The Sun
The Sun
ASTROPHYSICS LIBRARY
Series Editors: I. Appenzeller, Heidelberg, Germany
G. Börner, Garching, Germany
A. Burkert, München, Germany
M. A. Dopita, Canberra, Australia
T. Encrenaz, Meudon, France
M. Harwit, Washington, DC, USA
R. Kippenhahn, Göttingen, Germany
J. Lequeux, Paris, France
A. Maeder, Sauverny, Switzerland
V. Trimble, College Park, MD, and Irvine, CA, USA
Springer-Verlag Berlin Heidelberg GmbH
Michael Stix
The Sun
An Introduction
Second Edition
With 225 Figures, Including 16 Color Figures,
and 27 Tables
13
Dr. Michael Stix
Kiepenheuer-Institut für Sonnenphysik
Schöneckstrasse 6
79104 Freiburg, Germany
Cover picture: Sunspot and solar photosphere, observed in the G band at 430 nm with the German Vacuum Tower
Telescope, Tenerife. An adaptive optics system of the National Solar Observatory, Tucson, was used, and the image
was further processed by a speckle technique at the Universität Göttingen. Courtesy O. von der Lühe
springeronline.com
© Springer-Verlag Berlin Heidelberg 2002
Originally published by Springer-Verlag Berlin Heidelberg New York in 2002
Softconer erprint of the hardcover 2nd edition 2002
The use of general descriptive names, registered names, trademarks, etc. in this publication does not imply, even in
the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and
therefore free for general use.
Typesetting: Frank Herweg, Leutershausen
Cover design: design & production GmbH, Heidelberg
Printed on acid-free paper SPIN 10973264 55/3141/ba - 5 4 3 2 1 0
To Jakob
Preface to the Second Edition
Since this introduction to the physics of the Sun first appeared in 1989, signif-
icant progress has been made in solar research. New insights grew especially
from the results of space missions, above all SOHO, the Solar and Heliospheric
Observatory, which in 1996 reached its orbit around the Lagrangian point L1
between the Earth and the Sun. Refined theoretical models have been ad-
vanced as well, but clearly, in my view, the lead was on the observational
side.
For the present edition I have retained the subdivision into nine chapters,
and much of the original text. However, all nine chapters have been revised
with many adjustments, and with a number of major additions. Chapter 1, for
example, now includes the 11-year variation of the solar luminosity, Chap. 2
presents an improved standard solar model and a discussion of the neutrino
experiments of the past 13 years. In Chap. 3 the sections on image recon-
struction and adaptive optics, on narrow-band filters, and on polarimetry
have been enlarged; the table of chemical element abundances in Chap. 4
has been revised. The later chapters take advantage of the advances made
in several directions, in particular helioseismology, numerical simulation, and
observation from space. Thus, we now know more about the internal rotation
of the Sun, about the hydrodynamics of the convection zone, and about the
source region of the solar wind, to name only a few prominent examples.
What has been said in the first edition about the intention of this book,
about the audience to which it is addressed, the style, equations, units, etc.,
remains valid for this edition. The preface of 1989 is therefore reprinted here.
As for the first edition, I have selected the subjects according to my personal
interest, and again I apologize for omitting topics that might be considered
very important by other colleagues working in the field. On the other hand
I think that a general introductory monograph on the Sun could never cover
the wealth of results published in the journals. Also, good books are on
the market that go into more detail in one or other direction, such as The
Sun from Space by K. R. Lang or Solar and Stellar Magnetic Activity by
C. J. Schrijver and C. Zwaan. The bibliographical notes at the end of each
chapter have been complemented, but by no means in an exhaustive manner.
Powerful programs for literature searches are available in the Internet.
VIII Preface to the Second Edition
Colleagues and students often inquired for solutions to the problems that
are posed in the text. In many cases I could not comply, because a collection of
written-up solutions did not exist. In fact, while preparing such a collection,
I realized that some of the problems had to be formulated more carefully.
On my home page – http://www.kis.uni-freiburg.de/∼stix/ – a set of
solutions can be found; comments and suggestions concerning these solutions
are most welcome.
I wish to thank Prof. Oskar von der Lühe and the colleagues and students
at the Kiepenheuer-Institut, as well as many colleagues at other institutions,
for illuminating discussions, for reading sections of the text, and especially
for help with the figures. Springer-Verlag provided the typographic style files
and useful advice, Peter Caligari and Reiner Hammer helped whenever I had
a computer problem, Ilsa David took care of the figure scanning, and Markus
Roth converted the first edition into TEX, so that I could proceed. Thanks
to all of them.
“The Sun will not break the rules”, says Heraclitus of Ephesus; if it does,
“the Erinyes, the aids of justice, will take vengeance”. And Giorgos Seferis,
in his 1963 Nobel speech at Stockholm, wondered whether a modern scientist
could take advantage of this message.
I think he can. In fact, have not most of us always regarded astronomy
in general, and solar physics in particular, as quite “normal” physics? As a
field where the same rules apply that are valid on Earth? Enigmatic spectral
lines have found their explanation without recourse to such mysterious new
elements as “nebulium” or “coronium”; likewise the present puzzles, above
all the “solar neutrino problem”, will certainly find their solutions within the
rules of general physics. Of course, we must keep in mind that we may not
yet know all those rules with sufficient precision. But in this case the message
is equally clear: the Sun will help us to comprehend the rules.
The present book aims at illustrating the application of the rules of physics
to a star like the Sun. When the publisher approached the Kiepenheuer-
Institut in search of an author, we had some debate about the best way
of achieving this aim. I felt that many good and up-to-date review articles
on many branches of solar physics were available, and that quite a number
of excellent monographs on special topics also existed. In addition, a large
number of beautifully illustrated non-specialist volumes about the Sun have
appeared in recent years. Thus, when it became clear that I was to be the
writer, I decided to try a coherent yet more technical text on the whole Sun,
which I believed had not been available for some time.
I have written this introduction for students of physics at an interme-
diate level. For example, I assume that the reader knows the basic laws of
thermodynamics and of hydrodynamics, and Maxwell’s equations. Even the
equation of radiative transfer is not rigorously deduced. Nevertheless, I have
tried to familiarize the reader with all equations on the grounds of their
physical meaning. Thus the audience may include scientists from many di-
verse branches, astronomers who perhaps have not yet specialized in solar
astronomy, and in fact everybody who is interested in the Sun and is not
afraid of formulas; yet I believe that the mathematics used is always quite
elementary.
X Preface to the First Edition
2. Internal Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.1 Construction of a Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.1.1 The Evolutionary Sequence . . . . . . . . . . . . . . . . . . . . . . . . 17
2.1.2 The Standard Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2 Age and Pre-Main-Sequence Evolution . . . . . . . . . . . . . . . . . . . . 20
2.3 Model Ingredients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.3.1 Conservation Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.3.2 Energy Transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.3.3 Element Diffusion in the Interior . . . . . . . . . . . . . . . . . . . 27
2.3.4 The Equation of State . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.3.5 The Entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.3.6 Nuclear Energy Sources . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.3.7 The Opacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
2.3.8 Boundary Conditions and Method of Solution . . . . . . . . 51
2.4 Results for a Standard Solar Model . . . . . . . . . . . . . . . . . . . . . . . 54
2.4.1 General Evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
2.4.2 Neutrinos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
2.5 Non-Standard Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
2.5.1 The Low-Z Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
XII Table of Contents
5. Oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
5.1 Observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
5.1.1 Five-Minute Oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . 181
5.1.2 The Spectrum of Solar Oscillations . . . . . . . . . . . . . . . . . 184
5.1.3 Low-Degree p Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
5.1.4 Line Width and Line Asymmetry . . . . . . . . . . . . . . . . . . . 193
5.2 Linear Adiabatic Oscillations of a Non-Rotating Sun . . . . . . . . 194
5.2.1 Basic Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
5.2.2 Spherical Harmonic Representation . . . . . . . . . . . . . . . . . 195
5.2.3 The Cowling Approximation . . . . . . . . . . . . . . . . . . . . . . . 197
5.2.4 Local Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
5.2.5 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
5.2.6 Asymptotic Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
5.3 Helioseismology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
5.3.1 Direct Modeling and Inversion . . . . . . . . . . . . . . . . . . . . . 211
5.3.2 Speed of Sound in the Solar Interior . . . . . . . . . . . . . . . . 213
5.3.3 Depth of the Convection Zone . . . . . . . . . . . . . . . . . . . . . 215
5.3.4 Chemical Constitution . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
5.3.5 Equation of State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
5.3.6 Internal Mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
5.3.7 Precise Determination of the Solar Radius . . . . . . . . . . . 220
5.3.8 Internal Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
5.3.9 Travel Time and Acoustic Imaging . . . . . . . . . . . . . . . . . 226
XIV Table of Contents
6. Convection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
6.1 Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
6.2 Mixing-Length Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
6.2.1 The Local Formalism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
6.2.2 Numerical Test Calculations . . . . . . . . . . . . . . . . . . . . . . . 243
6.2.3 Overshooting: A Non-local Formalism . . . . . . . . . . . . . . . 245
6.3 Granulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
6.3.1 The Observed Pattern . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
6.3.2 Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
6.3.3 Mean Line Profiles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
6.4 Mesogranulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
6.5 Supergranulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
6.5.1 The Velocity Field and the Network . . . . . . . . . . . . . . . . 267
6.5.2 Convective Nature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
6.5.3 The Effect of Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
6.6 Giant Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
6.6.1 Tracer Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
6.6.2 Spectroscopic Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
6.7 Bibliographical Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
7. Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
7.1 Axis of Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
7.2 Oblateness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
7.2.1 Origin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
7.2.2 Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
7.3 Rotational History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
7.3.1 The Initial State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
7.3.2 Torques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
7.3.3 Evolution of the Solar Rotation . . . . . . . . . . . . . . . . . . . . 286
7.4 The Angular Velocity of the Sun . . . . . . . . . . . . . . . . . . . . . . . . . 288
7.4.1 The Internal Angular Velocity . . . . . . . . . . . . . . . . . . . . . 288
7.4.2 The Angular Velocity at the Surface . . . . . . . . . . . . . . . . 289
7.4.3 Meridional Circulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
7.4.4 Correlation of Flow Components . . . . . . . . . . . . . . . . . . . 296
7.5 Models of a Rotating Convection Zone . . . . . . . . . . . . . . . . . . . . 297
7.5.1 Conservation of Angular Momentum . . . . . . . . . . . . . . . . 297
7.5.2 Mean-Field Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298
7.5.3 Explicit Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
7.6 Bibliographical Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
Table of Contents XV
8. Magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
8.1 Fields and Conducting Matter . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
8.1.1 The Induction Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
8.1.2 Electrical Conductivity on the Sun . . . . . . . . . . . . . . . . . 307
8.1.3 Frozen Magnetic Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
8.1.4 The Magnetic Force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310
8.2 Flux Tubes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312
8.2.1 Concentration of Magnetic Flux . . . . . . . . . . . . . . . . . . . . 312
8.2.2 Observational Evidence for Flux Tubes . . . . . . . . . . . . . . 318
8.2.3 Vertical Thin Flux Tubes . . . . . . . . . . . . . . . . . . . . . . . . . . 327
8.2.4 Curved Thin Flux Tubes . . . . . . . . . . . . . . . . . . . . . . . . . . 333
8.2.5 Thermal Structure of Photospheric Tubes . . . . . . . . . . . 340
8.3 Sunspots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
8.3.1 Evolution and Classification . . . . . . . . . . . . . . . . . . . . . . . 342
8.3.2 Sunspot Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
8.3.3 Sunspots and the “Solar Constant” . . . . . . . . . . . . . . . . . 354
8.3.4 Dots and Grains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 356
8.3.5 Oscillations in Sunspots . . . . . . . . . . . . . . . . . . . . . . . . . . . 359
8.3.6 The Evershed Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
8.4 The Solar Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364
8.4.1 Global Magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364
8.4.2 Mean-Field Electrodynamics . . . . . . . . . . . . . . . . . . . . . . . 368
8.4.3 The Kinematic αΩ Dynamo . . . . . . . . . . . . . . . . . . . . . . . 372
8.4.4 The Magnetohydrodynamic Solar Dynamo . . . . . . . . . . . 378
8.5 Bibliographical Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 382
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 453
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 485
1. Characteristics of the Sun
We all have been fortunate enough at times to see and to contemplate the
starry sky on a clear night. We have seen bright stars and faint stars. And
of course we have been told that stars can be bright either because they are
close to us or because their luminosity is really large.
Perhaps not everyone has realized that the stars differ not only in bright-
ness, but also in color. The stellar colors are clearly discernable with the
naked eye: some are bluish, others are reddish. If you have not seen this,
watch for it next time. As with an ordinary lamp, the color varies with tem-
perature. Hence, when we observe the color of a star, what we see is in fact
a measure of the temperature at the surface of that star.
When an astronomer studies a certain star for the first time, he determines
the luminosity and the surface temperature, if this is possible in some way:
these are the coordinates of the Hertzsprung–Russell diagram. Having placed
a star in this diagram he will in most cases be able to tell its radius, its method
of energy generation and energy transport, the periods and growth rates of its
pulsations, etc. Even the average rotation of stars, their capability to conserve
1.1 Distance
then, again with the help of Kepler’s law, converted to the light time for unit
distance, τA , i.e., to the travel time for the astronomical unit (to be defined
presently). The result is
τA = 499.004782 ± 0.000006 s . (1.3)
The time τA can be determined with such a small standard error that
in 1976 the International Astronomical Union adopted it as one of the pri-
mary quantities in the system of astronomical constants. The unit distance
or astronomical unit A in this system is dynamically defined in such a way
that the semi-major axis a (in astronomical units), and the sidereal period
T (in days), of an elliptical orbit of a body of mass m (in units of m ) obey
the relation 4π 2 a3 /T 2 = k 2 (1 + m), where k, the Gaussian gravitational con-
stant, by definition has the value 0.01720209895. As the velocity of light, also
by definition (since 1983), is c = 299792458 m/s, we obtain from (1.3) the
astronomical unit
A = 149597870 ± 2 km . (1.4)
The error essentially arises from the inhomogeneity of the radar reflecting
layer and from the uncertain distance of this layer to the center of gravity of
the orbiting body.
When the sidereal year and the mass of the Earth–Moon system are
substituted into the defining equation for the astronomical unit, one finds
a = 1.000000036. Thus, strictly speaking, the semi-major axis of the Earth’s
orbit, or the Earth’s mean distance from the Sun, exceeds the astronomical
unit by a few parts in 108 . However, for all purposes of solar physics the dif-
ference is of no concern. In most cases it is quite sufficient to use the rounded
value 1.496 × 1011 m for the mean solar distance.
The variation of the solar distance is between 1.471×1011 m (at perihelion,
in January) and 1.521 × 1011 m (at aphelion, in July). At these distances,
the angle of 1 – often used by astronomers as a length unit for the solar
atmospheric fine structure – corresponds to 710 and 734 km, respectively, at
the center of the Sun’s disc.
Problem 1.1. In which sense of the word “mean” is the semi-major axis of
an elliptical orbit the mean distance?
1.2 Mass
Once distances in the solar system are known, we may use Kepler’s law once
more in order to determine the Sun’s mass, m . The high precision of time
and distance measurements yields, however, an equally high precision only
for the product, Gm , of the solar mass with the constant of gravitation:
Gm = (132712438 ± 5) × 1012 m3 s−2 . Laboratory measurements of G give
4 1. Characteristics of the Sun
This is the present value of m . The Sun’s luminosity L , i.e., the total
energy radiated into space, corresponds to a mass loss ṁ = L /c2 , or
about 4 × 109 kg/s (based on the value of L given below). As we shall see in
Chap. 9, the solar wind carries another 109 kg/s away. Thus, during the Sun’s
life of about 1.5 × 1017 s (cf. Chap. 2) the total mass loss has been less than
1027 kg; this is of the same order as the uncertainty of today’s m , and is
usually neglected. However, mass loss may have been (or will be) significant in
very early or very late stages of the Sun’s evolution when stronger winds did
(or will) blow than presently. Of course, during the period of star formation,
the accretion of mass was an important, and at times dominant, process.
1.3 Radius
In order to determine the radius r of the Sun, we must use the solar dis-
tance and a measurement of the angular diameter of the visible disc. For the
photoelectric method we define this diameter as the angular distance between
the inflection points of the intensity profiles at two opposite limbs. Recent
results from ground-based and balloon-borne instruments (Sofia et al. 1994,
Neckel 1995, Laclare et al. 1996, Wittmann 1997, Brown and Christensen-
Dalsgaard 1998) cluster remarkably close around the value 959.63 of the an-
gular semidiameter at mean solar distance, which was already obtained from
(visual) heliometer measurements during the 19th century (Auwers 1891),
and which is used until today in the yearly Astronomical Almanac. Visual
drift-scan measurements yield values that are larger by ≈ 0.5 because the
human eye sees the limb at an intensity which is lower than the intensity at
the inflection point (Wittmann 1997); however, this seems not to be the case
with visual methods where two images are brought into contact, such as the
old heliometer measurements or the visual results of Laclare et al. (1996).
The statistical error of some of the new results is as small as 0.01 ; how-
ever, the differences between the diverse experiments indicate a systematic
uncertainty of at least 0.1 . Hence, we may take the mentioned canonical
value, but a conservative error estimate is appropriate. With (1.4), we thus
have
In addition, using the value of Gm given above, we find the gravitational
acceleration at the solar surface, g = Gm /r 2 , or
2
g = 274 m/s . (1.8)
1.4 Luminosity
Fig. 1.2. Absorption in the Earth’s atmosphere. The edge of the shaded area
marks the height where the radiation is reduced to 1/2 of its original strength. UV
ultraviolet; V visible; IR infrared
1.4 Luminosity 7
1374
HF
1372 0.2%
1370
Solar Irradiance (Wm−2)
ACRIM I
1368
VIRGO
SOVA2
1366
1369
1368
1367
1366
1365
1364
1363
78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99
Year
Fig. 1.3. Total solar irradiance, measured by space experiments between 1978 and
1999 (upper panel). The lower panel shows a composite constructed by adjusting
different measurements within the overlapping periods, after correction for degra-
dation and operational effects as described in Fröhlich and Lean (1998). Courtesy
C. Fröhlich
magnitude that the star would have at a distance of 10 parsec, i.e., ten times
the distance from which the astronomical unit subtends an angle of 1 . The
absolute (bolometric, i.e., integrated over the whole spectrum) magnitude of
the Sun is M = 4.74.
The name “solar constant” for S is misleading. We shall see in Chap. 2
that the theory of stellar evolution tells us that the luminosity has increased
from about 72% of today’s L to the present value over the Sun’s life of about
4.6 × 109 years. Moreover, the same instruments on the Solar Maximum Mis-
sion and other satellites that measured the absolute irradiance quoted above,
recorded relative variations of up to 0.2% over several days, and variations of
order 10−6 over time intervals of minutes. We must discuss the latter in the
context of the solar oscillations in Chap. 5, and we shall see in Chap. 8 that
the former provide a clue to the energy balance in sunspots. On an interme-
diate time scale, Fig. 1.3 clearly demonstrates a variation with the 11-year
8 1. Characteristics of the Sun
activity cycle of the Sun, as first reported by Willson and Hudson (1988).
This variation is in phase with the cycle: the maxima of the irradiance and of
the number of sunspots coincide. The amplitude of the variation is ≈ 0.09%,
peak-to-peak.
Since the solar radius is known, we may, as another measure of the lumi-
nosity, define the effective temperature Teff by
4
L = 4πr 2 σTeff , (1.12)
where σ = 5.67051 × 10−8 W/m2 K4 is the Stefan–Boltzmann constant. For
the Sun we have
Teff = 5778 ± 3 K . (1.13)
The uncertainty is predominantly due to the uncertainty in the luminosity.
The effective temperature is, in addition to the gravitational acceleration, the
second essential parameter for the structure of a stellar atmosphere.
Problem 1.2. Show that the energy flux F (λ) is π times the intensity aver-
aged over the visible solar disc.
1
¯ I(μ, λ)
I(λ) = 2I(1, λ) μdμ . (1.16)
I(1, λ)
0
¯
Hence, in order to obtain F (λ) or, which is the same, π I(λ), one possibility
¯
is to measure I(λ) directly. Diffuse light, i.e., radiation emerging from all
parts of the solar disc, must be used in such an experiment. The second
possibility is to employ an imaging system, and to measure both the central
intensity, I(1, λ), and the limb-darkening function, I(μ, λ)/I(1, λ); for the
latter, a relative measurement is sufficient.
Instead of the wavelength λ, often the frequency ν is used as the inde-
pendent argument. It is common to write Iλ or Iν , with the argument as a
subscript. For a distribution such as the intensity this subscript at the same
time indicates the appropriate interval: a transformation of the intensity in-
tegral (over wavelength or frequency) yields Iλ dλ = −Iν dν; since λ = c/ν
and dλ = −cν −2 dν, we also have λIλ = νIν .
The visible and near-infrared parts of the spectrum contain most of the so-
lar energy. For these two spectral regions both types of experiments have
been carried out, mostly from high mountain sites such as the Jungfrau-
joch Observatory in Switzerland, or from aircraft. The required correction
for atmospheric extinction is done by the classical Langley method, i.e., by
measurements at various elevations of the Sun (or various amounts of air
mass), and subsequent extrapolation to zero air mass.
Problem 1.3. How does the atmospheric extinction depend on the zenith
angle?
Figure 1.4 shows the visible spectrum. On the red side of the maximum
¯
the mean intensity I(λ) approximately follows a black-body spectrum, with
T = Teff = 5778 K. However, even at the low resolution used in this figure,
some of the stronger absorption lines, such as Hα at λ = 656.3 nm, are clearly
discernible. On the blue side of the maximum, absorption lines dominate the
spectrum.
From 1961 onwards, D. Labs and H. Neckel measured the mean intensity
by means of the limb-darkening function and the absolute intensity I(1, λ) at
disc center. The latter, which is also shown in Fig. 1.4 (upper curve), can be
used to calibrate a solar spectrum with higher resolution. The Fourier trans-
form spectrometer (Sect. 3.3.2) has a sensitivity which varies in a very smooth
fashion over a wide spectral range and therefore yields intensity spectra that
are particularly suited for such calibration. In their experiments, Labs and
Neckel selected bands of 2 nm width by means of a double monochromator
(Sect. 3.4.6) and compared the integrated intensity of each of these bands to
10 1. Characteristics of the Sun
¯
Fig. 1.4. Central intensity I(1, λ) – upper curve – and mean intensity I(λ) – lower
curve –, averaged over 2-nm bands. Data from Neckel and Labs (1984). The smooth
curve is a black-body spectrum for T = 5778 K, the effective temperature of the Sun
the corresponding intensity of a standard lamp. The lamp itself was calibrated
in the laboratory by means of the radiation of a black body of given tem-
perature. The measurements show a scatter of about 0.2%, and systematic
errors are estimated to be less than 0.5%.
Next to the visible spectrum the infrared, which is shown in Fig. 1.5, makes
the second largest contribution to the solar energy output: about 44% of the
electromagnetic radiation is emitted at wavelengths longer than 0.8 μm. The
spectrum is approximately thermal, although high-resolution spectra show a
large number of lines. The infrared spectrum is rather well represented by
the Rayleigh–Jeans (i.e., long-wavelength) approximation of Planck’s law,
2hc2 2ckT
Bλ = , i.e., Bλ , (1.17)
λ5 (ehc/λkT − 1) λ4
so that
In particular, in the far infrared above 10 μm, the double logarithmic spec-
trum of Fig. 1.5 deviates very little from a straight line. The brightness
temperature, TB , defined by Iλ = Bλ (TB ) [where Iλ is the observed abso-
lute intensity at wavelength λ, and Bλ (T ) is the Kirchhoff–Planck function],
1.5 Spectral Energy Distribution 11
Problem 1.4. Calculate the spectral irradiance for the wavelength and
brightness temperature values given in the text. Estimate the error arising
from the Rayleigh–Jeans approximation of Planck’s law.
Problem 1.5. Use the Rayleigh–Jeans approximation for an estimate of
the entire solar irradiance above any given wavelength. Show that about
0.06% of the Sun’s energy is emitted in the far infrared, at λ > 10 μm.
Figure 1.6 illustrates the solar radio spectrum, which begins in the microwave
region at λ = 1 mm. As is common in radio astronomy, the energy flux is given
per frequency interval rather than per wavelength interval. The thermal quiet-
Sun spectrum, which smoothly continues from the infrared into the radio
region, therefore has a slope of −2 in the double logarithmic representation.
This slope is, however, not constant everywhere: between λ = 1 cm and
λ = 1 m there is a transition of the brightness temperature from about
104 K to about 106 K!
Superimposed on the quiet solar radio emission is a spectrum of great
variability. The s-component (for slowly varying) is correlated to the solar
11-year activity cycle. Indeed, the radio flux at 10.7 cm is used as an index of
the solar activity. The spectrum of the s-component is approximately thermal,
and its flux normally is 1 or 2 orders of magnitude below the quiet-Sun flux.
On the other hand, there are rapid bursts of radio emission, on time scales of
12 1. Characteristics of the Sun
seconds to days. During such events the flux may exceed the quiet-Sun level
by several orders of magnitude, with substantial deviations from the thermal
spectrum. The frequency of occurrence of radio bursts is again correlated to
the 11-year cycle.
The solar radio emission was first detected by G. C. Southworth and by
J. S. Hey around 1942. Its absolute flux is now generally measured with an
uncertainty of about 10%. Since the Earth’s atmosphere has a broad radio
window (Fig. 1.2), ground-based antennas can be used.
The solar ultraviolet spectral irradiance is shown in Fig. 1.7. As in the blue
and violet parts of the visible spectrum, absorption lines are the dominant
feature down to about 210 nm, although only a few of the strongest lines, such
as the Mg I line at λ = 285.2 nm or the Mg II h and k lines at λ = 280.3 nm
and λ = 279.6 nm, can be identified at the low spectral resolution of Fig. 1.7.
The sharp decrease near λ = 210 nm and the continuum which follows to-
wards shorter wavelengths are due to the ionization of Al I. The whole band
between 200 nm and 150 nm is approximately represented by a brightness
temperature of 4700 K.
Below 150 nm, emission lines dominate the spectrum. Most prominent
is the Lyman α line of hydrogen, a line of about 0.1 nm width centered at
1.5 Spectral Energy Distribution 13
Fig. 1.7. Solar spectral irradiance in the ultraviolet, averaged over 1-nm bands. The
solid and dashed curves are black-body spectra. Data from Heath and Thekaekara
(1977)
121.57 nm; the average irradiance in this line alone is 6 mW/m2 , and this is
as much as in the whole spectrum below 150 nm besides Lyman α!
Towards shorter wavelengths, the ultraviolet irradiance is increasingly
variable. For example, on a time scale comparable to the Sun’s synodic period
of rotation (27 days), changes of up to 25% are observed at 120 nm. These
changes are partly true temporal variations and partly a manifestation of
the non-uniform distribution of the sources in the solar atmosphere; these
sources pass across the visible hemisphere as the Sun rotates. Still larger in
amplitude – up to a factor 2 – is a variation which correlates with the 11-year
activity cycle.
Absorption in the Earth’s atmosphere, mainly by O2 , makes it necessary
to use rockets or satellites for observations in the ultraviolet. A major addi-
tional problem for absolute measurements are the standards. Recent results
(Thuillier et al. 1997) still have an uncertainty of 2.2% at 350 nm, and of
4% at 200 nm. For the total luminosity of the Sun the errors – and also the
mentioned variations – are of minor importance: the bands from 300 nm to
330 nm, 210 nm to 300 nm, and 150 nm to 210 nm contribute only 1.5%, 1%,
and 0.01%, respectively.
range, from about 8000 K to 4 × 106 K, can thus be studied. We shall see in
Chap. 9 that EUV emission lines are a major source of information for the
transition zone from the chromosphere to the corona.
Fig. 1.8. Solar spectral irradiance between 1 nm and 30 nm. Shaded intervals rep-
resent the range of results from various rocket experiments between 1965 and 1972.
Adapted from Manson (1977)
The corona is also a source of soft x-rays (0.1 nm to 10 nm). Figure 1.8
illustrates the variability of the solar spectral irradiance at these short wave-
lengths. In addition to the rocket experiments there has been continuous
monitoring from space: between solar maximum and minimum the SOLRAD
satellites found a variation by a factor of 200 in the 8–20 Å band, and by a
factor of 20 in the 44–60 Å band.
The largest variations, often several orders of magnitude, occur during
solar flares. Hard x-rays, below λ = 1 Å, and even γ-rays are emitted in these
events, which we shall briefly discuss in Chap. 9.
For comparison with other stars it is often convenient to express the spectral
energy distribution in the rough form of color indices. These are defined as
ratios of broad-band integrals over wavelength, viz.
∞ ∞
U−B = −2.5 log S(λ)EU (λ)dλ − log S(λ)EB (λ)dλ + CUB , (1.19)
0 0
∞ ∞
B−V = −2.5 log S(λ)EB (λ)dλ − log S(λ)EV (λ)dλ + CBV . (1.20)
0 0
1.6 Bibliographical Notes 15
α = l/HP (2.1)
The standard model of the Sun can be defined as the model which is based
on the most plausible assumptions and on the best available physical input.
Above all, it is spherically symmetric, i.e., all physical quantities depend only
on the distance, r, from the center. This means that the internal rotation
is sufficiently slow (comparable to that observed at the surface), and that
the internal magnetic field is sufficiently weak (in the solar interior a field of
sunspot strength would be weak in this sense), that the forces resulting from
these two agents have negligible influence.
In the standard model the abundances of the elements other than hy-
drogen and helium are usually combined into a number Z, which is their
fractional abundance by weight, so that
X +Y +Z =1 , (2.4)
where X and Y , the fractional abundances of hydrogen and helium, vary
with depth in the star and with time, because hydrogen is fused into helium,
and because – to a small amount – the heavier helium diffuses towards the
solar center, cf. Sect. 2.3.3. Such gravitational settling also leads to a slight
variation of Z. For the calculations of the present text only helium settling
is taken into account, while Z is kept constant. I have chosen Z = 0.02,
which closely represents the heavy element abundances observed on the solar
surface (cf. Chap. 4), and which is typical for a population I star. But (2.4)
is slightly modified to X + Y + Y3 + Z = 1, where Y3 , the mass fraction of
3
He, also evolves as a function of depth and time, starting from the constant
Y30 = 4 × 10−5 . The correct value of Y30 is not precisely known, but Rood
et al. (1984) and Balser et al. (1999) found values similar to the one used
here when they studied the 3 He+ hyperfine line at 8.7 GHz in galactic H II
regions.
20 2. Internal Structure
The detailed abundances of the heavy elements are not considered explic-
itly, with two exceptions: The opacity tables (Sect. 2.3.7) – these are based
on abundances determined spectroscopically at the solar surface or inferred
from elsewhere in the solar system (Table 4.2) –, and the CNO energy gener-
ation rate which depends on the initial abundances of 12 C and 14 N (cf. Sect.
2.3.6) in the solar core.
The relatively large content of heavy elements comes from the fact that the
Sun was formed as a second-generation star out of matter already processed
in stars of earlier generations. Some of the latter, the population II stars, are
still observed today, in particular in the very old globular clusters. The heavy
element content of these stars is generally smaller than the Sun’s by an order
of magnitude, or even more.
In the standard model, convection and mixing will occur only if the
Schwarzschild criterion for the stability of stratification, to be discussed in
Chap. 6, is violated. If this is the case, the model will employ the mixing
length formalism already mentioned. Convection and radiation in unstable
layers, and radiation alone in stably stratified layers, will be the only mecha-
nisms of energy transport (a small amount of heat conduction will be included
into the radiative part).
The standard model is not rigorously defined. A number of modifications
have been presented in the literature, e.g. slightly different values of Z, or
a more detailed treatment of the heavy elements and their ionization than
is given here, or various versions of the equation of state (see below). More
drastic modifications are considered as non-standard; some of them will be
treated at the end of this chapter.
Meteorites are the oldest bodies in the solar system studied so far in the
laboratory. Their age is determined from the decay of radioactive isotopes,
such as 87 Rb, which has a half-life of 4.8 × 1010 years. The abundances of
both the parent and the daughter (87 Sr) isotopes are measured relative to the
(constant) abundance of 86 Sr, and samples from many different meteorites,
in particular chondrites, yield essentially the same age: (4.55 ± 0.05) × 109
years. The decay of uranium and thorium into lead isotopes confirms this
result. According to recent determinations the error might be even 10 times
smaller (Wasserburg 1995). Hence the age of meteorites, or, more precisely,
the time elapsed since the condensation and chemical differentiation of these
bodies began, is rather well known.
Problem 2.1. Show how two samples with different Rb/Sr abundance ratios
can be used to determine their (common) age.
2.2 Age and Pre-Main-Sequence Evolution 21
The problem is, then, to relate this epoch to the start of the Sun’s main-
sequence life, i.e., to the ignition of hydrogen burning in the center of the
homogeneous initial star. To this end one should know the sequence and
duration of various processes during the formation of the Sun and the solar
system.
First, an interstellar cloud of, say, 104 solar masses is triggered into its
gravitational collapse. That is, some external agency must have generated
an initial compression of the cloud so that self-gravitation could overcome
the internal gas pressure. The external agency could have been the shock
wave of the galactic gas in a spiral arm: the fact that most young stars are
found in or near spiral arms supports this hypothesis. It could also have been
the shock wave generated by a nearby supernova explosion. This alternative
would explain the large abundance of the 26 Mg isotope found in connection
with 27 Al in the Allende meteorite of 1969: 26 Mg is a decay product of 26 Al,
which has a half-life of only 7 × 105 years and should have disappeared if
it was synthesized long before the formation of the solar system. If, on the
other hand, the 26 Al came from a nearby supernova explosion then it would
naturally yield the anomaly of 26 Mg.
The collapse begins when the initial compression is sufficiently strong for
gravitation to win against the internal pressure. This is the case if the Jeans
criterion is satisfied, i.e., if
Gmc RT
> , (2.5)
r μ
where R is the gas constant, and μ, T , mc , r are the cloud’s mean molecular
weight, temperature, mass, and radius.
Problem 2.2. Transform (2.5) into a condition for the cloud mass, at given
interstellar density and temperature. Show that about 103 solar masses
is the minimum required for instability, at typical interstellar conditions:
T = 50 K and ρ = 10−20 kg/m3 . Show that, as the collapse goes on, the
conditions become more favorable for further collapse.
Problem 2.3. Calculate the free-fall time tff , i.e., the time which a spheri-
cally symmetric cloud of initial density ρ0 and negligible internal pressure
needs for complete collapse.
the Sun was formed would had been conserved during a spherically symmet-
ric collapse, the field would have been amplified proportionally to ρ2/3 , to a
value of 1010 G or more. That is, virtually all the magnetic flux must have
been lost as well.
Problem 2.4. Use the solar model of Table 2.4 to calculate the Sun’s moment
of inertia, and the solar spin angular momentum. Compare the result with
the orbital angular momenta of the planets.
Now it happens that the magnetic field provides a very effective lever
arm for the torque required to remove angular momentum – as we shall see
in Chap. 9 in the context of the solar wind. A plausible hypothesis, therefore,
is that during a first phase of the collapse magnetic braking takes place. Later,
when the density has become sufficiently large (and the temperature is still
sufficiently low) so that neutral particles can form by recombination of the
originally ionized material, the cloud is no longer coupled to the magnetic
lines of force. The field may then escape back into the interstellar space.
A further complication is that, of course, the cloud must fragment in order
to form the 103 to 104 stars of a typical galactic cluster. Spontaneous frag-
mentation may occur because, as the density increases, the Jeans criterion is
satisfied for smaller masses – provided there is no significant increase of the
temperature, cf. Problem 2.2. The presence of molecules prevents such tem-
perature rise, because molecules effectively radiate the energy gained during
the collapse back into space. Hence subsystems of the original cloud can begin
their own collapse. The Sun, with its planets, must have formed out of such
a subsystem.
The details of the successive removal of angular momentum and magnetic
flux, and of the process of fragmentation, are not known. Probably there was
an intermediate state on a stellar accretion disc, with two jets emanating
perpendicular to the disc in the two directions along the axis. Observations
in the infrared have revealed such discs around several young stars, e.g.,
Wega, β Pictoris, or HL Tauri, and jets also have been observed. As far as
theoretical models and numerical simulations exist, they confirm that the
duration of this whole early history is still essentially the free-fall time, tff .
For a single fragment such as the pre-solar nebula the density is typically
10−17 to 10−15 kg/m3 , and the free-fall time is correspondingly shorter. De-
tailed numerical simulations exist for this phase (e.g., Bodenheimer 1983).
While the bulk of the fragment collapses in free fall, the central part be-
comes optically thick, heats up, and thus is able to set itself into hydrostatic
equilibrium: a protostar is formed.
It takes less than 106 years for the accretion of virtually the whole envelope
on to the hydrostatic core. At the end of this phase, a low-density, slowly
contracting cool star (that is: a gaseous sphere in hydrostatic equilibrium) is
born. Its effective temperature is ≈ 3000 K, its radius is about four times the
2.2 Age and Pre-Main-Sequence Evolution 23
present solar radius, and its luminosity is several times the present luminosity.
In the Hertzsprung–Russell diagram it evolves with decreasing luminosity
along the Hayashi line, the location of fully convective stars. The central
temperature is still below 106 K, much too low for the major nuclear reactions
to take place at a significant rate. Instead, the source of the luminosity is the
release of gravitational energy during the slow contraction.
According to the virial theorem (cf. Problem 2.10) the slowly contracting
star gains an amount of internal energy which is equivalent to the energy
radiated into space. Hence it becomes hotter, until finally hydrogen burning
sets in (Sect. 2.3.6). The Sun has then arrived at the zero-age main sequence,
i.e., at the main sequence of chemically homogeneous stars.
The duration of the slow contraction phase is the ratio of supply to con-
sumption. This time scale is also called the Kelvin–Helmholtz time,
Gm2
tKH = , (2.6)
rL
and is about 3 × 107 years for the Sun. An example of an evolutionary track
in the Hertzsprung–Russell diagram is shown in Fig. 2.1. It begins with the
formation of the core (prior to this event the object is optically thin and an
effective temperature cannot be defined) and covers the protostar collapse
and the final hydrostatic approach to the main sequence.
As far as the Sun’s age is concerned, the conclusion is as follows: The
initial cloud collapse, and the removal of angular momentum and magnetic
flux, took about 3 × 107 years. After a rapid collapse a time span of similar
length was spent in the final hydrostatic contraction. Sometime during this
whole evolution (or shortly after its end) the material from which meteorites
are formed condensed. Thus, the uncertainty in the dating of the origin of
meteorites with respect to the Sun’s pre-main-sequence evolution is larger
than the uncertainty in the age of the meteorites itself. For the solar age
we may therefore take the result of Wasserburg (1995), but with a more
conservative error estimate:
This is the time span which the Sun has spent on the main sequence, convert-
ing hydrogen to helium. We shall see that the Sun’s radius and luminosity
both slowly increased during this main-sequence evolution, until finally the
present r and L were reached.
We now discuss the laws and equations which govern the solar structure and
evolution.
Problem 2.5. Use (2.9) to find a lower bound for the pressure at the center
of the Sun. Compare the result with Table 2.4.
Problem 2.6. Use (2.10) to show that a violation of (2.9) leads to a signifi-
cant change of the solar radius in the dynamical time scale
tD (Gρ̄)−1/2 ≈ 1 hour , (2.11)
where ρ̄ is the mean density.
Problem 2.7. Calculate the Sun’s gravitational energy, i.e., the total work
required to disperse the solar matter over distances r r . Use Table
2.4 for the function r(m).
We now turn to the energy balance. Let L(m) be the luminosity generated
inside the sphere of mass m. We then define the energy generation ε per unit
mass through
∂L ∂S
=ε−T , (2.12)
∂m ∂t
where T is the temperature and S is the specific entropy (entropy per unit
mass). The second term on the right ensures that, in addition to explicit
sources, heating or cooling can affect the luminosity. As far as the main-
sequence evolution is concerned, the Sun is very close to thermal equilib-
rium. The change in specific entropy is then very small in comparison to ε.
The function ε(ρ, T ) itself represents the nuclear energy sources and will be
considered in Sect. 2.3.6.
At any depth in the Sun, the energy flux, F , is defined as the luminosity
per unit area. Since we consider energy transport by radiation (FR ) and by
convection (FC ), we may write
F = FR + FC = L/4πr2 . (2.13)
Let us first consider radiative energy transport. In the interior of stars
the mean free path of a photon is very small, so small in fact that the places
26 2. Internal Structure
where the photon is emitted and absorbed have nearly the same temperature.
That is, the conditions of local thermodynamic equilibrium (cf. Chap. 4) are
very closely satisfied. Consequently, in the equation of radiative transfer we
may replace the source function by the Kirchhoff–Planck function
2h ν3
Bν (T ) = , (2.14)
c2 ehν/kT − 1
and obtain
dIν
cos θ = −κν ρ(Iν − Bν ) . (2.15)
dr
Here θ is the angle to the local vertical direction, and Iν and κν are the
intensity and the absorption coefficient (cross section per unit mass) at fre-
quency ν. Of course, we must not use Iν = Bν , because a completely isotropic
radiation field would permit no net transport. But the expansion
cos θ dBν
Iν = Bν − (2.16)
κν ρ dr
satisfies (2.15) up to a term proportional to cos2 θ, which does not contribute
to the flux. The energy flux is
∞
F = Fν dν , (2.17)
0
The opacity is a harmonic mean, i.e., the average of 1/κν . That is, more
energy is transported at frequencies where the matter is more transparent.
It is also a mean weighted with dBν /dT ; this means that more energy is
2.3 Model Ingredients 27
In the solar interior there is a large central region, of radius ≈ 0.7r , where
energy is transported merely by radiation. Hence energy transport causes
no mixing of matter in this region. The chemical composition would remain
the same except for nuclear transmutations. However, driven mainly by the
gravitational force, heavier particles tend to diffuse towards the solar center,
while lighter ones tend to diffuse in the opposite direction, away from the
center. This gravitational settling is a very slow process: its characteristic time
exceeds the solar age by a factor 104 , hence it makes only a small difference
28 2. Internal Structure
for the interior chemical composition of the present Sun. Nevertheless the
process has been taken into account in many recent model calculations. This is
appropriate in view of the high precision required for some helioseismological
applications, or for the prediction of the solar neutrino flux.
Diffusion in the Sun has been treated already by Aller and Chapman
(1960); Noerdlinger (1977) first showed that it lowers significantly the helium
abundance at the solar surface, and raises the central helium abundance. In
the present text I shall follow the treatment of Bahcall and Loeb (1990) and
of Thoul et al. (1994).
In plasma physics a problem such as the present one is often treated in
terms of a multi-component fluid, where each component is subject to a set of
conservation equations that include the coupling to the other components. As
an illustrative example we consider a three-fluid model consisting of hydrogen,
helium, and electrons. Since the model is applied to the solar interior, where
T > 2 × 106 K, the assumption of complete ionization is in order.
The conservation of the hydrogen and helium particles, with densities nH
and nHe , is expressed – in spherical symmetry – by
∂nH 1 ∂ 2 ∂nHe 1 ∂ 2
=− 2 r nH vH , =− 2 r nHe vHe , (2.22)
∂t r ∂r ∂t r ∂r
where vH and vHe are the diffusion velocities and where, on the right-hand
sides, the particle density changes due to nuclear reactions must be added in
a calculation of the Sun’s evolution.
The coupling between the three fluid components appears in the momen-
tum conservation equations. The hydrogen and helium fluids are coupled
together by a friction force fH,He mediated by Coulomb collisions, that is
particle encounters with interactions due to their electric charge. Compared
to the momentum transfer during such ion-ion interactions, the momentum
transfer in ion-electron collisions is smaller by a factor (me /mi )1/2 , hence
the latter type of friction is neglected. On the other hand there is coupling
between ions and electrons, mediated by the electric field. With PH , etc.,
denoting the partial pressures, the three momentum conservation equations
are
∂PH
− + nH (eE − mH g) − fH,He = 0 , (2.23)
∂r
∂PHe
− + nHe (2eE − mHe g) + fH,He = 0 , (2.24)
∂r
∂Pe
− − ne eE = 0 . (2.25)
∂r
Since diffusion is a very slow process, we have omitted the acceleration terms.
In addition, because of the small electron mass, the gravitational force has
been omitted in the equation for the electrons. We recognize that the colli-
sional friction between H and He has effects that are equal in magnitude, but
opposite in sign, on these two fluid components.
2.3 Model Ingredients 29
We eliminate the electric field by (2.25), and express the diverse partial pres-
sures in terms of the total pressure P and the hydrogen mass fraction X
(which is easy in the case of the present three-fluid model with full ioniza-
tion), and so obtain the diffusion velocities as linear functions of ∂P/∂r and
∂X/∂r.
Thermal Diffusion. In addition to the gradients of pressure and species
concentration, a temperature gradient constitutes a driving force for diffusion.
This is because diffusion is governed by collisions, and particles coming from
the direction of larger T have a slightly larger thermal velocity than particles
coming from the opposite direction, of smaller T . It follows that the lighter
species, here hydrogen, is driven in the direction of decreasing temperature,
relative to the heavier species, helium. But there are not only consequences
for the momentum transfer during the collisions, but also for the transfer of
energy. Hence for each particle species an equation of energy conservation
must be considered, in addition to Eqs. (2.22) to (2.25) above. For a multi-
component fluid Thoul et al. (1994) have solved this whole set of conservation
equations numerically. For the solar application they also give approximate
expressions of the resulting diffusion velocities of hydrogen, helium, oxygen,
and iron; for hydrogen the result is
5/2
r 2 ρ0 T ∂ ln P ∂ ln T ∂ ln CH
vH = AP + AT + AH , (2.28)
τ0 ρ T0 ∂r ∂r ∂r
30 2. Internal Structure
For the solution of our differential equations we must express the density ρ
in terms of P an T . It is more common, however, to consider the pressure P
as a function of T and ρ, which is known as “the equation of state”. We shall
do the same here, too, but we must recall at the end that the relationship
will be solved for ρ.
In a rigorous treatment the equation of state is obtained from a minimiza-
tion of the free energy F via
∂F
P =− . (2.30)
∂V T,n
Problem 2.8. Confirm (2.34) to (2.36). Consider explicitly any one of the
heavy elements, and modify (2.34) and (2.36) accordingly.
where gij is the statistical weight of the jth state, and Eij is the energy of
that state, relative to the ground state. For an isolated atom or ion the sum
(2.38) has an infinite number of terms and diverges, but most of the terms
have energies close to the ionization energy and are therefore cut off in a
dense plasma where the particles perturb each other and so, in effect, lower
the ionization potential. In fact, a common approximation is to replace the
partition functions by the weights of the respective ground states, i.e., 2, 1,
and 2 for H, He, and He+ .
It must also be mentioned that, as the electron pressure increases to-
wards the stellar interior, ionization would finally decrease inwards according
to (2.37). But the mutual interaction of the particles leads to additional ion-
ization (“pressure ionization”) which, in a crude approximation, is simulated
by simply assuming complete ionization at all depths below the level where
complete ionization is reached (or where ionization begins to decrease) ac-
cording to (2.37).
the Sun this ratio is of order 0.1 or smaller (Fig. 2.2). That is, the electrostatic
interaction is not negligible, but small enough for the Debye–Hückel treatment
to be valid. [The Debye–Hückel treatment is outlined in physics texts such
as Landau and Lifschitz (1966 a, p. 246). The presentation here, essentially
following Clayton (1968), will be short.]
In the neighborhood of an ion, the density nZ of any other species with
charge eZ (we set Z = −1 for electrons) deviates from the mean density n̄Z
in accord with a Boltzmann distribution:
eZV
nZ = n̄Z exp − . (2.39)
kT
The potential V in turn is determined via the Poisson equation from the
combined densities of all charged particles. The result is
eZ
V = exp(−r/rD ) , (2.40)
4πε0 r
where
1/2
ε kT
rD = 2
0 2 (2.41)
e Z n̄Z
is the Debye radius, and the sum is over all charged species. The density
dependence of rD follows from n̄Z ∝ ρ.
We may now calculate the total additional energy density UES , due to the
potential correction VES of V [expand the potential (2.40)!], by
1 e3 ( Z 2 n̄Z )3/2
UES = eZ n̄Z VES = − . (2.42)
2 8πε0 (ε0 kT )1/2
The additional energy density is negative. In correspondence, the pressure
correction
34 2. Internal Structure
1
PES = UES (2.43)
3
is also negative: due to the electrostatic interaction of the charged particles
the plasma becomes “softer”, i.e., at a given temperature and density the
pressure is smaller than it would be for a perfect gas.
Problem 2.11. Expand the exponential in (2.39), and solve the Poisson
equation to confirm (2.40) and (2.41). Use the condition of mean charge
neutrality, the abundances X and Y , and the ionization degrees of hy-
drogen and helium to derive the explicit dependence of rD and UES on
density. Use the integrability condition for the entropy to confirm (2.43).
Show that the condition of small electrostatic energy (compared to kT )
is equivalent to a large Debye radius (compared to r̄).
ρNA (1 + X)/2, where NA is the Avogadro number, and this electron density
must be equal to the integral over (2.44). Hence
∞
ρNA (1 + X)/2 = ne (p) dp (2.45)
0
We shall now consider the second “equation of state”, which determines the
entropy. Only changes of the entropy are of interest, either in time [cf. (2.12)],
or in depth – as we shall see when we treat the theory of stellar convection in
Chap. 6. Hence the differential form common in thermodynamics is adequate,
but in place of the extensive variables U and V (internal energy and volume)
the intensive variables P and T are used. We write
dT dP
dS = cP − ∇a , where (2.47)
T P
∂S
cP = T and (2.48)
∂T P
∂ ln T
∇a = (2.49)
∂ ln P S
are the specific heat at constant pressure and the adiabatic temperature gradi-
ent (more precisely: the double-logarithmic isentropic temperature gradient).
In order to obtain the entropy we must determine cP and ∇a as functions of
P and T . We have already, in the preceding section, considered the function
ρ(P, T ) and thus can derive
∂ ln ρ
δ=− . (2.50)
∂ ln T P
Let us again neglect the radiation pressure and consider only the perfect gas
term in the equation of state; then
∂ ln μ
δ =1− , (2.51)
∂ ln T P
where μ(P, T ) is given by the ionization equilibrium (2.37). Using (2.50) we
may express the adiabatic gradient in terms of the specific heat (cf. Problem
2.12):
2.3 Model Ingredients 37
Pδ
∇a = . (2.52)
T ρcP
In order to find cP itself, we use the definition (2.48) and the basic relation
T dS = dU + P dV , where U is the specific internal energy and V = 1/ρ is the
specific volume (both per unit mass, in contrast to (2.42), where UES was an
energy per unit volume). Then
∂U Pδ
cP = + , (2.53)
∂T P ρT
where again (2.50) has been used.
It remains to determine the specific internal energy U as a function of P
and T . Its main contributions are the kinetic energy of free particles and the
energy of ionization. If only hydrogen and helium are considered, then
3RT 1
U= + [nH+ χH + nHe+ χHe + nHe++ (χHe + χHe+ )] . (2.54)
2μ ρ
The ionization energies, χH . . . , and the densities, nH+ . . . , of the various
ions have been defined in the preceding section. The temperature dependence
of the ion densities is defined by means of the three Saha equations (2.37).
Figure 2.4 shows cP and ∇a for a calculated solar model. Deviations from
the perfect gas expressions cP = 5R/2μ and ∇a = 2/5 mainly occur in the
layer where H and He are partially ionized, because there the degrees of ion-
ization, which enter (2.54), depend on temperature. The main excursion is
caused by the ionization of H and the first ionization of He, while the small
hump near P = 109 Pa is due to the second ionization of He.
Problem 2.12. Use the integrability condition for the enthalpy in order to
prove (2.52).
Problem 2.13. Calculate the corrections to cP and ∇a which arise from the
electrostatic interaction in the Debye–Hückel treatment.
In this section we shall treat the energy generation ε defined in (2.12). Since
we already know the density ρ as a function of P and T , we may determine
ε(ρ, T ); as will be clear presently, this is more natural than ε(P, T ) would be.
When it became clear that the Earth (and by implication the Sun) was
much older than the Kelvin–Helmholtz time, energy sources other than grav-
itation or cooling had to be studied. Atkinson and Houtermans (1929) were
among the first to consider nuclear energy, in particular the energy set free
by the conversion of hydrogen to helium. Of all nuclear energy sources this
“hydrogen burning” is the only lasting possibility for a star like the Sun.
First of all, hydrogen is by far the most abundant element. Second, the mass
defect of the helium nucleus – compared to the four constituent nucleons –
is much larger than the mass defect of all other nuclei compared to their
building blocks; therefore the energy release, 6.683 MeV per nucleon, is the
largest available among all nuclear reactions. Third, the hydrogen nucleus has
the smallest possible electric charge; the electrostatic barrier which must be
overcome (or tunneled through) before a nuclear reaction can occur is there-
fore not as prohibitive as for heavier nuclei. Even at the high temperature of
1.5 × 107 K in the Sun’s center this is a very important point: the thermal
particle energy kT is only ≈ 1.3 keV, while the height of the electrostatic
barrier generally is of order 1 MeV!
Hydrogen-Burning Reactions. The most important reactions in the Sun
are those of the proton-proton chains, ppI, ppII, and ppIII. Let us use the
notation X(a, b)Y , common in nuclear physics, where X is a target, a an
incident particle, b an emitted particle (two particles may also appear in place
of a or b), and Y is the residual nucleus. Thus, the reactants are before the
comma, the resulting particles are after the comma. The hydrogen isotopes
are designated by 1 H or p and 2 H or d, helium isotopes are 4 He or α and 3 He,
and the other particles have their usual abbreviations. A superscript * means
a nucleus in an excited state. The energy release per reaction is Q = Q + Qν ,
where Q is the part delivered to the thermal bath in which the reactions
occur, and Qν is the energy carried away by a neutrino. The pp chains are
then described by the scheme shown in Table 2.1.
The ppII chain occurs as an alternative to the third reaction of ppI, and
ppIII occurs as an alternative to the second reaction of ppII. The branching
ratios depend on the reaction rates, and therefore on temperature. According
to model calculations for the present Sun the global ratios are 85.2 : 14.8 for
2.3 Model Ingredients 39
the I : (II+III) branching, and 14.8 : 0.019 for the II : III branching (Problem
2.19).
The total energy release per produced α particle is 26.732 MeV in each
case – note that the first two reactions must be counted twice for each α
particle synthesized in ppI. The neutrino energies Qν (and hence also the
corresponding Q values) are averages. The averages are over continua for
the p+p reaction and for the β-decay of 8 B, and over two lines at 384 keV
and 862 keV for the electron capture of 7 Be. The neutrino spectrum will be
considered again in Sect. 2.4.2 below.
Besides the pp chains, the CNO cycle is of some importance for the solar
energy generation. In fact, when the reactions of the CNO cycle were first
investigated (von Weizsäcker 1938; Bethe 1939), available solar models had
central temperatures as high as 1.9×107 K, which made the energy generation
of these reactions dominant. In contrast, present calculations show that this
is true only for more massive main-sequence stars, while for the Sun the
is now the reaction rate per unit mass; the Xi = X, Y3 , . . . are the mass
fractions, and the Ai are the atomic weights. The time derivatives in (2.72)
to (2.74) are for fixed mass shells, i.e. Lagrangian derivatives (cf. Problem
2.15). Notice that for the CNO cycle in equilibrium n14 N n12 C , so that
X14 XC + XN , where XC and XN are the initial mass fractions of carbon
and nitrogen.
Using the rates (2.75), we may now write down the energy generation per
unit mass:
ε= Qik rik , (2.76)
where the sum is over all the reactions. Due to the assumed equilibria of
deuterium, 7 Be and 7 Li, etc., several of the rik coincide, and their energies
Qik may therefore be combined into a single term.
Figure 2.5 shows ε as a function of r in the Sun. Almost all of the energy
is generated within 25% of the radius (i.e., 1.5% of the volume). Even more
concentrated towards the center is the small contribution of ≈ 1.2% made by
the CNO cycle.
Reaction Rates. The rate at which a nuclear reaction takes place depends
on the relative velocity v = |v i − v k | of the reactants, on their distribution
in velocity space, and on the cross section σ(v), i.e., the probability that the
reaction occurs at a given particle flux density. If dni (v i ) and dnk (v k ) are
the distributions of the two reactants with velocities v i and v k , then the rate
per unit mass is
1
rik = vσ(v) dni dnk , (2.77)
ρ(1 + δik )
where the integrals are over the two velocity spaces. In the stellar interior
the conditions of local thermodynamic equilibrium are satisfied to a very
high accuracy; the velocity distributions are therefore Maxwellian. The cross
section depends only on the relative velocity v, but not on the velocity V
of the center of gravity. We may therefore transform (2.77) into an integral
over the v and V spaces, and integrate over V . Introducing the energy E =
mv 2 /2, where m = mi mk /(mi + mk ) is the reduced mass, substituting the
Maxwellian distributions, and using (2.75) we find
∞
(8/mπ)1/2
λik = exp(−E/kT )Eσ(E) dE . (2.78)
(kT )3/2
0
The remaining integral can be evaluated most easily when the target
nucleus has a resonance at a particular energy ER , i.e., when σ(E) is large
in a close neighborhood of ER . We may then treat σ as if it was a δ function
and simply set E = ER in the other factors. For such reaction rates, the
dependence on temperature is of the characteristic form
λik ∝ T −3/2 exp(−ER /kT ) . (2.79)
For the Sun resonant reaction rates are of minor importance, although reso-
nances do exist for the p reactions of 7 Li, 7 Be, 12 C, 13 C, 14 N and 15 N. Even at
the temperature of the solar center the exponential factor in (2.79) is so small
that these contributions are negligible; hence they are omitted in Table 2.3.
For non-resonant reactions the cross section σ(E) is determined by a
combination of quantum-mechanical calculation and laboratory experiment.
First, one finds
1 m Zi Zk e2 π
σ(E) = S(E) exp − . (2.80)
E 2E ε0 h
It is plausible that the cross section is proportional to the square of the par-
ticle wavelength, λp , which in quantum mechanics is related to the energy by
λ2p = h2 /2mE. This explains the 1/E factor in (2.80). The exponential factor
is a consequence of the electrostatic barrier between the two reactants with
charges Zi e and Zk e. The wave function of a particle tunneling through such
2.3 Model Ingredients 45
b = m/2Zi Zk e2 π/ε0 h
we find the maximum at
Emax = (bkT /2)2/3 ;
its height is exp(−3Emax /kT ) .
Problem 2.17. Approximate the Gamow peak by a Gaussian, and show that
its full width at 1/e of the maximum is 4(Emax kT /3)1/2 . Calculate Emax
and the width of the Gamow peak for the rates which appear in (2.69) to
(2.71) at T = 1.55 × 107 K, the temperature of the Sun’s center.
Table 2.3. Nuclear reaction rates for the Sun, NA λ, in reactions per second and
per (mole/cm3 ), after Caughlan and Fowler (1988). T9 is the temperature in units
of 109 K
−2/3 −1/3
1
H(p,e+ ν)d NA λpp = 4.01 × 10−15 T9 exp(−3.380 T9 )
1/3 2/3
×(1 + 0.123 T9 + 1.09 T9 + 0.938 T9 )
−2/3 −1/3 1/3
3
He(3 He,2p)α NA λ33 = 6.04 × 1010 T9 exp(−12.276 T9 ) × (1 + 0.034 T9
2/3 4/3 5/3
−0.522 T9 − 0.124 T9 + 0.353 T9 + 0.213 T9 )
−2/3 −1/3
3
He(α, γ)7 Be NA λ34 = 5.61 × 106 a−5/6 T9 exp(−12.826a1/3 T9 ),
where a = 1 + 0.0495 T9
−1/2
Be(e− , νγ)7 Li NA λe7 = 1.34 × 10−10 T9
1/3
7
[1 − 0.537 T9
+3.86 T9 + 0.0027 T9−1 exp(2.515 × 10−3 /T9 )]∗
2/3
−2/3 −1/3
7
Li(p,α)α NA λ17 = 1.096 × 109 T9 exp(−8.472 T9 )
−2/3 −1/3
−4.830 × 108 a−5/6 T9 exp(−8.472a1/3 T9 ),
where a = 1 + 0.759 T9
−2/3 −1/3
7
Be(p,γ)8 B NA λ17 = 3.11 × 105 T9 exp(−10.262 T9 )
−2/3 −1/3 1/3
14 15
N(p,γ) O NA λp14 = 4.90 × 10 T9
7
exp(−15.228 T9 ) × (1 + 0.027 T9
2/3 4/3 5/3
−0.778 T9 − 0.149 T9 + 0.261 T9 + 0.127 T9 )
∗
λe7 must not exceed 1.51 × 10−7 /ne (see text)
Fig. 2.6. Cross-section factor S(E) for the reaction 3 He(α,γ)7 Be, in keV×barn
(1 barn = 10−28 m2 )
2.3 Model Ingredients 47
As the last ingredient required for the solution of the equations of stellar
structure, (2.8), (2.9), (2.12), and (2.19), we must now determine the opacity
κ as a function of variables which are already known. As in the case of ε, we
choose ρ and T as independent variables for κ.
The opacity is an average over frequency ν; the dependence on ν must
therefore be determined first. This is done by a calculation of the probability
that a photon of energy hν is absorbed or scattered by individual atoms, ions,
or electrons. A time-dependent quantum-mechanical perturbation method is
generally used, and the results are given in form of cross sections for the
diverse processes. The present section is a short summary of such results.
Bound–Bound Absorption. Photons with a discrete frequency ν can be
absorbed when an atom or ion undergoes a transition between two states of
energy difference hν. In analogy to classical oscillators the cross section per
atom is written in the form
e2
σbb (ν) = f φ(ν) , (2.83)
4ε0 me c
where me is the electron mass; f is called the oscillator strength and contains
the calculated transition probability; φ(ν) is a normalized shape function
of the absorption, or the line profile. The two most important profiles are
the Gaussian Doppler broadening that arises from the Maxwellian velocity
distribution of the absorbing particles,
1
φD (Δν) = √ exp[−(Δν/ΔνD )2 ] , (2.84)
πΔνD
and the Lorentz profile due to collision broadening,
γ
φC (Δν) = . (2.85)
(2πΔν)2 + γ 2 /4
48 2. Internal Structure
Here Δν = ν − ν0 is the frequency distance from the line center. The widths
of the profiles are determined by the constant of collisional damping γ (or
twice the effective collision frequency), and by the Doppler width
The Doppler profile dominates in the core of the line, while the slowly
decaying damping profile is characteristic for the wings. In general, both
Doppler and collision broadening act together, and the two functions must be
folded. We shall treat the resulting line profile in more detail in the context of
line formation in the solar atmosphere (Chap. 4). For the opacity in the Sun’s
interior, line broadening is especially important because κ is a harmonic mean
where not the lines themselves but the windows between the lines contribute.
For good transparency windows must be kept clean.
Bound–Free Absorption. Bound-free absorption, or photoionization, can
occur when the photon energy exceeds the ionization energy χ of an atom or
ion. For hydrogen, or for hydrogen-like atoms and ions, χ = me e4 Z 2 /8(ε0 hn)2 ,
where Z is an effective charge number and n is the principal quantum number.
The cross section per atom in state n is
me e10 Z 4 gbf (n, ν)
σbf (ν) = √ , (2.86)
48 3πε50 ch6 n5 ν 3
where the Gaunt factor gbf , which only weakly depends on n and ν, results
from the quantum-mechanical calculation of the ionization probability. For
each atom (or ion) and state considered, (2.86) is valid for ν > χ/h, while
σbf = 0 for smaller ν. The characteristic ionization edges and adjacent ν −3
continua can be seen in the example shown in Fig. 2.7.
Bahcall and Ulrich 1988). This may appear little; but, according to (2.19), the
opacity governs the temperature gradient, and thus a reduction of κ leads to a
smaller central temperature. As we shall see in Sect. 2.4.2 (Problem 2.19), the
branching of the pp chains and hence the production of energetic neutrinos
by 8 B decay is extremely sensitive to changes of temperature.
Calculation of the Rosseland Mean. Before the integral (2.20) can be
calculated we must multiply, for each frequency ν , the cross sections for the
various processes by the number density of the respective absorbers, and add
all the results. Thus, for bound-bound, bound-free, and free-free absorption,
we need the number density nj of atoms (or ions) in the state from which
the transition to a higher state (or to the continuum) occurs. These number
densities are provided by the Boltzmann and Saha equations. For scattering
by electrons, we need the electron density, ne , which can be approximated
by the expression (2.45) for full ionization, because electron scattering is
important only in the deep interior. For the three processes involving an atom
(or ion), a correction for stimulated emission, i.e., a factor 1 − exp(−hν/kT ),
must also be applied. An example for the total absorption coefficient obtained
in this way is shown in Fig. 2.7. The 20 most abundant elements, cf. Sect.
4.4, have been taken into account for this calculation.
Let us substitute hν/kT = u in the integral (2.20). The radiative opacity,
or Rosseland mean absorption coefficient, per unit mass is then given by
∞
1 15u4 eu /4π 4 (eu − 1)2
=ρ du , (2.91)
κ (1 − e−u ) j σj nj + σs ne
0
where the sum is over all bound-bound, bound-free, and free-free processes,
with the corresponding cross sections σj and initial-state number densities nj .
Most extensive opacity calculations have been performed by groups at the
Los Alamos Scientific Laboratory (Huebner et al. 1977) and at the Lawrence
Livermore National Laboratory (Iglesias and Rogers 1996). The OPAL tables
of the latter group contain Rosseland mean opacities that explicitly include
the contributions from 19 of the most abundant elements, in addition to
hydrogen and helium. Since hydrogen is converted into helium during the
solar main-sequence evolution, it is necessary in principle to calculate such
an opacity table for each depth in the Sun, and for each solar age. Usually
a small number of tables, and linear interpolation between them, is used.
The Los Alamos tables extend down to kT ≈ 1 eV (11600 K); the OPAL
tables of Iglesias and Rogers (1996) reach further, down to ≈ 5600 K. Com-
plementary tables for the range of lower temperature have been prepared
by Kurucz (1979) and by Alexander and Ferguson (1994). For the present
Sun, Fig. 2.8 shows the opacity as a function of depth, together with its
temperature derivative.
2.3 Model Ingredients 51
Let us first collect the equations which describe our model of the solar interior.
We have four first-order differential equations:
∂r 1
= , (2.8)
∂m 4πρr2
∂P Gm
=− , (2.9)
∂m 4πr4
∂L ∂S
=ε−T , (2.12)
∂m ∂t
⎧
⎪ 3κL
∂T ⎨ − (in a stable layer) , (2.19)
= 256π 2 σr4 T 3
∂m ⎪ ⎩
(∂T /∂m)C (in an unstable layer) , (2.21)
ρ = ρ(P, T ) , dS = dS(P, T ) ,
ε = ε(ρ, T ) , κ = κ(ρ, T ) .
The four constitutive relations have been treated in the preceding sections,
and are here written in symbolic form only. For the differential equations we
must apply four boundary conditions.
Two of the four boundary conditions will be imposed at the center, m = 0.
These conditions are
The boundary values adopted by the solutions of the four differential equa-
tions are rs , Ps , Ls , and Ts . We do not know these values beforehand (except
for the present Sun), but we shall be able to derive 2 relationships between
them; these relationships are the desired boundary conditions.
For this purpose we must consider the atmosphere, here defined as the
layer r > rs , or τ < 2/3. As the energy sources lie deep in the interior,
we have L ≡ Ls in the whole atmosphere. Moreover, we assume (or know
from observation) that the atmosphere is geometrically thin, so that r ≡ rs .
Finally, the atmosphere contains very little mass, so that m ≈ m . In this
case the use of m as an independent coordinate would lead to very inaccurate
results. We therefore replace m by a more rapidly varying quantity such as
the optical depth introduced in (2.93) and set m = m whenever it occurs
as a coefficient.
Being consistent with the simplifications made, we discard the equations
for r and L, (2.8) and (2.12). The pressure equation, (2.9), is transformed
into
∂P Gm
= 2 , (2.94)
∂τ rs κ
and, integrated, leads to the first of our desired surface conditions:
2/3
Gm 1
Ps = dτ (2.95)
rs2 κ
0
For the study of the atmosphere itself this model, called the Eddington ap-
proximation, is much too coarse. But for the present purpose, the derivation
of a boundary condition for the interior model, it is quite sufficient.
2.3 Model Ingredients 53
1
K= I cos2 θ dΩ ,
4π
where the integrals are over the full solid angle Ω. The equation of transfer
dI
cos θ =I −B (2.98)
dτ
then immediately leads to
dF dK F
= 4π(J − B) , = . (2.99)
dτ dτ 4π
If (2.96) is substituted into (2.97) we find J = 3K = I0 and F = 4πI1 /3.
Since F is the constant energy flux, the first of (2.99) yields J = B, while
the second leads to
3
I0 = Fτ + b .
4π
The constant of integration b is now obtained from the “radiation condition”,
namely from the requirement that, at τ = 0, the net radiation into the upper
half-space is the total flux F . Equivalently, we could require that there is no
net radiation into the lower half-space. In both cases we find b = F/2π. Since
I0 = B = σT 4 /π, we have the final result
Ls
T4 = (3τ /4 + 1/2) (2.100)
4πσrs2
for the atmospheric temperature; in (2.100) we have expressed the flux in
terms of the luminosity and radius. This relation is used to evaluate the
integral (2.95), where we need κ(τ ) instead of κ(T ); taken at τ = 2/3, (2.100)
is the second boundary condition at the surface.
We see from (2.100) why τ = 2/3 has been chosen as the “surface” of our
interior model: At this depth the temperature equals the effective tempera-
ture,
In the preceding sections we have collected almost all the information needed
to calculate the evolution of the Sun from an initial homogeneous star to
the present state. The only missing detail is the temperature gradient in the
convectively unstable layers; this will be supplemented in Chap. 6, where
we shall treat the mixing-length theory of convection. However, the results
presented in the present chapter already employ this theory.
Figure 2.9 shows the Sun’s evolution in the familiar Hertzsprung–Russell
diagram. The approach from the Hayashi line to the zero-age main sequence
took ≈ 5 × 107 years, which is of the order of the Kelvin–Helmholtz time.
During its main-sequence life of 4.57 × 109 years the Sun has changed very
little in effective temperature and luminosity. We may conclude that such
evolutionary effects can explain only part of the spread of the stars within
the main-sequence band (another reason for the finite width of the main
sequence is that the initial chemical composition is not exactly the same for
all stars). The asterisks in Fig. 2.9 represent visual binary stars for which
Teff and L have been determined with the best possible accuracy [taken from
Audouze and Israël (1985, p. 252)].
For the Sun itself, and even more so for the Earth, the change of the
solar luminosity during its main-sequence life is, however, substantial. The
increase from about 0.7L to the present value often has been considered as a
great puzzle. If a reduction of the “solar constant” by a factor 0.7 happened
today, it would lead to dramatic cooling and very probably to a complete
ice cover of the Earth. But if the Earth was completely iced over once, it
would remain so even if the incoming radiation was increased to the present
level – a consequence of the increased albedo. Yet the present Earth is not
ice-covered, and geological evidence suggests that it never was.
The solution to this puzzle probably lies in the evolution of the Earth’s
atmosphere. For example, if the early concentrations of ammonia and (or) of
carbon dioxide were higher than today, the greenhouse effect in the infrared
could have compensated for the lower solar luminosity (e.g., Wigley 1981).
As shown in Fig. 2.10, the major part of the luminosity increase is due
to an increase of the radius from about 0.87r to the present radius. During
this evolution the effective temperature changed by only about 125 K, and
hence contributed less to the luminosity change (the neutrinos, also shown
in Fig. 2.10, will be considered in the following section).
A model of the present Sun, of age 4.57 × 109 years, is compiled in Ta-
ble 2.4. It is a standard model, based on OPAL tables for the equation of
state and opacity, on nuclear reaction rates as listed in Table 2.3, and on the
local mixing-length theory as described in Chap. 6; it includes diffusion of
helium as outlined in Sect. 2.3.3. The levels in the table are not equidistant
in any of the variables; they are chosen so that both the outer layers and
the central region are reasonably resolved. The surface (the first line) is at
optical depth τ = 2/3. The pressure and the density change very rapidly in
the outer layers, while the luminosity and the hydrogen content vary only in
the deep interior. The mean molecular weight μ first decreases inwards due
to ionization of the abundant elements hydrogen and helium, but increases
again near the center, where half of the hydrogen has already been converted
into helium.
The solar mass is strongly concentrated towards the center. The outer
convection zone, which extends from immediately below the surface down to
r/r ≈ 0.71, contains only about 2.5% of the total mass. And only about
10% of the mass lies outside r/r = 0.5, although 7/8 of the volume is there.
– There is no interior convection zone in this model.
A number of other properties of the present Sun’s interior have already
been discussed in earlier sections, and are illustrated in Figs. 2.2 to 2.5, and
2.8. Here we only add the distribution of 3 He. Consistent with the lifetime
variation of this isotope inside the Sun, there has been an accumulation near
r = 2 × 108 m (Fig. 2.11). The height of this peak is almost 100 times the
original uniform abundance Y30 = 4 × 10−5 adopted for this model.
2.4.2 Neutrinos
Since about 1970 the main motivation for the calculation of solar interior
models has been the persistent discrepancy between the predicted and mea-
sured flux of neutrinos from the Sun. We must therefore treat the solar neu-
trino production in some detail.
First it must be mentioned that, in addition to the reactions already
discussed in Sect. 2.3.6, an important source of neutrinos is the “pep reaction”
−
p(pe , ν)d , Qν = 1.442 Mev . (2.101)
As a three-particle reaction, the pep reaction occurs rather infrequently, only
at about 0.25% of p(p,e+ ν)d (which we abbreviate by p+p). For the energy
production the pep reaction is therefore unimportant; in fact it constitutes a
small loss because the total energy goes into the neutrino. Nevertheless the
58 2. Internal Structure
pep reaction has some importance for the neutrino experiments since Qν is
markedly above the threshold of 814 keV of the 37 Cl detector (Sect. 3.6.2).
Figure 2.12 shows an energy spectrum of solar neutrinos, as predicted by a
standard solar model for the distance of 1 astronomical unit. The continuous
neutrino spectra shown in this figure are the usual β-decay spectra, which
account for the conservation of energy and angular momentum. On the other
hand the three-particle pep reaction and the electron capture of 7 Be produce
monoenergetic neutrinos. The 7 Be neutrinos occur in two lines at 862 keV
(89.7%) and 384 keV (10.3%), because the resulting 7 Li nucleus can be either
in the ground state or in its first excited state.
The spectral distribution of the solar neutrinos strongly depends on tem-
perature T , as a consequence of the dependence of the pp branching on T
(Problem 2.19). This is illustrated in the evolution of the neutrino flux that
would be detectable by 37 Cl (Fig. 2.10). As the central temperature evolves
from 1.35 × 107 K to 1.56 × 107 K, this flux, which largely arises from ppIII,
increases much more than the luminosity, which almost exclusively is gener-
ated in ppI and ppII. In contrast to this, the flux of neutrinos that would be
detectable by a gallium experiment follows much more closely the evolution
of the solar luminosity. The threshold of 71 Ga at 233 keV is low enough to
allow the capture of a substantial fraction of neutrinos from the p+p reaction
which is largely responsible for the energy generation.
Of course the total neutrino flux can be predicted without detailed know-
ledge of the Sun’s interior: since two neutrinos are produced along with each
4
He nucleus, i.e., on the average, for every 25 MeV of the Sun’s radiative
output, the total flux can directly be obtained from the luminosity.
2.4 Results for a Standard Solar Model 59
Problem 2.19. Write the nuclear reaction rates listed in Table 2.3 in the
form r = r0 T η and calculate r0 and η at various temperatures between
107 K and 1.6 × 107 K. Determine the branching ratios for the pp chains
for the solar center (use Table 2.4).
The neutrino fluxes from the various sources can be calculated as integrals
over the reaction rates, rj , as given by (2.75):
m
1
Φj = rj dm ; (2.102)
4πA2
0
Table 2.5. Neutrino fluxes, cross sections (from Bahcall and Ulrich 1988, Bahcall
et al. 1996, and Bahcall 1997), and predicted capture rates for the chlorine and
gallium experiments
Other standard models predict capture rates that are similar to those
listed in Table 2.5. The uncertainty partly comes from uncertainties in the
opacity and in the heavy element abundance, which affect the temperature
and hence the relative frequency of the pp chains. Another source of possible
errors are the reaction rates themselves (Bahcall et al. 1998). For the flux
of 7 Be neutrinos the largest uncertainty comes from the 3 He-4 He reaction,
mainly due to the need of extrapolating the S factor down to low energy, cf.
60 2. Internal Structure
Fig. 2.13. Solar neutrinos: Prediction from the standard solar model BP98 of
J. N. Bahcall and M. H. Pinsonneault (high columns) and experimental results
(lower columns), for the 37 Cl (left), water (middle), and 71 Ga detectors (right). The
shading indicates the contributions to the theoretical prediction from the diverse
nuclear reactions in the Sun, as indicated at the bottom
Sect. 2.3.6; for the 8 B neutrinos the largest uncertainty lies in the S factor
of the 7 Be + p reaction.
The neutrino fluxes from the decays of 13 N and 15 O listed in the last two
lines of Table 2.5 are not the same because the CNO cycle has not yet reached
equilibrium. Their total contribution to both the 37 Cl and 71 Ga experiments
is, however, so small that the error would be insignificant if equilibrium had
been assumed.
The experimental results for 37 Cl and 71 Ga, as well as for the water
Čerenkov detectors (Sect. 3.6.2), are shown in Fig. 2.13, together with the
theoretical predictions. Three main conclusions can be drawn from these re-
sults: The first is that solar neutrinos have indeed been detected. This directly
confirms the concept of nuclear energy generation in the Sun. The second is
that the experimental result is significantly below the theoretical expecta-
tion based on the standard solar model. This is the often quoted neutrino
discrepancy, or the “solar neutrino problem”. The third is that the neutrino
discrepancy depends on neutrino energy; this conclusion is derived from the
different energy thresholds of the three detector types.
2.5 Non-Standard Models 61
Two ways of resolving the neutrino discrepancy have been proposed. The
first is to modify the solar model in such a manner that the relative frequency
of the pp reaction chains is modified, especially so that the ppIII chain is
reduced and hence less 37 Cl captures and less neutrino-electron scattering
events in the water detectors are predicted. This has been attempted with
non-standard solar models some of which will be presented in Sect. 2.5 below.
However, this “astrophysical” solution of the solar neutrino problem appears
to fail in a quantitative test. The total flux from the ppIII branch is ∝ Tc18 ,
where Tc is the temperature at the Sun’s center (Bahcall 1989). A model
with a, say, 3% reduction of Tc would therefore correctly predict the results
of Kamiokande and Super-Kamiokande. On the other hand, the yield of the
ppII branch is ∝ Tc8 , which means that the same temperature reduction would
not sufficiently reduce the flux of 7 Be neutrinos for a correct prediction of
the gallium results (Fig. 2.13).
The alternative is to reconsider the physics of the neutrino. Already in
1957 B. Pontecorvo discussed the possibility of neutrino oscillations, that
is the change from one neutrino flavor into another. In the solar case this
could mean that, during their transit from the Sun to Earth, the neutrinos,
which all are generated as electron neutrinos, partially change their flavor
and become μ or τ neutrinos. Such neutrino oscillations are not part of the
standard model of elementary particles; on the other hand, they are not
forbidden by any fundamental principle. A necessary condition is that at
least one type of neutrino must have a finite rest mass, which would be
consistent with measurements at nuclear reactors. Even with a very small rest
mass (mν c2 1 eV) the MSW effect, so named after the work of Mikheyev
and Smirnow (1985) and Wolfenstein (1978), would permit efficient neutrino
oscillations in the presence of other matter, i.e., in the deep interior of the
Sun. In either case, both μ and τ neutrinos would escape detection in the
present experiments, and the capture rate would be greatly reduced (see e.g.,
the discussion of Bethe 1986).
First results from the heavy-water detector at the Sudbury Neutrino Ob-
servatory (Sect. 3.6.2) indeed indicate that neutrino flavor transformation
is the answer. The charged-current reaction (3.97), which is sensitive only
to electron neutrinos, yields a significantly smaller flux than the elastic-
scattering reaction (3.96) which, in addition to being sensitive to νe , has
a small sensitivity to the two other flavors (Ahmad et al. 2001).
some of the non-standard models because they shed light on the internal
constitution of a star like the Sun.
the rotation of the solar surface! Detailed calculations of solar models with a
depth-dependent angular velocity (Bartenwerfer 1973) confirm this estimate.
A solar model with a such a rapidly rotating core is excluded by obser-
vations. One point is the oblateness of the core, which of course is invisible
but which adds a quadrupole moment to the gravitational field. The outer
layers of the Sun would be deformed, and the solar surface would show an
oblateness in excess of the observed ≈ 10−5 that is due to the (slow) surface
rotation (cf. Chap. 7). The other point is the observed rotational splitting of
solar oscillation frequencies which only allows for a small core that rotates at
most twice as rapidly as the solar surface (Chaplin et al. 1996 b, 1999; Lazrek
et al. 1996; Charbonneau et al. 1998).
The model with a strong internal magnetic field is formally related to the
model with the rapid internal rotation. Assumed is a randomly oriented,
small-scale field which, as a net effect, applies pressure to the material which
it permeates. This magnetic pressure (cf. Chap. 8) acts in complete analogy
to the before-mentioned centrifugal potential. For a 10% effect the necessary
field strength follows from
B 2 /2μ = P/10 . (2.104)
With a central pressure of 1016 Pa (Table 2.4) this means that the required
field strength for a noticeable effect is of order 109 G. Again detailed calcu-
lations of Bartenwerfer (1973) confirm this estimate.
The problem here is that a small-scale field would not survive for 4.6×109
years. Due to ohmic dissipation the electric current that supports a field of
scale l decays at a rate proportional to l−2 . If l r the lifetime of the field
would be much smaller than the Sun’s age (cf. Chap. 8).
Mixing in the solar core would replace some of the helium accumulated near
the center by hydrogen and thereby reduce the mean molecular weight μ.
A lower temperature is then sufficient to give the same thermal pressure
P ∝ T /μ, and this eventually leads to lower neutrino detection rates in the
37
Cl and water experiments, by the same reasoning as in the case of the other
non-standard models. The mixing has been proposed both as a statistically
stationary process (Schatzman et al. 1981) and as a time-dependent, inter-
mittent process (Dilke and Gough 1972). In the latter case, an instability of
the 3 He accumulated near r = 2 × 108 m (Fig. 2.11) would cause the mixing.
The hydrostatic equilibrium would immediately adjust itself to the changed
μ, but the thermal adjustment to the new stratification would take a Kelvin–
Helmholtz time (2.6), of order 3 × 107 years. In such a time-dependent model
64 2. Internal Structure
An analytical version of the minimization method for the free energy has
been obtained by Däppen (1980), who also quotes earlier authors who em-
ployed (2.30). In particular, Däppen introduces an analytical approximation
for the partition functions which appear in the Saha equations (2.37); this
approximation is valid in the framework of the Debye–Hückel treatment.
More detailed studies of the electrostatic interaction between the charged
particles in a stellar plasma have of course been made. Here I mention only
Stewart and Pyatt’s (1966) treatment of the lowering of ionization potentials,
with its relation to the concept of “pressure ionization”, and the general
discussion of Rogers (1984) of the ionization equilibrium and its effect on the
equation of state in the solar interior. More references can be found in this
latter paper. The influence on solar models has been investigated by Ulrich
(1982). An attempt beyond the Debye–Hückel approximation has been made
by Meister et al. (1999).
As far as the thermodynamic relations are concerned the present text
follows Baker and Kippenhahn (1962). As a general text in thermodynamics
I found the monograph of Callen (1985) most useful. In the context of the
equation of state, thermodynamic quantities have been treated by Däppen
et al. (1988) and Mihalas et al. (1990).
Nuclear reaction rates are most thoroughly described in the book of Clay-
ton (1968). In particular the exponential factor arising from the penetration
of the Coulomb barrier is discussed in detail. Rates at energies close to the
Gamow peak have been measured by Junker et al. (1998). Adelberger et al.
(1998) give a critical discussion. Electron screening, first treated by Salpeter
(1954), is described in Chap. 17 of Cox and Giuli (1968). Attempts to reduce
the solar neutrino discrepancy also include modifications in the treatment of
electron screening (Dzitko et al. 1995, Dar and Shaviv 1996), but Gruzinov
and Bahcall (1998) find a slight increase of the neutrino rates.
For the calculation of opacities I refer to the review of Huebner (1986)
and to the results of the Opacity Project (Seaton 1995, Berrington 1997). A
collection of Los Alamos tables has been compiled by Weiss et al. (1990). As
far as “collective effects” on electron scattering are concerned, the original
estimate by Diesendorf (1970) has been revised by Boercker (1987). The
effects of opacity changes on the solar model have been pointed out by Kim
et al. (1991), Faulkner and Swenson (1992), Charbonnel and Lebreton (1993),
and by Tripathy and Christensen-Dalsgaard (1998).
An extensive description of the program used here to calculate the solar
main sequence evolution has been given by Kippenhahn et al. (1967). In one
form or another, this code has been used by many different groups. When
results are compared one should be aware of this. The iteration used es-
sentially is a multi-dimensional Newton–Raphson method, and is also called
the Henyey method because in the context of stellar evolution it was first em-
ployed by Henyey et al. (1959). Many related codes have been developed since
then. The programs of, e.g., Eggleton (1971, 1972), Christensen-Dalsgaard
66 2. Internal Structure
3.1 Limitations
3.1.1 General Difficulties
Problem 3.1. An area of 0.3 × 0.3 with average intensity (Fig. 1.4) on
the solar surface is observed with a 60-cm telescope and a high-resolution
spectrograph. The desired exposure time is 1 ms, and the bandwidth is
10 mÅ at λ = 5000 Å. How many photons are detected if the efficiency of
the entire telescope-spectrograph-detector system is 1%?
The answer is: 400 photons, i.e. the photon noise is 1/20 or 5%. But often
the requirements are still more stringent: One would like to observe in a dark
sunspot, where the intensity is an order of magnitude less, or in the center
of a strong spectral line with, e.g., one tenth residual intensity, or both at
the same time. Or one would like to observe polarization of, say, 1% or less!
These examples illustrate that even in solar physics the number of available
photons can be a problem.
The methods to overcome this problem are the same as in general as-
tronomy: First, a larger aperture is used to collect more photons. Currently
the largest solar telescope in operation is the McMath–Pierce telescope of
the National Solar Observatory at Kitt Peak, USA, with a clear aperture of
1.52 m; the largest presently in planning is ATST, the Advanced Technology
Solar Telescope, with a 4 m aperture. Second, the least possible fraction of
photons should be lost. This is achieved especially by the use of image inten-
sifiers or highly efficient detectors. The most common detector is the CCD
or “Charge-Coupled Device”, a detector based on the internal photoelectric
effect in semiconductors. Such a detector has a quantum efficiency of up to
90%, which must be compared to the 5–25% (depending on wavelength) of a
photomultiplier, or the ≈ 1% of a photographic emulsion!
Another limitation that solar physics has in common with general as-
tronomy is atmospheric extinction. Of course this can only be overcome by
observing from above the atmosphere. Numerous solar experiments have been
carried out with the help of rockets or satellites since about 1950. Most no-
table are a series of Orbiting Solar Observatories (OSO 1 – OSO 8) between
1962 and 1975, the Apollo Telescope Mount (ATM) on the Skylab mission
(May 1973 to January 1974), the Solar Maximum Mission (SMM, 1980 –
1989), and Spacelab 2 (1985). More recent space missions are:
Ulysses. The Ulysses mission is the first spacecraft that explores the inter-
planetary space at high solar latitudes. It was launched on October 6, 1990
and, through a close fly-by at Jupiter in February 1992, entered a 6-year ec-
centric orbit that is highly inclined with respect to the ecliptic. The maximum
southern latitude of −80o with respect to the solar equator was first reached
in September 1994, the maximum northern latitude of 80o in September 1995.
Ulysses carries a number of instruments to study in situ the solar wind and
the interplanetary magnetic field at high solar latitude: composition, parti-
cle energies, plasma waves and radio waves, etc.. Other instruments observe
dust grains in the interplanetary space, or monitor bursts of solar x-rays and
cosmic γ-rays. The Ulysses mission is described in Astron. Astrophys. Suppl.
Ser., Vol. 92 (1992).
Yohkoh. The Yohkoh satellite is an observatory for the study of x-rays and
γ-rays from the Sun. It was launched on August 31, 1991 into a near-Earth
orbit, and carries a telescope for soft x-rays, a telescope for hard x-rays (up
to 100 keV), and spectrometers for a wide range of wavelengths from soft
x-rays up to γ-rays of 100 MeV. Its main purpose is the investigation of flares
and of transient phenomena in the solar corona. The Yohkoh instruments are
described in 6 articles in Solar Physics, Vol. 136 (1991).
SOHO. The Solar and Heliospheric Observatory was launched on December
2, 1995 and inserted into its orbit around the Lagrangian Point L1 on March
20, 1996. There, ca. 1.5 × 106 km sunward from the Earth, its instruments
have uninterrupted sunlight, which is especially important for the three in-
struments that record the global oscillations of the Sun. Five instruments on
3.1 Limitations 69
the SOHO satellite take images and spectra in diverse wavelength bands from
the ultraviolet to soft x-rays, whereby the transition layer between the chro-
mosphere and the corona, and the corona itself, can be studied; in addition,
the corona is monitored out to a distance of more than 10r by a triple-
coronagraph. Three further instruments investigate the solar wind in situ, in
particular the energy distribution of its constituents. The SOHO mission is
described in detail in a volume edited by Fleck et al. (1995).
TRACE. The Transition Region And Coronal Explorer, launched on April 2,
1998, carries a 30-cm normal-incidence telescope that images fields of 8.5 × 8.5
on the Sun. Diverse wavelength bands between 28 nm and 250 nm are used,
so that the continuum as well as the emission of highly ionized atoms can
be studied, characteristic for temperature ranges from 6000 K to 1000 000 K.
The images have a spatial resolution of better than 1 ; images of the whole
Sun are composed in a mosaic. Since the orbit is Sun-synchronous, long un-
interrupted observations of the transition region and the corona are possible.
Handy et al. (1999) describe the TRACE mission.
The second serious limitation to solar observations, which also has its ori-
gin in the Earth’s atmosphere, is seeing. Seeing is the degradation of image
quality by fluctuations of the refractive index in the light path. As such it
is of course a general astronomical problem. However, in the context of solar
observations the difficulties are aggravated by the Sun itself. The Sun causes
thermal convection in the entire troposphere, it heats the ground around
the observatory and thus initiates local convection, it heats the observatory
building, in particular the dome, and finally the telescope into which it must
shine. Any heated surface is a source of air instability. Since the viscosity of
air is small, the resulting convective motion will have a large Reynolds num-
ber and therefore be turbulent. The concomitant temperature fluctuations,
up to 0.1 K in the free atmosphere but frequently much larger in and around
the building and telescope, affect the refractive index and hence generate
wave-front aberrations. The refractive index varies according to
P/P0
n − 1 = 2.79 × 10−4 , (3.1)
T /T0
where P0 = 1 bar and T0 = 273 K. Since atmospheric turbulence is nearly
in pressure equilibrium, it is predominantly the temperature fluctuation that
causes the seeing.
Image degradation by seeing is a very complicated process, but often
three different aspects can be identified. Blurring is the defocusing effect of
air schlieren with an index of refraction that varies from place to place. The
whole image looses its sharpness. If the image essentially remains sharp but
is rapidly shifted back and forth we speak of image motion. If substantial
70 3. Tools for Solar Observation
parts of the image remain sharp but are shifted relative to each other then
we have image distortion. The frequency spectrum of image motion typically
reaches up to 100 Hz (e.g., Brandt 1969). Exposures of 10−2 s and faster can
therefore substantially reduce this aspect of solar seeing.
Point Spread Function and Modulation Transfer Function. Let us
consider a real image formed in the x, y-plane by a solar telescope. At each
point the intensity I(x, y) has contributions which “really” belong to neigh-
boring points of the (non-existing) perfect image. Introducing the point spread
function PSF(x, y; ξ, η) we have
∞
I(x, y) = I0 (ξ, η) PSF(x, y; ξ, η) dξdη . (3.2)
−∞
Fig. 3.1. Modulation transfer function for diffraction and seeing: (a) constant qm ,
or given telescope aperture, as a function of seeing; (b) constant s0 , or given seeing,
as a function of aperture. The label on the curves is s0 qm in both cases
Here z is the zenith angle and k is the wave number of the light. The factor
k 2 arises from the fact that the phase change is proportional to that wave
3.1 Limitations 73
where the numerical factor originates from Fried’s comparison of the seeing-
limited resolution with the diffraction limit of a circular aperture. Thus, the
Fried parameter r0 is the aperture of an imaginary diffraction-limited tele-
scope which has the same resolving power as a large, optically perfect tele-
scope used under the conditions of the actual seeing. As the seeing is variable,
r0 depends on the observatory site and on time; according to (3.13) its de-
pendence on wavelength is r0 ∝ λ6/5 .
Problem 3.5. Draw the modulation transfer function for the case of equal
half-widths of the diffraction and seeing point spread functions. What is
the value of s0 qm ?
Site Selection. A substantial part of the seeing originates in the free atmo-
sphere and is outside the astronomer’s control. The first step in the construc-
tion of a solar observatory, therefore, is to select an appropriate site where
the effects of the atmospheric turbulence are tolerable.
The search for a site may include in-situ measurements in the free at-
mosphere. For example, K. O. Kiepenheuer used an aircraft equipped with a
micro-thermal sensor to measure temperature fluctuations of down to 0.01 K
with a spatial resolution of 10 cm. Flights over and around the Canary Is-
lands (Fig. 3.2) revealed that, above the inversion layer, an intermediate
wind speed, say 5 m/s to 15 m/s, is favorable: Not violent enough to shake
the observatory building, it blows the locally generated turbulence off the
mountain.
Specially equipped radiosondes have also been used for in-situ measure-
ments of atmospheric temperature fluctuations. The results can be converted
into a value of the Fried parameter r0 as long as the distance between
two points of measurement falls into the inertial range of the atmospheric
turbulence, i.e., lies between the large outer scale of energy input and the
small inner scale of energy dissipation (in the Earth’s atmosphere this means
roughly between 30 m and 3 mm). For the Canary Islands the radiosonde
results (Fig. 3.3) show that most often 20 cm < r0 < 30 cm. These values
74 3. Tools for Solar Observation
are rather large, but the measurements do not include the seeing caused by
ground-level turbulence.
A method that includes the effect of all layers is the spectral-ratio method
proposed by von der Lühe (1984). It is based on a large number of short-
exposure images of the same object. If this object is structured on scales below
the seeing limit, such structure will be manifest in the Fourier transform, e.g.,
for the ith image
Fi (q) = F0 (q)Si (q) , (3.14)
where F0 is the transform of the object. The average |Si |2 over all images
will still be non-zero at spatial frequencies q for which the average transfer
Fig. 3.3. The Fried parameter r0 deduced from radiosonde measurements of tem-
perature fluctuations. Circles include measurements up to 10 km, crosses up to
17 km. After Barletti et al. (1977)
3.1 Limitations 75
function Si (q) has already dropped to an insignificant value. The spectral
ratio
|Fi (q)|2 |Si (q)|2
ε(q) = = (3.15)
|Fi (q)|2 |Si (q)|2
therefore has a cutoff at a certain spatial frequency that is related to the
Fried parameter r0 . Results based on this method are shown in Fig. 3.4. The
values are markedly smaller than those of Fig. 3.3, typically 10 cm.
The Fried parameter is a statistical property and therefore relevant for
long-exposure observation (i.e., long compared to the time scale of the seeing).
Occasionally the conditions might be more homogeneous than indicated by
the above results; the resolution of short-exposure images will be much better
during such moments.
Fig. 3.5. Solar granulation, in a field of 30 000 km × 20 000 km on the Sun. The
image has been taken at the Swedish Vacuum Solar Telescope on the Roque de los
Muchachos, La Palma, Canary Islands. Exposure 13 ms in the blue spectral region,
468 ± 5 nm. Courtesy P. N. Brandt, G. W. Simon, and G. Scharmer
From many test measurements it now appears that the most favorable
site for a solar observatory is on a high mountain (above the inversion layer)
on an island surrounded by a large water surface, which makes the incoming
air mass as homogeneous as possible. The Canary Islands and Hawaii seem to
satisfy these conditions very well and at the same time have enough sunshine
to justify the construction of large observatories. Images of the solar granula-
tion (Fig. 3.5) confirm this conclusion. Of course, photographs of comparable
quality have been obtained for some time at other observatories, notably at
Pic du Midi. But more persistent seeing conditions, which are required for
evolutionary studies of the small-scale photospheric structure, are met at the
island sites.
Local Turbulence. Beyond a careful selection of a good observatory site
nothing can be done about the seeing originating in the free atmosphere. But
the effects of ground-level turbulence can considerably be reduced by the use
of a high building. Tower telescopes were first constructed by G. E. Hale at
Mt. Wilson, California; his 20 m and 50 m solar tower telescopes, completed
in 1908 and 1911, were pioneering instruments. Since then these telescopes
provided important data, predominantly of velocity and magnetic fields on
the solar surface, and they are still in use today.
Local turbulence develops during the morning hours when the Sun heats
the ground around the observatory. This effect is reduced if the observatory
3.1 Limitations 77
Fig. 3.6. Big Bear Solar Observatory of the New Jersey Institute of Technology.
Courtesy H. Zirin
is surrounded by water, which has a high thermal inertia. Big Bear Solar
Observatory, situated on an artificial island in Big Bear Lake, California,
USA (Fig. 3.6), is an example. The scintillation measured at this site is
particularly small. Other lake sites are the Udaipur Solar Observatory in
Rajasthan, India, the Huairou Solar Station of the Beijing Observatory in
China, and the San Fernando Observatory at Northridge, California, USA.
Open Structure and Domeless Design. The design of the observatory
building is another important matter. We have already seen that a wind of
intermediate strength is welcome because it clears the site from local tur-
bulence. Exposing the coelostat (Sect. 3.2.2) to the open air may therefore
be better than protecting it by a dome. Examples are the the 60-cm tower
telescope at Kitt Peak, Arizona, and the very similar 70-cm telescope (Fig.
3.7) at Izaña, Tenerife, which have a cylindrical dome that can be completely
retracted, but also partially opened to act as a wind screen. An additional
78 3. Tools for Solar Observation
Fig. 3.7. The German Vacuum Tower Telescope at the Observatorio del Teide,
Izaña, Tenerife. The dome is partially retracted. Courtesy D. Soltau
Fig. 3.9. The Richard B. Dunn Telescope at Sacramento Peak, New Mexico, USA.
Courtesy National Solar Observatory, AURA, Inc.
more rapidly in helium than in air because of the larger thermal conductivity.
Both effects reduce the internal seeing.
Other Passive Measures. White paint on the whole building reflects the
sunlight and so helps to avoid unwanted heating. Mostly TiO2 is used, which
has a high reflectivity in the visible, but is nearly black in the far infrared
and hence radiates the building’s own heat away.
Thermal problems often are worst at the edges of optical elements. These
problems can therefore be alleviated by using oversize optical elements, i.e.,
by larger windows, lenses, mirrors, etc., than would correspond to the ac-
tual aperture of the telescope. The aperture is then defined by an additional
diaphragm (“stop”) behind (or in front of) the main focusing objective or
mirror. A white annular diaphragm just outside the main lens has been used
with great success at several observatories.
3.1 Limitations 81
tensifiers can reduce the exposure time by a factor of 10 or more, and thus
help in an indirect way. Figure 8.9 shows a spectrum taken with the aid of
an image intensifier.
There are several methods of finding an estimate of the best image of an ob-
ject from degraded images. We briefly mention two of these image-restoration
methods, and show examples of solar applications.
Speckle Interferometry. Speckles are seen as a grainy interference pattern
in images formed by laser light reflected from a diffusive surface, or in images
of astronomical objects taken with short exposure, say 10 ms. Labeyrie (1970)
first suggested that information beyond the seeing limit may be obtained
from such short-exposure images. Taking the absolute magnitude of (3.14)
and averaging we obtain the power spectrum of the object,
|Fi (q)|2
|F0 (q)|2 = . (3.16)
|Si (q)|2
In this expression the phase of the object Fourier transform has been
lost. Nevertheless it is possible to recover essential information about certain
objects, e.g., the angular separation of two stars, if |Si (q)|2 , the speckle
transfer function, is known. In the stellar case this function can be obtained as
the power spectrum of a point source if such is available within the isoplanatic
angle. The solar case is more complicated because there are no point sources
on the Sun. Therefore models have been derived (e.g. Roddier 1981) that
depend on the statistical properties of the atmosphere as well as on the
diffraction of the actual telescope.
In order to restore the image of an extended object, such as an area on
the solar surface, the Fourier phase is required in addition to the Fourier
amplitude. For this purpose Knox and Thompson (1974) have proposed to
use the autocorrelation of the image transform (3.14), or the cross spectrum
Fi (q)Fi∗ (q + Δq) = F0 (q)F0∗ (q + Δq) · Si (q)Si∗ (q + Δq) , (3.17)
where the left-hand side is an average over a large number of short-exposure
image transforms, while the right-hand side contains the cross spectrum of
the object and the corresponding signal transfer function Si (q)Si∗ (q + Δq).
Again models have been calculated that include atmospheric and telescopic
properties (von der Lühe 1988). A most important result is that the signal
transfer function is real. Therefore the phase difference occuring in (3.17),
φ(q) − φ(q + Δq), is due to the object cross spectrum alone. An integration
then yields the Fourier phase of the object. The amplitude is determined as
above, and an inverse Fourier transform restores the image.
A further extension of the Knox-Thompson method is speckle masking, so
named because it corresponds to the application of certain masks to speckle
84 3. Tools for Solar Observation
Fig. 3.11. Image restoration by speckle masking. Sunspot (left) and anomalous
granulation (right). Speckle images taken at λ ≈ 550 nm by C. R. de Boer and
C. Denker with the 70-cm Vacuum Tower Telescope at Izaña, Tenerife. Both pic-
tures cover a field of 24 × 17 . Courtesy F. Kneer
which, by (3.14), is the product of the object bispectrum and a transfer func-
tion that is again real and can be constructed on the base of an atmospheric
model and the telescopic diffraction (von der Lühe 1985). Hence the Fourier
phase of the object can be recovered. Figure 3.11 shows results obtained with
the speckle-masking technique.
Phase Diversity. In its simplest version the phase-diversity method uses
two images taken simultaneously of the same object to derive two unknown
functions, namely the desired intensity distribution I0 and the distorted wave
front in the entrance pupil of the telescope. We consider two convolutions of
type (3.2), each with additive noise Ni , i.e.,
Ii = I0 ∗ PSFi + Ni , i = 1, 2 . (3.19)
The two point spread functions PSFi contain information from the same
atmospheric degradation. They differ from each other because the phase of
the complex wave field is changed (“diversified”) in a well-defined manner for
the formation of the two images. One way to achieve this is to take one image
in the focal plane and the other slightly out of focus, at a distance Δz, say;
this was first proposed by Gonsalves and Chidlaw (1979), the application
to solar images has been described by Löfdahl and Scharmer (1994). The
equivalent phase difference at the edge of the entrance pupil is
2
πΔz D
Δφ = , (3.20)
4λ f
3.2 High-Resolution Telescopes 85
Fig. 3.12. Image restoration of a sunspot section, 7.7 × 7.7 , by phase diversity.
Original images, focussed (left) and defocussed (middle), and restored image (right).
Exposure 48 ms, Vacuum Tower Telescope, Izaña, Tenerife; bandpass 2.6 nm at
λ = 569.5 nm. Courtesy A. Tritschler
as can be shown, e.g., by means of spherical waves that originate from the
focus and from a point displaced from the focus by Δz, and propagate towards
the entrance pupil. In practical work Δz should be chosen such that the phase
difference at the edge of the aperture is at least of order π, say, so that the
two point spread functions differ in a significant manner, and a solution of the
system (3.19) can be found. In the presence of noise this solution is obtained
by a simultaneous minimization of |Ii − I0 ∗ PSFi |2 , for i = 1, 2.
The phase-diversity method relies on isoplanatic degradation, i.e. on the
uniformity of the point spread function over the field of view. In reality this
condition is satisfied only over small sections of a few arcseconds angular ex-
tent. Larger images must therefore be composed of sections that are restored
separately. Figure 3.12 illustrates the method. The phase-diversity method
can be generalized to several instead of two images, and to different means
of diversifying the phases. The method has also been combined with speckle
interferometry (Paxman and Seldin 1993).
where f is the focal length, r the solar radius and A the distance of the
Sun. If d = 46.5 cm as in the example above, then f must be 50 meters.
Such a long focal length is a major characteristic of many solar telescopes.
This is in contrast to common astronomical telescopes for which the light-
gathering power [proportional to (D/f )2 for extended objects and to D2 for
point sources] is of greater importance.
A large focal length makes a freely steerable solar telescope inconvenient.
Special mirror arrangements are therefore used in order to guide the sunlight
into a fixed direction: the heliostat, the coelostat, and the turret, which we
shall describe in the following section. For a tower observatory the fixed
direction is mostly vertical, and the height of the tower is often adapted to
the desired focal length.
Another possibility for obtaining a solar image of order 0.5 m in diame-
ter is to use a smaller primary focal length, and then to form a secondary
image that is large enough for high-resolution work. The diameter d of the
secondary image then defines the effective focal length, feff , via (3.21). A
great number of solar telescopes now in use are of this type. Generally these
telescopes are steerable; however, in order to feed a spectrograph or other
heavy post-focus instruments a fixed beam is still necessary. Mostly a Coudé
arrangement is chosen where the light is directed into the polar axis.
In all cases the required scale of the solar image essentially influences the
design of the telescope. The development of detectors with finer resolution, of
order a few microns, is a step forward. Not only the telescope itself but also,
for example, the accompanying spectrograph can thus be more compact. For
space experiments this is a particular advantage.
The design of solar telescopes is somewhat facilitated by the fact that
only a part of the sky must be within reach: the Sun always stays in the
declination range ±23.5o with respect to the celestial equator.
Fig. 3.14. Coelostat of the German Vacuum Tower Telescope at the Observatorio
del Teide, Izaña, Tenerife. Right, primary mirror, mounted on the polar axis; upper
left, secondary mirror, at the variable mast. Courtesy W. Schmidt
88 3. Tools for Solar Observation
reflects the light into the desired direction, which is horizontal in some, but
vertical in most cases. The vertical coelostat has an advantage in that the
thermal stratification is perpendicular to the light path; it also naturally fits
into the tower observatory. The height of the secondary mirror remains fixed
during the day, but must be adjusted over the year to the Sun’s varying dec-
lination. Figure 3.14 shows a rather low position (summer). In order to avoid
the second mirror’s shade to fall at a particular time of the day onto the first
mirror, the latter can be moved sidewards on a track.
The angles of incidence on the coelostat mirrors are time-dependent, but
in general smaller than for the heliostat. At any given declination and time of
the day there is a coelostat position on the track which minimizes the angles
of incidence and thus the instrumental polarization.
The coelostat mirrors do not rotate the image. For this reason today
coelostats are much more in use than heliostats, even though there is an
additional reflection.
The Turret. The turret, also called the “sunseeker”, is an altazimuth mirror
arrangement which is very compact and thus minimizes the local turbulence
generated by the tower top. It consists of two spheres, each containing a flat
mirror. One moves about a vertical axis in the azimuthal direction and carries
with it the other, which seeks the Sun’s elevation (Fig. 3.15). Altazimuth
telescope mounts are more complicated than equatorial mounts in that both
Fig. 3.15. Turret of the Richard B. Dunn Telescope at Sacramento Peak, New
Mexico, USA (Dunn 1969). E.S. is the elevation sunseeker. Courtesy National Solar
Observatory, AURA, Inc.
3.2 High-Resolution Telescopes 89
We now discuss the solar telescopes which have a focal length f large enough
for high-resolution work in the primary focus, with presently available detec-
tors. The typical aperture, D, of a solar telescope is between 0.5 m and 1.5 m,
hence large f also means large focal ratio, N = f /D. The long-primary-f
telescopes are fixed in the observatory building; they are fed by steerable
mirrors as explained in the preceding section.
Let us, as a typical example, consider the 60-cm Vacuum Tower Telescope
at Kitt Peak. This telescope will be replaced in the near future by SOLIS,
a specialized tool for Synoptic Optical Long-term Investigations of the Sun,
including a 54 cm telescope with a vector-spectromagnetograph. Still, the
Vacuum Telescope at Kitt Peak is well-suited for demonstration because of its
clear-cut design and good documentation (Livingston et al. 1976 a); moreover,
since its construction in 1973 it has been of invaluable service to many solar
astronomers through its magnetograms, which were regularly recorded and
distributed, e.g., in the Solar Geophysical Data.
A north-south section across the building and telescope is shown in
Fig. 3.16. The tower is located at the southeast edge of Kitt Peak and has
a completely retractable wind screen, so that the coelostat is well exposed
to the wind. The tower platform is kept rather small by the use of a curved
track for the coelostat mirror (the “Zeiss arrangement”).
The light enters the vacuum tank through a 86 cm diameter, 10 cm thick
fused-quartz window. Inside the tank the beam is folded, first by the focusing
main mirror, and then by another flat mirror. The main mirror is spherical,
with f = 36 m, stopped to an aperture of 60 cm. The f ratio, N = 60, is
typical, cf. the solar telescopes listed in Table 3.1. The inclination of the
mirror is 1.25o .
Spherical Aberration. Thanks to the large focal ratio, spherical aberration
is negligible for this telescope. In principle, spherical aberration limits the
usable aperture D, because off-axis rays have a shorter focus than rays close
to the axis (the paraxial rays). Hence with increasing D the point spread
function widens.
In the following we shall apply Strehl’s criterion for the quality of the
image. This criterion demands that the central intensity of the point spread
function is at least 80% of the central intensity if there was pure diffraction
(i.e., that the Strehl ratio is at least 0.8). In the case of spherical aberration
this condition coincides with the Rayleigh λ/4-criterion, namely that the
90 3. Tools for Solar Observation
Fig. 3.16. The Kitt Peak 60-cm Vacuum Tower Telescope (Livingston et al. 1976 a).
Courtesy National Solar Observatory, AURA, Inc.
3.2 High-Resolution Telescopes 91
D ≤ 512λN 3 . (3.22)
For D = 60 cm, N = 60, and visible light, (3.22) is amply satisfied. For this
reason spherical main mirrors are commonly used for solar telescopes with
large N . These mirrors are easier to manufacture, and therefore cheaper, than
parabolic mirrors.
Coma. Also still tolerable is the coma. Increasing with angular distance from
the center of the field, the coma appears as a radially elongated asymmetric
point spread function, and hence limits the usable field of view. If we apply
again Strehl’s criterion (which for the coma is equivalent to wave-front aber-
rations of no more than 0.6λ) and use the results of optical aberration theory,
we find the angular radius of the usable field (in radians) to be
w0 = 19.2λN 2 /D , (3.23)
or about 3o for the Kitt Peak tower telescope at λ = 500 nm. Thus the
criterion is satisfied, even if we take the 1.25o inclination of the main mirror
into account, although the margin is not as wide as for the criterion (3.22)
for spherical aberration.
Astigmatism. The severest optical aberration of our example telescope is
astigmatism. Astigmatism could be avoided by the use of an off-center piece
cut from a parabolic mirror instead of an inclined sphere. With the inclined
sphere the telescope is a “schiefspiegler”; astigmatism occurs because even for
an axial incident beam the mirror appears no longer rotationally symmetric,
but rather as an ellipsoid. Rays reflected from the “major” (and “minor”)
axes (perpendicular and parallel to the direction of inclination) are focussed
in two elliptical patches slightly outside (inside) the mean focal plane. Like
the coma, astigmatism generally limits the usable field of view. For a spherical
mirror Strehl’s criterion (now equivalent to wave-front aberrations of no more
than 0.17λ) determines the angular radius of the usable field (in radians):
w0 = (2.7λN/D)1/2 . (3.24)
In our case w0 = 0.67o at λ = 500 nm, i.e., w0 is little more than half the
inclination of the main mirror!
At the Kitt Peak tower telescope the main mirror’s astigmatism is largely
compensated by cylindrical bending of the flat mirror in the vacuum tank
which folds the beam down towards the exit window. The bending amounts
to a 1.2λ deformation of this mirror.
92 3. Tools for Solar Observation
Table 3.1. Examples of high-resolution solar telescopes with large primary focal
length. Coel coelostat (V vertical, H horizontal), D aperture (L Lens, M mirror),
f focal length
Telescope Year Type D [cm] f [m] N = f /D
Mt. Wilson, 150 ft Tower 1911 Coel, V L 30 45 150
Einstein-Turm, Potsdam 1924 Coel, V L 60 14 23
Kitt Peak, McMath-Pierce 1962 Heliostat M 160 82 51
Meudon 1968 Coel, V M 60 45 75
Sac. Peak, R.B. Dunn Tel. 1969 Turret M 76 55 72
Crimea 1973 Coel, V M 90 50 56
Kitt Peak, Tower 1973 Coel, V M 60 36 60
Baikal, Russia 1978 Siderostat L 76 40 53
Sayan, USSR 1979 Coel, H M 80 20 25
German Tower, Tenerife 1987 Coel, V M 70 46 66
Swedish Tel., La Palma 2002 Turret L 97 20.8 21
(Fig. 3.16). The position of this image is tracked automatically. In this way
signals are generated which are used to correct the tilt (in two directions)
of the second flat coelostat mirror. Because thermal expansion can cause
the guide and telescope optics to drift apart, a laser system monitors and
compensates for departure.
Variants. A number of other high-resolution telescopes with long primary
focus are listed in Table 3.1. They are all fixed telescopes, but with varia-
tions. For example, the German Vacuum Tower Telescope on Tenerife, which
is very similar to the Kitt Peak vacuum tower, has a double building. An
inner tower carries the coelostat, the telescope, the guiding system, and the
spectrograph. An outer, completely separated tower houses ancillary equip-
ment, the elevator, etc.; its main purpose, however, is to protect the telescope
from being shaken by the wind, which is generally stronger on the Canary Is-
lands than at Kitt Peak. Another difference in detail is that the astigmatism
originating from the slight inclination of the main mirror is corrected by a
cylindrical bending of the main mirror itself.
Although Table 3.1 is by no means complete, it is representative in that
the vertical coelostat is the most popular configuration. There are essen-
tially two reasons for this. First, coelostats have no image rotation. Second,
the vertical arrangement is less vulnerable to the seeing conditions: it avoids
ground-level turbulence and also has the advantage that the light path is per-
pendicular to the stratification of air layers with different indices of refraction.
Problem 3.6. Write the field limitations (3.23) and (3.24) in seconds of arc
for λ = 500 nm and for D given in centimeters. Apply the result to the
telescopes of Table 3.1; consider in particular the 0.84o tilt of the main
mirror of the Vacuum Tower Telescope at Izaña.
94 3. Tools for Solar Observation
Fig. 3.20. The French-Italian solar telescope THEMIS at the Observatorio del
Teide, Izaña, Tenerife. 1 Entrance window, 2–3 Ritchey-Chrétien system, 4 exit
window, 5–7 secondary optics, 8–13 spectrograph with camera, 14 universal bire-
fringent filter
|Δ| ≤ 2N 2 λ . (3.25)
3.2 High-Resolution Telescopes 97
While this upper bound is rather generous for the telescopes with large N
described in the preceding section, we find (at λ = 500 nm) |Δ| ≤ 30 μm for
the example Gregory telescope!
A flat mirror behind the heat trap reflects the light into the declination
axis. The position of this mirror must be stable in spite of the near-by heat
concentration. A second flat mirror (the Coudé mirror) deflects the light
into the polar axis. These two mirrors must be carefully aligned in order
to avoid a circular wandering of the solar image. At zero declination the two
mirrors are oriented perpendicular to each other. In this case the instrumental
polarization is minimal.
Guiding of the Gregory Coudé Telescope is achieved with the help of
an auxiliary telescope, rigidly attached to the main tubus, which allows the
observation of the solar limb. According to the momentary limb position,
signals are generated which control the telescope drives at the declination
and polar axes.
Variants. Examples of other solar telescopes with short primary focal length
are given in Table 3.2. Figure 3.19 shows the domeless solar telescope at Hida,
Japan, designed by Zeiss with the aim to avoid dome-generated seeing. The
Hida telescope has a Gregory system similar to the one described above. The
mounting is altazimuthal. A 45o flat mirror, as in a Newton telescope, reflects
the light into the elevation axis, which at its end has the Gregory mirror.
THEMIS, the Télescope Héliographique pour l’Etude du Magnétisme et des
Instabilités Solaires, is a solar telescope especially designed for polarimetry
(Fig. 3.20). The polarization modulator (Sect. 3.5.4) is located on the main
Table 3.2. Examples of high-resolution solar telescopes with short primary focal
length. E equatorial, A altazimuth; D aperture (L Lens, M mirror); fp , feff primary
and effective focal length. The ATST is in the study phase
Telescope Year Mount D [cm] fp [m] feff [m]
Okayama Coudé 1968 E M 65 6.0 37
Pic du Midi 1972 E L 50 6.45 35
Big Bear 1973 E M 65 3.5 50
Domeless Coudé, Hida 1979 A M 60 3.15 32.2
Multi-Channel, Huairou 1984 E M 60 1.5 6/12
Gregory Coudé, Izaña 1985 E M 45 2.48 25
THEMIS 1996 A M 90 3.15 15
Dutch Open Telescope 1997 E M 45 2.0 44
SOLIS VSM a 2002 E M 54 0.8 3.3
GREGOR b 2005 A M 150 2.50 60
ATST c d 2008 A M 400 10–20 80–120
LEST design study e 1990 A M 240 5.5 16.8/163.4
a
The Vector-Spectromagnetograph (Keller 2001), b von der Lühe et al. (2001),
c
Beckers (1995), d Keil et al. (2001), e Engvold and Andersen (1990)
98 3. Tools for Solar Observation
optical axis, so that for this particular purpose the configuration is completely
axisymmetric and hence free of instrumental polarization. With an entrance
window of 1.1 m diameter and 6.6 cm thickness, THEMIS is not evacuated
but has a helium filling.
New Projects. The fine structure of the magnetic field at the solar surface
requires polarimetric investigations at smaller spatial scale than that resolved
by existing solar telescopes. Large telescopes, with apertures up to 4 m, have
therefore been proposed. A detailed study (Engvold and Andersen 1990) has
been made for LEST, a Large Earth-Based Solar Telescope. The aperture is
2.4 m and, according to (3.7), the diffraction limit is 0.05 at λ = 500 nm. It
is a modified Gregorian system, with a second and a third focus as listed in
Tab. 3.2, and with an adaptive optics system. The alignment requirements
are rather strict for this instrument. Coma and defocusing are the major
sources for optical image degradation. All mirrors are equipped with an ac-
tive alignment system. In addition, the two ellipsoidal mirrors are especially
shaped to correct the optical aberration of the primary. This can be demon-
strated by the calculation of spot diagrams. The spot diagram shows where
rays emanating at various angles from an object point hit the image plane.
As illustrated in Fig. 3.21, the coma at F 1 becomes intolerably large even as
little as 3 off the optical axis, but is substantially reduced in the secondary
and tertiary images.
The LEST study employs a helium-filled telescope tubus. Even the smaller
SOLIS telescope will have a helium filling, which permits an entrance win-
dow of only 6 mm thickness. The alternative is an open structure (Fig. 3.8):
Fig. 3.21. Spot diagrams calculated for the three foci of LEST (Andersen et al.
1984). Courtesy LEST Foundation
3.3 Spectrographs and Spectrometers 99
the German telescope GREGOR for the Observatorio del Teide, which will
replace the 45 cm evacuated Gregory telescope described in this section, and
the proposed ATST (Advanced Technology Solar Telescope) will be open so-
lar telescopes.
Problem 3.7. Calculate the field limitation due to coma and astigmatism for
the primary images of the telescopes of Table 3.2. What are the tolerances
for the primary and secondary foci?
Problem 3.8. Suppose there are dark and bright filaments with 20% inten-
sity contrast and about 0.25 apart in a sunspot penumbra. How much
of this contrast would be transmitted by a 2.4-m telescope under perfect
seeing conditions?
The solar telescope projects the image of the Sun into its focal plane. This
is also the location of the entrance slit of the spectrograph. A light cone of
width D/f = 1/N enters, is made parallel by the collimator, and falls onto
the grating. In order to make the best possible use of the available light and
of the optical elements, precisely the whole cone should be collimated. It
follows that the focal ratio of the collimator should be the same as that of
the telescope:
fS /DS = f /D . (3.26)
Most modern solar spectrographs have reflection gratings. The dispersed
reflected light is refocussed either by the collimator itself as in the Littrow
arrangement (Fig. 3.16), or by a separate camera as in the Czerny–Turner
system (Fig. 3.18). In the latter case the collimator and camera both operate
slightly off-axis and therefore introduce coma and astigmatism. In a symmet-
ric configuration the coma is cancelled; thus generally collimator and camera
have equal focal lengths, fS , the “focal length of the spectrograph”.
Dispersion and Spectral Resolution. Let us consider a reflection grat-
ing, and let a be the distance between two grooves (Fig. 3.22), the grating
constant. If α is the angle of incidence, and β(λ) the angle of reflection, then
the directions β of maximum intensity are given by the grating equation
mλ = a(sin α + sin β) , (3.27)
where m is an integer, the order of the spectrum. Differentiation of (3.27)
yields the angular dispersion
dβ m
= . (3.28)
dλ a cos β
100 3. Tools for Solar Observation
The linear dispersion in the spectral plane is dx/dλ = fS dβ/dλ; hence, with
β α,
dx a cos α
fS = . (3.29)
dλ m
Available gratings have, e.g., 632 grooves/mm1 , and are typically used in
the 5th order at α ≈ 60o . The required dispersion dx/dλ is 4 mm/Å if, for
example, we want to resolve 5 mÅ and have a detector in the spectral plane
with pixels of 20 μm. With these numbers (3.29) gives fS = 6.3 m. On the
other hand, DS is determined by the dimensions of the grating. A 15 cm ×
30 cm grating is common, and yields a projected aperture of 15 cm × 15 cm
(α = 60o ). The focal ratio of the spectrograph is thus fixed and, by (3.26),
the focal ratio of the example telescope is equally fixed:
N = 42 .
Indeed, the telescopes listed in Tables 3.1 and 3.2 all have (effective) focal
ratios of this order of magnitude. We now see that this is a consequence of the
available detectors and spectrograph gratings. Detectors with smaller pixels
permit smaller focal ratios, and hence more compact instruments.
Problem 3.9. Show that the dispersion of a grating depends only on λ and
the angle of incidence (use β α).
Problem 3.10. An aperture DS permits a diffraction-limited angular reso-
lution of order λ/DS , cf. (3.7). Using this, show that the resolving power
λ/Δλ = nm, where m is the order of the spectrum and n is the total
number of grooves in the grating. What is λ/Δλ for the above example?
Problem 3.11. The entrance slit has a finite width, and its image must be
folded with the spectrum. How narrow a slit must be used in order to
obtain the 5 mÅ resolution of the example considered in the text? If the
seeing is 1 and the telescope has a focal length of 30 m, by how much
should the entrance slit be opened so that no light is wasted? What is the
spectral resolution in this case?
Problem 3.12. The various orders m overlap each other in the spectrum.
Show that a wavelength interval which does not overlap with itself is lim-
ited by Δλ < λ/m.
The advantage of the overlapping spectra is that lines from different re-
gions of the solar spectrum can be recorded simultaneously in a single cam-
era. At the same time, of course, much unwanted light from various orders is
superimposed onto the desired spectrum. For this reason the echelle spectro-
graph has a predisperser. Working at low spectral resolution, the predisperser
images the spectrum either perpendicular or parallel to the main direction
of dispersion. In both cases the wanted and unwanted spectral parts can be
separated by masks or filters.
and recombiner are identical glass plates, with external reflection and trans-
mission coefficients re and τ , then a monochromatic input emerges with an
amplitude
A = rc re τ eiωt (e−ikx1 + e−ikx2 ) , (3.30)
where k = 2π/λ; x1 and x2 are the lengths of the two paths, measured
from some arbitrary common origin before the beamsplitter. The emergent
intensity is
η
I = AA∗ = [1 + cos k(x2 − x1 )] , (3.31)
2
where η = 4rc2 re2 τ 2 is the efficiency of the instrument.
In general the input to the spectrometer is not monochromatic, but has
an intensity distribution B(k). We use x = x2 − x1 for the path difference,
and denote the term that is independent of x by I0 ; then
∞
1
I(x) = I0 + ηB(k) cos kx dk . (3.32)
2
0
Problem 3.13. The Doppler Compensator. Show that a glass plate of thick-
ness d and refractive index n laterally translates a beam by an amount
⎛ ⎞
1 − sin α ⎠
2
Δx = d sin α ⎝1 − , (3.33)
n2 − sin2 α
where α is the angle between the plate normal and the incident light. This
device has been used to compensate for an unwanted Doppler shift of a
spectral line; at the same time, that shift is determined by a measurement
of α. Calculate the tilt of a 4-mm glass plate which compensates the
Doppler shift caused by v = 10 m/s. Take the Fe I line at 557.6 nm in a
spectrum of 10 mm/Å dispersion, and use n = 1.5 for the refractive index
of the glass.
Two examples of spectral line shifts which are of interest in solar physics
are the gravitational redshift, ΔλG = λGm /r c2 , and the Doppler shift,
ΔλD = λv/c. In particular the Doppler shift, which occurs whenever the dis-
tance between the source of radiation and the observer changes (with relative
velocity v), has been an important object of solar research since about 1960,
and special instruments have been designed for its measurement. These in-
struments all work with the solar spectrum, or at least with a small part of
104 3. Tools for Solar Observation
it, and hence belong to the section on spectroscopy; because of their common
purpose we treat them here in a separate subsection.
Of course, the whole profile of a spectral line may be recorded and subse-
quently be used to determine the exact position of the line. However, often
this position is the only information of interest. Its direct measurement can
save much labor in data collection and data analysis. Nevertheless we must
keep in mind that sometimes knowledge of the whole profile is indeed neces-
sary, especially when the line profile is not symmetric.
The Resonance-Scattering Spectrometer. A very accurate method to
measure line shifts is the resonant-scattering solar spectrometer. Figure 3.24
illustrates the principle (e.g., Brookes et al. 1978). The instrument works
with a selected single spectral line; the sodium line D1 at 589.6 nm and the
potassium line at 769.9 nm have mostly been used so far.
In order to improve the signal-to-noise ratio and to reduce unwanted
heat, a filter first blocks all the sunlight except a band of, say, 1 nm
around the line under investigation (Fig. 3.24 a). Next the light is lin-
early polarized, and in this form enters the electro-optic modulator. This
is a special crystal which is uniaxial in the absence of an electric field.
Mostly potassium-dihydrogen-phosphate (KH2 PO4 – “KDP”) or ammonium-
dihydrogen-phosphate (NH4 H2 PO4 – “ADP”) crystals are used. Brookes et
al. use a deuterated KDP. These crystals can be made biaxial by means of an
electric field applied parallel to the optical axis. If the electric field strength
is properly chosen (a voltage of several kV is typically required), the wave
polarized parallel to the slow axis is retarded by λ/4 so that the combined
light emerges in circular polarization. If the electric field is reversed the fast
and slow axes change their roles, and the circular polarization is of opposite
sense. Application of an AC voltage therefore yields alternating right-handed
and left-handed circular polarization.
The heart of the resonance spectrometer is the vapor cell. An external
magnetic field parallel to the incident beam splits the states of the atoms in
the cell so that unpolarized light would show the longitudinal Zeeman effect,
visible in the resonant-scattered light. For the circular polarized light, how-
ever, only one of the two shifted components (i.e., one of the σ components)
is present in the scattered light, because only the properly polarized atoms
respond. Now, the spectral lines of the metal vapor in the cell are much nar-
rower than the broad line originating in the (much hotter) solar atmosphere,
and the scattering cross sections – S(λ) in Fig. 3.24 b – can therefore be lo-
cated at the two sides of the solar line profile; it is only necessary to choose
the proper strength of the external magnetic field. If the solar line is shifted,
the intensity of the light scattered in the two components is not the same.
The difference can be determined by separating the photomultiplier signals
(currents or photon counts) measured during the two phases (left and right)
of the incident polarized light. For small shifts we may assume that the solar
line profile is linear; if, in addition, the line is symmetric, then the velocity is
Nl − Nr
v=k , (3.34)
Nl + Nr
where Nl and Nr are the two photomultiplier signals, and k is a constant.
The denominator in (3.34) corrects for intensity variations.
The resonance-scattering spectrometer illustrated in Fig. 3.24 employs a
magnetic field parallel to the incident beam. As an alternative, a perpendicular
field can be used. The scattered light must then be observed through right
and left circular analyzers (Grec et al. 1976).
As with the interferometer, no slit is necessary for the resonance-scattering
spectrometer. A large entrance beam is available and thus the signal-to-noise
ratio is high. Since, in contrast to the Doppler compensator, the instrument
works without an imaging system, it is particularly suited for the measure-
ment of Doppler shifts in the mean solar radiation.
Indeed, since 1976 G. R. Isaak and his collaborators have been able to
observe the spherically symmetric pulsations of the Sun (together with the
pulsations of low degree, cf. Chap. 5) with a resonance-scattering spectro-
meter – one of the most important discoveries in recent solar physics (Claverie
et al. 1979). Since then, several of these instruments have been installed at
various observatories around the world, and the SOHO satellite also carries
one with the GOLF experiment.
The wavelength measured in the resonance spectrometer is absolute, be-
cause of the direct comparison to atomic states. After correction for the
Earth’s orbital and rotational motions, absolute solar wavelengths can be de-
termined to an accuracy of about 1 mÅ. However, the amplitudes of the solar
pulsations can be determined without knowledge of the absolute line posi-
tion. Because they have specific frequencies, overtones with less than 1 cm/s
amplitude have now been identified – this corresponds to a Doppler shift of
less than 1 μÅ!
106 3. Tools for Solar Observation
Problem 3.14. Determine the constant k in (3.34) from the known line
profile.
Problem 3.15. The widths of the solar Na D 1 line at λ = 589.6 nm and of
the solar potassium line at λ = 769.9 nm are 0.05 nm and 0.02 nm, respec-
tively, at the steepest points of the profile. Calculate the magnetic field
strength required to place the Zeeman σ components at these points (for
the Zeeman splitting, see Sect. 3.5.1).
Broad-band filters, with a bandwidth of order 100 nm, essentially are colored
glass plates and have already been mentioned in Chap. 1 in the context of
the UBV system of stellar colors. We do not discuss these filters further
because they are quite generally used in astronomy. Neither shall we discuss
the so-called interference filters, which consist of a multitude of thin dielectric
layers with alternate high and low index of refraction, and have transmission
windows typically 0.5 to 20 nm wide (such filters are often used in order to
preselect a certain spectral band). Instead, we shall concentrate on the filters
with very narrow bandwidths, 0.1 nm and less, which are exclusively used for
solar observation because only here is sufficient light available.
A filter with a very narrow bandwidth was first described by Lyot (1933)
and, independently, by Öhman (1938). It is called, variously, the polarization-
interference monochromator, the birefringent filter, or just the Lyot filter, and
consists of an alternating sequence of polarizers and birefringent crystals, cf.
Fig. 3.25.
All the polarizers are mounted parallel to each other. The crystals are
uniaxial, cut parallel to their optical axis and positioned in such a way that
this axis makes an angle of 45o with the axis of the polarizers. The thickness
of the nth crystal plate is
en = 2n−1 e , (3.35)
where e is the thickness of the first plate (n = 1).
Let us first consider the effect of a single crystal and its two enclosing
polarizers. An incident wave, of amplitude A, becomes linear polarized, and in
the crystal can√be decomposed into two perpendicular components, each with
amplitude A/ 2. One has its electric vibration parallel to the optical axis;
this component propagates with velocity c/ne of the extraordinary ray. The
other vibrates in the perpendicular direction, and propagates with velocity
c/no of the ordinary ray. After traveling across the first plate of thickness
e there is a phase difference δ = 2πe(no − ne )/λ, where λ is the vacuum
wavelength. The difference J = no − ne is called the birefringence of the
material.
Of the two components the second polarizer transmits only the respective
parts that are parallel to its own direction; these parts each have an amplitude
A/2. Let φ be a common phase; the interference of the two waves leaving the
crystal then yields the amplitude
A A
cos(φ + δ) + cos φ = A cos(δ/2) cos(φ + δ/2) . (3.36)
2 2
That is, we have again a vibration with phase φ (and shift δ/2); but its am-
plitude A = A cos(δ/2) is now modulated in dependence of λ. The intensity
I = A2 cos2 (δ/2) , (3.37)
has maxima at δ = 2kπ, and is zero at δ = (2k − 1)π, where k is any integer.
The wavelengths of the maxima are λ = eJ/k; for large k, the distance of
one maximum to the next is eJ/k 2 .
Now take N plates, with thicknesses (3.35), between N + 1 polarizers.
After passage of the last polarizer the intensity is
Fig. 3.26. Intensity transmitted by the elements of a Lyot filter. After Lyot (1944)
Fig. 3.27. Upper part: transmitted intensity for 6 individual birefringent crystals,
each between two polarizers. Middle: combined transmittance for the filter elements
No. 6, 6+5, 6+5+4, . . . . Lower part: transmittance of the complete filter, for various
temperatures (in centigrade). For comparison a solar spectrum, ranging from about
Na D on the left to Hα on the right. After Lyot (1944)
Problem 3.16. The birefringences of quartz and calcite are 0.0092 and
−0.172, respectively. What is the required thickness of the thickest plate
if λ = 656.3 nm (Hα), and Δλ = 0.05 nm is desired? How many plates are
necessary if the nearest spectral windows of the filter should be no closer
than ±25 nm?
110 3. Tools for Solar Observation
A better way of tuning the Lyot filter is the addition of a quarterwave plate to
the birefringent crystal, together with the possibility to rotate the subsequent
polarizer. The quarterwave plate is itself a birefringent (uniaxial) crystal. As
the main crystal, it is cut parallel to its optical axis and mounted in such a
way that this axis is perpendicular to the main axis of the filter. The difference
is that the quarterwave plate has its optical axis at 45o relative to the optical
axis of the crystal, and that it retards the two rays relative to each other
exactly by λ/4. If we now rotate the exit polarizer relative to the entrance
polarizer (which remains fixed to the two crystals) by an angle α, then the
intensity of the emergent linear polarized light will depend on λ (because of
the λ-dependent retardation in the main crystal) and on α (because of the
rotation). Instead of (3.37) we obtain (Problem 3.18)
I = A2 cos2 (δ/2 + α) , (3.43)
δ/2 + α = kπ . (3.44)
Hence, a rotation by an angle α = π shifts an intensity maximum to the
same position where the adjacent maximum was at α = 0, and rotation
angles α < π suffice to cover the whole λ range between two maxima. For a
given wavelength λ, the required angle of rotation is therefore
πeJ
α=− mod(π) . (3.45)
λ
If only the thickest element of the Lyot filter is tunable then only the
narrowest maxima (line No. 6 of Fig. 3.26) can be shifted. The wavelength
range over which the filter can be used is then comparable to the width of
the filter window itself. Obviously, tuning over a wide range of wavelengths is
possible if the whole filter is subdivided into single tuning elements, so that
the maxima in each line of Fig. 3.26 can be shifted to any desired wavelength.
In such an assembly, a quarterwave plate must be added to each crystal, and
each exit polarizer (together with the following crystal and quarterwave plate)
must be made rotatable.
3.4 Filters and Monochromators 111
The problem is that simple quarterwave plates work for one specific wave-
length only. This is satisfactory if we are only interested in the narrow spectral
region around a single solar line, which can be covered by tuning of, say, the
two thickest filter elements.
A successful “universal birefringent filter” became possible only after the
development of achromatic quarterwave plates (e.g., Beckers 1971 b). These
plates are themselves combinations of different materials, e.g., quartz and
magnesium fluoride, which compensate their respective spectral variations
of the retardation. The compensation is achieved in a similar way as the
compensation of the chromatic aberration in an achromatic objective.
The spectral region over which the filter can be tuned is limited to the
range in which the quarterwave plates can be made achromatic. The Zeiss
universal filter has 9 tuning elements, is tunable within the range 450 nm to
700 nm, and has a transmission window 0.0095 nm to 0.025 nm wide, depend-
ing on wavelength. A complementary tuning element narrows the window by
another factor 1/2 (Beckers et al. 1975).
It is a complicated problem to find, for a given wavelength, the exact
rotation angles of all the tuning elements, and to actually rotate the ele-
ments into their precise positions. November and Stauffer (1984) describe a
computer-controlled fully tunable 9-element filter.
Problem 3.18. For a single tuning element write down the vibrations of
the electric vector upon entrance to, an emergence from, the birefringent
crystal and the quarterwave plate. Prove relation (3.43).
Fig. 3.28. Tunable Michelson interferometer. Left: Wide-field function, with two
glass blocks between the beam splitter and the mirrors of the two arms. Right:
Polarizing interferometer, with quarterwave plates placed before the mirrors, and
orthogonally polarized output beams. After Title and Ramsey (1980)
element (Sect. 3.4.2), this feature has been added to the Michelson inter-
ferometer as well: The input beam is polarized at 45o with respect to the
vertical, and the beam splitter is made polarizing by means of a multi-layer
coating in the splitting plane so that it transmits and reflects, respectively,
the components parallel and perpendicular to the plane of incidence. A quar-
terwave plate is placed in front of each mirror and, since it is traversed twice,
rotates the direction of polarization by 90o . As a result one has two orthogo-
nally polarized output beams, as shown in Fig. 3.28. Tuning is then possible
with an additional rotatable half-wave retarder. The temperature of the in-
terferometer is controlled by a thermostat.
The Michelson interferometer has periodic transmission maxima. The
maxima can be narrowed by a combination of two such interferometers, with
free spectral ranges in the ratio 1 : 2. A prefilter selects the spectral region
around a certain solar line; both GONG and MDI use the nickel line at
676.8 nm. The interferometer itself is tuned to several positions within the
profile of that line, so that its Doppler shift can be determined.
The interferometer that is named after its inventors (Perot and Fabry 1899)
has been used since about 1970 for narrow-band solar spectroscopy. The
core of this interferometer consists of two plates with parallel surfaces of
reflectance R, enclosing a layer of width d with a medium of refractive index
n (Fig. 3.29). The device can be used to define a wavelength standard and is
called etalon for this reason.
An incident beam is multiply reflected between the two parallel surfaces,
but with each reflection a fraction T of the intensity is transmitted, and
3.4 Filters and Monochromators 113
all these transmitted fractions interfere in the outgoing beam. If the angle of
incidence is θ, then the path difference between two successive beam fractions
is Δ = 2nd cos θ, and the phase difference is
δ = 2πΔ/λ = 4πnd cos θ/λ . (3.47)
For an incoming wave of form exp(iωt) the transmitted and reflected (ab-
solute) wave amplitudes are T 1/2 and R1/2 , respectively; thus, counting all
the reflections and the concomitant phase differences, the outgoing wave is a
geometric series,
Aeiωt = T eiωt + T Rei(ωt+δ) + T R2 ei(ωt+2δ) + . . . (3.48)
or
T
A = T (1 + Reiδ + R2 e2iδ + . . . ) = . (3.49)
1 − Reiδ
The intensity of the outgoing beam is, therefore,
T2 T2
I = AA∗ = = (3.50)
1 − 2R cos δ + R2 (1 − R)2 + 4R sin2 (δ/2)
or
1
I = Imax δ
, (3.51)
1+ 4R
(1−R)2 sin2 2
where Imax = T 2 /(1 − R)2 . This result was first obtained by G. B. Airy in
1831. For two values of R the transmitted intensity as a function of phase δ
is shown in Fig. 3.30. The transmission is periodic; the mth maximum is at
δ = 2mπ or, according to (3.47), at wavelength λ = 2nd cos θ/m. The free
spectral range, or distance between two successive maxima, is FSR = λ/m =
λ2 /2nd cos θ.
The transmission peak becomes narrower as R approaches 1. If the width
is small in comparison to the free spectral range, we may expand the sin(δ/2)
in (3.51) and obtain for the full width Δλ at half-maximum transmission
114 3. Tools for Solar Observation
Δλ = FSR/F , (3.52)
where
√
π R
F= (3.53)
1−R
is the finesse of the interferometer. The finesse is the ratio of the free spectral
range to the spectral resolution.
In the absence of absorption we have T = 1−R; in this case Imax = 1, i.e.,
the intensity at the transmission maxima is equal to the incident intensity,
independently of the value of R. If a fraction A of the intensity is absorbed
with each reflection, then T = 1 − R − A, and Imax = [1 − A/(1 − R)]2 .
The free spectral range of a Fabry–Perot interferometer is rather small.
For λ = 500 nm, d = 1 mm we have FSR ≈ 0.13 nm. Therefore, further fil-
ters are necessary for the selection of a particular transmission maximum.
Cavallini et al. (1987) combine a Fabry–Perot interferometer with a grating
spectrograph, Bendlin et al. (1992) describe a combination with a universal
birefringent filter. Because of the high maximum transmission of the Fabry–
Perot (in comparison to other filters or spectral devices) the combination of
two or three such interferometers is most advantageous. An assembly of three
Fabry–Perot interferometers for the solar observatory at Culgoora, Australia,
has been described by Ramsay et al. (1970). Since the transmission maxima
and free spectral ranges of the etalons are essentially determined by the plate
separations d, the ratios of these separations must be chosen such that the
unwanted transmission windows are closed. Darvann and Owner-Petersen
(1994) have studied this problem in detail. Prefilters must still be applied in
order to select a certain spectral band, but their pass bands may be wider,
with better transmission. Technically, the challenge is to control the paral-
lelism of the etalon plates and the plate-separation ratios with high precision.
Kentischer et al. (1998) have built a two-etalon instrument and describe its
upgrade to a triple system (which has been realized in 2001). By variation of
the plate separation their spectrometer can be used in the wavelength range
3.4 Filters and Monochromators 115
420–750 nm, with a spectral resolution of 250 000 or 150 000, for field-of-view
diameters 50 and 100 , respectively.
Fig. 3.32. Double monochromator. E entrance slit, M middle slit, A exit slit; S1 ,
S3 collimators, S2 , S4 camera mirrors; G1 , G2 gratings; P1 –P4 plane mirrors. After
Labs and Neckel (1962)
Problem 3.19. Calculate the width of the exit slit for a spectral band of 20 Å.
3.4 Filters and Monochromators 117
3.5 Polarimetry
The presence of a magnetic field can be deduced from the Zeeman splitting of
spectral lines. However, this is possible only for a sufficiently strong field. If
the field is weak, we must resort to the fact that the Zeeman components are
polarized. We have already seen how the resonance-scattering spectrometer
and the magneto-optic filter utilize this polarization.
In the present section we shall describe the splitting and polarization of
Zeeman components in the solar spectrum, and how the Zeeman effect can
be used to extract information about the magnetic field in the region on the
Sun where the spectral line has been formed.
We recall that in the case of a weak magnetic field (weak in the sense that LS
– or Russell–Saunders – coupling is the appropriate description) L, S, J, and
MJ are the quantum numbers that define the state of an atom. L characterizes
the total orbital angular momentum of the electrons (more precisely: of the
outer electrons which are treated explicitly), S is the analogue for the spin,
3.5 Polarimetry 119
1564.8 1565.0
λ [nm]
V/Ic [%]
These rules for the polarization of the normal Zeeman triplet are valid for
an absorption line originating in an optically thin layer. For a line in emission,
the sense of circular polarization is reversed, and (for the transverse Zeeman
effect) “perpendicular” and “parallel” must be exchanged. The reason is of
course that only emission lines have intensity by themselves, while what is
seen in an absorption line is merely the residual intensity.
The polarization rules apply also in the more general (anomalous) case
where S = 0 and where g ∗ depends on the quantum numbers: for ΔMJ = ±1
there is σ radiation, for ΔMJ = 0 we have π radiation.
When the observer is neither in the direction of B nor in a direction
perpendicular to B all the Zeeman components are seen. We shall treat this
case below; we shall then also drop the restriction that the spectral line
originates in an optically thin layer.
Solar spectral lines are Doppler and collision broadened (cf. Sect. 2.3.7).
This has the consequence that the magnetic field B must be of order 0.15 T
3.5 Polarimetry 121
Problem 3.20. Show that P ≤ 1 always holds. What is the state of po-
larization if only Q, or only U , or only V does not vanish? What is the
meaning of the sign of Q, U , and V ?
I = IC (1 − τ ) − IC τ (η + + η − )/2 , (3.61)
where
η ± = η(λ ± ΔλB ) , (3.62)
and η(λ) = κl (λ)/κC is the ratio of the line absorption coefficient κl (λ),
to the absorption coefficient of the continuum, κC , in the vicinity of the
3.5 Polarimetry 123
line. Since the Doppler and collision broadening is the same in the magnetic
and non-magnetic cases, we may use the same absorption profile η(λ). The
second term of (3.61) describes the absorption in the two σ components.
This absorption renders the emergent radiation circularly polarized, and may
therefore be described in terms of the Stokes parameter V . The polarization
is antisymmetric with respect to the line center, as expressed by the sign of
V (λ):
In general the magnetic field has an inclination, γ, with respect to the line
of sight, and an azimuth, φ, cf. Fig. 3.36. In this case all four Stokes profiles
will be affected. Let us once again consider first an optically thin absorption
layer, but let the incident light, of intensity I, be already polarized, with
Stokes parameters Q, U , and V . We define the Stokes vector
124 3. Tools for Solar Observation
⎛ ⎞
I
⎜ Q ⎟
I=⎜ ⎟
⎝ U ⎠ . (3.66)
V
Across a thin layer the total change, ΔI, ΔQ . . . , in each of the four param-
eters must be a linear combination of these variables themselves. Thus
ΔI = −τ I − τ ηI . (3.67)
The first term on the right of this equation is the continuous absorption,
the second describes the absorption due to the line. The elements of the line
absorption matrix η depend on the magnitude of the magnetic field via the
shifted profiles (3.62); in addition, they explicitly depend on the field angles
γ and φ. We shall not go through the the detailed derivation of η here, but
rely on the work of Unno (1956).
Unno follows the classical theory of H. A. Lorentz and treats the absorb-
ing electrons as linear oscillators. In this picture the atomic state with its
Zeeman-split energy levels is represented by a precession of the linear oscil-
lators around the magnetic lines of force. The frequency shift, as well as the
directional characteristics of the line absorption, is then a consequence of this
precession. Unno’s result is
⎛ ⎞
ηI ηQ ηU ηV
⎜ ηQ ηI 0 0 ⎟
η=⎜ ⎝ ηU 0
⎟ , (3.68)
ηI 0 ⎠
ηV 0 0 ηI
where
1 1
ηI = η sin2 γ + (η + + η − )(1 + cos2 γ) ,
2 4
1 1
ηQ = η − (η + + η − ) sin2 γ cos 2φ ,
2 4
(3.69)
1 1 + −
ηU = η − (η + η ) sin2 γ sin 2φ ,
2 4
1 +
ηV = (η − η − ) cos γ .
2
Although we have not deduced the absorption matrix, we can see a num-
ber of its properties. In the diagonal we have always ηI , i.e., the amount of
absorption proportional to the respective parameter itself is the same for all
four Stokes parameters. The reason is that they all are intensities and be-
have as such. Outside the diagonal all entries are zero, except for the first line
and column; that is, of the three polarization parameters Q, U , and V each
3.5 Polarimetry 125
changes only in proportion to itself and to the intensity, while the intensity
has changes in proportion to all four parameters. Additional entries to the
absorption matrix, due to magneto-optic effects, will be mentioned below.
For γ = 0 we recover the special case of the longitudinal Zeeman effect
(3.61) and (3.63), for γ = π/2 and φ = 0 we find the transverse effect repre-
sented by (3.64) and (3.65).
Transfer of Polarized Radiation. The solar atmosphere is not an opti-
cally thin layer. Moreover, there is not only absorption, but also emission
and scattering of radiation. The general theory of radiative transfer in an at-
mosphere permeated by a magnetic field is complicated; here we shall again
follow Unno and assume local thermodynamic equilibrium. For clarity, we
shall give an index λ to the Kirchhoff–Planck function B(T, λ) in order to
distinguish it from the magnetic field B.
We write the equation of transfer in vectorial form,
dI
cos θ = (1 + η)(I − B λ ) , (3.70)
dτ
where I is the four-dimensional Stokes vector (3.66), and B λ ≡ (Bλ , 0, 0, 0);
τ is the optical depth in the neighboring continuum, and θ is the angle to
the local vertical direction. The absorption on the right of (3.70) immediately
follows from (3.67) because the atmosphere can be envisaged as a sequence of
thin layers. The emission, on the other hand, follows from the assumption of
local thermodynamic equilibrium: in each of the four equations (3.70) there
must be a term formally corresponding to the I term, but with I replaced by
−Bλ . Of course, (3.70) is a generalization of the one-dimensional equation of
transfer (2.15), which we have already used in Sect. 2.3.2.
There is a number of special cases in which Unno’s equations (3.70) can
be solved analytically, or at least can be reduced to the problem of evaluating
integrals. One such case is the optically thin absorption layer, which yields
(3.67). Further cases have been considered by Mattig (1966). In all these cases
an atmospheric solar model, i.e., T (τ ) and therewith Bλ (τ ), is a necessary
ingredient. In addition, the absorption profile in the absence of a magnetic
field, η(λ), must be known, as described in Chap. 4.
Longitudinal Magnetic Field. In this case we have γ = 0; hence ηQ =
ηU = 0, and
1 + 1 +
ηI = (η + η − ) , ηV = (η − η − ) . (3.71)
2 2
The system (3.70) then becomes (with μ = cos θ)
126 3. Tools for Solar Observation
dI
μ = (1 + ηI )(I − Bλ ) + ηV V ,
dτ
dQ
μ = (1 + ηI )Q ,
dτ
(3.72)
dU
μ = (1 + ηI )U ,
dτ
dV
μ = (1 + ηI )V + ηV (I − Bλ ) .
dτ
We may set Q = U = 0; if there is linear polarization at large optical
depth, we easily find
⎛ ∞ ⎞
1 + ηI
Q(0, μ) = Q(∞, μ) exp ⎝− dτ ⎠ (3.73)
μ
0
and an analogue expression for U . In any case, the interesting variables are
I and V . We define X(τ, μ) = I + V and Y (τ, μ) = I − V , and obtain
dX
μ = (1 + η + )(X − Bλ ) ,
dτ
(3.74)
dY
μ = (1 + η − )(Y − Bλ ) .
dτ
Each of these two equations is exactly of the type described in Chap. 4 in the
context of radiative transfer in a spectral line. We may therefore immediately
take the solution (4.7) derived there:
where I0 (λ) is the line profile in the absence of the magnetic field. Thus,
according to the definition of X and Y , we find
1
I= [I0 (λ + ΔλB ) + I0 (λ − ΔλB )] , (3.77)
2
1
V = [I0 (λ + ΔλB ) − I0 (λ − ΔλB )] . (3.78)
2
Figure 3.37 shows I and V profiles, together with I0 (λ), for a typical value of
the ratio vB = ΔλB /ΔλD of the Zeeman displacement to the Doppler width
of the line.
Transverse Magnetic Field. Here we have γ = 90o . Let us additionally
assume φ = 0. This corresponds to a special choice of the system of coordi-
nates which describe the transverse field components; this choice is possible
if the azimuth φ does not vary with optical depth. The absorption coefficients
are now ηU = ηV = 0, and
1 1
ηI = η + (η + + η − ) ,
2 4
1 1 (3.79)
ηQ = η − (η + + η − ) .
2 4
The equations for I and Q are treated in the same manner as the equations
for I and V in the previous case. Introducing X(τ, μ) = I + Q and Y (τ, μ) =
I − Q we find that X(0, μ) is identical to the original line profile, I0 , and that
⎛ τ ⎞
∞
2 + η+ + η− 2 + η +
+ η −
Y (0, μ) = Bλ exp ⎝− dτ ⎠ dτ . (3.81)
2μ 2μ
0 0
1 ∂2η 2
ηQ = − v sin2 γ cos 2φ 4
+O(vB ),
4 ∂v 2 B
(3.83)
1 ∂2η 2
ηU = − v sin2 γ sin 2φ 4
+O(vB ) ,
4 ∂v 2 B
∂η 3
ηV = vB cos γ +O(vB ),
∂v
and the expansion
2
I = I0 + vB I2 + . . . ,
2
Q= vB Q2 + . . . ,
(3.84)
2
U= vB U2 + . . . ,
V = v B V1 + . . . .
The transfer equations, to the lowest (zero) order, give
3.5 Polarimetry 129
dI0
μ = (1 + η)(I0 − Bλ ) ; (3.85)
dτ
that is, I0 is our original line profile.
To the first order we obtain the circular polarization:
dV1 ∂η
μ = (1 + η)V1 + cos γ(I0 − Bλ ) , (3.86)
dτ ∂v
which can be solved by an integral of the form which we have had before.
Without writing down this solution we note that for a weak magnetic field
the leading term of the Stokes parameter V is proportional to B cos γ, i.e.,
to the longitudinal (line-of-sight) component of the (possibly inclined) field
vector.
The second order yields the linear polarization:
dQ2 1 ∂2η
μ = (1 + η)Q2 − sin2 γ cos 2φ(I0 − Bλ ) ,
dτ 4 ∂v 2
(3.87)
dU2 1 ∂2η
μ = (1 + η)U2 − sin2 γ sin 2φ(I0 − Bλ ) .
dτ 4 ∂v 2
Again the solutions can be written in form of our now well-known integral. We
see that, to leading order, the Stokes parameters Q and U are proportional
to B 2 sin2 γ, i.e., to the square of the transverse component of the (possibly
inclined) magnetic field.
Example calculations of Jefferies et al. (1989) show that the weak-field
approximation yields reasonable results if vB ≤ 0.5.
Magneto-Optic Terms. There are several effects that cause further terms
in the matrix η, Eq. (3.68). The coupling between the Stokes parameters aris-
ing from those terms must be taken into account if high accuracy is desired.
One of those effects is the anomalous dispersion of an electromagnetic wave
propagating through a magnetized plasma, in the present context first treated
by Rachkovsky (1962 a); in a longitudinal field, for example, there is Faraday
rotation of the direction of linear polarization. A second effect is radiative
scattering: for scattered light the state of polarization generally depends on
the angles of the incident and scattered rays relative to the direction of B.
General Solution. For an arbitrary depth-dependence of the absorption
matrix Unno’s equations, supplemented by the magneto-optic terms, must
be integrated numerically. Diverse algorithms to do this have been described;
for the interpretation of observed Stokes parameters, however, an inverse
problem must be solved: The profiles I(λ), V (λ) . . . are given, and the at-
mospheric structure and motion, the magnetic field, etc., are wanted. One
way is to guess the magnitude and direction of the magnetic field in the solar
atmosphere, to calculate the elements of the absorption matrix, and then to
solve (3.70). If the emergent Stokes profiles agree with the observed profiles
we may say that the field has been determined, otherwise we must iterate.
130 3. Tools for Solar Observation
In this context the response function is a useful concept. For any physical
variable x, say the atmospheric temperature, the magnetic field strength, or
a velocity component, the response function Rx is an integral kernel that
determines how the Stokes profiles respond to a change of x, viz.
∞
δI(λ) = Rx (λ, τ )δx (τ )dτ . (3.88)
0
Real Stokes profiles show a great variety of forms, and often differ sub-
stantially from the schematic cases of Fig. 3.37. For example, the two wings
of V (λ) may be asymmetric, in amplitude as well as in area (Fig. 3.38, first
row). These asymmetries, δa and δA, are usually defined as the difference in
amplitude and area, respectively, between the blue and red wings, divided by
the sum. Another quantity of interest is the wavelength at which V (λ) = 0
between the two wings. Interpreted in terms of the Doppler effect this zero-
crossing velocity may amount to several km/s, cf. the examples in the second
row of the figure. If the resolution element of the observed area contains an
unresolved inhomogeneous magnetic field, e.g., a tube of magnetic flux in a
field-free environment, then the zero-crossing velocity refers to the magnetic
part alone. The third row of Fig. 3.38 shows more irregular V profiles; some
resemble the theoretical profiles of Q and U , possibly due to a superposition
of V profiles from both magnetic polarities, Doppler-shifted relative to each
other.
If the circular polarization V (λ) is integrated over a spectral band in-
cluding lines with asymmetric V profiles, then a non-zero net polarization is
obtained. Illing et al. (1975) have measured such broad-band polarization, and
first suggested a model containing layers of different velocity and magnetic
field strength as an explanation.
For the numerical treatment of Unno’s equations it makes no difference
whether or not the field magnitude B and the angles γ and φ are functions
of depth. The inversion method could be extended even to the case of an
atmospheric and magnetic structure that is horizontally inhomogeneous. This
would require to solve the transfer equations in three dimensions, instead
in terms of optical depth τ only. A great difficulty, however, is the limited
resolution of the polarimetric observations, which often precludes the retrieval
of detailed information from the inversion. We must keep in mind that a
certain amount of circular polarization can be generated in two different
ways: either by a weak uniform field, or by a stronger field that occupies
only a fraction of the resolved area at the solar surface. We postpone the
discussion on how to discriminate between these two cases to Chap. 8.
An example of Stokes parameters computed for a transverse field and
depth-dependent field azimuth φ is shown in Fig. 3.37. In Fig. 3.39 the max-
imum of the V profile is shown as a function of B and cos γ. An atmosphere
T (τ ) appropriate to a sunspot umbra (Hénoux 1969) has been employed in
3.5 Polarimetry 131
Fig. 3.38. Circular polarization V (λ) of two Fe I lines (at 630.15 nm and
630.25 nm), in % of the intensity IC of the continuum near the lines. The upper row
shows large amplitude and area asymmetries, the middle row large shifts vZC of the
zero-crossing velocity, and the lower row shows “irregular” profiles. Wavelength λ is
in nm; solid vertical lines mark the rest positions of the two iron lines, dashed lines
mark the cores of the corresponding I(λ) profiles. From measurements of small-scale
magnetic flux concentrations in the photosphere, made with the Advanced Stokes
Polarimeter at Sacramento Peak Observatory. Adapted from Sigwarth (1999)
line profile to some specified form – the output signal is directly converted
to a magnetic field strength. Magnetograms have been obtained regularly
since 1974 with the V -polarimeter installed at the Kitt Peak Vacuum Tower
Telescope, which was taken as an example telescope in the present text (Liv-
ingston et al. 1976 a, b, Jones et al. 1992). This magnetograph employs a
Kerr cell in combination with a λ/4 plate instead of an electro-optical crys-
tal, but otherwise very much follows the principles described above. Another
alternative of modulating the polarization is the rotating λ/4 plate first used
by Kiepenheuer (1953). The modulation must be either sufficiently fast so
that the two intensities that are subtracted from each other are measured
under the same atmospheric seeing distortion, or the measurement must be
strictly simultaneous by means of a beam-splitting device. If the full Stokes
vector is desired, the diverse states of polarization must all be encoded sepa-
rately by appropriate phase retarders within each modulation cycle. Another
important ingredient is a polarizer that can be mounted in front of the po-
larimeter; by introducing a well-defined amount and kind of polarization the
whole apparatus can be calibrated.
Polarimeters belong to the standard equipment of most solar observato-
ries. The review of Beckers (1971 a) lists 22 magnetographs, half of which
are vector magnetographs. Examples of more recent instruments are the Ad-
vanced Stokes Polarimeter (ASP; Elmore et al. 1992) at Sacramento Peak
Observatory, as well as the La Palma Stokes Polarimeter (LPSP) and the
134 3. Tools for Solar Observation
Fig. 3.41. Stokes I and V spectra obtained with a Fourier transform spectrometer.
From Stenflo et al. (1984)
Let us use the result (3.78) for an estimate of the expected circular po-
larization for a given (weak) longitudinal magnetic field (following Schröter
1973). For small vB (3.78) yields V = vB dI/dv = −vB IC dr/dv, where IC is
the continuum intensity, and r = 1 − I/IC is the relative line depression. Let
us assume a Gaussian profile with central depression r0 , i.e., r = r0 exp(−v 2 ).
And let us place the two windows at the steepest part of this profile, that is
at v = ±2−1/2 . At these positions we have r = 0.607r0 and dr/dv = ±0.86r0 .
3.5 Polarimetry 135
The difference of the two V signals, divided by the sum of the two I signals,
yields the degree of circular polarization:
V 0.86r0
= vB . (3.89)
I 1 − 0.607r0
A frequently used line for polarimetry is the Fe I line at λ = 525.02 nm
(although this line poses problems because of its temperature sensitivity),
with g ∗ = 3, ΔλD = 42 mÅ, and r0 ≈ 0.7. We substitute (3.55) for ΔλB and
obtain
V
≈ 9.6 × 10−4 B , (3.90)
I
where B must be measured in gauss. We see that a field of 10 G still produces
about 1% circular polarization, which can easily be measured with a good
polarimeter.
Since we have already seen that for a weak magnetic field V ∝ B cos γ we
may generalize (3.90) for an inclined weak field to V /I ≈ 9.6 × 10−4 B cos γ.
A rough estimate how much linear polarization we may expect in solar
spectral lines can be obtained as follows. We set again φ = 0, and compare
(3.87) to (3.86). In combination with (3.84) we see that we obtain Q instead
of V if we make the replacement
∂η 1 2 ∂2η
vB cos γ → − vB sin2 γ .
∂v 4 ∂v 2
If, in addition, we take the line profile r(v) in place of the absorption profile
[which is permitted only for the optically thin layer, cf. (3.67)], then
Q 1 IC d2 r 2
v sin2 γ .
I 4 I dv 2 B
Let us use the Gaussian profile as above. The maximum of |d2 r/dv 2 | occurs
at v = 0, and has the value 2r0 . At v = 0 we have IC /I = (1 − r0 )−1 , so that
the degree of linear polarization becomes
Q r0
v 2 sin2 γ . (3.91)
I 2(1 − r0 ) B
For the Fe I λ 525.02 nm line this finally gives (with B in gauss)
Q
≈ 10−6 B 2 sin2 γ . (3.92)
I
This rough estimate indicates that linear polarization originating from a
weak magnetic field is very difficult to measure. A transverse field with a
strength of 100 G yields ca. 1% linear polarization; by comparison, the same
amount of circular polarization is obtained from a longitudinal field of 10 G,
according to the estimate (3.90).
136 3. Tools for Solar Observation
Polarization does not necessarily require a magnetic field in the region where
the light originates. Polarized radiation is also produced by scattering, as is
well-known from the blue light of the sky. A necessary condition for scat-
tering polarization is the anisotropic illumination of the scattering particles.
Thomson scattering by free electrons in the corona is the most prominent so-
lar example (Sect. 9.1.3). But also at lower levels in the atmosphere one can
observe the effect near the limb. The polarization is linear; in the continuum
it is due to Rayleigh scattering by neutral hydrogen atoms and to Thomson
scattering by electrons; in spectral lines it is due to coherent scattering in
bound-bound transitions of atoms.
Fig. 3.42. Spectral regions around three Mg I lines, from records made near the
northern solar limb. Notice especially the lines of MgH that are rather weak in the
intensity spectrum (top), but very conspicuous in the linear polarization Q/I (lower
spectrum). From Stenflo et al. (2000 b)
flo and Keller 1997); Gandorfer (2000) has published a spectral atlas that
displays Q/I for the wavelength range 462.5 nm ≤ λ ≤ 699.5 nm. The differ-
ences to the intensity spectrum are so remarkable that one may speak of the
“second solar spectrum” (Fig. 3.42), with molecular lines, e.g., from C2 and
MgH, and other unusual features (Stenflo et al. 2000 ab).
The Hanle effect causes a decrease of the scattering polarization and a ro-
tation of the plane of polarization, and occurs when a magnetic field is present
in the observed part of the atmosphere. This effect was first investigated in
the laboratory by Hanle (1924) in the context of resonant fluorescence. One
can understand the Hanle effect if one realizes that the anisotropic radiation
field leads to the polarized scattered light by means of polarizing the states of
the scattering atoms. An atomic state is polarized if there is a non-equilibrium
population of its magnetic sublevels with well-defined phase relations between
these sublevels (quantum interference). In general, both the upper and the
lower state will be polarized. A magnetic field perturbs the polarization of the
atomic states (in the classical picture, it causes a precession of the oscillating
dipoles). Even a quite weak magnetic field may lead to depolarization via the
Hanle effect, and there is no cancellation of the contributions of unresolved
magnetic dipoles such as occurs with the circular polarization caused by the
Zeeman effect. Thus, the Hanle effect bears new possibilities to investigate
the solar magnetic field (e.g., Stenflo 1982; Landolfi and Landi degl’Innocenti
1986; Bianda et al. 1998, 1999).
The active cavity radiometer operates in such a way that a constant tem-
perature difference of about 1 K is maintained between the cavity and the heat
sink. This is achieved by a servo loop which uses the signal of the temperature
sensor. Thus the heating is automatically increased when the shutter closes
the cavity. Let Pr and Po be the electrical power in the reference (closed) and
observational (open) phases. The incident radiation flux density, S, is then
obtained from
SAc (αc + ρρc ) = Pr − Po + C . (3.93)
Here Ac is the area of the aperture, αc and ρc are the absorptance and re-
flectance of the cavity for solar radiation, and ρ is the fraction of the radiation
not initially absorbed which is reflected back into the cavity. The additional
term C comprises several small corrections, e.g., heat conducted to, or away
from, the cone by air or by the electrical lead, or radiation from the cone to
the surroundings.
The specular black coating in the cavity has an absorptance of αs =
0.9 ± 5%. As the cone angle is 30o , there are six internal reflections of an
axial ray before reflection out of the aperture. Theoretically, the absorptance
of the cavity is therefore
αc = 1 − (1 − αs )6
or 0.999 999 with an uncertainty of 0.0003%. In practice, an absorptance of
0.99943 ± 0.00002 has been measured. Hence, the term ρρc is also a small
correction.
Because of a possible temperature drift a second identical cavity and
thermal impedance is connected to the heat sink. Its temperature is allowed
to drift passively. This cavity always views the heat sink, just as the primary
cavity in the shutter-closed phase. It can therefore be used to eliminate the
effects of the heat sink temperature drift.
The ACRIM flown on the Solar Maximum Mission consisted of three
identical pyrheliometers of the kind just described. They were operated with
different shutter frequencies, and thus the effect of cavity surface degradation
could be controlled to some extent. The instrument measures the absolute
irradiance with an uncertainty of order 0.2%, but relative variations much
smaller than this have been recorded.
3.6 Special-Purpose Instruments 139
The Earth is essentially transparent for neutrinos, and the same is of course
true for all terrestrial materials from which neutrino detectors could be built.
Fortunately, the flux of neutrinos from the Sun is intense. Two neutrinos are
produced per synthesized α particle, i.e., for roughly every 25 MeV generated
in the Sun. At the Earth we thus have approximately 6.5×1014 neutrinos/m2 s,
cf. Table 2.5. Also fortunately, there are a few nuclei which have a sufficiently
large cross section to capture one or the other neutrino out of the solar flux;
the yield is about one neutrino in every 1020 . Among these nuclei are the 37 Cl
and the 71 Ga isotopes. Although other target atoms have been investigated,
these two are the most important ones used so far for real solar neutrino
detectors. A 37 Cl experiment has been in operation since 1968, and results
from two gallium experiments have been obtained since 1991. In addition to
the neutrino capture by nuclei, neutrino-electron scattering has been used
successfully in two water Čerenkov detectors since 1987.
All solar neutrino experiments are located in underground laboratories.
This is necessary because muons, produced in the Earth’ atmosphere by
cosmic rays, would cause unwanted signals in the detector and therefore must
be shielded off. Only at a depth of 1 to 2 km, equivalent to 3 to 6 km water,
the muon flux is sufficiently attenuated. The results of all solar neutrino
experiments have been discussed in Sect. 2.4.2.
37
The Cl Experiment. The basic reaction of this experiment is
decay
ν + 37 Cl ⇐⇒
capture
37
Ar + e− . (3.94)
The energy threshold for this reaction is 814 keV. Above this energy the cross
section of 37 Cl for neutrino absorption increases by several orders of magni-
tude, cf. Fig. 3.44, or the values listed in Table 2.5 (for the continuous solar
neutrino spectra the cross sections in this table are averages over energy).
Fig. 3.44. Cross section for neutrino capture. Left: 37 Cl, after Kitchin (1984); right:
71
Ga, after Hampel (1986)
140 3. Tools for Solar Observation
The reason is that the 37 Ar nucleus has a large number of excited states into
which the target atom can go if the neutrino energy is sufficiently high.
The 37 Cl experiment, carried out by R. Davis, measured the flux of solar
neutrinos for the first time in 1968 (e.g., Bahcall and Davis 1976). The target
consists of ca. 615 t of tetrachloroethene, C2 Cl4 , which is an inexpensive
cleaning fluid. The fluid is contained in a 400 m3 tank, 1500 m underground
in the Homestake gold mine near Lead, South Dakota, USA.
A typical run goes as follows. The tank is left alone for about 100 days.
As the half-life of 37 Ar is 35 days, the number of 37 Ar atoms has then almost
reached the saturation level. The argon atoms can be assumed to be freely
dissolved in the liquid, because the neutrino provides ample energy to break
up the original C2 Cl4 molecule. Helium is then pumped through the tank;
argon, a noble gas like helium, is carried along, and subsequently collected by
a charcoal trap at liquid nitrogen temperature. As a check a small amount of
a different Ar isotope, say 3×1019 atoms of 36 Ar, is added to the tank prior to
the helium pumping. Since the charcoal trap recovers the majority of these
“carrier atoms” (≥ 90%), and since all isotopes of an element chemically
behave in the same way, it is reasonable to conclude that all the 37 Ar atoms
have been recovered as well.
The argon is placed into a miniature proportional counter where the decay
is monitored. The 37 Ar nucleus captures an inner electron, – the right-to-left
reaction of (3.94). At the same time a 2.8 keV electron is emitted; by this
signature the 37 Ar decay is distinguished from other events in the counter.
Besides the test with the 36 Ar carrier gas, two other tests confirm that
37
Ar is being recovered efficiently. One is to irradiate the tank with a cali-
brated neutron source, the other to directly inject 500 atoms of 37 Ar. In both
cases the recovery is satisfactory.
The 71 Ga Experiment. If the neutrino energy exceeds the threshold of
233 keV, the neutrino can be captured according to
decay
ν + 71 Ga ⇐⇒
capture
71
Ge + e− . (3.95)
As Fig. 3.44 illustrates, the cross section for this reaction also increases with
increasing neutrino energy. But because of the relatively low energy threshold
the main solar contribution comes from the pp reaction, cf. Table 2.5, where
again energy-averaged cross sections are listed for the various solar spectra.
The isotope 71 Ga has a half-life of 11.4 days. The gallium target is therefore
exposed for about 3 weeks; within this period the number of 71 Ge atoms
reaches ca. 70 % of the saturation level.
One of the gallium experiments, the Russian–American SAGE in the Bak-
san underground laboratory in the Caucasus, uses 57 t of metallic liquid Ga
(gallium has a melting temperature of 29.8o ). The other experiment, the col-
laboration GALLEX of groups in France, Germany, Israel, and Italy, is per-
formed in the Gran Sasso underground laboratory in Italy; its target is a tank
with 101 t of GaCl3 /HCl solution, containing 30.3 t of gallium. As in the 37 Cl
3.6 Special-Purpose Instruments 141
experiment, the 71 Ge atoms are extracted from the tank. Subsequently their
decay according to (3.95) is measured in a miniature proportional counter.
Again the typical signature of the 71 Ge decay in the counter is used to dis-
criminate other events.
In order to ensure that the extraction is quantitative, a stable Ge isotope
is added as a “carrier” to each experimental run of GALLEX. As a further
test a known quantity of 71 As has been introduced into the tank; this arsenic
isotope decays with a half-life of 2.9 days into 71 Ge, which has been fully
recovered. Both gallium detectors have been tested successfully by calibrated
artificial 51 Cr neutrino sources, consisting of chromium irradiated with neu-
trons (Fig. 3.45). The isotope 51 Cr decays by electron capture and thereby
emits neutrinos of 430 keV and 750 keV, which is in the low-energy part of
the solar neutrino spectrum.
νx + e− → e− + νx . (3.96)
This reaction occurs with all three neutrino flavors, although with greatly
reduced sensitivity to νμ and ντ . The scattered electron emits Čerenkov light
if its velocity v exceeds the velocity of light in water, c/n, where n = 1.34 is
the refraction index of water. The light is emitted along a circular cone around
the path of the electron. The half-angle θ of the cone is given by cos θ = c/nv
and is 41.9o in the relativistic limit v c. Photomultiplier tubes at the walls
of the detector monitor the Čerenkov light so that the trajectory and the
timing of the electron can be determined. This allows to infer the energy and
the direction of the incoming neutrino. Thus the water Čerenkov detector is
a neutrino telescope, in contrast to a radiochemical experiment that has no
directional information.
Solar neutrinos have been measured with two water Čerenkov detec-
tors, Kamiokande since 1987 and Super-Kamiokande since 1996, both in the
Mozumi mine near Kamioka, Japan. The detectors consist of cylindrical tanks
filled with 4500 t and 50 000 t water, respectively, and a large number of pho-
tomultiplier tubes mounted at the walls, over 11 000 in the case of Super-
Kamiokande. The detector volume is divided into an inner part, in which
the neutrino-electron scattering is observed, and an outer (“anti-detector”)
part to discriminate events originating from outside. The time resolution is
a few ns; the angular resolution for neutrinos of the solar energy range is
ca. 28o .
The energy threshold is ≈ 7.5 MeV for Kamiokande, and ≈ 5 MeV for
Super-Kamiokande. For smaller neutrino energy the measurement is spoiled
by events arising from radioactive impurities dissolved in the water. Hence
the two water detectors are sensitive only to the 8 B neutrinos from the solar
reaction chain ppIII.
Kamiokande was originally built for the search of nucleon decay. Such has
not been discovered, but by improving the water purity the energy threshold
was lowered so that it could be used as a detector for solar neutrinos. The
name may stand for both endeavors: Kamioka Nucleon Decay Experiment,
or Kamioka Neutrino Detection Experiment.
The Sudbury Neutrino Observatory. A Čerenkov detector using heavy
water instead of ordinary water has been built 2000 m underground in a
nickel mine in Sudbury, Ontario, Canada. The detector is a spherical acrylic
(transparent) vessel of 12 m diameter, containing 1000 t of D2 O. Surrounding
this vessel is an outer container filled with 7000 t of H2 O that acts as shielding.
The walls of this outer container hold the 9500 photomultiplier tubes for the
detection of the reaction products.
In addition to neutrino-electron scattering, the following two deuteron
reactions can be monitored in the heavy-water tank:
3.6 Special-Purpose Instruments 143
νe + d → p + p + e− , (3.97)
νx + d → p + n + νx . (3.98)
In the visible part of the spectrum the solar corona radiates 105 to 106 times
less intensely than the solar disc. Under normal circumstances this faint radi-
ation is buried below stray light (originating from the disc) which is generated
mainly in the Earth’s atmosphere, but also in the telescope.
We may observe the corona at the occasion of a solar eclipse, because then
the bright solar disc is occulted by the moon outside the atmosphere and so
both the atmospheric and the telescopic contributions to the stray light are
largely eliminated.
But even without an eclipse it is possible to see the solar corona under
special conditions: the observatory site must have a sky as dark as possible,
i.e., as little atmospheric scattered light as possible, (high mountain sites
often are “coronal” in this sense), and a specialized telescope must be used
– the coronagraph invented by B. Lyot around 1930.
The essence of the coronagraph is to produce an artificial eclipse by means
of an occulting disc inside the telescope, and to take special precautions to re-
duce the instrumental scattered (and diffracted) light. Figure 3.46 illustrates
the principle of Lyot’s original coronagraph.
144 3. Tools for Solar Observation
The light enters the telescope through an elongated tube (H) which pro-
tects the objective (A) from dust. The objective itself is made of a single lens
so that the number of scattering surfaces is minimized. The primary image
is formed at B. From here the central part (the solar disc) is reflected at
an inclined surface (J) through a window (K) out of the telescope – reflec-
tions from the window itself leave through a second window (K ). Behind the
occulting surface (J) there is a field lens (C) which images the objective at
A A on a diaphragm D. Light scattered at the objective and diffracted at
its edges is blocked in this way. The center of the diaphragm is occupied by a
small screen (E) which stops the solar image formed by secondary reflections
at the surfaces of the objective. The objective lens itself must be cleaned
frequently, and for that purpose can be taken out together with a side cover
(L), to which it is mounted. The setting (I) of this lens has a concave surface,
so that all light falling on it is also reflected out of the tube. The final image
of the corona is produced at B B by an achromatic objective (F).
Many variants of Lyot’s coronagraph have been built. Of course, it is most
advantageous not only to reduce the stray light in the instrument, but also
to eliminate the atmospheric stray light. Coronagraphs have therefore been
flown on space missions. Numerous coronal photographs have been taken with
an instrument flown with Skylab (May 1973 to January 1974). The Large-
Angle Spectroscopic Coronagraph (LASCO) on the SOHO satellite consists
of three instruments with occulting discs of different size so that the corona
can be imaged at various distances, namely 1.1–3, 1.5–6, and 3.7–30 solar
radii, respectively (Brueckner et al. 1995). The two instruments for the outer
corona have an external occulting disc in front of the entrance aperture,
similar to the Skylab coronagraph; the one for the inner corona is internally
occulted as in Fig 3.46. Instead of an occulting disc an (external and/or
internal) occulting edge may be used for the investigation of a segment of the
corona. The UV Coronagraph Spectrometer (UVCS) on board of SOHO is
of this type (Kohl et al. 1995).
3.7 Bibliographical Notes 145
ent multilayer coatings to take images in two narrow spectral bands on either
side of a coronal emission line; a Doppler shift of the line thus manifests itself
in different intensities of the two images.
I have put no emphasis on light detectors, except in the context of the
scale of solar images. General texts on astronomical instruments, e.g., Kitchin
(1984), or the monographs of Holst (1998) and Howell (2000) treat this sub-
ject.
Diffraction gratings are extensively treated by Stroke (1967). An intro-
ductory monograph on spectroscopy has been written by Kitchin (1995).
Howard (1974) describes a variant of the Doppler compensator: instead of
tilting a glass plate, the pair of exit slits is moved until each of the two slits
gets an equal amount of light; the velocity signal is then taken from the shaft
which drives the slit motion.
The theory of birefringent filters is outlined in the classical papers of Lyot
(1944) and Evans (1949). Harvey et al. (1988) and Scherrer et al. (1995)
treat the GONG and MDI versions of the tunable Michelson interferometer;
earlier versions, also called “Fourier Tachometers”, have been described by
Beckers and Brown (1978), Brown (1981), and Evans (1981). For the Fabry–
Perot interferometer the monographs by Tolansky (1948) and Vaughan (1989)
provide introductory as well as special material. For the solar application the
Fabry–Perots are mounted either in the collimated beam, as shown in Fig.
3.29, or in the telecentric beam near the image plane; Darvann and Owner-
Petersen (1994) and Kentischer et al. (1998) discuss the specific advantages
of these two mountings.
A magneto-optical filter that is related to the Lyot filter has been inves-
tigated first by Öhman (1956). Beckers (1970) reviews the diverse types of
magneto-optical filters.
For the study of polarized light it seems necessary to distinguish between
strictly monochromatic light, and light with a bandwidth small compared
to the width of solar lines. As Shurcliff (1962) remarks, the definition of the
Stokes parameters in terms of electromagnetic theory “is awkward in that one
must assume that the light is sufficiently monochromatic that, at any time, a
definable phase angle γ exists between the instantaneous scalar components
. . . of the electric field, yet the light must be sufficiently polychromatic that
the unpolarized state is not precluded”.
It is also awkward, of course, that the classical theory of atoms is still
used to derive the absorption coefficients (3.69) for polarized light in the pres-
ence of a magnetic field. Unno’s result, which was independently derived by
Stepanov (1958), rests on formulae which were originally presented by Sears
(1913). Fortunately, a quantum-mechanical treatment (Landi degl’Innocenti
and Landi degl’Innocenti 1972) confirms the classical result.
The monograph by Stenflo (1994) covers the theory of polarized radiation
as well as the application to the Sun, earlier reviews were written by Evans
(1966), Beckers (1971 a), and Stenflo (1971). The volume edited by Trujillo-
3.7 Bibliographical Notes 147
Bueno et al. (2002) covers solar polarimetry, but also other astrophysical ap-
plications. The inversion of polarimetric data has been treated by Wittmann
(1974 a), Landolfi et al. (1984), Skumanich and Lites (1987), Rees et al. (1989,
2000), Ruiz Cobo and del Toro Iniesta (1992), and Bellot Rubio et al. (2000).
The magneto-optic effects are included in these treatments; their significance
has been demonstrated by Landolfi and Landi degl’Innocenti (1982), although
earlier results of Rachkovsky (1962 b) and Beckers (1969 b) had indicated
that the solar Stokes profiles are not largely altered. The volumes edited by
November (1991), Rutten and Schrijver (1994), and Stenflo and Nagendra
(1996) contain reviews and detailed contributions.
A major difficulty in the interpretation of polarimetric data arises from
the inhomogeneity of the solar magnetic field. Stenflo (1971) has discussed the
problem. Solanki (1993) reviews the effects of the small-scale flux concentra-
tions that will be treated in Chap 8, and Sigwarth et al. (1999) present results
with a spatial resolution of 1 . Grossmann-Doerth et al. (2000), Martı́nez
Pillet (2000), and Steiner (2000) demonstrate that magnetic discontinuities
such as occur in penumbral channels, or in a canopy-like transition from a
non-magnetic to a magnetic layer in the atmosphere, can cause diverse line
profiles such as shown in Fig. 3.38. On the other hand, magnetic structure
on scales much smaller than the photon free path has been proposed as the
origin of the “irregular” line profiles of polarized radiation (Sánchez Almeida
et al. 1996). Keller (1995) shows how the speckle technique can be combined
with polarimetry in order to measure the solar magnetic field with diffraction-
limited resolution.
The measurement of vector fields has been documented in the proceedings
edited by Hagyard (1985). Mickey et al. (1996) and LaBonte et al. (1999)
describe an imaging vector magnetograph based on a tunable Fabry–Perot
filter. The instrumental polarization introduced by coelostat mirrors has been
considered by Capitani et al. (1989). Stenflo (1982, 2001) and Landolfi and
Landi degl’Innocenti (1986) discuss the Hanle effect and its possible use to
measure the magnetic field in prominences and other places of the outer solar
atmosphere.
A description of the water-flow pyrheliometer and results obtained from
it can be found in Aldrich and Hoover (1954).
A number of neutrino detectors other than those described in Sect. 3.6.2
have been proposed. The proceedings of a Brookhaven conference (Friedlan-
der 1978) and the books of Bahcall (1989), Klapdor-Kleingrothaus and Zuber
(1997), and Schmitz (1997) provide detailed introductions to this field.
4. The Atmosphere
We now turn to a more detailed study of the solar atmosphere, which, in the
present chapter, will be understood as the combination of the photosphere
and the chromosphere.
The photosphere is a layer of little more that one hundred kilometers
thickness where – going inwards – the solar gas changes from almost com-
pletely transparent to completely opaque. Virtually all the light which we
receive from the Sun originates in the photosphere. Therefore, most of the
information which we have about the Sun stems from this layer.
Fortunately the transition from optically thin to optically thick depends
on position on the Sun, and on wavelength – in a most pronounced way in
the spectral lines. This circumstance allows us to study the thermodynamic
state of the solar atmosphere, its chemical composition, and a variety of
hydrodynamic and hydromagnetic phenomena which abound on the Sun.
In fact, in a number of strong spectral lines the transition to the optically
thick state occurs high in the atmosphere. When we observe the Sun in such
a line we therefore see the upper part of the solar atmosphere, the chromo-
sphere. This layer has obtained its name from the colorful appearance which
it presents at the time of a solar eclipse.
In the present chapter we shall concentrate on results concerning the
mean atmosphere, i.e., averaged over the horizontal coordinates on the solar
surface. Examples are the run of temperature and pressure with height in the
atmosphere, or the abundances of the elements. In the subsequent chapters,
we shall then treat spatially resolved and local phenomena.
Hence
dIν
μ = Iν − Sν (4.2)
dτν
describes the variation of the intensity with depth; Sν is the source function,
i.e., the ratio between emission and the absorption coefficient, and μ = cos θ
(the intensity depends on τν , μ, and ν; as for the other variables, κν , Sν . . . ,
the argument ν is often written as an index).
A formal solution of the radiative transfer equation, i.e., a reduction to
the problem of evaluating an integral, can be obtained in the following way:
Multiply (4.2) by μ−1 exp(−τν /μ) and integrate; the result is
where τ0ν is the optical depth at some level of reference. If we integrate from
τν = 0, where we observe, to τ0ν = ∞, i.e., deep into the star, then we have
the total emergent intensity
∞
1
Iν (0, μ) = Sν (τν ) exp(−τν /μ) dτν . (4.4)
μ
0
Equation (4.4) can be used in a variety of ways. For instance, we may as-
sume a known source function and predict the emergent intensity. Or, which
is more interesting in the context of solar atmospheric models, we may derive
the source function from the (absolute) intensity, if the latter is measured as
a function of ν and (or) μ. This requires an inversion of the integral (4.4).
Problem 4.1. Assume that there are neither absorbers nor emitters between
two points of a radiation field. Show that the intensity, according to its
definition in Sect. 1.5.1, is the same at these two points; convince yourself
that this is consistent with the equation of transfer, (4.2).
ionization and excitation of the atoms are distributed according to the Saha
and Boltzmann equations for that same T ; and the radiation field has the
homogeneous and isotropic black-body form given by the Kirchhoff-Planck
function, again for the same T :
2hν 3 1
Bν (T ) = . (4.5)
c2 ehν/kT − 1
No temperature gradient exists in thermodynamic equilibrium, and it is ob-
vious that this situation is realized virtually nowhere.
Very often, however, the conditions of local thermodynamic equilibrium
(LTE) are very closely satisfied. That is, at a certain place, a single tempera-
ture T does suffice to describe the statistical particle velocities, the population
of the atomic states, and the local ratio of emission to absorption of radiation.
In LTE the most important simplification of the radiative transfer problem
is the relation Sν = Bν (T ).
Whether or not LTE can be assumed depends on the thermalization
length. This is the distance over which a particle or photon emitted in a
certain collision or transition has undergone sufficient further collisions or ab-
sorption/emission processes so that it can no longer be distinguished within
the respective distribution. In LTE the thermalization length must be shorter
than the distance over which the temperature of the gas changes markedly.
The preceding definition of the local thermodynamic equilibrium makes
clear that, at the same place in the atmosphere, LTE may be a good assump-
tion for one particular process or species, but may be completely wrong for
another. Consequently, the substitution Sν = Bν may be allowed when the
formation of a certain spectral line is considered, but may fail for other lines.
Often the difference becomes apparent only after departures from LTE are
explicitly taken into account. A crude rule of thumb is that the continuum
in the visible and infrared, the wings of most spectral lines, and the entire
profiles of weak lines are formed in LTE, while possible departures from LTE
must be considered for the line cores, and for strong lines.
Departures from LTE occur in particular when radiative interactions are
too rare to establish the distributions enumerated above. The thermalization
length is large, and the distributions are influenced by photons coming from
great distances, where different conditions (e.g., different velocity or atomic
state distributions) might prevail. Thus the state is “non-local”, and often
called “non-LTE”. There might still be an equilibrium everywhere. However,
the equilibrium is no longer characterized by a single temperature. In the solar
atmosphere a characteristic situation is that the electrons satisfy a Maxwell
distribution, with electron temperature Te , because collisions are still frequent.
On the other hand, the population of atomic levels depends on radiative
processes, which become rare in a rarefied gas. These populations therefore
must be described by statistical equations. Therefore, instead of non-LTE
more descriptive names are SE, for statistical equilibrium, or KE, for kinetic
equilibrium.
152 4. The Atmosphere
In principle we could use expression (4.4) for the emergent intensity regardless
whether we consider a frequency ν in the continuum or in an absorption line.
However it is often advantageous to realize that in an absorption line τν is the
optical depth calculated from the continuum and line absorption coefficients,
i.e., dτν = dτC + dτl , or
where τ is now the optical depth in the continuum. As we saw, (4.7) is most
useful in the context of radiative transfer in magnetically sensitive lines.
It remains to determine the line absorption coefficient. We have already,
in Sect. 2.3.7, written the cross section per atom as a function of frequency.
We restrict our attention to the two mechanisms of Doppler and collisional
line broadening, as expressed by profiles (2.84) and (2.85). At each frequency
of the collisional profile we must apply the Doppler-broadening formula (or
vice versa), i.e., we must fold the two expressions and obtain
∞
γ exp(−(ν − ν )2 /ΔνD2
)
φ(ν) = √
dν . (4.8)
πΔνD [2π(ν − ν0 )] + γ /4
2 2
−∞
Fig. 4.1. Solar spectrum, with Mg b2 line. From Beckers et al. (1976)
∞
e−y
2
a
H(a, v) = dy . (4.13)
π (v − y)2 + a2
−∞
The Voigt function is normalized such that for small a its value at the line
center is 1, i.e.,
H(a, 0) = 1 + O(a) , (4.14)
while φ(ν) is normalized so that the integral over all frequencies is 1.
For solar spectral lines we generally have a 1. Inspection of (4.13)
shows that in this case the Doppler profile dominates near the line center ν0 .
On the other hand, the Gaussian rapidly drops with distance from ν0 , and
hence the line wings are shaped according to the collisional, or “damping”,
part of the profile. An example where the two parts of the line profile can
clearly be distinguished is shown in Fig. 4.1 (although a strong line such as
Mg b2 cannot very accurately be treated under the assumption of LTE!).
√
Problem 4.2. Show that the integral over v of H(a, v) is π. Prove (4.14).
The various terms are separated in this way because, in LTE, the third and
fourth factors, respectively, are determined from the Saha and Boltzmann
equations, (4.59) and (4.64) below, in combination with the normalization
conditions (4.60) and (4.65). The second factor of (4.16), the relative (num-
ber) abundance of the element considered, will be taken as known for the
present purpose.
The emergent line intensity can now be calculated if two atomic constants
are known: the oscillator strength f and the damping constant γ.
Laboratory experiments, or quantum-mechanical calculations of transi-
tion probabilities, yield the f values. Standard tables have been prepared by
Wiese et al. (1966) and Wiese et al. (1969), but new results appear constantly.
The damping constant γ generally depends on density and temperature, as
well as on atomic properties. We do not discuss this matter further, since this
is done in extenso in many texts on spectroscopy. Also, we shall not treat
line-broadening mechanisms other than the two considered above, although
some can be important (e.g. Stark-effect broadening for hydrogen lines).
We must however add a complementary remark on the Doppler width,
ΔνD . The thermal random motion of atoms normally does not suffice to
explain quantitatively the “Doppler core” of spectral line profiles. Therefore,
the effect of a turbulent gas motion is added, and we write
ν0 2RT
ΔνD = + ξt2 . (4.17)
c A
The parameter ξt is called “microturbulence”, or “microvelocity” (as opposed
to the “macroturbulence” which will be discussed in Sect. 4.3.3 below).
In the higher part of the solar atmosphere we must consider departures from
local thermodynamic equilibrium. We shall treat such departures here for a
single species only. An example is silicon, which is an important absorber
in the ultraviolet, cf. Fig. 4.5. In general both the continuum and the line
absorption will be affected, because the population of the atomic levels from
which bound electrons are excited (either to higher bound levels or to the
“free”, or continuum, state) may no longer obey Boltzmann’s law.
We shall however still assume that collisions between particles occur with
sufficient frequency for the velocity distributions to become Maxwellian. In
the atmosphere, our present concern, this is a reasonable assumption, but we
must keep in mind that in more tenuous environments, such as the corona
or the solar wind, it will certainly be violated. In the present treatment
the various particles all will have Maxwellian velocity distributions with the
4.2 Radiative Transfer – Statistical Equilibrium 155
same temperature; since collisions with electrons play a dominant role for the
atomic populations discussed below, this temperature will also be called the
electron Temperature, Te . In the following we mean this parameter whenever
we speak of temperature (a term which strictly would be reserved for LTE).
Vernazza et al. (1973) have outlined the equations describing the statis-
tical equilibrium in the context of solar atmospheric models. Here I shall
follow their work, but also draw from Schleicher (1976). We shall consider an
atom with an electron for which N discrete bound states, plus a continuum
state (i.e., the state of ionization) are possible. For brevity we shall omit the
indices denoting the chemical element and the state of ionization. To begin
with, we enumerate the various transitions between the atomic states and
their characteristic constants.
as the number of induced emissions per unit time, volume, frequency inter-
val, and solid angle. Again, BUL is an atomic constant, called the Einstein
coefficient of induced emission. Notice that BUL Iν has the same dimension
as AUL and can be interpreted as an inverse lifetime.
The third radiative transition is absorption, also called radiative excita-
tion. It is the opposite of induced emission. Initially the atom is in the lower
state. When exposed to radiation of frequency ν, the atom takes a photon
out of the radiation field, and thereby is excited into the upper state. The
number of these radiative excitations will be proportional to the number nL
of atoms in the lower state, and to the intensity Iν . As before, we must allow
for the finite width in frequency of the absorption process, so that
nL BLU Iν φ(ν)/4π (4.20)
is the number of absorbed photons per unit time, volume, frequency interval,
and solid angle; BLU is the Einstein coefficient for radiative excitation.
The profile φ(ν) is of course the absorption profile which we have already
considered in Sects. 2.3.7 and 4.1.3.
Problem 4.3. Relate the Einstein coefficient BLU to the oscillator strength
defined in (2.83).
Problem 4.4. Relations between Einstein coefficients. Consider the special
case of thermodynamic equilibrium, where the levels U and L are pop-
ulated according to (4.64), where Iν = Bν (T ), and where the principle
of detailed balance is valid, i.e., the number of upward transitions must
be equal to the number of downward transitions. Show that under these
circumstances the following relations must hold:
gU BUL = gL BLU (4.21)
gU 2hν 3
AUL = 2 BLU . (4.22)
gL c
Notice that, because the Einstein coefficients are atomic constants, rela-
tions (4.21) and (4.22) are generally valid, and not only in the special case
of thermodynamic equilibrium!
4.2 Radiative Transfer – Statistical Equilibrium 157
The Einstein coefficients AUL , BUL , and BLU can be calculated theoreti-
cally as transition probabilities, or may be determined by laboratory experi-
ments. For the present text we shall take them as known quantities.
Problem 4.5. In analogy to Problem 4.4, use the special case of local ther-
modynamic equilibrium (where the principle of detailed balance is valid)
to show that (4.24) can generally be replaced by the expression
n∗j hν 2hν 3
αj (ν)nC ∗ exp − + Iν hν , (4.25)
nC kT c2
where
3/2
n∗j h2 ne gj E C − Ej
= exp (4.26)
n∗C 2πme kT 2uC kT
is the LTE ratio of the populations of levels j and C (EC is the ionization
energy, uC is the partition function of the ionized state, cf. Sect. 2.3.4).
158 4. The Atmosphere
4.2.4 Collisions
dIν
μ = −κρIν + ε . (4.30)
dr
According to the diverse processes enumerated above, we must distinguish
between line absorption and continuum absorption; that is, we write
κ = κl + κC , (4.31)
where κl is the line absorption coefficient, and κC is the continuum absorption
coefficient. We also divide the source term of (4.30), ε, into line emission εl
and continuum emission εC :
ε = ε l + εC . (4.32)
Let us first consider line radiation. Knowing the Einstein coefficients for
the radiative transitions we can write down the net gain of the intensity at
frequency ν through these transitions. For the particular transition between
levels L and U the number of involved photons is given by expressions (4.18)
to (4.20). Each photon carries an energy hν; hence
hν
[nU (AUL χ + BUL Iν ψ) − nL BLU Iν φ] (4.33)
4π
is the change of energy, at frequency ν, per unit time, volume, frequency
interval, and solid angle. Instead to the volume we may as well relate this
change to area times length; in this way we recognize that (4.33) is just the
right-hand side of the transfer equation – as far as line radiation is concerned
– i.e., identical to
−ρκl Iν + εl . (4.34)
We replace εl by a line source function Sl , defined by Sl = εl /ρκl , compare
expressions (4.33) and (4.34), and find
hν
κl = (nL BLU φ − nU BUL ψ) , (4.35)
4πρ
nU AUL χ
Sl = . (4.36)
nL BLU φ − nU BUL ψ
Expression (4.35) makes clear why stimulated emission is also called “neg-
ative absorption” (e.g., Aller 1963): it simply enters as a negative contribution
to the line absorption coefficient.
At this point we introduce another simplifying concept, known as complete
redistribution. This concept consists of the assumption that the frequency
profiles of the three radiative processes are all equal:
χ(ν) = ψ(ν) = φ(ν) . (4.37)
This assumption is valid only under special circumstances; for example in
the extreme case of exact coherence, where a photon is emitted at precisely
160 4. The Atmosphere
the same frequency at which it was absorbed; or in the other extreme case
of complete incoherence where the frequency of emission is completely inde-
pendent of the frequency of absorption. In general these conditions are not
satisfied, and we must deal with the individual profiles; this case is called
partial redistribution. Here we shall be content with (4.37), which makes the
line source function (4.36) independent of frequency:
nU AUL
Sl = . (4.38)
nL BLU − nU BUL
Under the approximation of complete redistribution the line absorption
coefficient (4.35) is proportional to the function φ(ν), which we shall consider
as given in form of a Voigt function (4.13).
Problem 4.6. Use (4.21) and (4.22) to eliminate the Einstein coefficients
from the line source function. Derive the line source function for the case
where several lines are treated simultaneously.
Next we consider the continuum. With expressions (4.23) and (4.25) above
we have already prepared what we need for the continuum part on the right-
hand side of the transfer equation. Therefore
n∗ 2hν 3
−hν/kT j
−ρκCj Iν + εCj ≡ αj (ν) nC e + Iν − nj Iν . (4.39)
n∗C c2
This defines the continuum absorption coefficient κCj and the continuum
source function SCj = εCj /ρκCj for transitions to and from level j:
−1
n∗
−(hν/kT ) j
κCj = ρ αj (ν) nj − nC e , (4.40)
n∗C
2hν 3 1
SCj = . (4.41)
c2 (nj /nC )(n∗C /n∗j ) exp(hν/kT ) − 1
The ratio n∗j /n∗C is again given by (4.26); obviously SCj = Bν in the case of
LTE.
Problem 4.7. Derive the continuum absorption coefficient κC and the con-
tinuum source function SC for the case where transitions to and from
several bound states are considered simultaneously.
We may now treat line and continuum radiation together. While the ab-
sorption coefficient is simply the sum of the diverse κ’s, the source function
is the weighted sum:
κ l S l + κC S C
S= . (4.42)
κ l + κC
4.2 Radiative Transfer – Statistical Equilibrium 161
On the left of this equation we have the transitions from the jth level to
other levels, on the right we have the transitions from those other levels to
level j (the sums include the bound states as well as the continuum).
For bound levels the rates Rji are given in terms of the Einstein coefficients
and the collisional rates:
Rji = Aji + Bji J¯ji + Cji ; (4.47)
the term Aji on the right must be deleted whenever level i is higher than
level j.
For transitions from level j to the continuum (level C) the rate follows
from an integration of (4.23) over all frequencies above the frequency νjC
needed for ionization, integration over solid angle, and addition of the colli-
sional rate:
∞
αj (ν)
RjC = 4π Jν dν + CjC , (4.48)
hν
νjC
162 4. The Atmosphere
Equations (4.46) and (4.51) determine the number densities that are
needed for the radiative transfer in statistical equilibrium. Instead of these
number densities it is illustrative and common to determine the departure
coefficients defined by
nj /n∗j
bj = , bC = 1 . (4.52)
nC /n∗C
It is important to realize that the number densities which are obtained
by solving the statistical equations themselves depend on the intensity (via
J¯ji ); that is, the absorption coefficient and the source function depend on
intensity. The equation of transfer is therefore generally non-linear, and is
solved by an iterative method.
Problem 4.8. Transform (4.46) into a system of equations for the departure
coefficients.
Fig. 4.2. The Sun on 11 August 1958, with limb darkening and several sunspot
groups. The dashed curve is the equator. Photograph Einsteinturm, Astrophysical
Observatory Potsdam, Germany
into deeper layers when we observe normal to the surface, and into shallower
depth near the limb where the line of sight is nearly tangential.
In fact we have already been acquainted with an atmospheric model which
predicts the limb darkening: the Eddington approximation of Sect. 2.3.8. At
τ = 0 this approximation yields an intensity variation which is linear in μ:
(where μ = cos θ). The comparison of this result with observations (Fig. 4.3)
shows that the model is quite satisfactory near the disc center, i.e., at great
depth, but fails near the limb.
164 4. The Atmosphere
We return to expression (4.4) for the emergent intensity. Let us now use wave-
length λ instead of frequency for the spectral coordinate, and let us assume
local thermodynamic equilibrium, i.e., Sλ = Bλ . Let us also, for the mo-
ment, restrict ourselves to the continuum, and neglect lines. The absorption
coefficient then varies in a smooth fashion.
We have already seen in Chap. 1 that the absolute intensity has been
observed over a wide range of the spectrum. Using this result we may in
principle derive both the temperature as a function of τλ and the variation of
κλ with λ. To see this consider first a fixed wavelength, say λ = 500 nm, where
the continuum can clearly be distinguished (Fig. 4.4) and which is often taken
as a reference wavelength. Due to the exponential factor in the integrand of
(4.4) the contribution at any given τλ is more strongly attenuated near the
limb (small μ) than near the center (μ = 1). Hence, taking observations at
various μ, the function Bλ (T (τλ )) is weighted in different ways, and may thus
Fig. 4.4. Normalized intensity of the solar spectrum near 500 nm. Data from the
Liège Atlas of Delbouille et al. (1973)
4.3 Atmospheric Models 165
where nij is the number density of species i in the jth state of ionization.
The ratios of the nij are determined by Saha equations, cf. (2.37), which we
write in abbreviated form
ni,j+1 /nij = fij (Pe , T ) ; (4.59)
4.3 Atmospheric Models 167
The Xi are the mass fractions of the chemical elements (denoted earlier by
X, Y , . . . ) and are considered as given for the present purpose. The system
(4.57) to (4.60) is then closed by the relations for the electron pressure
Pe = ne kT , (4.61)
the density
ρ = P μ/RT , (4.62)
and the mean molecular weight
μ= ni Ai /(ne + ni ) (4.63)
i i
(not to be confused with μ = cos θ!). Equation (4.62) implies that only the
gas pressure is taken into account, which may suffice for the present pur-
pose. Finally, for κC (λ) we also need the number nijk of each atom (or ion)
in the diverse excited levels (ith species, jth ionization state, kth excited
level). Since we have assumed LTE, these numbers are given by Boltzmann
distributions,
nijk = nij0 (gijk /gij0 ) exp(−Eijk /kT ) , (4.64)
and by
nijk = nij . (4.65)
k
Of course, the first guess of T (τ500 ) may not provide the best fit to the
observed intensities. Then we must iterate.
As a practical matter we may integrate (4.56) in the inverted form, be-
cause when τ500 1, equidistant steps in pressure make more sense than
equidistant steps in τ500 ; or we may use ln τ500 or ln P as an independent
variable; or we may transform (4.56) into a differential equation for the elec-
tron pressure Pe . But in principle the scheme is always as outlined by (4.55)
to (4.66).
168 4. The Atmosphere
Fig. 4.6. Model atmospheres of the Sun. From Vernazza et al. (1976)
It is also important to realize that often not all the information is needed,
and that only the dominant absorbers must be treated with care. In the
infrared, for example, only the H− ion must be taken into account.
A solar model that has been obtained on the basis of continuum inten-
sities alone is the “Bilderberg Continuum Atmosphere”, or BCA, so named
after a 1967 conference hotel near Arnhem, Holland (Gingerich and de Jager
1968). The BCA is depicted in Fig. 4.6 as a dotted curve. Although this
model disagrees in detail with the later results, also shown in Fig. 4.6, it does
show a temperature minimum and a temperature rise for τ500 < 10−4 . The
information concerning these higher layers comes from both the ultraviolet
and the infrared. Only in these spectral regions the atmosphere is sufficiently
opaque that the higher layers can be observed in the continuum. But in the
infrared, near 1.6 μm, is also the spectral region where the solar atmosphere
is most transparent, i.e., where we see the deepest of all observable layers.
A Model Including Lines. In the continuum, high layers of the solar at-
mosphere are opaque, and therefore observable, only in the ultraviolet and
4.3 Atmospheric Models 169
Fig. 4.7. Level diagram for a Si I model atom. Ver- Fig. 4.8. Departure coef-
tical distances are in energy; numbers on the solid ficients for the 8-level Si I
lines (the radiative transitions) are wavelengths in atom. From Vernazza et al.
μm. From Vernazza et al. (1976) (1981)
Fig. 4.9. Temperature as a function of depth in the solar atmosphere, and approxi-
mate depths where diverse continua and lines originate. From Vernazza et al. (1981)
Fig. 4.10. Observed and calculated EUV intensity. From Vernazza et al. (1981)
172 4. The Atmosphere
rors in the atmospheric model. Lines of hydrogenic atoms are also complicated
due to their broadening by the linear Stark effect.
In order to reconcile the model with the observations it may be neces-
sary to introduce a “macroturbulence” or “macrovelocity”. In contrast to
the microturbulence mentioned earlier, which broadens the Doppler core of
the absorption coefficient and so affects κl and Sl , the macroturbulence con-
sists of a normalized distribution (mostly a Gaussian is chosen) of width
Δλmac = λVmac /c, which is folded with the line profile after the calculation
is completed. It adds no net absorption, i.e., the total area between the con-
tinuum and the line profile remains unchanged (so does the equivalent width,
which is that area divided by the continuum intensity). For the two Na D lines
Fig. 4.12 demonstrates how the profiles are influenced; in this case the best
agreement with observed profiles is obtained with Vmac = 2 km/s. The name
macroturbulence is derived from the idea that in the solar atmosphere large
moving parcels of gas introduce local Doppler shifts which are not resolved
observationally.
It is often useful to know which layers of the solar atmosphere contribute
to the emergent intensity in a certain line profile. For this purpose we must
evaluate not only the total emergent intensity, i.e., the integral over optical
depth, but also the integrand itself. The latter is also called the contribu-
tion function. Figure 4.13 is an example: for the two Na D lines it shows, at
each wavelength, where the contribution exceeds 1/3 of its maximum, and
the height of the maximum itself. It is remarkable that around 80 mÅ from
the line center the contribution to these lines is more or less divided into two
layers, with a clear gap at 500 km. The reason is the gradient of temperature
in the solar atmosphere, and the concomitant gradient in the width of the
Doppler core. Hence the absorption profile is broad in the highest layer be-
cause the Doppler core is broad, it is narrow in the intermediate layers where
the Doppler core is narrow and collisions are still unimportant, and it is again
broad in the deepest layer because there collisional “damping” generates the
broad wings.
174 4. The Atmosphere
Table 4.1. Atmospheric parameters for the quiet Sun. From model C of Vernazza
et al. (1981). Z + n means Z × 10n
h τ500 T ξt nH ne P PG /P ρ
[km] [K] [km/s] [m−3 ] [m−3 ] [Pa] [kg/m3 ]
2080 3.51−7 8180 8.55 6.54+16 3.78+16 1.80−2 0.689 1.53−10
2070 3.77−7 7940 8.50 6.96+16 3.78+16 1.84−2 0.681 1.63−10
2050 4.30−7 7660 8.42 7.71+16 3.79+16 1.94−2 0.670 1.80−10
2016 5.20−7 7360 8.22 9.08+16 3.81+16 2.12−2 0.662 2.12−10
1990 5.90−7 7160 8.01 1.03+17 3.86+16 2.28−2 0.660 2.42−10
1925 7.72−7 6940 7.63 1.38+17 4.03+16 2.78−2 0.662 3.23−10
1785 1.21−6 6630 6.92 2.60+17 4.77+16 4.51−2 0.677 6.08−10
1605 1.96−6 6440 5.85 6.39+17 6.01+16 9.33−2 0.726 1.49−09
1515 2.42−6 6370 5.26 1.05+18 6.46+16 1.41−1 0.760 2.45−09
1380 3.29−6 6280 4.51 2.27+18 7.60+16 2.77−1 0.805 5.32−09
1280 4.08−6 6220 3.92 4.20+18 7.49+16 4.79−1 0.842 9.82−09
1180 5.08−6 6150 3.48 7.87+18 8.11+16 8.53−1 0.869 1.84−08
1065 6.86−6 6040 2.73 1.71+19 9.35+16 1.73+0 0.914 4.00−08
980 9.15−6 5925 2.14 3.15+19 1.04+17 3.01+0 0.944 7.36−08
905 1.24−5 5755 1.70 5.55+19 1.05+17 5.04+0 0.963 1.30−07
855 1.55−5 5650 1.53 8.14+19 1.06+17 7.21+0 0.969 1.90−07
755 2.54−5 5280 1.23 1.86+20 8.84+16 1.53+1 0.978 4.36−07
705 3.29−5 5030 1.09 2.94+20 7.66+16 2.28+1 0.982 6.86−07
655 4.45−5 4730 0.96 4.79+20 8.09+16 3.50+1 0.985 1.12−06
605 7.02−5 4420 0.83 8.12+20 1.11+17 5.52+1 0.988 1.90−06
555 1.46−4 4230 0.70 1.38+21 1.73+17 8.96+1 0.991 3.23−06
515 3.01−4 4170 0.60 2.10+21 2.50+17 1.34+2 0.993 4.90−06
450 1.02−3 4220 0.53 3.99+21 4.52+17 2.57+2 0.995 9.33−06
350 5.27−3 4465 0.52 9.98+21 1.11+18 6.80+2 0.995 2.33−05
250 2.67−2 4780 0.63 2.32+22 2.67+18 1.69+3 0.994 5.41−05
150 1.12−1 5180 1.00 4.92+22 6.48+18 3.93+3 0.985 1.15−04
100 2.20−1 5455 1.20 6.87+22 1.07+19 5.80+3 0.980 1.61−04
50 4.40−1 5840 1.40 9.20+22 2.12+19 8.27+3 0.975 2.15−04
0 9.95−1 6420 1.60 1.17+23 6.43+19 1.17+4 0.970 2.73−04
−25 1.68+0 6910 1.70 1.26+23 1.55+20 1.37+4 0.969 2.95−04
−50 3.34+0 7610 1.76 1.32+23 4.65+20 1.58+4 0.970 3.08−04
−75 7.45+0 8320 1.80 1.37+23 1.20+21 1.79+4 0.971 3.19−04
is probably one reason why some results obtained from center-to-limb studies
disagree with results from the disc center alone.
As long as the horizontal scale of the fluctuations is large in comparison
to the scale of vertical stratification, it is meaningful to construct individual
atmospheric models for the various features. Vernazza et al. (1981) have
calculated such models. Their model C, representative for the quiet Sun,
is listed in Table 4.1. This table does not include the transition layer to the
corona. There the more recent models of Fontenla et al. (1990, 1993) should
be consulted; these models, based on additional information from helium
lines, do not need the temperature plateau at ≈ 2120–2260 km (Fig. 4.9) to
account for the Lyman α emission (cf. Sect. 9.1.2).
Models with different temperature profiles have also been “mixed”. That
is, as a crude way to consider the horizontal variation, the various emergent
spectral intensities have been added, with the weights chosen so that the
result gives a better match to observational data. An example is the “mean
model”, which, in addition to model C, is shown in Fig. 4.10. We shall return
to such multi-component model atmospheres when we discuss – in the context
of the granular velocity field – asymmetries of solar spectral lines.
The atmospheric models presented in this section are called semi-empirical
because their temperature is adapted in order to reproduce the observed spec-
tral intensity. The law of conservation of energy has not been used. If we had
done that, we would have been able to predict the temperature as a function
of height, i.e., to calculate a theoretical model. In Sect. 9.4 we shall address
this much more difficult problem.
For the treatment of model atmospheres we have so far pretended that the
abundance of the chemical elements were known. For example, we could have
used the abundances measured elsewhere in the solar system, e.g., in mete-
orites. But the assumption that those are valid for the Sun itself of course
needs justification. Therefore, the solar abundances must be determined di-
rectly in an analysis of the solar spectrum.
The most common approach is spectrum synthesis. That is, the number
densities ni [in (4.16) or (4.60)] are treated as adjustable free parameters.
Together with the other quantities (temperature, turbulence parameter, etc.)
they must be chosen in such a way that the intensity of as many spectral lines
as possible is correctly modeled.
176 4. The Atmosphere
the Sun, but the error margin is large. On the other hand, the photospheric
depletion of lithium by about two orders of magnitude, first recognized by
Greenstein and Richardson (1951), is firmly established.
The light elements Li, Be, and B can be destroyed by nuclear reactions
with protons. In addition to those already listed in Sect. 2.3.6, the actual
reactions are
6
Li(p, 3 He)α , 9
Be(p, α)6 Li ,
(4.68)
10
B(p, α)7 Be , 11
B(p, γ)3α .
A case of special interest is the “burning” of lithium and beryllium. The
reaction rates for these two elements become significant (in view of the solar
age) at temperatures of 2.5 × 106 K and 3 × 106 K, respectively. Thus the
strong (but not total) depletion of Li and the absence of Be depletion can
provide a hint to the depth of the outer convection zone of the Sun: Convec-
tion (including convective overshooting into the subadiabatic range below the
proper convection zone) and the concomitant mixing should reach the layer
where T ≈ 2.5 × 106 K, but essentially avoid the layer where T ≈ 3 × 106 K.
We may conclude from this argument that the solar convection zone should
have a depth of ≈ 2 × 108 m, in rough agreement with model calculations
which show that the stratification becomes subadiabatic, and hence stable
against convection, at about this depth (Sect. 6.2), and in agreement with
the helioseismic result (Sect. 5.3).
However, a detailed calculation shows that the problem of Li depletion is
not fully solved: During the ≈ 5 × 107 years of the Sun’s evolution from the
Hayashi line toward the zero-age main sequence the outer convection zone
may have reached a depth where T ≈ 3.5 × 106 K, and all the lithium may
have been lost (e.g., Ahrens et al. 1992, Schlattl and Weiss 1999). But the
Sun is not completely depleted of lithium; moreover, there is evidence that
Sun-like stars, younger than the Sun but also on the main sequence, have less
lithium depletion or even the “normal” abundance. Hence it appears that Li
depletion takes place during the main-sequence evolution. Since in this phase
the outer convection zone has not been sufficiently deep for the necessary
mixing, some additional mixing below the convection zone should occur. – In
terms of energy generation the contribution of lithium burning is small and
of no concern.
4.4.3 Helium
Helium deserves a special comment. Although a line of this element had been
discovered as early as 1868 by N. Lockyer (and ascribed to “helium” because
at the time it was unknown on Earth) the spectroscopic determination of
its abundance has remained rather inaccurate, with an error of order 0.2 in
log A. The reason is that all important helium lines fall into the ultraviolet
4.5 Bibliographical Notes 179
or infrared parts of the spectrum, and that these lines are produced in the
chromosphere and corona under conditions which largely deviate from LTE.
In view of the spectroscopic difficulties it is fortunate that a most accurate
method to determine the solar photospheric helium abundance is helioseis-
mology, as first proposed by Gough (1984 c). In a zone where an abundant
element is partially ionized the specific heats cP and cV both increase because
of the energy required for ionization, while their difference remains nearly un-
changed. Therefore the ratio of those specific heats, which is approximately
equal to the adiabatic exponent Γ1 , becomes closer to 1. Since the velocity
1/2
of sound is proportional to Γ1 , it is clear that the frequencies of oscillation
must be influenced by a change of the abundance. The second ionization of
helium occurs sufficiently deep in the Sun so that its signature on the eigen-
frequencies is not spoiled by the uncertainty of near-surface convection, and
is clearly discernable. Thus, the value log A = 10.93±0.004 listed in Table 4.2
is based on helioseismic inversion (Sect. 5.3), which yields the mass fraction
Ys = 0.248 ± 0.002 of helium at the solar surface, and thus
nHe /nH = Ys /4Xs , (4.69)
where Xs is the surface mass fraction of hydrogen.
The value Ys = 0.248 is very close to the result Ys = 0.245 obtained for the
theoretical solar model described in Sect. 2.4, where the helium abundance
has been calibrated by observed properties of the Sun, and where the process
of He diffusion towards the solar center (Sect. 2.3.3) has been taken into
account. Without such diffusion the calibration would yield Ys ≈ 0.28. This
demonstrates that it is appropriate to include that process into the standard
solar model.
A new branch of solar research has been developed since 1975, when it
was discovered that the photospheric periodic motions previously known as
“5-minute oscillations” have a spectrum of discrete frequencies. A large num-
ber of such modes, with periods in the range 2–15 minutes, have been ob-
served. They are identified as acoustic waves, or p modes, where the pressure
gradient is the main restoring force. The discrete mode pattern is a conse-
quence of reflecting boundaries and can be used to obtain information about
the Sun’s interior.
Gravity provides the restoring force for a second type of oscillation, with
lower frequencies. These internal gravity modes depend on a stable stratifica-
tion: they are supported by the radiative interior of the Sun, with a discrete
spectrum of g modes, which, however, has not yet been confirmed by observa-
tion; on the other hand, there are hints that they exist above the convection
zone in the stable solar atmosphere.
5.1 Observations
Fig. 5.2. Spectrum and velocity curves for the lines Mg b2 and Ti I λ = 517.37 nm,
obtained at Sacramento Peak Observatory. The arrows indicate the velocity in km/s
and the distance on the Sun in km. From Evans and Michard (1962)
5.1 Observations 183
Problem 5.1. How does the contrast pattern of Fig. 5.1 change if the two
Doppler plates are added instead of subtracted?
Problem 5.2. Consider the Fourier transform of a signal which has data gaps
as a consequence of the day/night cycle. Show that a line at ν has side
lobes at ν ± 11.57 μHz. Where are the side lobes if the data are collected
by a satellite in a 90-minute orbit with no continuous sunlight? Calculate
the height of the side lobes as a function of the length of the gaps.
In practice the integral is replaced by a sum and the rules of Fourier analysis
must be observed. The power spectrum is
P (kx , ky , ω) = f f ∗ . (5.3)
For the solar oscillations the important fact is that the power is not evenly
distributed in the kh , ω-plane, but instead follows certain ridges. As we shall
see in Sect. 5.2.4, each of these ridges corresponds to a fixed number of
wave nodes in the radial direction. The ridges were theoretically predicted
by Ulrich (1970), and first observed by Deubner (1975) with the Domeless
Coudé Telescope at Capri. Figure 5.4 shows a later example, where up to
15 ridges can be identified in the velocity power spectrum. The same ridge
pattern is seen in the power spectrum of intensity variations, measured, e.g.,
in Ca II K filtergrams (Fig. 5.7).
We shall see that the pressure perturbation essentially provides the restor-
ing force for the 5-minute oscillations. Hence the discrete modes into which
these oscillations are decomposed, and which we shall treat in the following
two sections, have been named p modes by Cowling (1941). The lowest ridge,
called fundamental or f mode, is an exception: this mode is essentially without
compression, and resembles a surface wave on deep water (Sect. 5.2.4).
In the three-dimensional kx , ky , ω-space the p-mode power is distributed
along trumpet-shaped surfaces. On the average these trumpet surfaces are
symmetric about the ω-axis, and a slice at constant ω forms a ring. However,
a large-scale stationary flow that underlies the wave field causes a deviation
of the rings from the circular symmetry; hence such a ring diagram can be
used to determine large-scale flows, both at the solar surface and underneath
the surface (Hill 1988).
Spherical Harmonics. Whenever the scan length becomes comparable to
the solar radius, the cartesian coordinates used so far must be replaced by
spherical polar coordinates (r, θ, φ). Instead of (5.2) we must use an expansion
in terms of spherical surface harmonics, i.e.,
∞
l
v(θ, φ, t) = alm (t)Ylm (θ, φ) . (5.6)
l=0 m=−l
5.1 Observations 187
Fig. 5.5. Node circles of spherical harmonics. After Noyes and Rhodes (1984)
terference of 107 or more single modes with randomly distributed phases. The
smallest amplitudes presently measured are of order 1 mm/s (Kosovichev et
al. 1997).
Problem 5.3. How many modes of oscillation lie between degrees l = 0 and
l = 1000, if 10 ridges are included and all m values are counted for each l?
Solar oscillations have also been observed as Doppler shifts in spatially unre-
solved spectra. The primary instrument used is the resonance-scattering spec-
trometer described in Sect. 3.3.3. Such an instrument yields a signal which
is averaged over the visible solar disc. Naturally the oscillations with smaller
horizontal wavelength (larger l) contribute less to this average because the
effects from areas of opposite phase tend to cancel each other.
Oscillations observed in spatially unresolved sunlight are sometimes called
“global” oscillations. In contrast to this definition, which would depend on the
particular observational setup, we shall rather attribute the word “global” to
those oscillations which can exist in certain regions extending over the entire
190 5. Oscillations
Degree l
Mode 0 1 2 3 4 5 6 7
p 7 1401.6 1443.7 1483.4
p 8 1500.3 1545.3 1587.5 1627.5
p 9 1641.0 1685.9 1727.8 1767.6
p 10 1778.0 1823.4 1866.2 1906.8
p 11 1815.4 1868.2 1914.8 1960.5 2004.2 2045.9
p 12 1823.6 1885.5 1947.5 2003.4 2051.7 2098.4 2142.6 2184.5
p 13 1957.3 2020.7 2084.1 2138.7 2188.4 2235.4 2279.9 2322.2
p 14 2093.5 2156.7 2218.4 2274.7 2324.2 2371.2 2415.6 2458.2
p 15 2228.6 2291.9 2352.2 2409.4 2458.6 2506.1 2551.1 2594.3
p 16 2362.5 2426.1 2486.8 2541.9 2593.0 2641.2 2687.0 2731.1
p 17 2496.6 2558.9 2620.3 2677.4 2728.3 2777.3 2823.7 2868.3
p 18 2629.6 2693.6 2754.5 2810.9 2864.2 2913.7 2960.6 3005.6
p 19 2764.4 2828.1 2890.1 2947.9 3000.1 3049.9 3097.2 3142.7
p 20 2899.3 2963.3 3024.1 3083.5 3135.9 3186.2 3233.9 3279.8
p 21 3033.8 3098.7 3160.0 3218.5 3271.6 3322.5 3370.8 3417.4
p 22 3168.6 3233.2 3296.1 3354.5 3408.0 3459.3 3508.2 3555.2
p 23 3304.1 3368.9 3431.2 3489.5 3544.4 3596.5 3645.9 3693.6
p 24 3439.8 3504.6 3567.2 3626.1 3681.4 3733.7 3783.7 3832.0
p 25 3576.3 3640.2 3703.3 3760.9 3818.9 3871.9 3922.1 3970.4
p 26 3711.5 3777.4 3837.8 3897.6 3956.3 4009.8 4060.7 4109.6
p 27 3847.4 3914.1 3975.5 4035.0 4093.9 4147.5 4198.9 4248.3
p 28 3984.9 4052.1 4112.9 4171.7 4231.2 4286.3 4338.1 4387.3
p 29 4121.9 4189.8 4249.3 4308.6 4371.0 4424.4 4476.5 4526.9
p 30 4257.4 4325.6 4387.3 4443.8
p 31 4394.9 4463.6 4524.2 4583.5
p 32 4532.3 4600.8 4656.9 4717.4
p 33 4668.6 4738.6
body of the Sun. We shall see that reflecting boundaries exclude some modes
from some parts of the Sun, but also that it is exactly the presence of these
boundaries which generates the discrete mode character of the oscillations
and thus manifests the global nature.
Even in a spatially unresolved Doppler signal single modes of oscillation,
with various (small) degrees l, can be distinguished. But for this purpose it
is a necessity to have good resolution in frequency; according to (5.1), this
means signals of sufficient duration in time. It is true that the first discovery
of low-l oscillations (Claverie et al. 1979) was made with signals of one and
two days length; but only the 5-day signal, obtained between 31 December
1979 and 5 January 1980 at the South Pole (Grec et al. 1983) revealed the
multitude and sharpness of lines in the solar low-l oscillation spectrum. Still
longer signals have been obtained since 1981 by G. R. Isaak and collaborators
who combined data recorded at the Canary Islands and at Hawaii; this was
the beginning of the above-mentioned BISON network. The large difference
5.1 Observations 191
3.0
2.5
2.0
ppm2
1.5
1.0
0.5
0.0
2000 2500 3000 3500 4000
frequency [μHz]
Fig. 5.9. Power spectrum of irradiance variations, measured with the VIRGO
instrument on SOHO in the 402 nm band. Courtesy W. Finsterle and C. Fröhlich
192 5. Oscillations
The lines in the solar oscillation spectrum have a finite width. The line width
is determined by the mode lifetime if this is shorter than the duration of the
observed signal. Indeed the lifetime of a coherent oscillation is limited both
because the oscillation may loose energy by some damping process and be-
cause the oscillation is continuously disturbed by the non-stationary velocity
field in the convection zone. In any case the damped harmonic oscillator is
an appropriate model, and we may expect a Lorentzian frequency profile as
in the case of collisional broadening of the lines in an optical spectrum, cf.
Eq. (2.85),
γ
φ(ν) = . (5.9)
[2π(ν − ν0 )]2 + γ 2 /4
The full width at half-maximum of the Lorentz profile is ΔνL = γ/2π. The
observed lines often fit very well to a Lorentz profile; the line widths obtained
from such fits increase with increasing oscillation frequency, as illustrated in
Fig. 5.11. The modes with the largest amplitude, with frequencies around
3 mHz, have a width of ≈ 1 μHz, the lines at lower frequencies are sharper,
the lines at higher frequencies broader. The lifetime of a mode is given by 2/γ,
or the inverse of π times the full width; hence the line widths shown in the
figure imply mode lifetimes between hours and months. The ratio Q = ν/ΔνL
is the quality factor of the oscillation. Solar p modes typically have quality
factors between 102 and 104 .
In general the lines found in the power spectra of solar oscillations do not
possess perfect Lorentzian profiles, but are slightly asymmetric. Duvall et al.
(1993 a) first derived such line asymmetries from a full-disc observation of
the intensity oscillations in the Ca II K line; later data, e.g. from the GONG
network and from SOHO/MDI, confirmed the result with respect to velocity
oscillations. Mainly the modes in the lower frequency range, ν < 3 μHz, show
an asymmetry, and it appears that the asymmetry is opposite in intensity
and velocity power spectra. If an asymmetric line is fitted with a symmetric
Lorentzian profile, a systematic error may be introduced in the determination
of the mode frequency.
We shall first assume that the perturbations occur sufficiently fast so that
everywhere in the star the adiabatic approximation
δP δρ
= Γ1 (5.10)
P0 ρ0
is satisfied. Here P0 and ρ0 are the undisturbed pressure and density, both
dependent on radius r, and δP and δρ are the respective Lagrangian pertur-
bations, i.e., the perturbations suffered by a fixed “parcel” or “element” of
material as it moves back and forth. The adiabatic exponent,
d ln P
Γ1 = , (5.11)
d ln ρ S
is related to the adiabatic sound velocity c by
c2 = Γ1 P0 /ρ0 , (5.12)
and also depends only on depth.
It is only in the atmosphere of the Sun that the radiative exchange be-
tween places of (due to the perturbation) different temperature is sufficiently
5.2 Linear Adiabatic Oscillations of a Non-Rotating Sun 195
ρ1 + ∇ · (ρ0 ξ) = 0 , (5.13)
∂2ξ ρ1
ρ0 + ∇P1 − ∇P0 + ρ0 ∇Φ1 = 0 , (5.14)
∂t2 ρ0
where ξ is the vectorial distance of a gas parcel from its equilibrium posi-
tion. The subscript 1 denotes Eulerian perturbations, i.e., perturbations at a
fixed position in the Sun; these are related to the Lagrangian perturbations,
denoted by a δ, through
δf = f1 + ξ · ∇f0 . (5.15)
Since the coefficients of our linear equations, which are known from the solar
equilibrium model, depend only on depth, we may separate the time and
angular dependencies. But first we notice, taking the curl of (5.14), that the
vorticity of the perturbation ξ has no vertical component:
196 5. Oscillations
∂2ξ
r · curl =0. (5.17)
∂t2
Since we are primarily interested in oscillations, with ∂/∂t ≡ iω = 0, this
means (in spherical polar coordinates)
∂ ∂ξθ
(sin θξφ ) − =0. (5.18)
∂θ ∂φ
The horizontal components, ξθ and ξφ , of the perturbation vector can there-
fore be obtained as derivatives of a single scalar function. This function, and
equally the radial component of ξ, will be expanded in terms of spherical
harmonics. Each term has the form
∂ ξh (r) ∂
ξ = eiωt ξr (r), ξh (r) , Ylm (θ, φ) . (5.19)
∂θ sin θ ∂φ
Ylm is the complex spherical harmonic of degree l and angular order m intro-
duced in the preceding section. Whenever physical quantities are discussed,
the real part of expression (5.19) must be taken.
The Eulerian perturbations of density, pressure, and gravitational poten-
tial are expressed in a corresponding manner:
(ρ1 , P1 , Φ1 ) = eiωt [ρ1 (r), P1 (r), Φ1 (r)] Ylm (θ, φ) . (5.20)
Here, for the sake of simplicity, the same symbols are used for the perturba-
tions and for their r-dependent factors.
It is a straight-forward matter to substitute (5.19) and (5.20) into the
above equations, and to eliminate the variables ξh and ρ1 . We obtain
1 d 2 ξr g 1 1 l(l + 1) l(l + 1)
2
(r ξr ) − 2 + 2
− 2 2 P1 − 2 2 Φ1 = 0 , (5.21)
r dr c ρ0 c r ω r ω
1 d g dΦ1
+ 2 P1 − (ω 2 − N 2 )ξr + =0, (5.22)
ρ0 dr c dr
1 d 2 dΦ1 l(l + 1) 4πGρ0 2 4πG
r − Φ1 − N ξr − 2 P1 = 0 . (5.23)
r2 dr dr r2 g c
ω 2 − ωA
2
N 2 − ω2
kr2 = 2
+ Sl2 2 2 , (5.34)
c c ω
where we have defined the acoustic cutoff frequency,
ωA = c/2H . (5.35)
Problem 5.6. Confirm (5.34). Write the dispersion relation for the normal-
ized frequency and wave number used in Fig. 5.12.
Problem 5.7. A generalization of (5.35). If the condition of an isothermal
atmosphere is dropped, the wave equations for the diverse variables ξr ,
P1 , etc., have different forms. Show that the acoustic cutoff is given by
c2 dH
2
ωA = 1−2 , (5.36)
4H 2 dr
if div ξ is the dependent variable.
sound wave to travel across a substantial part of the atmosphere, i.e., across
a few scale heights. The whole atmosphere then has enough time to continu-
ously adjust itself to a modified hydrostatic equilibrium, i.e., the exponential
stratification is complemented by a slowly varying (in time) contribution. The
vertical variation of such a perturbation is itself exponential, since kr2 < 0
according to (5.34). The corresponding intermediate frequency region of the
kh , ω-plane is called the evanescent-wave region.
In the evanescent region there is a particular mode for which div ξ = 0. In
general this additional constraint is not satisfied simultaneously with (5.27)
and (5.28), but for a special frequency, namely
ω = gkh (5.37)
this is possible because those two equations coincide. Equation (5.37) char-
acterizes the dispersion of surface waves in deep water. This wave mode is
marked in Fig. 5.12 by the dashed parabola. In the diagnostic diagrams (Figs.
5.4, 5.6, and 5.7) it appears as the lowest ridge, the fundamental, or f mode.
In the third, low-frequency, region of the kh , ω-plane kr2 is again positive,
so that propagating waves are possible. For small ω, and fixed non-zero l
(or kh ), the dispersion relation simplifies to
Sl2 N 2
2
kr = 2 −1 (5.38)
c ω2
or [for large l, by (5.26) and (5.30)], to
kh2
ω2 = N 2 ≡ N 2 sin2 Θ . (5.39)
kr2 + kh2
Waves with this dispersion relation are known as internal gravity waves. Their
frequencies depend only on the angle Θ between the propagation vector and
the vertical; they cannot propagate exactly in the vertical direction. Inspec-
tion of the dispersion relation (5.34) shows that the region of internal grav-
ity waves, or g modes, is bounded by ω = N for large l, and by the line
ω = Sl N/ωA for small l. The Brunt-Väisälä frequency thus turns out to be
the critical frequency for internal gravity waves. In particular, it is necessary
that N 2 > 0. We shall see in the Chap. 6 that this is also the condition of a
stable stratification in the star. The larger N 2 , the more solid is the stability,
and the higher are the frequencies of the internal gravity waves.
Problem 5.10. Consider a cool layer embedded between two hot zones. Cal-
culate the cutoff frequencies, and choose a mode which can propagate in
the two hot zones, but is evanescent in the cool layer. Compare this mode
to a particle which tunnels through a potential barrier.
Fig. 5.13. Critical frequencies in the Sun, according to (5.24), (5.30), and (5.35).
The dashed curves show Sl2 for l = 2, 20, and 200. Courtesy M. Knölker
202 5. Oscillations
Fig. 5.14. Depth of internal reflection, accord- Fig. 5.15. Acoustic ray paths
ing to (5.40), as a function of degree l, for p of two solar p modes with dif-
modes with oscillation frequencies 1.5, 2.25, 3, ferent depths of reflection in
3.75, and 4.5 mHz (from above). The curves the solar interior. The dashed
are calculated for the standard solar model of circle marks the base of the
Chap. 2 convection zone
Now, by (5.40), rt depends on l and ω only in the form l(l + 1)/ω 2 , the same
combination which appears on the right of (5.42). The integral, therefore, is a
function of l(l + 1)/ω 2 . Application of (5.26), at r = r , then yields Duvall’s
law.
Regions for Internal Gravity Waves. As illustrated in Fig. 5.13, there
are two regions where N 2 > 0 and, therefore, internal gravity waves can
propagate: the atmosphere and the radiative core. In the convection zone,
where N 2 < 0, these waves are evanescent. In the core below the convection
zone internal gravity waves are trapped and thus form a discrete frequency
spectrum of g modes. But the existence of g modes on the Sun has not yet
been confirmed; at ν = 200 μHz, upper limits are 1 cm/s in velocity and
5 × 10−7 in relative intensity variation (Appourchaux et al. 2000).
204 5. Oscillations
On the other hand, the phase diagram, Fig. 5.3, indicates the existence
of internal gravity waves in the solar atmosphere. The signature of these
waves is the negative phase difference at low frequencies, as first observed by
Schmieder (1976). To see this, we use (5.26) to rewrite the dispersion relation
(5.34) for the isothermal atmosphere in the form
kh2 (ω 2 − N 2 ) kr2 1
+ = 2 , (5.43)
ω (ω − ωA ) ω − ωA
2 2 2 2 2 c
which, for constant ω 2 , is a quadric surface in k space. The vector of phase
propagation, k, is the radius vector. The group velocity is the gradient of ω
in k space, and is perpendicular to the surfaces ω 2 = const.
In the region of propagating acoustic waves (Fig. 5.12) the surfaces (5.43)
are oblate ellipsoids of revolution with respect to the kr axis, because ω 2 > ωA
2
2 2
(and ω > N ). In this case the vertical components of phase velocity and
group velocity have the same sign. The propagating waves can form standing
waves (with almost no phase difference as in the range 2–5 mHz of Fig. 5.3),
or they will have an upwards propagating phase, such as for ν > 5 mHz in
that figure.
By contrast, (5.43) represents a one-shell hyperboloid of revolution for in-
ternal gravity waves, where ω 2 < N 2 (and ω 2 < ωA 2
). Now the r components
of the phase and group velocities have different signs. An internal gravity
wave excited from below, with upwards propagating energy, will therefore
exhibit a downwards propagating phase. In Fig. 5.3, at frequencies around
1 mHz, such downward phase propagation is seen; these waves are therefore
interpreted as internal gravity waves in the solar atmosphere.
Problem 5.11. Draw the surfaces ω 2 = const. for acoustic and internal grav-
ity waves. Mark the vectors of the phase and group velocities.
The Surface. The Sun has an extended atmosphere; its corona reaches far
into interplanetary space. In principle it would be necessary to treat the oscil-
lations, either propagating or evanescent, in this entire atmosphere, but this
is not practical for obvious reasons. Somewhere we must place a boundary on
the equilibrium model and impose boundary conditions on the perturbations.
We shall see that these conditions depend on the type of boundary chosen
for the solar model itself.
One possibility is to treat the atmosphere as isothermal and infinite, and
to utilize the acoustic cutoff, ωA , in the atmosphere. The atmospheric disper-
sion relation (5.34) gives kr = ±iα, with positive α; (5.31) and (5.32) then
tell us that the boundary condition must consist in the choice kr = +iα,
so that the exponentially growing contribution is discarded. Hence it follows
from (5.31) that
1/2 1/2
d(ξr ρ0 )/dr = −αξr ρ0 (5.45)
ne is called the effective polytropic index. Some of the coefficients of the oscil-
lation equations have a singularity at the surface of this model. In particular
we have N 2 ∝ (r − r)−1 . Therefore, the boundary condition to be imposed
on the oscillations simply is that all perturbations remain finite at r = r .
This approach has proved useful in analytic work, in particular in combina-
tion with the Cowling approximation. The asymptotic results to be described
in Sect. 5.2.6 below are an example.
A third method is to cut off the atmosphere of the solar equilibrium model
at a finite density. The surface of a model of this type would resemble a water
surface, where the pressure falls off to zero. The boundary condition then is
that the perturbed surface is also a zero-pressure surface. Hence we must
impose
δP = 0 . (5.47)
This boundary condition has been used in most numerical calculations.
Since the full set of (5.21) to (5.23) is normally solved in numerical work,
a second outer boundary condition must be imposed. To this end we require
that the internal gravitational potential smoothly matches onto the external
gravitational potential. For the latter, we must select the solution of ΔΦ1 = 0
which (for any given l) behaves like r−l−1 for r → ∞. Thus, the boundary
condition at the model surface becomes
dΦ1
r + (l + 1)Φ1 = 0 . (5.48)
dr
We may wonder which of the three types of boundaries best represents
the Sun. For the determination of eigenfrequencies alone it probably does not
matter too much, if only the place of the boundary is chosen at the right level
in the atmosphere. As test calculations demonstrate, the boundary should be
placed at optical depth τ = 10−4 , or even further outwards. The model eigen-
frequencies then become more or less independent of the place and type of
boundary condition. This is in particular true for the lower overtones (smaller
frequencies), and the reason is of course that these are already reflected by
their acoustic cutoff at some internal level, so that the actual boundary condi-
tion, placed in the evanescent region, has little effect: it must merely suppress
the exponentially growing contribution to the eigenfunction.
The Center. Although the center of the Sun does not represent a physical
surface, we must impose a boundary condition there. This is a consequence
of the use of spherical polar coordinates, which introduces a formal singular-
ity at r = 0 into the oscillation equations (5.21) to (5.23). As in the case of
the singular surface, the boundary condition is that all perturbations remain
finite at r = 0.
5.2 Linear Adiabatic Oscillations of a Non-Rotating Sun 207
Problem 5.12. Expand the diverse perturbations into Taylor series around
r = 0 and show that, for l = 0,
ξr ∝ rl−1 , P1 ∝ rl . (5.49)
2
P1 P1 ω − Sl2 P1
+ (A + B) + 2
+ BA + A
ρ0 ρ0 c ρ0
N2 2N
= − 2 (ω 2 ξr ) + B + ω 2 ξr , (5.50)
ω N
and
!
P1 P1 N2 2
ω ξr −
2
+A = (ω ξr ) , (5.51)
ρ0 ρ0 ω2
A = −N 2 /g , (5.52)
B = 2/r − g/c2 (5.53)
have been introduced.
In the outer domain the expansion parameter is Sl2 /ω 2 ; we derive from
(5.27) and (5.28) the two equations
2
2c ω − N2 2c
ξr + A + B + ξr + + B
+ B A + ξr
c c2 c
!
Sl2 P1 2Sl P1
= 2 2 + A+ , (5.54)
ω c ρ0 Sl ρ0
and
P1 S 2 P1
+ c2 (ξr + Bξr ) = l2 . (5.55)
ρ0 ω ρ0
In each domain we now have a second-order differential equation. The terms
on the right are multiplied by the respective (small) expansion parameter,
while the left-hand sides of (5.50) and (5.54) are rather similar to the equa-
tion that governs Bessel functions. In particular, note the singularities in the
coefficients of the undifferentiated variables P1 /ρ0 and ξr . Bessel functions
are therefore appropriate for the expansion of P1 /ρ0 and ξr , and the correct
choice of their index takes care of the singularities. The leading terms of these
expansions are, in the inner domain
1/2 1/2
P1 c vi 1
= ki Jl+1/2 (ωvi ) − Fi (ui )Jl+3/2 (ωvi ) (5.56)
ρ0 ρ0 r ω
and
1/2
vi
ω 2 ξr = −ki ωJl+3/2 (ωvi )
r(cρ0 )1/2
!
l+1
+ Hi (ui ) − Jl+1/2 (ωvi ) , (5.57)
ui
where
r
dr
ui = vi = , (5.58)
c
0
Fi and Hi are such that (5.50) and (5.51) are satisfied, and ki is a constant.
In the outer domain the expansions begin with
1/2
vo 2
ξr = ko Jne (ωv o ) − F (u )J
o o ne +1 (ωv o ) , (5.59)
r(cρ0 )1/2 ωvo
5.2 Linear Adiabatic Oscillations of a Non-Rotating Sun 209
and
1/2 1/2
P1 c vo
= −ko ωJne +1 (ωvo )
ρ0 ρ0 r
!
ne + 1 vo
+ Ho (uo ) − Jn (ωvo ) , (5.60)
2uo 2 e
where ne is the polytropic index at the surface (cf. the preceding section), ko
is another constant, and uo and vo are defined by
r
dr
2u1/2
o = vo = . (5.61)
c
r
The inner and outer domains overlap. In the interval that is common
to both domains the two expansions (5.56), (5.57) and (5.59), (5.60) must
approximate the same solution. This matching requirement firstly determines
the ratio ko /ki (one of the two constants remains arbitrary since the whole
problem is linear and homogeneous), and secondly yields the desired condition
for the eigenfrequencies ω. For large arguments, i.e., at large distance from
the respective boundaries, the Bessel functions can be represented in terms of
trigonometric functions. If this is done, the matching condition is reduced to
the form cos(q(ω)) = 0, so that q = π(2n − 1)/2 with an integer n; excluding
terms of order ω −2 or smaller, q(ω) is given by
3 π 1
q(ω) = ω(vi + vo ) − l + ne + − Vio . (5.62)
2 2 ω
We write ω = 2πνn,l , substitute vi and vo , and finally obtain
l ne 1 Vio
νn,l = Δν n + + + + 2 , (5.63)
2 2 4 2π νn,l
where we have defined the asymptotic frequency splitting
⎛ r ⎞−1
dr ⎠
Δν = ⎝2 . (5.64)
c
0
radial oscillations (l = 0), and Vandakurov (1967) first considered the general
case. Evaluation of (5.64) yields
Δν ≈ 136 μHz , (5.65)
twice the spacing of the main peaks in the power spectrum of low-l p modes. It
is now clear why this value was chosen as bandwidth of the spectral sections in
the echelle diagram, Fig. 5.10. The line identifications on the top of Fig. 5.8
also can be obtained from (5.63) to (5.65) (or, of course, from numerical
calculations).
The close coincidence of the frequencies νn,l and νn−1,l+2 seen in the
observational results now turns out to be a degeneracy in the leading term of
(5.63). It is worth mentioning that the asymptotic frequencies do not depend
on the choice of the matching point. As Tassoul (1980) shows, this is even
true for the second-order correction Vio . Due to this correction the degeneracy
is not exact; Vio is related to the small separation
δνn,l = νn,l − νn−1,l+2 . (5.66)
Various estimates have been made at δνn,l , which all depend on the derivative
of the sound velocity. Christensen-Dalsgaard (1998) gives
r
(4l + 6)Δν 1 dc
δνn,l − dr . (5.67)
4π 2 νn,l r dr
0
Chap. 2 yields T0 = 35.4 min; the model of the 1989 edition of this book gave
T0 = 34.8 min, and model 1 of Christensen-Dalsgaard (1982) [as quoted by
Gough (1984 b)] T0 = 34.5 min. The differences in the depth of the convection
zone, i.e., in the value of rv , apparently do not matter greatly because in the
vicinity of rv the Brunt-Väisälä frequency N , and therefore the contribution
to the integral in (5.68), is small.
5.3 Helioseismology
The measurement and interpretation of travel times of earthquake signals
traditionally has been used to study the Earth’s interior. In addition to this
classical method, the measurement of free oscillations has become possible
since the great Chilean earthquake of May 22, 1960. Mostly this more recent
branch of seismology has been applied in solar (and indeed stellar) research,
although the travel-time method also has its solar analogue.
ξ = (r/c)2 , (5.73)
ξ
1 dr
F (u) = (ξ − u)1/2 dξ , (5.74)
r dξ
u
or
214 5. Oscillations
ξ
2 dF/du
r = r exp − du . (5.77)
π (u − ξ)1/2
ξ
This is an implicit equation for ξ(r) and therefore, by (5.73), for c(r). The
important point is that c(r) can be determined from (5.77) and the known
function F (u) without recourse to a calculated solar model.
The result of the original inversion is shown in Fig. 5.17. In the range
0.4r < r < 0.9r it agrees well with the theoretical model (model 1 of
Christensen-Dalsgaard 1982). Outside that range this inversion is not reliable,
as was shown by inverting the eigenfrequencies of the model itself. One rea-
son is that (5.42) was based on the local dispersion relation ω 2 = c2 (kr2 + kh2 )
which breaks down near the reflecting boundaries. Another reason, responsi-
ble for the uncertainty below 0.4r , is that not enough frequencies of modes
penetrating deep into the core were available.
Figure 5.18 shows a more recent result, obtained numerically according to
the recipes outlined above: the difference δc2 (r) between the squared sound
velocities of a seismic and a standard solar model. Although this difference is
remarkably small, there are significant excursions: a peak just below the con-
vection zone, and a dip around r/r = 0.2. Both the peak and the dip occur
in regions where the mean molecular weight μ varies, due to the gravitational
settling of helium and to the hydrogen burning in the core, and it has been
found that smoothing the gradient of μ, e.g., by some mild mixing, would
reduce the excursions (Christensen-Dalsgaard 1997). In the central region,
where the accuracy is least, the sound-speed deviation is still less than 1%.
The temperature of the solar center cannot be determined with the same
precision as the sound speed, because of its dependence on the mean molec-
ular weight: T ∝ μc2 . Error estimates of ≈ 2% have been obtained by Antia
and Chitre (1995) and by Takata and Shibahashi (1998) in a combined evalu-
ation of helioseismic results and solar model calculations. The uncertainty lies
both in the abundance of the elements and in the opacity. However, a central
temperature of 1.57 × 107 K with an uncertainty of 2% is still a sufficiently
severe constraint to dismiss the non-standard solar models of Sect. 2.5.
Fig. 5.18. Relative difference between the squared sound speed as inferred from
2 months of MDI data and the standard solar model of Christensen-Dalsgaard et
al. (1996). Vertical bars are error estimates, horizontal bars show the resolution in
depth. From Kosovichev et al. (1997)
decreases with increasing α. As Ulrich and Rhodes (1977) point out, it is the
entropy which physically characterizes the convective solar envelope, while α
merely serves as a technical parameter.
The entropy excess ΔS can also be viewed as the entropy “jump” across
the outermost layers of the convection zone, because only there ∇ − ∇a is
substantial, cf. Table 6.1. The fact that theoretical models tend to predict
too large frequencies of high-degree p modes therefore means: the theory of
solar convection, as outlined in Sect. 6.2, somewhat overestimates the entropy
jump ΔS.
Unfortunately, α is a free parameter only for an envelope model, which
is not subject to central boundary conditions, but not for a full solar model.
As explained in Chap. 2, α is calibrated (together with the original helium
abundance, Y0 ) by the requirement that the model of the present Sun must
have the observed radius (and luminosity). Hence we can only vary a third
quantity, and hope that the calibration after such a variation leads to an
increased α. Partial success in this direction has come from the electrostatic
correction in the equation of state, see Sect. 5.3.5 below, and – to a larger
extent – from more detailed opacity calculations (the OPAL tables, Sect.
2.3.7). For example, the standard model of the present text requires α = 1.81,
as compared to α = 1.38 for the model of the 1989 edition. Accordingly its
convection zone is deeper, with rv /r ≈ 0.708 (compared to 0.74 of the earlier
model), very close to the result (5.80) of the helioseismic inversion.
In Fig. 5.17 the base of the convection zone can be recognized by a slight
change of the slope of c2 (r) near r = 0.7r . We differentiate c2 (r), given by
(5.12); in order to simplify the argument we use the perfect-gas law P =
ρRT /μ, with μ = const. and Γ1 = 5/3 (ionization is unimportant near the
base of the convection zone). Then
dc2 5R dT 5Gm
= =− 2 ∇, (5.79)
dr 3μ dr 3r
where ∇ = d ln T /d ln P , and where the hydrostatic equilibrium condition has
been used in the second relation. Since ∇ ∇a ≈ 0.4 in the deep part of the
convection zone, the expression W = (r2 /Gm) dc2/dr is approximately −2/3
there. In the radiative zone the absolute magnitude of W is smaller. Hence the
value of rv can be determined as the radius in the seismic solar model where
W (r) reaches the constant −2/3. In this manner Christensen-Dalsgaard et
al. (1991) and Basu and Antia (1997) obtained
with errors of ±0.003 and ±0.001, respectively. With the smaller error, this
translates into (199 700 ± 700) km for the depth of the convection zone.
5.3 Helioseismology 217
Fig. 5.20. Echelle diagram with p modes for l = 0 and l = 1. The frequencies are
ν = ν0 + ν̂. Circles: observations, from Jiménez et al. (1988), cf. Table 5.1; solid and
dashed lines: theoretical results of Noels et al. (1984), for Z = 0.02 and Z = 0.018,
respectively
Z than observed was allowed because of the possible accretion of dust at the
solar surface). We must therefore find other means to repair the discrepancies
of up to ≈ 10 μHz visible in Fig. 5.20. One is the equation of state.
The lowest ridge in the diagnostic diagram of solar oscillations is the f mode.
This mode has no nodes in the radial direction. It resembles √ a surface wave
on deep water, with dispersion according to (5.37), or ν = gkh /2π. The
horizontal wave number
is asymptotically related to the spherical harmonic
degree l through kh l(l + 1)/r. Because of the simple dispersion relation
the f-mode frequencies are essentially independent of the internal structure of
the Sun, except for the radius. The mass variable m(r) introduced in Chap. 2
varies very little in the observable layers, so we may use g = Gm /r2 and see
that ν ∝ r−3/2 . Since the frequencies can be determined with high precision,
and since the product Gm is known from the dynamics of the planetary
system, the radius can be determined (Schou et al. 1997, Antia 1998).
In principle such a radius determination could be more precise than the
common procedure where the apparent angular diameter is measured and the
radius is calculated from the Sun’s distance (Sect. 1.3). The problem with the
f-mode method is only that the depth in the solar atmosphere that marks the
radius obtained in this way must be defined precisely. It has been practical
not to use the dispersion relation but instead to calculate solar models and
their f-mode eigenfrequencies for different model radii. If the model radius
is fixed at optical depth τ = 2/3, the result is, with a conservative error
estimate, rτ =2/3 = (6.957 ± 0.001) × 108 m.
+2iωα ρ0 ξ ∗ · (v 0 · ∇)ξ dV . (5.86)
Radial Shear. Let us first consider the case where Ω depends only on the
radial coordinate. We substitute (5.19) into (5.88) and obtain, after a few
integrations by parts
r $ %
ρ0 Ω |ξr − ξh |2 + [l(l + 1) − 2]|ξh |2 r2 dr
0
Δωα = −m r & '
ρ0 |ξr |2 + l(l + 1)|ξh |2 r2 dr
0
r
≡ Kα (r)Ω(r) dr (5.89)
0
(Gough, 1981). The rotational splitting kernel, Kα (r), depends on the eigen-
function with index α (which was omitted for brevity). Therefore, any given
Δωα samples the angular velocity in the depth range corresponding to ξ α .
The integrals appearing in (5.89) are independent of m. Hence, in the
case of purely radial shear, Δωα is strictly linear in m. Since −l ≤ m ≤ l, we
obtain a multiplet of 2l + 1 frequencies, with equidistant spacing.
The modes with m = ±l have the largest frequency shifts. These modes
have a sectoral pattern, as shown in Fig. 5.5 for the case m = l = 10. Ob-
servationally, the sectoral modes have been isolated by Duvall and Harvey
(1984) by means of the same cylindrical lens that had been used to isolate
the zonal (m = 0) modes, cf. Sect. 5.1.2: here the lens is oriented paral-
lel to the solar axis of rotation, so that the image is optically averaged in
north-south direction. The entrance slit to the spectrograph lies parallel to
5.3 Helioseismology 223
Fig. 5.23. Power spectra for p modes with l = 3, m = +3 (a) and m = −3 (b).
The labels are (l, n); subscripts u and l mark the upper and lower side lobes arising
from night gaps in the data (Duvall and Harvey 1984). Courtesy National Solar
Observatory and NASA
Problem 5.18. Show that the amplitudes of sectoral surface harmonics are
very small except in a band along the equator, of width l−1/2 in latitude.
Sectoral modes with large l therefore essentially sample the equatorial fre-
quency, νrot (r, π/2), of solar rotation.
224 5. Oscillations
where νrot = Ω/2π, and Kα is readily obtained from (5.89) and (5.91). For
any fixed depth r0 we now take a linear combination
D(r, r0 ) = aα (r0 )Kα (r) . (5.93)
α
Optimal kernels for the internal solar rotation have indeed been calculated;
examples are shown in Fig. 5.24. In the outer part of the Sun, the optimal
kernels are predominantly combinations of p-mode kernels. In the inner part,
p modes of low degree are of some value, but the inclusion of g-mode kernels
would considerably improve the depth resolution.
Unfortunately the linear combinations of observed splittings on the right
of (5.95) involve much cancellation, and hence an increase of relative errors.
Thus, there is a tradeoff between resolution in depth and attainable accuracy.
(Duvall et al. 1986). Here Δνα = Δωα /2π, Pi is the ith Legendre polyno-
mial, and L = [l(l + 1)]1/2 . An expansion in powers of m would be equivalent,
but (5.96) is preferred because the Pi are orthogonal, and the coefficients
ai , therefore, independent (actually Clebsch-Gordon coefficients are used be-
cause orthogonality is required for a discrete data set, cf. Ritzwoller and
Lavely 1991). As it turns out the argument m/L takes care of most of the
where Δ is the distance between two points r 1 and r 2 on the solar surface,
f (r i , t) is an oscillation signal at point r i , and τ is the time delay; the integral
is taken over the time of observation. The signal f can be the velocity or the
intensity variation. For a given Δ the covariance function attains maxima for
certain delay times τ , namely for those that correspond to the travel times
of the signal from r 1 to r 2 (or vice versa) via acoustic rays with 1, 2, . . .
reflections at the base of their respective cavities. Figure 5.15 shows examples
5.3 Helioseismology 227
of such rays, and Fig. 5.14 gives the depth of reflection as a function of degree
l and frequency ν.
The travel time of an acoustic wave varies due to the variation of the
sound speed c(r) along the path, and due to local flows with velocity v(r):
ds
τ= , (5.99)
c(r) + v(r) · n(r)
where s is a coordinate along the path from r 1 to r 2 , and n(r) is a unit
vector tangent to the path. We write c(r) = c0 (r) + δc(r) and consider the
case where both δc(r) and v(r) · n(r) are small in comparison to c0 (r), the
sound speed of a horizontally uniform reference state. Then
ds δc(r) + v(r) · n(r)
τ= − ds . (5.100)
c0 (r) c20 (r)
Here we can take the integral along the path that would be traveled in the
reference state because, according to Fermat’s principle, τ is an extremum,
and first-order perturbations of the path do not contribute to the variation
of τ . Equation (5.100) is an inversion problem: we want to know the functions
δc(r) and v(r) that appear in the integrand. This time there is no explicit
parameter dependence. Nevertheless a solution is possible because a large
number of travel times and wave paths are evaluated at the same time.
As first pointed out by Gough and Toomre (1983), the effects of the sound-
speed and flow variations can be separated by measuring the travel times τ +
and τ − of signals traveling in opposite directions between two points. For,
these two travel times respond in the same way to the variation δc, but in
opposite ways to the local flow v. Thus we have two inversion problems:
1 + −
δc(r)
τ +τ =− ds (5.101)
2 c20 (r)
and
1 + v(r) · n(r)
τ − τ− = − ds . (5.102)
2 c20 (r)
The time-distance technique has been applied to the supergranulation,
which is seen at the solar surface as a network of convective cells of 20 000–
30 000 km extent. The result (Fig. 5.26) indicates that the flow structure
continues several thousand km below the surface, and that associated tem-
perature variations of up to 2%, positive in the upflow and negative in the
downflow regions, confirm the convective nature of the supergranulation (cf.
Sect. 6.5.2).
The temperature variation deduced for the supergranular pattern rests
on the relation δT /T 2δc/c. In general, however, the sound speed may be
perturbed additionally by a magnetic field. This complicates the inversion,
but also opens the possibility of inferring the magnetic field underneath the
Sun’s surface.
228 5. Oscillations
Fig. 5.26. A vertical cut showing the supergranular flow and the associated temper-
ature variation below the Sun’s surface (the vertical scale is stretched by a factor
of ≈ 16). Time-distance inversion of data obtained with the Michelson Doppler
Imager on SOHO. After Duvall et al. (1997)
where the angular brackets denote an azimuthal average over an annulus with
radius Δ, and where the sum is taken over all annuli within the observed area
(the “aperture”), with weighting factors w(Δ) that express the change of
amplitude during the transit of each contributing wave along its path. Each
contribution is taken at time t + τ (Δ), i.e., retarded by a travel time that
can be determined by a correlation analysis if r 1 is on the visible surface,
but otherwise must be calculated from a solar model and acoustic ray theory.
According to (5.40), modes with equal l/ν have approximately the same
depth of internal reflection, and hence approximately the same travel time
for a given distance Δ.
The acoustic signal can be reconstructed at all points within a given
section of a plane or of a volume. Using data from the TON network, Chou
et al. (1999) derived an acoustic image of an active region extending several
ten thousand kilometers below the surface. With measurements of MDI on
SOHO, Lindsey and Braun (2000) were able to image an active region on the
far side of the Sun; for this purpose they used acoustic signals having one
additional reflection from the solar surface between the points of measurement
and the target.
5.4 Excitation and Damping 229
We first mention the κ mechanism, which has been identified as the driving
agent of stellar pulsations, although it is probably less important in the solar
case. The κ mechanism acts like a valve: suppose that in a phase of compres-
sion the opacity increases. The compressed layer then absorbs energy out of
the radiative flux toward the stellar surface, and thus will be heated in excess
of the mere adiabatic heating. The subsequent expansion will be stronger than
the preceding one was. Zhevakin (1953) demonstrated that this mechanism
of overstability drives the pulsation of δ Cephei and related variable stars,
where it is particularly effective in the layer of the second helium ionization.
The crucial parameter measuring the opacity variation is
∂ ln κ
κT ≡ (5.104)
∂ ln T P
and is shown in Fig. 2.8. It has a high maximum in the layer of partial hy-
drogen ionization. In this layer there is strong driving, but we must include
the contributions from all layers in order to see whether a particular mode
is excited or damped. Indeed Ando and Osaki (1975) found positive growth
rates for a wide range of solar oscillations with periods around 5 min. On
the other hand there is not only excitation, but also damping, by radiative
losses in the optically thin atmosphere and by interaction with the instation-
ary convective motion. Hence the overstabilities could not be confirmed by
subsequent studies of Goldreich and Keeley (1977 a) and Balmforth (1992).
There is another reason why the excitation of solar p modes by means
of the κ mechanism appears unlikely. The excited (or damped) oscillator is
symbolically represented by an equation of the form
ξ¨ − 2γ ξ˙ + ω02 ξ = 0 , (5.105)
where the net effect of all excitation and damping yields the coefficient γ.
Now suppose excitation wins for a particular mode, that is, γ is positive. Then
there is unlimited growth of this mode, because (5.105) is homogeneous and
linear. It is only by means of the non-linear terms, which have been neglected
in (5.105) as well as in our full system of oscillation equations, that the growth
could be held. But before these terms can take effect the amplitude should
be sizable, unlike the small amplitudes observed on the Sun. Cepheids are
different: their pulsation amplitude, of order δr/r ≈ 0.1 or larger, is indeed
limited by non-linear effects.
230 5. Oscillations
Fig. 5.27. Excitation of solar oscillations in a model similar to that of Kumar and
Lu (1991). Left: Potential V (r), with step at r = a, and point source at r = r0 ;
right: power at r = a as a function of frequency, with resonances below and above
the acoustic cutoff νA (the maxima below ν = νA have been cut at finite hight)
232 5. Oscillations
Kumar (1994) concluded that the solar oscillations are excited in a layer only
about 100–200 km below the surface.
A close inspection of Fig. 5.27 shows a slight asymmetry of the peaks in
the calculated power spectrum. This asymmetry is caused by the diverse in-
terferences of the excited and reflected waves; it was also noticed by Gabriel
(1992), Nigam and Kosovichev (1998), and Rosenthal (1998), and contributes
to the line asymmetry first discovered by Duvall et al. (1993 a). An additional
contribution comes from the coupling of p modes with nearly equal frequen-
cies by a large-scale convective flow (Roth and Stix 1999).
Although the average depth of the acoustic excitation may be several
100 km below the surface, individual sources of acoustic waves have been
localized at the surface itself. These sources lie preferentially in the dark
narrow downstream regions in between the bright granules (Brown 1991,
Rimmele et al. 1995, Espagnet et al. 1996, Kiefer et al. 2000 b, Khomenko
et al. 2001). In general the velocity in these downstream regions is higher
than the velocity within the granules. It appears that the local sources, also
called “acoustic events”, feed sufficient energy into the solar oscillations as to
compensate for the energy loss by interaction with small-scale motion or by
leakage (Goode et al. 1998, Nigam and Kosovichev 1999, Strous et al. 2000).
We may ask why acoustic oscillations have not yet been observed below,
say, 1 mHz, although p modes with still lower frequencies are possible accord-
ing to theory. Here we must distinguish between the high-l and low-l cases.
At high l, there are no eigenmodes at smaller frequencies, i.e., we see the
smallest possible radial order n, including the fundamental, n = 0. At low
l, the oscillations of low frequency have their upper reflection at a level that
lies so deep below the photosphere that in the observable layer the amplitude
has become too small to be detected with present methods (we recall that
all observed oscillation signals emerge from the region of evanescent waves,
i.e., from within the “tunnel”). The lowest mode of radial pulsation, with
a period of slightly more than 1 hour, and a number of other low-l, low-n
modes very probably are not yet detected for this reason.
results of the local analysis are described in the classic treatise of Eckart
(1960) and by Moore and Spiegel (1964). The properties of internal gravity
waves are also discussed in Landau and Lifschitz (1966 b, p. 49).
Much of the material on helioseismology is collected in conference proceed-
ings, such as those edited by Hill and Dziembowski (1980), Gough (1983 a,
1986), Belvedere and Paternò (1984), Noels and Gabriel (1984), Ulrich (1984),
and Christensen-Dalsgaard and Frandsen (1988). Some of these conferences
are built around the results of major experiments like GONG and SOHO
(Brown 1993, Ulrich et al. 1995, Hoeksema et al. 1995, Korzennik and Wil-
son 1998, Wilson 2001), others concentrate on special topics such as the
convection-oscillation interaction (Pijpers et al. 1997) or the helioseismic
diagnostics of convection and activity [Vols. 192 and 193 (2000) of Solar
Physics].
Duvall et al. (1988) and Libbrecht and Kaufman (1988) present tables of
observed oscillation frequencies which are more exhaustive than Table 5.1.
Frequencies and mode spectra, as well as beautiful illustrations of various
results, can also be found on the Internet pages of the mentioned projects.
Helioseismic results concerning the standard solar model – equation of
state, convection-zone depth, diffusion of elements, etc. – have been ob-
tained also by Kim et al. (1991), Christensen-Dalsgaard and Däppen (1992),
Berthomieu et al. (1993), Christensen-Dalsgaard et al. (1993), Guzik and Cox
(1993), and Guenther et al. (1996).
Lynden-Bell and Ostriker (1967) derived the variational principle for the
case of rotation, a field treated also in the monograph of Tassoul (1978).
Sobouti (1980) calculates eigenfrequencies of rotating stars by an expansion of
the eigenfunctions in terms of those of the non-rotating star; Clement (1986)
employs the variational principle to calculate the eigenfrequencies. A general
approach, based on perturbation theory, to treat the effects of rotation and
of a large-scale flow has been developed by Ritzwoller and Lavely (1991) and
Lavely and Ritzwoller (1992, 1993). Schou et al. (1998) discuss the inversion
of the rotational frequency splitting, and show results based on data from
SOHO/MDI.
In addition to the permanent excitation by turbulent convection, flares
and coronal mass ejections (Chap. 9) can lead to occasional seismic events
on the Sun. Kosovichev and Zharkova (1998) describe a case where propa-
gating circular wave packets were observed through their Doppler effect for
35 minutes after a flare.
Damping of solar oscillations by the turbulent velocity field of the solar
convection zone has been discussed in terms of viscous damping (Gold-
reich and Keeley 1977 ab, Stix et al. 1993). Amplitudes and damping rates
of stochastically excited solar and stellar oscillations have been estimated
by Houdek et al. (1999). In addition to the damping effect, the turbulence
also slightly shifts the p-mode frequencies as it changes the mean structure
near the surface of the solar model (Christensen-Dalsgaard and Thompson
234 5. Oscillations
The energy generated in the solar core must be carried to the surface. In
Chap. 2 radiation was treated as one possible way of energy transport. An-
other is convection, where internal energy, sometimes including latent heat,
is carried along with the motion of matter. We know that convection is quite
efficient in distributing heat in rooms, and we shall see in the present chap-
ter that convection is also an efficient carrier of solar energy, at least in the
outer layer between a depth of ≈ 200 000 km and the surface. The stratifica-
tion of this outer layer is unstable, much like the stratification of the Earth’s
troposphere which becomes unstable when the ground is heated.
The mixing-length concept, where parcels of gas are envisaged to travel
a certain distance and then to dissolve and to deposit their excess heat, will
be used to describe the convective flux of energy.
As opposed to oscillatory motions, where parcels of gas move back and
forth, the convective motion normally is overturning, although not neces-
sarily stationary. We shall consider the overturning character as typical for
convection. Thus, the cellular structure seen at the solar surface, from the
small-scale granular velocity field to the pattern of giant cells, will be treated
as a convective phenomenon in the present chapter. This widens the mean-
ing of the word “convection”, because most of the observable layer is stably
stratified, and the energy transport is by radiation rather than by convection.
6.1 Stability
When a parcel of gas (also called a “bubble” or “blob”, or an “eddy” or “fluid
element”) is adiabatically lifted from its equilibrium position, it will be either
heavier or lighter than its new environment. In the latter case it continues to
rise, i.e., the original equilibrium was unstable, in the former case it returns
to its initial position, which means that the original equilibrium was stable.
A criterion for such stability or instability was first derived by Schwarzschild
(1906).
Let us consider a parcel of gas which has traveled a vertical distance δr,
cf. Fig. 6.1. The density of the parcel is now ρ∗ , and must be compared to the
density ρ∗0 of its new environment. We assume that the motion is sufficiently
fast so that the parcel behaves adiabatically, but still sufficiently slow so that
at each point its internal pressure has adjusted itself to the local ambient
pressure. These assumptions are reasonable whenever the time scale of energy
exchange is long compared to the sound travel time across the parcel; in the
optically thick solar interior this is always the case.
The density difference at the level r + δr can be expressed in terms of the
density gradient: expansion of ρ∗ and ρ∗0 yields, to first order, the condition
for instability,
dρ dρ
ρ∗ − ρ∗0 = δr − <0, (6.1)
dr a dr
where (dρ/dr)a is the density gradient under adiabatic conditions. We restrict
the discussion now to the case where the pressure is given by the perfect gas
equation, P = ρRT /μ. But we shall allow for variations of the mean molec-
ular weight μ arising from ionization and from the change of the chemical
constitution due to hydrogen burning. Making use of the pressure equilibrium
we thus obtain
dT dT T dμ T dμ
< + − . (6.2)
dr dr a μ dr μ dr a
Inside the Sun dT /dr < 0. For instability to occur the temperature gradient
must be steeper, i.e., more negative, than the adiabatic temperature gradient
plus the corrections arising from the change of μ.
In the solar core the material is fully ionized, and we may assume that it
remains so under an adiabatic perturbation. Then (dμ/dr)a = 0. The gradient
dμ/dr is the result of helium accumulation towards the center and is negative.
Therefore this gradient is stabilizing. This is of course expected since helium
is heavier than hydrogen. However, the μ gradient is not really needed for
stabilization of the Sun’s core. The models actually calculated have a stable
core even if the μ gradient in (6.2) is neglected. Of course, there might be other
sources of instability: a magnetic field, or a large gradient in angular velocity,
cf. Chap. 7. In a discussion of such phenomena it must not be forgotten that
the molecular weight variation always acts against the instabilities.
In the outer layers of the Sun – the “envelope” – the mean molecular
weight varies because the abundant elements hydrogen and helium are par-
tially ionized. In this case μ is a function of P and T . Let us assume that
within the upwards traveling parcel there is instantaneous adjustment of the
6.2 Mixing-Length Theory 239
ionization equilibrium. Then the molecular weight in the parcel is the same
function of P and T as the exterior μ, that is,
dμ ∂μ dP ∂μ dT
= + , (6.3)
dr ∂P T dr ∂T P dr
and
dμ ∂μ dP ∂μ dT
= + . (6.4)
dr a ∂P T dr a ∂T P dr a
We substitute these expressions into (6.2), recall that we have assumed pres-
sure equilibrium, and find the Schwarzschild criterion for convective instabil-
ity of the solar envelope,
dT dT
< . (6.5)
dr dr a
The molecular weight does not enter explicitly because the factor δ ≡ 1 −
(∂ ln μ/ ln T )P , cf. (2.51), appears on both sides and is positive. Instead of
(6.5) one often writes
∇ > ∇a , (6.6)
where ∇ = d ln T /d ln P , and ∇a is the adiabatic value of ∇, defined in (2.49).
Notice that these double-logarithmic temperature gradients must not be con-
fused with the symbol ∇ that abbreviates the div, grad, or curl operators.
The criterion for instability is deduced here without recourse to the hy-
drodynamic equations governing the problem. A linear stability analysis of
those equations leads however to the same result (Lebovitz 1966).
ΔT = (∇ − ∇ )T α/2 . (6.11)
The second of these relations follows from the assumption of pressure equi-
librium, ΔP = 0, between the parcel and the ambient gas; the factor δ, given
by (2.51), takes care of a possible variation of the mean molecular weight μ.
We substitute (6.10) and assume that ∇, ∇ , δ, g, and HP all are constants
over a mixing length. The integral of (6.13) then yields
gδ
(∂δr/∂t)2 = (∇ − ∇ )(δr)2 , (6.14)
HP
which means that the work done by the buoyancy force appears as kinetic
energy of the parcel. It has been customary to multiply the right of (6.14) by
1/2 in order to account for some expenditure of buoyancy work due to friction.
Then, using again l/2 for the mean of δr, we obtain the mean convection
velocity
1/2
gδ
v=l (∇ − ∇ ) . (6.15)
8HP
Radiative Loss. During its rise the parcel radiates energy into its environ-
ment. For this reason the gradient ∇ differs from the adiabatic gradient ∇a .
In order to assess ∇ we write for the radiative flux across the surface of the
parcel
16σT 3 ΔT 8ασT 4
FR = − = (∇ − ∇) , (6.16)
3κρ d 3κρd
where ΔT has been substituted by (6.11); d is the distance over which ΔT
drops to zero (essentially the parcel’s diameter).
The convective flux (6.12) can be written in the form
The first term is the convective flux which we would have under ideal adi-
abatic conditions, the second (negative) term determines how much smaller
than this the real convective flux is. We must multiply this second term by an
242 6. Convection
effective cross section q of the parcel, and we must multiply FR by the par-
cel’s surface S. Both results express the same quantity, namely the radiative
loss per unit time. Hence
16σT 3 S
(∇ − ∇) = ρcP vq(∇a − ∇ ) . (6.18)
3κρd
There is some arbitrariness in this derivation concerning the quantities
S, d, and q. In order to remain consistent with Vitense’s work, we choose
Sl/qd = 9/2, corresponding, e.g., to spherical parcels and d = (8/9)l. Other
choices could be made as well, but the result would be numerical factors
of order 1 that occur in combination with the mixing length, much like the
factor 1/2 which was introduced in (6.15). In the end, the whole difference
would be a different calibration of the dimensionless parameter α in (6.7).
We substitute (6.15), and obtain
∇ − ∇a = 2U (∇ − ∇ )1/2 , (6.19)
where
√
24 2σT 3 P 1/2
U= . (6.20)
cP κgl2 δ 1/2 ρ5/2
The Cubic Equation. We return to our basic equation (6.8). We define a
“radiative” gradient
3κρHP L
∇R = (6.21)
64πr2 σT 4
which, in the absence of convection, is identical to the true gradient ∇. In
a convection zone, ∇R is the fictitious gradient that would be needed to
transport all the energy by radiation; hence it simply measures the total
energy flux.
Substitution of (6.9), (6.12), (6.15), and (6.21) into (6.8) yields
9
∇ − ∇R + (∇ − ∇ )3/2 = 0 . (6.22)
8U
This relation, in combination with (6.19), is a third-order algebraic equation
determining ∇ in a convection zone. In the code describing the internal struc-
ture of the Sun (Chap. 2) this equation is numerically solved. It yields the
convective temperature gradient which, by (2.8) and the definition of HP , is
related to ∇:
dT T∇
=− . (6.23)
dm C 4πρr2 HP
Problem 6.2. Introduce x = (∇−∇a +U 2 )1/2 , and derive the cubic equation
9
(x − U )3 + x2 − U 2 − ∇R + ∇a = 0 . (6.24)
8U
Show that this equation has only one real root.
6.2 Mixing-Length Theory 243
A Solar Model. For the standard solar model described in Chap. 2 some of
the variables in the convection zone are listed in Table 6.1. In the outer part
hydrogen is partially ionized. A large amount of latent heat (the ionization
energy) is therefore available. This effectively lowers ∇a , cf. Fig. 2.4, and
so furthers the convective instability. It also raises cP , see again Fig. 2.4,
and hence raises the efficiency of the convective energy transport. For these
reasons the convection zone of the Sun sometimes is called the “hydrogen
convection zone”. Nevertheless, the ionization of helium also contributes [ηi
in Table 6.1 is defined as the number of atoms in the (i + 1)th state of
ionization, divided by the total number of atoms of the element in question].
We also see from Table 6.1 that a very small temperature difference ΔT
and a very small superadiabaticity, Δ∇ ≡ ∇ − ∇a , are sufficient for FC to
carry the entire solar luminosity over most of the convection zone. As a
consequence, the stratification is essentially adiabatic, and the mixing-length
formalism mainly serves to select the right adiabat for the stellar model.
The convection velocity v calculated according to (6.15) is about 2 km/s
near the surface, but takes on much smaller values, 500 m/s and less, already
at small depth. The numbers in the table must however not be understood too
literally because v, unlike FC , is subject to the arbitrary factor 1/2 introduced
above.
The total depth of a convection zone calculated with the recipe outlined
here slightly depends on the physical input to the model, especially on the
opacity which enters expression (6.21) for the radiative gradient. With the
OPAL opacity table the depth is ≈ 200 000 km, consistent with the helioseis-
mological result (5.80).
Problem 6.3. Show that the convective energy flows down the gradient
dS/dr of the specific entropy. Give an interpretation in terms of a diffusion
process, with a “turbulent diffusivity”, κt vl/2.
Problem 6.4. Convince yourself that ∇R > ∇ > ∇ > ∇a in a convection
zone. Show that the efficiency of convection
∇ − ∇ cP ρ2 lκv
Γ ≡ = (6.25)
∇ − ∇a 24σT 3
is very large in the deeper part of the convection zone and that, therefore,
∇ ∇a is a very close approximation.
The advent of large and fast computers made it possible to simulate numer-
ically turbulent flows and their capability to transport heat. It is true, these
three-dimensional and time-dependent hydrodynamical calculations are still
too expensive (in terms of computer time and storage) to be incorporated
244 6. Convection
into a full stellar evolution code. They can nevertheless be used to test some
of the assumptions made in the heuristic approach of the preceding section.
The calculations of Chan and Sofia (1987, 1989) and Kim et al. (1996)
are of particular interest. These authors solve the Navier-Stokes equation
together with the mass and energy equations for a compressible flow in an
atmosphere having a prescribed energy flux density at its bottom. The cal-
culated flow is turbulent, and three results are relevant to the mixing-length
theory:
First, the correlation
v1 v2
A = 2 1/2 2 1/2 (6.26)
v1 v2
between the vertical velocity components at two levels, 1 and 2, in the at-
mosphere is calculated. The brackets denote an average over time and (or)
the horizontal coordinates. The length over which A drops by a substan-
tial amount (e.g., to 0.5) can be identified with the vertical distance after
which a convective parcel looses its identity, i.e., with the mixing length. The
calculations show that, at various levels in the atmosphere, and also for var-
ious density stratifications, the correlation length of A is the same multiple
of the pressure scale height. This confirms the assumption (6.7). A similar
proportionality to the density scale height has not been found.
The second result of relevance concerns the rms value of the temperature
fluctuation ΔT and the mean square value v 2 of the vertical velocity com-
ponent. Both of these quantities turn out to be proportional to T Δ∇, with
suitable constants of proportionality. Thus, we have a justification of equa-
tions (6.11) and (6.15), although it is necessary to set ∇ = ∇a and δ = 1 for
a detailed comparison.
Finally, Chan and Sofia calculated the convective energy flux from the
exact formula
FC = ρcP vr δT , (6.27)
where δT is the temperature fluctuation. They found that the average vr δT
is proportional to the rms values of vr and δT (which we identify with v and
ΔT , respectively). In combination with the previous result this confirms the
heuristic expression (6.12) of the convective flux.
The results of Chan and Sofia generally support the application of the
mixing-length formalism, in particular in the deeper part of the convection
zone. Only near the surface the mixing-length assumptions are less well jus-
tified; there the numerical simulations imply a steeper temperature gradient,
and this result is supported by an improved agreement between predicted
and observed p-mode frequencies (Kim et al. 1996, Schlattl et al. 1997).
and
r2
1 2 gδ
v (r2 ; r1 ) = ΔT (r; r1 )dr . (6.29)
2 T
r1
Expression (6.28) replaces the temperature excess (6.10) of the local theory
(if the radiative loss is neglected). Equally (6.29), which is the buoyancy work
integral along the path from r1 to r2 , replaces (6.15).
For rising parcels, we set r2 = r1 + l/2, for sinking parcels, r2 = r1 − l/2,
where l is the mixing length. This is consistent with the local formalism
6.2 Mixing-Length Theory 247
where δr = l/2 was used. We should add a factor 1/2 on the right of (6.29)
on account of friction, as above. For the convective flux at level r we write
FC = f cP ρv(r; r + l)ΔT (r; r + l) + f cP ρv(r; r − l)ΔT (r; r − l) . (6.30)
The first term is the contribution of sinking parcels, the second that of rising
parcels. The factors f and f (with f + f ≤ 1) may be interpreted as the
fractions of a given horizontal area (at level r) filled with sinking and rising
parcels, respectively. These factors also may absorb the somewhat arbitrary
scaling applied above to the velocity v.
We must substitute (6.30) into equation (6.8) that describes the total
energy transport. For the mixing length l, which defines the range of inte-
gration in (6.28) and (6.29), we again use (6.7). As before, the dimensionless
parameter α can be adapted – together with the initial helium content, Y0 –
to yield the correct radius and luminosity of the present Sun.
Overshooting at the Base of the Convection Zone. Let us apply the
non-local treatment to the lower part of the solar convection zone. Here, in
the optically thick regime, we can safely use the diffusion-like radiative part
(6.9) of the energy flux. Deep in the convection zone we may also set ∇ = ∇a ,
cf. Problem 6.4.
We start the integration of the solar model at the surface with the usual
local mixing-length formalism, and switch over to the non-local treatment at
some level rc . As we are interested in downward overshooting, let us consider
only sinking parcels. We then adjust f in (6.30) in such a way that FC
continuously matches to its local form that is used above rc (Pidatella and
Stix 1986, Skaley and Stix 1991).
With this prescription, but otherwise with the same program as described
in Chap. 2, solar models have been calculated. Within the computational
error the same values of Y0 and α are found by calibrating the present lumi-
nosity and radius; the factor f in (6.30) is ≈ 0.2. The main characteristics of
these models is a temperature gradient that continues to follow the adiabatic
value, ∇a (γ −1)/γ ≈ 0.4, for some distance below the intersection with the
radiative gradient, ∇R , as illustrated in Fig. 6.2. The reason is that at that
intersection the sinking parcels still have kinetic energy and so continue their
descend into the stably stratified layer. Since radiative exchange is weak,
the layer of convective overshooting becomes nearly isentropic. In fact the
6.3 Granulation
6.3.1 The Observed Pattern
Fig. 6.3. Spectroscopy of solar granulation. Upper panel : a stripe of 60 × 5.3 Mm2
on the solar surface, with the spectrograph slit along its middle line, and a vertical
black mark that serves to position the spectrum. Middle: the spectral band 491.186–
491.219 nm (λ increasing upwards), with the Ni I line at 491.203 nm. Lower panel :
intensity Ic of the continuum near the line, with variations up to ±20 % around
the mean; vertical velocity v, with variations up to ±1.5 km/s; and full line width
at half-maximum, corresponding to Δv = 1.9–3.4 km/s (combined thermal veloc-
ity and microturbulence ξt ). The spectrum was taken on 30 July 1999 with the
echelle spectrograph of the Vacuum Tower Telescope, Izaña, Tenerife (Nesis et al.
2001), exposure 0.6 s; a correlation tracker was used for stabilization. Courtesy H.
Schleicher
The bright granules are upwards-moving, hot parcels of gas, while the
dark intergranular lanes represent cooler, downwards-moving, material. This
is clearly demonstrated in high-resolution spectra such as Fig. 6.3: these
spectra show “line wiggles”, i.e., the lines are blue-shifted in the brighter
sections of the continuum, and red-shifted in the darker sections. Moreover,
it appears that the width of the spectral line (lowest curve of Fig. 6.3) has
maxima at those positions where the gradients of I and v are largest. The
increased line width is ascribed to an increased non-thermal contribution ξt
(Eq. 4.17), and has been interpreted as shear-generated turbulence (Nesis et
al. 1993, 1996).
Intensity Contrast. The intensity contrast is important because it is a
direct indication of the temperature fluctuations. For small fluctuations, and
a Planck spectrum, we have δI/I = 4δT /T. We may use the maximum and
250 6. Convection
minimum observed intensities, Imax and Imin , to define the maximum relative
contrast
Imax − Imin
C=2 . (6.31)
Imax + Imin
More detailed information is provided by the power spectrum of the in-
tensity fluctuation, PI (kh ), where kh = (kx2 + ky2 )1/2 is the horizontal wave
number. The integral of PI over all wave numbers kh is related to the relative
rms intensity fluctuation, defined as
(δI/I)rms = (I − I)2 1/2 /I . (6.32)
For the best ground-based observations the rms intensity fluctuation may
exceed 0.1, e.g., an 11-h sequence of granulation images obtained in a 10-nm
spectral band around 468 nm on 5 June 1993 at the Swedish Vacuum Solar
Telescope, La Palma (Simon et al. 1994). Figure 6.4 shows the intensity
distribution of a typical image. Generally PI , in particular at large kh , is
attenuated by atmospheric seeing, but also by the telescope. The telescopic
attenuation can be overcome by a sufficiently large aperture, while seeing
can be avoided by observing from above the atmosphere. For ground-based
measurements a correction is necessary; the modulation transfer function, the
MTF, should be known for this purpose.
During a partial solar eclipse the intensity profile of the Moon’s limb
provides an opportunity to empirically determine the line spread function
LSF(x) and, as its transform, the modulation transfer function MTF. An
advantage of this method is that the same photographs contain both the
lunar limb profile and the granulation to be investigated. One can therefore
be sure to obtain the right MTF, for each observation. Also, in this way both
the telescopic and the atmospheric effects are taken into account at once.
An intensity power spectrum, observed and corrected (i.e., divided by
the MTF) by Deubner and Mattig (1975), is shown in Fig. 6.5. Except at
very small kh , around 1 Mm−1 , the correction is substantial. Moreover, again
except for small kh values, the correction strongly depends on the wings of the
6.3 Granulation 251
spread function. In order to show this, the spread function was approximated
by a pair of Gauss functions,
1 1
LSF(x) exp(−(x/a)2 ) + exp(−(x/b)2 ) . (6.33)
a b
Two cases were considered. In both LSF(x) had a narrow core represented
by a = 180 km and a = 200 km, respectively. On the other hand, rather
different wings were chosen, represented by b = 540 km and b = 1500 km.
The second case (crosses in Fig. 6.5) closely approximates the wings of the
observed spread function, and indeed yields about the same power (up to
kh ≈ 8 Mm−1 ) as the exact numerical calculation. The first case (circles)
suppresses the extreme wings of LSF(x), with the result that almost half of
the total power is lost.
The weight of the wings of the line spread function, and the great diffi-
culty in measuring these wings correctly, has lead to a wide range of results
concerning the solar intensity fluctuations. Also, there is a wavelength depen-
dence of these fluctuations (Problem 6.5). For (δI/I)rms , as defined in (6.32),
Deubner and Mattig (1975) find the value 0.128 at λ = 607 nm, Durrant
et al. (1983) derive 0.113 at λ = 556 nm, and Nordlund (1984 a), who used
Deubner and Mattig’s observed data, but a spread function with even more
pronounced wings than in the second case above, obtained 0.20!
The intensity contrast C, defined in (6.31) above, of course exceeds
(δI/I)rms . The excess is roughly a factor 2, according to a large number
of observations.
Problem 6.5. Approximate the emergent intensity by Wien’s law, and show
that
C (c2 /λT 2 )ΔT , (6.34)
where ΔT is the temperature difference between granules and intergran-
ular space, and c2 ≡ hc/k = 1.438 cm K. Estimate ΔT from the numbers
given in the text.
252 6. Convection
Fig. 6.6. Distribution of distances between centers of adjacent granules [a, after
Bray and Loughhead (1977)], and of granular diameters [b, after Roudier and Mul-
ler (1986)]
that later connects to the intergranular lanes. Finally the granules split into
two or more fragments, which fade away, or grow to become new granules.
When a large granule develops a dark spot in the center, a ring-shaped
convection cell is formed: an exploding granule. The ring grows, and dis-
integrates into a number of segments (Fig. 6.7). The phenomenon is quite
common: On a movie taken during a space experiment Title et al. (1986)
recognized 44 examples within 1600 s in an area of 40 × 40 . Because events
near the edge of the frame, or at the beginning or end of the movie, may have
escaped detection, the true density of exploding granules may be still larger.
The lifetime of solar granules can be measured by a correlation analy-
sis. The time required for the auto-correlation of the photospheric intensity
variation to decay to 1/e of an initial value is ≈ 6 minutes according to a
number of studies, including the balloon experiment Stratoscope (Bahng and
Schwarzschild 1961) and the space results of Title et al. (1986). Individual
granules may however live much longer than this. In a time sequence from
the balloon experiment Spektro-Stratoskop Mehltretter (1978) could identify
single granules for over 8 minutes both backwards and forwards from any
given instant of time; Dialetis et al. (1986) found a mean lifetime of 12 min
for 200 individually tracked granules. Individual tracking, in particular per-
formed manually, can account for the proper motion of granules and hence
yields a longer lifetime. On the other hand, the result also depends on the
criteria set for the birth and death of a granule. Hirzberger et al. (1999 a),
using an automatic technique and more restrictive criteria, have tracked 2643
granules through an 80-min sequence of white-light images obtained at the
Swedish Vacuum Solar Telescope, La Palma. Figure 6.8 shows their result:
an exponential distribution of lifetimes; the mean is 6 minutes.
The Velocity Field. To measure the granular velocity (via the Doppler
effect) a spectrum (Fig. 6.3) is necessary. Such is more difficult to obtain
than granulation photographs, because a typical exposure time is 1 s, and
good seeing conditions normally do not last that long. It is also necessary
to recall that spectral lines originate higher in the solar atmosphere than
the continuum. Therefore, the intensity and the velocity generally refer to
different atmospheric levels.
Let us first concentrate on the rms value of the vertical velocity. This
component is determined from spectra obtained at the center of the solar
disc. The results of Durrant et al. (1979), listed in Table 6.2, illustrate three
major problems.
First, a correction must be made for the telescopic and atmospheric image
degradations. Durrant et al., who used observations made during the partial
solar eclipse on 29 April 1976, achieve this correction by means of a modula-
tion transfer function obtained from the profile of the Moon’s limb, see above.
The correction factor typically is between 3 and 4! It should be noted that
here the velocity is corrected by means of an MTF obtained from intensity
fluctuations. This is a plausible procedure since the spatial smearing (along
the slit) “mixes” adjacent red- and blue-shifted parts of the line and so re-
duces the rms velocity signal; Mehltretter (1973) shows that the procedure
is legitimate if the Doppler shift is small in comparison to the line width.
Table 6.2. Vertical rms velocity, in km/s, in the solar atmosphere. For the height
of line formation see Fig. 6.9. From Durrant et al. (1979)
especially since the contribution functions are calculated only for the line
centers. Nevertheless, several detailed studies suggest that both the vertical
and the horizontal rms velocities decrease with height, so that the granular
velocity field extends only over a few hundred km. Figure 6.10 shows some ob-
servations, and results of mixing-length models with convective overshooting
of comparable magnitude.
The horizontal granular velocity is observed near the solar limb. Alter-
natively, we may measure the line-of-sight velocity as a function of angular
distance, θ, from the disc center. Assuming that the vertical and horizontal
velocity components are uncorrelated we have
2
vrms = μ2 vr,rms
2
+ (1 − μ2 )vh,rms
2
, (6.36)
Fig. 6.11. Velocity and temperature of the average granule. The arrow length is
proportional to |v|, with maximum 1.6 km/s. From Ruiz Cobo et al. (1996)
6.3 Granulation 257
exists only in the deepest observable layers. Higher up, the (still coherent)
vertical velocity pattern is interpreted as convective overshooting.
Ruiz Cobo et al. (1996) have used high-resolution spectra obtained at
disc center to derive the line-of-sight (vertical) velocity component for a large
number of granules. The horizontal component has been determined under
the assumptions of mass conservation and azimuthal symmetry. Thus they
derived the velocity distribution for an average granule shown in Fig. 6.11. In
addition, the temperature field has been obtained from the spectral intensity;
it shows a conspicuous reversal of the isotherms at a height of ≈ 140 km.
6.3.2 Models
Problem 6.6. Consider a mass flow ρv that horizontally varies like sin kh x,
and has a vertical scale height Hm . Show that the ratio of horizontal to
vertical velocity components is
vh /vr 1/(kh Hm ) . (6.40)
For given Hm , therefore, the ratio vh /vr increases with increasing cell size.
ρ = ρ(P, T ) , (6.41)
which contains the temperature; the latter is determined by the energy equa-
tion
∂H ∂P
ρ + ρv · ∇H = Q + + v · ∇P . (6.42)
∂t ∂t
6.3 Granulation 259
Here, H = U + P/ρ is the enthalpy per mass, and Q is the rate of energy
gain (or loss) per volume. The specific internal energy U is assumed to be a
known function of P and T .
If Q = 0, then (6.42) describes adiabatic changes. Let us restrict the
attention to radiative exchange of heat. Then
Q = −∇ · F R = ρ κν (Iν − Sν )dΩdν . (6.43)
Fig. 6.12. Numerical simulation of solar granulation. Left: Velocity (arrows) and
temperature (color, blue ≡ cool, red ≡ hot) of an individual granule. Right: Emer-
gent intensity of a 6 × 6 Mm2 field of simulated granulation, with superimposed
contours marking supersonic flow. From Stein and Nordlund (1998 a)
temperature field at one instant of time. The simulation resembles the ob-
served pattern: bright material rising in isolated, and dark material sinking in
connected, but otherwise irregular, areas. At the surface the horizontal often
exceeds the vertical flow velocity, and occasionally is supersonic and forms
shocks near the edges of granules (Fig. 6.12, right). There is some evidence
for supersonic flows in spectroscopic data (Nesis et al. 1992); perhaps the
bright edges of some granules shown in Figs. 3.5 and 3.11 are the intensity
signature of shocks (de Boer et al. 1992).
The horizontal temperature variation of the simulation generally is much
larger than observed. Nordlund suggests that most of this variation is masked
by the temperature sensitivity of the opacity, which raises the visible surface
over the granules and lowers it in the intergranular lanes. In the higher levels
the temperature variation diminishes and even reverses its sign. This effect,
which equally occurs in mean models based on mixing-length theory (Kiefer
et al. 2000 a), is attributed to adiabatic cooling of granules that overshoot
into the stable upper photosphere. It corresponds to the above-mentioned
loss of coherence between intensity and upwards velocity.
The striking resemblance between the numerical simulation and images of
the solar surface demonstrates that granulation is the visible manifestation
of convection. What is the driver of the phenomenon? A numerical study of
Rast et al. (1993) shows that about two-thirds of the enthalpy transported
by granular convection is carried as latent heat of partially ionized hydrogen.
The ionization energy is contained in the function U (P, T ), and the ensuing
buoyancy force is contained in the term ρg. Thus, ionization plays an impor-
tant role in the driving, in a similar way as in the heuristic theory of Sect.
6.2, except that single parcels, the granules, are now spatially resolved.
6.3 Granulation 261
For strong lines the dependence on χ is less (except for the increasing scatter)
because these lines are formed higher in the atmosphere, where, as we have
seen, the correlation between intensity and upward velocity essentially is lost.
The Limb Effect. When the mean line position is measured as a function
of position on the Sun it is found that the convective blueshift decreases from
the center towards the limb of the solar disc. Of course the Doppler shift
caused by the Sun’s rotation must be eliminated in order to isolate the effect,
for example by measuring along the central meridian. The decrease of the
blueshift is called the limb effect. It is to be expected because the granular
contrast decreases towards the limb and because the vertical velocity makes
an increasing angle with the line of sight so that the Doppler shift decreases.
The projection effect may even generate a redshift near the limb: the receding
horizontal motion at the far side of granules is seen against bright granules
behind.
Mean shifts of the Ti I line at 571.3 nm in sunspots have been measured by
Beckers (1977). With respect to the laboratory wavelength there is only the
predicted gravitational redshift, independent of the position on the solar disc.
Since the convective energy transport is inhibited in sunspots (Sect. 8.3), this
indirectly confirms that the blueshift with its center-to-limb variation is of
convective origin.
Line Asymmetry. In addition to their Doppler shift, mean line profiles
show an asymmetry. This asymmetry also has its origin in the granular con-
vection, as demonstrated in Fig. 6.14. Schematically, the line is composed of
granular and intergranular contributions. Both of these are already asymmet-
ric by themselves because the center of the profile originates at greater height
than the wings, and therefore has a different Doppler shift due to the depen-
dence of the velocity field on height. Averaging over the blueshifted granular
6.3 Granulation 263
Fig. 6.14. Origin of the line asymmetry. Granular and intergranular regions (left)
contribute different profiles (center). The average is an asymmetric line, with a
C-shaped bisector (right, solid). The dashed, symmetric, line would result in the
absence of convection. From Dravins et al. (1981)
and the redshifted intergranular contributions leads to the small net blueshift
already described, but this blueshift is more pronounced in the line center
than in the wings (the extreme wings may actually show a redshift), and thus
a further asymmetry is introduced.
The line asymmetry is commonly represented by the bisector, which is the
curve dividing any iso-intensity line across the profile into two halves. Often
the bisector has the form of a “C”. This C shape depends in a systematic
manner on line strength and on the lower excitation potential (Fig. 6.15),
as well as on other line parameters. Moreover, it varies with position on the
Fig. 6.15. Average bisectors, from observations at disc center, for solar lines of
various strengths (left) and various lower excitation potentials (right). Adapted
from Dravins (1982)
264 6. Convection
Fig. 6.16. Center-to-limb variation of the bisector for the Fe I line at λ = 557.61 nm.
The bars show the rms variation at cos θ = 1.0 (left) and 0.2 (right). After Brandt
and Schröter (1982)
solar disc: Figure 6.16 demonstrates this effect, but also shows the general
limb effect as discussed above. Further, the oscillatory velocity field (Severino
et al. 1986; Gomez et al. 1987) and the presence of a magnetic field in the
solar atmosphere (Livingston, 1982) have an effect on the mean line profile,
especially its asymmetry.
The main virtue of the analysis of mean line profiles is the possibility to
investigate stellar granulation, which cannot be spatially resolved by direct
observation. For such investigations the Sun plays a key role because only
here are both spatially resolved and mean profiles available.
We have explained the convective blueshift, the limb effect, and the line
asymmetry in a qualitative way by rising granules and sinking intergranu-
lar matter. Early quantitative calculations were based on compositions of
normal one-dimensioned model atmospheres, each with is own temperature
stratification, as described in Chap. 4, and with its own Doppler shift. The
“3-stream” model of Voigt (1956) had redshifted and blueshifted fractions,
and a fraction at rest; the “2-stream” model of Schröter (1957) had redshifted
and blueshifted fractions only. More detailed calculations of mean line profiles
are possible with the hydrodynamic model described in Sect. 6.3.2. In such a
(3-dimensional) model the equation of radiative transfer, (6.44), is integrated
along various selected rays to obtain theoretical line profiles as a function of
position. The average, in space and (or) in time, is then the synthetic profile
to be compared with the mean observed profile. The main characteristics de-
scribed qualitatively in this section find their theoretical explanation in this
way. Even the parameters of “micro” and “macro” turbulence (Chap. 4),
which were introduced to explain non-thermal line broadening and line
strengthening, appear to be unnecessary when the full 3-dimensional velocity
field is taken into account.
6.4 Mesogranulation 265
6.4 Mesogranulation
At first sight the solar granulation displays the irregular, sometimes polygo-
nal, pattern shown in Fig. 3.5, without any conspicuous organization at larger
scales. A closer inspection however reveals that such organization might in-
deed exist.
An example is the distribution of “active” granules, i.e., granules which
for some time continue the reproductive sequence: expansion, fragmenta-
tion, expansion of fragments, fragmentation of fragments . . . . Using pho-
tographs taken at the Pic-du-Midi observatory, Oda (1984) found that these
active granules form a network, with typical mesh size 10 , as shown in
Fig. 6.17. The network is correlated with the photospheric brightness distri-
bution. Hirzberger et al. (1999 b) found a related distribution for the explod-
ing granules; their study was based on white-light images from the Swedish
Vacuum Solar Telescope, La Palma.
The pattern shown in Fig. 6.17 is reminiscent of convection cells. A ve-
locity field with the same cell size in fact had been discovered prior to Oda’s
work by November et al. (1981), and christened mesogranulation. Before the
mesogranulation can be seen in the Doppler signal, some data reduction is
necessary, as demonstrated in Fig. 6.18: the larger-scale (typically 40 ) su-
pergranulation must be subtracted, and the oscillatory velocity field must be
removed by taking an average over time. For the Mg I line at λ = 517.3 nm,
which was used to measure the velocity shown in Fig. 6.18, the supergranu-
lar rms velocity is ≈ 40 m/s, comparable to the mesogranular flow, and the
oscillatory signal is ≈ 500 m/s, i.e., larger by an order of magnitude!
As far as small scales, i.e., the granules, are concerned, the velocity signal
is smoothed both by the seeing during the observation and by the time aver-
aging. What is left of the granular contribution seems to have little effect on
the mesogranulation: the signal is equally clear in the images with 1 × 1
and 3 ×3 resolution [(e) and (f) of Fig. 6.18; nevertheless, Wang (1989) and
Fig. 6.18. Chromospheric network (a), and velocity maps (b–f ; upflow dark, down-
flow light) showing super- and mesogranulation. The maps (b) to (d) contain the
total velocity signal, at various resolutions as indicated. In (e) and (f) the super-
granular velocity (d) is subtracted so that the mesogranulation is seen more clearly.
From November et al. (1981)
6.5 Supergranulation
6.5.1 The Velocity Field and the Network
When Hart (1956) spectroscopically measured the solar rotation she found a
velocity field that fluctuates across the visible disc. Moreover, the autocorre-
lation function of these fluctuations had a secondary maximum at a mean lag
d ≈ 26000 km between the two correlated points, in addition to the primary
maximum at d = 0.
Subsequently Leighton et al. (1962), using their technique of photographic
Doppler spectroheliograms (Sect. 3.4.7), found a cellular pattern, distributed
uniformly over the Sun, with a typical cell diameter of 1.6×104 km and a mean
spacing of ≈ 3 × 104 km between cell centers. They named this pattern the
supergranulation. A study of Schrijver et al. (1997) shows that the distribution
of supergranulation cells is in fact very similar to the distribution of granules,
in spite of the large difference in length scale. As Fig. 6.19 demonstrates, the
flow is almost invisible at the disc center by means of the Doppler effect; i.e.,
it is predominantly horizontal. Nevertheless it can be investigated at disc
center by the method of local correlation tracking, where the granulation
(and even the mesogranulation) that is carried along is used as a tracer. The
measured horizontal velocity is most frequently in the range 300–500 m/s,
directed radially from the center to the periphery of the cells; the distribution
derived by Shine et al. (2000) extends to ≈ 1 km/s.
The vertical velocity of supergranulation cells is smaller. A photoelec-
tric measurement, made by Küveler (1983) at the Locarno solar observa-
268 6. Convection
Fig. 6.19. Dopplergram of the entire Sun, Fig. 6.20. Velocity contours (la-
showing the supergranulation (dark : ap- bels in m/s) of a supergranulation
proaching, bright: receding flow). From cell, observed at disc center. Adapted
SOHO/MDI, courtesy G. W. Simon from Küveler (1983)
observation during the arctic summer (in Thule) gave 25 h for the network
observed in Hα (Rogers 1970). The material from Thule has also been an-
alyzed in a morphological study by Janssens (1970). He found that, on the
average, the network cells loose their identity after 21 hours.
The larger the scale of a velocity field, the more sensitive is the flow to the
action of the Coriolis force that arises in a rotating system. The rotational
effect is estimated by the inverse of the Rossby number
Ro = u/(2Ωl) , (6.45)
which is essentially the ratio of the inertia and Coriolis forces; u and l are a
typical velocity and a typical scale, and Ω is the average rate of rotation.
For supergranules u ≈ 500 m/s and l ≈ 107 m; with Ω ≈ 3 × 10−6 s−1
we obtain Ro ≈ 10. The effect of rotation, which is manifest in an azimuthal
(vortex-type) velocity component within each cell, therefore is of order 1/10:
considering the dependence of the Coriolis force on latitude ψ we may ex-
pect an azimuthal motion of order sin ψ × 50 m/s. This is small, but some
spectroscopic evidence has been found by Kubičela (1973). For 32 supergran-
ulation cells he found that the zero-velocity line across a cell generally is not
6.6 Giant Cells 271
perpendicular to the radius pointing from the disc center towards the cell.
Instead, there is an average angle of ≈ 80o . The sign of the deviation is such
as expected from the Coriolis effect. An azimuthal velocity component was
also found by Zhang et al. (1998) who traced the proper motion of magnetic
flux concentrations within supergranulation cells; they found that the mag-
nitude of this component increases from zero at the cell center to ≈ 400 m/s
near the cell boundaries, and that the direction is consistent with the action
of the Coriolis force.
Another effect of solar rotation is the apparent decrease with latitude of
the mean cell size of supergranules. High-latitude cells are smaller by ≈ 10%
than cells in the equatorial zone (Rimmele and Schröter 1989), an effect that
also has been found for the cells of the chromospheric emission network by
Brune and Wöhl (1982) and confirmed by Münzer at al. (1989). Presently
there is no theoretical explanation for this dependence on latitude of the
supergranular cell size.
Patterns on the solar surface with typical length scales of 108 m and more are
commonly called “giant cells”. Other names are “giant granulation”, indicat-
ing a physical analogy to the convective cells of smaller scale, and “global
convection”, which expresses the possible extent of the giant cells over the
entire depth of the convection zone.
Giant cells are most elusive: although numerous observational studies have
been made, no typical and persistent pattern has yet been established. Yet
such giant cells are most interesting: with l ≈ 108 m and u ≈ 100 m/s (see
below) their Rossby number (6.45) is much smaller than 1, and the effect of
solar rotation upon giant cells should therefore be significant. Indeed, as we
shall see in Chap. 7, giant cells possibly play a key role in the theory of the
Sun’s differential rotation.
Giant convective cells have been inferred from the peculiar motions of smaller-
scale features across the solar surface. Such tracers include the supergranula-
tion and the related chromospheric network, inhomogeneities in the magnetic
field distribution, and the visible consequences of such inhomogeneities, i.e.,
sunspots, filaments, etc. In fact, it was the appearance of complexes of (mag-
netically) active regions which, during the 1960s, first led to the concept of
a convective pattern consisting of giant cells (Bumba 1967).
A particularly nice, although very rare, example of a regular distribution
of filaments is shown in Fig. 6.22. As filaments are supported by the mag-
netic field, and the magnetic field is carried along and concentrated by the
272 6. Convection
Fig. 6.22. Distribution of solar filaments on 11 June 1972 (Wagner and Gilliam
1976). Courtesy National Solar Observatory, AURA, Inc.
Fig. 6.23. Giant cell velocity field, derived during 1976 from the proper motion of
Ca II mottles. From Schwan and Wöhl (1978)
6.6 Giant Cells 273
For the period April to June 1975 Schröter and Wöhl (1976) in this way
found a regular pattern of 4 cells distributed along the solar equator (i.e.,
longitudinal wave number m = 2). Only a narrow latitude zone (±12o ) was
observed; therefore a divergence or a vorticity of the cell flow could not clearly
be distinguished. But velocity differences of up to 80 m/s were measured, and
flows crossing the equator have been identified. In another study, made during
1976 (Schröter et al. 1978), the latitude range was extended to ±30o . Again
flows crossing the equator were found; in addition, areas of clearly divergent
motion could be identified, cf. Fig. 6.23, which shows a spherical harmonic
representation of the observed velocity vectors.
Fig. 6.24. East-west velocity on the Sun, derived from the Doppler shift of the
Fe I line at 525.02 nm, during September 1979. Contour levels are ±10, ±20 . . . m/s
(shaded eastward). From Scherrer et al. (1986)
274 6. Convection
govern the Reynolds stresses, the convective flux and related correlations (see
also Grossman 1996, Canuto and Dubovikov 1998; the latter paper also com-
ments on the asymmetry between up- and downdrafts). Chan and Sofia (1996)
consider the problem of closure with respect to the higher-order moments. Ly-
don et al. (1992, 1993) have replaced the mixing-length equations altogether
by relationships between the dynamic and thermodynamic quantities which
they derived from numerical simulation; without adjustable parameters, they
were able to predict the radius of the present Sun within an error of ≈ 1 %. All
these alternatives also allow for convective overshooting (it should be noted
here that the proper fluid-dynamics term for this phenomenon is penetrative
convection if, as in the case treated in Sect. 6.2.3, the mean stratification is
substantially affected). Zahn (1991) gives a general discussion.
Spruit et al. (1990) and Spruit (1997) review stellar convection in com-
parison to laboratory experiments and numerical simulation. The main con-
clusion is that the observed pattern is essentially determined by the cooling
at the solar surface. The downdrafts seen as intergranular lanes form isolated
channels at greater depth, merging to successively larger scales. Such plumes
may dive through the entire convection zone (Rieutord and Zahn 1995). At
the surface their locations may appear as especially long-lived intergranular
“holes” (Roudier et al. 1997). Numerical studies of convective overshooting
show penetrative flows over a sizable fraction of a pressure scale height, both
above and below the unstable region (Singh et al. 1994, 1995, 1998). Asplund
et al. (2000 a) consider the effects of grid resolution in numerical simulations,
while the anelastic approximation is discussed by Lantz and Fan (1999).
Balthasar (1984, 1988) has compiled parameter lists for 143 asymmetric
spectral lines obtained with a Fourier transform spectrometer. Canfield and
Beckers (1976) and Dravins (1982) describe the effects of unresolved motions
on spectral lines. Kaisig and Durrant (1982) employ a perturbation analysis
to model the line asymmetry, and Dravins et al. (1986) extend the work of
Dravins et al. (1981) to Fe II lines. Márquez et al. (1996) found that internal
gravity waves and a spatially variable microturbulence parameter yield a
good match to observed asymmetric line profiles. Asplund et al. (2000 bc)
find that the observed width and asymmetry of Fe lines can be matched in
three-dimensional radiative-hydrodynamical simulations without recourse to
micro- and macroturbulence concepts, and Gadun et al. (1999) reproduce a
number of observed line-profile features even with two-dimensional models.
Stellar line asymmetries and wavelength shifts, and implications on stellar
granulation, were investigated by Dravins (1987 a), and first stellar bisectors,
among others for α Cen A and Procyon, have been determined by Dravins
(1987 b).
Reviews on resolved velocity patterns are those of Beckers (1981) and
Beckers and Canfield (1976). Müller et al. (2001) study over 42 000 individ-
ual granules with an automatic tracking technique and find a mean lifetime
of 7.5 min, slightly larger than the result of Hirzberger et al. (1999 a). The
276 6. Convection
strong dependence of the granular intensity contrast on the form of the spread
function has been confirmed by Collados and Vázquez (1987). A related dis-
cussion, on the restoration of the granular velocity, has been presented by
Mattig (1980). Incidentally, there is not only atmospheric and instrumental
smearing of granular information; through repeated scattering within the so-
lar atmosphere horizontal fluctuations may be reduced in magnitude, i.e., the
Sun itself sets limits to the resolution (Kneer 1979). Nevertheless, sharp edges
caused, e.g., by the magnetic structure of the atmosphere may be recognized
down to a scale of order 10 km (especially in polarized light) if a sufficiently
large telescope is used (Bruls and von der Lühe 2001).
Meso- and supergranulation velocities, in particular their variation with
height in the solar atmosphere, have been spectroscopically studied by
November et al. (1979, 1982). The rotational effect on the supergranulation
was measured by time-distance helioseismology (Duvall and Gizon 2000). Hill
(1989) uses his method of ring diagrams to deduce flows of order 40 m/s on
scales larger than the supergranulation. Volumes 192 and 193 (2000) of Solar
Physics cover the helioseismic diagnostics of solar convection.
Bumba (1987) once more emphasizes the possible relationship between
large-scale inhomogeneities of the solar magnetic field and giant-cell convec-
tion. A special spectroscopic search for giant cells at high solar latitude was
made by Cram et al. (1983) and Durney et al. (1985). The strong rotational
influence should render the giant convective cells elongated in the direction
parallel to the axis of rotation (cf. Sect. 7.5.3) and thus, it was argued, their
visibility should be best in the vicinity of the poles. But the evidence is
marginal, with an upper bound of ≈ 5 m/s. In contrast to the theoretical
prediction Ribes et al. (1985) report the discovery of an east-west oriented
convective structure. Proper motions of sunspots, measured in spectroheli-
ograms from the Paris observatory, are the basis of this finding.
The amplitude, or the upper bounds for the amplitude, of giant cells is
about an order of magnitude smaller than predicted by the global circulation
models described in Sect. 7.5.3. A possible answer to this discrepancy is a
confinement of the giant cells to the lower part of the convection zone, and
a screening by the above-lying, smaller-scale convection (Stix 1981 a; van
Ballegooijen 1986).
7. Rotation
by Carrington, suggest that i is slightly smaller: Balthasar et al. (1986 a), us-
ing sunspot group observations made between 1874 and 1976 at Greenwich,
and between 1947 and 1984 at Kanzelhöhe, Austria, found i = 7.137o ±0.017o .
For recurrent spots alone (those which are seen a second time after passing
the Sun’s far side) the Greenwich data yield i = 7.12o ± 0.05o (Balthasar et
al. 1987), which is also smaller than Carrington’s result. Carrington’s , on
the other hand, lies within the range of modern determinations.
Heliographic Coordinates. Once the axis of solar rotation is known the
solar latitude is easily defined as the angular distance ψ from the equator.
The distance to the North Pole is θ = π/2 − ψ; this quantity, also called the
colatitude, is often used in theoretical work.
Longitude is more difficult since there are no fixed points on the Sun.
Following R. C. Carrington, we divide the time into intervals of 27.2753 days,
called Carrington rotations because this period was determined by Carring-
ton as the mean synodic rotation period of sunspots. These intervals have
consecutive numbers: rotation number 1 commenced on 9 November 1853,
number 1978 was completed on 30 June 2001. At the date of commencement
of a new rotation the center of the solar disc has longitude φ = 0.
The coordinates ψ and φ are measured in a system envisaged to rotate
with the Sun. In addition to these we may define coordinates β and λ in
an ecliptic system. These angles are related to ψ and φ and to the rotation
elements trough (cf. Waldmeier 1955, p. 42)
Convince yourself that the solar North Pole is best visible in September,
and the South Pole is best visible in March.
7.2 Oblateness 279
7.2 Oblateness
7.2.1 Origin
As a rotating, non-rigid body the Sun must have an oblateness. At the surface,
where the oblateness can be observed, there are two sources for such an
effect: one is the rotation of the surface layer itself, the other a possible
internal quadrupole moment. We shall now derive expressions for these two
contributions.
For simplicity, consider an inviscid star in the state of pure steady rigid
rotation, i.e., (in spherical polar coordinates)
v = (0, 0, rΩ sin θ) , (7.6)
and
ρv · ∇v = −∇P − ρ∇Φ . (7.7)
The equilibrium condition (7.7) is written in a frame at rest, and Φ is the
gravitational potential. The surface of our rotating star is characterized by a
constant value of the pressure P . Substitution of (7.6) into (7.7) then implies
that, at the surface, the expression
1
Φ − (rΩ sin θ)2 (7.8)
2
is a constant. In addition, there must be, at the surface, a continuous transi-
tion of Φ to the outer gravitational potential, which consists of the familiar
monopole and a small quadrupole:
r
2
Gm
Φ = Φext ≡ − 1 − J2 P2 (θ) . (7.9)
r r
Here J2 is the quadrupole moment, and P2 (θ) is the second Legendre poly-
nomial.
We approximate the oblate surface by
where r is now the mean solar radius and c is a small quantity related to
the oblateness:
Δr 1 Ω 2 r 3
= + J2 . (7.13)
r 2 g 2
In the derivation of (7.13) we have ignored that the Sun rotates differ-
entially. On the other hand, we have admitted a quadrupole moment, which
possibly arises from a core rotating more rapidly than the surface. This ap-
parent inconsistency can however be removed in a more general treatment
which allows for differential rotation (cf. Problem 7.2). Then the value of Ω
to be used in (7.13) is a mean angular velocity at the surface.
If we take Carrington’s synodic rotation period (Sect. 7.1), we obtain the
synodic angular velocity Ωsyn = 2.67 × 10−6 s−1 , and Ωsid = 2.87 × 10−6 s−1
as the sidereal rate. The contribution of the solar surface rotation to the
oblateness then is 1.04 × 10−5 .
Problem 7.2. For a rotating star described by (7.6) and (7.7) introduce an
effective gravitational acceleration, defined by ρg eff = ∇P. Show that the
following three statements are equivalent: (a) g eff can be derived from
a potential, Ψ ; (b) Ω is constant on cylinders, i.e., Ω = Ω(s), where
s = r sin θ; (c) the surfaces P = const. and ρ = const. coincide with each
other (and with the surfaces Ψ = const.).
7.2.2 Measurements
Observed values of Δr/r are listed in Table 7.1. The first value given is
considerably in excess of the surface contribution. It would mean that J2 ≈
2.5 × 10−5 ; with this quadrupole moment the gravitational potential of the
Sun would deviate from the spherically symmetric form by an amount that
would cause the perihelion of Mercury’s orbit to rotate by about 3.5 per
century. This in turn would require a modification of Einstein’s general theory
of relativity which already accounts for the full 43 per century not explained
by classical celestial mechanics and special relativity. Moreover, the larger
quadrupole moment, if due to a rapidly spinning core, would imply an internal
angular velocity about 20 times larger than Ω at the surface, in contradiction
to the results obtained from helioseismology. On the other hand, the more
recent oblateness measurements are quite consistent with the notion that the
Sun’s visible oblateness is essentially caused by the surface rotation alone. In
fact the internal rotation derived from helioseismology (Sect. 7.4.1) has been
used by Pijpers (1998) to estimate J2 = (2.18 ± 0.06) × 10−7 .
The measurement of Δr/r is difficult. Dicke et al. project the Sun onto
an occulting disc which is slightly smaller than the solar image. The light
passing the edge is scanned by two diametrically opposed rotating windows,
and the signal measured in this way is converted into information concerning
7.3 Rotational History 281
∗
as quoted by Lydon and Sofia (1996)
the Sun’s shape. Hill and Stebbins also use two windows at diametrically
opposed edges of a solar image. At a fixed position angle, these windows
are scanned across the solar limb; the scan signal is fed into a servo system
to adjust the distance between the two windows. The adjustment itself is
measured interferometrically. The whole apparatus can be turned so that
equatorial and polar solar diameters can be measured. The further results
listed in Table 7.1 have been obtained with the balloon-borne Solar Disk
Sextant of Maier et al. (1992), with a rotatable scanning heliometer at the
Observatoire de Pic du Midi, and with the Michelson Doppler Imager during
roll maneuvers of the SOHO satellite.
An oblateness of 10−5 corresponds to a difference of 14 km between the
equatorial and polar solar diameters, or to 0.02 . This is an order of magni-
tude less than can be resolved with the best solar telescopes during excellent
observing conditions! Clearly, a large number of measurements must be made
before a statistically meaningful result is obtained. Even more difficult is the
assessment of all possible systematic errors. The latter may arise both in the
instrumentation (e.g., inaccurate centering of the solar image) and on the
Sun itself (e.g., latitude-dependent brightness of the limb).
We have seen in Chap. 2 that the total mass of a star, its initial chemical
constitution, and its age essentially determine the present state. There, by
“initial” we meant the beginning of the hydrogen-burning phase, i.e., of the
star’s main-sequence life. For the evolution of a rotating star we should now,
additionally, know the initial angular velocity.
There are several observational and theoretical aspects indicating that the
initial Sun had a more rapid rotation than the present Sun. The first is that
the specific angular momentum of the material that eventually formed the
282 7. Rotation
Fig. 7.1. Lithium abundance, surface rotation, and Ca II emission versus stellar
age. From Skumanich (1972)
Sun was much larger than that of the whole present solar system. Although
much of this angular momentum must have been lost by magnetic braking
during an early phase, this loss probably was not complete. The T Tauri
stars, which populate a region of the Hertzsprung–Russell diagram passed by
the Sun before it arrived on the main sequence, rotate with surface velocities
around 15 km/s (the group with m/m < 1.25; see Bouvier et al. 1986),
as compared to the Sun’s 2 km/s. During their final approach to the main
sequence these stars contract and therefore even accelerate their rotation.
The second piece of evidence is that for otherwise similar main-sequence
stars the rate of rotation decreases with increasing age. This is demonstrated
in Fig. 7.1, together with two other age-dependent stellar properties: the
lithium abundance, and the emission in the core of the Ca II lines H and K.
A third argument is indirect. Early main-sequence stars, spectral types O
to F, rotate more rapidly, by factors up to 100, than later main-sequence stars,
types F to M (Fig. 7.2). The reason is that magnetic braking, first proposed
by Schatzman (1959), is only effective among the latter. Stars earlier than,
say, F5, have no deep outer convection zone; hence they cannot generate a
magnetic field that would provide the lever arm of the braking mechanism.
By implication, the initial Sun was a fast rotator, magnetically braked down
to the present state.
7.3 Rotational History 283
7.3.2 Torques
Magnetic Braking. Any matter that escapes from the surface of the ro-
tating Sun takes with it some angular momentum. The essence of magnetic
braking is that this angular momentum does not correspond to the solar sur-
face but to a distance, called the Alfvén radius rA , far above this surface. In
a simplified view we may consider the magnetic lines of force as a lever arm
that out to rA forces the escaping material to rotate rigidly with the Sun.
284 7. Rotation
very small indeed, although the oblateness measurements would still permit
a central rotation rate several times the surface value.
The question of stability is more difficult. Large |∂Ω/∂r| means a strong
shearing motion which, at the low viscosity of the solar gas, must be unsta-
ble. Another instability arises from the accumulation of 3 He and the ensuing
(perhaps intermittent) growth of internal gravity waves in the core. A third
is of baroclinic nature: a finite angle between the surfaces of constant tem-
perature and constant pressure makes potential energy available which may
drive a flow, comparable to the driving of winds in the terrestrial atmosphere.
The effect of these and many other instabilities is that the generated flows
transport angular momentum and hence tend to smooth out gradients of Ω.
Internal Magnetic Torque. But the instabilities also mix matter, which
brings about two complications. First, as model calculations show, the mix-
ing associated with sufficient transport of angular momentum would cause a
much stronger than observed lithium depletion. Second, because of the gra-
dient in the mean molecular weight, the mixing would require work, so much
work in fact that it possibly does not occur after all. For these reasons Spruit
(1987) favors the torque exerted by an internal magnetic field. It is readily
seen that a very weak field is sufficient to slow down the Sun’s core:
Let Bp and Bt be the poloidal and toroidal components of an axisym-
metric internal field (i.e., with lines of force in meridional planes and along
parallel circles, respectively). To Bt belongs a poloidal current density of or-
der Bt /(rμ) which, together with Bp , forms an azimuthal Lorentz force of
order Bp Bt /(rμ) per unit volume. Multiplication by r (the lever arm) yields
the torque density, and again by r3 (the volume) yields the total torque. To
be effective within a time t this total torque must be of order ΘΩ/t, i.e.,
ΘΩ/t Bp Bt r3 /μ , (7.16)
Bt Bp Ωt , (7.17)
and therefore,
Bp (μρ)1/2 r/t . (7.18)
As typical values for the solar interior take ρ = 103 kg/m3 and r = 3 × 108 m.
For t to be less than the solar age we then need
Bp ≥ 10−10 T . (7.19)
If the field is stronger then the braking of the core proceeds more quickly.
Incidentally, the magnetic torque also acts as a restoring force of a torsional
286 7. Rotation
Problem 7.3. Derive the magnetic torque per volume from the general ex-
pression r × (j × B), and write down the correct differential equations
approximated by (7.16) and (7.17).
Fig. 7.3. Left: angular velocity Ω(r) at an age of 3 × 107 yr, from calculations
without magnetic torque, for a total initial angular momentum of 50, 16.3, and
5 × 1042 kg m2 /s (thin curves, from above), and the final angular velocity of the
present-day Sun in all three cases (thick curve). Right: the final rotational frequency
Ω(r)/2π at the equator (solid), 45o latitude (dashed), and at the poles (dash-dotted)
for a model that includes the magnetic torque in the interior; this model also in-
cludes the Λ effect which leads to differential rotation in the convection zone. After
Pinsonneault et al. (1989) and Rüdiger and Kitchatinov (1996)
Problem 7.4. Assume that the magnetic torque in Eq. (7.20) is negligible
and show that a substantial change in Ω occurs in a time r2 /ν, the time
scale of diffusion.
Problem 7.5. Rayleigh’s dynamical instability. In a rotating star, exchange
two rings of matter with different radii. Show that this exchange liberates
energy if the angular momentum per unit mass decreases with distance s
from the axis of rotation, i.e., if
∂ 2 2
(s Ω) < 0 . (7.21)
∂s
[In the solar core this instability is inhibited by the entropy gradient, cf.
Chap. 6; in addition, the radial change of Ω appears to be too small to
satisfy (7.21). For a stability analysis of the internal rotation inferred from
helioseismic data see Charbonneau et al. (1999)].
288 7. Rotation
In this section we shall collect results concerning the angular velocity Ω(r, θ)
of the Sun. In view of the theory to be presented in Sect. 7.5, we shall also
include some data for meridional circulation, as well as for the large-scale
velocity field already treated in Sect. 6.6.
As explained in Sect. 5.3.8, the solar rotation removes the (2l + 1)-fold degen-
eracy of the frequencies of a non-radial p mode of degree l. The frequencies
are split into (2l + 1)-fold multiplets. Since different modes of oscillation
penetrate into different depths of the Sun, the frequency splitting can be
inverted; the angular velocity as a function of depth can thus be obtained.
Moreover, multiplets with non-equidistant frequency spacing contain infor-
mation on the dependence of Ω on latitude. The reason for this dependence is
that the spherical harmonics with |m| ≈ l (the outer multiplet components)
are rather concentrated towards the equator, while harmonics with |m| l
(the inner components) cover the entire sphere, cf. Problem 5.15.
The results (Fig. 7.4) show small changes of the angular velocity between
the surface and ≈ 0.95r . Except for this near-surface region, the angu-
lar velocity within the convection zone has approximately the same latitude
dependence as at the surface. At the base of the convection zone, around
Fig. 7.4. Solar rotation rate νrot = Ω/2π, as a function of r, for three heliographic
latitudes. An inversion of data obtained with the Michelson Doppler Imager on
SOHO. The vertical line marks the base of the convection zone; arrows indicate the
rate measured spectroscopically at the surface. From Kosovichev et al. (1997)
7.4 The Angular Velocity of the Sun 289
600
500
400
300
200
0.00 0.05 0.10 0.15 0.20
rt /R
Fig. 7.5. Solar rotation rate νrot = Ω/2π, as a function of r, derived from the
experiments GOLF (filled circles) and MDI (open circles) on the SOHO satellite.
After Bertello et al. (2000)
Table 7.2. Coefficients for the Sun’s differential rotation (sidereal, in degrees/day)
A B C
Newton and Greenwich; recurrent spots, 14.368 −2.69
Nunn (1951) 1878–1944
Howard et al. Mt. Wilson; all spots, 14.522 −2.84
(1984) 1921–1982
Balthasar et al. Greenwich; all spots, 14.551 −2.87
(1986 b) 1874–1976
Snodgrass Mt. Wilson; Doppler 14.050 −1.492 −2.606
(1984) shifts, 1967–1984
Snodgrass and Mt. Wilson; Doppler signal 14.71 −2.39 −1.78
Ulrich (1990) correlations, 1967–1987
Komm et al. Kitt Peak; magnetogram 14.42 −2.00 −2.09
(1993 a) correlations, 1975–1991
Brajša et al. SOHO/EIT; bright points, 14.6 −3.0
(2001) 1998–1999
Timothy et al. Skylab; coronal holes, 14.23 −0.4
(1975) 1973
7.4 The Angular Velocity of the Sun 291
Fig. 7.8. Annual residual sunspot rotation, 1921–1982, averaged over latitudes
30o S to 30o N. Vertical lines mark the minima of the solar activity cycle (cycle
numbers in boxes). From Gilman and Howard (1984 a)
294 7. Rotation
Fig. 7.9. Variation of the solar differential rotation, derived from Doppler shifts
measured at Mt. Wilson Observatory. Velocity contours are ±1.5, ±3, ±6 m/s; areas
with velocity ≥ +1.5 m/s are shaded. After Howard and LaBonte (1983)
removed by this procedure, on the other hand spatial variations of small lati-
tudinal extent [which are not represented by (7.22)] are made visible. Among
these variations is a steady narrow equatorial retardation (the equator rotates
slower than the adjacent low latitude zones by about 0.04 degrees/day), and
a wave-like periodic variation of the differential rotation: alternating faster
and slower latitude zones, migrating towards the equator in each of the two
hemispheres. Moreover, helioseismic studies (Howe et al. 2000 b) show that
this variation is not confined to the solar surface, but extends downward at
where the brackets mean an average over longitude φ. The observations in-
dicate that Qθφ is positive in the northern hemisphere, and negative in the
southern, with a nearly linear variation in the latitude range covered by
sunspots (e.g., Pulkkinen and Tuominen 1998). It will soon be clear – cf.
(7.27) below – that this sign means transport of angular momentum from
both sides towards the equator. Averages over latitude of |Qθφ | are listed in
Table 7.3. Although there are differences between the various tracer results
we conclude that there exists a real effect of order 103 m2 /s2 .
The viscosity of the solar gas is very small, so that we may begin with
Euler’s equation in order to describe the conservation of momentum:
∂v 1
+ v · ∇v = − ∇P − ∇Φ , (7.25)
∂t ρ
which we have already considered in the special case of pure rotation, ex-
pressed by (7.6). Let us, for the moment, neglect fluctuations of the density
ρ, as in the anelastic approximation (strictly, this is not allowed, especially
since we know that the convective motion is driven by density fluctuations).
The conservation of mass is then expressed by the equation of continuity in
the steady form
∇ · (ρv) = 0 . (7.26)
We now consider the mean longitudinal motion, i.e., the average over
the φ component of (7.25). The gradients do not contribute to this average
(Problem 7.8). We multiply by the lever arm s = r sin θ, substitute (7.24),
make use of (7.26), and write sΩ in place of vφ ; the result is
∂
(ρs2 Ω) + ∇ · (ρs2 Ωv m + ρsuφ u) = 0 , (7.27)
∂t
and expresses the conservation of angular momentum. At any given place, the
density ρs2 Ω of angular momentum changes either because angular momen-
tum is carried along by meridional circulation v m , or because its transport
is mediated by the Reynolds stresses (Reynolds 1895),
Qij = ui uj . (7.28)
298 7. Rotation
These two means of transporting angular momentum play the key roles in
most theories of the Sun’s differential rotation, and we see now why it is
important to gather as much observational information as possible about
them.
The observed variation in time of the Sun’s angular velocity is so much
smaller than the spatial variation (in latitude) that it is reasonable to restrict
the attention to steady models. Then the diverse transport processes must
compensate each other. Moreover, we consider the convection zone in isola-
tion, with no flux (of matter or angular momentum) across its surfaces at rv
and r . That is, at these boundaries,
vr = 0 , Qrφ = 0 . (7.29)
Theories of the differential rotation can be classified according to their
treatment of the Reynolds stresses Qij . In principle these quantities must be
calculated from the velocity components ui , which means that the fluctuating
part of (7.25) should be solved. Such is not feasible at present. In fact one
may ask whether it is necessary to know all details of the turbulent flow
in order to explain a large-scale phenomenon such as the Sun’s differential
rotation.
In the following two sections we shall deal with two kinds of models: those
which employ the Reynolds stresses as mean quantities only, and those which
explicitly attempt to calculate ui , and therefrom Qij .
Problem 7.8. Confirm (7.27); show that the gradients occuring in (7.25)
make no contribution.
Problem 7.9. Show that u · ∇u may be written in terms of a divergence
of the Reynolds-stress tensor.
Let us now follow the work of Rüdiger (1980) and divide the Reynolds stresses
into their “diffusive” and “non-diffusive” parts. The diffusive part was already
mentioned in the context of the internal solar rotation: it models the smooth-
ing effect of turbulence upon large-scale gradients. In the present case this
means that the transport of angular momentum occurs downwards along the
gradient of Ω, i.e.,
(Qrφ , Qθφ ) = −νt s∇Ω (7.30)
(the factor s = r sin θ ensures that Qrφ and Qθφ vanish at the poles, as
they must according to their definition, and that the parameter νt has the
dimension of a diffusivity). It is easy to see, cf. Problem 7.4, that there will
be uniform rotation if there is no meridional circulation and no contribution
other than (7.30) to the Reynolds stresses. For νt = 109 m2 /s, the state of
uniform rotation would be reached within a few years.
7.5 Models of a Rotating Convection Zone 299
their existence to the rotation itself. Recalling the definition (7.28) of the Qij
we see that Λr could be a function of r alone, but that Λh should have at
least a sin2 θ-dependence (the factor cos θ takes care of the antisymmetry of
the Coriolis effect upon u, and the dimensionless parameter ŝ is included in
order to allow for a difference in the vertical and horizontal smoothing effect
of the turbulence).
Historically the development was slightly different. In view of their
smoothing effect the Reynolds stresses were written as a linear function of
vi and the derivatives of vi . Upon substitution into the averaged Euler
equation a term resembling the viscous force in the Navier-Stokes equation
was then obtained. In its simplest form the diffusivity thus defined is a scalar
quantity, and is identical to the coefficient νt introduced above. Going back
to the definition (7.28) of Qij an order-of-magnitude estimate immediately
shows that
νt lu , (7.32)
where l and u are the typical scale and velocity of the field u (in the convection
zone, νt ≈ 109 m2 /s).
It was however noticed by Biermann (1951 a) that in a stellar convection
zone the transport should generally be different in the radial and horizontal
directions. Therefore, νt was replaced by a tensor νij which, in spherical polar
coordinates, was given the form (Kippenhahn 1963)
⎛ ⎞
1 0 0
νij = νt ⎝ 0 ŝ 0 ⎠ , (7.33)
0 0 ŝ
which is known under the name “anisotropic viscosity”.
A detailed calculation shows that ŝ is the parameter already introduced
above (Kippenhahn writes s for ŝ) and that (7.33) is identical to (7.31) pro-
vided
Λr = 2νt (ŝ − 1) , Λh = 0 . (7.34)
We shall now consider this special case. The necessity of a meridional
circulation is seen in the following way. Assume first that v m = 0 and substi-
tute (7.31) into the (steady) angular momentum balance. A little calculation
shows that the angular velocity Ω depends on r. The centrifugal force arising
from such Ω generally is not conservative, and hence cannot be balanced by a
pressure gradient. Turning around the argument of Problem 7.2, we conclude
that the assumption of pure rotation, i.e., v m = 0, was false. Thus we do have
meridional circulation, and may use it to sustain the observed ∂Ω/∂θ. This
is the essence of the anisotropic viscosity models (Kippenhahn 1963; Köhler
1970).
A variety of such models has been calculated; their circulation and angular
velocity pattern often look similar to the case shown in Fig. 7.12. The models
7.5 Models of a Rotating Convection Zone 301
There is no hope (and, as already said, perhaps no need) to simulate all details
of turbulent convection in a numerical calculation. But perhaps we are able
to simulate the interesting part. This part consists of the largest cells in the
convection zone. For these the rotational influence is strong, i.e., the Rossby
number (6.45) is small; we may expect that the ensuing distortion of the cells
produces the Reynolds stresses that we need to maintain the right Ω(r, θ).
Strictly speaking, the velocity field is now divided into three parts. The
axisymmetric part is the same as in (7.24) above. But the part u is subdi-
vided: the larger-scale structure is treated explicitly, which is possible with a
reasonable number of grid points or harmonics; the smaller-scale structure is
parameterized as in the mean-field models, but in the crudest possible way,
namely by a scalar turbulent diffusivity νt .
Calculations of this kind have been performed especially by P. A. Gilman
and G. A. Glatzmaier since about 1975. They use the anelastic approximation
already mentioned in Sect. 6.3.2 in the context of granulation. Further, they
define a reference state of the convection zone. This state is characterized by
302 7. Rotation
Fig. 7.13. “Global convection”. Left: snapshots of the radial velocity at three levels
in the convection zone (dotted : solar surface, dashed : equator). Right: enlarged
segments of 50o × 66o , showing radial velocity and temperature. The arrow points
to an area of high vorticity. From Miesch et al. (2000)
purely diffusive energy transport, cf. Problem 6.3, and by a pure rotation with
a mean angular velocity Ω0 . Depending on the rates κt and νt of turbulent
diffusion, and on Ω0 , the reference state itself becomes unstable, and global
convection sets in. In the numerical models this global convection is simulated
by solving the time-dependent equations of motion.
The results of such a calculation depend on the model parameters, in
particular on the turbulent diffusivity νt . Miesch et al. (2000) treat two cases,
with νt = 1.9 × 109 m2 /s and νt = 3 × 108 m2 /s, respectively, for a spherical
shell confined between r = 0.62r and 0.96r . Their simulation covers the
main part of the convection zone, including a layer of convective overshooting
at its base, but excluding a layer immediately below the photosphere where
the scales are too small to be resolved. In the case with the larger diffusivity
7.6 Bibliographical Notes 303
Problem 7.10. Assume that the balance of forces is dominated by the equi-
librium between pressure gradient and Coriolis force (known in meteorol-
ogy as the geostrophic balance), and assume P = P (ρ). Prove the Taylor–
Proudman theorem, namely that, in the rotating frame, ∂v/∂z = 0,
where z is the coordinate parallel to the axis of rotation.
(1978), and the articles of Gilman (1974) and Stix (1989) also give general
reviews; the books of Tassoul (1978, 2000) are comprehensive introductions.
The rotation elements, i and , have also been measured spectroscopi-
cally (Wöhl 1978), but with lesser precision. The measurements of the Sun’s
oblateness and their implications are discussed by Dicke (1974), Godier and
Rozelot (2000) include the differential rotation into their theoretical study.
Ulrich (1986) gives an account of the stresses for the internal transport of
angular momentum.
Schröter (1985) reviews the surface measurements, and especially the
sources of errors, which are plentiful. Beck (2000) compares the results ob-
tained in a large number of studies based on spectroscopic, tracer, and he-
lioseismic techniques. As a new helioseismic result, variations of the angular
velocity at the base of the convection zone, with an amplitude of order 1 %
and a time scale of 1.3 years, have been reported by Howe et al. (2000 a).
A critical discussion of the wave-like variation of the angular velocity at the
surface (Fig. 7.9) is due to Snodgrass (1985). Hathaway (1987) proposes a
general representation of the surface velocity field in terms of spherical har-
monics, and discusses how to separate the various contributions, i.e., rotation,
circulation, giant cells, etc.
Many studies based on sunspots or other tracers have confirmed that the
meridional circulation diverges from the zones of sunspot occurrence, and its
concomitant dependence on the cycle phase (e.g. Tuominen and Kyröläinen
1982, Komm et al. 1993 b, Nesme-Ribes et al. 1993, Pulkkinen and Tuomi-
nen 1998); Wöhl and Brajša (2001) show that the magnitude of the velocity
increases with the distance from the spot zone. The spectroscopic determina-
tions of the circulation velocity have been reviewed by Cavallini et al. (1992)
and interpreted as a real temporal variation.
Rüdiger (1989) treats the foundations for the mean-field theory of the so-
lar differential rotation. The Λ effect is quenched by rapid rotation (Kitchati-
nov and Rüdiger 1993), which has implications for the prediction of stellar
differential rotation. In addition to the Λ effect, the anisotropy of the con-
vective heat transport and the meridional circulation have been included
in models of rotating convection zones by Kitchatinov and Rüdiger (1995,
1999) and Rüdiger et al. (1998). Mean-field models of the solar differential
rotation have also been calculated by Pidatella et al. (1986). Durney (1987)
has outlined a model in which the rotation makes the excess, ∇ − ∇a , of
the temperature gradient over the adiabatic gradient dependent on latitude.
Thus he generalizes the mixing-length theory to the case of a rotating con-
vection zone. Not only the mixing length, but also the lateral dimension of
the convecting parcels must be specified in such a model; if these quantities
are known, the Reynolds stresses can be estimated. A rigorous discussion of
the Reynolds stress tensor can also be found in Gough (1978). Miesch (2000)
reviews the diverse models that describe the convection-rotation coupling.
8. Magnetism
For at least 2000 years astronomers have been attracted by solar activity:
At first mainly the appearance and the variation of sunspots, later sunspot
structure, prominences, coronal variations, eruptions with all their terrestrial
consequences from aurorae to blackouts in radio transmission, – the literature
on these phenomena is enormous.
It is now known that all solar activity is an immediate consequence of
the existence of a magnetic field on the Sun. In the present chapter we shall
discuss this field. The main emphasis will be on the photosphere and the
convection zone. The chromosphere and the corona, where the magnetic field
often is the dominant agent, will be treated in the subsequent chapter.
Magnetism on the Sun occurs on all scales. The diameters of the smallest
“tubes” of magnetic flux are at or below the present limit of spatial resolution,
while the largest manifestations of the field, the mean-field components, may
cover an entire hemisphere. In three sections we shall go from the smallest
to the largest scales. But first let us recollect some fundamental facts about
the interaction of electrically conducting matter with a magnetic field.
We begin with Maxwell’s equations for the magnetic field B, the electric field
E, and the electric current density j:
div B = 0 , (8.1)
curl B = μj , (8.2)
curl E = −Ḃ , (8.3)
where the dot denotes the time derivative, and μ is the magnetic permeability
(which will always be taken as that of free space, i.e., μ = 4π × 10−7 Vs/Am).
In (8.2) we have neglected the displacement current, which is an excellent
approximation for non-relativistic, or “slow” phenomena (Problem 8.1).
If the magnetic and electric fields are embedded in a material with electric
conductivity σ, then the current density is σ times the electric field strength.
j̃ = σ Ẽ . (8.4)
If the motion, say v, is non-relativistic (|v| c) the contribution of the
convected volume charge to the electric current density is negligible, and the
transformation to the frame at rest is j̃ = j and Ẽ = E + v × B. Hence
j = σ(E + v × B) (8.5)
is the form of Ohm’s law which we shall use in the present chapter.
It is an easy matter to eliminate E and j from (8.2), (8.3), and (8.5). We
obtain the induction equation
where
1
η= (8.7)
μσ
is called the magnetic diffusivity [because, for v = 0, (8.6) is a diffusion
equation].
The first term on the right of (8.6) describes the inducing effect of the
material motion upon the magnetic field, while the second term manifests
the “ohmic” decay of the field due to the finite electrical resistance. By order
of magnitude we may compare the two terms if we replace the diverse vectors
by their absolute magnitudes, and the curl operator by 1/l, where l is the
scale of field variation in space. We obtain
Rm = vl/η (8.8)
as the ratio of the induction term over the decay term; Rm is called the
magnetic Reynolds number; it can also be understood as a ratio of two time
scales, namely the time scale of ohmic decay
τD = l2 /η (8.9)
and the advection time scale, l/v. In solar physics one often speaks of “high
conductivity”; the precise meaning of this is Rm 1, or that τD is very much
longer than l/v or than any other characteristic time scale of interest.
Problem 8.1. Give an estimate of the displacement current which was neg-
lected in (8.2).
8.1 Fields and Conducting Matter 307
corona. For a crude estimate of σ in the solar convection zone let us assume
complete ionization and apply (8.10). With all constants substituted, we have
σ 0.003T 3/2 A/Vm. We may use the results of Chap. 6 (Table 6.1) to see
that this is a very large conductivity indeed. Figure 8.1 shows the magnetic
Reynolds number (8.8) calculated with convection velocities and scales ob-
tained in the mixing-length theory.
Problem 8.2. Show that Λ (kT /EES )3/2 and use the results of Fig. 2.2 in
order to assess the value of ln Λ in the Sun’s interior.
Fig. 8.2. Electrical conductivity σ in the solar atmosphere; photosphere and facula
after Kopecký and Soytürk (1971); sunspot after a formula of Kopecký (1966), and
data from Avrett (1981 a)
8.1 Fields and Conducting Matter 309
Second, when collisions are rare and the magnetic field is strong, the con-
ductivity becomes anisotropic. Expressions (8.10) and (8.11) above would
then give the conductivity for a current along the field, while the “transver-
sal” conductivity would be substantially reduced (because of the guiding of
electrons around the magnetic lines of force). The anisotropy of σ becomes
important in the chromosphere and the corona. Third, the effect of plasma
turbulence (i.e., a variety of waves depending on the electromagnetic restoring
force which acts on charged particles) is to reduce the conductivity, mostly
in combination with an anisotropy. Plasma turbulence must not be confused
with the hydrodynamic turbulence of conducting matter which increases the
diffusivity of the average magnetic field. This important aspect will be treated
in Sect. 8.4.2 below.
The magnetic Reynolds number Rm shown in Fig. 8.1 is large, and a large
Rm is also obtained even if we take the smallest possible σ in a sunspot,
say 1 A/Vm (Fig. 8.2), together with l = 100 km (the order of the scale
height) and v = 1 km/s (a typical measured velocity). Thus we see that the
solar plasma generally has high conductivity, and in this limit the important
concept of a frozen magnetic field has wide applications.
where the second integral is taken along the closed curve S. We use B · (v ×
ds) = (B × v) · ds, apply Stokes’ theorem to convert the line integral into
an integral over F , take the limit dt → 0, and obtain
dΦ & '
= Ḃ − curl (v × B) · df . (8.14)
dt
F
By virtue of the induction equation (8.6), the right-hand side is zero for
σ → ∞. That is, in the limit of infinite conductivity the total magnetic flux
enclosed by the curve S is conserved. Since the choice of S is arbitrary, and
since the constant flux can be represented by a fixed number of field lines,
we may say that the field lines behave as if they were firmly attached, or
“frozen”, to the moving fluid. The concept of a frozen field will help to com-
prehend many of the solar phenomena treated in the present chapter, as well
as in the subsequent one.
Problem 8.3. Use the equation of continuity to show that for a frozen field
the equation
d B B
= ·∇ v (8.15)
dt ρ ρ
holds, a result first derived in 1946 by C. Walén (cf. Cowling 1976). Give
an interpretation.
Problem 8.4. A velocity field v = (−αx, −αy, 2αz) with α > 0 is given.
Find the steady solution of the induction equation (with constant η) un-
der the assumption that B points into the z-direction. What is the central
field strength of the flux tube generated by the converging flow if the total
flux is given? Calculate the radius of the circle that encloses 90 % of the
flux (Moffatt, 1978).
A single particle which carries a charge e and moves with velocity v across a
magnetic field B experiences a force
ev × B . (8.16)
This force is perpendicular to both v and B. It causes the well-known gyra-
tion of the particle around the magnetic lines of force.
In the present chapter we shall not study individual particle motions,
nor shall we separately consider individual species with their diverse charges,
8.1 Fields and Conducting Matter 311
i.e., ions, electrons, and neutrals. Instead we shall treat the solar plasma as
a single fluid. In combination with the neglect of the displacement current in
(8.2), and the non-relativistic form (8.5) of Ohm’s law, this constitutes the
magnetohydrodynamic approximation.
The volume force exerted by a magnetic field on a conducting fluid is the
Lorentz force,
j×B . (8.17)
We do not rigorously derive this expression here. But it is clear that the
product ev in (8.16) describes the transport of electric charge, i.e., the moving
particle contributes to the electric current. The sum over those products for
all charged particles contained in a small volume δV , divided by δV , yields
the net current density j. Since in the one-fluid model the forces that are felt
by the individual charged particles are communicated to the fluid as a whole
by means of collisions, we obtain a net force that is proportional to j, namely
(8.17).
The volume force (8.17) can be divided into a magnetic pressure gradient
and a magnetic tension: using (8.2), and B for the absolute magnitude of B,
we find
B2 B
j × B = −grad + (B · grad ) . (8.18)
2μ μ
Instead of pressure and tension we also speak of Maxwell stresses. With the
definition
1 1
Bik = δik B − Bi Bk
2
(8.19)
μ 2
the ith cartesian component of the volume force is
∂
− Bik . (8.20)
∂xk
As described by the first term of (8.18), a bundle (“rope”, “tube”) of
magnetic flux applies a lateral pressure to the gas into which it is embed-
ded. To obtain equilibrium, this magnetic pressure must be balanced by the
gas pressure. A pressure of 104 Pa is typical for the solar atmosphere; such
pressure may balance a field of up to ≈ 0.15 T, which indeed is the strength
of the field concentrations seen on the solar surface. In sunspots, where we
see into the somewhat deeper level corresponding to ≈ 2 × 104 Pa, the field
strength may reach a maximum value about twice as high.
The effect of magnetic tension is that the lines of force have a desire
to shorten themselves. This effect (but also the pressure effect) provides a
restoring force to perturbations. The fluid is thus able to support particular
wave motions, in addition to those which depend on the compressional and
gravitational restoring forces.
312 8. Magnetism
Problem 8.5. Confirm (8.18) to (8.20). Convince yourself that the two terms
of (8.18) have opposite components in the direction of B. Write the vol-
ume force in cylindrical coordinates and calculate the magnetic force for
the field of Problem 8.4.
vx = −u sin(πx/l) cos(πz/l) ,
vy = 0, (8.21)
vz = u cos(πx/l) sin(πz/l) .
Fig. 8.4. Concentration of magnetic flux at the edges of convection cells. Each box
is a vertical cross section; the flow is clockwise, the numbers give the time in units
of 5l/8u. From Galloway and Weiss (1981)
8.2 Flux Tubes 313
The magnetic Reynolds number, here ul/η, is 250; the boundary condi-
tions are such that the field is vertical at all times at all boundaries. Initially
the field is homogeneous and of strength B0 . The clockwise flow first takes
this field along as if it was frozen into the fluid. But later on the lines of
force are more and more deformed and finally sharply bent. This means that
the diffusion term of the induction equation (8.6), which contains the now
growing second derivative, is no longer negligible. Dissipation, in the figure
visible in the form of field-line reconnection, sets in and finally dominates
almost the whole volume. The magnetic flux is thus expelled from the cell
interior, and accumulated in sheets near the cell edges.
one of the hexagonal cells is B0 l2 initially, and Bd2 in each of the final
tubes. We equate the initial and final fluxes, use (8.22), and obtain
B Rm
1/2
B0 (8.23)
B Rm B0 (8.24)
in a flux tube. The numerical results confirm these estimates, although a non-
negligible fraction of the total flux appears at the top center of a hexagonal
cell (see the lower right of Fig. 8.6).
The process of field amplification is rapid: it takes only a time of order
l/u, the advection or “turnover” time, to produce the sheets or tubes from
the initial homogeneous field. The subsequent expulsion of flux from the cell
interior is somewhat slower and depends on Rm . In the examples shown in
Figs. 8.4 and 8.6 the final state is essentially reached after about 5 l/u.
8.2 Flux Tubes 315
In the following, we shall only discuss the flux tubes, although sheets,
which are produced in a two-dimensional flow geometry, may as well exist
on the Sun, cf. the chain-like crinkles described in Sect. 8.2.2 below or the
elongated intergranular feature reported by von der Lühe (1987).
The main question now is the final strength up to which the field is am-
plified. Expressions (8.23) and (8.24) only give the final field in terms of the
initial field B0 . However, as the magnetic force at some stage must become
strong enough to counteract the motion, there should be a limit which is
independent of B0 . Several estimates have been made at this dynamical field
limitation. Here I mention only the simplest of them: namely that there is
equipartition between the kinetic and magnetic energy densities when the
Lorentz force (8.17) begins to matter (the dynamic regime). In this case
B = Be ≡ (ρμ)1/2 u . (8.25)
Φc = lη(ρμ)1/2 . (8.26)
The equipartition field (8.25) and the critical flux (8.26) are listed in
Table 8.1 for the three prominent cell types that we have treated in Chap. 6.
It is interesting to compare the flux values given there to the fluxes of observed
magnetic features. The latter range from ≈ 109 Wb for the smallest flux tubes
to ≈ 1014 Wb for an active region.
For comparison, the field BP which corresponds to equilibrium between
magnetic and gas pressure (for the deep part of the respective convection
cells) is also listed in Table 8.1. Generally BP is much larger than the equipar-
tition field (8.25).
Table 8.1. Critical flux and magnetic field strength in the Sun’s
convection zone, after Galloway and Weiss (1981). The magnetic
Reynolds number Rm is calculated for ηt = 2 × 107 m2 /s
For the evaluation of Φc in Table 8.1 a diffusivity has been adopted which
is orders of magnitude larger than the value corresponding to the conductiv-
ity given in Sect. 8.1.2. The argument used by Galloway and Weiss (1981)
is that there is still a turbulent motion within the flux tube that tends to
diffuse the field (cf. the result of Sect. 8.4.2 below). The actual value taken,
ηt = 2 × 107 m2 /s, is however not derived from a rigorous theory; it corre-
sponds to the observed decay rate of sunspots (Sect. 8.3.2). It must be said
that the whole problem of diffusion in flux tubes is not really understood: On
the one hand, we argue that the field B is strong enough to exclude a flow of
order u from the tube, on the other hand we need motions inside the tube.
And what is worst, those internal motions should be larger than u in order
to provide the desired diffusive effect!
Problem 8.6. Consider magnetic sheets and tubes with average distance l
from each other. How does the rms field strength depend on Rm in the
two cases?
eigenvalues positive, which is necessary for stability. For the lower boundary
at a depth z1 = 5000 km the result is βc = 1.83, for z1 = 195 000 km (about
the total depth of the convection zone) βc = 1.51. From these values of βc and
from (8.28) and (8.31) it is straight forward to see that stable static flux tubes
are possible for a minimum field strength of order 0.1 T. At such strength the
field is capable of suppressing the convective instability. Numerical work con-
firms the transition into the strong-field state (Grossmann-Doerth et al 1998),
although this final state appears to be time-dependent rather than stationary
or static. An example is shown in Fig. 8.7.
Fig. 8.7. Vertical cut through the solar atmosphere, with gas flow (arrows) and
concentrated magnetic field (thin solid curves) in a two-dimensional numerical sim-
ulation. The thick curve near z = 0 marks optical depth τ = 1, the temperature
scale (in K) is given at the top. Courtesy O. Steiner
The general picture of the photospheric magnetic field can now be sum-
marized. The field is very weak in the major fraction of the solar surface,
but has values around or above 0.1 T in local spots, the flux tubes. We may
say that the field is intermittent. The first cause for this field structure is
concentration of flux by converging motions in the highly conducting solar
plasma. The second cause is the convective collapse, which transforms the
weak concentrations into strong and narrow tubes. We shall see in the fol-
lowing section that observations confirm this picture (in fact the observations
preceeded the theoretical interpretation outlined here!).
Fig. 8.8. Pores and a small sunspot with a one-sided penumbra, in a 32 × 24
area. Notice the regions showing the fuzzy “abnormal” granulation. From the Dutch
Open Telescope, La Palma
line clearly shows the Zeeman splitting. Because, according to (3.55), ΔλB /λ
is proportional to λ, the splitting is easily detectable. A field strength of
0.17 T has been measured in this example.
Magnetic knots are abundant in the vicinity of sunspots. Up to a distance
of 8 × 107 m from a spot’s center, there are ≈ 10 knots per 100 granules. The
polarity of the knot field predominantly is opposite to the polarity of the spot
to which they “belong”. The combined net magnetic flux of all knots is of
the same order as the flux in the spot. This suggests that the lines of force
which leave a sunspot essentially return to the solar surface and re-enter in
the surrounding knots.
The lifetime of magnetic knots is at least 1 hour. Since the knots mostly
are associated with downward motion, and are situated in the dark inter-
granular lanes, it is suggestive to assume a close relationship between their
existence and the converging granular and supergranular velocity field.
0.17 T
Fig. 8.10. A magnetic knot in the vicinity of a sunspot near disc center. The
profile V (λ) of the Fe I line at 1564.8 nm has been measured at the center of the
small circle. From the Tenerife Infrared Polarimeter at the German Vacuum Tower
Telescope. Courtesy R. Schlichenmaier
8.2 Flux Tubes 321
exit slits of the magnetograph (Fig. 3.40), and thus the measured V no longer
increases.
The second reason why the field strength is not immediately measured in
the magnetograph is the uncertainty about spatial resolution. In Fig. 8.12,
for example, features down to about 1 are discernable; but we do not know
whether we measure a field of uniform strength (over the resolved area) or a
field that occupies only a fraction of the resolved area (but is much stronger
there), or even a field that has sign reversals within the resolved area: What
is measured is the average field strength B over the resolved area or, which
is equivalent, the total magnetic flux enclosed by that area.
In order to obtain information about the true field structure and field
strength, the above-mentioned saturation of the polarimeter signal can be
exploited. It is necessary for this purpose to measure simultaneously in at
least two lines having different Landé factors g ∗ . Except for the g ∗ values,
the two lines should be formed in the same layer of the solar atmosphere,
and should have the same excitation energy of the lower level so that their
temperature dependence is the same.
Let us concentrate on the longitudinal Zeeman effect. As long as the field
is weak we may expand (3.78):
dI0
V (λ) = ΔλB , (8.36)
dλ
which means that V (λ) is proportional to the field strength B and has its
maximum value Vmax at the wavelength where the line profile is steepest. In
addition, (8.36) says that V (λ) is proportional to g ∗ , cf. (3.55). The two iron
lines at 524.706 nm and 525.022 nm shown in Fig. 3.41 satisfy the conditions
mentioned, and have g ∗ values of 2 and 3 respectively, but obviously their
circular polarization is not in the ratio 2 : 3. Instead, the two Vmax values are
almost equal. Hence (8.36) is not applicable; the V profile saturates because
ΔλB becomes comparable to the line width ΔλD . The example of Fig. 3.41
is of particular beauty because the measurement has been made with the
Fourier transform spectrometer and the saturation therefore has nothing to
do with the position of a magnetograph exit slit.
If saturation can be recognized, a crude estimate at B is obtained by
setting ΔλB = ΔλD , and by using g ∗ = 3, ΔλD = 42 mÅ (Sect. 3.5.4), and
(3.55). The result is B = 0.11 T. The comparison of the V profiles of two lines
thus leads to the conclusion that strong unresolved field concentrations must
exist on the Sun. Since the average field B generally is smaller by orders of
magnitude, we must conclude further that the field strength is high only in
a small fraction of the observed area. Numerous studies confirm this result.
The next question is the structure of the strong unresolved field. Is there
a single large flux concentration or a larger number of small flux tubes? A
possible answer comes from counts of bright points (see below) or spicules
(Chap. 9), which are seen in filtergrams and are related to the magnetic field.
8.2 Flux Tubes 323
Fig. 8.13. The solar filigree at distances +7/8 Å (lower left), −7/8 Å (lower right),
and +2 Å (upper right) from the center of Hα. The continuum (upper left) shows
“abnormal granulation” (Dunn and Zirker, 1973). Courtesy National Solar Obser-
vatory, AURA, Inc.
white light near the solar limb (Fig. 8.14), and with the help of filters also
elsewhere on the disc. Observing at disc center during extremely good seeing,
Mehltretter was able to resolve these features into a number of small bright
points. The facular points also have sizes of about 0.25 or less, and clearly
sit within the dark intergranular lanes (Fig. 8.15). Simultaneous photographs
in Ca II and in the wing of Hα suggest a close correspondence to the grains
and crinkles. We may assume that the two phenomena represent the same
solar feature, the filigree.
8.2 Flux Tubes 325
Fig. 8.16. G-band bright points. CCD image obtained at the R. B. Dunn Solar
Telescope at Sacramento Peak Observatory, with a filter of 1 nm width centered
at 430.5 nm; the exposure time was 6 s, and an adaptive optics system, locked on
the small pore in the center, was used. The contrast and resolution is best in the
central part, but gradually deteriorates as one moves towards the edges, away from
the isoplanatic patch. The displayed area is approximately 40 × 40 . Courtesy
T. R. Rimmele
The sharpest images of bright points have been obtained in the G band
(Fig. 8.16). The G band is a band head at ≈ 430.5 ± 1 nm which is dominated
by absorption lines due to electronic transitions of the CH molecule, along
with changes of the rotational and vibrational energy. Due to the enhanced
temperature in the visible layer of a small magnetic element (Fig. 8.26) there
is enhanced dissociation of CH, and therefore less absorption, as illustrated
in Fig. 8.17. The high contrast of the G-band images is due to the large
sensitivity to temperature variations of the CH dissociation; the bright-point
intensity may exceed the average quiet-Sun intensity by up to 30 %.
326 8. Magnetism
Fig. 8.17. Mean spectrum of several G-band bright points (thick curve), and mean
spectrum of the quiet Sun (thin curve). The CH lines are weaker in the bright-point
spectrum, while the Ca I line at 430.254 nm and the Fe I line at 430.318 nm have
about the same strength in both spectra. After Langhans et al. (2001)
Magnetograms such as Fig. 8.12 do not resolve a flux tube of the size of
a crinkle or bright point. But counts made by Mehltretter have shown that
the number density of facular points is roughly proportional to the mean
magnetic field strength, see Fig. 8.18. If we divide this mean field strength
by the number density of facular points, we obtain the average magnetic flux
per point. Mehltretter’s result is 4.4 × 109 Wb, which, with a field of 0.1 T as
deduced above, leads to d ≈ 200 km for the diameter of the flux tubes. The
fact that the visible features become larger at greater height (i.e., towards
the core of Hα) can be explained by the spatial divergence of the magnetic
lines of force.
Finally, a direct proof that bright points are magnetic features was pre-
sented by Keller and von der Lühe (1992). Using speckle interferometry they
were able to reconstruct images in polarized light with a diffraction limited
In this and the following section we shall describe the magnetic field and
the motion in an isolated tube of magnetic flux, i.e., within a bundle of
lines of force that is surrounded by a plasma with zero magnetic field. The
complete treatment of such a configuration would consist of a solution of the
appropriate equations both in the tube interior and in the exterior space, and
of combining these two solutions by a fit across the tube boundary. In the
special cases where such treatment has been applied it confirms the results
obtained by the method of approximation used in the following.
The approximation of thin or slender flux tubes is valid as long as all
length scales along the tube are large in comparison to the tube diameter. In
the present case, where we consider an axisymmetric vertical tube, with its
axis in the z-direction (upwards, cf. Fig. 8.19), and with a cross section of
radius RT (z), this means that
RT /H 1 , kRT 1 (8.37)
for all z; H is the scale height of any quantity, e.g., pressure, in the tube,
and k is the vertical wave number of any perturbation propagating along the
tube.
We shall here neglect all dissipative processes. That is, we assume that the
plasma is inviscid and of perfect electrical conductivity, and that all changes
of state occur adiabatically. The set of equations which must be solved is
then: the equation of motion,
ρ̇ + ∇ · (ρv) = 0 , (8.39)
Ḃ = curl (v × B) , (8.40)
and the condition of adiabatic change, cf. (8.30), (8.35),
etc., while the s and φ components of v and B have only odd terms, e.g.,
Let us make two modifications of the scheme just outlined. The first is
to replace vφ by the angular velocity Ω = vφ /s; Ω has an even expansion
such as (8.42). The second is the introduction of a vector potential A for the
poloidal field components Bs and Bz . Thus
where
A = (0, A, 0) , (8.45)
and
∂A 1 ∂
Bs = − , Bz = (sA) . (8.46)
∂z s ∂s
In this way we reduce the number of variables by one, and at the same time
ensure that always div B = 0. From (8.46) it is clear that A has an odd
expansion of type (8.43).
We should stress here that the expansions (8.42), (8.43) must not be
confused with a linearization. Even in the lowest order our problem will keep
the essential non-linearities of the full magnetohydrodynamic equations (8.38)
8.2 Flux Tubes 329
(Landau and Lifschitz 1967, p. 265). Let n̂ be the local unit vector normal to
the boundary, and n and τ coordinates in the normal and tangential directions
(Fig. 8.19). Because of the perfect conductivity there is no flow across the
magnetic field, i.e., vn = 0 at s = RT , and v = (0, v τ ) in the n, τ -system.
Hence
vs = vτ sin α , (8.55)
vz = vτ cos α . (8.56)
Using (8.55) to (8.57), and expression (8.18) for the magnetic force, we
may calculate the projection of the equation of motion, (8.38), onto the n-
direction. We obtain
∂ B2
P+ =C , (8.58)
∂n 2μ
where C is a combination of tangential derivatives, time derivatives, and
other bounded terms. Integration of (8.58) across the boundary from n = −ε
to n = +ε yields (in the limit ε → 0) the condition that P + B 2 /2μ must be
continuous, or that, at s = RT ,
B2
P+ = Pe . (8.59)
2μ
The external magnetic field is zero; Pe is the external pressure. We again use
our expansion, this time at s = RT , and the transformation (8.55), (8.56) to
obtain
2 2 2 1 2 2
P0 + A1 + RT P2 + (16A1 A3 + A 1 + Bφ1 ) = Pe . (8.60)
μ 2μ
The radius RT may be eliminated in terms of the total flux
RT
Φ = 2π Bz s ds = 2πRT (z)A(z, RT (z)) 2πRT
2
A1 . (8.61)
0
2
The last expression on the right is of sufficient accuracy because RT itself is
a second-order term.
If we go to the second order in the expansion we must use (8.60) as it
stands; for problems which we treat only in the lowest order we may drop the
2
term proportional to RT . Notice that in the latter case the total flux does
not matter: That is, to leading order, tubes with various total fluxes behave
in the same way on their axis, s = 0.
8.2 Flux Tubes 331
which means that it takes four pressure scale heights for the tube radius to
expand by a factor e. In the solar atmosphere this is about 500 km. Thus, a
tube which is thin at optical depth τ500 = 1 may no longer be thin [in the
sense of (8.37)] at the level of the temperature minimum. On the other hand,
(8.67) ensures that even for tubes that violate (8.37) in the photosphere the
present approximation will be quite satisfactory in the underlying convection
zone.
Problem 8.7. For given T (z) and given total flux Φ calculate explicitly the
functions Bz0 (z), P0 (z), and RT (z) for a static thin vertical tube.
332 8. Magnetism
Fig. 8.20. Longitudinal tube wave Fig. 8.21. Torsional tube wave
8.2 Flux Tubes 333
Problem 8.9. Calculate the second-order (in s) terms of the torsional Alfvén
wave in a thin flux tube without stratification.
The axis of the flux tube constitutes a curve in 3-dimensional space. Let
l be the length measured along this curve. At each point we may define a
set of three orthonormal vectors, namely the tangent vector l̂, the principal
normal n̂, and the binormal b̂ = l̂ × n̂ (Fig. 8.22). The differential geometry
of space curves tells us that, going along the curve, these three unit vectors
change according to Frenet’s equations:
∂ l̂/∂l = κn̂ , (8.76)
∂ n̂/∂l = −κl̂ + τ b̂ , (8.77)
∂ b̂/∂l = −τ n̂ , (8.78)
where κ is the curvature (the inverse of the radius of curvature), and τ the
torsion of the curve.
The motion of the flux tube is governed by (8.38). We decompose the
volume force, F , on the right of that equation into its components parallel
to the three unit vectors, thereby essentially following Spruit (1981a). We do
this in the lowest order of the thin-tube approximation, in which the field is
in the direction of l̂, and the Lorentz force (8.17) is perpendicular to l̂. In the
direction of l̂ we therefore have
Fl = −∂P/∂l + ρ l̂ · g . (8.79)
For the transverse components we use the form (8.18) of the Lorentz force.
To lowest order B = B l̂ and
∂ 1 ∂B 2
B · ∇B = B (B l̂) = l̂ + κ B 2 n̂ , (8.80)
∂l 2 ∂l
where (8.76) has been used. Let n and b be coordinates in the directions of
n̂ and b̂. Then
1 ∂B 2 κ
Fn = −∂P/∂n + ρ n̂ · g − + B2 , (8.81)
2μ ∂n μ
8.2 Flux Tubes 335
and
1 ∂B 2
Fb = −∂P/∂b + ρ b̂ · g − . (8.82)
2μ ∂b
Now we have already seen in the preceding section that, to lowest order,
the internal pressure plus the magnetic pressure are laterally balanced by
the external pressure Pe . This result holds for curved tubes as long as the
curvature is weak (here expressed through κRT 1). We may therefore
replace P + B 2 /2μ by Pe . Using ∇Pe = ρe g we thus obtain
Fn = (ρ − ρe ) n̂ · g + κ B 2 /μ , (8.83)
Fb = (ρ − ρe ) b̂ · g . (8.84)
We have P < Pe and, assuming that the temperature is the same inside and
outside the tube, ρ < ρe . Hence F points into the opposite direction of g, i.e.,
upwards. This upwards directed force is called magnetic buoyancy. Its effect
is a rise of the horizontal flux tube as a whole (Parker 1955 a, Jensen 1955).
336 8. Magnetism
Let α be the angle between the tube axis and the vertical, so that tan α =
∂ξ/∂z ≡ ξ . As ξ is small, α is also small, and l̂ = (−ξ , 1), n̂ = (−1, −ξ ) to
first order in α. Using again (8.77) we find κ = ξ . Hence
B 2
Fn = (ρ − ρe ) gξ + ξ . (8.88)
μ
Before we substitute this into the equation of motion we must find a way to
incorporate the perturbation of the external medium. This perturbation was
neglected in our discussion of the convective collapse in Sect. 8.2.1 above and
also in the derivation of (8.83), but this time the flux tube will move fully into
the external medium and this certainly will have some effect. An approximate
way is to neglect both compression and curvature at this point. Then we may
use the result valid for a circular cylinder in a liquid, namely that the inertia
of the cylinder is apparently increased by the mass of the displaced liquid
(Basset 1961, Vol. I, p. 186). This result is plausible: obviously some of the
surrounding material must be accelerated to give room for the transverse
motion of the tube. Following this reasoning, and making use of the small
amplitude of the perturbation, we obtain (ρ + ρe ) ∂vn /∂t = Fn , or, with
vn = ξ˙ and (8.72),
ρ − ρe ρ
ξ¨ = gξ + c2 ξ . (8.89)
ρ + ρe ρ + ρe A
8.2 Flux Tubes 337
equation (8.77) then gives κ = −∂ 2 ζ/∂x2 . Hence, (8.79) yields the following
equation for the longitudinal perturbation:
1 ∂P1 ∂ζ
ξ¨ = − −g . (8.95)
ρ0 ∂x ∂x
The transverse equation follows from the perturbation of (8.83):
ρ1 ζg(1 − ∇) ∂ 2 ζ B02
ζ̈ = − g− + , (8.96)
2ρ0 2HP (1 + 1/β) ∂x2 2μρ0
where, in the same way as explained above in the context of transverse tube
waves, the external medium has been included by dividing the force Fn by
(ρ + ρe ) instead by ρ only. The external equilibrium density has been re-
placed by ρ0 , and ρe1 = −ρ0 ζ/Hρe = −ρ0 ζ(1 − ∇)/[HP (1 + 1/β)] has been
substituted.
The remaining equations are the linearized forms of the lateral equilib-
rium (8.28), the condition of adiabatic variation (8.41), and the frozen-field
condition, which we take in the form (8.94):
P1 2B1 ζ
+ =− , (8.97)
P0 βB0 HP
P1 ρ1
=γ , (8.98)
P0 ρ0
Ḃ1 ρ˙1 ∂ ξ˙
− = . (8.99)
B0 ρ0 ∂x
Following Spruit and van Ballegooijen (1982), we seek solutions of form
exp(iωt + ikx), introduce a dimensionless frequency ω̃ by ω̃ 2 = ω 2 βHP /g and
a dimensionless wave number α̃ = kHP , and obtain for the system (8.95) to
(8.99) the dispersion relation
ω̃ 4 + A ω̃ 2 + C = 0 , (8.100)
where [with ∇a = (γ − 1)/γ]
γβ α˜2 1 1 β2 ∇ − ∇a
A = −α̃2 − + − + β2 , (8.101)
1 + βγ/2 2 γ (1 + β)(2 + βγ) 2(1 + β)
and
, -
β β 1
C = γ α̃2 α̃2 − + β(∇ − ∇a ) . (8.102)
1 + βγ/2 2(1 + β) γ
Problem 8.12. Confirm (8.100) to (8.102) and show that the condition for
real ω̃ 2 , namely A2 > 4C, is always satisfied.
8.2 Flux Tubes 339
6 8 10 12 14 16 18 20
80 80
300
10000
100
50
20
10000
60
latitude [deg]
60
0
30
50
10000
100
40 40
10000
20
20 m=1
20
m=2
10000
stable
6 8 10 12 14 16 18 20
magnetic field: B0 [104 G]
Fig. 8.24. Instability of toroidal flux tubes, as function of magnetic field strength
and heliographic latitude (Ferriz Mas and Schüssler 1995), for ∇−∇ad = −2.6·10−6 .
The labels on the contours are growth times in days, m is the azimuthal wave
number of the unstable modes. Courtesy P. Caligari
340 8. Magnetism
of the order 1013 Pa (Table 2.4 or 6.1); and |∇a − ∇| ≈ 10−6 . . . 10−5 . Using
(8.105) and models of overshooting convection as described in Sect. 6.2.3,
Pidatella and Stix (1986) showed that stable flux tubes with a field strength
of several tesla are permitted in such a layer.
A general stability analysis must include the effects of rotation. In this
case the critical field strength depends on latitude. Instead of (8.100) one
obtains a dispersion relation of sixth order for a generally complex frequency
(Ferriz-Mas and Schüssler 1995), which must be solved numerically. Non-
axisymmetric perturbations now become especially important; Figure 8.24
shows a typical result with unstable modes of azimuthal wave number m = 1
and m = 2. In a non-axisymmetrically perturbed tube the gas flows from the
outward displaced crests along the tube toward the troughs. Hence the crests
become lighter and continue to rise. Parker (1966) has considered this process
first in the context of the galactic magnetic field. Therefore the instability
discussed here has been named the Parker instability.
If a facular or G-band bright point represents a thin flux tube, and a pore
represents a thick flux tube, then we have yet to explain why the former
appears as a bright feature, while the latter is dark (and why magnetic knots
are entirely invisible in the continuum).
A strong magnetic field inhibits the convective transport of energy, as
was first recognized by Biermann (1941) in the context of sunspots. The
reason is that the field is frozen to the plasma, but too strong to suffer
the kinematic effects discussed in Sect. 8.2.1 above; it rather prevents the
overturning motion. A consequence is a reduced energy input from below
and hence a reduced temperature at any given level, both in the thin and the
thick tube, as illustrated in Fig. 8.25. The difference between thin and thick
tubes is that the lateral influx of heat, which is only by radiative transport
Fig. 8.25. Temperature profiles across a very thin (a) and a moderately thin (b)
flux tube. The lower dashed curve is for the geometric level where τ = 1 in the
photosphere, the upper for the (deeper) level where τ = 1 in the tube. The solid
curve indicates the temperature actually observed. Adapted from Spruit (1981 b)
8.2 Flux Tubes 341
and therefore inefficient, cannot heat the thick tube (except a thin outer
part), but does suffice to heat the thin tube up to a temperature that is
above the photospheric temperature (at a geometrically higher level). Now,
because the opacity decreases with decreasing temperature and decreasing
density (the density is reduced in the tube) we see to deeper geometrical
levels in the tube than in the surrounding photosphere. The result is that in
the thin tube we see material which is hotter, and thus brighter, than the
photospheric material, while the opposite is true for the thick tube. Spruit
(1976), who first presented this argument, found that the transition from
thin bright tubes to thick dark tubes should occur around 600 km for the
tube diameter. In this intermediate region the magnetic knots should then
be found.
The influx of heat from the surrounding photosphere into the tube pro-
duces a “hot wall”, which becomes visible when the tube is inclined. As a
consequence, tubes that are invisible or dark near the disc center may appear
as bright features, e.g., faculae, towards the limb.
More detailed calculations confirm the qualitative reasoning of Fig. 8.25.
Semi-empirical models of thin flux tubes, of the kind explained in Chap.
4, have been constructed by Solanki (1986) and have a higher temperature
than the surrounding photosphere at the same optical depth, as anticipated
in Fig. 8.25a. In order to obtain an intensity for the tube models Solanki took
the profile V (λ) of the Stokes parameter describing the circular polarization,
measured with a Fourier transform spectrometer, and inverted (8.36):
λ
1
IV (λ) = IC + V (λ ) dλ . (8.106)
ΔλB
λ1
8.3 Sunspots
It has been known since the work of G. E. Hale in the early 20th century
that sunspots possess a magnetic field. Indeed, ever since the field has been
considered as the primary cause of the spots’ darkness.
Because the magnetic field is source-free, the total magnetic flux through
the solar surface is always zero. What we can expect, then, is eruption of flux
in form of loops. These loops lead to the characteristic bipolar structure of
spots. Nevertheless, such structure is not a necessity. Often the flux of one
polarity is more concentrated than that of the other, so that only a single,
“unipolar” visible spot is formed.
Let us discuss the various forms of sunspots and sunspot groups in terms
of the Zürich classification (Waldmeier 1955). This classification is largely a
time sequence which, if all stages come about, may take several months. It
distinguishes unipolar and bipolar configurations, the size and complexity of
the spot (group), and the presence or absence of a penumbra, i.e., the ray-
structured halo of intermediate mean intensity that in general surrounds the
dark central part, the umbra, of large spots. The sunspot darkness depends
on wavelength, cf. the intensity scans shown in Fig. 8.27.
Fig. 8.27. Intensity profiles across a sunspot, at three wavelengths. Adapted from
Wittmann and Schröter (1969)
Fig. 8.28. Arch-filament system (center) in a small sunspot group, seen in Hα. Slit-
jaw photograph (the black line is the entrance slit to the spectrograph), A. Bruzek,
Domeless Coudé Telescope, Capri
344 8. Magnetism
For more than half of the sunspot groups the evolution terminates, after
a day or a few days, with state A or B. They never develop a penumbra. For
those that do the Zürich classification continues.
Mostly, but not always, the larger spots have a penumbra, while the
smaller ones have none. Experienced observers have seen spots without a
penumbra as large as 11 000 km in diameter, and spots with penumbra as
small as 2500 km in (umbral) diameter (McIntosh 1981). The typical radial
extent of the penumbra is 5000 km. It appears as if the penumbra forms out
of radially aligned granules surrounding the umbra, a process that may be
completed within one hour.
Again, the growth of a sunspot group may end with state C or D, and
be followed immediately by the decay states explained below. On the other
hand there are further growing groups, classified as E and F.
E: Large bipolar group, extending over more than 10 o on the Sun. Both prin-
cipal spots have a penumbra and often a complex structure. Numerous
small spots are present.
F: Very large, and very complex, bipolar group, extending over ≥ 15 o .
The following three classes describe the decay of the group, and of the
last remaining single spot.
The H spot often is rather symmetric, and stays without change for several
weeks, except for a very slow decrease in size. It is the “theoreticians’ spot”.
Figure 8.29 shows examples. It has been demonstrated (e.g., Bumba 1963)
that the slow decay of the area, A, of many H spots proceeds at a rate that
is nearly constant:
Fig. 8.29. Sunspots on their path across the solar disc. From Wittmann and
Schröter (1969)
More recent studies have included other spot types, in addition to the
long-lived H spots. Moreno-Insertis and Vázquez (1988) and Petrovay and
van Driel-Gesztelyi (1997) found that a decay rate proportional to the in-
stantaneous spot radius s generally is a better approximation than (8.107):
Ȧ −CD s/s0 , (8.108)
where s0 is the maximum radius, and CD ≈ 32 millionths of the visible
hemisphere per day, i.e., CD ≈ 1.1 × 109 m2 /s. The rate (8.108) yields a
concave (parabolic) decay law for the sunspot area, and it includes the fast
initial decay phase. As the decay proceeds, the ratio of the umbral to total
radius approximately remains constant at a value somewhat less than 1/2
(e.g., Martı́nez Pillet et al. 1993).
McIntosh (1981, 1990) has added two more parameters to the Zürich
classification. These parameters have to do with the shape and complexity of
the largest spot within the group, and whether the group is compact or open,
i.e., densely or sparsely filled with spots (like star clusters). The additional
classification has proved valuable in the prediction of flares (Sect. 9.5): the
more complex the configuration, the more likely is the magnetic instability
that gives rise to the flare.
Sunspots and sunspot groups are only the most conspicuous part of what
is called an active region. The active region comprises field concentrations
346 8. Magnetism
where s is the distance to the spot’s center, and s∗ is the spot radius, including
the penumbra (Beckers and Schröter 1969). The angle of the field vector with
the local vertical is (π/2)(s/s∗ ), as already found by Hale and Nicholson
(1938), and confirmed by later observations. According to these results the
field strength at the outer penumbral boundary of an H spot is one-half of
the central field strength, and the field is horizontal at that boundary.
We conclude this section by defining, as a measure of solar activity, the
sunspot relative number introduced by R. Wolf in 1848:
R = k (10 g + f ) , (8.110)
where g is the number of spot groups and f is the total number of spots
(an isolated spot is also a group). The calibration constant k accounts for
the instrumentation and seeing conditions at the individual observatory. By
coincidence, R is nearly proportional to the total area on the Sun covered
by spots. Since the field strength in sunspots is always of the same order
(≈ 0.2 − 0.3 T), R is also a rough measure of the total absolute magnetic flux
penetrating the visible hemisphere within sunspots.
The solar astronomer distinguishes two types of “sunspot models”. The first
essentially consists in the physical quantities, notably temperature, as func-
tions of height in the sunspot atmosphere, and is determined empirically by
the methods outlined in Chap. 4. The second one is the magnetohydrostatic
model, which is derived theoretically and aims to explain the balance of forces
and energy.
8.3 Sunspots 347
than the limb-side penumbra; this is clearly seen in the examples shown in
Fig. 8.29.
Tables of empirical sunspot models have been published, e.g., by Avrett
(1981 a), and by Maltby et al. (1986). In these tables the pressure at τ500 = 1
exceeds the normal photospheric pressure (Table 4.1). If there was no Wilson
effect this would be in contrast to the magnetohydrostatic model treated be-
low. The geometrical depression, however, does permit the desired pressure
deficit of the spot at any given (geometrical) height.
f yy − y 4 + 2μΔP = 0 , (8.114)
where f is a constant which depends on the spot’s total magnetic flux, and
ΔP (z) is the pressure difference between the spot’s exterior and its axis.
If ΔP (z) were known we could integrate (8.114). In a consistent model,
however, ΔP (z) must be determined by the magnetic field itself: the reduced
convective energy flux leads to a decreased thermal pressure in the spot. It
is this latter part of the model which is difficult to formulate. Deinzer (1965)
found a solution in terms of the mixing-length theory outlined in Chap. 6.
The mixing length served as a free parameter and was reduced inside the spot
in comparison to the exterior (where he took l/HP = 1). Alternatively – and
the results shown in Tab. 8.3 are calculated in this way – the spot’s effective
temperature is chosen freely, whereupon the mixing length is obtained as an
eigenvalue. Each of the models also predicts a Wilson depression, zD , of the
Table 8.3. Effective temperature, mixing length/scale height, central field strength,
and Wilson depression for model sunspots of total flux 5 × 1013 Wb (Deinzer 1965)
Teff [K] l/HP B [T] zD [km] Teff [K] l/HP B [T] zD [km]
right order of magnitude. At large depth, > 107 m say, the field in Deinzer’s
model becomes independent of depth.
Problem 8.14. Show that (8.112) and (8.113) satisfy div B = 0, and that
a = const. describes a line of force. Which geometrical meaning has the
function ζ(z) ? Confirm (8.114) and, for given total flux, determine the
constant f . Calculate f for a field distributed according to (8.109).
The model equations have been solved as a free-surface problem: the geo-
metrical position of the two interfaces, the magnetic field, and the thermo-
dynamic properties of the three domains were determined in such a way that
the magnetohydrostatic balance was satisfied; in addition, the observed mag-
netic field configuration (8.109) and the observed mean heat fluxes in the
three domains were matched. Jahn and Schmidt have calculated their mod-
els for diverse values of the total magnetic flux, total depth, and lateral inflow
of heat. The models require a reduced convective heat transport in the um-
bra and penumbra; they predict separate values of the Wilson depression in
these two domains. Another result is that the penumbra has a depth that is
comparable to its horizontal extent and contains more than 50 % of the total
magnetic flux, in consistence with (8.109). No penumbra was found if the
total magnetic flux was below ≈ 3 × 1013 Wb, because in such small spots the
heat flux available to the penumbra was insufficient. If a penumbra existed,
the magnetopause had a minimum inclination of ≈ 25o from the vertical.
Stability. A perturbation of a magnetohydrodynamic system typically prop-
agates with the Alfvén velocity (8.72). The time a perturbation needs to tra-
verse a sunspot is of the order of 1 hour, or even smaller. Thus, the spot must
be stable against perturbations; i.e., the perturbation must decay, because
if it was growing it would disrupt the whole configuration in a time short
compared to the spot’s lifetime.
Decay of sunspots. We have seen that the observed decay of the sunspot
area can be approximated in two ways: either with a constant rate, or with
a rate proportional to the radius. It will be clear presently that the constant
rate (8.107) means that the destruction mechanism works over the entire cross
section of the spot, while a rate proportional to s, such as (8.108), indicates
that the spot is destructed mainly along its periphery.
In both cases the essential assumption is that the spot decays in a diffu-
sive way, and that a turbulent diffusivity ηt , due to irregular motions in the
deeper part (several 1000 km), describes the process in an adequate manner.
A general derivation of ηt will be described in Sect. 8.4.2 below. Here we
simply use the induction equation (8.6), with v = 0 and η replaced by ηt .
For a vertical field Bz that is independent of depth this means
1 ∂ ∂Bz
Ḃz = sηt . (8.115)
s ∂s ∂s
We first assume that ηt is constant. In this case a solution to (8.115) is
(Meyer et al. 1974, Krause and Rüdiger 1975)
Φ s2
Bz = exp − , (8.116)
4πηt t 4ηt t
where Φ is the total flux. According to this solution, the central field strength
Bm is very large for small t, whereas the field strength in real sunspots is lim-
ited to ≈ 0.3 T by the lateral magnetohydrostatic equilibrium. For this reason
Meyer at al. (1974) used an empirical expression for the flux Φ∗ through the
visible spot, namely
Φ∗ 0.4 A Bm (8.117)
(Bray and Loughhead 1964, p. 216), and a constant Bm . Since the visible
spot is a consequence of reduced energy transport, it is reasonable to assume
that Φ∗ (t) is that part of the total original flux Φ which at time t is within a
circle where Bz exceeds a value Bc , the minimum field strength to affect the
convective energy transport. If sc is the radius of that circle then
sc
Φ∗ = 2π Bz s ds = Φ − 4πηt Bc t , (8.118)
0
which, by (8.117), yields the desired linear decrease of the spot area:
8.3 Sunspots 353
The critical field strength Bc is not precisely known; Meyer et al. (1974)
take Bc /Bm = 1/2. Comparing (8.119) with (8.107) we then find the required
value of the diffusion constant in the spot,
ηt ≈ 107 m2 /s . (8.120)
What happens to the energy flux that is blocked by a sunspot? Solar observers
often have searched for bright rings, i.e., areas of enhanced energy flux around
spots. The evidence is casual, and entirely insufficient to account for the full
spot deficit (Bray and Loughhead 1964).
Measurements made with the Active Cavity Radiometer (Sect. 3.6.1) show
that we must not expect bright rings around spots. The two largest dips in
the curve of Fig. 8.33, which occurred in early April and late May 1980, both
coincide with the appearance of sunspot groups. Moreover, size and intensity
of the spots approximately explain the magnitude of these dips. We must con-
clude that the sunspot flux is truly missing. It reappears neither as a bright
ring nor as small excess flux distributed over a large fraction of the solar disc
(the latter possibility had occasionally been proposed as a replacement of
the evasive bright ring). This result has been confirmed by numerous further
coincidences of spots with luminosity dips.
Problem 8.15. Calculate the expected drop of the solar energy flux at 1 as-
tronomical unit for a sunspot of area A and intensity contrast C, at an-
gular distance θ from the disc center. Assume that the limb darkening is
the same for spot and undisturbed photosphere and use the result (4.53)
of the Eddington approximation (Foukal, 1981).
Fig. 8.33. Solar irradiance variation, measured by the ACRIM instrument on the
Solar Maximum Mission (Willson et al. 1981). Courtesy Jet Propulsion Laboratory,
California Institute of Technology, Pasadena, California
8.3 Sunspots 355
An interesting implication for stars other than the Sun is that transient
drops of their luminosity may be interpreted in terms of stellar spots.
For the Sun, let us apply a model of time-dependent diffusive energy
transport (Foukal 1981; Spruit 1981 c). We may thereby understand that the
missing spot energy is stored in the convection zone.
For the present purpose we neglect the small difference between ∇ and
∇a discussed in Sect. 6.2.1. The result of Problem 6.3 is then, in vectorial
form
1
F C = − ρlv T ∇S . (8.124)
2
We need this vectorial form of the convective transport because with the
sunspot as obstacle the situation is of course no longer spherically symmetric.
We also need time-dependence, because at the time when the spot is formed
the obstacle is “switched on”, and we want to see how the energy begins to
flow around it. Hence
∂S
ρT = −div F C (8.125)
∂t
is the energy equation to be solved (for the sake of simplicity, we neglect the
radiative energy flux). As boundary condition we require that, at any time,
the flux is radiated into space according to the surface temperature. For the
perturbation caused by the obstacle this means
δFr = 4 σ T 3 δT . (8.126)
Foukal and Spruit have solved the problem posed by (8.124) to (8.126),
and have calculated the total energy released by the model surface. The main
result is that as soon as the obstacle is introduced this energy drops by an
amount corresponding exactly to the surface fraction of the obstacle. To re-
turn to the original luminosity, it would be necessary to heat the surface
slightly in order to compensate for the reduction of its radiating part. How-
ever, because of the large thermal inertia (or heat capacity) of the layers to
be heated this recovery takes very long.
A measure for the time required to reach thermal equilibrium is the in-
ternal energy U divided by the luminosity L :
tth U/L , (8.127)
For the whole star this thermal time scale is equivalent (Problem 2.10) to the
Kelvin–Helmholtz time originally introduced in terms of the gravitational
energy.
Using (8.127) we can define the thermal time scale in dependence of the
total depth D of the layer considered. Foukal (1981) takes D = 7500 km for
his model and finds tth ≈ 1 year. Spruit (1981 c) uses the entire convection
zone and obtains tth ≈ 2 × 105 years (which demonstrates the large heat
capacity of the deeper layers). Both values exceed the lifetime of sunspots.
356 8. Magnetism
Fig. 8.35. Left: Number of umbral dots as a function of the diameter, from Sobotka
et al. (1997 a). Right: Brightness distribution of umbral dots. Courtesy A. Tritschler
bility treated in Sect. 6.1 cannot develop in the same manner, because the
magnetic field provides a restoring force which is absent outside the spot.
Instead of the overturning convective motion a growing oscillatory motion
is possible under certain conditions. Such oscillatory convection has been
proposed as an explanation for umbral dots. Once the instability is fully
developed (which we must assume in view of its short time scale) the phe-
nomenon needs not be periodic. Thus, the idea is that parcels of hot gas
squeeze through the magnetic lines of force from beneath and so cause the
visible structure of the spot.
The inhomogeneous sunspot model of Parker (1979 b) is an extreme, but
consistent view of the situation just described. In this model the spot is a
cluster of flux tubes, intermixed with columns of field-free gas. The dots
and grains would be the uppermost tips of those columns. Support for this
model comes from the difficulty to stabilize the deeper, more vertical, part
of the sunspot field against the flute instability, cf. Sect. 8.3.2 above. If the
field-free columns are isolated, then the oscillatory convection is the only
means of transporting energy in the spot; if they are interconnected (while
the magnetic tubes are isolated), then the convection could take on a more
normal form, including some overturning flow. The true subsurface structure
of sunpots is not known at present; the approach with a reduced mixing length
is a substitute for a comprehensive magnetohydrodynamic description.
For the penumbra, a more detailed picture has been proposed by Schmidt
(1991): Convective energy transport by the interchange of entire magnetic
flux tubes over a large depth range. Such flux tubes would move back and
forth from the outer edge of the penumbra (the magnetopause of Fig. 8.31)
to the penumbral photosphere. The part of the tube that just arrives at the
surface has the highest temperature and moves towards the umbra, as the
observed penumbral grains do. Below we shall further discuss this model in
the context of the Evershed effect.
8.3 Sunspots 359
outgoing power is less than the incoming power, sometimes reduced to one-
half. Apparently there is wave absorption in sunspots, but actually it may be
conversion of acoustic into magnetohydrodynamic modes of oscillation (e.g.,
Bogdan 2000).
Three-Minute Oscillations and Umbral Flashes. The three-minute os-
cillations in sunspots are distinct in that the spot has resonances in this
range of periods. As explained in the context of the Sun’s global oscillations,
resonances (or modes) are the consequence of reflecting boundaries. And as
there, the reflecting boundaries are provided by the steep gradient of the
atmospheric parameters that are important for wave propagation.
The speed of sound is proportional to T 1/2 and therefore increases both
towards the interior and towards the corona. The Alfvén velocity, propor-
tional to ρ−1/2 , increases outwards. As model calculations show, a waveguide
is formed between the inwards increasing sound speed and the outwards in-
creasing Alfvén velocity. In a quantitative treatment this yields the three-
minute resonance. Geometrically this resonance occurs rather deep, in the
sunspot’s photosphere. This is not necessarily in contrast to the observation
that the three-minute oscillation has the largest amplitude in the spot’s chro-
mosphere: because of its density dependence the kinetic energy is at least five
times larger in the photosphere (Abdelatif et al. 1984; Lites and Thomas
1985).
The three-minute oscillations also have a signature in the intensity of
spectral lines. Most conspicuous are the umbral flashes, discovered even prior
to the velocity signal (Beckers and Tallant 1969). The flashes are seen in
the sunspot chromosphere as periodic brightenings in the emission cores of
the Ca II lines (Fig. 9.2). In a sequence of K-line filtergrams one can see
them arising in the central part of the umbra and then moving towards the
umbra/penumbra boundary (Moore 1981). The umbral flashes seem to occur
whenever the velocity oscillation reaches a large amplitude, of order 5 km/s.
Excitation of the resonant umbral oscillations is possible by means of the
short-period tail of the global solar oscillations (the p modes), i.e., from out-
side the spot. The exciting amplitude is, however, small; moreover, within the
umbra the flashes move outwards. Therefore a more plausible hypothesis is
excitation from within the sunspot itself. The above-mentioned irregular ve-
locity field beneath the visible layer offers a broad spectrum of perturbations.
From these, the spot atmosphere can select its own resonant modes.
Running Penumbral Waves. In stable, circular sunspots the phenomenon
of running penumbral waves can be observed. These waves were discovered
by Zirin and Stein (1972) in the intensity of Hα, and by Giovanelli (1972)
in the Doppler shift of the same line. In a filtergram taken in off-center
Hα, for example, the periodic Doppler shift produces alternating dark and
bright bands. The bands emerge from the umbra/penumbra boundary and
propagate radially outwards across the penumbra. The period is in the range
3 to 5 minutes; the horizontal phase velocity is typically 10 to 20 km/s.
8.3 Sunspots 361
undergoes a shock, thereby adjusting its pressure to the given end pressure.
Since the flux tubes laterally are in magnetohydrostatic equilibrium, where a
balance of type (8.59) holds, the pressure at each foot point may be derived
from the measured strength of the magnetic field. If a flux tube connects the
inner penumbra, with B1 ≈ 0.15 T, to a magnetic knot or a pore outside
the sunspot with, say, B2 ≈ 0.2 T, then we have P1 > P2 , which drives the
desired flow.
The moving-tube model of Schlichenmaier et al. (1998 ab) is more gen-
eral than the siphon model, as it includes time dependence and does not
rely on a given pressure difference. Magnetic flux tubes emerging from the
deep penumbra are able to transport heat to the penumbral photosphere.
Figure 8.31 schematically shows this convection by the interchange of entire
flux tubes. The path of a single tube starting at the magnetopause has been
modeled in the thin-tube approximation. The tube first rises adiabatically. At
the point where it meets the photosphere it sharply bends into the horizontal.
At this point a high temperature is sustained by the upflow of hot gas within
the tube; this has been interpreted as a penumbral grain (Fig. 8.38). The
model also yields a horizontal pressure gradient along the tube which drives
an outward flow, as in the siphon model. As the gas in the tube reaches the
optically thin photospheric environment, it cools down, and the tube becomes
10 km/s
2000 km 6 10 14
Temperature (1000 K)
Fig. 8.38. The moving-tube model for the Evershed flow in a sunspot penumbra
(lower panel). Penumbral grains (upper panel) are interpreted as hot upflows within
emerging flux tubes. Courtesy R. Schlichenmaier
8.3 Sunspots 363
transparent itself. Thus the bright grain tails off, and the horizontal outflow
is accelerated to its largest velocity in the outer, dark part of the tube.
The observation that the Evershed flow is so sharply confined to the
penumbra may find an explanation such that the flow channels become invis-
ible at the outer penumbral edge. The downflow patches found near that edge
support this view. But even a horizontal channel may cut across the τ = 1
surface, because near the magnetopause this surface is inclined upwards.
We shall now describe the solar magnetic field on the largest possible scale,
that is the field of the Sun as a whole. At the same time, we shall deal with
very long time scales, notably the 11-year cycle of the Sun’s magnetic activity.
The best-known indicator of the solar cycle is the sunspot relative number,
(8.110), yearly means of which are shown by the lower curve of Fig. 8.40. The
cycle was first discovered by Schwabe (1844). Each cycle, lasting from one
minimum to the following, is given a number, beginning with the minimum
around 1755 (Wolf evaluated the notes of many earlier solar observers and
so was able to establish the cycle backwards in time). But not only the spots
appear in cycles. Figure 8.40, for example, shows the latitude where the polar
crown of prominences appears, and the frequency of polar faculae.
G. E. Hale had measured the magnetic field in sunspots for the first time
in 1908. By 1923 numerous spots, of three consecutive cycles, had been ob-
served, confirming the polarity rules formulated by Hale et al. (1919):
8.4 The Solar Cycle 365
The magnetic orientation of leader and follower spots in bipolar groups re-
mains the same in each hemisphere over each 11-year cycle.
The bipolar groups in the two hemispheres have opposite magnetic orienta-
tion.
The magnetic orientation of bipolar groups reverses from one cycle to the
next.
These rules are illustrated in Fig. 8.41. They mean that the magnetic
cycle does not last 11 years, but 22 years. Only very few exceptions to the
polarity rules are known.
Fig. 8.41. Polarity rules for sunspots: N, north, S, south. R and V indicate the red
and violet σ-component of the Zeeman triplet; the preceding spot (in the sense of
rotation) is at the right of each pair. The curves indicate the migration in latitude
of the spot zones. After Hale (1924)
Fig. 8.42. Butterfly diagram of sunspots, according to Harvey (1992 b), and con-
tours of radial mean field (from Mt. Wilson and Stanford magnetograms). The
contours are for ±20 μT, ±60 μT, ±100 μT, . . . , solid positive, dashed negative.
From Schlichenmaier and Stix (1995)
vations of Howard and LaBonte (1981) the total flux at any one time is of
order 1015 Wb. For the horizontal extent of the available area we take the
latitude range where sunspots occur, corresponding to, say, 2.5 × 108 m. If
we take the depth of the convection zone, 2 × 108 m, for the vertical extent,
then Bt ≈ 0.02 T. We should not take more than the depth of the convection
zone because the field is oscillatory and therefore has a very small skin depth
toward the core (Problem 8.16). On the other hand, the toroidal flux may
well occupy only a small fraction of the convection zone, notably the layer of
convective overshooting at its bottom, with a thickness of order 5 × 106 m.
This yields the estimate Bt ≈ 1 T. Depending on whether the latitudinal flux
distribution is smooth or in the form of concentrated flux tubes, the actual
field strength may be still larger, possibly reaching 10 T. As we have seen,
tubes of this field strength become unstable even in the stably stratified layer
of overshooting convection (Fig. 8.24). After the onset of the instability the
tube rises in loops toward the surface, as illustrated in Fig. 8.43.
Problem 8.16. Calculate the skin depth for a field which alternates with a
period of 22 years when the conductivity σ is given by (8.10), and for an
“effective” σ derived from (8.145).
B = B + b , (8.129)
where B may be understood as an average over longitude or, more generally,
as an ensemble average (i.e., an average over a large number of realizations
of a stochastic variable). In any case, we shall assume that the operation of
averaging commutes with the operations of differentiation and integration.
The field b is the fluctuating part of B, and obeys b = 0. Likewise we write
v = v + u . (8.130)
By v we shall mean a motion on the Sun which is of global scale, notably
the solar differential rotation. The field u describes the irregular, or turbulent,
convective motion treated in Chap. 6.
In this section the dynamo problem will be treated in its kinematic form,
where v is assumed to be given, and independent of B. It is however not
necessary to know all details of the velocity field. We shall be content to have
the mean v, as well as the statistical properties of u, i.e., mean values, cor-
relations, etc. The kinematic mean-field dynamo was first treated by Parker
(1955 b), Steenbeck et al. (1966), and Steenbeck and Krause (1969); the fol-
lowing outline rests on their work, and on the treatise of Moffatt (1978).
Substitution of (8.129) and (8.130) into the induction equation (8.6), and
separation of the mean and fluctuating parts yields
∂
B = curl (v × B + E − η curl B) , (8.131)
∂t
and
8.4 The Solar Cycle 369
∂b
= curl (v × b + u × B + G − η curl b) , (8.132)
∂t
where
E = u × b , (8.133)
G = u × b − u × b . (8.134)
The mean electric field E is the crucial quantity. If it was known, we
could solve (8.131) for B. In principle E must be calculated in terms of
B by substituting into (8.133) a solution b of (8.132). That is too difficult
in general, but inspection of (8.132) and (8.134) immediately shows that there
must be a linear relationship between b and B, and hence also between E
and B. We write the latter relationship as an expansion:
The coefficients αij and βijk are pseudo-tensors, because they relate an
axial vector B to a polar vector E. In the kinematic approach they are
statistical properties of the velocity field u, and are independent of B.
First-order Smoothing. The mean electric field E has been calculated
explicitly in the special case where the term G in (8.132) is negligible. This
neglect of second-order terms (in the fluctuating quantities) has been called
the first-order-smoothing approximation or, since no higher than second-order
terms appear in (8.133), the second-order-correlation approximation. This
approximation is justified if either of the following conditions is met:
ul/η 1 or (8.136)
uτ /l 1 (8.137)
(Steenbeck and Krause 1969). Here u is a typical magnitude of u, and l and τ
are typical scales of the variation of u (and b) in space and time, respectively.
Condition (8.136) ensures that G is much smaller than the last term on the
right of (8.132); this is of course the case of small magnetic Reynolds number.
If (8.137) holds, then G is negligible in comparison to ∂b/∂t.
Unfortunately neither of the two conditions is satisfied on the Sun. The
magnetic Reynolds number is large, and observations suggest that uτ /l ≈ 1
rather than uτ /l 1. It is nevertheless instructive to proceed. In order to
obtain qualitative results let us, in addition to G, omit the term containing
v in (8.132). Moreover, due to the high conductivity we may omit the
diffusion term in that equation. A solution to (8.132) is then
t
b= curl (u × B) dt . (8.138)
−∞
370 8. Magnetism
The integrand is to be taken at time t ; the initial field, b(−∞), need not be
written down because certainly it is uncorrelated to u(t) and hence does not
contribute to the subsequent calculation of E.
Problem 8.17. Use (8.138) to calculate E in the special case where u repre-
sents “weakly” isotropic turbulence; that is the case where the statistical
properties of u are invariant under rotation but not generally under re-
flection of the frame of reference. Show that (8.135) takes the form
E = αB − β curl B + . . . , (8.139)
where
∞
1
α=− u(t) · curl u(t − t ) dt , (8.140)
3
0
and
∞
1
β= u(t) · u(t − t ) dt . (8.141)
3
0
α ±lΩ , (8.146)
where Ω is the mean angular velocity of the Sun. The sign of α is opposite
to the sign of the helicity. Its magnitude is less than the estimate (8.146) if u
and curl u are not perfectly correlated. Moreover, α depends on the scale l,
which widely varies in the convection zone. Values of α between a few cm/s
and ≈ 100 m/s have been derived by various authors.
Numerical Determination of α. As an alternative to first-order smooth-
ing, numerical simulation of magneto-convection has been used to calculate
the tensors α and β which appear in (8.135). In this way it is possible to
go beyond the limitations set by (8.136) and (8.137), although the magnetic
Reynolds number Rm of the solar convection zone is still much larger than the
values used in actual calculations. Moreover, a large Rm implies small scales
of the magnetic field, which cannot be resolved in a global numerical model,
such as shown in Fig. 7.13. Therefore the calculations have been made in a
rectangular box, intended to simulate the situation in a particular limited
section of the solar convection zone.
The result shown in Fig. 8.44 is obtained for the simplest case where the
vectors of angular velocity and gravity are parallel to each other; the depth
range covers several density scale heights and includes a stably stratified
region in the lower part. Thus, the location of the computational box would
be on the southern axis of the Sun, at the base of the convection zone. An
external mean horizontal field B H is given. After the magnetohydrodynamic
simulation has settled to a statistical equilibrium, the component of E in the
direction of B H is evaluated and yields
372 8. Magnetism
<w.u>/3u~m. D: H!U,m.
0 .30
0 .20
0.10
-0.00
-O"~
.,
-0.20
-0.5 L:----~::-~_:_:~___,_:_- -0.30 _
0.0 0.5 1.0 1.5 0.0 0.' ..0
Fig . 8.44. Numerical !;i mlllatioll of magneto-convection ncar the base of the f;Olar
convection zone. Left: normalized helicity. night: componcnt 011 of the 0 tcusor.
The abscissa z points downward: sign n~ versals occur at z ~ 0.7. dose to tile \e,·e1
z = I where ~he stratificn tioll becomeS ~tab lc. Shading indicates tile variatioll. the
curves give avcragt'5 and error estimates for a calculation last.ing ~ 20 coherence
times of the turbulent flow. PrOIlI Ossendrijver et a1. (2001 )
(8.1 ,17)
By applying different. ext.ernal llIean fields , otller components of the 0" tensor
have been determiued. Tile present example confirms t he close relation~hip
of the a effect with the helieit.y of tile convect.ive flow: In the southern hemi-
sphere there is positive helicity within the convection zone, and accordingly
negat.ive Ol!; both q uallt.it.ies reverse their signs near t he transition to the
st.able stratification, where sillking parcels are converl.ed into hori7.0ntall'y
diverging flows.
The mean-ficld induction C<luation has been the basis for much work 011
the solar uymullo, as well as for planet.ary, stellar, anu galact ic dynamos.
But although some beautiful results have been obtained , t he a pproximations
mauc mllst llO~ be forgott.en. On t he SUIl , first-order smoot hing is marginal!r
valid at best, and the capability of numerical siltlulation is limited. We should
also be aware that, whenever a t urbulent diffusivity 1)t is llsed (e.g .. Sects.
8.2.1 and 8.3.2). we deal with the mean field (B " rather than with B itself.
(in spherical polar coordinates r, θ, φ). In accord with what was said in the
preceding section, let α be antisymmetric with respect to the equatorial plane,
i.e.,
α(r, π − θ) = −α(r, θ) . (8.148)
The angular velocity will be taken as a symmetric function,
Ω(r, π − θ) = Ω(r, θ) , (8.149)
and, besides rotation, no mean motion shall exist:
v = (0, 0, Ω r sin θ) . (8.150)
We separate the mean field into its poloidal and toroidal parts
B = B p + B t , (8.151)
where
B p = curl (0, 0, A(r, θ, t)) , (8.152)
B t = (0, 0, B(r, θ, t)) . (8.153)
The existence of a vector potential of B p follows from div B = 0 and from
the axisymmetry; the function B had already been introduced as Bt in Sect.
8.4.1.
The mean-field induction equation can now also be separated into its
poloidal and toroidal parts. In the simplest case, where ηt is a constant, one
obtains
Ȧ = α B + ηt Δ1 A , (8.154)
∂Ω ∂ 1 ∂Ω ∂
Ḃ = (A sin θ) − (rA sin θ)
∂r ∂θ r ∂θ ∂r
1 ∂ ∂ 1 ∂ α ∂
− α (rA) − 2 (A sin θ) + ηt Δ1 B , (8.155)
r ∂r ∂r r ∂θ sin θ ∂θ
where
Δ1 = Δ − (r sin θ)−2 . (8.156)
Equations (8.154) and (8.155) demonstrate the crucial role of the α effect.
With α = 0, the poloidal field would decay exponentially, because its equation
is of diffusive nature. And with A gone, B would decay as well, for the same
reason.
By (8.155), the toroidal field is generated from a parent poloidal field by
means of differential rotation and α effect. In solar models the latter effect is
commonly neglected in this equation. The condition for this is
|α| r
2
|∇Ω| . (8.157)
374 8. Magnetism
0.8
0.6
Fig. 8.45. Contours of constant angular
0.4 velocity in a meridional section of the Sun,
derived by the Global Oscillation Network
Group, GONG. The labels are r/r , red in-
0.2
dicates fast rotation at low latitude, blue
slow rotation at high latitude. The dashed
0.0 curve marks the base of the convection
0.0 0.2 0.4 0.6 0.8 1.0 zone. Courtesy M. Roth
Let us consider the special case αΩ0 < 0. This would be the situa-
tion in the transition region from the convection zone to the radiative core:
the tachocline found by helioseismology (Figs. 7.4 and 8.45) is a layer with
∂Ω/∂r > 0 (at low latitude, where sunspots occur), and – on the northern
solar hemisphere – we have α < 0 at this depth, see above. Without loss
of generality we may assume k > 0. The frequency ω, which is generally
complex, is then
must not be negative. Hence the solution with the upper sign in (8.166) and
(8.167) can be discarded. The other solution is marginally stable if
|kαΩ0 /2|1/2 = ηt k 2 , (8.168)
and grows exponentially for larger |kαΩ0 /2|. As in the more general case
treated above, condition (8.168) states that the product of α and Ω0 must
exceed a critical magnitude, i.e., that the dynamo number must exceed a
critical value.
For the marginal (or unstable) solution the frequency of oscillation is
Re(ω) = −|kαΩ0 /2|1/2 , (8.169)
Fig. 8.46. Oscillatory kinematic αΩ dynamo. The meridional cross sections show
contours of constant toroidal field strength on the left, and poloidal lines of force
on the right. Arrows indicate strength and sign of the polar field, the time scale is
adjusted to 11 years for each half-cycle. From Stix (1976 b)
Problem 8.22. Discuss the phase of the poloidal relative to the toroidal field
in the two cases α > 0, Ω0 < 0 and α < 0, Ω0 > 0. Derive the phase re-
lation for the solar field from Figs. 8.41 and 8.42 (Stix 1976 a; Yoshimura
1976).
two parities the latter corresponds to the main characteristics of the observed
mean field. It obeys
For αnorth ∂Ω/∂r < 0 this antisymmetric field is the preferred solution of the
numerical models, in the sense that it becomes marginal at a smaller critical
dynamo number than the symmetric field. This lends further support to the
above choice of α(r, θ) and Ω(r, θ).
angular velocity Ω(r, θ) derived from global convection models and the angu-
lar velocity of the Sun, as inferred by helioseismology, but also to differences
of the helicity between the earlier models and the Sun.
The continuous loss of unstable and buoyant magnetic flux from a dy-
namo within the convection zone has led to the idea that the dynamo mainly
operates in a layer of overshooting turbulence at the base of the convection
zone, where the situation is more stable (Sect. 8.2.4). But the details of such
a deep-seated dynamo are unknown, although a number of models have been
offered.
Dynamic Systems. One of the important dynamic aspects of solar mag-
netism is the instability and buoyancy of horizontal flux tubes in the convec-
tion zone. As the flux tubes drift upwards they acquire a tilt that is observed
in the emergent bipolar regions; this tilt yields a contribution to the mean-
field equation that is equivalent to the α effect. Adding a non-linear term
to equation (8.155), Leighton (1969) was able to construct a dynamo that
works at finite field amplitude. The resulting field resembles the solar field,
but, since α and Ω are freely given, the model is essentially still kinematic.
The same is true for models that simulate the dynamic effects by a quenching
or cutoff at large B of the α effect and (or) the shear, but otherwise use
the equations of the αΩ dynamo (Stix 1972; Yoshimura 1975 a).
As a dynamic system the magnetohydrodynamic dynamo is capable of
chaotic behavior. Such can be seen from the numerical integrations mentioned
above, but still more clearly by considering the following system of coupled
ordinary differential equations (Weiss et al. 1984):
Ȧ = 2DB − A , (8.174)
1
Ḃ = iA − iΩA∗ − B , (8.175)
2
Ω̇ = −iAB − νΩ . (8.176)
The first two of these equations are scaled versions of (8.162) and (8.163);
D is the dynamo number. The spatial dependence has been eliminated by a
Fourier expansion, of which only the leading terms are retained; A∗ is the
complex conjugate of A. Equation (8.176) is the φ component of the equation
of motion, truncated in an analogous manner, and ν = νt /ηt . This equation
describes the effect of the Lorentz force, here −iAB, upon the angular veloc-
ity, Ω(t).
System (8.174)–(8.176) is a complex generalization of a system of three
ordinary differential equations first studied by Lorenz (1963) as a model of
turbulent convection. Like the Lorenz system it has chaotic solutions but,
because it is complex, it also has solutions that are periodic in time. We may
identify the latter with the dynamo wave of Sect. 8.4.3.
The critical dynamo number for the complex Lorenz system is D = 1.
Above D = 1 there exists a periodic solution, for which the period of Ω
380 8. Magnetism
is 1/2 times the period of A and B. In this respect the periodic solution
resembles the cyclic variation of Ω described in Sect. 7.4.2. If D is sufficiently
large and ν < 1, the periodic solution becomes unstable; it is replaced first
by multiply periodic and finally by chaotic solutions. Figure 8.48 (left panel)
shows an example where we find “normal” cycles with variable amplitude, but
also periods of low activity, much as the Maunder Minimum in the sunspot
record of the 17th century (Fig. 8.49, lower panel).
During the period from about 1645 to 1715 very few spots were seen on
the Sun. Eddy (1976) has named this prolonged sunspot minimum after E. W.
Maunder who had considered the subject in the 1890’s. Recent studies (Ribes
and Nesme-Ribes 1993, Hoyt and Schatten 1996) confirm the nearly complete
absence of sunspots. On the other hand, records of aurorae (Schröder 1988)
and the concentration of the 10 Be isotope in ice cores (Beer et al. 1998) indi-
cate that the activity cycle actually might have persisted with low amplitude
during the whole period, although the cycle maxima and minima deduced
from these two sources differ greatly.
The complex Lorenz system has been generalized by including the spatial
variation in two dimensions, depth and latitude (Tobias 1997, Knobloch et
al. 1998). Again an irregular modulation of the field amplitude is obtained;
in addition, the butterfly diagram appears as a characteristic of the αΩ dy-
namo, as shown in Fig. 8.48 (right panel). As a further interesting feature the
model exhibits the non-linear interaction of the magnetic field modes with
different parities with respect to the equator. Such interaction is absent in
linear models with symmetric Ω and antisymmetric α. A close inspection of
the butterfly diagram of Fig. 8.48 reveals a small contribution of even parity.
Indeed, the solar mean magnetic field does not strictly follow the antisymmet-
ric parity defined by (8.172) and (8.173), and the few spots observed during
the Maunder Minimum predominantly occured on the southern hemisphere!
Problem 8.23. Find the analytic periodic solution of the complex Lorenz
system (8.174)–(8.176).
8.4 The Solar Cycle 381
Fig. 8.49. Upper panel : butterfly diagram derived from an αΩ dynamo with a
stochastic contribution to the α effect; time in diffusion time units. After Ossendri-
jver (2000). Lower panel : sunspot relative number, adapted from Eddy (1977)
where A is the vector potential of the magnetic field B. From the induction
equation (8.6) and from its integrated form
∂A
= v × B − η curl B − grad ψ (8.179)
∂t
(where ψ is a scalar function) we derive
dHm ∂B ∂A
= A· +B· dV
dt ∂t ∂t
V (8.180)
= −2 η B · curl B dV + surface terms .
V
Problem 8.24. Calculate the surface terms that appear in Eq. (8.180). Under
which boundary conditions do these terms vanish?
static models of flux tubes, while Parker (1979 c) and Choudhuri (1986) dis-
cuss details of the cluster model. Moreno-Insertis and Spruit (1989) investi-
gate the stability of sunspots to convective motion; the interchange instability
of flux tubes is examined by Bünte (1993) and Bünte et al. (1993). Oscilla-
tions in sunspots have been reviewed by Staude (1999) and Bogdan (2000).
Montesinos and Thomas (1997, with further references therein) have elab-
orated the siphon-flow model of penumbral grains, and Degenhardt et al.
(1993) find some evidence for the shocks that occur in the flow according to
that model.
Solar variability and its influence on Earth has been reviewed by Lean
(1997). Fligge et al. (2000) relate the irradiance variation to the variable
magnetic features on the Sun’s surface. As for the luminosity dips caused
by sunspots, it appears that they are caused predominantly by young spots,
during the phase of rapid development (Pap 1985). Section 1.6 gives further
references on this subject.
The solar cycle and the solar dynamo are subjects of the proceedings
edited by Bumba and Kleczek (1976), Harvey (1992 a), Krause et al. (1993),
Proctor et al. (1993), and Núñez and Ferriz-Mas (1999), of the reviews of
Cowling (1981), Rädler (1990), or Stix (1981, 1987 b, 1991, 2001). Mean-
field electrodynamics is the subject of Roberts and Soward (1975 a,b), and
Knobloch (1977, 1978). The layer of overshooting turbulence at the base of
the convection zone has been treated by van Ballegooijen (1982), Schmitt
et al. (1984), and Pidatella and Stix (1986). Schüssler (1983, 1987 a) gives
arguments why the dynamo should operate in this layer.
Since helioseismology has revealed the variation of Ω near the base of the
convection zone, αΩ dynamo models employ this tachocline. With an analyt-
ical model Parker (1993) illustrates that the dynamo takes on the character
of a surface wave; Charbonneau and McGregor (1997) extend this interpre-
tation to the spherical geometry. Dikpati and Charbonneau (1999) apply the
observed rotation profile Ω(r, θ) to the earlier model of Leighton (1969), and
Belvedere et al. (2000) consider the asymptotic case of large dynamo number
(Ruzmaikin et al. 1988) in the light of the helioseismic result. It is another
problem whether the toroidal field generated by the differential rotation has
the form of isolated flux tubes, or a more even distribution. Rempel et al.
(2000) and Wissink et al. (2000 a) have considered the stability of a magnetic
layer and the breakup of such a layer into individual tubes.
There are further observations that may provide keys for a better under-
standing of the Sun’s dynamo. One is the variation of sunspot darkness over
the cycle (Albregtsen and Maltby 1978). A second is the apparent change of
the number density of granules with the solar cycle (Macris and Rösch 1983)
and, probably related, a variation of the atmospheric temperature gradient
(Holweger et al. 1983). A third is the meridional circulation (Sect. 7.4.3),
which may severely alter the propagation of the dynamo wave, as a calcula-
tion of Choudhuri et al. (1995) shows. Finally, the p-mode eigenfrequencies
8.5 Bibliographical Notes 385
vary with the phase of the solar cycle, as first noticed by Woodard and Noyes
(1985) and Fossat et al. (1987), and confirmed later (e.g., Howe et al. 1999);
at 3 mHz, for example, the frequencies increase by ≈ 0.4 μHz from activity
minimum to maximum. Dziembowski et al. (2000) discuss possible reasons for
this variation, one being changes in the structure of the turbulent convection
and the ensuing variation of the convection-oscillation interaction (Zhugzhda
and Stix (1994).
Apart from questions concerned with the 11-year solar cycle, there are
other, more fundamental, problems in dynamo theory. One is related to the
origin of the small-scale magnetic flux that is found outside active regions,
notably in the ephemeral active regions (Martin and Harvey 1979). It was
suggested earlier that this kind of magnetism is “waste” from the global dy-
namo (Golub et al. 1981). More recently, numerical simulations led Cattaneo
(1999; see also Cattaneo and Hughes 2001) to advance the hypothesis of lo-
cal field amplification by turbulent convection. Other questions arise in the
context of magnetic helicity conservation. Brandenburg et al. (2002) review
this subject, Berger and Ruzmaikin (2000) especially discuss the solar case.
9. Chromosphere, Corona, and Solar Wind
In the preceding chapters we have learned about convection and rotation, the
main causes of solar magnetism. And we have learned about this magnetism
itself. We shall now recognize that the Sun’s magnetic field is the main source
of almost all structure and variability which we find in the outermost layers:
the chromosphere, the corona, and the solar wind. Moreover, we shall see
that the physical state and the extent of these outer layers are most certainly
determined by the solar magnetic field.
Thus, inhomogeneity and time-dependence are characteristics of the top-
ics to be treated in this chapter. Nevertheless, we shall often retreat to the
conveniences of spherically symmetric and stationary models. Although such
models are idealized and far from reality, they may serve to illustrate basic
concepts such as the expanding hot corona, or may yield order-of-magnitude
estimates which otherwise would be difficult to obtain.
To begin with, we shall review the most important observations.
For a few seconds, just after the beginning and before the end of a total
eclipse, the solar limb presents a most colorful view (chromosphere: the “col-
ored sphere”). The spectrograph reveals the flash spectrum which shows –
in emission – a large number of the lines that are dark in the normal so-
lar spectrum, in particular the strongest ones. A prominent example is the
red Hα line at 656.3 nm. These lines are seen in emission because at their
frequencies the solar atmosphere is still opaque at a level where, even for a
tangential line of sight, it is transparent in the continuum (the visible limb
lies at τ500 ≈ 0.004, cf. Sect. 1.3). That is, in the lines we see the hot solar
atmosphere, while in the continuum we look into cool and dark space.
Mottles and Spicules. The large optical thickness of the chromospheric
lines also allows us to observe the chromosphere on the disc, by means of a
narrow-band filter. The Hα filtergram of Fig. 6.22 is an example. Another is
Fig. 9.1, taken in the wing of the Hα line profile. Even more than the limb
Fig. 9.2. Solar spectrum of the Ca II K line region. The spectral band is from
393.17 nm to 393.56 nm, the vertical extent is 160 Mm on the Sun. Photograph
Schauinsland Observatory, Kiepenheuer-Institut, Freiburg (Grossmann-Doerth et
al. 1974)
390 9. Chromosphere, Corona, and Solar Wind
In fact it is much stronger in the LTE models than in the models based on
statistical equilibrium only.
Finally, as we tune the filter towards the line core, we arrive so high in
the atmosphere that the assumption of LTE completely breaks down. The
source function is smaller than the Planck function; as a result we have the
intensity minimum at the line center.
The intensity minima in the red and violet wings of the K line are usually
designated by K1r and K1v , the emission peaks by K2r and K2v , and the
central minimum by K3 .
It is evident from the spectrum shown in Fig. 9.2 that the Ca II emission
strongly varies with position on the solar surface. Filtergrams, or spectrohe-
liograms such as the one shown in Fig. 9.4, reveal that there are mainly two
sources of Ca II emission: larger areas, the plages, which spatially coincide
with the active regions, and the finer chromospheric network, which we have
already identified with the cell boundaries of the supergranular velocity field.
The close relationship between plages and active regions, and between
the network and the supergranulation (which, as we have seen, accumulates
magnetic flux, cf. Fig. 8.11), suggests that the magnetic field plays a key role
for the chromospheric emission. The chromospheric heating must be greatly
enhanced in the areas where the magnetic field strength is large. We postpone
the discussion of chromospheric heating but note that a study of Skumanich
et al. (1975) shows that the total Ca II K emission in the network is roughly
proportional to the local strength of the magnetic field.
The connection between Ca II emission and magnetism allows us to study
magnetic activity of stars other than the Sun. Indeed, from measurements
initiated in 1966 by O. C. Wilson (Wilson 1978), and extended by Baliunas
et al. (1995), it is possible to identify activity cycles for about 50 late-type
stars. Moreover, as the sources of Ca II emission are unevenly distributed on
the stellar surface, it is also possible to measure the rotational modulation
of the emission. Stellar rotation that is too slow to be visible in the typical
rotational line-broadening could still be detected in this way (Vaughan et
al. 1981). Rapidly rotating young stars usually have high average levels of
calcium emission, but no smooth cyclic variation; slow rotators like the Sun
9.1 Empirical Facts 391
clearly exhibit cycles, while their average emission is lower. When the Sun is
observed as a star, i.e., in the light integrated over the disc, the cyclic as well
as the rotational variation of the Ca II emission is clearly discernable.
Because of the importance of the outward temperature increase, especially
for the Ca II emission, a more precise definition of the chromosphere is used
frequently: the layer between the temperature minimum and the level where
T = 25 000 K. In the one-dimensional mean model, e.g., Fig. 4.9, this layer
comprises some 2000 km; Fontenla et al. (1990, 1993) give tables of such
models. On the other hand the spicules, which also have a chromospheric
temperature, cover a range of ≈ 5000 km when observed at the limb. This
apparent discrepancy once more illustrates how much care must be taken
with the application of a mean model!
392 9. Chromosphere, Corona, and Solar Wind
Fig. 9.5. Solar spectrum in the extreme UV, obtained from a quiet-Sun region with
the spectrograph SUMER on the SOHO satellite. Prominent lines, and continua
with black-body fits, are identified. The relative uncertainty of the spectral flux is
≈ 0.2. Courtesy K. Wilhelm
The “layer” between the relatively cool (≈ 104 K) chromosphere and the hot
(≈ 106 K) corona is called the transition region. It is appropriate to think of
a temperature regime rather than a geometric layer. This is not only because
of the extreme spatial inhomogeneity of this region, but also because the
transition is so sharp that there is virtually a discontinuity in T (and hence
in ρ, because the pressure, P ∝ ρT , must remain continuous).
The ultraviolet spectrum from 50 nm to 160 nm is most appropriate to
investigate the transition region. It contains numerous emission lines, and
emission continua, originating from ions existing at various temperatures in
the above-mentioned range (Fig. 9.6). Observation from space is necessary to
acquire an ultraviolet spectrum. Examples of successful instruments are the
Harvard College Observatory spectrometer on Skylab (Reeves et al. 1977),
and the High Resolution Telescope and Spectrograph (HRTS) of the Naval
Research Laboratory (Brueckner 1981) flown on rockets and on Spacelab 2.
9.1 Empirical Facts 393
Fig. 9.7. Lyman α filtergram (Bonnet et al. 1980), showing the structure of the
transition region with a resolution of 1 . This photograph (exposure 1 s) was ob-
tained during a rocket flight by the Laboratoire de Physique Stellaire et Planétaire,
Paris, in cooperation with Lockheed Corporation and NASA
394 9. Chromosphere, Corona, and Solar Wind
alistic description. Again we know from limb observations that the sources
of transition-zone radiation are distributed over several thousand km. Pic-
tures such as the Lyman α filtergram shown in Fig. 9.7 directly demonstrate
the substantial inhomogeneity. Moreover, there are material flows (Doppler
shifts), and there is variation in time.
The spatial inhomogeneity becomes less pronounced in the higher-tempera-
ture part of the transition-zone radiation. In particular, the chromospheric
network, which is clearly visible in the lines of the lower transition zone, grad-
ually disappears as we go towards the corona. The common interpretation of
this effect is that the governing agent, namely the magnetic field, becomes
more homogeneous as it fans out from the photospheric flux tubes into the
corona; as the coronal gas pressure is small the confinement into such narrow
tubes is no longer possible.
In the upper part of the transition region the mean temperature profile flat-
tens (Fig. 9.6), and we come to an extended region with T ≈ 106 K, the
solar corona. Strictly we should now distinguish between the ion and elec-
tron temperatures, Tion and Te , but presently we discuss the mere fact that
the coronal temperature is so much higher than the temperature of the pho-
tosphere, which was recognized around 1940, earlier than the corresponding
information about the chromosphere and transition region was available. Still,
the discovery came rather late in the history of solar astronomy, perhaps be-
cause it is in contrast to the simple picture of a hot star in a cool interstellar
environment. A hot corona requires energy to be pumped from low to high
temperature. We shall see below that even today the problem of coronal
heating is far from a satisfactory solution.
White-Light Corona. The optical coronal radiation, as observed during
total eclipses, is traditionally divided into the K corona and the F corona. The
F corona dominates outwards from, say, 2 or 3 solar radii. Its spectrum shows
the normal dark Fraunhofer lines of the photospheric spectrum, and its name
is derived from these lines. The continuum of the outer corona also resembles
the photospheric spectrum, and is only weakly polarized. Hence the F corona
can be explained as photospheric light scattered on dust particles. As zodiacal
light it can be observed far into interplanetary space. In the following we shall
not be concerned further with this type of coronal radiation.
The K corona derives its name from the German “Kontinuum”. Its contin-
uous spectrum also resembles the photospheric spectrum, but the Fraunhofer
lines are absent (Problem 9.2 will make clear why). The light of the K corona
is highly polarized, which indicates that it arises from Thomson scattering
by free electrons.
The intensity of the white-light corona drops steeply with distance from
the Sun, and strongly depends on the position angle. As an average over
9.1 Empirical Facts 395
position angle Baumbach (1937), using data from 10 eclipses between 1905
and 1929, derived the formula
I 0.0532 1.425 2.565
= 10−6 + + , (9.1)
IC x2.5 x7 x17
where IC is the intensity at disc center, and x is the normalized radius
r/r , projected onto the plane of the sky. Of the three terms in (9.1) the first
is mainly due to the more slowly decaying F corona, while the second and
third essentially describe the rapidly decaying K corona. Even during good
(“coronal”) eclipse conditions the sky has a radiance of ≈ 10−9 IC ; thus,
according to (9.1), the mean corona becomes invisible at x ≈ 4, the K corona
even at smaller x.
From the observed intensity an estimate of the density of scattering
electrons can be obtained under the assumption of a spherically symmet-
ric corona. Let EK be the total emission, per volume and solid angle, of
scattered radiation, and let y be the coordinate along the line of sight, which
is perpendicular to x. Then
ρ2 ≡ (r/r )2 = x2 + y 2 , (9.2)
and the observed intensity is
∞ ∞ ∞
ρ E (ρ)
I(x) = EK (ρ) dy = 2 EK (ρ) dy = 2
K dρ . (9.3)
ρ2 − x2
−∞ 0 x
Since EK is unknown one must invert (9.3), i.e., calculate the inverse Abel
transform
∞
1 dI/dx
EK (ρ) = −
dx . (9.4)
π x2 − ρ2
ρ
Now compare this result to the total expected scattered light, that is, set
1
EK (ρ) = σs ne I (θ) dΩ . (9.5)
4π
The integral is over the angle subtended by the solar disc as seen from the
distance ρ, I (θ) is the intensity of the Sun’s radiation, ne the electron den-
sity, and σs the cross section of Thomson scattering, defined in (2.90). Finally,
we solve (9.5) for ne . Notice that, in addition to the assumption of spherical
symmetry, it is now assumed that the scattering is isotropic. This assumption
is false but may suffice for an order-of-magnitude estimate.
Allen (1947) used the method outlined here to derive the electron density
of the mean K corona. His result is approximated by
1.55 2.99
ne (ρ) = ne0 + , (9.6)
ρ6 ρ16
396 9. Chromosphere, Corona, and Solar Wind
x Theoretical Observed
where ne0 = 1014 m−3 . The two contributions correspond to the last two
terms of (9.1).
The correct treatment of the scattering problem must of course include
the anisotropy, and will predict the polarization. If the scattering angle was
exactly 90o , we would have a polarization of 100%, a consequence of the fact
that both the incident and the scattered waves are transverse. The direction
of the polarization would be tangential to the Sun (z, say). For two reasons
the scattering angle is not 90o : first, because the observer integrates over the
line of sight; and second, because for any point within the corona the solar
disc subtends a finite area on the sky.
If these geometrical effects are taken into account, and the observed in-
tensity distribution is used, one obtains the theoretical results listed in Table
9.1. The degree of polarization given there is
Iz − Ix
P = , (9.7)
Iz + Ix
where Iz and Ix are the intensities in the two perpendicular directions of
polarization. The observed polarization is listed for comparison. There is
close agreement if the F corona is eliminated.
Scattering on free electrons is independent of wavelength. However, be-
cause the photospheric limb darkening is more pronounced at shorter wave-
lengths, the Sun looks slightly more “point-like”. The polarization of the
white-light corona is therefore slightly larger at smaller λ, in particular close
to the Sun, at small x.
Problem 9.3. Use the electron density (9.6), assume a barometric stratifi-
cation, and derive the coronal electron temperature.
Like the chromosphere, the solar corona is closely related to the Sun’s
magnetism. A particular aspect of this relationship has been known for a long
time: the variation in shape from activity minimum to activity maximum.
Fig. 9.8. The solar corona at activity minimum (upper) and maximum (lower
image), photographed with a radially graded filter (transmission range 104 ) during
the eclipses of 3 November 1994 in Putre, Chile, and 16 February 1980 in Palem,
India. The vertical is heliographic north in both images. Courtesy A. Lecinski, High
Altitude Observatory, NCAR, Boulder, Colorado
398 9. Chromosphere, Corona, and Solar Wind
Table 9.2. Three coronal emission lines. From Billings (1966), with identi-
fications of Edlén (1942); χ is the ionization potential of the preceding ion
the temperature turns out too low, because then fewer electron collisions
would suffice to maintain the ionization equilibrium).
Similar to the ionization equilibrium is the equilibrium that governs the
line emission: excitation by collisions with electrons, and spontaneous emis-
sion of radiation. The lines are “forbidden” in the sense that, at normal
densities, the probability of radiative transition would be much smaller than
the probability of collisional de-excitation. In the tenuous corona collisions
are rare, and the opposite is true. Formally, the selection rule ΔL = ±1 for
the total orbital angular momentum is violated for most of the transitions in
question, cf. the examples of Table 9.2.
Because the coronal emission lines are so intense the combination of a
coronagraph with a narrow-band filter is especially rewarding. Both tools
have been developed by B. Lyot in the 1930s, so that coronal observations
could be made on a routine basis independently of eclipses. Extensive ob-
servations of the emission-line corona (or E corona), have been made by
Waldmeier (1951). His studies not only covered an entire cycle of solar activ-
ity, but also for the first time showed the existence of the dark coronal holes
which are so important in modern coronal research (Waldmeier 1957, p. 275).
Problem 9.4. Estimate the coronal ion temperature from the width (≈ 1 Å)
of the coronal emission lines.
The Radio Corona. We have already seen, cf. Fig. 1.6, that the quiet
Sun possesses a radio spectrum of thermal nature, and that the brightness
temperature changes from ≈ 104 K to ≈ 106 K in the wavelength range 1 cm
to 1 m. Clearly, we may now attribute this transition to the temperature
rise from the temperature minimum to the corona. The coronal thermal con-
tinuum mainly originates from free-free transitions.
With increasing wavelength the radio spectrum becomes dominated by
the various types of radio bursts. We postpone this subject to Sect. 9.5.1, but
note already here that the radio bursts can be used as a probe of the corona.
The reason is that in a plasma such as the corona, electromagnetic waves
cannot propagate if their frequency lies below the plasma frequency
1/2
e ne
νP = . (9.8)
2π ε0 me
Waves with lower frequency will be absorbed or reflected: on Earth, for ex-
ample, we can receive distant radio stations because reflection occurs in the
ionosphere. The cutoff has its complete analogue in the cutoff for acous-
tic wave propagation in a stratified atmosphere, which we have treated in
Chap. 5. In the corona, the electron density decreases with distance from the
Sun. Therefore, radio signals with longer wavelength must arise from sources
which lie further outwards than the sources of short-wavelength signals.
400 9. Chromosphere, Corona, and Solar Wind
Problem 9.5. Consider oscillations of free electrons relative to the ions (at
rest) in a one-dimensional model. Use the equation of continuity, the mo-
mentum balance, and Maxwell’s equations to show that such plasma os-
cillations occur with the frequency (9.8).
Problem 9.6. The radio heliograph at Nançay, France, consists of a cross-
shaped array of antennas, with branches of 3200 m and 1250 m in the
North-South and East-West directions, respectively. Observations are
made at wavelengths between 60 cm and 2 m. Use (9.6) and (9.8) to cal-
culate the height of the coronal sources. What is the angular resolution
for the two directions?
Ultraviolet and X-rays. The spectrum shown in Fig. 9.5 demonstrates the
dominance of emission lines in the extreme ultraviolet region. This continues
toward shorter wavelengths. Spectral atlasses have been obtained from the
SOHO instruments CDS and SUMER and cover the ranges 30–60 nm and
67–161 nm, respectively (Brekke et al. 2000, Curdt et al. 2001); an earlier
table compiled by Billings (1966) contains more than 200 identified emission
lines between 1.37 nm and 105.87 nm, originating from ions such as N VII,
O VIII, or Fe XVII. Lines from even higher states of ionization, e.g., from
Fe XVIII to Fe XXIII, are seen during flares.
Profiles of EUV emission lines far above the limb have been measured with
the UV Coronagraph Spectrometer (UVCS) on board of SOHO (Kohl et al.
1997, 1998). Large line widths have been found especially for emission from
coronal features having a magnetic field perpendicular to the line of sight, e.g.,
streamers, helmets, or coronal-hole regions above the limb. The interpretation
in terms of a thermal velocity distribution yields values of the transverse
kinetic temperature, T⊥ , of up to 2 × 108 K for the diverse ions at a distance
of about 4r . In addition to the line width, the thermal velocity distribution
parallel to the magnetic field and the velocity of radial outflow from the
Sun have been determined by a measurement of the Doppler dimming effect:
the intensity of resonantly scattered light is diminished if the velocity of the
scattering medium relative to the source of the incoming radiation is of the
order of (or larger than) the thermal velocity of the scattering particles. Thus,
with increasing speed of the solar wind in the corona, the resonance frequency
becomes Doppler shifted away from the frequency of the incoming photons,
and the part of the line emission that is due to resonant scattering becomes
weaker. Using a given spherically symmetric density model of the corona,
and applying the technique to lines of diverse ions, Cranmer et al. (1999 b)
and Cranmer (2000) were able to derive parallel kinetic ion temperatures T
and the outflow velocity. As a result they find that the wind speed reaches
200–300 km/s as close as one solar radius from the solar surface, and that the
ion temperatures are highly anisotropic, with T⊥ T . The latter result will
be discussed below (Sect. 9.4.2) in the context of the heating problem.
9.1 Empirical Facts 401
Figure 9.9 is a TRACE image that shows plasma loops in a spectral band
of width ≈ 0.6 nm around 17.1 nm, containing emission lines of Fe IX and
Fe X. The finest discernable threads have a width of ≈ 106 m, which is close
to the resolution limit; nevertheless it is remarkable that this width does not
change much along the loops. Moreover, often these loops seem to be nearly
isothermal, in the case of Fig. 9.9 at ≈ 106 K, and for large loops the density
and the pressure appear to decrease with a scale height that is several times
the scale height derived from the temperature (Lenz et al. 1999, Aschwanden
et al. 2000). This indicates that the loops are not in hydrostatic equilibrium.
In addition to the line spectrum, the soft x-ray region (below 10 nm, say)
is characterized by a continuum of thermal form, arising from free-bound and
free-free (bremsstrahlung) transitions. Figure 1.8 gives the intensity and its
variation. Of course, the thermal x-ray radiation can be used to assess once
more the high coronal temperature.
Since the cool photosphere has a small intensity at wavelengths below
150 nm, an x-ray telescope sees the corona with all its rich structure in front
of the solar disc. Numerous coronal x-ray images were taken with the Wolter
telescope flown on Skylab in 1973–1974, with the normal-incidence x-ray
telescope NIXT on rocket flights (Golub et al. 1990), and with the soft x-ray
telescope of the Yohkoh mission (a modified Wolter telescope). Figure 9.10
shows examples; another example was already presented in Fig. 7.7.
402 9. Chromosphere, Corona, and Solar Wind
Fig. 9.10. Images of the Sun in soft x-rays (0.3–6 nm), obtained in 1973 by the
Skylab mission (left, Vaiana et al. 1973), and in 1991 by Yohkoh (right)
The first evidence for a continuous particle stream emanating from the Sun
came from the ion tails of comets. These tails roughly point into the direc-
tion opposite to the Sun, but Hoffmeister (1943) found that there is a small
systematic deviation: the ion tail slightly trails (in the sense of the comet’s
motion), so that there is a small angle, normally < 5o , between it and the
solar radius vector. Biermann (1951 b), realizing that the radiation pressure
404 9. Chromosphere, Corona, and Solar Wind
of the sunlight could not account for the acceleration observed in the ion
tails, proposed a corpuscular radiation instead, and so was able to explain
Hoffmeister’s discovery: the deviation angle of the tail is given by the ratio
of the comet’s transverse orbital velocity component to the velocity of the
corpuscular radiation. From the distribution of comets it could be inferred
that the postulated corpuscular radiation, later termed the solar wind, would
blow continuously, and into all directions. What is more, the large variations
in space and time of the phenomenon could already be seen by means of its
variable effects upon the comets.
Problem 9.7. A comet’s ion tail deviates by 4o from the radial direction.
Its transverse orbital velocity is 30 km/s. What is the speed of the solar
wind? Convince yourself that this wind speed is supersonic.
In order to measure the solar wind in situ, a spacecraft must fly above the
Earth’s magnetosphere, i.e., in an orbit several thousand km above ground, or
in an interplanetary orbit. The data collected in 1962 by Mariner 2 on its way
to Venus fully confirmed the earlier predictions: a high-speed (supersonic),
continuous, but rather variable, flow of ionized matter. Electrons, protons,
and α particles (3–4 percent, relative to the protons) are the predominant
constituents.
As Fig. 9.11 illustrates, the solar wind velocity v often remains high
(≈ 700 km/s) during a few consecutive days; the particle density np varies in
anti-phase to v. The actual numbers show that the particle flux density np v
of the slow wind is 2–3 times that of the fast wind. It has been shown that
the high-speed streams have their sources in those parts of the solar corona
where the magnetic lines of force are open, i.e., essentially in the coronal
holes. Figure 9.11 also shows the tendency of the high-speed streams to recur
Fig. 9.11. Three-hour averages of solar wind velocity (thick) and proton density
(thin), as observed by Mariner 2. Adapted from Hundhausen (1972)
9.1 Empirical Facts 405
Fig. 9.12. The source region of the fast solar wind. This SOHO/EIT image in the
19.5-nm pass band shows several coronal holes. The inset is a SUMER velocity map
obtained from the Doppler shift of the Ne VIII line at 77.04 nm, with outflow ve-
locities (blue) of 5–20 km/s. The contours of the chromospheric network are derived
from the Si II emission (Hassler et al. 1999)
after 27 days, which is the synodic equatorial period of rotation of the large-
scale solar magnetic field. The sources of the solar wind coincide with the M
regions postulated earlier as the sources on the Sun of recurrent geomagnetic
perturbations.
With the spectrograph SUMER on board of the SOHO mission it has
been possible to identify the source region of the solar wind in more detail
(Fig. 9.12). The Ne VIII line at 77.04 nm, which is formed at T ≈ 600 000 K
in the transition region, shows a Doppler shift corresponding to an outflow
velocity of up to 20 km/s. The measurement has been made in the second
spectral order, either with respect to an unshifted chromospheric line (Hassler
et al. 1999) or with respect to the line position at the solar limb where no
Doppler shift is expected because the outflow is transverse (Peter 1999).
406 9. Chromosphere, Corona, and Solar Wind
As the inset of the figure illustrates, the outflow clearly originates in the
coronal holes. Moreover, Hassler et al. were able to prove a relationship to the
chromospheric network; the largest outflow velocities occur at the junctions
of the network cell boundaries. Earlier evidence for an outflow from coronal
holes, but with less spatial resolution, had already come from the Doppler
shift of Si XI, O V, and Mg X lines measured during rocket flights (Cushman
and Rense 1976, Rottman et al. 1982). A coronal outflow was also inferred
from the Doppler shift of radar echos from the Sun (Abel et al. 1963).
The solar wind has been probed in situ by numerous space experiments.
The distinction between the low-speed and high-speed forms of the solar
wind has become clear through many space experiments, especially the two
Helios missions launched in 1974 and 1976. Figure 9.13 illustrates the more
recent measurements made by the Ulysses spacecraft during solar minimum
conditions: at heliographic latitudes higher than ±20o the wind velocity was
≈ 750 km/s, with only small fluctuations; this characterizes the high-speed
wind that originates in the coronal holes at the northern and southern poles
of the Sun. In contrast to this, the wind is generally slower but much more
variable in the low-latitude range. Most conspicuous is the variation caused
by the rotating sector structure of the wind, with the recurrent streams of
higher and lower speed. Within a fixed latitude interval, Ulysses encountered
a larger number of such streams when the spacecraft was near its aphelion,
Fig. 9.13. Solar wind velocity versus heliographic latitude, measured by Ulysses.
Left: south-bound sections of the orbit, at distances 2–5 astronomical units from
the Sun; right: north-bound section, including the fast passage of the perihelion
near 1 astronomical unit. Courtesy S. T. Suess
9.1 Empirical Facts 407
and therefore slow, while a smaller number was seen during the fast passage
of the perihelion (Fig. 9.13, left and right, respectively).
A relatively close approach to the Sun was reached by Helios 1 and He-
lios 2, which had their perihelia at 0.31 and 0.29 astronomical units, respec-
tively. The results listed in Table 9.4 are from these two spacecraft. The
temperatures given there are directly derived from the velocity distributions
of the diverse particles. The temperature anisotropy is defined as T /T⊥ ,
where T and T⊥ characterize the distributions parallel and perpendicular to
the magnetic field. Because collisions are rare (cf. Table 9.3), and because the
magnetic field guides the particles, it is quite natural that such an anisotropy
occurs, and that the protons have a different temperature than the electrons
(in the low-speed solar wind, where v ≈ 300 km/s, the proton temperature is
only about 4 × 104 K at 1 astronomical unit, and has an anisotropy of ≈ 2,
while the electron temperature is still ≈ 1.5 × 105 K). There is little change
of the solar wind velocity between 0.3 and 1 astronomical units.
Fig. 9.14. Element abundances as a function of the first ionization potential, rel-
ative to oxygen and normalized to the photospheric ratio. The dashed line in an
approximation to the slow-wind and flare-particle abundance ratios (except for Kr
and Xe). From Geiss (1998)
Of the results described in the preceding section the most important one
was the high temperature of the outer solar atmosphere. Due to the high
temperature there are a large number of free electrons which, as we have
seen in Sect. 8.1.2, give rise to a large electrical conductivity. In a similar
way the electrons provide a large thermal conductivity.
Below we shall mainly be interested in the conduction of heat in the
transition zone and the corona. Therefore, we may use the result obtained by
Spitzer (1962) for fully ionized gases, namely
√
640 ε20 2π k(kT )5/2
κ = √ εδT . (9.11)
me e4 Z ln Λ
This expression has been derived in analogy to the electrical conductivity
(8.10). As there, the temperature dependence reflects the electron velocity
distribution. The factor ε (< 1) arises because an electric field is produced
by the distorting effect (on the electron velocity distribution) of a temperature
gradient. The effect of this is to reduce the transport of heat by the electrons.
The last factor, δT , of (9.11) corresponds to γE in expression (8.10) and
corrects for the error made by the assumption of a Lorentz gas. For a charge
number Z = 1 Spitzer gives ε = 0.419 and δT = 0.225. If we use ln Λ = 20
for the Coulomb logarithm, as it is appropriate to the corona, and substitute
all the constants in (9.11), we obtain κ = 9.2 × 10−12 T 5/2 WK−1 m−1 .
In the presence of a magnetic field the electrons are guided along the field
lines, and (9.11) applies only to the heat flow parallel to the field; this is
indicated by the subscript. Transverse to the field the conductivity is greatly
reduced (here heat is mainly conducted by protons, which have much larger
gyration radii).
The large difference in parallel and transverse conduction of heat, together
with the phenomenon of the frozen-in field, explains why we can directly see
various details of the coronal structure: both matter and heat are forced to
follow the magnetic field. Hence the polar plumes (Fig. 9.8), for example,
indeed outline the magnetic field.
The Heat-Conduction Corona. We shall treat the problem of coronal
heating below in Sect. 9.4. For the moment, let us simply assume that heat
is deposited at some level r0 in a spherically symmetric corona, and that we
may model this heat input by prescribing the temperature T0 at this level
(e.g., T0 = 106 K). We also assume that there is no other heat transport than
conduction, i.e.,
F = −κ ∇T , (9.12)
and
410 9. Chromosphere, Corona, and Solar Wind
div F = 0 . (9.13)
Substitution of (9.11) immediately yields
d dT
r2 T 5/2 =0. (9.14)
dr dr
We know that in the photosphere (r = r ) as well as far away from the Sun
(r → ∞), the temperature is much smaller than T0 ; hence for the transition
region and the corona the error introduced by setting T (r ) = T (∞) = 0 is
small. The solution of (9.14) then is
2/7
1 − r /r
T = T0 for r ≤ r0 , (9.15)
1 − r /r0
and
−2/7
r
T = T0 for r ≥ r0 , (9.16)
r0
and is depicted in Fig. 9.15 for the case r0 /r = 1.2. This heat-conduction
temperature profile correctly models both the steep gradient in the transition
region and the rather flat decline towards the interplanetary space.
must match the interstellar pressure for r → ∞. Since the interstellar pressure
is many orders of magnitude smaller than P0 , this requirement means that
the integral in (9.18) must diverge. Clearly, this is the case if T (r) declines
more rapidly than 1/r. The heat conduction profile (9.16) does not qualify
under this condition. In other words, the heat-conduction corona cannot be
embedded into interstellar space under pressure equilibrium.
Problem 9.8. Calculate P (∞) for T (r) given by (9.16). Take T0 = 106 K,
P0 = 10−2 Pa, and compare the result to the interstellar pressure at
T = 50 K and ρ = 10−20 kg/m3 (cf. Problem 2.2).
9.2.2 Expansion
The difficulty encountered by the static corona model was removed when
Parker (1958) replaced the hydrostatic equilibrium by a stationary dynamic
equilibrium. In order to demonstrate the essential point let us again consider
the case of spherical symmetry, with a velocity field of the form
v = (v(r), 0, 0) (9.19)
in spherical polar coordinates. Also, as the profile (9.16) is in any case too
flat for hydrostatic equilibrium, we may as well seek a dynamic equilibrium
for the simpler case T = const.
Instead of (9.17) we must now solve the equations of continuity and mo-
mentum balance, and the equation of state, viz.
d
(ρr2 v) = 0 , (9.20)
dr
dv 1 dP Gm
v =− − , (9.21)
dr ρ dr r2
P = ρRT /μ . (9.22)
Elimination of ρ and P yields the differential equation for v,
1 dv 2 2c2 Gm
(v − c2 ) = − , (9.23)
v dr r r2
where c = (RT /μ)1/2 is the isothermal speed of sound. Using a coronal
temperature we readily see that the right-hand side of (9.23) is negative close
to the Sun, and positive at large distance from the Sun. At the distance
Gm
rc = , (9.24)
2c2
the critical radius, there is a change of sign. Of course, the left-hand side of
(9.23) must also reverse its sign at rc ; this leads to the classification of four
distinct types of solution v(r), as illustrated in Fig. 9.16.
412 9. Chromosphere, Corona, and Solar Wind
Two of the solutions start in the corona with supersonic velocity. This is in
contrast to observation, so these two types must be rejected. Of the other two,
one always stays subsonic, and at large r even comes to rest. This also is not
observed; in addition, it can be shown that this type of solution cannot match
the interstellar pressure, just as the hydrostatic solution. What remains is the
singular solution that turns from subsonic to supersonic as it passes through
the critical radius. We must adopt this solution and, following Parker, we
shall call it the solar wind. It is the only solution that satisfies all boundary
conditions, including the requirement of matching the interstellar pressure
(Problem 9.9). The adjustment to the interstellar pressure is accomplished
at a distance of order 100 astronomical units from the Sun in a termination
shock, where the velocity turns subsonic again. Still further away, at the
heliopause, the wind merges into the interstellar medium.
The true solar wind is more complicated than the simple model treated
here. In fact such a model generally fails to predict the exact values of the
speed and mass flux of the wind. One must consider the effects of rotation
and the magnetic field, and the variation with the angular coordinates and
with time. And, instead of the assumption of isothermal expansion, one must
solve a real energy equation, including heating by waves and energy loss by
heat conduction back to the chromosphere. However, before we touch some
of these complications let us return to the more general question of the deter-
mination and the effects of the magnetic field in the Sun’s outer atmosphere.
Problem 9.9. For T = 106 K calculate the critical radius and the speed of
sound. Integrate (9.23) and discuss the behavior of v(r) and P (r) for the
various solutions at large distance from the Sun. Estimate v(r) at 1 as-
tronomical unit for the solar wind solution.
9.3 The Magnetic Field in the Outer Atmosphere 413
In principle the magnetic field could be measured by means of the Zeeman ef-
fect. However, the difficulty which already exists in the photosphere, namely
that the Zeeman splitting is small in comparison to the line width, becomes
almost unsurmountable in the outer atmosphere where the temperature is
higher (and the lines broader) and the field is weaker (i.e., the splitting
smaller). But measurements have been made in the relatively cool promi-
nences. The circular polarization in quiescent prominences indicates a longi-
tudinal field component of order 1–10 mT, while active prominences generally
have a field strength above 10 mT. In the corona itself, Lin et al. (2000) were
able to measure the circular polarization of emission lines. From the Stokes
parameter V (λ) of the Fe XIII near-infrared line at 1074.7 nm they could infer
a field strength of 1–3 mT at a height of ≈ 108 m above active regions.
A second possibility, also mainly applied to prominences, is the Hanle
effect. The Hanle effect provides information on the transverse component
of the magnetic field and thus complements the longitudinal Zeeman mea-
surement. The field strength obtained in prominences is again of order 1 to
10 mT. It is important to notice that the Hanle effect yields magnitude and
direction, but not the sign of the transverse field component. This ambiguity,
which the Hanle effect has in common with the transverse Zeeman effect [cf.
the dependence on γ and φ of the absorption matrix (3.69)], has interesting
consequences for the prominence models which will be discussed below.
Some information about the coronal magnetic field is obtained from ra-
dio observations. Since the electrons emitting gyromagnetic radiation spiral
around the lines of force, observation of moving radio sources directly yields
the geometry of the guiding field. The strength and direction of the field can
also be inferred by measuring the intensity and polarization of the radiation.
Unfortunately there are severe limits on spatial resolution at radio frequen-
cies. Above active regions Dulk and McLean (1978) find values of order 10 mT
at the base of the corona; at r/r = 2 a typical field strength is 0.1 mT.
The most accurate magnetic field measurements have been made in situ.
The Helios results have already been included in Table 9.4 above. The num-
bers given there apply to high-speed streams of the solar wind in or near the
ecliptic plane; outside these streams the inclination is somewhat larger, so
that 45o is an average value at 1 astronomical unit. Figure 9.17 illustrates
the field measurements made earlier by IMP-1 (“interplanetary monitoring
platform”), and their extrapolation back to the solar surface. The spiral pat-
tern results from the fact that the field is frozen into the solar wind (Sect.
8.1.3), in combination with the effect of solar rotation. Since the wind speed
does not change much over most of the distance Sun–Earth (cf. Table 9.4, or
Problem 9.9), the spiral is nearly Archimedian.
414 9. Chromosphere, Corona, and Solar Wind
The sector boundaries shown in Fig. 9.17 are the intersections of a warped
heliospheric current sheet with the ecliptic plane. This current sheet separates
the two magnetic polarities that dominate the two polar coronal holes. During
phases of minimum activity such as the years 1963/64 these polar holes extend
to rather low latitude, cf. Fig. 9.8. The warped current sheet, first proposed
by Svalgaard and Wilcox (1976), has been confirmed by the out-of-ecliptic
mission Ulysses (Fig. 9.13).
The main outcome of the preceding section is the hardship, and often im-
possibility, of obtaining the magnetic field in the outer solar atmosphere.
On the other hand, photospheric magnetograms are being made on a rou-
tine basis and, as far as the line-of-sight component of B is concerned, with
good reliability. This situation has led solar astronomers to extrapolate the
photospheric measurements.
The magnetic field measured in the photosphere can be ascribed to a sub-
photospheric electric current. In addition to this current, there is, in general,
an electric current system in the atmosphere itself which also constitutes a
source for the field. But, since we know very little about the atmospheric cur-
rent, we shall ignore it in a first approximation. In this case the atmospheric
field has a potential,
B = −∇Φ , (9.25)
9.3 The Magnetic Field in the Outer Atmosphere 415
and
(rw /r)l+1 − (r/rw )l
fl (r) = . (9.31)
(rw /r )l+1 − (r /rw )l
The radial dependence is such that fl (rw ) = 0. That is, at r = rw we
have a magnetic field pointing in the radial direction. The justification of this
boundary condition is that, at some distance from the Sun, the solar wind
becomes strong enough to force the frozen-in field lines into its own direction
416 9. Chromosphere, Corona, and Solar Wind
Fig. 9.18. Photograph of the eclipse of 30 June 1973, with overlay of a potential
magnetic field. From Altschuler et al. (1977)
of flow. Inspection of coronal images such as Fig. 9.8 suggests that this already
happens in the outer corona. A typical choice is rw = 2.6 r (Altschuler et al.
1977); the subscript w stands for “wind”. The surface r = rw is also called
the source surface because it is equivalent to an electric current in r ≥ rw ,
which adds to the sources of the potential field in r < rw .
The functions fl are normalized to fl (r ) = 1. Therefore, the amplitudes
of the diverse multipoles are entirely given by the expansion coefficients glm
and hml . These coefficients are determined by a least-square fit to the mag-
netic field observed in the photosphere, at r = r . Multipole coefficients up to
N = 90 have been calculated in this manner. The larger the truncation index
N , the more details of the magnetic structure can be represented. An example
with N = 25 is shown in Fig. 9.18. The agreement between the magnetic field
and the coronal structure is fairly good. In particular the distinction between
the closed and open field regions, made in Sect. 9.1.3 above, now becomes
justified. Even better agreement between field and coronal structure can be
obtained by the introduction of a non-spherical source surface (Levine et al.
1982).
Problem 9.10. Confirm that (9.30), with (9.31), is a solution to the Laplace
equation, (9.29). Calculate the energy distribution over the multipoles.
9.3 The Magnetic Field in the Outer Atmosphere 417
Problem 9.11. Show that, for given boundary conditions, the potential field
represents the field of minimum magnetic energy.
Let us now drop the assumption of a potential field. Thus we allow an electric
current, of density j, to flow; with the current goes, at least in principle, a
magnetic volume force j ×B. Now in the outer solar atmosphere the pressure
P is everywhere small in comparison to the magnetic pressure, B 2 /2μ (the
plasma parameter β = 2μP/B 2 is small). Moreover, the dynamic pressure
exerted by the solar wind is small compared to B 2 /2μ unless we go beyond
the Alfvén radius already introduced in Sect. 7.3.2. Hence there are no forces
available to compensate the magnetic force. In an equilibrium the magnetic
force must therefore essentially balance itself, i.e.,
j×B =0 . (9.32)
Condition (9.32) characterizes the force-free magnetic field (Lüst and
Schlüter 1954). Alternatively we may write
curl B = αB , (9.33)
where α in general is a function of the spatial coordinates. However, applica-
tion of the divergence operator immediately yields
B · ∇α = 0 , (9.34)
i.e., α is constant along the field lines.
In the solar atmosphere a force-free magnetic field, i.e., a solution to
(9.33), can be calculated if, in addition to the normal component, Bz , the
value of α is known for each “foot point” of a field line (Sakurai 1981).
Because α does not change along the field line, an independent value can be
prescribed only at one foot point of any closed line.
The vertical component of (9.33) yields
α = (∂By /∂x − ∂Bx /∂y)/Bz . (9.35)
Thus, the desired value of α could be obtained by observing Bz and the
horizontal components, Bx and By , as functions of position on the solar
surface, and by subsequent differentiation. The difficulty of this approach
lies not only in the problem of measuring transverse field components (Sect.
3.5.4), but also in the high spatial resolution and accuracy required: without
such accuracy, the process of differentiation may just produce noise. The
attempts to calculate a force-free field therefore are mostly based on assumed
rather than measured α distributions. Alternatively, one may adjust α in such
a way that the coincidence of the visible structure with the magnetic field
geometry in the corona is optimized.
418 9. Chromosphere, Corona, and Solar Wind
flows in the solar atmosphere has been demonstrated by the work of Brandt
et al. (1988), cf. Fig. 9.20.
An interesting aspect of the force-free magnetic field is that its energy
exceeds the minimum energy of the potential field belonging to the same
photospheric boundary conditions (Problem 9.11). Increasing twist therefore
means increasing storage of field energy. Seehafer (1994) has proposed that
this energy build-up can be described in the framework of mean-field theory,
with the force-free field as the mean field whose twist is increased by fluc-
tuations originating in the photosphere. If, at some stage, the twisted field
becomes unstable, the stored excess energy might be released in an explosive
event. We shall return to this in Sect. 9.5.
9.3.4 Prominences
We now come to a special case where the Lorentz force j × B must not be
balanced by itself, but rather is needed to maintain a static equilibrium. We
mean the magneto-static equilibrium in a quiescent prominence.
According to their state of ionization, and hence their temperature, promi-
nences belong to the chromosphere. At the limb, they offer a most beautiful
view when observed in chromospheric lines (in emission), notably the lines
of neutral hydrogen (Fig. 9.7). On the disc they are seen in filtergrams or
spectroheliograms taken in the same lines; here they appear as dark, thin,
and rather long filaments (Fig. 6.22). Typical values for thickness, height,
and length are 5000 km, 50 000 km, and 200 000 km, respectively.
Since the prominences protrude far above the average solar chromosphere,
they are surrounded by coronal material and may thus be considered as
part of the corona. Because the temperature in the prominence is ≈ 104 K,
about one hundred times smaller than in the surrounding corona, the lateral
pressure equilibrium demands that the density is about one hundred times
the coronal density.
Often prominences remain rather invariable for weeks or even months. In
the following, we shall therefore consider a static equilibrium. With T ≈ 104 K
a pure hydrostatic equilibrium must be excluded: the scale height would
be only ≈ 300 km, which is in conflict with the large vertical extent of the
prominence. The problem is exactly the same which the solar astronomer had
when he thought the corona was cool. In the corona, the solution came with
the recognition of the high temperature. In the prominence the solution must
lie in the support by an electromagnetic force.
Photospheric magnetograms show that prominences predominantly follow
the lines Br = 0. In fact this coincidence is so well documented that it is
possible to outline the polarity of the large-scale solar magnetic field simply
on the basis of Hα observations (McIntosh 1979). The assumption that the
field lines connect the two polarities somewhere in the chromosphere/corona
then yields the model of Kippenhahn and Schlüter (1957), Fig. 9.21 a. Due to
the weight of the prominence material, the crest of the arcade-like magnetic
420 9. Chromosphere, Corona, and Solar Wind
Fig. 9.21. Models for the magnetic support of a prominence (hatched rectangle).
Adapted from Anzer (1987)
field becomes slightly depressed. The ensuing magnetic tension provides the
balancing force.
More formally, let us consider an idealized prominence as a thin sheet of
infinite length. In the magneto-hydrostatic equilibrium,
dP
− − ρg + (j × B)z = 0 , (9.36)
dz
we may discard the pressure gradient because in any case it is by far insuf-
ficient, as already explained. The remaining terms are integrated across the
sheet (in the x direction):
1 ∂Bz
g ρ dx = (j × B)z dx Bx dx . (9.37)
μ ∂x
In the last expression on the right we have replaced j according to Maxwell’s
equation and, because the sheet is thin, we have only retained the derivative
with respect to x. Now the normal component, Bx , is continuous across the
sheet, and may be taken out of the integral. Therefore
1
g ρ dx Bx [[Bz ]] , (9.38)
μ
where [[Bz ]] is the jump, or discontinuity, of the vertical field component
across the sheet.
For a prominence of 5000 km thickness take ρ = 10−10 kg/m3 and
Bx = 10−3 T. Equation (9.38) then yields [[Bz]] ≈ 2 × 10−4 T, which is
a quite moderate change of the vertical field in the prominence. Too moder-
ate, in fact, to be checked by observation.
But we may ask whether the horizontal field, Bx , really has the direction
indicated in Fig. 9.21 a. Because of the above-mentioned ambiguity of trans-
verse field measurements this question is difficult. Most limb prominences
are seen edge-on, or almost edge-on, i.e., with their long (y) axis more or less
parallel to the line of sight. Figure 9.22 illustrates the two possibilities of the
field vector. The two angles, αV , and αF , are known from field measurements;
9.3 The Magnetic Field in the Outer Atmosphere 421
we just do not know which of the two is the true angle of the field with the
prominence, and which is false.
Leroy et al. (1984) used a statistical argument to find a preferred field
angle. Suppose there exists such a preferred value of α, perhaps due to an
alignment effect of the solar differential rotation. Let the two possible angles
be observed for a large sample of prominences, and be plotted as functions
of β, the angle with the line of sight; β itself is determined by observing the
prominence as a filament on the disc, with the assumption that it remains
constant during the passage to the limb. Then, since
αF = −αV − 2β (9.39)
(Fig. 9.22), the true angles, αV , will accumulate around the (constant) pre-
ferred angle, while the false ones, αF , will accumulate along the inclined
straight line given by (9.39). For a sample of 120 relatively high prominences
the result is shown in Fig. 9.23: the preferred angle is found to be ≈ −20o .
That is, the field has a large component By parallel to the prominence, and
the orientation of Bx is such that the field preferentially points from negative
to positive photospheric polarity, i.e., opposite to the orientation of Fig. 9.21 a.
Alternative configurations of prominence fields are illustrated in Figs.
9.21 b and 9.21 c. Thanks to a magnetic neutral point, these two models
(Kuperus and Raadu 1974; Malherbe and Priest 1983) predict the trans-
verse field direction found by Leroy et al. (1984). In the open form, Fig.
9.21 c, they also account for the frequent coincidence of a prominence with
an above-lying coronal streamer if the latter is interpreted by a current sheet
(tangential magnetic field discontinuity). Still, the magnetostatic equilibrium
in the prominence itself is the same as in the original Kippenhahn–Schlüter
model, and is given by (9.38).
In contrast to the sample presented in Fig. 9.23, prominences of smaller
height (< 30 000 km) tend to have Bx components that conform to Fig. 9.21 a.
Since the total number of low prominences is much larger than the number
of high prominences, it may well be that the Kippenhahn–Schlüter configu-
ration correctly describes the majority of cases (Leroy 1988).
As long as the kinetic energy of the solar wind is small in comparison to the
energy of the magnetic field, the latter determines how the wind may blow.
The field keeps some parts of the corona closed altogether. In other parts,
the open regions, it channels the wind along the magnetic lines of force.
Locally the channeling of the solar wind may be described by replacing
the spherically symmetric equation of continuity (9.20) by a relation of form
F ρ v = const., where F is the cross section of a magnetic flux tube, and v is
the velocity component along that tube. Because the flux, BF , is also a con-
stant, one may determine F from B. Hence, it is possible to solve the wind
equations in the case where the field is known. In a consistent treatment,
however, the field B must be calculated together with the wind velocity v in
9.3 The Magnetic Field in the Outer Atmosphere 423
From Problem 2.4 we know J = 1.7 × 1041 kg m2 /s. We divide this by J˙ and
obtain the (present) time scale of the Sun’s rotational braking,
J/J˙ ≈ 7 × 1010 years . (9.41)
Had we used the larger value of vφ quoted above, the result would have been
smaller by a factor of 10. In any case we may conclude that within the Sun’s
life solar-wind braking has been significant, in particular if we recognize that
early in the Sun’s history the braking was probably stronger than today
(Sect. 7.3).
The simplest model that describes the solar braking is axisymmetric and
stationary (Weber and Davis 1967). Moreover, the attention is restricted to
the equatorial region where sin θ ≈ 1.
The azimuthal balance of forces is
vφ rΩ , (9.51)
i.e., rigid rotation with the Sun; on the other hand, at large distance from
the Sun
vφ rA
2
Ω(1 − vA /vr )/r , (9.52)
9.4 The Energy Balance 425
Problem 9.13. Introduce the diverse sin θ factors, which in the present
section were omitted, and show that the solar losses of mass and angular
momentum are related by
2
J˙ = ΩrA 2
ṁ , (9.53)
3
which is a more precise form of (7.14).
The outer solar atmosphere radiates and expands; this takes energy. Although
the total need is only of order 3 × 1022 W, about 10−4 of the solar luminosity,
it is most instructive to consider the diverse processes of energy transport
and energy loss in some detail.
By far the largest energy loss is chromospheric radiation. It occurs mainly
in lines of Ca II and Mg II, in Lyman α, and the H− continuum. A list of the
main chromospheric emitters, according to the empirical models of Vernazza
et al. (1981), is given in Table 9.5. The listed losses are integrals over height
of the emission rates per volume. The latter are shown, as functions of height,
in Fig. 9.24. The lower, broad peak is essentially due to Ca II and Mg II; the
upper, narrow peak is Lyman α radiation and occurs where the temperature
is high enough for the n = 2 level to be populated but not high enough for
complete ionization for hydrogen. These results are typical for the average
quiet Sun, model C of Vernazza et al. (1981). In the chromospheric network
(their model F), the emission rates are higher by factors 2 to 3, and the height
distribution is slightly shifted downwards.
Ca II H 490 Mg II h 430
K 640 k 520
866.2 nm 460 H Lyman α 340
849.8 nm 550 H− bound-free 170
854.2 nm 680 free-free 220
426 9. Chromosphere, Corona, and Solar Wind
The coronal energy balance is more complex. In a steady state, the energy
flux has zero divergence:
1 2 Gm ρ
div v ρv + H − − κ∇T + F R + F H = 0 . (9.54)
2 r
The contributions to this balance are, from left to right: convection of energy
by the solar wind velocity v, thermal conduction, radiation, and heating.
The convected energy consists of kinetic energy, enthalpy, and gravitational
energy. Thermal conduction is, to a smaller part, outwards from the temper-
ature maximum, and, to the larger part, inwards toward the transition layer
and upper chromosphere where it helps to cover the radiation loss in Ly-
man α. Thus, the chromosphere acts as a heat sink for the corona, although
at the same time it provides the mass reservoir for the coronal expansion.
Further out in the solar wind radiation as well as heating become unim-
portant. The fluxes corresponding to the other terms in (9.54) are listed for
1 astronomical unit in Table 9.7. Kinetic energy dominates. The specific en-
thalpy,
5
H= nk(Tp + Te ) , (9.55)
2
has contributions from protons and electrons, and contains both the internal
energy density and the pressure (because work has to be done to expand the
coronal material). The density of gravitational energy is negative; hence the
corresponding flux is inwards. The final entry in Table 9.7 is the thermal con-
duction (here only outwards) by electron collisions. Conduction by protons
is negligibly small.
Problem 9.14. Show that, in spite of its inward direction, the flow of gravi-
tational energy has a positive divergence, and so constitutes a loss. Relate
this loss to the solar mass loss, and estimate its flux density per surface
area on the Sun.
9.4.2 Heating
The term div F H in (9.54) represents the heating of the outer solar atmo-
sphere. Besides the total magnitude of this heating, its spatial distribution
428 9. Chromosphere, Corona, and Solar Wind
Fig. 9.25. Left: Temperature as a function of height: start model (thin dashed),
time average (thick solid), lower and upper edges (thin solid) of the range attained
in the calculation, the static model A (dash-dotted) of Fontenla et al. (1993), and
a “semi-empirical” model (thick dashed) that fits the mean calculated emission.
Right: Evolution of the calculated Ca II H line profile, with shocks occuring every
150 s on the average. From Carlsson and Stein (1995, 1997)
the Ca II lines that are characteristic for the chromosphere (Fig. 9.25, right).
They find the remarkable result that short intervals of very high temperature
are caused by the acoustic shocks. Because of the non-linear dependence of
the Planck function on T this leads to a high emissivity, which can explain the
bright calcium grains seen in spectrograms (Lites et al. 1993) and filtergrams
(von Uexküll and Kneer 1995). They also find that the average chromospheric
temperature continues to decrease outwards from the level where the semi-
empirical models have a temperature minimum (Fig. 9.25, left). Nevertheless,
from the calculated emission they deduce a “semi-empirical” mean temper-
ature that resembles the models shown in Sect. 4.3. This latter result was
however criticized by Kalkofen et al. (1999) who find that only an increasing
mean temperature can account for all of the chromospheric emission, and that
acoustic waves with periods shorter than 100 s, which were not considered by
Carlsson and Stein, would yield the necessary heating.
Magnetohydrodynamic Waves. While acoustic waves seem to be ade-
quate to heat the lower chromosphere, these waves probably cannot solve the
problem of coronal heating. Their energy is dissipated before they reach the
corona. Also, in the corona the spatial association of strong (closed) magnetic
field configurations with large radiative and conductive losses is so obvious
that we must expect the field to play a more central role. The heating could
be due either to the dissipation of magnetohydrodynamic waves, or to direct
dissipation of an electric current.
430 9. Chromosphere, Corona, and Solar Wind
a resonance occurs. At this resonance the ions are accelerated, and hence
that species is preferentially heated. For any given wave frequency, and an
outwards decreasing magnetic field strength B, the ions with smaller q/m
will be heated first, while for those with larger q/m the wave-energy flux is
reduced. The apparent decrease of the ion temperature with increasing q/m,
as deduced from the width of coronal emission lines, supports this interpre-
tation (Tu et al. 1998, Kohl et al. 1998). Marsch et al. (1982) had already
discussed the ion-cyclotron resonance in the solar wind and suggested that
the higher transverse temperature of the protons (Table 9.4) results from this
resonance.
Electric Current Dissipation. Coronal heating by direct current dissipa-
tion may occur even for slow photospheric motion. This mechanism has been
advanced by Parker (1972, 1983) and has been called topological dissipation.
The name explains the principle: coronal field lines of opposite direction are
brought into close contact because of the random motion of their photospheric
foot points. Parker points out that the narrow sheets where the current den-
sity j is large are unstable. The magnetic lines of force reconnect, just as we
9.5 Explosive Events 431
have already seen in the case of magneto-convection, Sect. 8.2.1; the current
in the sheet is dissipated. The rate of energy dissipation per volume is j 2 /σ,
where σ is the electrical conductivity. Since j is large, this may be significant
in spite of the large value of σ, and in spite of the small volume fraction
occupied by the current sheets. Presumably, the field line reconnection takes
place in form of numerous unresolved events, called nanoflares by Parker
(1988). Parker estimates the energy of a single nanoflare to 1017 J and less,
which is only 10−3 (or less) of the energy released during the x-ray bursts
(microflares) observed by Lin et al. (1984), but is perhaps manifest in the
smallest individual emission spikes during those bursts.
Flow Acceleration by Waves. Acoustic and magnetohydrodynamic waves
are not only able to heat the outer solar atmosphere. These waves can also
directly exert a net force and so contribute to the acceleration of an outward
flow. In a mean-field approach, such as already described in the context of
the solar differential rotation and the solar dynamo, the fluctuating wave
field u leads to a mean wave pressure ρ0 u2 /2 whose gradient may help driving
the solar wind, or the upflow seen in spicules. Belcher (1971) and Alazraki
and Couturier (1971) suggested that Alfvén waves might accelerate the solar
wind in this manner. The original model of Parker (1958), where the coronal
expansion was based only on the thermal pressure gradient, is thus improved
in that the acceleration occurs closer to the Sun, and yields a higher wind
speed than was possible with increased heating alone (which essentially lead
to enhanced heat conduction back to the Sun).
While wave pressure is an effect that equally occurs for damped and un-
damped waves, Haerendel (1992) has pointed out that damped Alfvén waves
may exert a net Lorentz force. In a dense plasma the transverse field and
current excursions of an Alfvén wave are 90o out of phase so that, averaged
over a wave period, the Lorentz force vanishes. But when collisions become
less frequent, and the gas is only partially ionized as in the solar chromo-
sphere, the ions and neutral particles slightly slip relative to each other. This
constitutes yet another damping mechanism for the Alfvén wave; moreover,
it yields a non-vanishing average Lorentz force in the longitudinal direction.
Haerendel has proposed that spicules might be driven by this force.
The nano- and microflares mentioned in the preceeding section, and the explo-
sive events defined by Dere et al. (1989), release energies of up to ≈ 1020 J, and
the smallest of them may not be observable individually. We shall now treat
more violent events. The largest have a total energy release of order 1026 J,
and the associated energy flux may exceed the photospheric flux which, on
the average, is ≈ 104 times the flux originating from the outer layers of the
Sun. In view of the common origin, namely an instability of a magnetic field
432 9. Chromosphere, Corona, and Solar Wind
configuration in the solar atmosphere, we shall now use the term explosive
event with a more general meaning for small and large explosions.
Fig. 9.26. Two-ribbon flare of 28 May 1972. (a): pre-flare state, 11.37 UT; (b):
flare at maximum intensity, 13.27 UT; (c): expanding ribbons, 14.25 UT; (d): post-
flare state, 29 May, 9.01 UT. Photographs in Hα, Domeless Coudé Telescope, Capri
(Bruzek 1979)
434 9. Chromosphere, Corona, and Solar Wind
For events at the solar limb the height in the atmosphere can be deter-
mined by spatially resolved observations. Using this and the form of the x-ray
spectra, one can explain the radiation as bremsstrahlung of electrons with en-
ergies 10–100 keV precipitating in beams from the corona down to the denser
atmosphere (e.g., Haug and Elwert 1985). The microwave emission can be
interpreted as synchrotron radiation of the same electron beams.
Large flares may be visible in white light. These white-light flares spatially
coincide with the area emitting in Hα, and are synchronous with the impulsive
phase.
After the impulsive phase thermal radiation dominates. Heating of the
chromosphere leads to increased excitation of the atomic levels responsible
for Hα and the other flaring lines. The heating may be so effective that part
of the chromosphere is “evaporated”, i.e., transformed into coronal material.
This material could be the hot plasma, with T ≈ 3 × 107 K, responsible for
the thermal radiation of the main flare phase. An alternative idea is that such
a hot plasma originates directly during the primary energy release high in
the atmosphere. In any case heat will be conducted downwards and so supply
energy to the Hα flare.
9.5 Explosive Events 435
Energetic Particles and γ-rays. The evidence arising from the diverse
radiation processes for electrons of 10–100 keV associated with solar flares
is confirmed by in-situ measurements of these particles in the solar wind.
In addition, large flares are accompanied by relativistic electrons and high-
energy protons (proton flares); tens of MeV are common, and exceptional
flares accelerate protons to energies exceeding 1 GeV.
The γ-rays, too, are characteristic for large events. Their spectrum shows
very intense line radiation. One strong line, at 511 keV, originates from the
annihilation of electron-positron pairs (positrons are emitted by radioactive
nuclei); another, at 2.223 MeV and still stronger, results from the capture of
a neutron by protons: n + p → d + γ. The presence of radioactive nuclei and
of neutrons indicates that nuclear reactions take place during flares. Further
evidence for such reactions comes from the abnormal abundance of many
nuclei, e.g., 3 He, measured in interplanetary space in the wake of flares.
Fig. 9.28. Coronal mass ejection and prominence eruption of 18 August 1980. The
originally stable helmet streamer expands with a nearly spherical front (frames 2–
4). The eruptive prominence first appears (frame 2) at the edge of the occulting
disc of the Solar Maximum Mission coronagraph, grows, and explodes in form of
bright thin filaments (frames 3–6). Courtesy High Altitude Observatory, NCAR,
Boulder, Colorado
Fig. 9.29. Prominence eruption and coronal mass ejection of 2 June 1998. The
front of the CME has passed the field of view, but a faint cone is still visible. The
bright, helical filaments are the remainders of the prominence. White-light image,
from the LASCO C2 coronagraph on SOHO. The white circle on the occulting disc
marks the solar surface. SOHO is a cooperation between ESA and NASA
Fig. 9.30. Coronal wave, expanding from the site of a coronal mass ejection on
12 May 1997. The images show intensity differences of successive exposures (UT
4.50 − 4.35, 5.07 − 4.50, 5.24 − 5.07), obtained by SOHO/EIT in the Fe XII band
at 19.5 nm. The front speed is ≈ 300 km/s. SOHO is a cooperation between ESA
and NASA
438 9. Chromosphere, Corona, and Solar Wind
The total energy emitted in an explosive event in the solar atmosphere may
exceed 1025 J. As a detailed case study (Sturrock 1980) shows, most of this
energy lies in the masses ejected in eruptive prominences and coronal mass
ejections, and in the magnetic field carried with these masses. Only a few
percent are radiated in the diverse spectral ranges.
The large majority of explosive events occurs in active regions, i.e., re-
gions of enhanced and complex magnetic field. And, as we have seen, there is
evidence that changes of the field configuration occur during flares. Finally,
the energy contained in the magnetic field is of the right order: take a volume
of 1023 m3 and a field strength of 0.01 T, then the energy is ≈ 4 × 1024 J, suffi-
cient for a typical event. For these reasons one must conclude that the nature
of the explosion is a sudden release of energy stored in the magnetic field.
Since, for given boundary conditions, the potential field has the minimum
energy (Problem 9.11), one would expect a field change towards a potential
configuration. So far such is not well-documented, however. Actually cases
have been studied (e.g., Wang et al. 1994) where the deviation from a poten-
tial field increases. It appears that new magnetic flux that emerges during
a flare enhances the magnetic energy, as already noticed by Schmidt (1964),
and so obscures the energy release during the instability proper.
The main questions, then, are the following. How is the energy stored in
the magnetic field? By which instability is it released? How are the particles
accelerated? And how are the masses of, e.g., coronal mass ejections set into
motion? These primary questions (Sturrock 1980) must be answered by the-
oretical models. Synchrotron radiation in microwaves, x-ray bremsstrahlung,
heating and the ensuing thermal radiation, etc., may then be considered as
secondary processes whose basic physics is known (although many quantita-
tive details remain open).
9.5 Explosive Events 439
Fig. 9.31. Two magnetic field configurations which possibly lead to flares. Left:
from Sturrock (1980); right: from Heyvaerts et al. (1977)
The magnetic instability itself must develop within seconds. This in the
time scale of the single spikes which we see in the impulsive phase. Various
magnetohydrodynamic instabilities have been proposed. A single twisted loop
like the one shown in Fig. 9.19 could be kink unstable, i.e., develop a number
of local, strongly twisted kinks that would allow rapid current dissipation.
On the other hand, a current sheet could be unstable to tearing, i.e., the
two adjacent regions of opposite field direction would tear magnetic flux
across the interface and so form alternating sequences of O-type and X-
type neutral points. Fig. 9.31 (left) shows a model that is characterized by
field-line reconnection at an X-type neutral point. According to this model,
440 9. Chromosphere, Corona, and Solar Wind
which was first proposed by Kopp and Pneuman (1976), a lateral inflow
towards the neutral point, and mass ejection from above the neutral point
is expected. Indeed Tsuneta (1996 b) and and Yokoyama et al. (2001) report
cases where the inflow can be inferred from the change of the soft x-ray and
extreme UV emission, as observed by Yohkoh and SOHO/EIT. Similarly,
an upward moving mass has been observed in soft x-rays in another case
(Tsuneta 1997). The model is also supported by Hα observations with a time
resolution Δt < 0.1 s (Wang et al. 2000); these observations indicate that
the initial energy release occurs at the top of a low-lying loop, followed by
emission from loop foot points on both sides.
Quantitative models of explosive events have been two-dimensional in
most cases. An unstable field configuration such as the current sheet of
Fig. 9.31 (left) is assumed, and the instability is triggered by introducing
a small region of anomalous, that is enhanced, electrical resistivity. Field-
line reconnection then proceeds in the tearing mode (e.g., Schumacher and
Kliem 1996, Magara and Shibata 1999, Roussev et al. 2001, Yokoyama and
Shibata 2001). Three-dimensional models are more difficult, because of the
increased computational expenditure and because of the complicated topol-
ogy of magnetic reconnection (e.g., Priest and Schrijver 1999, Birn et al. 2000,
Galsgaard et al. 2000, Schumacher et al. 2000, Brown and Priest 2001).
Schrijver and Zwaan (2000), and in many conference proceedings, e.g., Bonnet
and Dupree (1981) or Schröter and Schüssler (1987), and the series Cool Stars,
Stellar Systems, and the Sun, which appeared since 1980 in a two-year cycle.
Alternative techniques of calculating a potential field in the corona have
been applied by Adams and Pneuman (1976), and by Elwert et al. (1982).
Pneuman et al. (1978) compared the potential field to the observed coronal
structure and generally found good correspondence. Hagyard and Pevtsov
(1999) discuss the use of vector magnetograms for deriving the factor α that
connects B and curl B in a force-free field.
Tandberg-Hanssen (1974, 1995) introduces solar prominences. Querfeld et
al. (1985) and Bommier et al. (1986 a) report further attempts to measure the
magnetic field vector in prominences, while Bommier et al. (1986 b) use the
depolarizing effect of collisions to derive the electron density. The volumes
edited by Jensen et al. (1979), by Ballester and Priest (1988), and by Priest
(1989) are entirely devoted to the physics of solar prominences. Cheng and
Choe (1998) construct models of current sheet and prominence formation
through photospheric plasma motion. The energy balance of prominences,
especially in view of their fibril structure, is discussed by Anzer and Heinzel
(1999) and Heinzel and Anzer (2001).
Durney and Pneuman (1975) describe how a given magnetic field guides
the solar wind. For the case of axial symmetry consistent solutions for wind
and field have been calculated by Pneuman and Kopp (1971), Endler (1971),
and Sakurai (1985). The latter of these includes the effect of solar rotation,
and extends the problem posed by Weber and Davis (1967; cf. Sect. 9.3.5)
to the full range of heliographic latitude. Durney and Stenflo (1972) employ
the assumption that in the past the total (absolute) magnetic flux across the
solar surface has varied in proportion to the Sun’s angular velocity Ω; they
thus calculate a rotational braking of the form Ω ∝ t−1/2 , which reproduces
the Ω(t) derived from stellar clusters (Fig. 7.1).
Chromospheric and coronal heating has been reviewed by Narain and
Ulmschneider (1990, 1996); Hammer (1987, 1988) discusses the physics of the
whole system chromosphere, transition layer, and corona. An important part
in the chromospheric energy balance could be played by molecules such as
CO (Ayres 1981; Kneer 1983): molecules are strong emitters, i.e., coolants; at
the same time, their formation is rather temperature-sensitive. This may lead
to a two-state situation where the chromosphere proper exists only in narrow
regions, perhaps confined in magnetic flux tubes, while the intermediate space
is rather cool. The cooling effectivity of CO has been disputed, however, by
Mauas et al. (1990).
Coronal heating by electric current dissipation is further elaborated by
Browning et al. (1986) and by Browning and Priest (1986). Parker (1987)
gives an illustrative account. Details of the ion-cyclotron resonance have been
worked out by Cranmer et al. (1999 a) and Cranmer (2000).
442 9. Chromosphere, Corona, and Solar Wind
Spicule models have been reviewed by Sterling (2000). Hammer and Nesis
(2002) point out that spicule-driving by low-pressure regions above the chro-
mosphere would require less fine-tuning to get the right velocity and height
than acceleration from below.
Flares are introduced by Švestka (1976). The study of explosive events
has been greatly advanced through the results of space missions, especially
Skylab, Yohkoh, and SOHO. The volume edited by Priest (1981) concentrates
on the magnetohydrodynamic aspects, that of McLean and Labrum (1985)
on the physics of the radio emission. Solar radio astronomy is also introduced
by Kundu (1965) and Krüger (1979). Gosling (1993) emphasizes that coronal
mass ejections cause geomagnetic disturbances. Reames (1999) reviews the
mechanisms of particle acceleration.
Reconnection of magnetic lines of force was first studied by Sweet (1958)
and Petschek (1964). Soward and Priest (1982) presented a detailed mathe-
matical model. The existence and stability of magnetohydrostatic equilibria
as possible pre-flare states has been reviewed by Schindler et al. (1983), while
Low (1984) treats special magnetohydrodynamic models of coronal mass ejec-
tions.
List of Symbols
A Vector potential
A 1. Amplitude of light wave
2. Astronomical unit
3. Atomic weight
4. 1012 times element abundance relative to hydrogen
5. Azimuthal component of vector potential
6. Constant in law of differential rotation
7. Area of sunspot
AUL Einstein coefficient for spontaneous emission
AZ Mean atomic weight of atoms heavier than helium
B Magnetic field vector
Bp Poloidal magnetic field vector
Bt Toroidal magnetic field vector
B 1. Kirchhoff–Planck function
2. Magnitude (or component) of magnetic field
3. Constant in law of differential rotation
BLU Einstein coefficient for radiative excitation
BP Field in equilibrium with pressure
BUL Einstein coefficient for induced emission
Bc Critical magnetic field strength
Be Equipartition field strength
B0 1. Heliographic latitude of disc center
2. Initial magnetic field strength
C 1. Intensity contrast of granulation
2. Constant in law of differential rotation
Cij Rate of collisional transition
D 1. Aperture (diameter) of telescope
2. Diameter of granule
3. Dynamo number
DS Aperture of spectrograph
E Electric field vector
E 1. Number fraction electrons/all particles
2. Energy of nuclear reaction
E u × b, vector of mean electric field
444 List of Symbols
NA Avogadro number
O(x) Term of first order in x
P 1. Pressure
2. Degree of polarization
3. Power spectrum
4. Position angle of northern direction on Sun
PES Electrostatic correction to pressure
PG Gas pressure
PI Ion contribution to gas pressure
PPG Perfect gas pressure
PR Radiation pressure
Pe 1. Electron pressure
2. Pressure external to flux tube
Ps Pressure at surface of solar model
P0 1. Equilibrium pressure
2. Pressure on the axis of flux tube
P1 Eulerian pressure perturbation
PSF Point spread function
Q 1. Energy release per nuclear reaction
2. Stokes parameter of linear polarized light
3. ν/Δν, quality factor
4. Energy gain per volume of granular flow
Qij Reynolds stress
Qν Energy of neutrino
Q Energy release exclusive of neutrino energy
R 1. Reflected fraction of intensity
2. Sunspot relative number
R Gas constant
RT Radius of flux tube
Rij Rate of transition from ith to jth state
Rm Magnetic Reynolds number
Ro Rossby number
S 1. Irradiance (energy flux) at mean solar distance
2. quantum number of spin angular momentum
3. Source function
4. Specific entropy
Si (q) Transfer function of ith image
Sl 1. Line source function
2. Limit frequency for acoustic oscillations
T 1. Absolute temperature
2. Orbital period
3. Period of solar oscillation
4. Transmitted fraction of intensity
TB Brightness temperature
446 List of Symbols
a 1. Grating constant
2. 4σ/c, radiation constant
3. Semi-major axis of elliptical orbit
b Fluctuating part of magnetic field vector
b̂ Binormal vector of thin flux tube
c 1. Velocity of light
2. Velocity of sound
3. Oblateness parameter
cA Alfvén velocity
cP Specific heat at constant pressure
cT Tube speed
cV Specific heat at constant volume
c2 hc/k, radiation constant
d 1. Diameter of solar image
2. Distance between centers of granules
3. Width of magnetic sheet or tube
d Deuteron
e 1. Elementary charge
2. Thickness of first crystal in Lyot filter
e− Electron
e+ Positron
List of Symbols 447
f 1. Focal length
2. Oscillator strength
3. Filling factor for convecting parcels
g Vector of gravitational acceleration
g 1. Gravitational acceleration
2. Landé factor of atomic state
gL Statistical weight of lower state
gU Statistical weight of upper state
g eff Vector of effective gravitational acceleration
gij Statistical weight of jth state of ith atom
glm Coefficient of multipole expansion
g∗ g factor for atomic transition
g Gravitational acceleration at present Sun’s surface
h 1. Planck constant
2. Height in solar atmosphere above the level where τ500 = 1
hml Coefficient of multipole expansion
i Inclination between ecliptic and solar equator
j Vector of electric current density
k Wave vector
k 1. Gaussian gravitational constant
2. Boltzmann constant
3. Wave number
kNy Nyquist wave number
kh Horizontal wave number
kr Radial wave number
l̂ Tangent vector of thin flux tube
l 1. Mixing length
2. Size of convection cell; length scale of fluctuation
3. Degree of spherical harmonic
m 1. Mass
2. Reduced mass
3. Apparent magnitude of a star
4. Order of spectrum
5. Longitudinal order of spherical harmonic
mH Atomic mass of hydrogen
me Electron mass
m Mass of present Sun
n̂ Principal normal vector of thin flux tube
n 1. Number of grooves in a diffraction grating
2. Principal quantum number
3. Index of refraction
4. Order (radial node number) of oscillations
5. Coordinate in normal direction
n Neutron
448 List of Symbols
Γ1 Adiabatic exponent
Δ 1. Focus tolerance
2. Laplacian operator
ΔS Total entropy change over convection zone
ΔT 1. Temperature difference between granules and intergranular space
2. Temperature excess of displaced parcel
Δr Difference between equatorial and polar solar radii
Δt 1. Time delay in Doppler spectroheliogram
2. Time resolution of observed oscillation
Δλ Wavelength distance to line center
ΔλB Zeeman splitting
ΔλD Doppler width of line (in wavelength)
ΔλG Gravitational redshift
ΔνD Doppler width of line (in frequency)
Δρ Density difference of displaced parcel
Δωα Rotational shift of oscillation frequency ωα
Δ∇ ∇ − ∇a , excess of temperature gradient over adiabatic gradient
Λ Ratio of Debye radius to 90o -impact parameter
Λh Coefficient of horizontal non-diffusive momentum flux
Λr Coefficient of vertical non-diffusive momentum flux
Θ 1. Characteristic angle of internal gravity waves
2. Moment of inertia
Φ 1. Gravitational potential
2. Magnetic flux
3. Potential of current-free magnetic field
450 List of Symbols
α 1. Helium nucleus
2. Ratio of mixing length to pressure scale height
3. Angle between flux tube axis and vertical
4. Coefficient of α effect in dynamo
5. Coefficient of force-free field
6. Angle of prominence with magnetic vector
αj Cross section of state j for photoionization
β 1. Ratio of gas pressure to total pressure
2. 2μP/B 2 , Plasma parameter
3. Contribution of turbulence to mean diffusivity
4. Angle of prominence with line of sight
γ 1. Angle of magnetic field with line of sight
2. Constant value of adiabatic exponent
3. Damping constant
4. Photon
δ −(∂ ln ρ/∂ ln T )P
δik Kronecker symbol
δP Lagrangian pressure perturbation
δr Radial displacement
δν Asymptotic frequency separation
δρ Lagrangian density perturbation
ε 1. Energy release per mass
2. Phase angle of polarized light wave
3. Element abundance relative to hydrogen
4. Correction factor in thermal conductivity
εC Continuous emission
εl Line emission
ε0 Dielectric constant of free space
η Absorption matrix for Stokes vector
η 1. Efficiency of Fourier spectrometer
2. Temperature exponent of nuclear energy release
3. Magnetic diffusivity
4. Ratio of line absorption to continuum absorption
5. Degree of ionization
List of Symbols 451
∇ 1. Gradient operator
2. d ln T /d ln P , double-logarithmic temperature gradient
∇R Radiative double-logarithmic temperature gradient
∇a Adiabatic double-logarithmic temperature gradient
∇ Mean double-logarithmic temperature gradient of parcels
Ecliptic longitude of ascending node of Sun’s equator
Subscript denoting present Sun
approximately equal (physical relations)
≈ approximately equal (numbers)
∝ proportional to
References
Abdelatif, T.E., Lites, B.W., Thomas, J.H. (1984): In Keil (1984), p. 141
Abel, W.G., Chisholm, J.H., James, J.C. (1963): In Space Research III, ed. by W.
Priester (North-Holland, Amsterdam), p. 635
Abraham, Z., Iben, I., Jr. (1971): Astrophys. J. 170, 157
Adams, J., Pneuman, G.W. (1976): Solar Phys. 46, 185
Adelberger, E.G., Austin, S.M., Bahcall, J.N., and 36 others (1998): Rev. Mod.
Phys. 70, 1265
Ahmad, Q.R., Allen, R.C., Andersen, T.C., and 175 others (2001): Phys. Rev.
Letters 87, 071301
Ahrens, B., Stix, M., Thorn, M. (1992): Astron. Astrophys. 264, 673
Alazraki, G., Couturier, P. (1971): Astron. Astrophys. 13, 380
Albregtsen, F., Maltby, P. (1978): Nature 274, 41
Albregtsen, F., Maltby, P. (1981): In Cram and Thomas (1981), p. 127
Aldrich, L.B., Hoover, W.H. (1954): Ann. Astrophys. Obs. Smithsonian Inst. 7
Alfvén, H. (1947): Mon. Not. R. Astron. Soc. 107, 211
Alfvén, H. (1954): On the Origin of the Solar System (Clarendon Press, Oxford)
Allen, C.W. (1947): Mon. Not. R. Astron. Soc. 107, 426
Aller, L.H. (1963): Astrophysics, The Atmospheres of the Sun and Stars, 2nd ed.
(Ronald Press, New York)
Altschuler, M.D., Levine, R.H., Stix, M., Harvey, J. (1977): Solar Phys. 51, 345
Altschuler, M.D., Newkirk, G., Jr. (1969): Solar Phys. 9, 131
Anders, E., Grevesse, N. (1989): Geochimica et Cosmochimica Acta 53, 197
Andersen, T.E., Dunn, R.B., Engvold, O. (1984): LEST Foundation, Technical Rep.
No. 7
Ando, H., Osaki, Y. (1975): Publ. Astron. Soc. Japan 27, 581
Antia, H.M. (1998): Astron. Astrophys. 330, 336
Antia, H.M., Chitre, S.M. (1995): Astrophys. J. 442, 434
Anzer, U. (1987): In Hillebrandt et al. (1987), p. 61
Anzer, U., Heinzel, P. (1999): Astron. Astrophys. 349, 974
Appenzeller, I. (1982): Fundamentals of Cosmic Physics 7, 313
Appenzeller, I., Tscharnuter, W. (1975): Astron. Astrophys. 40, 397
Appourchaux, T., Fröhlich, C., Andersen, B., and 12 others (2000): Astrophys. J.
538, 401
Arvesen, J.C., Griffin, R.N., Jr., Pearson, B.D., Jr. (1969): Applied Optics 8, 2215
Aschwanden, M.J., Nightingale, R.W., Alexander, D. (2000): Astrophys. J. 541,
1059
Aschwanden, M.J., Schrijver, C.J., Alexander, D. (2001): Astrophys. J. 550, 1036
Asplund, M., Ludwig, H.-G., Nordlund, Å., Stein, R.F. (2000 a): Astron. Astrophys.
359, 669
Asplund, M., Nordlund, Å., Trampedach, R., Allende Prieto, C., Stein, R.F.
(2000 b): Astron. Astrophys. 359, 729
454 References
Asplund, M., Nordlund, Å., Trampedach, R., Stein, R.F. (2000 c): Astron. Astro-
phys. 359, 743
Athay, R.G. (1972): Radiation Transport in Spectral Lines (Reidel, Dordrecht)
Athay, R.G. (1976): The Solar Chromosphere and Corona: Quiet Sun (Reidel, Dor-
drecht)
Atkinson, R. d’E., Houtermans, F.G. (1929): Z. Phys. 54, 656
Audouze, J., Israël, G., eds. (1985): The Cambridge Atlas of Astronomy (Cambridge
University Press, Cambridge)
Auwers, A. (1891): Astron. Nachr. 128, 361
Avrett, E.H. (1981a): In Cram and Thomas (1981), p. 235
Avrett, E.H. (1981b): In Bonnet and Dupree (1981), p. 173
Axford, W.I., McKenzie, J.F. (1992): In Solar Wind Seven, ed. by E. Marsch and
R. Schwenn (Pergamon Press, Oxford), p. 1
Ayres, T.R. (1981): Astrophys. J. 244, 1064
Barletti, R., Ceppatelli, G., Paternó, L., Righini, A., Speroni, N. (1977): Astron.
Astrophys. 54, 649
Bartenwerfer, D. (1973): Astron. Astrophys. 25, 455
Basset, A.B. (1961): A Treatise on Hydrodynamics, republished (Dover, New York)
Basu, S., Antia, H.M. (1994): Mon. Not. R. Astron. Soc. 269, 1137
Basu, S., Antia, H.M. (1997): Mon. Not. R. Astron. Soc. 287, 189
Baturin, V.A., Mironova, I.V. (1998): In Korzennik and Wilson (1998), p. 717
Baumbach, S. (1937): Astron. Nachr. 263, 121
Beck, J.G. (2000): Solar Phys. 191, 47
Becker, U. (1954): Z. Astrophys. 34, 129
Beckers, J.M. (1969 a): A Table of Zeeman Multiplets, Sacramento Peak Observa-
tory, Rep. AFCRL-69-0115
Beckers, J.M. (1969 b): Solar Phys. 9, 372
Beckers, J.M. (1970): Applied Optics 9, 595
Beckers, J.M. (1971 a): In Howard (1971), p. 3
Beckers, J.M. (1971 b): Applied Optics 10, 973
Beckers, J.M. (1976): Astrophys. J. 203, 739
Beckers, J.M. (1977): Astrophys. J. 213, 900
Beckers, J.M. (1981): In Jordan (1981), p. 11
Beckers, J.M. (1993): Solar Phys. 145, 399
Beckers, J.M. (1995): In Kuhn and Penn (1995), p. 145
Beckers, J.M., Bridges, C.A., Gilliam, L.B (1976): Rep. AFGL-TR-76-0126 (II)
Beckers, J.M., Brown, T.M. (1978): Osserv. Mem. Oss. Astrofis. Arcetri 106, 189
Beckers, J.M., Canfield, R.C. (1976): In Cayrel and Steinberg (1976), p. 207
Beckers, J.M., Dickson, L., Joyce, R.S. (1975): Sacramento Peak Observatory Rep.
AFCRL-TR-75-0090
Beckers, J.M., Milkey, R.W. (1975): Solar Phys. 43, 289
Beckers, J.M., Schröter, E.H. (1968): Solar Phys. 4, 142
Beckers, J.M., Schröter, E.H. (1969): Solar Phys. 10, 384
Beckers, J.M., Tallant, P.E. (1969): Solar Phys. 7, 351
Beer, J., Tobias, S., Weiss, N. (1998): Solar Phys. 181, 237
Belcher, J.W. (1971): Astrophys. J. 168, 509
Bellot Rubio, L.R., Rodrı́guez Hidalgo, I., Collados, M., Khomenko, E., Ruiz Cobo,
B. (2001): Astrophys. J. 560, 1010
Bellot Rubio, L.R., Ruiz Cobo, B., Collados, M. (2000): Astrophys. J. 535, 475
Belvedere, G., Godoli, G., Motta, S., Paternó, L., Zappalá, R.A. (1976): Solar Phys.
46, 23
Belvedere, G., Kuzanyan, K.M., Sokoloff, D. (2000): Mon. Not. R. Astron. Soc.
315, 778
Belvedere, G., Paternó, L., eds. (1978): Workshop on Solar Rotation, University of
Catania
Belvedere, G., Paternó, L., eds. (1984): Mem. Soc. Astron. Ital. 55, No. 1–2
Bendlin, C., Volkmer, R., Kneer, F. (1992): Astron. Astrophys. 257, 817
Bennett, K., Roberts, B., Narain, U. (1999): Solar Phys. 185, 41
Berger, M.A., Ruzmaikin, A. (2000): J. Geophys. Res. 105, 10481
Berger, T.E., Löfdahl, M.G., Shine, R.A., Title, A.M. (1998): Astrophys. J. 495,
973
Berger, T.E., Schrijver, C.J., Shine, R.A., Tarbell, T.D., Title, A.M., Scharmer, G.
(1995): Astrophys. J. 454, 531
Berrington, K.A., ed. (1997): The Opacity Project, Vol. 2 (Institute of Physics,
Bristol and Philadelphia)
Bertello, L., Henney, C.J., Ulrich, R.K., and 8 others (2000): Astrophys. J. 535,
1066
456 References
Berthomieu, G., Cooper, A.J., Gough, D.O., Osaki, Y., Provost, J., Rocca, A.
(1980): In Hill and Dziembowski (1980), p. 307
Berthomieu, G., Provost, J., Morel, P., Lebreton, Y. (1993): Astron. Astrophys.
268, 775
Bethe, H.A. (1939): Phys. Rev. 55, 434
Bethe, H.A. (1986): Phys. Rev. Letters 56, 1305
Bethe, H.A., Critchfield, C.L. (1938): Phys. Rev. 54, 248
Bianda, M., Solanki, S.K., Stenflo, J.O. (1998): Astron. Astrophys. 331, 760
Bianda, M., Stenflo, J.O., Solanki, S.K. (1999): Astron. Astrophys. 350, 1060
Biermann, L. (1932): Z. Astrophys. 5, 117
Biermann, L. (1937): Astron. Nachr. 263, 185
Biermann, L. (1941): Vierteljahresschr. Astron. Ges. 76, 194
Biermann, L. (1946): Naturwiss. 33, 118
Biermann, L. (1951 a): Z. Astrophys. 28, 304
Biermann, L. (1951 b): Z. Astrophys. 29, 274
Biermann, L. (1977): In Spiegel and Zahn (1977), p. 4
Billings, D.E. (1966): A Guide to the Solar Corona (Academic Press, New York)
Birn, J., Gosling, J.T., Hesse, M., Forbes, T.G., Priest, E.R. (2000): Astrophys. J.
541, 1078
Biskamp, D. (1993): Nonlinear Magnetohydrodynamics (Cambridge University Press,
Cambridge)
Bodenheimer, P. (1983): Lectures in Appl. Math. 20, 141
Bodenheimer, P. (1995): Ann. Rev. Astron. Astrophys. 33, 199
Bodenheimer, P., Yorke, H.W., Różycka, M., Tohline, J.E. (1990): Astrophys. J.
355, 651
Boercker, D.B. (1987): Astrophys. J. 316, L95
Bogdan, T.J. (2000): Solar Phys. 192, 373
Böhm-Vitense, E. (1958): Z. Astrophys. 46, 108
Bommier, V., Leroy, J.L., Sahal-Bréchot, S. (1986 a): Astron. Astrophys. 156, 79
Bommier, V., Leroy, J.L., Sahal-Bréchot, S. (1986 b): Astron. Astrophys. 156, 90
Bonnet, R.M., Bruner, E.C., Jr., Acton, L.W., Brown, W.A., Decaudin, M. (1980):
Astrophys. J. 237, L47
Bonnet, R.M., Dupree, A.K., eds. (1981): Solar Phenomena in Stars and Stellar
Systems (Reidel, Dordrecht)
Boothroyd, A.I., Sackmann, I.-J., Fowler, W.A. (1991): Astrophys. J. 377, 318
Bord, D.J., Cowley, C.R., Mirijanian, D. (1998): Solar Phys. 178, 221
Bouvier, J., Bertout, C., Benz, W., Mayor, M. (1986): Astron. Astrophys. 165, 110
Bracewell, R. (1965): The Fourier Transform and Its Applications (McGraw-Hill,
New York)
Braginsky, S.I. (1965): In Reviews in Plasma Physics, ed. by M.A. Leontovich,
Consultants-Bureau, New York, Vol. 1, p. 205
Brajša, R., Wöhl, H., Vršnak, B., Ruždjak, V., Clette, F., Hochedez, J.-F. (2001):
Astron. Astrophys. 374, 309
Brandenburg, A., Dobler, W. (2001): Astron. Astrophys. 369, 329
Brandenburg, A., Dobler, W., Subramanian, K. (2002): Astron. Nachr. 323, 99
Brandenburg, A., Schmitt, D. (1998): Astron. Astrophys. 338, L55
Brandt, P.N. (1969): Solar Phys. 7, 187
Brandt, P.N. (1970): Joint Org. Solar Obs., Ann. Rep. 1970, p. 50
Brandt, P.N., Ferguson, S., Shine, R.A., Tarbell, T.D., Scharmer, G.B. (1991):
Astron. Astrophys. 241, 219
Brandt, P.N., Mauter, H.A., Smartt, R. (1987): Astron. Astrophys. 188, 163
Brandt, P.N., Scharmer, G.B., Ferguson, S., Shine, R.A., Tarbell, T.D., Title, A.M.
(1988): Nature 335, 238
References 457
Cohen, E.R., Taylor, B.N. (1987): Rev. Mod. Phys. 59, 1121
Collados, M., Vázquez, M. (1987): Astron. Astrophys. 180, 223
Cowling, T.G. (1934): Mon. Not. R. Astron. Soc. 94, 39
Cowling, T.G. (1941): Mon. Not. R. Astron. Soc. 101, 367
Cowling, T.G. (1976): Magnetohydrodynamics (Adam Hilger, Bristol)
Cowling, T.G. (1981): Ann. Rev. Astron. Astrophys. 19, 115
Cox, A.N., ed. (2000): Allen’s Astrophysical Quantities, 4th ed. (Springer, New
York, Berlin, Heidelberg)
Cox, A.N., Guzik, J.A., Kidman, R.B. (1991): Astrophys. J. 342, 1187
Cox, J.P. (1980): Theory of Stellar Pulsation (Princeton University Press, Prince-
ton)
Cox, J.P., Giuli, R.T. (1968): Principles of Stellar Structure (Gordon and Breach,
New York)
Cram, L.E., Durney, B.R., Guenther, D.B. (1983): Astrophys. J. 267, 442
Cram, L.E., Thomas, J.H., eds. (1981): The Physics of Sunspots (Sacramento Peak
Observatory, Sunspot, New Mexico)
Cranmer, S.R. (2000): Astrophys. J. 532, 1197
Cranmer, S.R., Field, G.B., Kohl, J.L. (1999 a): Astrophys. J. 518, 937
Cranmer, S.R., Kohl, J.L., Noci, G., and 27 others (1999 b): Astrophys. J. 511, 481
Curdt, W., Brekke, P., Feldman, U., Wilhelm, K., Dwivedi, B.N., Schüle, U.,
Lemaire, P. (2001): Astron. Astrophys. 375, 591
Cushman, G.W., Rense, W.A. (1976): Astrophys. J. 207, L61
Dialetis, D., Macris, C., Prokakis, T., Sarris, E. (1986): Astron. Astrophys. 168, 330
Dicke, R.H. (1974): Science 184, 419
Dicke, R.H., Goldenberg, H.M. (1967): Phys. Rev. Letters 18, 313
Dicke, R.H., Kuhn, J.R., Libbrecht, K.G. (1985): Nature 316, 687
Dicke, R.H., Kuhn, J.R., Libbrecht, K.G. (1986): Astrophys. J. 311, 1025
Diesendorf, M.O. (1970): Nature 227, 266
Dikpati, M., Charbonneau, P. (1999): Astrophys. J. 518, 508
Dilke, F.W.W., Gough, D.O. (1972): Nature 240, 262
Dodd, R.T. (1981): Meteorites (Cambridge University Press, Cambridge)
Dorfi, E. (1989): Astron. Astrophys. 225, 507
Dravins, D. (1982): Ann. Rev. Astron. Astrophys. 20, 61
Dravins, D. (1987 a): Astron. Astrophys. 172, 200
Dravins, D. (1987 b): Astron. Astrophys. 172, 211
Dravins, D., Larsson, B., Nordlund, Å. (1986): Astron. Astrophys. 158, 83
Dravins, D., Lindegren, L., Nordlund, Å. (1981): Astron. Astrophys. 96, 345
Dulk, G.A., McLean, D.J. (1978): Solar Phys. 57, 279
Dulk, G.A., McLean, D.J., Nelson, G.J. (1985): In McLean and Labrum (1985), p. 53
Duncombe, R.L., Fricke, W., Seidelmann, P.K., Wilkins, G.A. (1977): Trans. IAU
XVI B, p. 49
Dunn, R.B. (1969): Sky and Telescope 38, 368
Dunn, R.B., ed. (1981): Solar Instrumentation: What’s Next? (Sacramento Peak
Observatory, Sunspot, New Mexico)
Dunn, R.B. (1985): Solar Phys. 100, 1
Dunn, R.B., Zirker, J.B. (1973): Solar Phys. 33, 281
Durney, B.R. (1987): In Durney and Sofia (1987), p. 235
Durney, B.R., Cram, L.E., Guenther, D.B., Keil, S.L., Lytle, D.M. (1985): Astro-
phys. J. 292, 752
Durney, B.R., Pneuman, G.W. (1975): Solar Phys. 40, 461
Durney, B.R., Roxburgh, I.W. (1971): Solar Phys. 16, 3
Durney, B.R., Sofia, S., eds. (1987): The Internal Solar Angular Velocity (Reidel,
Dordrecht)
Durney, B.R., Stenflo, J.O. (1972): Astrophys. Space Sci. 15, 307
Durrant, C.J., Mattig, W., Nesis, A., Reiss, G., Schmidt, W. (1979): Solar Phys.
61, 251
Durrant, C.J., Mattig, W., Nesis, A., Schmidt, W. (1983): Astron. Astrophys.
123, 319
Duvall, T.L., Jr. (1982): Nature 300, 242
Duvall, T.L., Jr., Dziembowski, W.A., Goode, P.R., Gough, D.O., Harvey, J.W.,
Leibacher, J.W. (1984): Nature 310, 22
Duvall, T.L., Jr., Gizon, L. (2000): Solar Phys. 192, 177
Duvall, T.L., Jr., Harvey, J.W. (1983): Nature 302, 24
Duvall, T.L., Jr., Harvey, J.W. (1984): Nature 310, 19
Duvall, T.L., Jr., Harvey, J.W., Jefferies, S.M., Pomerantz, M.A. (1991): Astrophys.
J. 373, 308
Duvall, T.L., Jr., Harvey, J.W., Libbrecht, K.G., Popp, B.D., Pomerantz, M.A.
(1988): Astrophys. J. 324, 1158
Duvall, T.L., Jr., Harvey, J.W., Pomerantz, M.A. (1986): Nature 321, 500
Duvall, T.L., Jr., Jefferies, S.M., Harvey, J.W., Osaki, Y., Pomerantz, M.A. (1993 a):
Astrophys. J. 410, 829
Duvall, T.L., Jr., Jefferies, S.M., Harvey, J.W., Pomerantz, M.A. (1993 b): Nature
362, 430
Duvall, T.L., Jr., Kosovichev, A.G., Scherrer, P.H., and 10 others (1997): Solar
Phys. 170, 63
462 References
Haber, D.A., Hindman, B.W., Toomre, J., Bogart, R.S., Thompson, M.J., Hill, F.
(2000): Solar Phys. 192, 335
Haerendel, G. (1992): Nature 360, 241
Hagenaar, H.J., Schrijver, C.J., Title, A.M. (1997): Astrophys. J. 481, 988
Hagyard, M.J., ed. (1985): Measurements of Solar Vector Magnetic Fields, NASA
Conf. Publ. 2374
Hagyard, M.J., Pevtsov, A.A. (1999): Solar Phys. 189, 25
Hale, G.E. (1924): Nature 113, 105
Hale, G.E., Ellerman, F., Nicholson, S.B., Joy, A.H. (1919): Astrophys. J. 49,
153
Hale, G.E., Nicholson, S.B. (1938): Publ. Carnegie Inst. No. 498 (Washington)
Hammer, R. (1982 a): Astrophys. J. 259, 767
Hammer, R. (1982 b): Astrophys. J. 259, 779
Hammer, R. (1984): Astrophys. J. 280, 780
Hammer, R. (1987 a): In Schröter and Schüssler (1987), p. 77
Hammer, R. (1987 b): In Schröter et al. (1987), p. 255
Hammer, R., Nesis, A. (2003): In Cool Stars, Stellar Systems, and the Sun: The
Future of Cool-Star Astrophysics, ed. by A. Brown, G.M. Harper, T.R. Ayres,
p. 613, http://origins.colorado.edu/cs12/proceedings/poster/hammer.ps
Hammerschlag, R.A. (1981): In Dunn (1981), p. 583
Hampel, W. (1986): In Weak and Electromagnetic Interactions in Nuclei, ed. by
H.V. Klapdor (Springer, Berlin, Heidelberg), p. 718
Handy, B.N., Acton, L.W., Kankelborg, C.C., and 45 others (1999): Solar Phys.
187, 229
Hanle, W. (1924): Z. Physik 30, 93
Hansen, C.J., Cox, J.P., Van Horn, H.M. (1977): Astrophys. J. 217, 151
Hardorp, J. (1982): Astron. Astrophys. 105, 120
References 465
Harrison, R.A., Sawyer, E.C., Carter, M.K., and 36 others (1995): Solar Phys.
162, 233
Hart, A.B. (1956): Mon. Not. R. Astron. Soc. 116, 38
Harvey, J.W. (1973): Solar Phys. 28, 9
Harvey, J. (1977): Highlights of Astronomy 4, Part II, p. 223
Harvey, J., and the GONG Instrument Development Team (1988): In Seismology
of the Sun and Sun-like Stars, ed. by E.J. Rolfe, ESA SP-286, p. 203
Harvey, J.W., Hill, F., Hubbard, R.P., and 14 others (1996): Science 272, 1284
Harvey, K.L., ed. (1992 a): The Solar Cycle, ASP Conf. Ser. 27
Harvey, K.L. (1992 b): In Harvey (1992 a), p. 335
Hassler, D.M., Dammasch, I.E., Lemaire, P., Brekke, P., Curdt, W., Mason, H.E.,
Vial, J.-C., Wilhelm, K. (1999): Science 283, 810
Hathaway, D.H. (1987): Solar Phys. 108, 1
Haug, E., Elwert, G. (1985): Solar Phys. 99, 219
Heasley, J.N., Milkey, R.W. (1978): Astrophys. J. 221, 677
Heath, D.F., Thekaekara, M.P. (1977): In White (1977), p. 193
Heinzel, P., Anzer, U. (2001): Astron. Astrophys. 375, 1082
Henning, H.M., Scherrer, P.H. (1986): In Gough (1986), p. 55
Hénoux, J.-C. (1969): Astron. Astrophys. 2, 288
Hénoux, J.-C. (1998): Space Sci. Rev. 85, 215
Henyey, L.G., Wilets, L., Böhm, K.H., LeLevier, R., Levee, R.D. (1959): Astrophys.
J. 129, 628
Herbold, G., Ulmschneider, P., Spruit, H.C., Rosner, R. (1985): Astron. Astrophys.
145, 157
Hess, W.N., ed. (1964): The Physics of Solar Flares, NASA Sp-50 (NASA, Washington)
Heyvaerts, J., Priest, E.R. (1983): Astron. Astrophys. 117, 220
Heyvaerts, J., Priest, E.R., Rust, D.M. (1977): Astrophys. J. 216, 123
Hill, F. (1988): Astrophys. J. 333, 996
Hill, F. (1989): Astrophys. J. 343, L69
Hill, H.A., Dziembowski, W.A., eds. (1980): Nonradial and Nonlinear Stellar Pul-
sation, Lect. Notes Phys., Vol. 125 (Springer, Berlin, Heidelberg)
Hill, H.A., Stebbins, R.T. (1975): Astrophys. J. 200, 471
Hillebrandt, W., Meyer-Hofmeister, E., Thomas, H.-C., eds. (1987): Physical Pro-
cesses in Comets, Stars, and Active Galaxies (Springer, Berlin, Heidelberg)
Hirzberger, J., Bonet, J.A., Vázquez, M., Hanslmeier, A. (1999 a): Astrophys. J.
515, 441
Hirzberger, J., Bonet, J.A., Vázquez, M., Hanslmeier, A. (1999 b): Astrophys. J.
527, 405
Hirzberger, J., Vázquez, M., Bonet, J.A., Hanslmeier, A., Sobotka, M. (1997): As-
trophys. J. 480, 406
Hoeksema, J.T., Domingo, V., Fleck, B., Battrick, B., eds. (1995): Helioseismology,
ESA SP-376, Vol. I, II
Hoekzema, N.M., Brandt, P.N. (2000): Astron. Astrophys. 353, 389
Hoekzema, N.M., Brandt, P.N., Rutten, R.J. (1998): Astron. Astrophys. 333, 322
Hoffmeister, C. (1943): Z. Astrophys. 22, 265
Hollweg, J.V. (1982): Astrophys. J. 257, 345
Holst, G.C. (1998): CCD Arrays, Cameras, and Displays, 2nd ed. (SPIE Optical
Engineering Press, Bellingham)
Holweger, H. (1967): Z. Astrophys. 65, 365
Holweger, H., Livingston, W., Steenbock, W. (1983): Nature 302, 125
Horn, T., Staude, J., Landgraf, V. (1997): Solar Phys. 172, 69
Houdek, G., Balmforth, N.J., Christensen-Dalsgaard, J., Gough, D.O. (1999): As-
tron. Astrophys. 351, 582
466 References
Kohl, J.L., Noci, G., Antonucci, E., and 23 others (1997): Solar Phys. 175, 613
Kohl, J.L., Noci, G., Antonucci, E., and 27 others (1998): Astrophys. J. 501, L127
Komm, R.W., Howard, R.F., Harvey, J.W. (1993 a): Solar Phys. 143, 19
Komm, R.W., Howard, R.F., Harvey, J.W. (1993 b): Solar Phys. 147, 207
Komm, R., Mattig, W., Nesis, A. (1991 a): Astron. Astrophys. 243, 251
Komm, R., Mattig, W., Nesis, A. (1991 b): Astron. Astrophys. 252, 827
Kopecký, M. (1957): Bull. Astron. Inst. Czech. 8, 71
Kopecký, M. (1966): Bull. Astron. Inst. Czech. 17, 270
Kopecký, M., Soytürk, E. (1971): Bull. Astron. Inst. Czech. 22, 154
Kopp, R.A. (1977): In Zirker (1977), p. 179
Kopp, R.A., Pneuman, G.W. (1976): Solar Phys. 50, 85
Korzennik, S., Wilson, A., eds. (1998): Structure and Dynamics of the Interior of
the Sun and Sun-like Stars, ESA SP-418, Vol. I, II
Kosovichev, A.G. (1996): Astrophys. J. 461, L55
Kosovichev, A.G., Duvall, T.L., Jr., Scherrer, P.H. (2000): Solar Phys. 192, 159
Kosovichev, A.G., Schou, J., Scherrer, P.H., and 31 others (1997): Solar Phys.
170, 43
Kosovichev, A.G., Zharkova, V.V. (1998): Nature 393, 317
Kotov, V.A. (1985): Solar Phys. 100, 101
Kotov, V.A., Haneychuk, V.I., Tsap, T.T., Hoeksema, J.T. (1997): Solar Phys.
176, 45
Koutchmy, S., Lebecq, C. (1986): Astron. Astrophys. 169, 323
Kovitya, P., Cram, L. (1983): Solar Phys. 84, 45
Krafft, M. (1968): Solar Phys. 5, 462
Krause, F. (1967): Habilitationsschrift, Universität Jena
Krause, F., Rädler, K.-H. (1980): Mean-Field Magnetohydrodynamics and Dynamo
Theory (Akademie-Verlag, Berlin)
Krause, F., Rädler, K.-H., Rüdiger, G., eds. (1993): The Cosmic Dynamo, IAU
Symp. 157 (Kluwer, Dordrecht)
Krause, F., Rüdiger, G. (1975): Solar Phys. 42, 107
Krieg, J., Wunnenberg, M., Kneer, F., Koschinsky, M., Ritter, C. (1999): Astron.
Astrophys. 343, 983
Krüger, A. (1979): Introduction to Solar Radio Astronomy and Radio Physics (Rei-
del, Dordrecht)
Kubičela, A. (1973): In Solar Activity and Related Interplanetary and Terrestrial
Phenomena, ed. by J. Xanthakis (Springer, Berlin, Heidelberg), p. 123
Kubičela, A. (1976): Solar Phys. 47, 551
Kudoh, T., Shibata, K. (1999): Astrophys. J. 514, 493
Kuhn, J.R., Bush, R.I., Scheick, X., Scherrer, P. (1998): Nature 392, 155
Kuhn, J.R., Penn, M.J., eds. (1995): Infrared Tools for Solar Astrophysics: What’s
Next? (World Scientific, Singapore)
Kumar, P. (1994): Astrophys. J. 428, 827
Kumar, P., Lu, E. (1991): Astrophys. J. 375, L35
Kundu, M.R. (1965): Solar Radio Astronomy (Interscience, New York)
Kuperus, M., Raadu, M.A. (1974): Astron. Astrophys. 31, 189
Kupke, R., LaBonte, B.J., Mickey, D.L. (2000): Solar Phys. 191, 97
Kurucz, R.L. (1979): Astrophys. J. Suppl. 40, 1
Küveler, G. (1983): Solar Phys. 88, 13
Laclare, F., Delmas, C., Coin, J.P., Irbah, A. (1996): Solar Phys. 166, 211
Lamers, H.J.G.L.M., Cassinelli, J.P. (1999): Introduction to Stellar Winds (Cam-
bridge University Press, Cambridge)
Landau, L.D., Lifschitz, E.M. (1966 a): Statistische Physik (Akademie-Verlag, Berlin)
Landau, L.D., Lifschitz, E.M. (1966 b): Hydrodynamik (Akademie-Verlag, Berlin)
Landau, L.D., Lifschitz, E.M. (1967): Elektrodynamik der Kontinua (Akademie-
Verlag, Berlin)
Landi degl’Innocenti, E., Landi degl’Innocenti, M. (1972): Solar Phys. 27, 319
Landolfi, M., Landi degl’Innocenti, E. (1982): Solar Phys. 78, 355
Landolfi, M., Landi degl’Innocenti, E. (1986): Astron. Astrophys. 167, 200
Landolfi, M., Landi degl’Innocenti, E., Arena, P. (1984): Solar Phys. 93, 269
Landolt-Börnstein (1981): New Series, Group VI, 2a (Springer, Berlin, Heidelberg)
Lang, K.R. (1999): Astrophysical Formulae, 3rd ed., Vols. I, II (Springer, Berlin,
Heidelberg)
Lang, K.R. (2000): The Sun from Space (Springer, Berlin, Heidelberg)
Langhans, K., Schmidt, W., Rimmele, T., Sigwarth, M. (2001): In Sigwarth (2001 a),
p. 439
Lantz, S.R., Fan, Y. (1999): Astrophys. J. Suppl. 121, 247
Lapwood, E.R., Usami, T. (1981): Free Oscillations of the Earth (Cambridge Uni-
versity Press, Cambridge)
Lavely, E.M., Ritzwoller, M.H. (1992): Phil. Trans. R. Soc. London A 339, 431
Lavely, E.M., Ritzwoller, M.H. (1993): Astrophys. J. 403, 810
Lazrek, M., Pantel, A., Fossat, E., and 10 others (1996): Solar Phys. 166, 1
Lean, J. (1997): Ann. Rev. Astron. Astrophys. 35, 33
Lebovitz, N.R. (1966): Astrophys. J. 146, 946
Lebreton, Y., Maeder, A. (1987): Astron. Astrophys. 175, 99
Ledoux, P. (1962): Bull. Acad. Roy. Belg., Cl. Sc., 5me série, 48, 240
Ledoux, P., Walraven, T. (1958): In Handbuch der Physik, ed. by S. Flügge
(Springer, Berlin, Heidelberg), Vol. LI, p. 353
Lehnert, B., ed. (1958): Electromagnetic Phenomena in Cosmical Physics, IAU
Symp. 6 (Cambridge University Press, Cambridge)
Lehnert, B. (1970): Cosmic Electrodynamics 1, 397
Leibacher, J.W., Noyes, R.W., Toomre, J., Ulrich, R.K. (1985): Scientific American
253, No. 3, p. 34
Leighton, R.B. (1969): Astrophys. J. 156, 1
Leighton, R.B., Noyes, R.W., Simon, G.W. (1962): Astrophys. J. 135, 474
Leka, K.D., Steiner, O. (2001): Astrophys. J. 552, 354
Lenz, D.D., DeLuca, E.E., Golub, L., Rosner, R., Bookbinder, J.A. (1999): Astro-
phys. J. 517, L155
Leroy, J.L. (1988): In Solar and Stellar Coronal Structure and Dynamics, ed. by
R.C. Altrock (Sacramento Peak Observatory, Sunspot, New Mexico), p. 422
Leroy, J.L., Bommier, V., Sahal-Bréchot, S. (1984): Astron. Astrophys. 131, 33
LeVeque, R.J., Mihalas, D., Dorfi, E.A., Müller, E. (1998): Computational Methods
for Astrophysical Fluid Flow, ed. by O. Steiner and A. Gautschy (Springer,
Berlin, Heidelberg)
Levine, R.H., Schulz, M., Frazier, E.N. (1982): Solar Phys. 77, 363
Libbrecht, K.G. (1988): Astrophys. J. 334, 510
Libbrecht, K.G., Kaufman, J.M. (1988): Astrophys. J. 324, 1172
Libbrecht, K.G., Woodard, M.F., Kaufman, J.M. (1990): Astrophys. J. Suppl. 74,
1129
Libbrecht, K.G., Zirin, H. (1986): Astrophys. J. 308, 413
Liggett, M., Zirin, H. (1985): Solar Phys. 97, 51
Lighthill, M.J. (1952): Proc. R. Soc. London A 211, 564
470 References
MacQueen, R.M., Sime, D.G., Picat, J.-P. (1983): Solar Phys. 83, 103
Macris, C.J., Rösch, J. (1983): Comptes Rendus 296, 265
Magain, P. (1986): Astron. Astrophys. 163, 135
Magara, T., Shibata, K. (1999): Astrophys. J. 514, 456
Maier, E., Twigg, L.W., Sofia, S. (1992): Astrophys. J. 389, 447
Malherbe, J.M., Priest, E.R. (1983): Astron. Astrophys. 123, 80
Maltby, P., Avrett, E.H., Carlsson, M., Kjeldseth-Moe, O., Kurucz, R.L., Loeser,
R. (1986): Astrophys. J. 306, 284
Mankin, W.G. (1977): In White (1977), p. 151
Manson, J.E. (1977): In White (1977), p. 261
References 471
Mariska, J.T. (1992): The Solar Transition Region (Cambridge University Press,
Cambridge)
Márquez, I., Bonet, J.A., Vázquez, M., Wöhl, H. (1996): Astron. Astrophys.
305, 316
Martin, S.F., Harvey, K.L. (1979): Solar Phys. 64, 93
Martı́nez Pillet, V. (2000): Astron. Astrophys. 361, 734
Martı́nez Pillet, V., Collados, M., Sánchez Almeida, J., and 8 others (1999): In
Rimmele et al. (1999), p. 264
Martı́nez Pillet, V., Moreno-Insertis, F., Vázquez, M. (1993): Astron. Astrophys.
274, 521
Mathys, G., Stenflo, J.O. (1987): Astron. Astrophys. Suppl. 67, 557
Mattig, W. (1966): In Atti del Conv. s. Macchie Sol., ed. by G. Barbéra (Firenze),
p. 194
Mattig, W. (1980): Astron. Astrophys. 83, 129
Mattig, W. (1983): Solar Phys. 87, 187
Mattig, W., Mehltretter, J.P., Nesis, A. (1981): Astron. Astrophys. 96, 96
Mauas, P.J., Avrett, E.H., Loeser, R. (1990): Astrophys. J. 357, 279
Maunder, W. (1922): Mon. Not. R. Astron. Soc. 82, 534
McIntosh, P.S. (1979): Annotated Atlas of Hα Synoptic Charts, Rep. UAG-70
(World Data Center A, Boulder)
McIntosh, P.S. (1981): In Cram and Thomas (1981), p. 7
McIntosh, P.S. (1990): Solar Phys. 125, 251
McLean, D.J., Labrum, N.R., eds. (1985): Solar Radiophysics (Cambridge Univer-
sity Press, Cambridge)
Mehltretter, J.P. (1973): Solar Phys. 30, 19
Mehltretter, J.P. (1974): Solar Phys. 38, 43
Mehltretter, J.P. (1978): Astron. Astrophys. 62, 311
Mein, P. (1971): Solar Phys. 20, 3
Meister, C.-V., Staude, J., Pregla, A.V. (1999): Astron. Nachr. 320, 43
Meyer, F. (1968): In Kiepenheuer (1968), p. 485
Meyer, F., Schmidt, H.U. (1968): Z. Angew. Math. Mech. 48, T218
Meyer, F., Schmidt, H.U., Simon, G.W., Weiss, N.O. (1979): Astron. Astrophys.
76, 35
Meyer, F., Schmidt, H.U., Weiss, N.O. (1977): Mon. Not. R. Astron. Soc. 179, 741
Meyer, F., Schmidt, H.U., Weiss, N.O., Wilson, P.R. (1974): Mon. Not. R. Astron.
Soc. 169, 35
Meyer, J.-P. (1985): Astrophys. J. Suppl. 57, 151
Michalitsanos, A.G., Kupferman, P. (1974): Solar Phys. 36, 403
Michaud, G., Charbonneau, P. (1991): Space Sci. Rev. 57, 1
Mickey, D.L., Canfield, R.C., LaBonte, B.J., Leka, K.D., Waterson, M.F., Weber,
H.M. (1996): Solar Phys. 168, 229
Miesch, M.S. (2000): Solar Phys. 192, 59
Miesch, M.S., Elliott, J.R., Toomre, J., Clune, T.L., Glatzmaier, G.A., Gilman,
P.A. (2000): Astrophys. J. 532, 593
Mihalas, B.W., Toomre, J. (1981): Astrophys. J. 249, 349
Mihalas, B.W., Toomre, J. (1982): Astrophys. J. 263, 386
Mihalas, D. (1978): Stellar Atmospheres, 2nd ed. (Freeman, San Francisco)
Mihalas, D., Auer, L.H., Mihalas, B.R. (1978): Astrophys. J. 220, 1001
Mihalas, D., Däppen, W., Hummer, D.G. (1988): Astrophys. J. 331, 815
Mihalas, D., Hummer, D.G., Mihalas, B.W., Däppen, W. (1990): Astrophys. J. 350,
300
Mihalas, D., Mihalas, B.W. (1984): Foundations of Radiation Hydrodynamics (Ox-
ford University Press, New York)
472 References
Norton, A.A., Ulrich, R.K., Bush, R.L., Tarbell, T.D. (1999): Astrophys. J. 518, L123
November, L.J. (1986): Applied Optics 25, 392
November, L.J., ed. (1991): Solar Polarimetry (Sacramento Peak Observatory,
Sunspot, New Mexico)
November, L.J., Stauffer, F.R. (1984): Applied Optics 23, 2333
November, L.J., Toomre, J., Gebbie, K.B. (1979): Astrophys. J. 227, 600
November, L.J., Toomre, J., Gebbie, K.B. (1982): Astrophys. J. 258, 846
November, L.J., Toomre, J., Gebbie, K.B., Simon, G.W. (1981): Astrophys. J. 245,
L123
Noyes, R.W., Rhodes, E.J., Jr., eds. (1984): Probing the Depths of a Star: The
Study of Solar Oscillations from Space, NASA/JPL 400-237
Núñez, M., Ferriz-Mas, A., eds. (1999): Stellar Dynamos: Nonlinearity and Chaotic
Flows, ASP Conf. Ser. 178
Nye, A.H., Thomas, J.H. (1974): Solar Phys. 38, 399
Querfeld, C.W., Smartt, R.N., Bommier, V., Landi degl’-Innocenti, E., House, L.L.
(1985): Solar Phys. 96, 277
Rachkovsky, D.N. (1962 a): Izv. Krym. Astrofiz. Obs. 27, 148
Rachkovsky, D.N. (1962 b): Izv. Krym. Astrofiz. Obs. 28, 259
Radick, R.R., ed. (1993): Real Time and Post Facto Solar Image Correction (Sacra-
mento Peak Observatory, Sunspot, New Mexico)
Rädler, K.-H. (1990): In Inside the Sun, ed. by G. Berthomieu, M. Cribier (Kluwer,
Dordrecht), p. 385
Ramsay, J.V., Kobler, H., Mugridge, E.G.V. (1970): Solar Phys. 12, 492
References 475
Rozelot, J.-P., Klein, L., Vial, J.-C., eds. (2000): Transport and Energy Conversion
in the Heliosphere (Springer, Berlin, Heidelberg)
Rozelot, J.-P., Rösch, J. (1997): Solar Phys. 172, 11
Rucklidge, A.M., Weiss, N.O., Brownjohn, D.P., Matthews, P.C., Proctor, M.R.E.
(2000): J. Fluid Mech. 419, 283
Rüdiger, G. (1980): Geophys. Astrophys. Fluid Dyn. 16, 239
Rüdiger, G. (1983): Geophys. Astrophys. Fluid Dyn. 25, 213
Rüdiger, G., Hasler, K.-H., Kitchatinov, L.L. (1997): Astron. Nachr. 318, 173
Rüdiger, G., Kitchatinov, L.L. (1996): Astrophys. J. 466, 1078
Rüdiger, G., von Rekowski, B., Donahue, R.A., Baliunas, S.L. (1998): Astrophys.
J. 494, 691
Rüdiger, G., Tuominen, I. (1987): In Durney and Sofia (1987), p. 361
Rüedi, I., Solanki, S.K., Stenflo, J.O., Tarbell, T., Scherrer, P.H. (1998): Astron.
Astrophys. 335, L97
Ruiz Cobo, B., del Toro Iniesta, J.C. (1992): Astrophys. J. 398, 375
Ruiz Cobo, B., del Toro Iniesta, J.C., Rodrı́guez Hidalgo, I., Collados, M., Sánchez
Almeida, J. (1996): In Cool Stars, Stellar Systems, and the Sun, ed. by R.
Pallavicini, A.K. Dupree, ASP Conf. Ser. 109, p. 155
Rutten, R.J., Milkey, R.W. (1979): Astrophys. J. 231, 277
Rutten, R.J., Schrijver, C.J., eds. (1994): Solar Surface Magnetism, NATO ASI Ser.
C, 433 (Kluwer, Dordrecht)
Ruzmaikin, A.A., Sokoloff, D.D., Starchenko, S.V. (1988): Solar Phys. 115, 5
Sackmann, I.-J., Boothroyd, A.I., Fowler, W.A. (1990): Astrophys. J. 360, 727
Sackmann, I.-J., Boothroyd, A.I., Kraemer, K.E. (1993): Astrophys. J. 418, 457
Sakao, T., Tsuneta, S., Hara, H., Shimizu, T., Kano, R., Kumagai, K., Yoshida, T.,
Nagata, S., Kobayashi, K. (1999): Solar Phys. 187, 303
Sakurai, T. (1979): Publ. Astron. Soc. Japan 31, 209
Sakurai, T. (1981): Solar Phys. 69, 343
Sakurai, T. (1985): Astron. Astrophys. 152, 121
Salpeter, E.E. (1954): Australian J. Phys. 7, 373
Sánchez Almeida, J., Landi degl’Innocenti, E., Martı́nez Pillet, V., Lites, B.W.
(1996): Astrophys. J. 466, 437
Schatzman, E. (1959): In The Hertzsprung–Russell Diagram, ed. by J.L. Greenstein,
IAU Symp. 10 (Mount Wilson and Palomar Observatories, Pasadena), p. 129
Schatzman, E., Maeder, A., Angrand, F., Glowinski, R. (1981): Astron. Astrophys.
96, 1
Scherrer, P.H., Bogart, R.S., Bush, R.I., 9 others, and the MDI Engineering Team
(1995): Solar Phys. 162, 129
Scherrer, P.H., Bogart, R., Hoeksema, J.T., Yoshimura, H. (1986): In Gough (1986),
p. 93
Scherrer, P.H., Wilcox, J.M., Christensen-Dalsgaard, J., Gough, D.O. (1983): Solar
Phys. 82, 75
Schindler, K., Birn, J., Janicke, L. (1983): Solar Phys. 87, 103
Schlattl, H., Weiss, A. (1999): Astron. Astrophys. 347, 272
Schlattl, H., Weiss, A., Ludwig, H.G. (1997): Astron. Astrophys. 322, 646
Schleicher, H. (1976): Dissertation, University of Göttingen
Schlichenmaier, R., Collados, M. (2002): Astron. Astrophys. 381, 668
Schlichenmaier, R., Jahn, K., Schmidt, H.U. (1998 a): Astrophys. J. 493, L121
Schlichenmaier, R., Jahn, K., Schmidt, H.U. (1998 b): Astron. Astrophys. 337, 897
Schlichenmaier, R., Schmidt, W. (1999): Astron. Astrophys. 349, L37
Schlichenmaier, R., Schmidt, W. (2000): Astron. Astrophys. 358, 1122
Schlichenmaier, R., Stix, M. (1995): Astron. Astrophys. 302, 264
References 477
Seaton, M.J., ed. (1995): The Opacity Project, Vol. 1 (Institute of Physics, Bristol
and Philadelphia)
Seehafer, N. (1994): Astron. Astrophys. 284, 593
Seehafer, N. (1996): Phys. Rev. E 53, 1283
Severino, G., Roberti, G., Marmolino, C., Gomez, M.T. (1986): Solar Phys. 104,
259
Severny, A.B. (1965): In Lüst (1965), p. 358
Seykora, E.J. (1993): Solar Phys. 145, 389
Shack, R.V., Platt, B.C. (1971): J. Opt. Soc. America 61, 656
Shaviv, G., Salpeter, E.E. (1973): Astrophys. J. 184, 191
Sheeley, N.R., Jr. (1964): Astrophys. J. 140, 731
Shelke, R.N., Pande, M.C. (1985): Solar Phys. 95, 193
Shibahashi, H., Noels, A., Gabriel, M. (1983): Astron. Astrophys. 123, 283
Shibahashi, H., Osaki, Y. (1981): Publ. Astron. Soc. Japan 33, 713
Shibahashi, H., Osaki, Y., Unno. W. (1975): Publ. Astron. Soc. Japan 27, 401
Shimabukuro, F.I. (1977): In White (1977), p. 133
Shine, R.A., Simon, G.W., Hurlburt, N.E. (2000): Solar Phys. 193, 313
Shklovskii, I.S. (1965): Physics of the Solar Corona, English Translation of 2nd ed.
(Pergamon Press, Oxford)
Shurcliff, W.A. (1962): Polarized Light (Harvard University Press, Cambridge, Ma.)
Sigwarth, M. (1999): Dissertation, Universität Freiburg
Sigwarth, M., ed. (2001 a): Advanced Solar Polarimetry: Theory, Observation, and
Instrumentation, ASP Conf. Ser. 236
Sigwarth, M. (2001 b): Astrophys. J. 563, 1031
Sigwarth, M., Balasubramaniam, K.S., Knölker, M., Schmidt, W. (1999): Astron.
Astrophys. 349, 941
Simon, G.W., Brandt, P.N., November, L.J., Scharmer, G.B., Shine, R.A. (1994):
In Rutten and Schrijver (1994), p. 261
Simon, G.W., Leighton, R.B. (1964): Astrophys. J. 140, 1120
Simon, G.W., Weiss, N.O. (1968): Z. Astrophys. 69, 435
Singh, H.P., Roxburgh, I.W., Chan, K.W. (1994): Astron. Astrophys. 281, L73
Singh, H.P., Roxburgh, I.W., Chan, K.W. (1995): Astron. Astrophys. 295, 703
Singh, H.P., Roxburgh, I.W., Chan, K.W. (1998): Astron. Astrophys. 340, 178
Skaley, D., Stix, M. (1991): Astron. Astrophys. 241, 227
Skumanich, A. (1972): Astrophys. J. 171, 565
Skumanich, A., Lean, J.L., White, O.R., Livingston, W.C. (1984): Astrophys. J.
282, 776
Skumanich, A., Lites, B.W. (1987): Astrophys. J. 322, 473
Skumanich, A., Smythe, C., Frazier, E.N. (1975): Astrophys. J. 200, 747
Slettebak, A. (1970): In Stellar Rotation, ed. by A. Slettebak (Reidel, Dordrecht),
p. 3
Snodgrass, H.B. (1984): Solar Phys. 94, 13
Snodgrass, H.B. (1985): Astrophys. J. 291, 339
Snodgrass, H.B., Kress, J.M., Wilson, P.R. (2000): Solar Phys. 191, 1
Snodgrass, H.B., Ulrich, R.K. (1990): Astrophys. J. 351, 309
Sobotka, M. (1999): In Motions in the Solar Atmosphere, ed. by A. Hanslmeier and
M. Messerotti (Kluwer, Dordrecht), p. 71
Sobotka, M., Brandt, P.N., Simon, G.W. (1997 a): Astron. Astrophys. 328, 682
Sobotka, M., Brandt, P.N., Simon, G.W. (1997 b): Astron. Astrophys. 328, 689
Sobouti, Y. (1980): Astron. Astrophys. 89, 314
Sofia, S., Heaps, W., Twigg, L.W. (1994): Astrophys. J. 427, 1048
Solanki, S.K. (1986): Astron. Astrophys. 168, 311
Solanki, S.K. (1993): Space Sci. Rev. 63, 1
References 479
Soltau, D., Schröter, E.H., Wöhl, H. (1976): Astron. Astrophys. 50, 367
Souffrin, P. (1966): Ann. d’ Astrophys. 29, 55
Soward, A.M., Priest, E.R. (1982): J. Plasma Phys. 28, 335
Spiegel, E.A. (1957): Astrophys. J. 126, 202
Spiegel, E.A., Zahn, J.P., eds. (1977): Problems of Stellar Convection, Lect. Notes
Phys., Vol. 71 (Springer, Berlin, Heidelberg)
Spiegel, E.A., Zahn, J.-P. (1992): Astron. Astrophys. 265, 106
Spitzer, L., Jr. (1962): Physics of Fully Ionized Gases, 2nd ed. (Interscience, New
York)
Spruit, H.C. (1976): Solar Phys. 50, 269
Spruit, H.C. (1981 a): Astron. Astrophys. 98, 155
Spruit, H.C. (1981 b): In Jordan (1981), p. 385
Spruit, H.C. (1981 c): In Cram and Thomas (1981), p. 480
Spruit, H.C. (1987): In Durney and Sofia (1987), p. 185
Spruit, H.C. (1997): Mem. Soc. Astron. Ital. 68, 397
Spruit, H.C., van Ballegooijen, A.A. (1982): Astron. Astrophys. 106, 58
Spruit, H.C., Nordlund, Å., Title, A.M. (1990): Ann. Rev. Astron. Astrophys.
28, 263
Spruit, H.C., Zweibel, E.G. (1979): Solar Phys. 62, 15
Staiger, J. (1987): Astron. Astrophys. 175, 263
Stanchfield II, D.C.H., Thomas, J.H., Lites, B.W. (1997): Astrophys. J. 477, 485
Stark, D., Wöhl, H. (1981): Astron. Astrophys. 93, 241
Staude, J. (1972): Solar Phys. 24, 255
Staude, J. (1999): In Schmieder et al. (1999), p. 113
Steenbeck, M., Krause, F. (1969): Astron. Nachr. 291, 49
Steenbeck, M., Krause, F., Rädler, K.-H. (1966): Z. Naturforsch. 21a, 361
Steffen, M., Ludwig, H.-G., Krüß, A. (1989): Astron. Astrophys. 213, 371
Steiger, R. von, Schwadron, N.A., Fisk, L.A., Geiss, J., Gloeckler, G., Hefti, S.,
Wilken, B., Wimmer-Schweingruber, R.F., Zurbuchen, T.H. (2000): J. Geophys.
Res. 105 27217
Stein, R.F., Nordlund, Å. (1998 a): Astrophys. J. 499, 914
Stein, R.F., Nordlund, Å. (1998 b): In Korzennik and Wilson (1998), p. 693
Steiner, O. (2000): Solar Phys. 196, 245
Steiner, O., Hauschildt, P.H., Bruls, J. (2001): Astron. Astrophys. 372, L13
Stenflo, J.O. (1971): In Howard (1971), p. 101
Stenflo, J.O. (1976): In Bumba and Kleczek (1976), p. 69
Stenflo, J.O. (1982): Solar Phys. 80, 209
Stenflo, J.O., ed. (1983): Solar and Stellar Magnetic Fields: Origins and Coronal
Effects, IAU Symp. 102 (Reidel, Dordrecht)
Stenflo, J.O. (1986): Mitt. Astron. Ges. 65, 25
Stenflo, J.O. (1994): Solar Magnetic Fields (Kluwer, Dordrecht)
Stenflo, J.O. (2001): In Sigwarth (2001 a), p. 97
Stenflo, J.O., Gandorfer, A., Keller, C.U. (2000 a): Astron, Astrophys. 355, 781
Stenflo, J.O., Harvey, J.W., Brault, J.W., Solanki, S. (1984): Astron. Astrophys.
131, 333
Stenflo, J.O., Keller, C.U. (1997): Astron, Astrophys. 321, 927
Stenflo, J.O., Keller, C.U., Gandorfer, A. (2000 b): Astron, Astrophys. 355, 789
Stenflo, J.O., Keller, C.U., Povel, H.P. (1992): LEST Foundation, Technical Rep.
No. 54
Stenflo, J.O., Nagendra, K.N., eds. (1996): Solar Phys. 164
Stepanov, V.E. (1958): Izv. Krym. Astrofiz. Obs. 18, 136
Sterling, A.C. (2000): Solar Phys. 196, 79
Stewart, J.C., Pyatt, K.D., Jr. (1966): Astrophys. J. 144, 1203
480 References
Timothy, A.F., Krieger, A.S., Vaiana, G.S. (1975): Solar Phys. 42, 135
Title, A.M., Frank, Z.A., Shine, R.A., Tarbell, T.D., Topka, K.P., Scharmer, G.,
Schmidt, W. (1993): Astrophys. J. 403, 780
Title, A.M., Ramsey, H.E. (1980): Applied Optics 19, 2046
Title, A.M., Tarbell, T.D., Simon, G.W., and the SOUP Team (1986): Adv. Space
Res. 6, No. 8, p. 253
Title, A.M., Tarbell, T.D., Topka, K.P., Shine, R.A., Simon, G.W., Zirin, H., and
the SOUP Team (1987): In Schröter and Schüssler (1987), p. 173
Tobias, S.M. (1997): Astron. Astrophys. 322, 1007
Tolansky, S. (1948): Multiple-Beam Interferometry of Surfaces and Films (Claren-
don Press, Oxford)
Tomczyk, S., Schou, J., Thompson, M.J. (1995 b): Astrophys. J. 448, L57
Tomczyk, S., Streander, K., Card, G., Elmore, D., Hull, H., Cacciani, A. (1995 a):
Solar Phys. 159, 1
Tripathy, S.C., Christensen-Dalsgaard, J. (1998): Astron. Astrophys. 337, 579
Tritschler, A., Schmidt, W. (2002): Astron. Astrophys. 382, 1093
Trujillo-Bueno, J., Moreno-Insertis, F., Sánchez, F., eds. (2002): Astrophysical Spec-
tropolarimetry (Cambridge University Press, Cambridge)
Tsinganos, K.C., ed. (1996): Solar and Astrophysical Magnetohydrodynamic Flows,
NATO ASI Ser. C, 481 (Kluwer, Dordrecht)
Tsuneta, S. (1996 a): Astrophys. J. 456, L63
Tsuneta, S. (1996 b): Astrophys. J. 456, 840
Tsuneta, S. (1997): Astrophys. J. 483, 507
Tsuneta, S., Acton, L., Bruner, M., and 10 others (1991): Solar Phys. 136, 37
Tsuneta, S., Hara, H., Shimizu, T., Acton, L.W., Strong, K.T., Hudson, H.S.,
Ogawara, Y. (1992): Publ. Astron. Soc. Japan 44, L63
Tu, C.-Y., Marsch, E., Wilhelm, K., Curdt, W. (1998): Astrophys. J. 503, 475
Tuominen, J. (1941): Z. Astrophys. 21, 96
Turck-Chièze, S., Lopes, I. (1993): Astrophys. J. 408, 347
Vaiana, G.S., Davis, J.M., Giacconi, R., Krieger, A.S., Silk, J.K., Timothy, A.F.,
Zombeck, M. (1973): Astrophys. J. 185, L47
Vandakurov, Yu.V. (1967): Astron. Zh. 44, 786, English Translation: Sov. Astron.
11, 630
Vaughan, A.H., Baliunas, S.L., Middelkoop, F., Hartmann, L.W., Mihalas, D.,
Noyes, R.W., Preston, G.W. (1981): Astrophys. J. 250, 276
Vaughan, J.M. (1989): The Fabry–Perot Interferometer (Adam Hilger, Bristol)
Vernazza, J.E., Avrett, E.H., Loeser, R. (1973): Astrophys. J. 184, 605
Vernazza, J.E., Avrett, E.H., Loeser, R. (1976): Astrophys. J. Suppl. 30, 1
Vernazza, J.E., Avrett, E.H., Loeser, R. (1981): Astrophys. J. Suppl. 45, 635
Vitense, E. (1953): Z. Astrophys. 32, 135
Voigt, H.-H. (1956): Z. Astrophys. 40, 157
Vrabec, D. (1971): In Howard (1971), p. 329
Wilhelm, K., Curdt, W., Marsch, E., and 13 others (1995): Solar Phys. 162, 189
Willson, R.C., Gulkis, S., Janssen, M., Hudson, H.S., Chapman, G. A. (1981):
Science 211, 700
Willson, R.C., Hudson, H.S. (1981): Astrophys. J. 244, L 185
Willson, R.C., Hudson, H.S., Fröhlich, C., Brusa, R.W. (1986): Science 234, 1114
Wilson, A., ed. (2001): Helio- and Asteroseismology at the Dawn of the Millennium,
ESA SP-464
Wilson, O.C. (1978): Astrophys. J. 226, 379
Winkler, K.-H., Newman, M.J. (1980): Astrophys. J. 236, 201
Wissink, J.G., Hughes, D.W., Matthews, P.C., Proctor, M.R.E. (2000 a): Mon. Not.
R. Astron. Soc. 318, 501
Wissink, J.G., Matthews, P.C., Hughes, D.W., Proctor, M.R.E. (2000 b): Astrophys.
J. 536, 982
Withbroe, G.L., Noyes, R.W. (1977): Ann. Rev. Astron. Astrophys. 15, 363
Wittmann, A. (1974 a): Solar Phys. 35, 11
Wittmann, A. (1974 b): Solar Phys. 36, 65
Wittmann, A.D. (1997): Solar Phys. 171, 231
Wittmann, A., Schröter, E.H. (1969): Solar Phys. 10, 357
Wöhl, H. (1971): Solar Phys. 16, 362
Wöhl, H. (1978): Astron. Astrophys. 62, 165
Wöhl, H., Brajša, R. (2001): Solar Phys. 198, 57
Woodard, M., Hudson, H.S. (1983): Nature 305, 589
Worden, S.P. (1975): Solar Phys. 45, 521