom's rest frame 412
n, n' Directions of radiation propagation 2
n Occupation-number solution-vector of statistical P,(,, ,, <)'}
equilibrium equations 138 PR(z, v, t), Monochromatic radiation pressure scalar 16
n, Number density in doubly excited state of dielectronic PR(v)
recombination process 135
P1(r, v, t)'} Thermodynamic equilibrium value of monochromatic
n, Number density of free electrons 94
p;(z, v, t) radiation pressure scalar 17 I
nt Number density ofatoms in state i 78
nf LTE value of number density of atoms in state i '79 p Radiation-pressure tensor 12 w
n/Jk Number density of atoms in excitation state i or ionization pd Photon destruction probability 333 0
stage j of chemical species k Component ij of radiation pressure tensor 12 0)
110 Pu
n, Proton number density 138 Pu Total transition rate from level i to level j 128 I
no Propagation vector of plane wave 8 P" Four-momentum 494
n(t) Turbulent eddy density 464 p Mean radiation pressure 13
nt(r, t) Particle density of particle species k in a gas 513 P(t) Power radiated from accelerating charge 82
nt(v) Number density of atoms in state i capable of absorbing P(uh) Probability distribution function for dimensionless velocity
radiation at frequency v 77 in turbulent atmosphere 464
il',(v) Ratio of population in substate v of state i to line profile, (P(w)) Average power radiated at circular frequency w by
n1(v)/<f,(v) 436 harmonic oscillator 82
N Total particle density (all species) 115 P,(t) Photon escape probability 334
N Number density of perturbers 282 P.t(r) Radial charge density 89
NJk Number density of atoms in all excitation states of q Heat delivered to a gas, per unit volume 517
ionization stage j of chemical species k Ill q, Generalized space coordinate 85
Nk Number density of atoms of chemical species k in all q. Sphericality factor in extended atmosphere 251
excitation and ionization states 114 Conductive heat flux 517
q,
NN Number density of nuclei (atoms + ions) 115 q(,) Hopf function 55
N, Third angular moment of monochromatic radiation field 502
CJu(TJ Collision rate i .... j, per atom in state i, per electron,
..,Vk,n Number of transitions k .... m 87 averaged over Maxwellian velocity distribution at
p Total pressure 170 temperature T 132
596 Glossary of Physical Symbols Glossary of Physical Symbols 597
q,(t) Exponential absorption factor for a turbulent atmosphere 465 .'Jf; Net radiative into level i 486
(q.(,))s Static average of q,(t) 466 R(x',x) Angle-averaged redistribution function in dimensionless
Q Integration constant in discrete-ordinate method 67 frequency uni ts 427
Q Factor correcting for ionization and radiation pressure R(x', n'; x, n) Redistribution function in dimensionless frequency units 418
effects on mean molecular weight of a gas 188 R(1•', v) Angle-averaged redistribution function 28
Q,J Collision cross-section in units of 1raa2 132 R(v', n'; v, n) Redistribution function for scattering process 27
Q(r,µ) Derivative of radial velocity along a line of sight in Sobolev R,(v', n'; v, n) Redistribution function in laboratory frame for coherent
method 479 scattering in atom's frame 418
Q,(s) Laplace transform of q.(,) 466 R,.i(,·) Radial wavefunction 89
r Radial distance from center of a star 3 R.(v', v) Angle-averaged redistribution function for atom moving
r Distance between two test points 4 with (dimensionless) velocity u 425
r Ratio of continuum to line opacity x,IX, 36 R,,(v', n'; v, n) Redistribution function for atom moving with velocity v 416
r Opacity ratio in schematic Lyman continuum problem 222 s Path length 31
r Position in a stellar atmosphere 2 s Spin quantum number 89
r, Core radius in extended atmosphere 251 SI! Electron scattering coefficient per gram of stellar material 554
r, Critical radius in transsonic wind 526 s Surface area 3
r, Sonic point radius 562 s Frequency-integrated source function 54
rA Alfvenic radius 535 s Poynting vector It
'o Mean interatomic distance 111 s Total spin angular momentum of atom 92
ro Radial coordinate of surface of constant radial velocity,-- s, Line source function 80
(zo2 + p2)t . 479 sm.. Maximum source function in finite slab I
347
ri Radius at Rosseland optical depth i'R =½ 256 Sd Discrete representation of source function at depth-point d w
'•i Stellar radius 11 in Feautrier difference-equation solution of transfer 0
Unit vector in radial direction 3 equation 157 --..J
Residual specific intensity, relative to continuum, in S(i,j) Line strength in transition i -+ J 88 I
r,(µ)
spectrum line at angle cos- 1 µ from disk center, 1 - a,(µ) 270 S(,, ,)'}
R Stellar radius 49 S(z, v), Source function, ri.!x. 35
gf Rydberg constant 89 s,.
9l Reynolds number 560 S(cx) Normalized Stark profile 296
R Stress-energy tensor of radiation field 498 S(-11) Angular distribution of intensity emergent from grey.
Dielectronic recombination rate d -+ b 135 atmosphere 70
Rdb
Radiative transition rate from state i to state j 77 Y(2') Strength of line within multiplet 92
RIJ
Photoionization rate of level i to continuum 123 Y'(.lt) Multiplet strength 92
R,.
Radiative de-excitation ratej-+ i scaled to equilibrium s.(s) Laplace transform of (q.(t))s 466
R11
value R11 = 117 Rj,lnf 129 Time 2
Rj, ltadiative de-excitation rate j -+ i per atom in upper state 129 Current optical depth scale in Avrett-Krook procedure 174
R,, Radiative recombination rate ,,: -+ i scaled to equilibrium Equivalent electron-scattering optical depth in expanding
value R,1 = 11: R~1/11f 131 atmosphere 561
R~, Radiative recombination rate ,,: -+ i per ion in ground state 130 t, Self-relaxation time for electrons in a plasma 122
R, Residual flux, relative to continuum, in spectrum line t, Average recombination time 122
profile, 1 - A. 269 T Absolute thermodynamic temperature 7
Ro Residual flux at center of infinitely opaque line 312 T Stress-energy tensor of electromagnetic field 497
598 Glossary of Physical Symbols Glossary of Physical Symbols 599

Color temperature 248 v, Critical velocity in transsonic wind 526


T,
Radiation temperature of core in expanding atmosphere 486 Escape velocity from stellar surface 550
T, VC(l;C

T, Kinetic temperature of electrons 122 v, Expansion velocity in radial direction 474


Effective temperature 49 v, Expansion velocity along a ray with impact parameter p 474
T,rr
Kinet.ic temperature of atoms and ions 123 Vo Most probable speed 110
T•
T, Radiation temperature 360 v., Terminal velocity in stellar wind 530
Total optical thickness of finite slab 36 v;'i. Fiducial thermal velocity 449
T,, T
Boundary temperature of atmosphere 61 iJ Average speed of convective elemcn ts 188
To
Temperature perturbation in Avrett-Krook procedure 174 v Acceleration 81
T,
T, Tridiagonal matrix representing:differential operator for v(z, v, µ), v,.,, Antisymmetric angle-average of specific intensity
angle-frequency point I in Rybicki difference-equation ½[I(v, +µ) - l(v, -µ)] 152
solution of transfer equation 157
v(r) Expansion velocity of atmosphere 449
TM Maxwell stress tensor 14
vh(r) Hydrodynamic velocity in turbulent atmosphere 464
T(k 2), T(X) Characteristic function in discrete ordinate method 66
V Volume 6
T(t,0) Time development operator 298
V Perturbation potential 86
TA(t,O) Time-development operator for atom 301
V Velocity of atmosphere in fiducial thermal velocity units,
Tp(l,0) Time-development operator for perturber 300 v/v;';, 449
T,(,, µ)} Radiation temperature 121 v;
V, The ith component of total particle velocity in a gas
The ith component of thermal velocity of a particle in a gas
5)3
513
TR(v),
TR <Vi> = v, The ith component offtuid (or mear. flow) velocity in a gas 513
I
T0 (1) Current temperature distribution in Avrett-Krook v.,. Matrix element of perturbation potential ( it,!i Viit,.) 86
procedure 174
v• Four-velocity 542
w
u Velocity in units of thermal velocity (m/2kT)t v 41'7 0
Discrete representation of u(z4 , v, µ) in Feautrier difference- v, Matrix containing depth-variation of quadrature weights ())
u, and profile functions at angle-frequency point i in Rybicki
equation solution of transfer equation 155 I
difference-equation solution of transfer equation 159
u, Discrete representation of depth-variation of u(z, v,, µ,) in
Rybicki difference-equation solution of transfer equation 157 v,1(1) Classical interaction potential 301
Symmetric angle-average of specific intensity v;,(c) Canonical transformation of classical interaction potential
u(z, v, µ), u,~ 152 to interaction representation 302
½(J(v, +µ) + J(v, -µ)]
Hydrodynamic velocity in turbulent atmosphere in units of w Electron-impact line width 306
Uh(t)
local thermal velocity vh(t)/v,h(t) 464 IV Doppler width corresponding to thermal velocity
(v 0/c)(2kT/m)t 417
u, Matrix giving depth-coupling to J at angle-frequency point
Quadrature weight 144
; in Rybicki difference-equation solution of transfer Wk
equation 159 \V1, Relative probabilities of line and continuum bands in
IVz

U(t,O) Time-development operator in interaction representation 302 picket-fence model 207


U1(r, 0, qi; w Dilution factor 120
n, I, m, s) Electron orbital 91 w Matrix in final system WJ = Q in Rybicki difference-
Partition function of ionization stage j of chemical species k Ill equation solution of transfer equation 160
u1k(TJ
V Frequency displacement from line-center measured in
279
w. Electron distribution function 292
Doppler widths, (v - 110)/Avo w, Ion distribution function 292
282
V Average relative velocity of colliding particles Wi, w. Equivalent width of spectrum line 270
V Velocity 7 w• Reduced equivalent width, W/2A 0AvO 318
600 Glossary of Physical Symbols Glossary of Physical Symbols 60)

W(r) Nearest-neighbor distribution function 290 IX;Jf v) Absorption cross-section at frequency v in bound-bound
W(t) Energy density in electromagnetic field 8 transition i --> j 128
W(P) Field-strength distribution function 291 IX1.(v) Absorption cross-section at frequency v in bound-free
transition i --> K 130
W(P, o) Field-strength distribution function allowing for shielding
effects 293 IXoR(T) Dielectronic recombination coefficient 135
wm Distribution function for velocities along line-of-sight 279 IXRR(T)
IX._(v)
Radiative recombination coefficient 131
WH(P) Holtsmark field-strength distribution function 294 Free-free absorption cross-section at frequency v 165
W,;,(µ), W,(µ) Equivalent width of spectrum line at angle cos - 1 11 from p Ratio of line to continuum opacity in pkket-fence model 207
disk-center 270 p Field strength in units of normal field strength, F/F 0 291
X Cartesian coordinate in horizontal direction 3 fJ Velocity in units of speed of light 493
X Frequency displacement from line-center measured in fl, Probability of penetration of core radiation to test point in
Doppler or damping widths 338 expanding atmosphere 478
x, I /k., where k. is a characteristic root 69 {J,, Fractional departure of monochromatic opacity from
;i; Minimum of absolute value of incident and scattered mean value 74
photon frequencies measured in dimensionless units from {J, Ratio of line opacity to continuum opacity,
line-center 428 x1M!x, = x,<J,,/(K, + u} 309
x Maximum of absolute value of incident and scattered /Jo Ratio of line to continuum opacity for line with Voigt
photon frequencies measured in dimensionless units from profile, p,. = PoH(a, v), Po = Xolx, 312
line-center 428 P(r} Photon escape-probability in expanding atmosphere 478
x. x.2 = 1/k.2, where k. is a characteristic root 69 y Classical damping constant 82
x. Generalized optical depth variable iri spherical atmosphere 251 y Ratio of specific heats for an ideal gas 186
x. Contribution to opacity from (fixed} overlapping transitions 397 y Convective efficiency parameter 189 I
Xo u0 /(1 - e-uo), where u0 = (hv 0 /kT) 313 y Ratio of radiation force to its limiting value in diffusion w
X,[f(t)] K-integral operator 41 approximation 255 0
Y· Cartesian coordinate in horizontal direction 3 y Fraction of all emission that occurs coherently in atom's co
y Ratio of abundance of helium to hydrogen by number 138 rest-frame 415 I
y Velocity gradient iJY/iJ, in uniformly expanding atmosphere 481
Y,J Net collisional bracket 132
Y,"'(0, <J,) Spherical harmonic 89
y Lorentz transformation factor (I - 02/c2)-t 493
Cartesian coordinate in vertical direction (normal to Y,, y Ratio of monochromatic opacity to mean value or
z
atmospheric layers) 3 continuum value 74
z Path-length along ray in extended atmosphere 247 ')'. Generalized noncoherent scattering coefficient in non-LTE
source function 223
Zo(P, x), Zo The z-coordinate of surface of constant radial velocity
y(z, p) Coefficient of frequency-derivative in comoving-frame
corresponding to frequency shift x 479
transfer equation using optical depth scale 504
z Total geometrical thickness offinite slab 36 ji(z, p) Coefficient of frequency-derivative in comoving-frame
z Charge number of atomic nucleus 91 transfer equation 504
Z1 Ionic charge 330 r Ratio of specific heats for non-ideal gas (i.e., including
Z11 Net radiative bracket in transition j --> i 129 ionization and radiation pressure effects) 186
IX Stark shift in A per unit normal field strength, 6,)./ F0 296 r Ratio of radiation force to gravity force 256
IX1 Relative abundance of chemical species k 115 r Reciprocal lifetime of excited state 277
IX,. Energy absorption cross-section per atom 94 r Total damping width of a line 278
IX. Ste11ar angular diameter 12 r. Ratio of radiation force from electrc., scattering only to
IX(t) Atomic wave function 300 gravity 256
602 Glossary of Physical Symbols Glossary of Physical Symbols 603

re Collisional damping width 282 t'lH(r) Error in integrated Eddington flux in Unsold-Lucy
fL, ru Reciprocal mean lifetime of lower and upper states of procedure 64
transition L +-+ U 277 t:.T(m) Temperature perturbation at depth (column mass) m 173
r11 Radiative damping width 282 M Difference in 0... between Sun and a star 327
rw Collisional damping width in Weisskopf theory 283 t'lv Frequency shift arising from Doppler effect 279
f3 Resonance damping width 287 t:.x Lowering of ionization potential 112
r6 Van der Waals damping width 326 t:.)., Classical damping width in wavelength units 276
r,i(T) Secondary temperature-dependent factor for collision rates 133 t'l.l.w Wavelength shift produced by perturber located at
'5, '5L, Ou, 011, Oc Reduced damping widths, f/2 in circular frequency units, Weisskopf radius 296
f/4n in ordinary frequency units 277 t'lvo Doppler width in frequency units 279
,5 Number of perturbers in Debye sphere 294 t:.vt Fiducial Doppler width 449
,5 Ratio of continuum to total opacity averaged over line t':,rd • I Optical depth increment between mesh-points d and d +
profile 351 in di!Tcrencc-cquation solution of transfer equation 154
oN, oN, Perturbation of total number density (at depth-point d) in t'lwu Frequency shift from line-center at boundary between
linearization procedure 118 impact and statistical-broadening regimes 289
oT, oT, Perturbation of temperature (at depth-point cl) in t'lww Frequency shift produced by perturber located at
linearization procedure 118 Weisskopf radius 289
6N Perturbation of total number density distribution in t'lwo Line shift in Lindholm theory 285
linearization method 184 t'lwo Normal frequency shift 290
6T Perturbation of temperature distribution in lin\:arization t'lw(I) Instantaneous frequency shift induced by collision with
method 184 perturber 283
o,J Kronecker c5-symbol 14 E Electric permittivity 7
on., c5n,,, Perturbation of electron density (at depth-point cl) in E Fraction of line emission that is thermal (classical theory) 35
linearization procedure 117
E, e' Collisional thermalization parameter in non-LTE source
c5n,. on,.,,} Perturbation of level-populations (at depth-point d) in function 337 I
6n, linearization procedure 118 E, Generalized thermal emission parameter in non-LTE source w
_.,
,5J., /Jjd• Perturbation of mean intensity (at depth-point d) in function 223
linearization procedure 143 £ Thermalization parameter in schematic Lyman continuum 0
6J.

c5(x)
Perturbation of depth-variation of mean intensity
(at frequency v1 ) in linearization procedure
Dirac delta-function
184
8
, ,.
...
17
problem
Non-LTE source function parameter
Photoionization coupling parameter in non-LTE source
224
226
I

c5(z) Ratio of Doppler width to fiducial Doppler width, function 359


t'lv0 (z)/t:.vt 449 17o Critical phase shift in Weisskopf theory 283
op, op, Perturbation of mass density (at depth-point d) in r,' Phase change in time-interval ds 284
linearization procedure 183 r,(t) Instantaneous phase shift induced by collision with
or, •• ori,. Perturbation of emissivity (at depth d, frequency v.) in perturber 283
linearization procedure 183 r,(t, s) Change in phase in time interval (t, t + s) 284
c5x•• c5x,. Perturbation of opacity (at depth d, frequency v.) in
,,1,. ,, >)'}
linearization procedure 183
ox. Change in opacity produced by departures from LTE 219 t/(Z, v, I), Emission coefficient 25
61j,, Perturbation of solution vector in complete linearization t/,
method 231 17(00), r,(p) Total phase shift produced in collision with pcrturber at
t'lB(r) Correction to integrated Planck function at depth r in impact parameter p · 283
Unsold-Lucy procedure 63 17'(r, V, t), 17'(v) Thermal emission coefficient 26
604 Glossary of Physical Symbols Glossary of Physical Symbols 605

17'(r, v, t), 1/'(v) Scattering emission coefficient 28 Vo Threshold frequency for continuum absorption 123
Vo Line-center frequency 278
11•(r, v, r)} Thermodynamic equilibrium value of emission coefficient 26 y Minimum of incident and scattered photon frequencies 426
11•(v)
ii Maximum of incident and scattered photon frequencies 426
0 5040/T 322
Vd Auxiliary vector used in developing Feautrier difference-
0 Polar angle between direction of pencil of radiation and equation solution of transfer equation 156
normal to atmosphere layers 3
e Line-of-sight velocity 279
0 Recombination source term in non-LTE source function 359
e. f Photon frequency in atom's rest frame 412
o.,r S040/T.rr 213
~1herm Most probable line-of-sight thermal speed at temperature
o... Excitation temperature parameter deduced from curve of
322
r... deduced from curve of growth 323
growth, S040/T...
9 Unit vector in direction of change in polar angle for
e,urb Most probable line-of-sight speed of small-scale "turbulent"
mass motions in atmosphere 323
orthogonal spherical coordinate system 16
ex Photon destruction probability at frequency x from line
0 Polar angle of a point on a spherical surface 3 center 350
0 Angle between incident and scattered photons 416 Non-LTE source function parameter 225
e.. ek
IC Thermal conductivity 517 Most probable line-of-sight velocity 279
eo
Continuum absorption coefficient 205
K,
? Average photon destruction probability in a line 351
re, Absorption-mean opacity 60 n(t) Perturber wave function 300
re, Planck mean opacity S9 n
,1,. ,, !)}
Gas pressure tensor 516
niJ The ijth component of gas pressure tensor 514
1C(z, v, r), Absorption ("true") coefficient 24
nt The ijth component of partial pressure tensor for particle
IC, species k in a gas 514
IC•(r, v, r)} Thermodynamic equilibrium value of"true" absorption p Charge density 7
IC•(v) coefficient 26 p Mass density 170 I
283 w
). .Wavelength 8 p Impact parameter in collision
.....
I. De Broglie wavelength 300 P1 Density matrix element
Density matrix for atomic states
298
301
.....
).~ Ratio of true absorption to total opacity PA
,c,/(IC, + a-,) = (l - p,) 148 PP Density matrix for perturber states 301
)., Fraction of total emission that is thermal in classical line- Pw Weisskopf radius 283
formation theory [(l - p) + eP.]/(1 + P.) 309 p. Ratio of scattering coefficient to opacity,
A Thermalization depth 335 a,/(IC, + a-,) = (l - ).,) 43
A, Discrete matrix representation of A-operator at (v,. µ1) 160 Po Effective impact parameter for collision-broadening 283
A,[f(r)] Lambda (mean-intensity) operator 41 Po Invariant mass density 542
µ, Jt' cos O(cosine of polar angle of pencil of radiation) 3 Poo Equivalent mass density of invariant mass density plus
µ Magnetic permeability 7 internal energy of fluid 542
µ Number of atomic mass units per free particle in a gas 519 Pooo Equivalent mass density of invariant mass density plµs
enthalpy offluid 542
Jt, Angle-point in discrete-ordinate method 65
89 (1 Continuum scattering coefficient 309
JIH Reduced mass of hydrogen atom
2 a, Thomson scattering cross-section for free electrons 106
v, v' Frequency
22 a,ot Total scattering cross-section 84
vlJ Frequency associated with transition i -+ j
a, Imaginary part of collision integral 285
I'n Threshold frequency for ionization from nth state of
hydrogen 99 <111 Stefan-Boltzmann constant 12
606 Glossary of Physical Symbols Glossary of Physical Symbols 607

(111, Real part of collision integral 285 <I>,[f(t)] Phi (flux) operator 41
a(,,,,''}
u(z, v, t), Scattering coefficient 24
X,
XIJ
Opacity in continuum
Line opacity in transition i -+ j, x,(v) = X1J<P,
35
316
u, XIJk Excitation potential of state i, relative to ground state, of
ionization stage j of chemical species k 110
UIJ(v} Cross-section for transition I -+ j induced by collisions with
electrons of velocity 11 132 i(lon, XI Ionization potential 94
T Mean time between collisions 282 x Mean opacity 56
T Proper time 542 Xe Chandrasekhar mean opacity 74
t, Continuum optical depth 271 XF Flux-weighted mean opacity 57
t, Optical thickness of convective element 189 XR Rosseland mean opacity 58
Tr Static line optical depth 453 Xo Line opacity assuming a Voigt profile, x,(v) = ;x: 0 H(a, v),
Xo = X1Jf(1r+!J.vo) 317
t, Effective impact time 288
111. Rosseland mean optical depth 58 ,1,. ,, ')}
;:(z, v, r), Extinction coefficient, opacity, total absorption coefficient 23
T1 Optical-depth perturbation in Avrett-Krook method 174
Mean optical depth x.
t 56
<(,,,)} x,(v), x,
I/Ip
Linc opacity
Numerical factor in expression for total phase shift
35
283
t(z, v), Monochromatic optical depth 34
lj,4 Solution vector at depth-point din complete linearization
T, method 230
to(ro) Line-of-sight optical depth through uniformly expanding 1/,(ri, ... , rN) Wave function of N-electron atom 84
envelope in Sobolev theory 479
1/,(v), it,. Line emission profile 27 I
<P Azimuthal angle of pencil of radiation around normal to
atmospheric layers 3 l/, 1(r, I) Time-dependent wave function of atomic state i 85 w
<P Scalar potential 7 it,•(v) Natural-excitation line emission profile 29 .....
<P, Ratio of Paschenjump to Balmer jump, Dp/D8 236 w Solid angle 3 I\)
cj, Unit vector in direction of change in azimuthal angle for w Circular frequency 8 I
orthogonal spherical coordinate system 16 Wmn Circular frequency associated with transition m -+ n 86
</J(s) Reduced autocorrelation function 284 w. Ratio ofline emission profile to aosorption profile, it,,/<!>. 437
<f>(v), </J. Line absorption profile 27 Wo Resonant frequency of an oscillator 82
</J,(r) Time-independent wave function or"atomic state i 85 Wo Line-center frequency 276
<I> Azimuthal angle of a point on a spherical surface 3 'v, 'v E Logarithmic temperature-pressure gradient in ambient
<!>ab Matrix element describing broadening of spectrum line atmosphere and convective elements 187
a-+ b 302 'i1 A• VR Adiabatic and radiative logarithmic temperature.-pressure
<I>,. Continuum photon-absorption rate coefficient, 4nr,.,/hv gradient 186
223
<I>(s) Autocorrelation function 275 VP Gradient with respect to momentum coordinates 33
<I>(x) J~.., </J(x) dx 479 EB Earth symbol 524
<I>,Jt(T) Saha-Boltzmann factor of excitation state i of ionization 0 Sun symbol 171
stage j of chemical species k relative to ground-state ~ Integral over all solid angles 5
population of ionization stage}+ 1, "Gt= n,no,J+i,t<I>rJt 113
<1> 11t(T) Saha-Boltzmann factor of excitation state i of ionization
stage J of chemical species k relative to total number
density in ionization stage}+ 1, n;jt = n,NJtdi,1k 113
<!>,(µ) Limb-darkening function 262

Potrebbero piacerti anche