INTERNATIONAL SERIES IN PHYSICS
F. K. RICHTMYER, CONSULTING EDITOR
Condon and Morse-
Qu ANTUM MECHANICS
Ruark and Urey-
ATOMS, MOLECULES AND QUANTA
Pauling and Goudsmit-
THE STRUCTURE OF LINE 8PEC1RA
INTRODUCTION TO MODERN PHYSICS
INTRODUCTION TO
MODERN PHYSICS
BY
F. K. RIGHTMYER
Professor of Physics at Cornell University
FIRST EDITION
FOURTH IMPRESSION
McGRAW-HILL BOOK COMPANY, INc.
NEW YORK AND LONDON
1928
COPYRIGHT, 1928, BY THE
McGRAW-HILL BooK CoMPANY, INc.
PRINTED IN THE UNITED STATES OF AMERICA
THE MAPLE PRESS COMPANY, YORK, PA.
To My Wife
PREFACE
For several years, the author has given at Cornell University,
and, occasionally, in summer sessions, elsewhere, a course of
lectures under the title "Introduction to Modern Physical
Theories." These lectures have been adapted, as far as possible,
to meet the needs of two groups of students: (1) those special
students in physics who, before entering the specialized graduate
courses, desire a survey of the origin and development of modern
physics in order the better to understand the interrelations of the
more advanced courses; and (2) those students who, pursuing
either academic or professional curricula and having had the
usual elementary undergraduate courses in physics, wish a
further bird's-eye view of the· whole subject. This book is
based upon these lectures and has been prepared, although rather
reluctantly, as a result of the importunities of former students
and other friends.
The purpose of this book is, frankly, pedagogical. The
author has attempted to present such a discussion of the origin,
development, and present status of some of the more important
concepts of physics, classical as well as modern, as will give to the
student a correct perspective of the growth and present trend of
physics as a whole. Such a perspective is a necessary basis-so
the author, at least, believes-for a more intensive study of any
of the various subdivisions of the subject. While for the student
whose interests are cultural, or who is to enter any of the pro-
fessions directly or indirectly related to physics, such as engi-
neering, chemistry, astronomy, or mathematics, an account of
modern physics which gives the origin of current theories is
likely to be quite as interesting and valuable as is a categorical
statement of the theories themselves. Indeed, in all branches of
human knowledge the "why" is an absolutely indispensable
accompaniment to the "what." "Why?" is the proverbial
question of childhood. "Why?" inquires the thoughtful ( !)
student in classroom or lecture hall. "Why?" demands the
venerable scientist when listening to an exposition of views held
vii
Vlll PREFACE
by a colleague. Accordingly, if this book seems to lay somewhat
greater emphasis on matters which are frequently regarded as
historical, or, if here and there a classical experiment is described
in greater detail than is customary, it is with a desire to recognize
the importance of "why."
If one were to attempt to answer all of the "why's" raised by
an intelligent auditor in answer to a categorical statement, such
.as, "The atom of oxygen is composed of eight electrons surround-
ing a nucleus containing four alpha particles," one would have to
expound a large part of physical science from Copernicus to
Rutherford and Bohr. To attempt a statement of even the more
important concepts, hypotheses, and laws of modern physics
and of their origin and development would be an encyclopedic
task which, at least in so far as concerns the aim of this book,
would fall of its own weight. Accordingly, it has been necessary
to seiect those parts of the subject which best serve our purpose.
This selection, as well as the method of presentation, has been
based upon the experience gained in giving the above-mentioned
lectures to numerous groups of students. Many very important
developments, particularly the more recent ones, either have
been omitted entirely or have been given only a passing comment.
And even in those parts of the subject which have been discussed,
there has been no attempt to bring the discussion strictly up to
date. Indeed, with the present rapid growth of physics, it
would be quite impossible for any book, even a special treatise,
to be strictly up to date. Happily, for our purpose, up-to-date-
ness is not an imperative requisite, since it is assumed that the
student who wishes the latest knowledge will consult the current
periodicals.
In this connection, it should be emphasized that this book is an
introduction to modern physical theories and is intended neither
as a compendium of information nor as a critical account of any
of the subjects discussed. In preparing the manuscript, the
author has consulted freely the many very excellent texts which
deal with the various special topics. Save for here and there a
very minor item, or an occasional novelty in presentation, the
book makes no claim to originality, except, perhaps, as regards
the viewpoint from which some parts have been written.
It is assumed that the student is familiar with the elementary
principles of calculus, for no account of modern physics can
dispense with at least a limited amount of mathematical dis-
PREFACE IX
cussion, if for no other reason than to emphasize the fact that, in
the progress of physics, theory and experiment have gone hand
in hand. Partly, however, for the sake of brevity and partly in
the attempt always to keep the underlying physical principles
in the foreground, numerous "short cuts'' and simplifications,
some of them perhaps rather questionable from a precise stand-
point, have been introduced. These elisions should cause no
confusion.
The student who, in his educational career, has reached the
point where he can, with profit, pursue a course based on such
a book as this, has passed beyond the stage where he assimilates
only the material presented in lecture or class and has come to
regard a "course" as a channel to guide his own independent
studies, branching out from the "course" in such directions as
his fancy or interests may lead him. It is hoped that students
reading this book will do likewise. Deliberately, the author has
not given a collected bibliography at the end of each chapter, or
a list of problems and suggested topics for study. Rather,
references, in most cases to original sources, have been given at
appropriate points in the text, and it is hoped that, starting from
these references, the student will prepare his own bibliography of
such parts of the subject as appeal to him. The advantage to
the student of such a procedure is obvious. Quite apart from
the value of the experience gained in making contact with, and
in studying, the literature of any subject, the reading of first-
hand accounts of at least some of the more important develop-
ments will give the student a better understanding of the subject
than can, in general, be gained by textbook study only. Accord-
ingly, he will find here and there throughout this book suggestions
of important articles which should be read in the original. Like-
wise, in many places the discussion has, of necessity, been brief,
and the student is referred to special treatises for further details.
Various supplementary questions and problems will also arise at
numerous points as the student reads the text.
There is no more fasc.inating story than an account of the
development of physical science as a whole. (Any scientist
would probablymake the same statement about his own science!)
Such a study leads to certain broad generalizations which are of
outstanding importance in evaluating current theories and con-
cepts. For example, one finds that, taken by and large, the
evolution of physics has been characterized by continuity. That
x PREFACE
is to say: With few exceptions, the ideas, concepts, and laws of
physics have evolved gradually; only here and there do we find
outstanding discontinuities. The discovery of photoelectricity,
of X-rays, and of radioactivity represent such discontinuities
and are correctly designated "discoveries." But we must use
"discover" in a quite different sense when we say that J. J.
Thomson "discovered" the electron. The history of the electron
goes back at least to Faraday. Thomson's experiments are all
the more brilliant because he succeeded in demonstrating, by
direct experiment, the existence of something, evidence con-
cerning which had been previously indirect. Then, there are the
respective roles played by qualitative and by quantitative work.
Numerous important discoveries have been made "by investi-
gating the next decimal place.'' Witness the discovery of argon.
And ever since Kepler proved that the orbits of the planets are
ellipses, relations expressible in quantitative form have carried
greater weight than those which could be stated only qualita-
tively. For example, Rumford's experiments on the production
of heat by mechanical means were suggestive. But Joule's
measurement of the mechanical equivalent of heat was con-
vincing. If, directly, or indirectly by inference, the author has
succeeded here and there in the text in pointing out such generali-
zations as these, one more object which he has had in mind will
have been accomplished.
The author wishes to take this occasion to acknowledge his obli-
gations to those who have aided in the preparation of this book:
to his wife, for assistance in preparing the manuscript and in
proof reading; and to his many students, whose generous appro-
bation of the lecture courses upon which the book is based has,
in a large part, inspired its preparation. He is particularly
indebted to Dr. J. A. Be~ker, of the Bell Telephone Laboratories,
Inc., for his invaluable aid in reading the manuscript, pointing
out numerous errors, and suggesting important improvements.
F. K. RICHTMYER.
ITHACA, N. Y.
July, 1928
CONTENTS
PAGE
PREFACE . . . vu
INTRODUCTION 1
CHAPTER I
HISTORICAL SKETCH, FIRST PERIOD: EARLIEST TIMES TO 1550 A.D.
1. The Greeks . . . 5
2. Thales of Miletus . 5
3. Pythagoras . 6
4. Philolaus. . . . . 6
5. Anaxagoras and Empedocles 7
6. Democritus . . 8
7. Plato . . 9
8. Aristotle. . . 9
9. Euclid . . . 12
10. Aristarchus. 12
11. Archimedes . . 13
12. From the Greeks to Copernicus . 13
13. The Copernican System . . . . 14
CHAPTER II
HISTORICAL SKETCH, SECOND PERIOD (1550-1800 A.D.): THE RISE OF
THE EXPERIMENTAL METHOD
1. The Growing Dissatisfaction with Authority . . . . . . 17
2. Galileo Galilei . . 18
3. Tycho and Kepler. . . . . . . . 25
4. The Experimental Method Spreads . 32
5. Sir Isaac Newton . . . . 34
6. Newton's Contemporaries . . . . . 46
7. Mechanics during the Eighteenth Century . . . . 46
8. Heat during the Eighteenth Century. . .... 46
9. Light during the Eighteenth Century . . . . . . 47
10. Electricity during the Eighteenth Century . . . . . . 49
CHAPTER III
HISTORICAL SKETCH, THIRD PERIOD (1800-1890 .A..D.): THE RISE OF
CLASSICAL PHYSICS
1. Heat and Energy . 53
2. Light . . . . . . . . . . 56
3. Electricity and Magnetism. . 60
XI
Xll CONTENTS
PAGE
4. Michael Faraday . . 60
5. Joseph Henry. . . . 70
6. James Clerk Maxwell 72
CHAPTER IV
THE ELECTROMAGNETIC THEORY OF LIGHT
1. The Electrostatic System of Electrical Units . . 78
2. The Electromagnetic System of Electrical Units. 80
3. Ratio of the Two Systems of Units . 81
4. Some Fundamental Formulre. . . . . . . . . 82
5. Maxwell's Differential Equations of the Electromagnetic Field 83
6. The Differential Equations of the Electromagnetic Wave 94
7. The Electromagnetic Wave. . . . . . . . 101
8. Flow of Energy in an Electromagnetic Wave 103
9. The Electromagnetic Theory of Light . . 108
10. The Discovery of Electromagnetic Waves 110
11. The Refraction of Light 111
12. The Dispersion of Light 116
13. Summary . . 121
CHAPTER V
SOME THEOREMS CONCERNING MOVING CHARGES
1. The Magnetic Field Produced by a Moving Charge. . . 123
2. The Force Acting on a Charge Moving in a Magnetic Field 126
3. The Energy Contained in a Magnetic Field Surrounding a Mov-
ing Charge. . . . . . . . . . . . . . . . 127
4. The Energy Radiated by an Accelerated Charge . . . . 128
5. Some Special Cases of Radiation by Accelerated Charges 133
CHAPTER VI
THE PHOTOELECTRIC EFFECT
1. The Discovery by Hertz . 136
2. Some Early Experiments. 137
3. A Problem . . . . . . . 139
4. The Laws of Electrolysis . 140
5. Dispersion of Light . 142
6. The Zeeman Effect . . . 143
7. The Discovery of the Electron by Sir J. J. Thomson. 148
8. "Photoelectrons" . . . . . . . . . . . . . 155
9. Relation between Photoelectric Current and Intensity of
Illumination of the Cathode . . . . . . . 157
10. Velocity Distribution Curves for Photoelectrons. . . . . . . 158
11. Relation between the Velocities of Photoelectrons and the
Frequency of the Light . . . . 162
12. Origin of the Photoelectrons . . . . . 165
13. Source of the Photoelectric Energy . 168
14. What Is the Photoelectric Mechanism? 170
15. The Photoelectric Effect and the Corpuscular Theory of Light . 173
CONTENTS Xlll
CHAPTER VII
THE ORIGIN OF THE QUANTUM THEORY
PAGE
1. Thermal Radiation . . . . . . . . . . . . . 177
2. Some Fundamental Concepts and Definitions. 179
3. The "Black Body" and Its Properties. . . . 183
4. Relation between Absorptivity and Emissive Power. 184
5. The Emissive Power of a Black Body 187
6. Pressure of Radiation . . . . . . . 190
7. The Stefan-Boltzmann Law. . . . . 194
8. Experimental Verification of the Stefan-Boltzmann Law 201
9. The Spectral Distribution of Black-body Radiation . . . 203
10. The Successes and the Failure of Classical Thermodynamics 205
11. Degrees of Freedom and the Equipartition of Energy 214
12. Relation between Energy per Degree of Freedom and Temper-
ature . . . . . . . . . . . . 218
13. The Rayleigh-Jeans Radiation Law . . . . . . . . . . 221
14. Planck's Radiation Law: The Birth of the Quantum Theory 229
CHAPTER VIII
THE QUANTUM THEORY OF SPECIFIC HEATS
1. The Empirical Law of Dulong and Petit . . . . . . . . . 248
2. Variation of Atomic Heats of Solids with Temperature. 250
3. The Classical Theory of the Specific Heats of Solids. 252
4. Einstein's Theory of the Atomic Heats of Solids 256
5. Characteristic Tern peratures . . 259
6. Characteristic Frequencies . . . . . . . . . . 260
7. The Nernst-Lindemann Formula for Atomic Heats . 267
8. Debye's Theory of Atomic Heats 268
9. Further Considerations. . 273
10. The Atomic Heats of Gases. . . . 278
11. The Suppression of Degrees of Freedom . 285
CHAPTER IX
SERIES RELATIONS IN LINE SPECTRA
1. Units and Methods of Measurement. . . . . 290
2. Early Search for Series Relations in Spectra . 292
3. Balmer's Formula for the Hydrogen Spectrum . 295
4. Rydberg's Formula for Spectral Series. 297
5. Relations between Series . . . 304
6. The Rydberg-Schuster Law. . . . 308
7. Relations between Doublet Series. 308
8. Relations between Triplet Series 310
9. Satellites . . . . . . . 311
10. Combination Lines . . . . . . 311
11. The Significance of Spectral-series Terms. 312
12. Spectral Series and Atomic Properties . . . . 314
XIV CONTENTS
PAGE
13. Enhanced or Spark Spectra . . . . . . . . . . . . . . . . . . 318
14. Band Spectra . . . . . . . . . . . . . . . . . . . . . 320
15. Effect of External Physical Conditions on Spectral Lines. . 321
CHAPTER X
THE NUCLEAR ATOM AND THE ORIGIN OF SPECTRAL LINES
1. Early Views on Atomic Structure . 330
2. The Thomson Atom . . . . . . . . . . . . . . . . . . . . 331
3. The Scattering of Alpha Particles in Passing through Matter . 337
4. Rutherford's Nuclear Atom. . . 340
5. The Phase Integral . . . . . . . . . . . . . 346
6. Bohr's Extension of the Nuclear Atom Model . . . . . . 350
7. Further Successes of the Rutherford-Bohr Atom Model 356
8. Elliptical Orbits in Bohr's Theory. 364
9. Fine Structure of Spectral Lines . 372
10. The Selection Principle. . . . . . . 380
11. Systems with More than One Electron. 383
12. The Absorption of Energy by Atoms. . . 383
13. Energy-level Diagrams. . . . . . . . 396
14. Notation; Inner Quantum Numbers. . . . . 402
15. Molecular Spectra . . . . .. 407
CHAPTER XI
THE ARRANGEMENT OF ELECTRONS IN ATOMS
PART I. THE STATIC ATOM MODEL . . . . . . . . 413
1. The Inert Gases: Atomic Numbers and Properties. 413
2. Some Chemical Properties of the Lighter Elements 418
• 3. The Heavier Elements. . . . . . . 425
p ART II. THE DYNAMIC ATOM MODEL . . . . . . 433
4. The Problem of the Distribution of Electrons in Orbits . 433
5. Some Examples Illustrative of the Spectroscopic Method. 434
6. Inner Quantum Numbers . . . . . . . . . . . . . 442
7. Complete Distribution Schemes for All Elements . . . . . . 445
CHAPTER XII
X-RAYS
1. Roentgen's Discovery . . . . . . . . . 447
2. Some Early Experiments and Theories 450
3. The Ether-Pulse Theory of X-rays . 458
4. Characteristic Secondary Radiation . . . 467
5. The Crystal Diffraction Grating. . . . . . . 472
6. The Experiment of Friedrich, Knipping and Laue. . . . . . 478
7. The X-ray Spectrometer. . . . . . . . . . . . . 480
8. Bragg's Discovery of Monochromatic Characteristic Radiations . . 483
9. Moseley's Law . . . . . . . . . . . . . . 485
10. The Continuofis X-ray Spectrum . . . . . . . . . 489
11. The Empirical Laws of the Absorption of X-rays . . . 497
CONTENTS xv
PAGE
12. Characteristic X-ray Spectra . . . . . . . 502
13. X-ray Energy-level Diagrams. . . . . . . 508
14. Fluorescence and the Photoelectric Effect . 511
15. The Scattering of X-rays; The Compton Effect. 517
16. The Refraction of X-rays . . . 528
CHAPTER XIII
THE NUCLEUS
p ART I. THE MASSES OF ATOMS 539
1. Positive Rays. . . 536
2. Isotopes . . . . . . . . 539
3. The Packing Effect . . 547
p ART II. RADIOACTIVITY . . 549
4. Becquerel's Discovery . . . . . . . 549
5. The Types of Radioactive Radiations 550
6. Sources of the Rays. . . . . . . . 555
7. The Radioactive Disintegration Series. 556
8. Gamma-ray Spectra. . 558
9. Nuclear Energy Levels. . . . . . . . . . . . . . . . . 561
10. The Structure of the Nuclei of the Radioactive Elements 563
APPENDIX I
(a) Atomic Numbers and Atomic Weights. . . . . . . . . . . 569
(b) The Periodic Table, Giving Atomic Numbers and Isotopes. . 570
(c) Bohr's Periodic Table of the Elements. . . . . . 571
APPENDIX II
Arrangement of Electrons in Orbits According to Foote . . . . . . . 572
APPENDIX III
Some Important Constants. . 575
INDEX . . . 577
INTRODUCTION
TO MODERN PHYSICS
INTRODUCTION
1. The term "modern physics," taken literally, means, of
course, the sum total of knowledge included under the head of
present-day physics. But by "modern physics,'' many writers
and speakers frequently mean that part of present-day physics
which has been developed during the past twenty-five or thirty
years; in contradistinction to "classical physics," by which is
meant the sum total of physics as it was known in, say, 1890.
The justification for the latter use of the term is to be found
partly in the fact that advances since 1890 have been very great
indeed and partly in the fact that some of these advances have
brought into question, or are in direct contradiction to, many of
the theories which, in 1890, were thought to be firmly and finally
established. For example, few, if any, physicists in 1890
questioned the wave theory of light. Its triumph over the old
corpuscular theory was thought to be final and complete, par-
ticularly_after the brilliant experiments of Hertz, in 1887, which
demonstrated, beyond doubt, the fundamental soundness of
Maxwell's Electromagnetic Theory of Light.
2. And yet, by an irony of fate which makes the story of
modern physics full of the most interesting and dramatic situa-
tions, these very experiments of Hertz brought to light a new
phenomenon-the photoelectric effect-which, together with a
series of discoveries 1 coming in rapid succession in the single
decade, 1887-1897, was the beginning of the development of the
now famous quantum theory. This theory is, in many of its
aspects, diametrically opposed to the wave theory of light.
Indeed, the reconciliation of these two theories, each based on
incontrovertable experimental evidence, may be said to be one
of the two great problems of modern physics; the other problem
being that of the structure of matter.
1 Namely, X-rays in 1895; radioactivity in 1896; the electron in 1897.
1
2 INTRODUCTION
3. It shall be the purpose of the following pages to give a brief
outline of the origin, development, and, in so far as may be
possible in this rapidly developing subject, the present status of
these two problems.
4. But a history of the United States cannot begin abruptly
with July 4, 1776. In like manner, if we would understand the
full meaning of the growth of physics since, say, 189.0, we must
have clearly in mind at least the main events in the development
of the subject up to that time. Accordingly, we shall begin our
study by a brief account of the history of physics up to a half-
century ago.
6. In presenting this brief historical survey, however, the
author has in mind another purpose, toward which he hopes the
reader will be, ultimately at least, sympathetic. Modern
scientists have, with few exceptions, grossly neglected to cultivate
the history of their respective sciences. How many physicists
can answer the questions: When was the law of the conservation
of energy first enunciated? Who was Count Rumford? Did the
concept of universal gravitation spring full-grown from the head
of that genius Newton? Indeed, when did Newton live?
Just as any good American should know the essential outline
of the history of his country, so any good physicist should know
the principal facts in the history of physics. For in that history,
in the lives of those men whose labors have given us our subject,
and in the part which physics has played in moulding human
thought and in contributing to modern civilization, the student
will find a story which is as full of human interest and inspiration
as is any subject of the curriculum.
What can be more inspiring than the life of Michael Faraday
and his whole-souled devotion to his work? Which have had
a greater effect on present-day civilization: the victories of
Napoleon or the electrons of J. J. Thomson? Was Roentgen
when he discovered X-rays seeking a new tool to help surgeons
set broken bones?
The physicist owes it to his science to possess such a knowledge
of the history of physics as gives him a correct perspective of the
development and present-day importance of the subject and, in
turn, enables him to acquaint his lay contemporaries with these
essential facts. If there is apathy on the part of the public
toward physics, the physicist himself is largely at fault, since he is
so absorbed in the interest of the present that he forgets th0
'INTRODUCTION 3
importance of the past. He would find it much easier to justify
to a popular audience the latest experiments on, say, the magnetic
spectrum of electrons emitted from targets radiated by X-rays,
if he prefaced his remarks by an account of the relation of
Faraday's work to the modern dynamo.
It is hoped, therefore, that the student of these pages who
proposes to follow physics as a profession, as well as the student
whose interest is largely cultural, will extend the following all too
brief historical sketch by independent study, particularly of
biography.
6. In order to make it easier to keep the essential facts m
mind, we may, somewhat arbitrarily, divide the history of
physics into four periods.
The FmsT PERIOD extends from the earliest times up to about
1550 A.D., which date marks, approximately, the beginning
of the experimental method. During this long era, there was, of
course, substantial advance in the accumulation of the facts of
physics as a result of the observation of natural phenomena,
particularly by the Greeks, whose authority was almost unques-
tioned for many centuries. But the development of physical
theories was rendered impossible, partly by the speculative, meta-
physical nature of the reasoning employed, but more particularly
by the almost complete absence of experiment to test the cor-
rectness of such theories as were proposed. The main charac-
teristic of this period, therefore, is the absence of experiment.
7. The SECOND PERIOD extends from 1550 to 1800 A.D. While
numerous basic advances were made during this period-by
such men as Gilbert, Galileo, Newton, Huyghens, Boyle, Benja-
min Franklin-its most important characteristic is the develop-
ment and the firm establishment of the experimental method as a
means of 'scientific inquiry, as is well illustrated by Galileo's
famous experiment (about 1590) of dropping two bodies of
unequal weight from the leaning tower of Pisa, thereby proving
by experiment the incorrectness of the assertion of Aristotle that
the heavier body would fall more rapidly-an assertion which had
been believed implicitly for nearly two thousand years.
It took two centuries after Galileo's experiment to overcome
prejudice, dogma, and religious intolerance and to bring universal
recognition, even among scientific men, to the basic principle that
. . . science can advanc'e only so far as theories, themselves based upon
experiment, are accepted or rejected according as they either agree with or
are contrary to other experiments devised to check the theory.
4 INTRODUCTION
8. The THIRD PERIOD, 1800-1890, is characterized by the
development of what is now called "classical physics." The
experiments of Count Rumford (about 1798) led ultimately to
our present kinetic theory of heat. The observations of Thomas
Young (1802) and his proposal of the principle of interference
(of two beams of light) resulted ultimately in the triumph of
Huygens' Wave Theory of Light over the corpuscular theory, as
supported by Newton. And the researches of Faraday gave
Maxwell the material for the crowning achievement of this
period, namely, the electromagnetic theory.
So profound were these, and many other, developments that,
by 1880, not a few physicists of note believed that ail the impor-
tant laws of physics had been discovered and that, henceforth,
research would be concerned with clearing up minor problems
and, particularly, with improvements of methods of measurement
so as "to investigate the next decimal place." They could not
have foreseen that the world of physics was on the eve of a series
of epoch-making discoveries, destined, on the one hand, to
stimulate research as never before and, on the other, to usher in
an era of the application of physics to industry on a scale pre-
viously unknown.
9. The FouRTH PERIOD dates quite definitely from the dis-
covery of the photoelectric effect, in 1887. In rapid succession,
followed the discovery of X-rays, in 1895; of radioactivity, in
1896; of the electron, in 1897; and the beginning of the quantum
theory, in 1900.
So varied and extensive have been the developments in both
pure and applied physics from that time to the present that it is
difficult · to characterize this period by a single appelation.
Hence, perhaps one may use the pleonasm "modern physics."
Only the historian of a century hence can properly evaluate the
growth of physics during the first part of the twentieth century.
We, of the present, are too close to it to grasp its full significance.
10. It is, obviously, far beyond the scope of this book to give
a detailed account of the history of physics during each of these
four periods. Rather, the material chosen for discussion, par-
ticularly the biographical notes, is to be taken as representative.
It should, also, be pointed out that so many of the advances in
physics have come, either directly or indirectly, from astronomy,
that frequent reference to the history of the latter subject will
be made.
CHAPTER I
HISTORICAL SKETCH, FIRST PERIOD : EARLIEST TIMES
TO 1560 A.D.
1. The Greeks.-It would be quite incorrect to say that the
Greeks were entirely unacquainted with physical laws and facts.
Relatively speaking, their contributions to the natural sciences
were far less than to mathematics, literature, art, and meta-
physics. But, in spite of their vague and misty philosophizing
concerning natural phenomena, and in spite of their failure to
test theory by experiment, the Greeks gave to the world much
of the physics that was known up to 1400 A.D. And in their
writings one finds, here and there, the germ of such fundamental
modern principles as the conservation of matter, inertia, atomic
theory, the finite velocity of light, and the like.
Certain of these ideas-for example, the heliocentric hypothesis
of Aristarchus-doubtless led to further developments centuries
later. Others-for example, the atomic hypothesis of Democri-
tus-probably had little or no connection with the modern
counterpart, largely on account of the speculative reasoning upon
which the Greek theories were based.
2. Thales of Miletus (624-547 B.c.), according to the later
writings of Aristotle, was acquainted with the attractive power of
magnets and of rubbed amber. According to some authorities,
he was able to predict eclipses, but others claim that he was
rather the first to explain the cause of eclipses. 1 Apparently, the
Chaldeans some 2,000 years B.c., had devised a rule-of-thumb
method for predicting eclipses, on the basis of their discovery
that, after a cycle of 18 years, eclipses repeat themselves in the
same order, a little later in the year. Other discoveries and
doctrines attributed to Thales 2 are the inclination of the ecliptic;
the spherical shape of the earth; and the measurement of the
angular diameter of the moon, as seen from the earth, as 7,f 2 0
part of a whole circle. Aristotle credited Thales, however, with
the doctrine that the earth was cylindrical in shape and rested on
1 HEATH: "Aristarchus of Samos," pp. 12-23.
2 RosENBERGER: "Geschichte der Physik," vol. I, p. 6.
5
6 HISTORICAL SKETCH, FIRST PERIOD [CHAP. I
water; and, according to Heath (loc. cit., pp. 21-23), there is
reason to believe that the angular diameter of the moon had been
measured by the Babylonians at least a thousand years before
Thales' time.
3. Pythagoras (580-500 B.c.) was one of the greatest of the
early Greek philosophers and the founder of the Pythagorean
school. He seems to have been the first to hold that the earth
was spherical, although the basis of his belief is not known.
According to Heath, his arguments were probably based on
"mathematico-esthetic" reasoning, since "the sphere is the
most perfect of all figures." Pythagoras himself, and probably
his immediate successors among the Pythagoreans, believed
that the universe was spherical in shape, with the earth at its
center, and that the sun, stars, and planets moved in independent
circles around the earth as a center.
4. Philolaus, one of the later Pythagoreans and a contempor-
ary of Socrates (470-399 B.c.), abandoned the geocentric hypoth-
esis and boldly reduced the earth to the status of a planet. According
to his system, the universe is spherical, finite in size, and has
at its centre a "fire" 1 which contains the "governing principle"
which directs the universe. The earth revolves around this
central fire in a circle, once per day, always keeping its face
outward, so that we never see the fire. .Beyond the earth,
revolving in similar but larger circles, come, in turn, the moon, the
sun, the five planets (in the order Mercury, Venus, Mars,
Jupiter, Saturn), and, last of all, the sphere containing the fixed
stars. The period of revolution of the earth about this central
"fire" is 1 day; that of the sun, about 365 days. (Day and night
are thus accounted for.) Schiaparelli 2 gives the periods of
revolution of these several bodies, as computed by Philolaus,
and compares them with modern data as follows (the terrestrial
day is the unit of time):
Period of revolution
Planet according to
Philolaus I Modern view
Saturn ............................ . 10, 752. 75 days 10, 759. 22 days
Jupiter ........................... . 4,301.10 4,332. 58
Mars ........................... ··· 693.71 686.98
Venus I
Mercury r . : ...................... . 364.5 365.26
Sun J
Moon ............................ . 29.50 29.53
1 This "fire" is not the sun, as some writers have apparently assumed.
2 Qu,oted in HEATH'S "Aristarchus,," p. 102.
SEC. 5] ANAXAGORAS AND EMPEDOCLES 7
5. Anaxagoras and Empedocles.-Almost contemporary with
Philolaus were Anaxagoras (500-428 B.c.) and Empedocles (484-
424 B.C.). According to Plato, Anaxagoras neglected his pos-
sessions in order to devote himself to science, and, indeed, he is
said to have affirmed that his purpose on earth was '' the in vestiga-
tion of the sun, moon, and heaven." Apparently, human nature,
as well as human curiosity, has not changed for 2,500 years, for,
to paraphrase a famous saying, the course of pure science has
never run smoothly. Anaxagoras was accused of impiety
because he taught that the sun was a red-hot stone and that the
moon was simply earth. For holding this doctrine, he was
banished from Athens-thereby becoming one of the first of a
long line of martyrs of science.
Anaxagoras is credited with the almost epoch-making discovery
that the moon does not shine by its own light but that "the sun
places the brightness in the moon" and "the moon is eclipsed
through the interposition of the earth." Also, "The sun is
eclipsed at new moon through the interposition of the moon.'' 1
Anaxagoras believed that the Milky Way is due to the shadow
which the earth casts in the sky when the sun, in its 'revolution
around the earth, is below the earth, thereby rendering visible the
fainter stars-apparently, the principle involved being somewhat
the same as that which makes it impossible to see the stars in
the daytime. No explanation is offered as to why the moon is
not, therefore, eclipsed when in the Milky Way. The theory was
later refuted by Aristotle on essentially correct grounds, namely,
that the shadow which the earth casts is a cone the apex of which
does not reach the stars, since the sun is larger than the earth and
is very distant, and the stars are still more distant. ·
To Anaxagoras is due the germ of the idea of the atomic hypoth-
esis of Democritus, who lived in the next generation. Anaxa-
goras denied the contention of the earlier Greeks regarding the
creation or destruction ·of matter and taught that these changes
consist of combinations or separations of invisible, small particles
(spermata) of matter. These particles themselves are unchange-
able and imperishable and differ from each other in form, color,
and taste-a doctrine which, qualitatively, foreshadows the law
of the conservation of matter.
1 HEATH: "Aristarchus," p. 78. Quotations cited from later Greek
writers.
8 HISTORICAL SKETCH, FIRST PERIOD [CHAP. I
Empedocles reduced the elements in the theory of Anaxagoras
to four-earth, water, air, and fire-through combinations and
separations of which the All exists. He also held definite views
concerning the phenomena of light. According to him, light is
due to the emission by the luminous or visible body of small
particles which enter the eye and which are then returned from
the eye to the body, the two "streams" giving rise to the sense
of form, color, etc.
Empedocles, according to Aristotle, proposed thd light "takes
time to travel from one point to another"-this theory, of course,
being based on abstract speculation, rather than on observation
or experiment. The theory was rejected by Aristotle, who stated
that "though a movement of light might elude our observation
within a short distance, that it should do so all the way from east
to west is too much to assume." 1
6. Democritus (460-370 B.c.) gave more definite, quantitative
form to the atomic hypothesis of Anaxagoras by postulating
that the universe consists of empty space and an (almost) infinite
number of indivisible and invisible particles which differ from each
other only inform, position, and arrangement. Bodies come into
being, and vanish, only through combinations and separations of
these atoms. This hypothesis Democritus regards as reason-
able, for, he argues, the creation of matter is impossible, since
nothing can come from nothing and, further, nothing which is
can cease to exist. Aristotle 2 puts this argument for the inde-
structibility of matter in the na:ive form: "If, then, some one of
the things which are is constantly disappearing why has not the
whole of what is been used up long ago and vanished away?"
But he rejects the atomic hypothesis, which, on the basis of
speculative reasoning alone, could not evolve beyond the point
where Democritus left it. It was twenty-two centuries later, when
Dalton, on the basis of experimental evidence, announced his law
of multiple proportions, which, nourished by further experi-
ments, rapidly grew into our present theory of the atomicity of
matter.
With Empedocles, Democritus held that vision arises through
the entrance into the eye of small particles-contrary to the
then prevalent belief that vision resulted in something which
Quoted in HEATH' s "Aristarchus," p. 93.
1
"De Generatione et Corruptione," Book I, Chap. III; translated by
-2
H. H. Joachim. The Clarendon Press (1922).
SEC. 8] ARISTOTLE 9
"reached out" from the eye, just as the sense of touch involves
reaching out our hands to grasp an object.
7. Plato (429-347 B.C.), although one of the greatest of the
Greek philosophers, did not make contributions to physics such
as are of interest in connection with the present discussion.
Quite the opposite, however, is true of Aristotle, a pupil of Plato.
8. Aristotle (384-322 B.c.) contributed so much to all branches
of knowledge-logic, rhetoric, ethics, metaphysics, psychology,
natural science-that it is difficult to sift out that which is
germane to a brief history of physics. Perhaps the most impor-
tant single fact is the tremendous influence which, as a result of
his intellectual brilliance and achievements in many branches of
learning, he exerted for many succeeding centuries in all branches,
physics included. Aristotle is justly entitled to fame for much
of his work; but, viewed from our twentieth-century vantage
point, not a little of his reasoning concerning the physical universe
must be regarded as piffle. For example, in "De Generatione
et Corruptione," he discusses the rather abstract question of
the "coming-to-be" and the "passing-away" of things and
argues for the indestructibility of matter by saying that (Book
II, Chap. X) "it is far more reasonable (to assume) that what is
should cause the coming-to-be of what is not than that what is not
should cause the being of what is." Which is understandable.
But then follows the curious argument: "Now that which is
being moved is, but that which is coming to be is not: hence,
also, motion is prior to coming-to-be . . . and we assert that
motion causes coming-to-be." But coming-to-be and passing-
away are two processes contrary to one another. Therefore,
says Aristotle, we must look for two motions, likewise contrary,
as the cause of both coming-to-be and passing-away. Since these
processes go on continuously, we must look for continuous motion.
Only motion in a circle is continuous, and motion in an inclined
circle has the necessary duality of opposing movements. Such
a motion is that of the sun, which, as it approaches (spring),
causes coming-to-be and, as it retreats (autumn), causes decay.
The motion of the sun along the ecliptic is, therefore, the cause of
generation and destruction.
And yet Aristotle frequently calls in observed facts to sub-
stantiate his speculation. For example, in "De Caelo" (Book
II, Chap. XIV), after proving, by a more or less abstract argu-
ment, that the earth is spherical, he says:
10 HISTORICAL SKETCH, FIRST PERIOD [CHAP. I
The evidence of the senses further corroborates this. How else would
eclipses of the moon show segments as we see them? . . . since it is
the interposition of the earth that makes the eclipse, the form of this
line [i.e., the earth's shadow on the moon] will be caused by the form of
the earth's surface, which is therefore spherical.
He, also, points to the apparent change in altitude of the stars
as one travels north or south and concludes that "not only is
the earth circular, but it is a circle of no great size." He then
quotes the figure 400,000 stades (about 10,000 miles) as the
circumference (diameter ?) of the earth, as determined by
"mathematicians" and remarks, "This indicates . . . that,
as compared with the stars, it [the earth] is not of great size."
Indeed, Aristotle, in spite of his own misty philosophising,
places, in theory if not in his own practice, great emphasis on
the importance of facts in connection with scientific development.
For, in a paragraph in "De Generatione et Corruptione" (Book
I, Chap. II), he states:
Lack of experience diminishes our power of taking a comprehensive
view of the admitted facts. Hence those who dwell in intimate associa-
tion with nature and its phenomena grow more and more able to formu-
late, as the foundation of their theories, principles such as to admit of
a wide and coherent development; while those whom devotion to
abstract discussions has rendered unobservant of the facts are too ready
to dogmatize on the basis of a few observations. 1
This is surely good doctrine even for twentieth-century scientists.
An attempt to summarize Aristotle's views on physics is
beyond the scope of this book, 2 but reference may be made to
two of his doctrines because of their bearing upon subsequent
history.
The first is his views on falling bodies. In "De Caelo"
(Book III, Chap. II), in the course of his proof that every body
must have a finite (i.e., neither infinite nor zero) weight, he says: 3
Suppose a body A without weight, and a body B endowed with weight.
Suppose the weightless body to move a distance CD, while Bin the same
time moves the distance CE, which w_ill be greater since the heavy thing
1 Taken from the English translation by H. H. Joachim.
2 A summary of Aristotle's works is given by W. D. Ross in his book
"Aristotle." Methuen and Company, London (1923). The reader will
also find it very interesting to consult the English translation of the works
of Aristotle, under the editorship of W. D. Ross. The Clarendon Press,
Oxford.
3 Quoted from the English translation by J. L. Stocks.
PLATE !.-Aristotle.
(Facino page 10)
SEC. 8] ARISTOTLE 11
must move further. Let the heavy body then be divided in the proportion
CE: CD (for there is no reason why a part of B should not stand in this
relation to the whole). Now if the whole moves the whole distance CE,
the parl must in the same time move the distance CD.
And in the "Physica" (Book IV, Chap. VI), in connection
with the proof that there is no void separate from the bodies,
appears an argument which, paraphrased by Ross, runs as follows:
Speed of movement varies in the ratio of the weight of the moved body
to the resistance of the medium. Therefore (a) that which passes
through a void [i.e., zero resistance] should take no time to do so, and (b)
a light body should take no longer to move through the void than the
heavy one. But in fact nothing moves in no time, and heavy bodies
alicays move faster than light ones.
These two references illustrate Aristotle's belief that a heavy
body will move (fall?) more rapidly than a light one, a doctrine
which was held unquestioned until Galileo's famous experiment
at the leaning tower of Pisa.
The second doctrine referred to is that of the motion of the
earth, sun, and planets. Aristotle, in his "De Caelo" (Book
II, Chap. XIV), after a series of abstract arguments, in the course
of which he states that "heavy bodies forcibly thrown quite
straight upward return to the point from which they started,
even if they be thrown to an infinite ( !) distance," concludes
'' that the earth does not move and does not lie elsewhere than
at the center." That is, the earth is immobile and in the center
of the universe, which, according to him consists of a series of
concentric spheres containing, and causing, the motion of
the sun, planets, and stars. The authority of Aristotle was so
great as completely to overshadow the theory of Philolaus that
the earth moves round a central fire and to render sterile the
brilliant work of Aristarchus, who, in the century following
Aristotle, proposed a cosmogony essentially in agreement with
the modern Copernican system. Concerning Aristotle's views
on other branches of physics, little comment is necessary. His
theory of vision coincided with that of Democritus and Empe-
docles, but he rejected the atomic theory of Democritus. In
his "Mechanica "-probably not written by Aristotle himself-
is found a discussion of levers, pulleys, wedges, and the like. The
doctrine of inertia is raised by the question (Chap. XXXIII):
"How is it that a body is carried along by a motion not its own,
if that which started it does not keep following and pushing it
along?" But his answer quite misses the point, for he says that
12 HISTORICAL SKETCH; FIRST PERIOD [CHAP. I
the original impelling force pushes one thing (i.e., air or other
medium in which the motion occurs), "and this in its turn pushes
along something else," and it finally comes to rest "when the
weight of the moving object has a stronger inclination downward
than the force of that which pushes it." And yet, as previously
quoted, Aristotle clearly recognized the effect on motion of the
"resistance of the medium."
9. Euclid (last half of fourth century B.c.), in addition to his
mathematical works, wrote two books on optics, the "Optics"
and the "Catoptics." These are largely mathematical and
assume vision to be due to particles emitted from the eye. He
treats of visual angles and apparent size of objects and correctly
enunciates the modern law of reflection.
10. Aristarchus (about 310-230 B.c.) was an astronomer
and mathematician rather than a physicist, but his enunciation
of a cosmogony identical with that proposed by Copernicus
2,000 years later commands our attention, because of its relation
to Newton's discovery of the law of gravitation. What led
Aristarchus to this discovery, we are not told. Indeed, no
mention of this hypothesis is made in his only extant work,
"On the Sizes and Distances of the Sun and Moon." But in a
book called "The Sand-reckoner" by Archimedes, 25 years
younger than Aristarchus, we are told that "Aristarchus of
Samos brought out a book" containing the hypothesis "that
the fixed stars and the sun remain unmoved; that the earth
revolves around the sun in the circumference of a circle, the sun
lying in the middle of the orbit"; and that "the sphere of the
fixed stars" is very great compared to the circle in which the
earth revolves. The theory is similar to that of Philolaus, in
that both, discarding the then-prevalent idea that the earth is
fixed in the center of the universe, make the earth simply one of
the planets. The theory is a far-reaching improvement over
that of Philolaus, in that it dispenses with all the unknown in the
universe-the central fire-and builds a cosmogony out of known
heavenly bodies only. But the prestige of Aristotle was too great,
and the geocentric hypothesis which he supported was so com-
pletely satisfactory to the ancient mind, that Aristarchus' theory
was practically lost for nearly 2,000 years. 1
1
HEATH, T. L.: "Aristarchus of Samos," an exceedingly interesting and
valuable book (previously referred to). It contains a review of early Greek
astronomy up to the time of Aristarchus, a discussion of his work, and a
translation of his only extant book, "On the Sizes and Distances of the Sun
and Moon."
SEC. 11] ARCHIMEDES 13
11. Archimedes (287-212 B.c.), whose name is known to every
student of elementary physics because of the famous principle of
hydrostatics which bears his name, was one of the most noted
physicists of antiquity. He was a man of great ability in what
would now be called "theoretical (or mathematical) physics"
as well as a practical engineer-a sort of ancient Lord Kelvin.
Of his nine extant works, six deal almost exclusively with mathe-
matics, mostly geometry, and three touch the subject of physics,
namely, "On the Equilibrium of Planes," "The Sand-reckoner,"
and "On Floating Bodies." In the first of these are developed
numerous propositions relative to the center of gravity. In the
second, he computes that 10 63 grains of sand would fill the sphere
of the universe, as fixed by Aristarchus. In "On Floating
Bodies," by a series of propositions, is laid the foundations of
hydrostatics. His Proposition 7 enunciates the famous principle:
"A solid heavier than a fluid will, if placed in it, descend to the
bottom of the fluid, and the solid will, when weighed in the fluid,
be lighter than its true weight by the weight of the fluid dis-
placed." Indeed, so apt and accurate was Archimedes' applica-
tion of mathematics, that he might be called the founder of
mathematical physics.
12. From the Greeks to Copernicus.-To give but a passing
comment to the 17 centuries between Archimedes and Coper-
nicus would seem to give the reader the impression that no
developments of moment occurred during that long period.
This impression is almost, but not quite, correct. The one impor-
tant exception is in the field of optics-geometrical optics, not
physical optics. Ptolemy of Alexandria (70-147 A.D.) collected
the optical knowledge of his time in book form. This treatise
discusses, among other things, reflection from mirrors-plane,
convex, concave-and, particularly, refraction, which Ptolemy
seems to have studied experimentally. He states, qualitatively,
the correct laws of refraction (i.e., the bending of the refracted ray
toward or away from the perpendicular on entering a more dense
or a less dense medium). He gives, in degrees, relative values of
angles of incidence and of refraction for air-water, air-glass, and
water-glass surfaces, and he describes an apparatus by which he
determined these quantities. He states that for a given inter-
face these two angles are proportional. He also mentions atmos-
pheric refraction as affecting the apparent position of stars-a
discovery attributed to Cleomedes, some decades earlier.
14 HISTORICAL SKETCH, FIRST PERIOD [CHAP. I
From Ptolemy to the Arabian, Al Hazen (who died about 1038
A.D.), is a span of 9 centuries-twice the total lapse of time from
the discovery of America to the present-during which there
was stagnation in almost all lines of intellectual pursuits. But
about the eighth century A.D., as an indirect result of religious
activity, the Arabs began to cultivate chemistry, mathematics,
and astronomy, in large part by translating into Arabic the works
of the Greeks. And in a few instances, the Arabs made original
contributions. It was under these influences that Al Hazen,
following the exposition set forth by Ptolemy and his predeces-
sors, produced a work on optics in /seven books. This treatise
sets forth a clear description of the optical system of the eye;
discusses the laws of concave, and convex mirrors, both spherical
and cylindrical; and carries the subject of refraction a little
farther than Ptolemy carried it by recognizing that the propor-
tionality between the angles of incidence and refraction holds
only for small angles.
But from the time of the Greeks until the sixteenth century,
there were practically no developments in physics worthy even
of passing comment. The authority of the Greeks was supreme.
Then came a period of intense intellectual activity-the Renais-
sance. And then were produced such men as Copernicus,
Tycho, Kepler, Galileo, Newton, who, with their contemporaries
and colleagues, in a space of hardly more than a century, com-
pletely broke the "spell" of Greek prestige and made possible
the beginnings of modern experimental science. In so far as
the heliocentric theory completely revolutionized man's concep-
tion of the universe, and his place in it, it is quite correct to
regard the work of Copernicus as beginning a new era in scientific
thought. But had it_. not been for other discoveries coming
immediately after Copernicus, such as the telescope, Kepler's
laws, Galileo's famous experiments on falling bodies, and many
others, it is quite possible that the theory of Copernicus would
have had the same fate as that of Aristarchus centuries earlier.
It is, therefore, fitting to regard the birth of the Copernican
theory as closing the first period in the history of physics.
13. The Copernican System.-Copernicus was born in Thoru,
in Polish Prussia, on Feb. 19, 1473. (He is, therefore, to be
thought of as a younger contemporary of Columbus.) At the
age of eighteen, he entered the University of Cracow, remaining
there three years, during which he studied, among other subjects-,
SEC. 13] THE COPERNICAN SYSTEM 15
mathematics and astronomy. In 1496, he entered the Uni-
versity of Bologna, where he studied astronomy under the famous
di Novara, who, by his support of the Pythagorean system of the
universe as against the Ptolemiac, probably gave Copernicus
the germ of the idea which later made him famous. After
studying at Padua and at Ferrara, from which he received the
degree of Doctor of Canon Law in 1503, he returned to his
boyhood home, Ermeland, wh~re he ljved most of the rest of
his life, as one of the leading canons at Frauenburg, near the
mouth of the Vistula. "He was a quiet, scholarly monk of
studious habits and with a reputation that drew to him earnest
students who received viva voce instructions from him." 1 He
died May 24, 1543.
Copernicus' chief contribution to science is his famous "De
Revolutionibus Orbium Coelestium," in the preparation and
revision of which he spent nearly 30 years, and which was printed
just at the close of his life. In this book, Copernicus boldly
proclaims the heliocentric world system, as against the Ptolemaic.
How he came to devise the new system we are not told; we can
only surmise. But it is almost certain that the idea of a moving
world was not original with him. He was doubtless familiar
with the theories of Aristarchus, 2 and, indeed, in the preface of
"De Revolutionibus," after pointing out the extremely unsatis-
factory status of astronomical theories, he says:
Wherefore I took upon myself the task of re-reading the books of all
the philosophers which I could obtain, to seek out whether any one had
ever conjectured that the motions of the spheres of the universe were
other than they supposed who taught mathematics in the schools.
He found that Cicero referred to the theory of Nicetus that the
earth moved; that "according to Plutarch, certain others had
held the same opinion." These "others" were Philolaus,
Heraclides, and Ecphantus. "When from this, therefore," he
continues, "I had conceived its possibility, I myself also began
to meditate upon the mobility of the earth." Copernicus was,
also, familiar with the work of two astronomers of the previous
generation, Piirbach (1423-1461) and Regiomontanus (1436-
1476), who, although they themselves accepted the Ptolemaic
theory, nevertheless convinced themselves that the actual posi-
1 LODGE, OLIVER: "Pioneers of Science." Quotation from "Copernicus
and the Earth's Motion."
! HEATH: "Aristarchus of Samos," p. 301.
16 HISTORICAL SKETCH 1 FIRST PERIOD [CHAP. I
tions of the planets in the sky differed very considerably from
the places computed from the existi;ng astronoJilical tables based
on Ptolemy. Further, and from the historian's point of view
very important, impetus was given to the study of astronomy by
the ever increasing demands of navigators, who were obliged to
trust entirely to the stars and the compass and who, therefore,
required as perfect a theory as possible of the motions of the
heavenly bodies.
Copernicus perceived that, by assuming that the earth is a
planet, like the others, and that all the planets move round the
sun, a great simplification, both philosophical and mathematical,
could be made with regard to the world system. He could, thus,
easily account for the seasons and for the apparent retrograde
motion, at times, of the planets. The rotation of the earth on
its axis caused the apparent daily motion of the sun, moon, and
stars, and he pointed out that, probably, the stars were too far
away for any (annual) motion of the ,earth to affect their apparent
places. He gave the correct order of the planets from the sun
outward.
In the absence of data to the contrary, he assumed that the
planets moved in uniform circular motion around the sun,
instead of in ellipses, as Kepler, using the accurate observations
of Tycho, later showed.
· Whatever the system as proposed by Copernicus lacked quanti-
tatively, it was correct, in its main outline, qualitatively. Its
reasonableness set a few men thinking and did much to usher in
a new era in science, an era which could come only when truth
could have the opportunity of standing alone, unaided or unhin-
dered by the "authority" of 2,000 years.
CHAPTER II
HISTORICAL SKETCH, SECOND PERIOD (1660-1800 A.D.):
THE RISE OF THE EXPERIMENTAL METHOD
1. The Growing Dissatisfaction with Authority.-Copernicus,
by announcing his heliocentric world system, may be said to
have spoken the prologue preceding the rise of the curtain on
modern physics revealing among the first actors, Galileo, Tycho
and Kepler. But certain details of the stage setting must not
be forgotten. A few men were beginning to show dissatisfaction
with existing conditions in science, dominated as it then was by
ancient Greek dogma and "authority." J
Two centuries before Copernicus, Roger Bacon (1214-1294),
British philosopher and scientist and a monk of the Franciscan
Order, taught that in order to learn the secrets of nature we must
first observe. He believed in mathematics and in deductive
science, but he clearly realized that only as these were based on
observed phenomena and tested by experiment could useful
knowledge result. ''In an age where experimenting was sure to
cost a man his reputation and was likely to cost him histlife, he
insisted on experimenting and braved all risks;" 1 with the result
that he spent 24 of the last 37 years of his life in prison-one of
the many "martyrs of science."
Then there was Leonardo da Vinci (1452-1519), Italian painter,
architect, sculptor, engineer, and philosopher, whose greatness as
a scientist has come to be appreciated only in recent years, for
his works were left in manuscript form and were probably not
widely known among his contemporaries-for which reason his
influence on early science is comparatively insignificant. Never-
t heless, da Vinci is generally regarded as one of the world's
greatest thinkers. His belief in the value of experiment is worthy
of the twentieth century: ''Experience,'' he writes'' never deceives;
it is only our judgment which deceives us." Or, again: "Before
making this case a general rule, test it by experiment two or three
: °WHITE, ANDREW D.: "History of the Warfare of Science with
Theology."
17
18 HISTORICAL SKETCH, FIRST PERIOD [CHAP. II
times, and see if the experiment produces the same effect." 1
Although expressed in the vague language of his time,
some of da Vinci's ideas concerning what we now refer to as
"force ' " "inertia ' " "acceleration ' " the "laws of motion ' " etc .,
were qualitatively correct. For example, as regards velocity
and acceleration, he says:
In the air of uniform density, the heavy body which falls acquires at
each stage of time a degree· of movement more than the degree of the
preceding time and likewise a degree of velocity more than the degree
[of velocity] of the preceding time. Then to each quantity doubled in
time the length of the descent is doubled, likewise the velocity of the
movement.
That is, in the case of a falling body, he held correctly that the
velocity is proportional to time of fall, but he did not grasp the
significance of this in connection with the total distance covered.
He wrote about levers, inclined pl;anes, pulleys, flight of birds, and
mechanical flight. Concerning perpetual motion, he wrote:" Oh,
speculators on perpetual motion, how many vain projects of the
like character you have created! Go and be the companions of
the searchers after gold." Rejecting the Ptolemaic theory, he
held that "the sun does not move." That he was not persecuted,
or even burned at the stake, as was Bruno, a century later, for
holding such revolutionary and, therefore(!), heretical, views is
probably due to the fact that his doctrines were given so little
publicity; for, holding no academic position, he did not teach,
and he published nothing. Could his successors have known
of his work, how great an impetus might have been given to
the growth of science! We shall see a similar situation in the
work of Henry Cavendish in the latter part of the eighteenth
century.
The unrest typified by the doctrines of Copernicus, Roger
Bacon, Da Vinci, and others culminated in the work of Galileo
and of Tycho and Kepler. Galileo is usually thought of as a
physicist, and Tycho and Kepler as astronomers. Although
Tycho was the oldest of the three, we shall discuss Galileo first.
2. Galileo Galilei (1564-1642).-Galileo is widely, and quite
correctly, regarded as the father of modern physics. To be sure,
physics had grandfathers and still more remote ancestors, but
1 HART, IvoR B.: "The Mechanical Investigations of Leonardo da Vinci.''
The Open Court Publishing Company, Chicago.
SEC. 2] GALILEO 19
none of them gave to physics, particularly to experimental
physics, so much as Galileo.
He was born in Pisa, Feb. 15, 1564-about the time of the
founding of the first permanent American_ settlement, St. Augus-
tine, Florida, in 1565. He was descended from a noble family,
and it is quite probable that he inherited from his father the
spirit of free inquiry which characterized his life. For in the
writings of the elder Galileo, who was well educated and was
an accomplished musician, appears the statement: "It appears
to me that they who in proof of any assertion rely simply on the
weight of authority, without adducing any argument in support
of it, act very absurdly."
At the age of twelve, Galileo was sent to the monastery of
Vallombrosa, near Florence, where he studied Greek, Latin,
Logic, and other "Humanities." He was apparently a brilliant
and a versatile student, for he excelled in classics and literature
and was something of a poet, musician, and art critic. He also
showed an aptitude for science and exhibited considerable
mechanical inventiveness.
It was decided that a medical career was more in keeping with
his intellectual attainments than were mercantile pursuits, and
young Galileo, at the. age of seventeen, was sent to the University
of Pisa to study medicine. It was here that he made his first
discovery and invention. One day, in 1581, he noticed the
regular oscillations of the great bronze (hanging) lamp, in the
cathedral at Pisa. Although the amplitude of these oscillations
became less and less, they were all performed in the same time,
as he determined by counting his pulse. Turning the process
around, he invented a "pulsometer," a ball-and-string (i.e.,
simple pendulum) device, whose length, when adjusted to
synchronism with the pulse, was a me81Sure of its frequency.
But the urge toward mathematics and science overcame the
pecuniary advantages of a medical career, and, eventually,
Galileo, at the age of twenty-six, became professor of mathe-
matics at Pisa, where he began a systematic investigation of
the mechanical doctrines of Aristotle. He soon demonstrated
experimentally to his own satisfaction that Aristotle was in error
in many of his assertions, and these errors he proclaimed ener-
getically from his professoriai chair. Now, to doubt Aristotle,
or any other ancient authority, was heresy, and Galileo soon
brought upon himself the enmity of many of those who should
20 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
have been his colleagues and supporters. To convince these
that, in at least one particular, it was easy to demonstrate that
Aristotle was in error, Galileo performed the most famous
experiment in all history, the dropping of two bodies of different
weight from the leaning tower of Pisa.
It had been pointed out above (Chap. I, Sec. 8) that Aristotle
had taught that a heavy body would fall more rapidly than a
light one. Galileo had, by experiment, proved this doctrine to
be wrong. He thought that he could convince his opponents
and, perhaps, shake their implicit faith in Aristotle, by an appeal
to their senses. So, one morning, he ascended the leaning tower
of Pisa, taking with him a 1-pound shot and a 10-pound shot
(some writers say 100 pounds). These he dropped simultane-
ously before the expectant multitude. They fell together and
struck the ground with a single thud. A few were convinced
that Galileo was right and that Aristotle was wrong. But the
vast majority rubbed their eyes, looked up their Aristotle, and
found, chapter and verse, that he said that the heavier body would
fall more quickly. Therefore Galileo must be wrong.
And then began a persecution which was to last Ga,Iileo' s
lifetime, increasing in severity as he grew older, and, finally,
resulting in imprisonment. To present the details of his stormy
life is far beyond the scope of this book. The reader is referred
to his biographers. 1
He was soon forced to quit Pisa and, in 1592, became Profes-
sor of Mathematics at the Venetian University of Padua, where
he remained 18 years, enjoying comparative liberty of thought
and teaching. His fame as a teacher spread all over Europe,
and his lectures were crowded. To this period belongs his
invention (1602) of the air thermometer; also, a military and
geometrical compass.
In October, 1608, a Dutch optician, Lipperhy, as a result of a
chance observation of an apprentice, had succeeded in "making
distant objects appear nearer but inverted" by looking through a
tube in which were mounted two spectacle lenses. News of
this invention-but not a sample of the instrument-reached
Galileo in June, 1609. Immediately grasping the principle
involved, he made a telescope and exhibited it in Venice "for
1 FAHIE, J. J.: "Galileo: His Life and Work" ( 1903).
BREWSTER, Sm DAVID: "Martyrs of Science" (1870).
LODGE, OLIVER: "Pioneers of Science."
PLATE 2.-Galileo.
(Facino paue 20)
SEC. 2] GALILEO 21
more than a month to the astonishment of the chiefs of the
republic." Within a short time, he made a better instrument,
with a magnifying power of 20 diameters, and by January, 1610,
a still more powerful one with a power of 30 diameters. 1
With this instrument, he made a number of the most funda-
mental discoveries. He observed the mountains on the moon
and made a reasonably close estimate of their altitude. He saw
that the number of fixed stars was vastly greater than could be
seen by the unaided eye and, thus, was able to explain the age-
long puzzle-the Milky Way. He saw that the planets appeared,
in his telescope, as luminous disks, while the stars still remained
points of light.
But the most important of these astronomical discoveries
was that of the moons of Jupiter. On Jan. 7, 1610, he directed
his telescope toward Jupiter and noticed three stars near the
planet, one on the west and two on the east. On the following
night, they were all on the west side and nearer to each other.
On Jan. 10, only two of the stars were visible, and they were
both on the east side! Galileo concluded that these "stars"
were really satelites of Jupiter, and, by Mar. 2, he had determined
their periods. These discoveries naturally made Galileo famous,
and he soon accepted an invitation to return to Pisa as '' First
Mathematician and Philosopher," at a very substantial increase
in salary, although at a sacrifice-and a very unfortunate one,
as it turned out-of his "academic freedom" in Padua. Con-
tinuing his astronomical investigations, he discovered the cres-
cent phases of Venus, sunspots and the rotation of the sun, the
faculm of the solar atmosphere, and the libration of the moon.
In 1612, he published his "Discourse on Floating Bodies."
At first, it seemed as if his fame had silenced all opposition
from the Church. But the support which his discoveries gave to
the hated Copernican theory and his vigorous attacks on Aris-
totelian philosophy roused his enemies to fury; with the result,
that, in 1615, he was hauled before the Pope and, under threat
of imprisonment and torture, was "enjoined . . . to relinquish
altogether the said opinion that the sun is the center of the world
and immovable . . . nor henceforth to hold, teach or defend it
m any way " Simultaneously, it was decreed that the
1 Galileo's telescopes were similar to the modern opera glass-a double-
convex (or plano-convex) object glass and a double-concave eyepiece.
Hence they ·had an erect image, unlike the Dutch instrument.
22 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
works of Copernicus "be suspended until they be corrected."
Galileo acquiesced in these decrees and was allowed to return to
Pisa, where he continued his researches along such lines as
would not give offense.
In 1623, one of Galileo's friends, Barberini, became Pope
Urban VIII, from whom Galileo received assurances of "pontif-
ical good will." Thereupon, thinking that the decree of 1615
would no longer be enforced, he began the writing of his great
book, "Dialogues on the Ptolemaic and Copernican Systems,"
which was published in 1632, under formal permission from the
censor, the "Master of the Sacred Palace," who "signed the
license with his own hand." 1
The form of these dialogues is ingeniously contrived to abide
by the letter of the decree of 1615. Three "characters" carry on
the discussion: Salviati, a Copernican; Simplicio, an Aristotelian;
and Sagredo, a witty, impartial, good-natured chairman. The
dialogues cover 4 "Days," during which the arguments for and
against each system are set forth with apparent impartiality
and without reaching any stated conclusion. Nevertheless,
the general effect of the book was "a powerful plea for
Copernicanism. '' 2
But, in spite of its enthusiastic reception by the public, its
form did not deceive his enemies, who were now determined that
he must be silenced. But how could this be accomplished as
long as the Pope was Galileo's friend? Very easy: Make the
Pope an enemy of Galileo! This was very effectually accom-
plished by representing (of course, falsely) to the Pope that the
Simplicio of the dialogues, whose ignorance was very apparent,
was simply a caricature of the Pope himself. In spite of the
absurdity of the argument-for Galileo would hardly have
risked offending Urban VIII, his one friend in the Church-the
Pope's weakness, vanity, arrogance, and ambition so beclouded
his senses that he was easily convinced that Galileo "had made
game of him.'' Whereupon he was ready to join Galileo's
enemies in persecuting that great scientist, ostensibly for "the
safety of the Church and the vindication of its decrees."
This tragic incident-the alienation of the Pope's friendship-
is illustrative of the fact that, in the seventeenth just as in the
1 FAHIE: "Galileo and His Works." Letter from Galileo to Cioli, Mar. 7,
1631.
2 FAHIE, Zoe. cit.
SEC. 2] GALILEO 23
twentieth century, much of the "warfare between science and
theology''-as Andrew D. White calls it1-has been based upon
personal ambition and revenge rather than upon a sincere wish
to uphold theological doctrines. In 1612, the Pope-he was
then Cardinal Barberini-upon a receipt of a copy of the work,
"On Floating Bodies," had written Galileo:
I have received your treaties . . . and shall read them with great
pleasure, both to confirm myself in my opinions, which agree with yours,
and to enjoy with the rest of the world the fruits of your rare intellect.
In 1613, Barberini had complimented Galileo very highly for
his work on sunspots and, in 1620, had composed some verses,
celebrating Galileo's discoveries, which he sent to Galileo,
writing, "I propose to add lustre to my poetry by coupling it
with your renowned name." Late in 1623, Galileo published
"II Saggiatore," which he dedicated to Barberini, then Pope
Urban VIII, who was so delighted with it that he "had it read
aloud to him at table." 2 The Pope as Galileo's friend could
apparently not only applaud but also heartily concur in Galileo's
doctrines and teachings. But the Pope as Galileo's enemy could
join, with all his might, in declaring those same doctrines to be
so dangerous and inimical to the (supposed!) teachings of Holy
Scripture as to warrant bringing Galileo before the Inquisition,
which, some 30 years before, had burned Bruno at the stake.
These personal motives, which have been the cause of so much
opposition to new scientific advances, are to be found not only
in all branches of the Church, Protestant as well as Catholic,
but also even among scientists themselves, as we shall see later.
Intolerance, even when it is sincere, is to be condemned; but
insincere intolerance is to be despised. Yet, one cannot fully
comprehend the forward march of science unless one recognizes
the seriousness of obstacles of this kind which have had to be
overcome. Great as is the fame of Galileo, how much more
might he have accomplished if the energy which he was forced
to spend in overcoming opposition might have been directed
toward his researches. Without doubt, the giant strides which
are being taken by science in this twentieth century are made
possible, in part at least, by the freedom which the scientist
now enJoys. How absurd it would be for anyone to criticise the
1 WHITE, ANDREW D.: "The History of the Warfare between Science and
Theology."
2 FAHIE, Zoe. cit.
24 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
Bohr theory of atomic structure because it is mentioned neither
by Aristotle nor the Bible!
With this array against him, i,t was inevitable that Galileo
should be called before the Inquisition. He was now sixty-seven
years old, impaired in health and in spirit, if not in intellect.
Bowing to the inevitable because of the magnitude of the forces
arrayed against him, which forces included his former friend the
Pope, and perhaps with the memory of Bruno all too clear, he
followed the advice of his friends and indicated his "free and
unbiased" willingness to recant, to "abjure, curse and detest
the said heresies and errors and every other error and sect con-
trary to the Holy Church" and he further agreed "never more in
future to say, or assert anything, verbally or in writing, which
may give rise to a similar suspicion.''
Perhaps out of respect for his age and infirmities, perhaps
because the Pope remembered, with a bit of shame, his former
friendship, Galileo was not imprisoned and tortured but was
"detained" in Rome, under suspended sentence. Ultimately,
he was allowed to return to his home at Arcetri, although still as
a prisoner. Here, during the last years of his life, he prepared
and, in 1636, published his "Dialogues on the Two New Sciences" 1
(i.e., Cohesion and Mot.ion).
These dialogues on "Motion" sum up Galileo's earlier experi-
ments and his more mature deliberations. He states that "if the
resistance of the media be taken away, all matter would descend
with equal velocity." He correctly describes, and deduces the
formulre of, uniformly accelerated motion. He shows that the
path of a projectile is parabolic under the limiting conditions
(1) that the path is small compared to the dimensions of the
earth and (2) that the motion either takes place in vacuo or is
comparatively slow. He states that, if all resistance were
1 See the excellent translation by Crew and de Salvio.
Some writers have severely censured Galileo for yielding to the Inquisition.
They say that "had Galileo added the courage of a martyr to the wisdom
of a sage . . . science would have achieved a memorable triumph" (see
BREWSTER: "Life of Newton." Whatever opinions on this question one
may hold, one fact stands out indisputable: Had Galileo not yielded he
would surely have been cast into the dungeon, and would probably have been
burned at the stake. His working days would have been over. We should
not have had handed down to us these dialogues on Motion, so fundamental
to our modern physics.
SEC. 3] TYCHO AND KEPLER 25
removed, a body projected along a horizontal plane would
continue to move forever. And, indeed, he paved the way for
the enunciation, by Newton, of the famous three laws of motion,
the foundation of mechanics. He describes an attempt to
measure the velocity of light by two observers, a mile apart, one
of whom uncovers a lantern as soon as he perceives the other to
cover up his lantern. No greater elapsed interval was observed
at a distance of a mile than at short distances. Not convinced,
however, that light travels instantaneously, Galileo recom-
mended that the experiment be tried, with the aid of a telescope,
at greater distances. It is, perhaps, worthy of remark that the
"toothed-wheel" experiment of Fizeau, performed in 1849, is, in a
sense, similar to this method proposed by Galileo.
Galileo's sight began to fail, and, by 1638, he was totally
blind. Nevertheless, his mind continued active, and, in 1641,
with the aid of two assistants, Torricelli and Viviani, he prepared
some additional material for his dialogues.
He died Jan. 8, 1642, only a few months before the birth of that
other great physicist of this second period, Isaac Newton.
How much the modern world owes to Galileo, perhaps few of
the present day realize. Physicists, at least, should be
acquainted with the main details not only of his scientific work
but also of his life. The student is, therefore, urged to peruse
one of the several biographies and, also, to read some of Galileo's
writings, if not in the original Latin, then in translation. Even a
short time spent in following his deductions and in reading first
hand something which he wrote will prove both interesting and
valuable.
3. Tycho (1546-1601) and Kepler (1571-1630).-While Galileo
was carrying on his investigations in Italy, two other European
scientists, Tycho and Kepler, were making astronomical studies
of the most fundamental character. The work of these two
men is particularly interesting, not only because of its direct
bearing on the development of physics but also, more particularly,
because of the mutual dependence of the work of each upon that
of the other, a relation very common in present-day science.
Tycho was the experimentalist, the observer, who supplied the
accurate data upon which Kepler, the theoretically inclined,
built his theory of planetary motion. Tycho's observations
were, of course, of empirical value, but, without a Kepler to
build a theory from them, they would have attracted hardly
26 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
more than passing notice. Kepler, in turn, might have theorized
to his heart's content, but without the accurate data of a Tycho,
those theories would ultimately have shared the fate of Aris-
totle's. Sometimes, theory precedes; sometimes, experiment.
But neither can get far without the other.
Tycho Brahe was born of a noble family in Scania (the southern
part of Sweden, belonging at that time to Denmark), on Dec.
14, 1546. When just past twelve years of age, he was sent to
the University at Copenhagen, where he studied rhetoric and
philosophy, in contemplation of a career as statesman, for which
his uncle, who assumed responsiblity for Tycho's upbringing,
desired him to prepare. But his attention was directed toward
astronomy by a partial eclipse of the sun, in 1560, and, during
the remainder of his three years at Copenhagen, he studied
Ptolemy, Euclid, and allied authors.
In 1562, the uncle, t4inking to remove him from the influences
which were rapidly taking him away from the life of a statesman,
sent him to Leipzig, where, in spite of instructions to the con-
trary, Tycho, at times in secret, continued his astronomical
studies. He soon discovered the incorrectness of the current
astronomical tables. In August, 1563, occurred a conjunction
of Saturn and Jupiter, of which Tycho made careful observations.
During this time, he dabbled in astrology, because of the
then very prevalent belief that human affairs are materially
influenced by the position of the planets; just as vegetable life
depends on the position of the sun in the ecliptic (i.e., winter
or summer).
After several years of study and travel, during which time
he acquired a notable reputation as an astronomer, Tycho was
settled by King Frederick II of Denmark upon the Island of
Hveen, in 1575. Here the king built the celebrated observatory
of Uraniborg and guaranteed Tycho an income to enable him to
carry on his observations. In return for this assistance, Tycho
made astrological calculations for the royal family and acted as
scientific advisor. The observatory was equipped with the
most elaborate and beautiful instruments, and for over twenty
years Tycho remained at Uraniborg, making systematic obser-
vations of the planets, constructing a star catalogue, and accumu-
lating other astronomical data, always with the highest possible
order of accuracy and with the hope ultimately of reforming
astronomy.
SEC. 3] TYCHO AND KEPLER 27
In 1597, Tycho was compelled to leave Uraniborg, having
incurred the displeasure of the government, because, it was
alleged, he was careless in carrying out the various duties incum-
bent on him in return for his grants of income. In 1599, he
entered the service of the German emperor, Rudolph the Second
and, in due course, made plans to establish an observatory at
Prague, with his instruments from Uraniborg and a corps of
assistants. These plans were hardly more than well underway
when Tycho suddenly died, Oct. 24, 1601.
Among the assistants attracted to Tycho's observatory at
Prague was a brilliant young mathematician, Johann Kepler,
with whom Tycho had corresponded for several years. Kepler
was born at Weil, in Wurtenburg, Dec. 21, 1571. At the age
of twenty, he took the Master's degree at the University of
Tubingen, where, under the mathematician Mastlin, he became
acquainted with the theories of Copernicus.
, In 1594, Kepler accepted a lectureship at Gratz. This post
he held for five years, devoting most of his time to astronomy.
Among the questions which he attacked, initially without suc-
cess, was that of a possible relation between the periods of
revolution of the then-known planets and their distances from
the sun. Being compelled to leave Gratz because of certain
religious decrees, Kepler finally became an assistant to Tycho at
Prague, late in 1600. This association lasted only a few months,
being cut short by Tycho's sudden death. Whereupon Kepler
succeeded Tycho as principal mathematician to the emperor
and undertook the completion of several projects which Tycho
had underway. Chief among these were the new astronomical
tables based on Tycho's elaborate observations. After many
delays, these were finally published as the Rudolphine Tables,
in 1628, in honor of Emperor Rudolph, who had established
Tycho at Prague.
Keple'r remained at Prague until 1612, when he accepted a
professorship at Linz. This post he held almost until his death
in November, 1630.
The contrast between Kepler's life and that of Tycho is strik-
ing, indeed. Tycho was of noble birth, was rich, vigorous, of
great mechanical ingenuity and experimental skill. Kepler
was sickly, lacked experimental ability, and was, throughout his
life, harassed by poverty and misfortune. Yet he possessed a
speculative imagination and mathematical perception of very
28 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
high order. He believed thoroughly in the Copernican system,
which Tycho rejected for a system of his own in which the sun
with his family of planets revolved once per day around the
stationary earth, the planets (except the earth) revolving around
the sun. It is one of the dramatic situations in science that
Tycho's data on planetary motions, taken in support of his own
theory, became, in the hands of Kepler, the clinching argument
for the Copernican system.
At Prague, Kepler made a special study of the motion of Mars,
using Tycho's observations taken so carefully at Uraniborg.
At first, he tried to reconcile the various recorded positions of
the planet by assuming that the circumsolar orbits of both Earth
and Mars were circles. He tried various orbits in various relative
positions with respect to the sun. None worked. By resorting
to the Ptolemaic notion of epicycles and deferents, some improve-
ment resulted, but still the observed positions differed from the
computed, in some cases by as much as 8 minutes of arc. 1 Kepler
knew that Tycho's observations could not be in error by that
amount. Some new concept regarding planetary motion was
necessary.
All of these attempts required enormous labor, but still, in
spite of the seeming hopelessness of the problem, Kepler per-
severed in his quest for the right orbit. Some progress was made
when he gave up uniform circular motion and assumed that the
speed varied inversely as the planet's distance from the sun.
This assumption is his famous "second law:" that the radius-
vector from the sun to the planet describes equal areas in equal times.
It worked approximately, but still there were systematic errors
which exceeded the possible errors of observation.
Finally, he cast aside the last traditions of the Ptolmaic system,
namely, that the planets move in circles, and tried, first, an oval
path and, then, an ellipse, with the sun at one focus. At last,
his years of computations bore fruit. The path was an ellipse.
Theory and observation agreed! And one of the most important
and far-reaching laws in all science had been discovered, all
because of a discrepancy of 8 minutes of arc between observation
and theory! In fact, one of the striking things in the growth of
science, particularly physical science, is the fact that a very great
many fundamental advances have come about because of just
1 About one-fourth the angular diameter of the moon viewed from the
earth.
SEC. 3] TYCHO AND KEPLER 29
such discrepancies, frequently very small ones, between observa-
tion and theory. We shall have occasion to note this, time and
agam.
Another problem with which Kepler had been wrestling for
many years was that of the possible relation between the periods
of revolution of the several planets around the sun and the radii
of their respective orbits. Long before he joined Tycho, he had
speculated on this problem, possibly because he felt, instinctively,
that the motions of the planets must be the result of some uni-
versal "cause." Of course, he rejected, as nonsense, the prevail-
ing opinion that there were six planets because the number
"six" was sacred! And he reasoned that it was not an acci-
dental coincidence that whenever a gap appeared in the series
of periods (as between Mars and Jupiter), there was a cor-
responding gap in the series of radii of the several orbits.
His first "solution" of a part of the problem is almost as
fantastic as the sacredness of the number six. While lecturing
at Gratz, he noted that the orbits of Jupiter and Saturn bore to
each other about the same relation as a circle inscribed in an
equilateral triangle bears to the circumscribed circle. 1 Starting
from this "observation" and noting that, since there are six
planets, there are five spaces between their orbits, he concluded
that there must be some connection between these five spaces
and the five regular geometrical solids. After some speculation,
he hit upon the following scheme:
Start with a sphere whose great circle represents the earth's
orbit. Describe around this a dodecahedron, and circumscribe
around this a sphere. A great circle of this sphere represents
the orbit of Mars. By similar use of a tetrahedron around the
orbit of Mars, we get to Jupiter; and from Jupiter to Saturn, by
a cube. An icosahedron inside the earth's orbit gets to the orbit
of Venus; and an octahedron inside that gets to the orbit of
Mercury. The scheme was not exact so far as giving relative
values for the radii of the orbits, but Kepler thought that this
might be due to erroneous values of the several planetary
distances.
Kepler was delighted with this scheme and published it, in
1596, in a book, "Mysterium Cosmographicum." It was this
book which brought Kepler to the attention of Tycho, who, while
1
The latter ratio (ratio of diameters) is 2: 1, while modern astronomical
d2.ta gives the former about 1.85: 1.
30 HISTORIC.AL SKETCH, SECOND PERIOD [CHAP. II
expressing appreciation of Kepler's industry, nevertheless
advised him "not to build up abstract speculations concerning
the system of the world but rather first lay a solid foundation in
observations and then, by ascending from them, strive to come to the
cause of things." 1
Kepler may have followed this advice, for it was many years
later (1618) when, again with Tycho's observations at hand, he
hit upon the true relation between the periods of the planets
and the radii of their orbits, a relation now known as Kepler's
'' third law,'' that the squares of the times of revolution around the
sun are as the cubes of the mean orbital radii.
Here, then, are the three laws of planetary motion which
Kepler handed down to posterity and which, sweeping away
epicycles, deferents, and other remnants of the Ptolmaic system,
paved the way for modern astronomy:
1. The planets move round the sun in orbits which are ellipses,
with the sun at one focus.
2. The radius vector (from sun to planet) sweeps over equal
areas in equal times.
3. The squares of the periods of revolution of the planets
round the sun are proportional to the cubes of the (mean) radii
of their respective orbits, or, in algebraic symbols,
r2 =
R3 constant
where T is the period of the planet and R is the mean distance
from the sun.
But what makes the planets move? Why do the outer ones
go more slowly? Is there "one moving intelligence in the sun,
the common center, forcing them all around, but those most
violently which are nearest?" Kepler speculated long on this
question, but he had only a faint glimpse of the truth, a qualita-
tive idea, which Newton, with Kepler's laws and Galileo's re-
searches on motion before him, later was able to put in quantitative
form in the famous theory of universal gravitation. The nature
of Kepler's speculations are indicated by the following extracts
from the Introduction to his "Commentaries on the Motion of
Mars," published in 1609:
Every corporeal substance, so far as it is corporeal, has a natural
fitness for resting in every place where it may be situated by itself beyond
1Quoted by HART in "Makers of Science," p. 82. The italics are the
author's.
SEC. 3] TYCHO AND KEPLER 31
the sphere of influence of a body cognate with it. Gravity is a mutual
affection between cognate bodies towards union or conjunction, so that
the earth attracts a stone much rather than the stone seeks the earth.
. . . If two stones were placed in any part of the world 1 near each
other, and beyond the sphere of influence of a third cognate body, these
stones . . . would come together in the intermediate pojnt, each
approaching the other by a space proportional to the comparative mass
of the other. If the moon and earth were not retained in their orbits by
their animal forces or some other equivalent the earth would mount to
the moon by a fifty-fourth part of their distance and the moon fall
toward the earth through the other fifty-three parts . . . assuming
that the substance of both is of the same density. [Then follows a
hint as to the cause of the tides.] If the attractive virtue of the moon
extends as far as the earth, it follows with greater reason that the attrac-
tive virtue of the earth extends as far as the moon and much farther;
and, in short, nothing which consists of earthly substance anyhow
constituted although thrown up to any height, can ever escape the
powerful operation of this attractive virtue . . .
Unquestionably, Kepler's rough, qualitative concept of universal
gravitation bears to the completed theory so beautifully devel-
oped by N e-v?ton much the same relation as the crude ideas of
planetary motion proposed by Copernicus on the basis of the
suggestion of Aristarchus bears to the quantitative statements
contained in Kepler's three laws. Thus it is, that, in the growth
of an idea or law, the qualitative phase precedes the quantitative.
Sometimes, both phases come from a single mind; but very fre-
quently, as, for example, in the case of planetary orbits and of
gravitation, the completed law is the result of the efforts of several
individuals, perhaps men in different generations, working '' in
series." The "quantum jump" seems to apply to ideas as well
as to electrons!
In passing, it may be mentioned that Kepler made substantial
contributions to the field of optics. He applied the law of
refraction of light at small angles, previously discovered by Al
Hazen, 2 to the path of a ray through a lens. And he proposed
the Keplerian or astronomical type of telescope, in which a real
image is formed, thus making possible accurate measurements
by means of "cross-hairs" in the focal plane of the objective.
He studied atmospheric refraction as affecting the apparent
1 That is, the universe.
! That is, that for small angles the angle of incidence is proportional to
:::.e angle of refraction (see p. 14).
32 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
position of the heavenly bodies and worked out an approximate
formula to allow for this error for all positions, from zenith to
horizon. He understood clearly the principle of total reflection
and how to determine what we now call the "critical angle."
He computed the properties of hyperboloid refracting surfaces.
And he was the first to propose the meniscus type of lens. In
fact, his two works on optics, the "Supplements to Vitello's
Optics" (1604) and the "Dioptrics" (1611), constitute by far
the most systematic and original treatises on the subject up to
that time.
But, again, the inevitable question: "What greater things
could Kepler have accomplished if his whole life had not been
harassed by ill health, poverty, and misfortune?" At least,
one misfortune was fortunate. Kepler's expulsion from Gratz
led to his association with Tycho, without which he would never
have had access to that mine of data which, in the end, brought
undying fame.
4. The Experimental Method Spreads.-The impetus given
to science by that great trio, Tycho, Kepler, and Galileo, resulted
in an ever increasing number of investigators in the generations
which followed. Of great significance, too, is the fact that, at
about this time, there were formed, in several European centers,
various learned societies which brought together, for argument
and discussion, men of kindred interests. The Lincean Society
was founded in Italy, in 1603; the Royal Academy of Sciences,
in France, in 1666; and the Royal Society for the Advancement
of Learning, in England, in 1662. The continued improvement
of the art of printing further facilitated the diffusion of scientific
knowledge. Developments, in physics and mathematics, of the
most fundamental importance were at hand. Mention may be
made of a few of them.
Gilbert (1544-1603), an English physician, had, in 1600,
published his famous work, "De Magnete," based largely upon
his own experiments, in which he showed the fallacy of such
popular fancies as the belief that lodestones lost their magnetic
power when rubbed with garlic and regained it again when
smeared with goat's blood. He was the first to recognize that
the earth is a great spherical magnet, and he actually magnetized
a small sphere of iron and showed that it behaved as the earth
behaves. He studied frictional electricity and showed that
other substances than amber are active.
SEC. 4] EXPERIMENTAL METHOD SPREADS 33
Among other workers in magnetism may be mentioned Kircher
(1601-1680), who, by measuring the force required to pull a
piece of iron from either pole of a magnet, demonstrated the
equality of the two poles; Cabeo (1585-1650), who showed that
an unmagnetized iron needle, floated freely on water, would place
itself along the earth's magnetic meridian; and Gellibrand
(1597-1637), who discovered the secular variation of the magnetic
declination.
In the field of optics, there was Scheiner (1575-1650), who
studied the optics of the eye; Snell (1591-1626), who discovered
the true law of refraction; Descartes (1596-1650), who investi-
gated, mathematically, the form of the interface between two
media, so that a point source in one should be accurately refracted
to a point image in the other; Cavalieri (1598-1647), who gave
the correct formula for the focal length of a thin glass lens in
terms of the radii of curvature of the two sides; and Johannes
Marci (1595-1667), who, by his investigations of the phenomena
of refraction of light by a prism, in part anticipated Newton's
discovery of the true nature of the spectrum.
Studies in acoustics were not wanting. For example, Mer-
senne (1588-1648), after having investigated the laws of vibra-
ting strings, determined, in absolute measure, the frequency of
a tone. He also measured the velocity of sound by observing
the time interval between the flash of a gun and the arrival of
the report. Similar measurements were made by Gassendi
(1592-1655), who showed, also, that the velocity of the sound
was independent of pitch.
There was considerable activity in the study of fluids. One
of the best-known workers in this field was Torricelli (1608-
1647), who studied the flow of liquids from orifices, discovered
the principle of the barometer, and observed variation in baro-
metric height with altitude. Every student of elementary
physics is nowadays acquainted with the "Torricellian vacuum."
Working quite independently of Torricelli, Guericke (1602-1686)
invented the air pump. Pascal (1623-1662), continuing the
work of Torricelli on the barometer, measured the difference in
barometric height between the base and the top of a mountain,
correctly explaining the reason for the difference, and, later,
announced the famous principle of hydrostatics which bears his
name.
34 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
Not only was physics, as a subject, beginning to assume
definite form, but even the now classical subdivisions comprising
mechanics, light, sound, etc., were beginning to crystalize out.
Then came a man
. . . towering head and shoulders above all his contemporaries, a
veritable giant among the giants, a man whose intellect and whose
contributions to knowledge are incomparably greater than those of any
other scientist of the past, that prince of philosophers, Sir Isaac Newton. 1
The other "giants" referred to, contemporaries of Newton, are
such men as Boyle, Huyghens, and Hooke-names well known
to every physicist.
5. Sir Isaac Newton (1642-1727) .-Newton was born in the
little hamlet of Woolsthorpe, England, on Christmas Day, 1642,
less than a year after the death of Galileo. At the age of twelve,
he was sent to the public school at Grantham, but at first he
showed no exceptional aptitude for his studies. An interesting
event, however, may have been the cause of his "turning over a
new leaf.'' It is related that a schoolmate, larger than young
Newton and his superior in scholarship, gave him a kick as they
were going to school one morning. Newton, resenting the
insult, challenged his tormentor to a fight and, though physically
inferior, soundly thrashed his opponent. This victory set him
thinking. He had proved his physical superiority, but he was
still below his antagonist in scholarship. Thereupon, an '' educa-
tional" contest ensued, in which Newton was not only successful
but also actually rose to the highest place in the class.
At the age of fifteen, Newton was removed from school to
assist his widowed mother in running the family estate at Wools-
thorpe. But he had little taste for farming. Rather, he was
interested in studying, particularly mathematics, and in devising
various mechanisms. He made a water clock, waterwheels,
sundials, and a working model of a windmill. And his uncle,
one morning, found him under a hedge studying mathema-
tics, when he should have been farming. Thereupon, Newton's
mother wisely decided that an educational career was more
suitable for her son, and he was sent back to school and, ulti-
mately, to Cambridge, which he entered in 1661.
Of his studies at Cambridge comparatively little is known,
save that he developed a fondness for mathematics and that
1 HART: "Makers of Science,"
PLATE 3.-Newton.
(Facing page 34)
SEC. 5] NEWTON 35
his reading included, among other works, Kepler's "Optics,"
Euclid, and Descartes' "Geometry." But his creative genius
soon began to appear, and before he took the degree of Bachelor of
Arts (January, 1665), he had discovered the binomial theorem,
developed the methods of infinite series, and discovered "flux-
ions," or the differential calculus.
Soon thereafter, an outbreak of the plague closed the Uni-
versity for some months, during which time Newton, at the family
estate at W oolsthorpe, began his speculations on the subject of
gravity, which, later, led to his enunciation of the inverse-square
law. It was here that the much-quoted "falling apple" episode
is said to have occurred, which is supposed to have given Newton
the basic idea of universal gravitation. Of course, a falling
apple may have started Newton's train of reasoning. Brewster,
in his "Life of Newton," seems inclined to credit the story and
states that he (Brewster) "saw the apple tree in 1814 and brought
away a portion of one of its roots" ( !) But Newton himself
makes no mention of the incident, and it seems far more probable
that, at Cambridge, he had read Kepler's qualitative discussion
of the general principle of gravitation (quoted on p. 31) in the
treatise on Mars. Certainly, Newton was familiar with the three
laws of planetary motion which Kepler announced in this book.
In 1667, Newton returned to Cambridge as Fellow of Trinity.
Two years later, at the age of twenty-six, he was appointed
Lucasian Professor of Mathematics, a chair which he held for
nearly 30 years. In 1703, finding his ever increasing activities
too onerous, he resigned his professorship, to devote himself to
his duties as Master of the Mint, to the scientific work of his
contemporaries, and to defending his own work against the
attacks of jealous rivals. In this same year, he was elected
President of the Royal Society, an office to which he was reelected
annually during the remaining 25 years of his life. In 1705, he
was knighted by Queen Anne.
Most of Newton's important scientific work was done before
he vacated the professorship in 1703, although he remained
thereafter "a power of the first magnitude in the world of
science." In his later years, he devoted much time to theo-
logical studies, particularly questions of Biblical criticism and
chronology. He died Mar. 20, 1727, at the ripe old age of
eighty-five. Throughout his life, he shunned publicity and
36 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
retained a modesty and simplicity which is indicated by a
sentiment uttered shortly before his death:
I do not know what I may appear to the world, but to myself I seem
to have been only like a boy playing on the seashore, and diverting
myself in now and then finding a smoother pebble or a prettier shell than
ordinary, whilst the great ocean of truth lay all undiscovered before me.
Any brief account of Newton's work must inevitably give a
very inadequate impression of his contributions to science. The
student is urged to read some of his writings first hand or, at
least, some extensive biographical discussion of his life and
work. We can here refer only to a very few of his researches on
optics and on mechanics.
Newton's work on optics arose out of an attempt to improve
lenses. The lenses made by Galileo and his immediate successors
all had spherical surfaces. Optical instruments made with such
lenses showed increasingly indistinct images as the powers of
those instruments were increased. And there seemed to be an
upper limit to the power of telescopes consistent with good
definition.
The inability of a lens with spherical surfaces to bring parallel
rays to a point focus was early recognized. In 1629, Descartes
had shown that lenses with hyperbolic or, under certain condi-
tions, parabolic surfaces should be free from the defect which we
now call "spherical aberration"; and many schemes were pro-
posed for grinding lenses with such surfaces. Newton found,
however, that such lenses produced only a very slight improve-
ment in the image, and he conjectured that, perhaps, the trouble
lay not in the lens but in the light itself.
Accordingly, he procured a prism of glass and, placing it over
a hole }i inch in diameter, through which sunlight was shining
into a darkened room, he observed the "very vivid and intense
colors" produced on a wall some 20 feet distant. Newton was
surprised to find that this "spectrum," as we now call it, was so
much longer than it was wide (13;!,i by 2J'S inches). The width
subtended at the hole an angle (about 31 minutes) correspond-
ing exactly to the sun's angular diameter. But the length-13;!,i
inches-could not be so explained. ·unaware of the true explana-
tion, he made various surmises as to the origin of the colors, such
as the varying thickness of the prism from apex to base, the
unevenness of the glass, a curvilinear motion of the light after
SEC. 5] NEWTON 37
leaving the prism, etc. One by one, experiment proved these
hypotheses wrong.
Finally, he isolated one ray, or "color," after another, by
suitable screens, and caused it to pass through a second prism.
In this way, he could measure the refrangibility of each ray.
And he found that the refrangibility increased from red to violet;
that, therefore, the first prism simply "sorted out" the colors,
which, in combination, made "white" light. In other words,
so-called "white light" is made up of the spectral colors-a very
elementary concept to us of the twentieth century but very new
and of far-reaching importance in 1666.
Newton at once saw that this phenomenon, this dispersion
of light, was the cause of his failure to effect any substantial
improvement in telescopes by use of paraboloidal lenses. "Chro-
matic" aberration could not be eliminated by mere change of
lens form.
Accordingly, he turned his attention to the reflecting type of
telescope, one form of which had been described by James
Gregory, in 1663. With his own hands, Newton made several
small instruments, one of which was presented to the Royal
Society in 1671. These telescopes do not seem to have been put
to any important scientific use, and, as Newton made no more,
the development of reflectors was dropped for 50 years.
Indeed, the development of refracting telescopes was retarded
many years because of Newton's singular error in assuming, on
the basis of a hurried experiment, that, in different media, dis-
persion was always proportional to refracting power. As noted
above, the power of refracting telescopes had reached the upper
limit set by chromatic aberration. And, in 1684, Newton
declared "the improvement of telescopes by refraction to be
desperate." Although, in the same year, Gr~gory had sug-
gested that, in imitation of the human eye, the object glasses
of telescopes might be composed of media of different density,
yet Newton's incorrect assumption seemed to make achromatic
telescopes an impossibility. And it was not until 1730, that
Hall, by a suitable combination of crown and flint glasses made
several achromatic combinations.
But Hall published no account of his work. Ne~Jrly 40 years
later, the principle of the achromatic combination was redis-
covered by Dolland, who, in 1758, presented to the Royal Society
a paper on the dispersive powers of different substances. He
38 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
was, thus, led to try combinations of crown and flint glasses,
with such success that, in spite of the previous work of Hall,
Dolland was granted a patent on achromatic lenses-an inven-
tion within the grasp of Newton, three-quarters of a century
before.
Newton's theories concerning the nature of light are of great
historical interest. And much has been written concerning
the extent to which he is supposed by some to have retarded the
development of optics, by espousing the corpuscular theory as
against the wave theory of his contemporaries, Huyghens
(1629-1695) and Hook (1635-1703). Accordingly, it may be of
interest to point out, by a quotation or two, that Newton,
apparently, was by no means dogmatic in his support of the
corpuscular theory and that later writers may have taken him, in
this regard, more seriously than he intended.
In a communication to the Royal Society, in 1675, concerning
'' An Hypothesis Explaining the Properties of Light'' Newton
states:
I have here thought fit to send you a description . . . of
this hypothesis . . . though I shall not assume either this or any other
hypothesis, not thinking it necessary to concern myself whether the
properties of light discovered by men be explained by this or any other
hypothesis capable of explaining them; yet while I am describing this, I
shall sometimes, to avoid circumlocutation . . . speak of it as if I
assumed it.
He then proceeds to describe "an retherial medium, much of
the same constitution with air but far rarer, subtiler and more
strongly elastic" but which vibrates "like air, only the vibrations
[being] far more swift and minute." This "aether"
. . . pervades the pores of crystal, glass, water and other natural
bodies yet it stands at a greater degree of rarity in those pores than in
the free mtherial spaces, and at so much a greater degree of rarity as the
pores of the body are smaller . 1
1 That is, the ether is less dense in transparent bodies than in free space.
Newton apparently makes this assumption by analogy with his erroneous
conception of the cause of the rise of liquids in capillary tubes. He assumes
this rise to be due to a reduced pressure of the air within the tube: the
smaller the tube the higher the liquid rises, and therefore the "greater
the rarity." The pores of transparent bodies must be very small. Hence,
the attenuation of the ether within them must be very great.
SEC. 5] NEWTON 39
He then supposes that
. . . light is neither mther, nor its vibrating motion, but something
of a different kind propagated from lucid bodies. They that will may
suppose it an aggregate of various peripatetic qualities. Others may
suppose it multitudes of unimaginable small and swift corpuscles of
various sizes springing from shining bodies ". . . and continually
urged forward by a principle of motion which in the beginning acceler-
ates them, till the resistance of the aetherial medium equals the force
of that principle much after the manner that bodies let fall in water are
accelerated till the resistance of the water equals the force of gravity.
Further, "light and mther mutually act upon one another, mther
in refracting light and light in warming mther."
Refraction is accomplished in this wise: When a beam of light
passes obliquely from, say, glass to air, it passes from rare ether
to denser ether through a transition layer of increasing density.
During the passage through this layer, the denser ether beyond,
perhaps by repulsion, causes a '' continual incurvation '' of the
beam of light toward the rarer ether; i.e., the ray is "bent from
the perpendicular,'' as we now say. The reverse occurs in
passing from air to glass. Evidently, the light, whatever it is
( !) , should go more rapidly in glass than in the air.
Just as bodies of various sizes . . . do by percussion excite sounds
of various tones . . . so, when the rays of light, by impinging on the
stiff refracting superficies, excite vibrations in the mther, those rays,
whatever they be,1 as they happen to differ in magnitude, strength, or
vigour, excite vibrations of various bigness,
giving rise to a sense of color. This theory, as proposed by
Newton, is a combination of the undulatory theory and the
corpuscular theory. .
Nearly 33 years later (1704), Newton published his optical
researches in book form in his well-known "Opticks." The
third edition of this appeared in 1721 and 1 without doubt, con-
tains his mature judgment on the subject. The first sentence of
this book reads: '' My Design in this Book is not to explain the
Properties of Light by Hypotheses, but to propose and prove
them by Reason and Experiment." He then, in some 300 pages,
gives his researches on refraction, reflection, colors of thin plates,
etc., and he concludes the book by "proposing only some queries
1 The italics are the author's.
40 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
in order to further search to be made by others.'' Among these
''queries'' are the following:
1. Do not bodies act upon light at a distance, and by their action
bend its rays . . . ?
5. Do not bodies and light act mutually upon one another, that is to
say, bodies upon light tn emitting, reflecting, refracting and inflecting 1
it; and light upon bodies for heating them, and putting their parts into a
vibratory motion wherein heat consists?
12. Do not the rays of light in falling upon the bottom of the eye
excite vibrations in the tunica retina; which vibrations, being propagated
along the solid fibres of the optick nerves into the brain, cause the sense
of seeing?
28. Are not all hypotheses erroneous in which light is supposed to
consist in pression or motion propagated through a fluid medium?
If light consists only in pression propagated without actual motion, it
would not be able to agitate and heat the bodies which refract and
reflect it, and . . . it would bend into the shadow. For pression or
motion cannot be propagated in a fluid in right lines beyond an obstacle
. . . but will bend and spread every way into the quiescent medium
which lies beyond the obstacle.
29. Are not the rays of light very small bodies emitted from shining
substances? For such bodies will pass through uniform mediums in
right lines without bending into the shadow, which is the nature of rays
of light . . . Pellucid substances act upon the rays of light at a dis-
tance in refracting, reflecting and inflecting them, and the rays mutually
agitate the parts of those substances in heating them; and this action at
a distance very much resembles an attractive force between bodies.
These quotations from Newton's "Queries" are illustrative of
his speculations concerning the nature of light. That they are
not intended as dogmatic assertions in the form of questions is
indicated by the two following quotations which raise queries
quite opposed to each other. In Query 18, after pointing out
that a thermometer suspended in a vacuum in a cold room will,
upon being transferred to the warm room, soon come to the
temperature of the warm room, he says:
Is not the heat of the warm room convey' d through the vacuum by
the vibrations of a much subtiler medium than air, which after the air
was drawn out remained in the vacuum? And is not this medium the
same with that medium by which light is refracted and reflected and by
whose vibrations light communicates heat to bodies? . . . And do
not hot bodies communicate their heat to contiguous cold ones by the
1 That is, diffracting.
SEC. 5] NEWTON 41
vibrations of this medium? . . . And doth it [the medium] not readily
pervade all bodies? And is it not (by its elastick force) expanded
through all the heavens?
Apparently, heat, but not light, might involve vibrations in
this medium. But later on, in Query 28, after discussing the
retardation of bodies moving through a fluid and pointing out
that the motion of the planets is '' regular and lasting,'' he says:
A dense fluid can be of no use for explaining the phaenomena of
nature, the motions of the planets and comets being better explain' d
without it. It serves only to disturb and retard the motions of those
great bodies, and make the frame of nature languish: and in the pores of
bodies it serves only to stop the vibrating motions of their parts, wherein
their heat and activity consists. And as it is of no use, and hinders the
operations of nature, and makes her languish, so there is no evidence for
its existence and therefore it ought to be rejected. And if it be rejected
the hypothesis that light consists in pression and motion propagated
through such a medium is rejected with it.
Then follows Query 29, quoted above.
Apparently, Newton was frankly raising questions. For, in
one of the concluding paragraphs of the books, he says:
In this third Book 1 I have only begun the analysis of what remains
to be discovered about light and its effects upon the frame of nature,
hinting several things about it and leaving the hints to be examined and
improved by the farther experiments and observations of such as are
inquisitive. And if natural philosophy in all its parts, by pursuing this
method shall at length be perfected, the bounds of moral philosophy will
also be enlarged.
But if Newton was only "leaving hints to be examined," he
left some very profound ones. For, in Query 30, he asks:
30. Are not gross bodies and light convertible into one another, and
may not bodies receive much of their activity from the particles of light
which enter their composition? . . .
The changing of bodies into light and light into bodies, is very con-
formable to the course of nature which seems delighted with
transmutations . . .
We begin, perhaps, to see an answer to this query in the modern
doctrine of the equivalence of matter and radiant energy-a
theory which accounts for solar and stellar radiation on the basis
of an actual "transmutation" of matter. The very fact that
Newton raised this question, particularly in such concrete form, is
1 That is, the third section of the book on "Opticks ."
42 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
an indication of the searching power of his intellect. And his
willingness to leave the arntwer "to such as are inquisitive," who
can base their conclusions on "farther experiments and observa-
tions'' is abundant proof of his sound scientific judgment.
Newton's researches on optics alone would have given him a
high rank, perhaps even premier place, among the scientists of
his time. But of still greater value was his work in mechanics.
In announcing that "every particle of matter in the universe
attracts every other particle with a force inversely proportional
to the square of the distance between the two particles''; in
showing that this one universal and comparatively simple law
governs not only the motions of the planets round the sun and
of the satellites round their planets but, probably, also the
relative motions of all the heavenly bodies, Newton gave to the
world a truth the importance of which in all branches of human
thought can hardly be overestimated. Of value to science, of
course, from microphysics to macrophysics. But consider the
effect on man's concept of nature and of his relations thereto of
realizing-indeed, of having proven to him, for the first time-
that the physical universe is governed by law, not by caprice; and
if the physical universe, why not the biological universe, even
the moral universe!
But how did Newton come to these discoveries? It is an
interesting story.
Newton early appreciated the correct quantitative significance
of the qualitative statements regarding motion, put forth by
Galileo, and, as a result, he formulated the three laws of motion
which bear his name. He was certainly familiar with Kepler's
quantitative laws of planetary motion and, probably, also, with
Kepler's speculations regarding gravitation as the cause of these
motions. It remained to put this "power of gravity" in quanti-
tative form. Accordingly, Newton attacked the problem of
finding out a law of gravitational attraction between two bodies,
as the sun and a planet, which would result in Kepler's third
law, namely, that the squares of the periods of rotation of the
planets round the sun are proportional to the cubes of their mean
distances from the sun.
The gist of Newton's reasoning is as follows: Suppose the
attraction between sun and planet to vary as the inverse xth
power of the distance between them. That is, let
F =GMm
{I rx
SEC. 5] NEWTON 43
where Fg is the force of attraction, G is some constant, M and
m the mass of sun and planet, respectively, and r their distance
apart. The centripetal1 force Fe required to keep the planet in
motion in its (nearly) circular orbit is
mv 2
Fe=-
r
where v is the speed of the planet in its orbit. Assuming the
gravitational attraction to be the origin of the centripetal force,
it foJlows that
And since, letting T equal the period of revolution,
21rr
v=-
T-
we have
G Mm _ 41r 2mr
--:;:x--r2
This can, at once, be written in the form
r2 41r2
rx+ 1 - GM
For a given family of planets around a central body of mass M,
T 2 /rx+ 1 is, therefore, constant for the several planets and is inde-
pendent of their masses. But Kepler's third law, based on obser-
vation, states that for the several planets T 2/r3 is constant. It
follows at once that
x = 2
and a gravitational attraction varying as the inverse square of
the distance gives this third law of planetary motion.
Newton saw that a test of this inverse-square law could easily
be made by comparing the acceleration of the moon toward the
earth, with the acceleration of free falling bodies at the surface
of the earth. He had shown that toward an external body a
sphere acts as if all its mass were concentrated at its center. It
was known that the distance between the moon and the earth's
center is about sixty times the earth's radius. By the inverse-
square law, therefore, the moon should "drop" toward the earth,
in 1 second, }13 0 2 as far as a body at the surface of the
1 Both Newton and Huyghens were familiar with acceleration toward the
center, of bodies moving in uniform circular motion. Newton's second law
of motion at once gave the law of centripetal force.
44 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
earth drops in 1 (i.e., the first) second. The latter distance
being, from observations on falling bodies, 16 feet, the former
should be 1~13 0 2 feet, or 16 feet in 1 minute.
But the acceleration of the moon could be determined directly
from the expression
v2
a=-
r
r
=41r2-
r2
where Tis the period of the moon's motion around the earth and
r is the radius of the moon's orbit. Now, r is equal to sixty times
the earth's radius, which was taken as 3,436 miles, on the then
common assumption that a degree of latitude is 60 miles. On this
basis, the moon is found to "drop" 13.9 feet toward the earth in
a minute, instead of 16 feet, as should be the case if the inverse-
square law were obeyed. The two results did not agree. Was
the inverse-square law wrong?
Newton was twenty-three years old at the time, and he
abandoned the work, not mentioning to anyone either his results
or his disappointment. Some years later, he learned of a
new and more accurate determination of the length of a degree
which had been made by Picard, who found, not 60, but more
nearly 70, miles. Newton, apparently in 1682, on the basis of
Picard's new value for the length of a degree, revised his computa-
tions on the moon's acceleration and, to his great joy, found that
it falls toward the earth 16 feet in a minute, just as predicted
by the inverse-square law. At last, he had discovered the true
law of gravitation. On the basis of this law, he could now derive
all three of Kepler's laws. Theory and observation checked
perfectly.
These results, together with some propositions on the motion
of the planets, were communicated, in 1683, to the Royal Society,
which requested permission to publish Newton's complete
researches on the subject of motion and gravitation. Newton
consented to prepare the material for publication, and, in 1687,
appeared the first edition of the "Principia," or, in full, PHILO-
SOPHIAE NATURALIS PRINCIPIA MATHEMATICA (Mathematical
Principles of Natural Philosophy) '' without exception the most
important work in natural philosophy extant." 1
1 HART: "Makers of Science."
SEC. 5] NEWTON 45
The original is in Latin, but English translations are available.
To these the reader is referred. The treatise is divided into
three books, the subject matter of each being presented by
propositions, theorems, and corollaries, the famous three laws
of motion being assumed as axioms. The first two books deal
with general theorems concerning the motions of bodies. For
example, Proposition XI of Book I solves the problem:
If a body revolves in an ellipse, it is required to find the law of the
centripetal force tending to the focus of the ellipse.
The solution is the inverse-square law. Proposition LXXI
proves, assuming the inverse-square law for small particles:
that a corpuscle placed without a [thin] spherical shell is
attracted toward the center of the shell with a force reciprocally pro-
portional to the square of its distance from that center.
Book III applies these general propositions to "demonstrate
the frame of the System of the World." For example, Proposi-
tion VI states:
. . . that all bodies gravitate towards every planet; and that the
weights of bodies towards any (given) planet, at equal distances from
the center of the planet are proportional to the quantities of matter
which they [i.e., the bodies] severally contain.
The entire treatise is characterized by the exposition of the
Principle of Universal Gravitation and its ramifications, without,
however, as Newton very carefully points out, attempting any
hypotheses as to the cause of gravitation.
He says:
To us it is enough that gravity really does exist, and acts according
to laws which we have explained, and abundantly serves to account for
all the motions of the celestial bodies and of our sea.
Of Newton's invention of the method of fl.uxions (i.e., the
calculus), of his very interesting miscellaneous writings, par-
ticularly those on theology, of the many controversies with his
contemporaries into which he was unwillingly drawn in defense
of his scientific work, we cannot take space to write. In urging
our readers to make further study of the life and works of this
renowned physicist, we may quote the words of his chief biog-
rapher, Sir David Brewster, who wrote:
46 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
The name of Sir Isaac Newton has by general consent been placed at
the head of those great men who have been the ornaments of their
species.
6. Newton's Contemporaries.-If we except the publication
and revision of his scientific writings, the productive period of
Newton's life ended about 1700. One finds his biography so
full of interest and inspiration that one is tempted to discuss,
similarly, the work of his contemporaries, themselves eminent
scientists-the Honorable Robert Boyle (1627-1691), "saint
and scientist," whose researches on the atmosphere have been
immortalized by '' Boyle's Law"; Huyghens (1629-1695), whose
wave theory of light was to triumph a century and a half later
over the corpuscular theory; Robert Hooke (1635-1703), pro-
ponent of the undulatory theory and originator of "Hooke's
law" of elasticity; Leibnitz (1646-1716), whose calculus ulti-
mately replaced Newton's fluxions. But, remembering that the
main business of this book is modern physics, we must pass on
to a rapid review of the developments of physics during the
eighteenth century.
7. Mechanics 1 during the Eighteenth Century.-Newton's
work had so thoroughly covered the fundamentals of mechanics,
that, during the eighteenth century, there were no new dis-
coveries of note. Rather, the period is characterized by the
extension of the principles laid down by Newton and by their
application to all sorts of problems. Among the prominent
workers during the eighteenth century, we find such names
as Daniel Bernoulli (1700-1782), who worked on hydrodynamics,
the kinetic theory of gases, and the transverse vibrations of
rods; Euler (1707-1783); Clairaut (1713-1765); D'Alambert
(1717-1783); and Lagrange (1736-1813); all students of theo-
retical mechanics, whose researches paved the way for important
advances made during the nineteenth century. One item of
note is the invention, about 1780, of the Atwood's machine by
George Atwood (1746-1807).
8. Heat during the Eighteenth Century.-Among the impor-
tant advances in heat during the eighteenth century may be
mentioned the development of thermometers and thermometric
scales and the discovery and study of latent heats and specific
1 The reader will find MACH'S "The Science of Mechanics" very valuable
in giving a correct historical perspective regarding the fundamentals of the
subject.
SEC. 9] LIGHT, 1700-1800 47
heats. Galileo had invented the thermometer, in 1597, and had
proposed a temperature scale, in 1613. The first mercury
thermometer was used by Kircher, in 1643, but it is to Fahren-
heit (1686-1736) that we owe the first reliable thermometers.
About 1724, Fahrenheit proposed the temperature scale now
known by his name. This was followed by the Reaumur scale,
in 1734, and by the Celsuis scale, in 1742, all based on the fixed
temperatures of melting ice and boiling water.
Discoveries in the field of latent and specific heats are due to
James Black (1728-1799), Professor of Chemistry at Glasgow
and Edinburgh. His measurements of the heat of fusion and
of vaporization of water form the basis of modern calorimetry
and gave definite form to the previously hazy distinctions
between temperature and heat.
With regard to theories of heat, there was retrogression during
the eighteenth century. From the quotations given above
(pp. 40 and 41), it is clear that Newton regarded heat as
intimately connected with the motion of the small particles of
which bodies are composed. This view seems to have been
shared by Newton's contemporaries. But early in the eighteenth
century there was a return to the "materialistic" or "caloric"
theory of heat, which held that heat is a subtle, highly elastic
fluid which could be extracted from or added to a body, thereby
causing the changes of its temperature. This heat fluid was
indestructible, its particles were self-repellant but were attracted
by ordinary matter, and it was all-pervading. The expansion
of bodies when heated was the natural result of "swelling" due
to forcing caloric into matter. The production of heat by per-
cussion was due to the releasing or '' pounding loose'' of some of
~ the caloric naturally condensed in or absorbed by the body,
thereby increasing the amount of free caloric within the body.
Black, whose experiments on calorimetry were mentioned above,
explained latent heats and specific heats on the basis of this
theory. Indeed, by the end of the eighteenth century, the
materialistic theory of heat was quite generally accepted.
9. Light during the Eighteenth Century.-Reference has
already been made (p. 37) to the development of achromatic
lenses by Hall and, later, by Dolland. This discovery paved
the way for substantial improvements in refracting telescopes,
but the impossibility of getting larger pieces of flint glass of good
optical quality set an unfortunate upper limit to the size of
48 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
refractors. And, consequently, astronomers turned their atten-
tion to reflectors for telescopes of large light-gathering power.
Especially noteworthy are the telescopes made by William
Herschel (1738-1822), the largest one being 4 feet in diameter.
Herschel's skill in constructing these instruments was surpassed
by his skill and diligence in using them. He discovered the
planet Uranus and two of its satellites, two satellites of Saturn, a
very large number of nebulre and double stars, the motion of the
solar system through space-indeed, he established a new stand-
ard of observational astronomy. He did for the solar system
what Copernicus, Kepler, and Tycho had done for the earth.
They gave the earth its proper place among the planets. He
gave the solar system its proper place among the stars.
An event of extraordinary importance in the history of sciences
was the discovery of the aberration of light, by Bradley, in 1728.
The absence of any measurable stellar parallax was one of the
stumbling blocks in the way of the Copernican system and,
therefore, was one of the outstanding problems of astronomy.
Tycho recognized that, viewed from opposite sides of the earth's
orbit, the stars should show a perspective displacement. His
careful observations convinced him that no such displacement so
great as 1 minute of arc existed. Later observers, likewise,
sought in vain.
Meantime, in 1675, Romer, a Danish astronomer, by a sys-
tematic study of the eclipses of Jupiter's moons as they pass
behind that planet, had found that the interval between eclipses
was longer when the earth was receding from Jupiter and shorter
when the earth was approaching Jupiter. When the earth is on
the opposite side of the sun from Jupiter, Romer found that the
eclipses were some 22 minutes 1 behind schedule, as compared~
with the occurrence of the eclipses when the earth and Jupiter
are on the same side of the sun. Romer correctly explained
this discrepancy by assuming that it takes light 22 minutes to
cross the earth's orbit. Frorn an approximate knowledge of the
diameter of the orbit, he computed that the velocity of light must
be about 190,000 miles per second. This was the first determina-
tion of the velocity of light.
In hopes of being able to measure stellar distances, Bradley
(1692-1762), in December, 1725, began systematic observations
on the position of a zenith star, ~ Draconis. If stellar p~rallax
1 The modern value is about 16Y2 minutes.
SEC. 10] ELECTRICITY, 1700-1800 49
existed, this star should be farthest south in December and
should then move north, reaching its maximum northerly posi-
tion 6 months later. The position of the star was found to
change, but not in the manner expected. It reached farthest
south in March and farthest north in September, the angular
distance between the two positions being about 40 seconds of
arc. Bradley continued his observations with a larger instrument
and on other stars, and, in 1728, he came to the conclusion that
the observed displacement was not due to parallax at all but to
an apparent shift in the star's position due to a combination of
the velocity of light with that of the earth in its orbit. 1 He was,
thus, enabled to deduce a value for the velocity of light which
was in substantial agreement with that determined by Romer a
half-century earlier. This discovery of Bradley's was the first
in the series which eventually led to the theory of relativity.
Theories as to the nature of light made no material progress
during the eighteenth century. Some writers are inclined
to ascribe this to the prestige given to the corpuscular theory
by the fact that it was supported by Newton, whose preeminence
"seemed to act like a spell," as had Aristotle's, centuries before.
But if so, then, likewise, the dynamic theory of heat held by
Newton should have been uppermost during the eighteenth
century, whereas, as has already been pointed out, the reigning
theory of heat during this period was the caloric theory. Lack
of progress in the theory of light was, of course, due to lack of
any crucial experiment, just as was the case with theories of
heat. Science has never progressed on the basis of speculation
only.
10. E!ectricity during the Eighteenth Century.-With the
possible exception of theoretical mechanics, electricity received
more attention during the eighteenth century than did any other
branch of physics. But until the discovery of electric currents,
at the end of the century, research was concerned with
electrostatics.
Stephen Gray (1670-1736) discovered that the difference
between conductors and non-conductors of electricity depends
on the material, and he showed that conducting bodies may be
electrified by insulating them. Du Fay (1698-1739) extended
Gray's experiments and showed that all bodies may be electrified;
1See any text on astronomy for further explanation of the cause of the
phenomenon.
50 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
that flames exercise a discharging power; and that there are two
kinds of electricity which he called "vitreous" and "resinous.''
He was, thus, led to propose the two-fluid theory of electricity to
explain attraction and repulsion. During the first half of the
eighteenth century, the electroscope was invented (by Hawksbee,
in 1705), frictional electric machines were developed, the leyden
jar discovered (1745), and there was considerable popular
interest in electrical phenomena.
During the latter half of the century, three names stand out
preeminent: Benjamin Franklin (1706-1790), Henry Cavendish
(1731-1810), and Charles A. Coulomb (1736-1806).
Franklin's experiments began about 1745. One of his first
observations was the effect of points "in drawing off and throwing
off the electrical fire." He proposed the one-fluid theory of
electricity, somewhat similar to the caloric theory of heat. This
theory supposed that all bodies naturally contain a certain
amount of this fluid. When a body had an excess of the fluid, it
was "plus," or positively, electrified. When it had a deficit, it
was "minus," or negatively, electrified. In form, this theory is
similar to our present electron theory of electrification, with signs
changed. But the Franklin theory can hardly be called the
forerunner of our modern theory, which grew out of experiments
of a very different kind. _
About 1750, Franklin began to speculate on the identity of
electricity and lightning, pointing out many similarities and
proposing, by means of a pointed iron rod, to "draw off the
fire from a cloud." Franklin's writings were publishedinEurope,
and, in 1752, Dalibard actually tried the experiment in Paris,
confirming Franklin's prediction. A short time later, Franklin,
to verify Dalibard's results, performed the famous kite experi-
ment, so well known to every schoolboy. This led to his study of
atmospheric electricity and to his invention of the lightning rod.
Franklin's researches occupied but a small portion of his long and
busy life, but they were sufficient to give him a high standing
among the scientists of the world.
Up to and including Franklin's work, studies of electrostatics
had been qualitative. Then came the quantitative researches
of Cavendish and of Coulomb.
Cavendish is known not only for his work in electrostatics but
also for his researches in chemistry and for the well-known
"Cavendish experiment," in 1798, in which he determined the
PLATE 4.-Franklin.
(Facin.() page 50!
SEC. 9] ELECTRICITY, 1700-1800 51
constant of gravitation. His electrical researches were very
extensive and covered the period from about 1770-1780, but
most of his work remained unknown, for he published only one
paper of importance: "An Attempt to Explain Some of the
Principal Phamomena of Electricity by Means of an Elastic
Fluid." This appeared in the Philosophical Transactions of the
Royal Society in 1771. He left behind a large amount of mate-
rial, however, in the form of manuscript notes. These were
edited and published, in 1879, by Maxwell, under the title "The
Electrical Researches of the Honorable Henry Cavendish,
F.R.S." In these experiments, Cavendish proved the inverse-
square law of electrostatic force; measured capacity and expressed
it correctly in "inches;" 1 recognized the principle of the con-
denser and measured the specific inductive capacity of several
substances; had a reasonably clear idea of the quantity which we
now call "potential"; and anticipated Ohm's law by 50 years.
His measurements were made by means of an electrometer of
the pith-ball type, the balls being ''made of pith of elder, turned
round in a lathe, about one-fifth of an inch in diameter and
suspended by the finest linen threads that could be procured,
about 9 inches long." Later, he used "a more exact kind of
electrometer '' consisting of cork balls suspended by gilt-covered
wheat straws. Had these important measurements been com-
municated to his scientific contemporaries, the history of elec-
tricity might have been substantially modified. Maxwell
quotes Franklin as saying of Cavendish: "It were to be wished
that this noble philosopher would communicate more of his
experiments to the world, as he makes many, and with great
accuracy.''
Coulomb's work in electricity grew out of his development of
the torsion balance, originally used for studying the torsional
elasticity of wires. In the period 1785-1789, he published seven
papers on electricity and magnetism in the Memoirs de l'Academie
Royale des Sciences. In these papers, he showed, by means of
the torsion balance, that electrostatic forces obey the inverse-
square law; that on conducting bodies, the charge exists only on
the surface; and that the capacity of a body is independent of
the nature of the material of which it is composed. Most of
1 Cavendish's unit was one-half the modern unit, since he used the diameter
of the sphere instead of its radius. His standard of capacity was a sphere
12.1 inches in diameter.
52 HISTORICAL SKETCH, SECOND PERIOD [CHAP. II
Coulomb's discoveries had been anticipated by Cavendish, but,
on account of the latter's failure to publish his work, the farmer's
experiments were taken as the basis of theoretical studies by
Poisson and others. Coulomb advocated the two-fluid theory
of electricity.
This second period comes to a close with rival theories contend-
ing in each of three of the subdivisions of Physics: the caloric
vs. the dynamic theory in heat; the corpuscular vs. the undulatory
theory in light; and the one-fluid vs. the two-fluid theory in
electricity. The very fact that these issues were raised, in rather
clean-cut fashion, is an indication of the tremendous strides
which had been taken since Galileo. But most important of all,
men had learned the value of experiment and observation and
the fallacy of blindly following "authority." Physics was ready
to profit by the discoveries about to be made by Rumford, Davy,
Young, Oersted, and Faraday.
CHAPTER III
HISTORICAL SKETCH, THIRD PERIOD (1800-1890 A.D.):
THE RISE OF CLASSICAL PHYSICS
1. Heat and Energy.-The law of the conservation of energy
is one of the most fundamental and far-reaching of all our
physical laws, and yet, curiously enough, it is of comparatively
recent origin, for it was not announced until the middle of the
nineteenth century. It began in qualitative form with the
experiments of Rumford, in 1798, and assumed its final and
quantitative form after the experiments of Joule, about 1847.
Rumford's 1 work marks the beginning of the end of the caloric
theory of heat. About 1798, while engaged in boring out some
cannon for the Bavarian government, he was impressed with
the large quantity of heat produced. What was its source?
If the caloric theory were correct, then this great evolution of
heat ought to result in a loss of "something," perhaps weight,
by the gun, the chips, or the tool. But no such loss could be
observed. Perhaps the caloric theory, so generally accepted
at that time, was wrong! If so, an experiment should answer
the question.
Accordingly, by a mechanism worked by two horses, he
caused a blunt steel boring tool to rotate, under great pressure,
on a piece of brass, the brass and tool being immersed in water.
In 2 hours, the water actually boiled. Apparently, heat would
be produced by this apparatus just as long as the horses kept
1 Count Rumford, or, as he was originally known, Benjamin Thompson,
was born in New England in 1753. He appears to have exhibited, very
early, a taste for science. When the War of the Revolution broke out in
177 5, he took sides with the British and fled to England, leaving behind him
in America his wife and child whom he never saw again. In 1778 he was
made a Fellow of the Royal Society. He left England in 1783 and became
a sort of military engineer to the Bavarian government, which conferred on
him the title "Count Rumford." It was in this capacity, in 1798, that he
performed the experiment of producing heat by friction. It is also of inter-
est to note that, in 1800, he founded the Royal Institution in London, which,
for a century and a quarter has been one of the leading scientific institutions
of the world. He died in France in 1814.
53
54 HISTORICAL SKETCH, THIRD PERIOD [CHAP. III
going, the quantity of heat which might, in this way, be generated
being independent of the properties of the brass or the iron. But
this conclusion, arrived at by experiment, was quite contrary to
the caloric theory, according to which there should be a finite
quantity of heat fluid in any body. Therefore, Rumford
reasoned:
. . . anything which any isolated body, or system of bodies, can
continue to furnish ":ithout limitation cannot possibly be a material
substance; and it appears to me to be extremely difficult, if not quite
impossible, to form any distinct idea of anything capable of being excited
and communicated in the manner heat was excited and communicated
in these experiments, except it be motion. 1
Rumford's conclusions-that heat results from motion-
were so revolutionary and so contrary to the prevailing caloric
theory, that they made but few converts. Sir Humphrey Davy,
Director of the Royal Institution, was sufficiently impressed
to experiment further. He observed that two pieces of ice,
kept below the melting point, could be melted by rubbing them
together, even if the experiment were performed in vacuum, but
it was not until 1812 that he was convinced that the "cause of
the phenomenon of heat is motion." Another convert was
Thomas Young, who, in 1807, espoused the kinetic theory on the
basis of Rumford's experiments.
But the majority of the supporters of the caloric theory were
unconvinced. Even Carnot (1796-1832), the founder of the
modern science of thermodynamics, when, in 1824, he proposed
the now famous Carnot's cycle, based his reasoning on the
caloric theory. A given quantity of caloric "falling" from a
higher to a lower temperature was analogous to a given quantity
of water falling from a higher to a lower level. Each was capable
of producing motive power. The kinetic theory had to wait for
a quantitative experiment or statement.
In 1842, R. J. Mayer (1814-1878) published a paper 2 in which,
partly on philosophical grounds, he announced the equivalence
of heat and energy and, from data on the specific heat of a gas
at constant volume and constant pressure, he deduced a value for
the mechanical equivalent of heat.
1 Rumford's paper appeared in the Philosophical Transactions of the
Royal Society for 1798.
2 Ann. der Chemie und Pharmacie, May, 1842.
SEC. 1] HEAT AND ENERGY 55
Meanwhile, Joule (1818-1889), in England, unacquainted
with Mayer's work, was carrying on a very careful series of
experiments, begun in 1840, in which he converted the mechanical
energy of a falling weight into heat by a paddle wheel revolving
in water and, thus, determined that 778 foot-pounds of work
would raise 1 pound of water 1°F. Joule announced his results
at a meeting of the British Association for the Advancement of
Science, in June, 1847. The paper would have passed almost
unnoticed, had it not been for William Thompson, later Lord
Kelvin, who, grasping the real significance of the proposed
theory, by his discussion made the paper the event of the meeting.
Quite independently of the work of Mayer and of Joule,
Helfi1:holtz (1821-1894), in 1847, read a paper before the Physical
Society, in Berlin, on "Die Erhaltungen der Kraft," in which, on
the basis of the impossibility of perpetual-motion machines, he
announced the law of the conservation of energy. The paper
was rejected for publication by the editor of the Annalen der
Physik! It was later published in pamplet form.
The caloric theory could not withstand these attacks, and, by
1850, the mechanical theory of heat and the doctrine of energy
and of the conservation of energy were, quite generally, accepted.
After half a century, Rumford's theory had come into its own.
Heat was simply one of the several forms of energy.
Upon this solid foundation, the science of heat grew apace,
during the next 50 years. In part, under the impetus of applied
science, there was accumulated a vast amount of data on the
thermal properties of solids, liquids, vapors, and gases. Ther-
mometry was put on a rational basis by Lord Kelvin's thermo-
dynamic temperature scale. Thermodynamics itself became
almost a separate science. The atomic and molecular theory of
matter, combined with the kinetic theory of gases, all but made
visible the actual motions which constitute thermal agitation.
Indeed, the theory of heat and energy, the "classical theory," as
we now call it, with all its ramifications in the various branches
of both pure and applied science, seemed to have reached its
final form.
But, in the closing years of the nineteenth century, a careful
study of the phenomena of radiation established experimental
laws which the classical theory could not explain. Either a
new theory or a modification of the existing one was necessary.
The latter alternative led to the development of the quantum
56 HISTORICAL SKETCH, THIRD PERIOD [CHAP. III
theory, which left a large part of the main structure of the clas-
sical theory intact, but which evolved radically new concepts to
explain the new facts.
2. Light.-The revival of the wave theory of light, by Thomas
Young (1773-1829), is one of the most important events in the
history of the early years of the nineteenth century. It will be
remembered that, since the time of Newton, the majority of
scientists had supported the corpuscular theory. In 1800,
Young published a paper in the Philosophical Transactions of the
Royal Society under the title "Outlines of Experiments and
Inquiries respecting Sound and Light," in which he pointed out
(1) that the observed fact that the velocity of light is quite
independent of the nature of the source-be this electric spark,
a faintly glowing body, or the sun-is very difficult to explain
on any corpuscular theory but follows, naturally, from the wave
theory, since, by analogy with sound, the velocity of the light
waves in the ether should be independent of intensity or fre-
quency ;1 and (2) that the dividing of a beam of light into a
refracted and a reflected ray at the interface between two media
was to be expected from the wave theory but had not been
satisfactorily explained by the adherents of the corpuscular
theory.
A year later, Young presented to the Royal Society a paper
"On the Theory of Light and Colors," in which he proposed the
principle of the interference of two wave trains as an explanation
of Newton's rings and the colors of thin plates. From Newton's
careful measurements of the thickness of the air layers necessary
to produce the several colors, Young was enabled to compute
wave lengths. He found the wave length of yellow light, for
example, to be 0.0000227 inches, a value in excellent agreement
with modern data. In subsequent papers, he described the
interference fringes which he had observed by placing hairs or
silk threads in front of a narrow slit illuminated from the rear;
he announced the change of phase on reflection; he explained
diffraction bands by the principle of interference, and he showed
that the spacing of these bands gave values of the wave length
agreeing with those obtained from Newton's rings and that,
1It is a very curious circumstance that, a century later, an argument of a
very similar kind proved to be a very serious embarrassment to the wave
theory. For it was observed that the velocity with which photoelectrons are
emitted under the influence of light is independent of the intensity of the
source (but depends on the wave length). See p. 169.
SEC. 2] LIGHT 57
therefore, both phenomena must be due to a common cause.
Again, quantitative measurements became an indispensable link
in the chain of reasoning.
But Young's paper aroused a storm of protest, even of derision
and abuse. He had dared to question Newton's corpuscular
theory! He was attacked not by the Church, as was Galileo, but
by some of his scientific, or, more probably, pseudoscientific,
contemporaries. His chief assailant was Henry Brougham,
afterward Lord Chancellor of England, who "reviewed" Young's
papers in the Edinburgh Review. The nature of Brougham's
attack is indicated by the following q-qotations, taken almost at
random:
This paper [i.e., Young's] contains nothing which deserves the name
either of experiment or discovery . . .
We wish to raise our feeble voice against innovations that can have
no other effect than to check the progress of science, and renew all those
wild phantoms of the imagination which Bacon and Newton put to
flight from her temple. We wish to recall philosophers to the strict
and severe methods of investigation . . .
We . . . have searched without success for some trace of learning,
acuteness and ingenuity, that might compensate his evident deficiency
in the powers of solid thinking, calm and patient investigation, and
successful development of the laws of nature by steady and modest
observation of her operations.
Young was accused of deliberately misquoting other investi-
gators and of many other "scientific" crimes.
Brougham's attacks, as full of abuse as they were devoid of
science, were so evidently based on personal spite that they
should have convinced no one. But though Young replied at
length in a privately published pamphlet, it was a long time
before public opinion was willing to receive his theories with an
open mind. Young's final vindication came through the work
of two Frenchmen, Fresnel and Arago.
The phenomenon of the polarization of light by Iceland spar
had been discovered by Bartholinus, in 1669. Newton had
tried to fit the corpuscular theory to polarization by assuming a
sort of structure to the corpuscles. But the explanation was
not convincing, and, indeed, polarization had proved an enigma
to both theories of light.
In 1808, Malus (1775-1812), almost by accident, discovered
the phenomenon of polarization by reflection. A year later, as a
58 HISTORICAL SKETCH, THIRD PERIOD [CHAP. III
result of further research, he announced the law of polarization
which bears his name: If a beam of plane-polarized light be
incident onto a crystal of Iceland spar so that its plane of polari-
zation makes an angle a with the principal plane of the crystal,
the intensity of the ordinary ray transmitted by the crystal is pro-
portional to cos2 a; and of the extraordinary ray, to sin2 a.
A little later, Fresnel (1788-1827), unaware of Young's work,
rediscovered the phenomenon of interference and performed the
famous experiment with the two mirrors. His confidence in
the wave theory as an explanation of the phenomenon was
strengthened upon learning from Arago of Young's discoveries
more than a decade earlier. But the difficulties of explaining
polarization were as great as ever.
Finally, Young, in a letter to Arago, written in January, 1817,
made the bold suggestion that all the phenomena of polarization
could be explained on the assumption that the light vibrations
were transverse rather than longitudinal. Quite independently,
Fresnel reached the same conclusion. Subsequent experiments
were in entire agreement with this assumption. For example,
Fresnel and Arago showed that two plane-polarized beams of
light could be made to interfere if their planes of polarization
were parallel; but no interference was observed if the beams were
polarized at right angles to each other. It had been difficult
enough for scientists and philosophers to conceive of a fluid ether
filling all space and transmitting longitudinal vibrations. But
now they were asked to believe the still more impossible-that
this medium, the ether, was a solid or, at least, had such of the
properties of a solid as are necessary for the transmission of
transverse waves, i.e., the properties of rigidity and density.
And yet the planets must still move through this "solid," per-
vading all space, with no measurable changes in their periods of
revolution. No wonder the scientific public was even more
incredulous!
Nevertheless, the indisputable experimental evidence of interfer-
ence phenomena forced the conclusion 1 that light is a wave
1 A close parallel to this situation is to be found in the origin, a century
later, of Rutherford's nuclear theory of the atom, based upon the experi-
ments of the scattering of alpha particles, a theory which was at once
accepted, in spite of the fact that it seemed to necessitate the abandonment,
or radical revision, of some of the fundamental concepts of electrodynamics
(see Chap. X, Sec. 4).
SEC. 2] LIGHT 59
motion, and the equally unambiguous evidence from experiments
on polarization required that this wave motion be transverse.
There seemed to be no escape from these conclusions.
But the crucial experiment was not performed until Foucault
(1819-1868) measured the relative velocities of light in water and
in air. It had been generally accepted that, on the basis of the
corpuscular theory, the velocity of light should be greater the
greater the optical density of the medium ;1 while the reverse
should be the case with the undulatory theory. And it had
been known for a long time that a determination of the relative
velocities of light in air and in some optically denser medium,
such as water, would decide between the two theories. But it
was not until 1850, that experimental methods had developed
to the point of making such a test feasible. In that year,
Fizeau and Foucault, in part, independently, by means of the
well-known, rotating-mirror method 2 previously proposed by
Arago, ascertained that light travels more slowly in water than
in air-a final triumph, so it seemed, for the wave theory.
Meantime, other evidence had been accumulating. Fraun-
hofer, in 1817, observed the dark lines in the solar spectrum.
About 1820, he made excellent determinations of the wave
length of the D-lines of sodium, by means of a grating made of
fine wires. Babinet, in 1829, proposed the wave length of light
as a standard of length. The phenomenon of dispersion was
explained, on the basis of the elastic-solid theory of the ether,
by Cauchy, who proposed his famous dispersion formula. In
fact, the new wave theory seemed capable of explaining all
optical phenomena.
From 1850, until the end of the third period (1890), the
wave theory held the field, undisputed. Each new discovery
seemed to strengthen its position. The frequent assertions that
the corpuscular theory was finally disposed of certainly seemed
justified, particularly after the development of Maxwell's
electromagnetic theory of light and its experimental verification.
But the corpuscular theory was not dead. It was only sleeping.
Some important discoveries in light from 1800 to 1890 not
previously mentioned are:
1 See EnsER: "Light for Students."
2 See any textbook on optics.
60 HISTORICAL SKETCH, THIRD PERIOD [CHAP. III
DISCOVERER
Three-color theory of vision (1807) . . . . . . . . . . . . . . . . . . . . . . . Young
Heat and light rays differ only in wave length (1807) ....... Young
Rotary polarization of quartz (1811) ..................... Arago
Polarization of scattered light (1813) ..................... Arago
Rotary polarization by liquids (1815)... . . . . . . . . . . . . . . . . . Biot
Light sensitivity of silver bromide (1826). . . . . . . . . . . . . . . . . Balard
Change of conductivity of selenium on illumination (1837) ... Knox
Doppler effect (1842). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Doppler
Beer's law of absorption (1852) .......................... Beer
Foundation of spectral analysis (1859) .................... Kirchhoff and
Bunsen
3. Electricity and Magnetism.-The history of electricity
during the nineteenth century is so extensive that even a sketchy
outline would fill a small volume. We shall, therefore, discuss
only a few of the more important events, particularly the works of
Faraday, Henry and Maxwell. In singling out these three names
for special biographical mention, there is no attempt to minimize
the great contributions of their contemporaries-such men as
Ampere (1775-1836), Oersted (1777-1851), Ohm (1787-1854),
'Wheatstone (1802-1878), Lenz (1804-1865), Stokes (1819-1903),
Lord Kelvin (1824-1908), Kirchhoff (1824-1887), Tait (1831-
1901), Hertz (1857-1894), and many others, whose names are
intimately associated with theoretical and experimental elec-
tricity. But the work of Faraday and that of Maxwell are so
closely related to each other and to the whole subject of modern
physics that we can best present that part of electricity and
magnetism in which we are particularly interested by giving an
account of the contributions of these two men.
4. Michael Faraday. 1 (a) Biographical Sketch.-Michael
Faraday was born Sept. 22, 1791, in a small village near London.
He was the son of a blacksmith, James Faraday, who died in
1810, following a long illness. After a very rudimentary educa-
tion, young Michael, to assist his mother in providing. for the
family, was engaged as errand boy to a bookseller and stationer,
in 1804. He performed his duties so conscientiously that, in
the following year, he was formally apprenticed to his employer
to learn the art of bookbinding.
During this apprenticeship, Faraday made good use of his
spare time by reading some of the books which passed through the
shop. He was particularly interested in works on science, such
1 See THOMPSON, SYLVANUS P.: "Michael Faraday: His Life and Work."
PLATE 5.~Faraday.
(Facing pa{)e 60)
SEC. 4] FARADAY 61
as the articles on electricity in the '' Encyclopedia Britannica''
and in Marcet's "Conversations in Chemistry." In connection
with the reading of the latter book, he showed one of the impor-
tant characteristics 1 of the great investigator-to-be by performing
such of the simple experiments described "as could be defrayed
in their expense by a few pence per week.'' He also made some
simple electrical apparatus.
Aside from his own reading, Faraday's only scientific education
consisted in a dozen lectures on natural philosophy by a Mr. Tatum,
in the years 1810 and 1811, and four lectures on chemistry by
Sir Humphrey Davy, in the winter of 1812. The very careful
and neatly written notes which he made of these lectures served
him a very useful purpose when, after having finished his appren-
ticeship and being quite unhappy at his trade of bookbinder,
he made bold to apply to Sir Humphrey Davy for a position,
however menial, at the. Royal Institution of which Davy was
then director. '' As proof of his earnestness'' Faraday sent
along with his application the notes which he had taken of
Davy's four lectures. Davy was so pleased with the letter and
the notes, th-at;eventually, in March, 1813, Faraday was engaged
as apparatus and lecture assistant at 25 shillings per week.
In October, 1813, Faraday accompanied Sir Humphrey and
Lady Davy on a trip to the Continent. During their 18 months'
absence, they visited many of the important scientific centers of
Europe and saw many of the continental scientists and something
of their methods. Naturally, Faraday was much impressed
and edified by what he saw and learned, but, assistant though
he was, he also left an impression because of his modesty, ami-
ability, and intelligence. Said one writer, "We admired Davy;
we loved Faraday."
On returning to England, in the spring of 1815, Faraday was
reengaged at the Royal Institution, at 30 shillings per week.
At first, he simply "assisted," but, under the encouragement of
Davy, he soon began original investigations, initially in chem-
istry. In 1816, he published his first paper, on "An Analysis of
Caustic Lime." In 1817, he published 6 papers; in 1818, 11;
and, in 1819, 19 ! These papers were concerned with such
subjects as the escape of gases through capillary tubes; the
production of sound in tubes by flames; the combustion of the
1 In later life, he wrote: "I was never able to make a fact my own without
seeing it.''
62 HISTORICAL SKETCH, THIRD PERIOD [CHAP. III
diamond; and the separation of manganese from iron. About
1820, he began his electrical researches. These, and others
growing out of them, continued for nearly 40 years.
Almost his entire scientific life was spent at the Royal Institu-
tion. In 1825, he was made Director of the Laboratory. Declin-
ing offers of positions elsewhere, turning away professimtal
occupations which might have made him wealthy, he gave to
his science and to the institution which he served a devotion
seldom if ever equaled. The secret of his success, which brought
him, during his lifetime, honors from all over the scientific world
and which immortalized his name by the long list of scientific
discoveries ascribed to him, is, perhaps, to be found in some
excerpts from his many notes:
Aim at high things, but not presumptuously.
Endeavor to succeed-expect not to succeed.
It puzzles me greatly to know what makes the successful philosopher.
Is it industry and perseverance with a moderate proportion of good sense
and intelligence? Is not a modest assurance or earnestness a requisite?
Do not many fail because they look rather to the renown to be acquired
than to the pure acquisition of knowledge . . . ? I am sure I have
seen many who would have been good and successful pursuers of
science, and have gained themselves a high name, but that it was the
name and the reward they were always looking forward to-the reward
of the world's praise. In such there is always a shade of envy or regret
over their minds and I cannot imagine a man making discoveries in
science under these feelings.
The reader is urged to study carefully Faraday's life and works,
particularly to read, as unexcelled examples of scientific exposi-
tions, portions of his "Experimental Researches in Electricity
and Magnetism."
Faraday's last scientific work was done about 1860. He died
in 1867.
(b) Earlier Work in Electricity and M agnetism.-It is neces-
sary to preface an account of Faraday's researches in electricity
and magnetism by an account of the development of "current"
electricity up to 1820. It will be remembered that all work in
electricity up to about 1790 was confined to electrostatics.
About that time, Galvani (1737-1798), as a result of a chance
observation that a frog's leg kicked convulsively when connected
with the terminal of an electric machine, was led to an extensive
study of "animal electricity." In the course of these experi-
SEC. 4] FARADAY 63
ments, Galvani observed that if the frog's leg were so suspended
that the exposed nerves touched a metal plate, say silver, then,
a contraction of the muscle occurred whenever the foot touched
another metal, say iron. He even observed slight muscular
contraction when both plates were of the same kind of metal.
This led him to believe that the nerve was the source of electricity
and that the metal served simply as conductor.
Volta (17 45-1827) continued and extended Galvani's experi-
ments and proved that the phenomenon was due to electricity
generated by contact of dissimilar metals, and, in March, 1800,
he communicated to the Royal Society a description of the first
battery for producing an electric current-the historically famous
Volta, or voltaic, "pile," consisting of zinc and copper plates
placed alternately and separated by blotting paper moistened
with brine. He also described the first voltaic battery, which
consisted of cups containing brine or dilute acid connected by
copper and zinc strips joined together.
This new source of electricity was received with a great deal
of interest, and many important discoveries soon followed.
Only a few weeks after Volta's communication, Nicholson and
Carlisle (in May, 1800) accidentally discovered the decomposition
of water by the electric current. Thinking to secure better
contact between two wires forming part of the circuit, they
joined the ends of the wires by a drop of water and, at once,
observed the formation of a gas which they recognized as hydro-
gen. In September, 1800, Ritter observed the decomposition
of copper sulphate, and, about the same time, Cruikshank showed
that many solutions of salts could be similarly decomposed.
In 1802, Davy discovered that the electrolysis of water yielded
2 volumes of hydrogen to 1 of oxygen. In the same year, Ritter
discovered the principle of the storage battery. Sodium and
potassium were discovered by Davy, by electrolysis, in 1807.
During this same period were discovered the heating effect of
the current, and the arc light.
It was early suspected that there was some relation between
electricity and magnetism, and, indeed, a few scattered observa-
tions had been made, such as the magnetism of steel needles by
lightning discharges. But the science of electromagnetism dates,
quite definitely, from the discovery, in 1820, by Oersted (1777---
1851), that a magnetic needle tends to set itself at right angles
to a wire through which an electric current is flowing.
64 J!ISTORICAL SKETCH, THIRD PERIOD [CHAP. III
Immediately, the brilliant French physicist, Ampere (1775-
1836), extended Oersted's discoveries; announced the well-
known "Ampere's rule" for the relative direction of current flow
and direction of deflection; discovered the mutual force action
of two parallel currents on each other; invented the electromag-
net; and made the first suggestions of the electric telegraph.
(c) Faraday's Discovery of the Principle of the Motor .-Faraday's
interest in electromagnetism dates from April, 1821, when
W ollaston attempted, at the Royal Institution, to make a wire
carrying an electric current revolve around its own axis when the
pole of a magnet was brought near. The experiment was unsuc-
cessful, but the phenomenon excited Faraday's interest, and he
determined to make a study of it. First, he read what had been
done by others and repeated many of their experiments. In the
course of these experiments, he observed that when the magnetic
pole was brought near the wire, "the effort of the wire is always
to pass off at right angles from the pole, indeed to go in a circle
around it . . . " 1
The following day, Sept. 4, he wrote:
Apparatus for revolution of wire and magnet. A deep basin with a
bit of wax at bottom and then filled with mercury. A magnet stuck
upright in wax so that pole [is] just above surface of mercury. Then
piece of wire, floated by cork, at lower end dipping into mercury and
above into silver cup.
On passing a current through the wire, it revolved continuously
around the magnet. This was the first electric motor! On
Christmas Day, 1821, Faraday succeeded in making a wire
carrying a current move under the influence of the earth's field.
It was decades before electric motors were to be built, but this
experiment of Faraday's clearly sets forth the fundamental
principle involved.
(d) Electromagnetic Induction.-Oersted's experiment, and the
subsequent developments, had clearly shown how, at will, "to
produce magnetism by electricity." Faraday seems to have
held it as one of the tenets of his scientific philosophy that every
physical relation (of cause and effect) has its converse. If
electricity can produce magnetism, then magnetism should pro-
duce electricity. And in his notebook for 1822 appears the
entry: "Convert magnetism into electricity." His repeated
1 Quotation from Faraday's laboratory notebook, Sept. 3, 1821.
SEC. 4] FARADAY 65
attempts failed. For example, in 1825, he tried what seemed to
be the obvious converse by looking for an electric current in a helix
of wire coiled around a magnet. Later, he tried to find a current
in a wire placed near another wire carrying current. In 1828,
is recorded another failure.
Faraday was not alone in his search for this converse effect.
Fresnel, as early as November, 1820, had announced that he had
succeeded in decomposing water by the current produced by a
solenoid wound around a magnet. Later, he withdrew the state-
ment. Ampere, likewise, made and subsequently withdrew a
similar statement regarding the production of current from a
magnet. Apparently, Ampere, later, became convinced that no
such effect existed. They were all looking for the production
of a steady current by placing a wire near a magnet.
But, several times, investigators were very near to the dis-
covery of induced currents. In 1822, Ampere and de la Rive
had observed a slight motion in a suspended copper coil when
a magnet was approached. In 1824, Arago observed the damp-
ing of the vibrations of a magnetic needle suspended over a
copper plate. This observation was extended by causing the
needle to revolve by revolving the copper plate underneath it,
air disturbances being, of course, eliminated. A little later, the
converse effect (the causing of the copper disk to revolve by
revolving the needle) was discovered. It was shown that this
"dragging" effect was greater the greater the electrical conduc-
tivity of the spinning plate. Even the effect of radial slits in
the copper disk, in reducing the dragging action on the magnet,
was observed. But, suggestive as these experiments were, the
real principle requisite to the "conversion of magnetism into
electricity'' remained undiscovered.
In the summer of 1831, Faraday attacked the problem for a
fifth time. This time, instead of placing a permanent magnet
inside a helix, he procured a soft iron ring 6 inches in external
diameter, on which he wound two coils of copper, A and B,
"separated by twine and calico." Coil A consisted of three
lengths of wire each 24 feet long, "insulated from each other and
capable of being connected as one length or used separately."
Coil B consisted of two lengths of 30 feet each. To detect a
possible current in coil B, he "connected its extremities by a
copper wire passing to a distance and just over a magnetic
needle." When coil A was connected to a battery, there was
66 HISTORICAL SKETCH, THIRD PERIOD [CHAP. III
"a sensible effect on the needle. It oscillated and settled at last
in original position. On breaking connection of side A with
battery, again a disturbance of the needle."
Slight as these momentary effects were, Faraday recognized
their importance, although he had been looking for a continuous
effect. On Aug. 30, he. writes, "May not these transient effects
be connected with causes of difference between power of metals
at rest and in motion in Arago's experiments?" 1
From this slender clue, Faraday proceeded rapidly to the
discovery of the real effect. On the "third day" of his experi-
ments, he wound a coil of wire around an iron cylinder and
placed the cylinder so as to join the N pole of one permanent
magnet with the S pole of another. The coil was connected to
a galvanometer:
Every time the magnetic contact at Nor S was made or broken there
was a magnetic action at the indicating helix [i.e., galvanometer]-the
effect being, as in former cases, not permanent but a mere momentary
push or pull.
On the fourth day, he showed that the presence of iron was
not necessary; that the effect was observed by the action of one
helix on another.
The fifth day's experiment is best described in his own words:
A cylindrical bar magnet . . . had one end just inserted into the
end of the helix cylinder; then it was quickly thrust in the whole length
and the galvanometer needle moved; then pulled out and again the
needle moved, but in the opposite direction. The effect was repeated
every time the magnet was put in or out, and therefore a wave of elec-
tricity was so produced from mere approximation of a magnet and not
from its formation in situ.
At last! He had "converted magnetism into electricity." The
essential requisite was relative motion, or a change of condition.
On the ninth day, he produced a continuous current by
turning a copper disk between the poles of a powerful electro-
magnet, the periphery of the disk being connected to its axis
through an indicating galvanometer. This was the now well-
known Faraday disk dynamo, the very first dynamo-electric
machine.
After only a few days' work in his laboratory, following,
however, years of patient and persistent experiment, Faraday
1 That is, the spinning copper disk mentioned above.
SEC. 4] FARADAY 67
had discovered a phenomenon for which the greatest scientists
of his time had sought in vain-electromagnetic induction, the
importance of which to subsequent generations is almost beyond
comprehension. It is to be doubted whether any single event
in all history has had a greater effect on the material aspects of
human society than has this discovery. For on it are founded
modern electrical science and industry-indeed, almost the whole
of our modern electrical age. Justly may Faraday be called
the father of electrical science.
Following this discovery, Faraday devised and tried various
electric machines to test and extend his newly discovered prin-
ciple. One of these machines, consisting of a rotating rectangle
of wire with a commutator attached, is, in every respect, the proto-
type of the modern dynamo. But his interest was always in
pure science, for he writes:
I have rather, however, been desirous of discovering new facts and
relations dependent on magneto-electric induction, than of exalting the
force of those already obtained; being assured that the latter would find
their full development hereafter.
Being unacquainted with mathematical symbols and methods,
Faraday always sought to explain his discoveries and to extend
his researches by purely physical reasoning. To the mathema-
tician, the law of magnetic attraction
F = m1m2
r2
may have been a sufficient explanation of the phenomenon, but
to Faraday, this gave a statement only of the magnitude of the
magnetic forces; it left the phenomenon itself quite unexplained.
Accordingly, he insisted that two magnetic poles, or two electric
charges, could act on each other only if the medium between the two
played some important part in the process. This insistence on the
importance of the medium ultimately led him to the very fruitful
concept, at first qualitative but later quantitative, of lines of force
and of the ''cutting'' of these lines as an essential process in electro-
magnetic induction. Three decades later, this concept became
one of the corner stones of Maxwell's electromagnetic theory.
Commenting on Faraday's laws of electromagnetic induction,
Maxwell wrote :
After nearly a half-century . . . , we may say that, though the
practical applications of Faraday's discoveries have increased and are
68 HISTORICAL SKETCH, THIRD PERIOD [CHAP. III
increasing in number and value every year, no exception to the statement
of these laws as given by Faraday has been discovered, no new law has
been added to them, and Faraday's original statement remains to this
day the only one which asserts no more than can be verified by experi-
ment, and the only one by which the theory of the phenomena can be
expressed in a manner which is exactly and numerically accurate, and
at the same time within the range of elementRry methods of exposition.
(e) The Laws of Electrolysis.-Faraday next turned his atten-
tion to proving that "Electricity, whatever may be its source 1 is
identical in its nature." He found, for example, that electricity
from a friction machine would deflect a galvanometer and would
cause chemical decomposition just as would electricity produced
by chemical action. And from his own experiments and those of
others, he concluded that the different "kinds" of electricity
differ not in character but only in degree.
This led him into the field of electrolysis. He first found that
many substances, such as certain chlorides, nitrates, and sul-
phates, are non-conductors when solid but are good conductors
when melted and that in the molten state they are decomposed
by the passage of current. This showed that water was not
essential to electrolysis. To clarify description of his experi-
ments he introduced the terms" electrode "" anode "" cathode "
"ion "' "anion " "cation " "electrolyte ' '' '' electrochemica]
' '
'
equivalent," etc.' A quantitative
' ' whole phenomena
study of the
resulted in his discovery of the laws of electrolysis which bear
his name and which are the basis of all present-day work in that
subject.
Further, Faraday clearly recognized that a definite quantity
of electricity is associated with each atom or ion in electrolysis,
and, had he been able to determine the number of atoms in unit
mass of any substance, he would have anticipated, by 60 years,
the determination of the fundamental charge e. For he says:
Equivalent weights of bodies are simply those quantities of them
which contain equal quantities of electricity; . . . it being the elec-
tricity which determines the combining force. Or, if we adopt the
atomic theory or phraseology, then the atoms of bodies which are
equivalent to each other in their ordinary chemical action, have equal
quantities of electricity naturally associated with them.
(f) The Conservation of Energy.-In connection with a proof of
the fact that the electricity from the voltaic pile results from
1 That is, whether frictional, chemical, thermal, or electromagnetic.
SEC. 4] FARADAY 69
chemical action and not from mere contact of one substance with
another, Faraday stated clearly the doctrine of the conservation
of energy several years before the statements of Mayer, Helm-
holtz, and Joule. In 1840, he wrote:
The contact theory assumes that a force which is able to overcome a
powerful resistance . . . can arise out of nothing . . . This would
indeed be a creation of power, and is like no other force in nature. We
have many processes by which the form of the power is so changed that
an apparent conversion of one into the other takes place . . . But in
no case is there a pure creation or a production of power without a corre-
sponding exhaustion of something to supply it.
(g) The "Faraday" Ejfect.-Reference has already been made
to Faraday's abhorrence of the doctrine of "action at a distance."
He believed that, if two electric charges attract each other, the
medium between the two takes part in the process. Presumably,
therefore, the medium between two such charges is in a different
state than it would be if the charges were not present, and if so,
such an altered state should be detectable by observing the
alteration in some physical property of the medium. As early
as 1822, Faraday had experimented with a beam of polarized
light passing through a transparent solution carrying a current,
to see whether the current caused any ''depolarizing'' action.
Although he repeated the experiment several times, in subsequent
years, the results were all negative.
In 1845, he returned to the problem again, studying it very
exhaustively, but with negative results. He then tried solid
dielectrics with plates of metal foil connected to a powerful
electric machine, to see whether, under electric strain, they
would show any optical effects. No results! 1
Faraday then substituted a magnetic field for the electrostatic
field, to see whether the former would cause any depolarizing
action on the beam of light. Various substances were tried, but
still with negative results. Finally, he placed in the magnetic
field a very heavy piece of lead glass, which he had made many
years earlier, and when the magnetic lines were parallel to the
direction of the beam of polarized light, he observed that the
plane of polarization was rotated. At last, he had found a
relation between magnetism and light. This magnetic rotation
is now known as the "Faraday effect." Again, his persistent
1 Years later, this effect was found by Kerr.
70 HISTORICAL SKETCH, THIRD PERIOD [CHAP. III
search, maintained during 20 years of repeated failures, was
rewarded by the discovery of an effect, of the existence of which
he had the most sublime confidence.
(h) Miscellaneous.-Among Faraday's other researches may
be mentioned: numerous investigations in chemistry; the lique-
faction of several gases formerly thought "permanent"; the
diffusion of gases through solids; self-induction; certain funda-
mental properties of dielectrics; diamagnetism; distinction
between anode and cathode in the electric discharge through
gases at low pressure; vibration of plates; regelation of ice;
alloys of steel; and optical glass. If further proof be needed of
the impress which Faraday left on physics, one need only recall the
many times his name is used in physical terminology. Thus, we
have the farad; the Faraday rotation (of a beam of polarized light);
the Faraday dark space (in an electric discharge at low pressure);
Faraday's Laws of Electrolysis; the Faraday cylinder; and the
Faraday "ice pail" experiment.
Well may this simple, modest, self-taught philosopher be given
a conspicuous place among the great benefactors of mankind.
6. Joseph Henry.-Any account of Faraday's work, however
brief, should be accompanied by at least a mention of the
researches of one of Faraday's illustrious contemporaries, the
American physicist, Joseph Henry, whose memory is honored
by the name of the unit of inductance, the henry, which bears
to electrokinetics a relation identical with that of the farad to
electrostatics.
Henry was born at Albany, New York, in 1799. Like Faraday,
he became an apprentice at an early age, and, also like Faraday,
his attention was called to science by the chance reading of a
book on "Experimental Phylosophy." He entered the Albany
Academy and, in 1826, became professor of mathematics.
During the short summer vacations, only a single month, he
experimented with such apparatus as he could make with his own
hands. Independently of, but almost simultaneously with,
Faraday, Henry discovered the laws of electromagnetic induc-
tion. But Faraday's results were published earlier than Henry's,
and Faraday is justly credited with the discovery.
Henry devoted considerable attention to the improvement of
electromagnets, invented by Sturgeon in England in 1825.
Henry insulated the wire by silk thread; wound the wire in
several layers instead of one; and introduced the ''spool'' con-
SEC. 5] HENRY 71
struction, which is standard today. He built one magnet which
could support over four hundred times its own weight.
Extending these principles, Henry designed and constructed
the first electric motor operating by an electromagnet. This
was a sort of "push and pull" device consisting of an electro-
magnet, supported on pivots, so that it could rock back and
forth between two permanent magnets, the current which ener-
gized the electromagnet being reversed at each ''rock'' by a
suitable combination of mercury cups.
Henry's experiments on the operation of an electromagnet,
located at a distance from the battery, led him to a clear under-
standing of the optimum conditions for securing maximum trac-
tive effect. Thus, if the magnet were a long distance from the
battery, both the cells of the battery and the "spools" of the
electromagnet should be connected in series. While, if only short
wires joined the magnet to the battery, t4e series connections
should both be replaced by parallel connections. These discov-
eries led directly to the commercial development of the telegraph.
In the course of these experiments, Henry's attention was
called to the intensity of the spark observed when a break is
made in a circuit containing a long coiled wire through which a
current is flowing. The effect was found to be intensified by
coiling the wire into a close helix; but was diminished, or even
absent, when the circuit consisted of only a short piece of uncoiled
wire. Repeated trials with many modifications led him to the
conclusion that the effect was due to the inductive action of the
coil of wire upon itself. And in 1832, he announced the discovery
of the phenomenon of self-induction in a paper 1 entitled "Elec-
trical Self-induction in a Long Helical Wire."
In 1832, Henry was appointed to the chair of natural philos-
ophy at Princeton. Here he continued his investigations on
induced currents, and, in 1842, he was led to the important
discovery that the discharge from a leyden jar is, under certain
conditions, oscillatory. The importance of this discovery was
not appreciated until over a decade later, when Lord Kelvin, as a
result of a theoretical study, reached the conclusion that the
discharge should be oscillatory.
The creative period of Henry's work in electricity ends with
1846, when he was appointed secretary of the Smithsonian
Institution. Here, in spite of his official duties, he found time
1 Amer. J our. Sci., vol. XXII~ p. 408 (1832).
72 HISTORICAL SKETCH, THIRD PERIOD [CHAP. III
for an active interest in science. He organized a systematic
method for the collection of meteorological data and for the
making of daily weather maps. He contributed much to the
improvement of lighthouses. And he was instrumental in
organizing, under a federal charter, the National Academy of
Sciences.
Henry died in 1878. When we consider his comparative isola-
tion from the scientific atmosphere of Europe and his meager
opportunities for carrying on research, we may well call him
the '• Dean of American Sci en tis ts.''
6. James Clerk Maxwell.-It would be difficult to pick out
two eminent scientists whose beginnings differed from each
other more than did Maxwell's and Faraday's. Faraday came
of very humble parentage; Maxwell, from a long line of dis-
tinguished ancestors. Faraday's early life was lived almost in
poverty; Maxwell's family had abundant means. Faraday
received only the most rudimentary education; Maxwell was
given every advantage of school and university. They differed
quite as much in their aptitude for scientific work. Faraday
was one of the greatest exponents of experimental science which
the world has ever seen; while Maxwell is one of the greatest
figures in the entire history of theoretical physics. And yet
both made indispensable and mutually supplementary contri-
butions to the classical theories of electromagnetics.
Maxwell was born in Edinburgh in June, 1831, the very year
when Faraday discovered electromagnetic induction. After
an unsuccessful experience with a private tutor, young Maxwell,
at the age of ten, was sent to the Edinburgh Academy.
He was a friendly boy, though never quite amalgamating with the
rest. But, however strange he sometimes seemed to his companions,
he had three qualities which they could not fail to understand: agile
strength of limb, imperturbable courage, and profound good nature. 1
He was an all-around good student, but in geometry he made a
particularly brilliant record. For, on Apr. 6, 1846, two months
before he was fifteen years old, he presented a paper, by invitation,
to the Royal Society of Edinburgh on "The Description of
Oval Curves and Those Having a Plurality of Foci." At the
same time, he was busy with experimental work on magnetism
and polarized light. When he left the Academy, in 1847, he
1 GLAZEBROOK: "James Clerk Maxwell and Modern Physics."
PLATE 6.-Maxwell.
(Facing page 72)
SEC. 6] MAXWELL 73
was "first in mathematics and in English and nearly first m
Latin."
In November, 1847, Maxwell entered the University of
Edinburgh, where, in addition to his studies in mathematics,
natural philosophy, and logic, he prepared material for two
important papers, the one on "The Theory of Rolling Curves,"
and the other on '' The Equilibrium of Elastic Solids' '-all this
before he was nineteen years old!
But his interests were in experimental work as well as in
theory. His description of his little laboratory on the family
estate at Glenlair, where he spent his vacations, is interesting.
In a letter to a friend, he writes:
I have regularly set up shop now above the wash-house at the gate, in
a garret. I have an old door set on two barrels, and two chairs, of which
one is safe, . . .
On the door there is a lot of bowls, jugs, plates, jam pigs, etc., con-
taining water, salt, soda, sulphuric acid, blue vitriol, plumbago ore;
also broken glass, iron, and copper wire, copper and zinc plate, bees'
wax, sealing wax, clay, rosin, charcoal, a lens, a Smee's galvanic appara-
tus, and a countless variety of little beetles, spiders, and wood lice,
which fall into the different liquids and poison themselves . . .
I am experimenting on the best methods of electrotyping. So I am
making copper seals with the device of a beetle. First, I thought a
beetle was a good conductor, so I embedded one in wax (not at all cruel,
because I slew him in boiling water in which he never kicked), leaving his
back out; but he would not do. Then I took a cast of him in sealing
wax, and pressed wax into the hollow, and blackleaded it with a brush;
but neither would that do. So at last I took my fingers and rubbed it,
which I find the best way to use the blacklead. Then it coppered
famously.
A little later, he writes of experiments with jelly, under strain,
exposed to polarized light; of "a notion for the torsion of wires
and rods"; of plans for studying the relation between optical
and mechanical constants, etc.
In 1850, Maxwell entered Trinity College, Cambridge, from
which he graduated, in 1854, with high honors. After a short
tenure as fellow at Trinity, he was appointed, in 1856, to the
Professorship of Natural Philosophy at Aberdeen, a post which
he held until 1860.
From 1860 to 1865, he was Professor at King's College, London.
It was from here that some of his most important papers were
74 HISTORICAL SKETCH, THIRD PERIOD [CHAP. III
published, such as "Physical Lines of Force" ( 1862), and his
greatest paper, '' A Dynamical Theory of the Electromagnetic
Field," read Dec. 8, 1864.
After a retirement of several years, Maxwell was, in 1870,
elected to the newly founded Professorship of Experimental
Physics at Cambridge. In this capacity, he superintended the
planning and equipment of the now famous Cavendish Labora-
tory, of which he was director until his untimely death, on
Nov. 5, 1879. A large proportion of Maxwell's papers, over one
hundred in number, may be grouped under three headings: color
vision; molecular theory; and electromagnetic theory.
The work on color vision was undertaken to make a quantita-
tive study of the physical facts of the theory of color sensations
proposed by Thomas Young, according to which any luminous
sensation is the result of exciting, in the eye, three primary
sensations, respectively, red, green, and violet. Young based
his theory on his own experiments on spinning colored disks,
by which he showed that any color may be matched by a mixture,
in proper proportions, of these three primaries. Maxwell first
verified this conclusion by experiments with his "color top."
Then, to give his experiments greater definiteness, he invented a
"color box," by means of which he could mix spectral colors,
instead of colored papers. He expressed his measurements in a
series of color equations, from which can be deduced the value of
any color in terms of the three fundamental colors chosen as
primaries. The results he expressed graphically both by the
use of the '' color triangle,'' previously em ployed by Young, and
by plotting the three elementary sensations on a wave-length scale.
In the preface to '' The Scientific Papers of James Clerk
Maxwell," the editor, W. D. Niven, writes (1890):
These observations, on which Maxwell expended great care and
labour, constitute by far the most important data regarding the com-
binations of color sensations which have yet been obtained, and are of
permanent value whatever theory may ultimately be adopted of the
physiology of the perception of colour.
Maxwell, himself, clearly points out that his measurements are
concerned with the physics, not the physiology, of color vision.
For, at the conclusion of his paper, "Experiments on Colors as
Perceived by the Eye," 1 he writes:
1 Trans. Roy. Soc. Edinburgh, vol. XXL
SEC. 6] MAXWELL 75
The laws of sensation can be successfully investigated only after the
corresponding physical laws have been ascertained, and the connection
of these two kinds of laws can be apprehended only when the distinction
between them is fully recognized.
Maxwell's work on molecular theory is so extensive as not to
admit of brief abstract. He deduced, on rigorous mathematical
grounds, the law of the distribution of velocities among the
molecules of a gas. From consideration of coefficients of vis-
cosity of gases, he determined the value of the "mean free path"
and showed that the value thus obtained was in substantial
agreement with the value which he computed from data on the
diffusion of gases. He proved that
. . . the physical condition which determines that the temperature
of two gases shall be the same, is that the mean kinetic energy of agita-
tion of the individual molecules of the two gases are equal.
And he brought to bear upon the whole subject mathematical
methods "in their generality and elegance, far in advance of
anything previously attempted on the subject." Indeed,
Maxwell is the co-founder with Clausius (1822-1888) of the
kinetic theory of matter.
The electromagnetic theory is so intimately connected with
modern physics as to warrant treatment in a separate chapter,
in which, however, we shall not follow Maxwell's method.
Maxwell states clearly that much of the experimental basis of
his theoretical researches in electromagnetism was derived
from Faraday's works. In the preface to the treatise, "Elec-
tricity and Magnetism," Maxwell writes:
Before I began the study of electricity I resolved to read no mathema-
tics on the subject till I had first read through Faraday's "Experimental
Researches on Electricity."
It is a noteworthy fact that in the early part of his speculations
Maxwell made frequent use of analogies. In his first electrical
paper on "Faraday's Lines of Force," published in 1856, he
writes:
In order to obtain physical ideas without adopting a physical theory
we must make ourselves familiar with the existence of physical analogies.
He then proceeds to develop the similarity between the electri-
cal field and the steady flow of an incompressible fluid. In this
76 HISTORICAL SKETCH, THIRD PERIOD [CHAP. III
analogy, the mathematical expressions descriptive of one phenom-
enon are applicable to the other, provided "velocity of fluid" be
correlated to '' electrical force" and "difference of fluid pressure"
to "difference of electrical potential."
In another connection, to get a mechanism to simulate the
observed facts that "lines of force tend to contract longitudinally
and to repel each other laterally," he likened a line of force to a
string of spherical cells or sacks, with elastic walls and filled
with an incompressible fluid, the whole string being in rotation
about its axis. Each cell assumes the form of an oblate spheroid.
Thus, the string tends to shorten longitudinally, and the equa-
torial bulge of each cell pushes neigh boring strings laterally.
If one end of such a string under tension be suddenly moved
laterally, the disturbance will travel with finite velocity along
the string.
By gradual steps, these analogies pointed out the real electro-
magnetic relations. And Maxwell clearly showed that all the
facts of electrodynamics could be attributed to the action of a
medium, the properties of which he deduced by rigorous mathe-
matical reasoning. Maxwell's "Treatise on Electricity and
Magnetism," published in 1873, which embodies these results,
ranks with Newton's "Principia" as one of the most important
books in all science.
The electromagnetic theory, as developed by Maxwell and
extended by his successors, must be regarded as the crowning
achievement of all but the last decade of the nineteenth century.
And it is still, after more than 50 years, a large factor, perhaps
even a dominating one, in present-day physics, whether one
judges its greatness by the fertility of its application to the
problems of pure science or by. the more practical test of its
value in the applied science of radio communication. For
this reason, it is proper to discuss the elements of the electro-
magnetic theory as an integral part of the fascinating story of
modern physics rather than as a part of the history of physics.
As previously mentioned, the third period in the history of
physics ends about 1890, with the discovery of photoelectricity,
the first of a group of phenomena the explanation of which
require radical revision of some of the basic tenets of classical
physics. Our discussion of the development of classical physics
SEC. 6] MAXWELL 77
from 1800 to 1890 has, of necessity, been confined to a few out-
standing topics, such as: the mechanical theory of heat and the
doctrine of the conservation of energy; the revival of the wave
theory of light and, as we shall see in Chap. IV, its consolida-
tion with the electromagnetic theory; the discovery of the
phenomena of electromagnetism; and the development of the
kinetic theory of matter. These fundamental principles seemed
to cover so thoroughly and so adequately the entire field of
physics that, as previously mentioned 1, a few scientists, some of
them noted physicists, were inclined to the belief that all the
important discoveries in physics had been made and that research
henceforth would be concerned with studying details and with
improving the technique of measurement "so as to investigate
the next decimal place." How different our outlook in 1928!
The investigator in science is frequently likened to the explorer
who penetrates the unknowns of the earth's surface. The
similarity is striking, but there is this very important difference:
"\Vhereas the surface of the earth is finite, the extent of the
unknowns in science may, so far as anything we now know, be
infinite, at least in comparison with the present attainments of
the human intellect. The Galileos, the Newtons, the Faradays,
and the Maxwells will always find unsolved problems. Indeed,
the progress of science is concerned, perhaps, not so much with
solving problems as with finding problems to solve.
1 Introduction, Sec. 8.
CHAPTER IV
THE ELECTROMAGNETIC THEORY OF LIGHT
It was pointed out, in Chap. III, that the wave theory of
light, revived by Young in 1800, was put on a solid foundation
by the discovery of the phenomena of interference and by the
introduction of the concept of transverse vibrations to explain
the phenomenon of polarization. This theory, as developed by
Young, Arago, Fresnel, and their contemporaries, postulated an
"ether" in which a light disturbance was propagated by trans-
verse vibrations of a purely mechanical nature. The obvious
difficulties more or less inherent in this "mechanical" ether in
large part disappeared with the development of the electro-
magnetic theory of light by Maxwell who, while still r(ttaining
all the advantages of a transverse-wave theory, substituted for
the mechanical vibrations of Young and Fresnel the vibrations
of electric and magnetic vectors or, more simply, an electro-
magnetic vibration. It shall be the purpose of this chapter to
develop the fundamental concepts of the electromagnetic theory
of light to the point where the student may the more readily
appreciate its conflict with the modern quantum theory, which
will be discussed in subsequent chapters.
1. The Electrostatic System of Electrical Units.-Two electric
charges whose magnitudes are designated by Q1 and Q2 act upon
each other by a force f given by
f = !e Q1Q2
r2
(1)
whm·e E is the dielectric constant of the medium between the
charges and r is their distance apart. The quantity e may be,
for convenience, defined as the "ratio of the capacity of a parallel
plate condenser with the medium concerned between its plates
to the capacity of the same condenser when a vacuum is substi-
tuted for the dielectric." The electrostatic system defines in
the fallowing order: quantity of electricity (Q); current (i); (electro-
static) field strength (F).
78
SEC. 1] ELECTROSTATIC UNITS 79
(a) The electrostatic unit of quantity of electricity is a u point''
charge 1 of such magnitude that if two such charges be placed 1
cm., in vacuo, from each other, the force of repulsion between the
two will be 1 dyne. Similarly, if there is a mutual force action
of Q dynes between an unknown point charge and a unit point
charge when the two charges are 1 cm. apart, in vacuo, the
unknown charge contains Q units of electricity. In equation (1),
f is in dynes if Q1 and Q2 are measured in electrostatic units, as
above defined, and r is given in centimeters. The customary
definitions of positive and negative charges are assumed.
(b) Current, symbol i, in the electrostatic system is measured
by the, number of electrostatic units of quantity per second
passing through a surface S. That is,
- dQ
i - dt (2)
where dQ is the quantity of electricity passing through Sin time
dt.
At any point, the current per square centimeter taken at right
angles to the direction of flow is called the current density, symbol
j, at that point. Both current and current density are vector
quantities. The current density at a point may, therefore, be
represented by a vector drawn from the point in the direction of
current flow, the length of the vector being proportional to the
magnitude of j. If, from the point in question as origin, a set of
rectangular coordinate axes be drawn, then the projection of j on
each of these axes may be taken as the component of j in the
direction of that axis. If the axes be x, y, and z, the components
of current density along these axes may be designated as jx, jy,
and j;, respectively.
The concept of current density and of the components of
current density may be more readily grasped by considering
the flow of current through a large rectangular mass of copper
or through the electrolyte in a large rectangular tank. Imagine
the current to enter at one corner and to leave at the opposite
corner. Between these two points the current will follow the
well-known "lines of flow." If the direction of flow at any point
P be found, then the current density at that point is the current
per square centimeter through a surface placed at right angles
to the direction of flow. If one corner of the tank be taken as an
1
That is, a charge placed upon a body whose dimensions are small com-
pared to the distance to neighboring bodies or charges.
80 ELECTROMAGNETIC THEORY [CHAP. IV
origin of coordinates x, y, and z, assumed parallel, respectively, to
the three edges of the tank, then the current density at P in the
x direction, i.e., jx, is measured by the current per square centi-
meter flowing through a surface containing P and placed parallel
to the yz plane. And, similarly, one gets jy and jz.
(c) The (electrostatic) field strength, symbol F, at a point is
defined as the "force in dynes acting on a unit electrostatic
charge placed at the point." Since Fis, thus, a vector quantity,
we may resolve it into components along the x, y, and z axes,
respectively. These components we shall call X, Y, and Z.
Thus,
X = F cos Bx
where Bx is the angle between the direction of F and the x-axis.
And so for Y and Z.
From the definition of field strength, it follows that the force
f acting on a charge Q placed in a field of strength F is
f = F ·Q (3)
2. The Electromagnetic System of Electrical Units.-Two
magnetic poles whose magnitudes are designated by m1 and m 2 act
upon each other by a force f given by
f = ~ m1m2 (4)
µ r2
where µ is the permeability of the medium between the poles and
r is their distance apart. The electromagnetic system defines,
in the following order: magnetic pole (m); magnetic field (H);
current (i'); quantity of electricity (Q').
(a) The unit magnetic pole is a "point" pole of such strength
that if two such poles be placed 1 cm., in vacuo, from each other,
the force of repulsion between the two will be 1 dyne. Similarly,
if there is a mutual force action of m dynes between an unknown
pole and a unit pole when the two poles are 1 cm. apart, in vacuo,
the unknown pole is said to have a strength of "m units." In
equation (4), f is given in dynes if m1 and m2 are measured in the
above units, and r is in centimeters. The customary definitions
of positive and negative poles are assumed.
(b) The (magnetic) field strength, symbol H, at a point is
defined as the "force in dynes acting on a unit magnetic pole
placed at the point." As in the case of the electrostatic field,
we may resolve H into components along the x, y, and z axes,
respectively. We shall call these components a, {3, and 'Y·
SEC. 3] RATIO OF UNITS 81
(c) The Electromagnetic Unit of Current.-The definition of
'"current'' in the electromagnetic system is based on the experi-
mental fact that a wire carrying a current experiences a lateral
force when placed in a magnetic field. The electromagnetic unit
of current is defined as a "current of such strength that a wire,
carrying the current and placed at right angles to a unit magnetic
field, experiences a force of 1 dyne per centimeter in a direction
perpendicular to both field and wire." The symbol for current,
in these units, is taken as i'. (The ampere is one-tenth of this
unit.)
Current density and the components of current density are defined
in the electromagnetic system just as in the electrostatic system.
As in the case of current, the electromagnetic quantities are
distinguished from the corresponding electrostatic quantities
by the prime ('). Thus, we have j', j:, j~, and j; for current
density and the corresponding components.
The electromagnetic unit of quantity of electricity is defined as
the ,., quantity of electricity carried per second through a surface
through which 1 electromagnetic unit of current is flowing." We
shall use the symbol Q' to designate quantity of electricity
measured in these units. Since the order in which current and
quantity are defined in the electromagnetic system is just the
reverse of the corresponding order in the electrostatic system,
it is logical to use the converse of equation (2) to determine
quantity of electricity in the electromagnetic system; i.e.,
Q' = f i'dt (5)
3. Ratio of the Two Systems of Units.-If we measure a given
current first in electrostatic units and then in electromagnetic
units, the numeric, i, in the former case, will be much larger than
the numeric, i', in the latter, since the electrostatic unit is much
smaller than the electromagnetic unit. The ratio i/i' of these
two numerics is a very important physical quantity, which 1s
customarily designated by the symbol c. Thus,
_ i _ number of electrostatic units in a given current ( )
c - if - number of electromagnetic units in the same current 6
In like manner,
Q
c = - (7)
Q'
82 ELECTROMAGNETIC THEORY [CHAP. IV
This quantity c, therefore, is equal to the number of electrostatic
units of current (or quantity) in 1 electromagnetic unit of current
(or quantity).
Because of the fundamental importance of this ratio, it has
been the subject of many researches by many different methods.
Among the more important determinations may be mentioned
the following:
Weber and Kohlrausch (1856) ............ c = 3 .107 X 10 10
Lord Kelvin (1869). . . . . . . . . . . . . . . . . . . . . 2. 82
Rowland (1889). . . . . . . . . . . . . . . . . . . . . . . . 2. 98
Abraham (1890).. .. . .. . . . . . . . . . . . . . . . . . 2. 991
Rosa and Dorsey 1 (1907) . . . . . . . . . . . . . . . . 2. 9971
4. Some Fundamental Formulae. (a) La Place's Law.-The
magnetic field dHP, at a point P, due to a short straight element
ds of a circuit through which a (constant) current i' is flowing
is given by
i'ds .
dHP = - r2 sm o (8)
where ds is the length of the element; r is the distance from P
to the element; and () is the angle between the element and the
line from the element to the point P. It should be remembered
that dH is perpendicular to the plane including 8.
(b) Field H Due to Current in a Long, Straight Wire.-If a
point P be a cm. distant from a very long,2 straight wire carrying
a current i', the magnetic field HP at P is given by
2i'
Hp= - (9)
a
This can be readily proven by integrating equation (8) over the
entire length of the wire.
(c) Work Done in Carrying a Unit Magnetic Pole Once around
a Wire Carrying Current.-A magnetic pole of strength m, placed
a cm. from a long, straight wire, will experience a force f given,
by use of equation (9), by:
2i'
f = mHP = m · - a (10)
An excellent account of the various methods for measuring c and a detailed
1
description of the method which they employed will be found in the article
by RosA and DORSEY, Bureau of Standards, Bull., vol. 3, pp. 433, 541,
and 605 ( 1907).
2 That is, the length of the wire is very large compared to a.
SEC. 5] MA.XWELL'S EQUA.TIONS 83
If the pole be now carried around a circular path of radius a, con-
centric with the wire, an amount of work W will be done which
is given by
2i'
W = m · - · 21ra
a
= 41ri' · m (11)
Since equation (11) shows that W is independent of a, it is
not necessary to restrict the motion of the pole to a circular path.
If, therefore, a magnetic pole of strength m, starting at a point P,
is caused to move along any path whatsoever which "links"
once with a current i' (i.e., if the pole ultimately returns to point
.P), the work done is given by equation (11). The work will
be positive or negative according as the pole moves against, or
in the direction of, the force.
5. Maxwell's Differential Equations of the Electromagnetic
Field.-(a) Consider a point O inside a conductor through which
a current is flowing. It may help
the student to visualize the
arrangement if the conductor be
thought of as the large tank of a'
c - - - - - - -,b
r
electrolytic solution with current
I
entering at one point and leaving
at another point, 0 being some-
where within the tank. With point
dy
:r'I
O as origin, construct a set of rec-
tangular axes Ox, Oy, and Oz (Fig.
1), so oriented, for simplicity, that
the current in the neighborhood of
O flows parallel to the z-axis, say z Frn. 1.
in the direction zO. Let the density
of this current be j;, assumed constant in the immediate
neighborhood of 0.
Consider a small elementary rectangle Oabc in the x-y plane, the
length of whose sides are dx and dy, and whose area, therefore, is
dx · dy. The current di; flowing through (i.e., in this case, at
right angles to) this rectangle is given by
di; = j;dxdy (12)
Further, assume that in the neighborhood of O there is produced
by the currents and, possibly, by external -magnetic poles a
magnetic field whose components a, a', {3, {3' along the respective
84 ELECTROMAGNETIC THEORY [CHAP. IV
sides of the rectangle are as shown in the figure. In general, a'
will differ from a, al}d {3' from (3.
If, now, a unit magnetic pole, starting from 0, be carried,
say counterclockwise, once around the rectangle, the path of
the pole will "link" once with the current di; flowing through
the rectangle. The work dW done during this process may be
computed in either of two ways: (a), by equations (11) and (12)
dW = 41rdi:
= 41rj: · dxdy; (13)
or (b) we may compute the work done in carrying the pole along
each of the four rectilinear paths Oa, ab, be, and cO and then
add. Thus,
dW = a· dx +
{3' • dy - a'· dx - (3 · dy (14)
We can simplify equation (14). Let the curve def, in Fig. 2,
show the variation of (3 with x in the
f neighborhood of point 0, (3 and {3' of
Fig. 2 being identically the same as (3
and {3' of Fig. 1, and dx of Fig. 2 being
the same as dx of Fig. 1. If the slope
d of this curve in the neighborhood of e
be ·a[3 / ax, it at once follows that
--~~----=-__._~~-x
dx
FIG. 2.
{3' = (3 + ax
a[3 dx (15)
and, similarly,
dad
a =a+-
I
ay
y (16)
If the values of a' and of {3' from equations (15) and (16) be put
in equation (14), there results
dW = a · dx + ({3 + !! dx )ay - ( a + !; dy )dx - {3 • dy
- - -aa)
= ( a{3
ax ay
dxdy (17)
It is obvious that equations (13) and (17) must give identical
values of dW. We may, therefore, write,
4~j;.axdy = (!! - !;)axdy
. ., a[3 aa (18)
• . 411"Jz = ax - ay
SEC. 5] MAXWELL'S EQUATIONS 85
In this equation, current is measured in electromagnetic units.
"\Ye may introduce current measured in electrostatic units by
using equation (6) according to which j; = jz/c. We may, then,
write,
41r . _ iJ/3 _ da
c Jz - ax ay (19)
(b) In setting up Fig. 1, we arbitrarily oriented the system of
axes so that Oz was parallel to the direction of current at 0.
For this orientation, j: and j~ are each zero. This limitation is
obviously unnecessary. If the coordinate axes have any arbi-
trary direction with respec~ to
the current, j: and j~ will not, yJJ
in general, be zero, and we shall
have for each of them an equa-
tion similar to equation (19).
We might derive each of these p
equations in the same way as
equation (19) was derived. But
it is more instructive to consider o
,,,,,,______._,-,-r---.x:, a
the symmetrical relations exist- // :/
/ I
ing among the several quantities. / •
Let Fig. 3 represent the coor-
dinate axes x, y, z, with which
Z.,T
are associated, respectively, the Fm. 3.
components a, {3, 'Y, of a mag-
netic field. Let Op be a line drawn so as to make equal angles
with the three axes. Equation (19) shows a relation between
jz and quantities belonging to the xy plane. An exactly similar
relation must exist between jx and quantities belong to the yz
plane. To find this relation, imagine the whole figure rotated
clockwise 120 degrees about Op as an axis, so that the new posi-
tion of the y-axis coincides with the previous position of the x-
axis. The direction of jx will then coincide with the previous
direction of jz, and we may represent the transitions by
. .
Jy --1' Jx
j 1· --1' Jz (20)
. .
Jz --1' Jy
where the symbol --1' means "takes the place of." We may
represent these cyclical relations by Fig. 4. Havin~ derived one
equation, such as equation (19), the other two equations may be
86 ELECTROMAGNETIC THEORY [CHAP. IV
written by means of the "replacements" indicated in ~ig. 4.
Thus, we have
41r . a'Y a[3
-J., = - - -- (21a)
c ,w ay az
41r . _ da _ d'Y
c Jy - az ax (21b)
41r. _ a[3 _ aa
c Jz - ax ay (21c)
Equations (21) are in entire harmony, as, of course, they
should be, with our elementary concepts. The left-hand
members contain currents. The right-hand members contain
space rates of variation of magnetic fields. For example, if a
current flows, let us say, along a very long wire, in the x direction,
p
I
C- I
I
+
- I +
!:hJJ - I +
~
I
G1 - I +
- I +
I
\
- I +
- I +
I
- I
I
+
"\. a p
~, a
1--B--j
1
c b
FIG. 4. Frn. 5.
the resulting magnetic field around the wire is independent of
x, but it does depend on y and z. This is in accord with equation
(2la). Further, the left-hand side of equation (2la) represents
the work which would be done in carrying a unit magnetic pole
once around an area 1 cm. square through which a current jx is
flowing. If ay = az = 1 cm., the right-hand side of equation
(21a) is easily seen to give exactly the same quantity of work.
(c) Displacement Currents.-Maxwell regarded electricity as
behaving like an incompressible fluid. Consider the circuit
shown in Fig. 5. A battery B is connected to a long resistance
ab on which is a sliding contact c. Points a and c are connected
through galvanometers G1 and G2 to the plates of a condenser C.
(Assume, for the moment, that there is a vacuum between the
plates.) If c, starting from coincidence with a, moves uniformly
toward b so as to increase uniformly the potential difference
SEC. 5] MAXWELL'S EQUATIONS 87
between the plates of the condenser, thereby increasing its charge,
a constant current, say i, will flow through the galvanometer
G1 , and an identical current through G2. Except for the conden-
ser, the circuit aG 1G2c might be regarded as a continuous circuit.
No current actually flows between the plates of the condenser;
but during the charging process, the electric field between the
plates is increasing at a uniform rate. We may regard the
phenomena which happen in the space between the plates to be
exactly the same as if a constant current i, the displacement
current of Maxwell, were really flowing between the plates. We
can correlate this rate of increase of electric field with the equiva-
lent (or displacement) current i, as follows:
Let the area of either plate of the condenser be A. (Assume
the distance between the plates very small compared to their
lateral dimensions.) Then, when, at any instant, there is a
charge Q on the plates, the number of lines of force per square
centimeter between the plates, which number is numerically
equal to the field strength F between the plates, is given by
F = 41r~ (22)
The simultaneous rates of change of F and Q are given by
dF = 41r·l__. dQ (23)
dt A dt
But
dQ .
- = i
dt
and
1 dQ .
A dt = J
where j is the equivalent current density through the surface. We
may therefore write equation (23) as
dF .
dt = 47rJ (24)
. 1 dF
: . J = 41r. dt (25)
That is to say, if, at any point in (empty) space, the electric
field Fis changing at a rate dF /dt, we may regard the phenomena
resulting therefrom as identical with what would happen if
a current of density j given by equation (25) were actually
flowing through the same region in the direction of the change
of F.
88 ELECTROMAGNETIC THEORY [CHAP. IV
If, instead of a vacuum, there is between the plates of the
condenser a medium whose dielectric constant is e, equation (22)
should be written
41r Q
F =-·- (26)
e A
and equation (25) would, then, become
• E dF
J = -·- (27)
41r dt
An alternative method of deriving equation (25) lays greater
emphasis on the concept of electricity as behaving like an incom-
pressible fluid. Consider, first, an analogy. Imagine a large
vessel, filled with an incompressible fluid, say water, into the
side of which projects a tube t, (Fig. 6). Over the inside end oft
b t
I
r
r
I
r
I
I
Frn. 6. FIG. 7.
is fitted a completely deflated toy balloon b. Describe around
b an imaginary sphere of radius r. Now, let a volume of water
V be inserted into b through t. The balloon will be inflated and
will occupy a volume V. An equal volume of water will be
displaced outward through the surface of the imaginary sphere.
And it is readily seen that the average displacement through
unit area of the sphere is V / (41rr2 ).
The electrical analogue is obvious. Suppose that an insulated
metal ball (b, Fig. 7) is being charged at a uniform rate by means
of the wire w. 'Ne describe the phenomenon by saying that the
charge remains on the ball and that there is produced an electric
field in the neighborhood. But we might consider that the
electricity behaves as does the water in the analogy and that, if
SEC. 5] MAXWELL'S EQUATIONS 89
the ball has received a charge Q, a quantity of electricity equal to
Q has been displaced outward through the surface of an imaginary
sphere of radius r concentric with b. This imaginary displace-
ment of electricity is equivalent to the actual change of electric
field. For, consider the ball to be surrounded by a medium
whose dielectric constant is €. Then, if, at a given instant, the
charge on the ball is Q, the electric field F at the surface of
the sphere will be (neglecting the dissymmetry introduced by the
wire)
F
1
= - -
Q (28)
€ r2
and the rate of increase of F as Q increases will be
dF 1 1 dQ
dt = ~ · r 2 • dt (29)
41r 1 dQ
= --;- · 41rr 2 • dt
But 41rr 2 is the area of the imaginary sphere; and dQ/dt, the cur-
rent flowing onto the ball, is also the equivalent (i.e., displacement)
current flowing outward through the entire surface of the sphere.
The equivalent current per unit area through the sphere, i.e.,
the current density j, is, therefore, given by
. 1 dQ
J = 41rr 2 • rfi
Accordingly, we may write equation (29) in the form
dF 4-1r .
- =-·J
dt €
or, as before (equation (27)),
• € dF
J =-·- (27)
41r dt
It should be emphasized that the direction of this equivalent
current density j is the same as the direction of the rate of change
of F. In the foregoing discussion,
we have considered the change of
F to be due to a change in the
>______ +k-~- r----------f
Frn. 8.
magnitude of the charge producing
F. But the field F at a given point, due to a charge Q, may be
changed by changing the position of the charge. Thus, consider
a charge +Q (Fig. 8) moving toward the right with a velocity
90 ELECTROMAGNETIC THEORY [CHAP. IV
v. The electric field at P will increase. Assuming the system
to be in a vacuum, the electric field FP at P is given by
Q
Fp -:- 2
r
at the instant when Q is distant r from P. At that instant, the
rate of change of F P is given by
dFP = _ 2 Q. dr
dt r 3 dt
. . dr - v
. dt -
. dFp = _ Qv
2 3
dt r
and, from equation (27), the equivalent current density jp at P
is given by
. ·e 2Qv
Jp = - - · - (30)
41r r 3
In the case just considered, the direction of F remains constant.
But we may have a change of F occasioned by a change in its
!Jr------r
-tQ
d ~.dJi'
~
Ij
L----r .z;;
Fm. 9.
direction without change of magnitude. In Fig. 9, let the charge
+Q pass at a uniform velocity v from point P1 to P 2 • When Q is
at P1, the field F1 at Pis
Q
F1 = -r2
When Q reaches P2, the field F2 at P has the same numerical value
but has changed in direction, and there has been a change 11F in
F, as shown. If 11t has been the time required for Q to pass from
P1 to P2, the average rate of change of F is given by
~F
11t
It is left to the reader to compute the current density of the
displacement current equivalent to this change in F. (For
simplicity, take the case where r is large compared to d.)
SEC. 5] MAXWELL'S EQUATIONS 91
In general, neither F nor its rate of change will be directed
along one of the coordinate axes x, y, z. And just as F has
components X, Y, Z so dF /dt has corresponding components
ax aY az
_,_,_
at at at
(d) We may now regard equations (21), which were developed
by considering the case of an actual current of density j, to apply
to the case where, instead of the actual current, we have a
so-called "displacement current" resulting from a (time) rate
of change of electric field, provided we make the substitutions
for the components of current density, in accordance with
equation (27), viz.,
• e ax
J.r; = 41r at
. e aY
Jy = 41r at
e az
Jz = 41r at
We then have
eaX
-- --
a,y iJ{3
(31a)
cat ay az
e aY aa a,y
=- - (31b)
cat az ax
e az a[3 aa
=- - - (31c)
cat ax ay
(e) These equations (31) are the first set of Maxwell's equations
of the electromagnetic field. It is to be noted that they express
the time rates of change of electric fields (left-hand members) in
terms of the space rates of change of magnetic fields.
These equations contain no essentially new physical principles;
they simply put in rigorous mathem~tical form some very ele-
mentary concepts. For, consider, again, a long, straight wire
w1w2 through which a constant current is flowing. Such a wire
is surrounded by a magnetic field. We say that the current
produces the field. But let us examine the phenomenon a little
more closely. We regard the current, nowadays, as made up of
a succession of moving charges, i.e., electrons-really, negative
charges. A negative charge moving toward the left is equiva-
lent to a positive charge moving toward the right. Let us repre-
92 ELECTROMAGNETIC THEORY [CHAP. IV
sent the current through the wire as a procession of positive
charges (Fig. 10) moving at uniform speed toward the right, say
in the positive direction of x. The magnetic field produced by
such a wire carrying a current is, as we know from experi-
/Ilent, represented by a series of concentric circles, drawn about
the wire as a center and in planes at right angles to the wire.
The direction of these circular fields is clockwise looking in the
positive x direction. One of these circles is represented as drawn
about point a on the wire. The magnetic field at some point, such
as P1, must, in reality, be the sum total of the effects at P1 due to the
motion of each of the individual charges. In other words, each of
the moving charg~s, say e1, produces a magnetic field at P1
irrespective of the motion or presence in the wire of any other
charge such as e2. That is, even if e1 were alone and were moving
;;l· .. :.~~...... .
z
~ ... " ... "~... x
Frn. 10.
in the direction indicated, it would produce a magnetic field at P 1 •
Now, the only means, so far as we know, by which a charge can
produce any effect at some outside point is through the agency
of the electric field which surrounds the charge. So long as the
charge e1 is stationary, although there is an electric field produced
at P1, there is no magnetic field. When, however, e 1 is moving
toward a, the electric field at P1 due to e1 changes, and we observe
that, simultaneously, there is produced at P 1 a magnetic field.
Viewed from this standpoint, therefore, we may regard the
production of the magnetic field around a wire carrying current
as due, in reality, to the change (with time) of the electric fields
due to the moving charges: Or, to generalize, a changing electric
field produces a magnetic field-which is exactly what equations
(31) tell us. Further, a glance at Fig. 10 shows that the direction
of the magnetic field at P 1 is at right angles to the direction of the
rate of change of electric field at that point due to the motion
of e1. This fact is, also, contained in equations (31).
(!) The Second Set of Electromagnetic Equations.-Experiment
shows us that there are mutually reciprocal relations between
electric and magnetic fields. According to the previous para-
SEC. 5] MAXWELL'S EQUATIONS 93
graph, a moving positive charge Q will produce magnetic fields
which are related to its direction of motion, as shown in Fig.
ll(a). On the other hand (Fig. ll(b)), a positive magnetic pole
m, moving toward the right, will generate a counterclockwise
current in a circular loop of wire placed at a. Of course, the
electric field F, produced by the motion of the pole, will exist,
whether the loop of wire be present or not. That is to say, a
changing magnetic field produces an electric field which is
constant if the rate of change of magnetic field is constant.
This is, of course, the law of induced electromotive forces dis-
covered by Faraday.
+~ - - 0-~----
(a)
~-- 8-F__
(b)
Fm. 11.
The phenomena represented in Fig. ll(a) are expressed
mathematically in equations (31). The phenomena represented
in Fig. ll(b) may be represented by an exactly analogous set
of equations which express time rates of change of magnetic
fields (left-hand members) in terms of space rates of change of
electrostatic fields (right-hand members) by making the three
following changes in equations (31)
1. Replace the components of electric field X, Y, Z in the
left-hand members by the corresponding components of
magnetic field a, {3, 'Y; and vice vers[! in the right-hand
members.
2. Replace the dielectric constant e by the permeability µ,
since it is seen from equation (1) and (4) that these two
quantities play analogous roles in the respective phenom-
ena.
3. Change the signs of each term of the right-hand members,
since, by comparing Figs. ll(a) and (b), it is seen that the
direction of the magnetic field in Fig. ll(a) (clockwise) is
opposite to the direction of the electric field in Fig. 11 (b)
(counterclockwise).
94 ELECTROMAGNETIC THEORY fCHAP. IV
There then results the second set of electromagnetic equations,
viz.,
µ
- ·at
- =
aa -(az _ aY) (32a)
c ay az
a[3
µ
-·- =
c at
-(ax_ az)
az ax (32b)
µ a'Y
- · at
c
- = -(~! - ~i) (32c)
Equations (32) might have been derived quite independent
of equations (31) by a method exactly analogous to the method
used for the derivation of equations (31). 1
6. The Differential Equations of the Electromagnetic Wave.-
Ca) We may combine equations (31) and (32), as follows:
1. Differentiate both sides of equation (31a) with respect
to t:
e ax 2
a 'Y2
a (3 2
c· at 2 = ayat - -az-at (33)
2. Differentiate equation (32b) with respect to z:
µ a2(3 a2x a2z
-c . azat = - az 2 + axaz (34 )
3. Differentiate equation (32c) with respect to y:
µ a2'Y a2 Y a2x
c· ayat = - axay + ay 2 (
35 )
4. Replace each of the two right-hand terms of equation (33)
by their equivalents, as determined from equations (34)
and (35):
(36)
5. By adding and subtracting a2 X/ax 2 within the bracket
on the right-hand side of equation (36), we have, after
rearrangmg,
µe
c2 •
a2
dt
x (aaxx + aayx + aazx) (aaxx + axay
2 =
2
2
2
a
2
a z)
+ axaz
2
2 -
2
2
2
Y
2
(37)
Magnetic fields no longer appear in equation (37). Instead,
we have second derivatives of electrostatic fields.
1 DRUDE: "Theory of Optics," translated by Mann and Millikan, p. 265.
SEC. 6] THE ELECTROMAGNETIC WA VE 95
(b) Let us now discuss the second bracket on the right-hand
side of equation (37). Since each term of this bracket is differ-
entiated with respect to x, we may write it as follows:
j_ (ax+ aY + az) (38 )
ax ax ay az
Consider a small, rectangular parallelepiped (Fig. 12), whose
sides are, respectively, dx, dy, dz, placed in any given location
with respect to a coordinate system x, y, z. Assume that there
are, in the neighborhood, electric
fields due to any distribution of !I
charges, and, to make the prob- I
lem concrete, assume that the )----
/
resultant field, near the parallele- I
piped, is directed in the general dy I
direction away from the origin. dx
Lines of force will then enter
the parallelepiped through the
left, the front, and the bottom
faces and will leave through the
right, the rear, and the top faces.
2
If there are no free charges within FIG. 12.
the parallelepiped, as many lines
must enter it as leave. Or, calling lines which enter "positive"
and lines which leave "negative," we may say that the sum of
lines entering plus those leaving is zero. We may express this
fact in the following way:
Let X and X' be the (average) intensities of the x components
of the electric field over the left- and the right-hand faces,
respectively. Then Xdydz lines enter the left-hand face and
X'dydz lines leave the right-hand face. Letting Y, Y' and Z, Z'
refer, similarly, to the other four faces, we may write:
(X - X')dydz + (Y - Y')dxdz + (Z - Z')dxdy = 0 (39)
By a method exactly analogous to that used in deriving equations
(15) and (16), we obtain relations between X and X', Y and Y',
Zand Z', as follows:
X' = X + ax
ax
dx
Y' = y + aY dy (40)
ay
Z' = Z + az
az
dz
96 ELECTROMAGNETIC THEORY [CHAP. IV
Putting these values of X', Y', Z' from equations (40) in equation
(39), we have, after canceling the identical terms having opposite
signs,
ax aY az)
( -ax + -ay + -az dxdydz = 0 (41)
But dx · dy · dz is the volume of the parallelopiped. Since this
volume is not zero, it follows that the parenthesis of equation ( 41)
must be zero. i.e.,
(42)
Now, the parallelopiped was arbitrarily located. Hence equation
(42) is true for ail locations, or for all values of the coordinates.
Therefore, the quantity (38) is equal to zero; i.e.,
~(ax+
ax ax
aY + az)
ay az
= 0 (43 )
In passing, attention should be called to the fact that, in
setting up equation (39), we assumed that there were no free
charges within the parallelepiped. If, however, we assume
that there are free (positive, say) charges within the parallele-
piped to the extent of p per unit volume, or p · dx · dy · dz within
the whole volume, the number of lines which leave will be in
excess of those which enter by 41rpdxdydz. We should, then,
write, instead of equation (39),
(X-X')dydz + (Y-Y')dxdz + (Z-Z')dxdy = -41rpdxdydz (44)
Equation (42) then becomes
ax + a Y + az = 41rp (45)
ax ay az
In vector notation, the left-hand side of equation (45), or of
equation (42), is spoken of as the divergence (div) of the electric
vector F. And equations (42) and (45) may be written, respec-
tively,
div F = 0 (42')
div F = 41rp (45')
(c) Equation (43) shows that, if we consider a region in which
there are no free charges, the second term on the right side of
equation (37) is zero. And we may write,
a2x a2x
+ -aayx + -aazX
2 2
µe
-c2 -at 2 = -ax 2 2 2 (46)
SEC. 6] THE ELECTROMAGNETIC WA VE 97
This equation (46) was derived by the series of steps given on
page 94. It is obvious th~t, if we start by differentiating the
equation (31b) with respect to time and equations (32a) and (32c)
with respect to z and x, respectively, we should obtain for the
component Y an equation similar to equation (46). Proceeding
in this way, one may obtain equations for each of the components
of the electric and the magnetic fields. These six equations are
as follows:
µe a2X a2X fJ2X a2X
- -2= -2+ -2+ -2
c2 at ax iJy az
(47a)
µe a Y2 2
aY aY 2 a2Y
c 2 at 2 = ax 2 + ay 2 + az 2 (47b)
µe a2z a2z a2z a2z
- -
c2 at =
2 -
ax 2+ -
ay 2 + -2
az (47c)
µe d 2a _ a2a a2a a2a
c2 at 2 - ax 2 + ay 2 + az 2 (48a)
µe a2{3 - a213 a2{3 a2{3
c 2 at 2 - ax 2 + ay 2 + az 2 (48b)
µe a 'Y
2
_ a 'Y
2
a 'Y
2
a 'Y
2
c2 at 2 - ax 2 + ay 2 + az 2 (48c)
(d) General Equations of Wave M otion.-1 t now remains to inter-
pret these equations (47) and (48). They can be shown to repre-
y ····(·J·;····-·V\···--·---~{y (:>-'J1
_:c __ ,,,,, , ___ _
-----x.i------- -------.J('2- - - - - -
Fm. 13.
sent the differential equations of a wave motion. Consider, first,
a simple case of a wave motion along a string ..
In Fig. 13, let a string be stretched initially so as to coincide
with the x-axis. Let the string be then given (say by a trans-
verse blow) a "disturbance," such as is shown at (1), and let this
disturbance travel, as a wave, toward the right, with a velocity v.
The magnitude of v will depend on the tension and the mass per
unit length of the string. The "wave form," that is to say, the
displacements in the y direction of the several parts of the string,
is, of course, some function, say the f function, of x. If we measure
x and y from 01, we may write
y = f(x1) (49)
98 ELECTROMAGNETIC THEORY [CHAP. IV
After a time, t, the disturbance will have reached position (2).
If we imagine the origin to have been carried along with the
wave, the new position of the origin will be at 02, a distance
v · t from 01. If we assume that the string is homogeneous and
that there is no damping, it can be shown that the form of the
wave at (2) will be exactly the same as the form at (1). That is,
measuring x from 02, the displacement of the string at (2) will
be exactly the same function of x 2 as it previously was of X1 at
position (1). And we may write, for position (2)
y = f(x2) (50)
If, however, we wish to continue to measure x from 01, we must
replace X2 in equation (50) by
X2 = X1 - Vt ( 51)
where X1 now refers to distance 01P 2 and to time t. Hence, we
may write equation (50):
y = f(x1 - vt) (52)
This equation represents the wave motion under consideration.
It is to be noted that y is here a displacement.
We now proceed to derive the differential equation of this
wave motion. Differentiate equation (52) with respect to t and
x, 1 respectively:
ay = -vf'(x - vt) (53a)
at
ay = f'(x - vt) (53b)
ax
.. ay ay
-v- (54)
· · at - ax
Differentiating again gives
2
a y2 = v2f" (x - vt) (55a)
at
2
a y2 = f" (x - vt) (55b)
ax
. a2y - 2 a2y
.. at2 - v ax2 (56)
(The functions f' and f'' are, of course, the first and second deri va-
tives of the function f.)
Equation (56) is the differential equation sought. It represents
a plane-polarized (transverse) wave traveling along the x-axis
and is seen to be independent of the function f of equation (52);
1 The subscript of x1 may now be dropped.
SEC 6] THE ELECTROMAGNETIC WA VE 99
i.e., equation (56) applies to any wave form. It is, also, obvious
that exactly the same line of reasoning might be applied to a
longitudinal wave, such as a sound wave, by replacing the trans-
verse displacement of the string, here designated by y, by either
the longitudinal displacement or by the pressure variation p
in the sound wave. Thus, we might write,
a2p a2p
at 2 -- v2 ax 2 (57)
for a sound wave, where p stands for the variation of pressure
from normal.
It is instructive to derive the equation for the wave motion
along a string which is not parallel to the x-axis.
!I
lJ c s
I\ I
I \I
d1J.--
- - - - - F(X;!/)
Frn. 14.
In Fig. 14, let the string OS be in the xy plane and make angles
X and µ with the axes. Let
cos X = l
cosµ = m
be the direction cosines. As in Fig. 13, let a disturbance travel
along the string with velocity v. If P (coordinates x, y) is the
position of any. point on the string before displacement, and P 1
the position of the corresponding point after a displacement
D = PP1, we may write, as before, calling OP. = R,
D = f(R)
where f(R) is the f function of R.
For the equation of the wave motion corresponding to equation,
(52) we may write,
D = f(R - vt) (58)
It is preferable, however, to consider separately, the horizontal
and the vertical components of D. Calling
Dx = Pd
Dv = Pc
100 ELECTROMAGNETIC THEORY [CHAP. IV
We have
Dx = fx(R - vt) (59a)
Dv = f y(R - vt) (59b)
where fx and fy are suitable functions. Now,
·: OP = R = lx + my
.". Dx = fx(lx + my - vt) (60a)
DY = fy(lx + my - vt) (60b)
Differentiate equation (60a) twice with respect to t, x, and y, in
turn. We then have
. a;ix vJ; (lx + my - vt)
= 2 1
(61a)
(61b)
(6lc)
Since Z2 +m 2 = 1, we have, by adding (b) and (c) of equations
(61),
a2Dx
ax2
+ a2D.:c
ay2
= f"(l
x x
+ my - t)
v
and, by substitution in equation (61a), this giVf;s
aatDx2 = v (aaxDx2 + aayDx)
2
2
2
2 2
(62 )
If the string be not confined to the xy plane but makes angles
X, µ, v with the three coordinate axes x, y, and z, we should have,
by an obvious extension of equation (62),
2
a D.c = '1; 2
2
(a Dx + a2Dx2 + a2D.:c) (63 a)
at 2 ax 2 ay 2 az
and, for the other components of the displacement,
a2Dy = v2 (a2Dy + a2ny + a2Du) (63b)
at 2 ax 2 ay 2 az 2
2 2 2 2
a Dz = v2 (a Dz + a Dz2 + a Dz) (63 c)
at 2 ax 2 ay az 2
(e) Equations (63) are identical inform with equations (47)
and (48), provided we put
c2
v2 = -
µ€
Since equations (63) were derived from a wave motion, it is
obvious that equations (47) and (48) represent wave motions, the
SEC. 7] THE ELECTROMAGNETIC WA VE 101
former an electric wave, the latter a magnetic wave, the common
velocity of which is
c
v =-- (64)
,Yµe
In other words, just as the mechanical disturbance in the string
(the displacement Din Fig. 14) was propagated along the string,
so equations (47) tell us that an electrostatic disturbance in a
medium is propagated through the medium with a velocity given
by equation (64). And, similarly, for a magnetic disturbance.
7. The Electromagnetic Wave.-The differential equations of
the electric wave (equations (47)) and those of the magnetic wave
(equations (48)) seem to be inde- a
pendent of each other. But it must :
I
be recalled that they were derived
by suitably combining the six equa-
tions (31) and (32) and that, there- Y
fore, they are not independent. The
significance of this dependence, in
so far as the wave motion is con-
cerned, may be shown as follows:
Consider a very large, plane, metal I
sheet, theoretically an infinite sheet, 1'
placed in coincidence with the yz 6
FIG. 15.
plane. The sheet is represented in
cross-section by the line ab in Fig. 15, the z-axis being perpen-
dicular to the plane of the paper. Imagine a uniformly distrib-
uted, constant current flowing through this sheet parallel to the
z-axis and flowing toward the reader. If the current be i~ per
centimeter of width of the sheet, it can be shown, by application
of equation (9) to the infinite "current sheet," ~hat there will be
produced in the region adjacent to the sheet a magnetic field
which will be in the y direction and which is given by 1
{3 = 21ri~
This field is represented at some point P1 (Fig. 15). Now,
allow the current through the sheet to vary at a rate ai~/at.
There will result a variation of {3 at a rate a[3 / at which rate, at
any given instant, will vary from point to point along the x-axis
because of the finite velocity of propagation of a magnetic dis-
turbance (equation (64)).
1 The derivation of this formula is left to the reader.
102 ELECTROMAGNETIC THEORY [CHAP. IV
Let us now apply the electromagnetic equations (31) and (32)
to this phenomenon. Since there are no x or z components of
the magnetic field,
a="f=O
and all their derivatives must be, also, zero; i.e.,
aa _ aa _ aa _ O
at - ay - az -
a"'_ a"'_ a"'_ 0
at - ax - ay -
Further, because of the infinite extent of the current sheet in
the yz plane, there can be no variation of any electric or magnetic
quantity with y or z, since the events which happen at P1 happen
simultaneously at all points in a plane through P1 parallel to
the current sheet. We have, therefore,
= 0
8{3
az
az _ aY _ ax _ ax _0
ay az az ay
Equations (31) and (32) thus become, respectively,
~ ax = 0 (65a)
c at
~ aY = 0 (65b)
c at
e az _ a13
c7it - ax (65c)
and
0=0 (66a)
!! a[3 = az (66b)
cat ax
0 = aY
-- (66c)
ax
Differentiating equation (66b) with respect to t and equation
(65c) with respect to x, and combining, we have
a213 c2 a2[3
0
(67)
at 2 = µ€ ax 2
And differentiating equation (65c) with respect tot and equation
(66b) with respect to x, we have
a2z c2 a2z (68)
SEC. 8] FLOW OF ENERGY 103
Equation (67) shows that we have a transverse magnetic
c
disturbance traveling in the x direction with a velocity _ 1- ; and
v µe
equation (68) shows that accompanying this magnetic disturbance
is a transverse electric disturbance, also traveling in the x direction
with the same velocity. Further, the vectors in these disturbances
y
I. z--
0/recf/on of Propctgcd/on
. x
of' 0.'ISflJrbance
z
Frn. 16.
or waves, i.e., {3 and Z, respectively, are perpendicular to each
other, as well as to the direction of propagation (Fig. 16).
These two mutually dependent waves constitute a plane-polarized
electromagnetic wave.
8. Flow of Energy in an Electromagnetic Wave.-(a) Consider
the direct current discussed in connection with Fig. 15 to be
a
I
~--------cA-------
I
I
_ _ _ Mcrgnef/c Vecfo~jl
6
- - - - --Rede of Flow of'Energ.!/; .E',
FIG. 17.
replaced by an alternating current. As before, let the infi-
nite current sheet ab (Fig. 17) be coincident with the yz plane,
and assume that the current in the sheet is represented by the
usual sine function of the form
i
. ' = I' sm
. Tt
21r
0 0
where T is the period of the alternating current, I~ is its
amplitude and i~ is its value at time t, per unit width of the sheet.
104 ELECTROMAGNETIC THEORY [CHAP. IV
The fluctuations of the magnetic field {3 in the neighborhood
will be periodic with those of the current, and we may write for
the value of {3 at some point 0, arbitrarily located on the x-axis,
R
JJ = • 21r
f3o sm Tt
(69 )
where f3o is the amplitude of {3. 1 On account of the finite velocity
of propagation of an electromagnetic disturbance, namely:
c
v =-- (70)
y'µf.
whatever happens at O will happen at some point P a little
later. If we take O as the origin of coordinates, then OP = x,
and the time required for a disturbance to pass from O to
P will be x/v. By subtracting this time interval x/v from t of
equation (69), we can express the variation of {3 at any point
P on the x-axis by
(3 = f3o sin ~(t -~) (71)
This is the equation of a plane-polarized magnetic wave train
moving in the positive direction of x. If we fix our attention
on a particular point, i.e., if we keep x constant, then this
equation shows that {3 is a simple harmonic function of t. If,
however, we take a ''snapshot'' of the conditions along the
x-axis at any instant, i.e., if we keep t constant and vary x,
the value of {3 is seen to be a simple harmonic function of x, as
shown in the full line of Fig. 17. Starting from x = 0, the
values of {3 are seen to be identical at the points x = 0, x = vT,
x = 2vT, x = 3vT, etc., as may be shown by substituting these
successive values of x in equation (71). The distance between
these points is called a wave length, X, i.e., vT = X, and an alter-
native form of equation (71) is
{3 = f3o sin 271" ( ~- ~) (72)
(b) We may derive the equation of the accompanying electric
wave as follows: We have seen (equations (67) and (68)) that
we have to deal only with the Z component of the electric force.
Differentiating equation (71) with respect to x gives
~/!.
ax
= - l 27r [3 0 cos 27r
v T T v
(t - ~)
(73)
1
Strictly speaking, there is a phase difference between equation (69) and
the preceding equation.
SEC. 8] FLOW OF ENERGY 105
Putting this value of a[3/ax in equation (65c) gives
~ az = _ _! 21r
c at v T
f3o cos 21r
T
(t - ~)
v
(74)
Integrating this equation and remembering that the constant
of integration must be zero, since it is assumed that there are
no permanent fields, we have 1
Z = - - c f3o sm. -21rT ( t - -x)v
€V
(75)
= . T
Z o sm 21r ( t - vx) (76)
by putting
c
Zo = - - f3o
€ l)
A comparison of equations (76) and (71) shows that the electric
wave is everywhere in phase with the magnetic wave.
(c) The magnitudes of the vectors {3 and Z are very closely
related. Inserting the value of (3 from equation (71) into
equation (75) gives
c
z= --{3
EV
(77)
which shows that the two vectors are proportional to each other.
A symmetrical relation between Z and {3 may be obtained by
putting into equation (77) the value of c from equation (70).
We then obtain.
(78)
If we are dealing with empty space, for which e and µ are unity,
Z is numerically equal to {3.
(d) Now, the setting up of either an electric or a magnetic
field involves energy, and it may be shown 'that the energy
per unit volume in an electric field of strength F is
f.p2
81r
and in a magnetic field of strength H is
µH2
81r
1
The reader will find it instructive to consider the origin and the signif-
icance of the minus sign which appears before the right hand member of
equation (75) and certain of the following equations.
106 ELECTROMAGNETIC THEORY [CHAP. IV
Squaring both sides of equation (78) and dividing by 81r gives
ez2 µ{32 (79)
81r = 81r
These two terms give, respectively, the electric and the magnetic
energy per unit volume at any point in the wave where the
fields are Z and (3. It is, therefore, seen that the energy carrieQ
by the electric wave is equal to that carried by the magnetic
wave, and the total energy w per unit volume at the point may
be written, variously, as
w
eZ 2 µ{3 2
= 41r = 41r = -
Vµe
~Z{3 (80)
This leads to an important theorem regarding the flow of
energy m an electromagnetic wave. Consider an elementary
volume, located, say, at P (Fig. 17)
!I. with faces of unit area and of thick-
ness dx, placed parallel to the yz
plane. Such an element of volume
J- x is shown in Fig. 18. By equation
I
(80), the electromagnetic energy dW
z in this element of volume is, since
FIG. 18.
the volume is dx,
dW = - y'µeZ(3dx (81)
41r
Since the wave moves to the right with velocity v, in a time
dt given by
dt = dx
v
all this energy dW will have passed through the right-hand face.
The rate of flow of energy through this (unit) area is, therefore,
V µe Z(3dx
dW 41r
dt - dx/v
__ vy'µe Z(3
41r
c
- - 41r Z[3 (82)
by equation (70). This is Poynting's theorem, and it states that
in an electromagnetic wave (of the type herein discussed) the
rate of flow of energy per unit area is proportional to the product
of the electric and the magnetic vectors.
SEC. 81 FLOW OF ENERGY 107
If, in equation (82), we put the value of Z given by equation
(77), we get
(83)
= --{3 O
• 2 -21r(t -
c2 2 sm
41r€V T
x)
-
v
(84)
Equation (83) shows that, irrespective of the sign of {}, the flow
of energy is always in the same direction, since {3 2 is always
positive. And equation (84) shows the variation of this energy
flow with both t and x. Thus, considering the flow of energy
through the slab in Fig. 18, i.e., keeping x constant in equation
(84), the rate of flow varies between the minimum rate zero and
• c2{3~
the maximum rate - - · Or, considering the rate of flow, at
47r€V
any given instant at various points along the x-axis, we have
the rate of flow given by the dotted curve in Fig. 17.
(e) Equation (84) gives the instantaneous rate of flow of
energy through a square centimeter at right angles to the direction
of propagation of the wave.
A quantity of greater impor- dW
tance is the average rate of flow dt
of energy. This corresponds
to what is ordinarily called
the intensity I of the wave,
i.e., the energy transferred 7i'me b
a
per second per square cen- Fm. 19.
timeter of wave front. The
average rate of flow of energy may readily be determined as
follows: If one integrates equation (84) over a half-period, one
has the total quantity of energy flowing through the square
centimeter in time T /2. This quantity of energy divided by
T /2 gives the desired average rate of flow. The graph of equa-
tion (84) is shown in Fig. 19 from a time a = 0 to a time b = T /2.
Designating the average rate of flow per square centimeter by
I, we may indicate the desired integration by
i rT/2 aw
I = T /2 Jo dt dt
=
1 c 2
T/2 41rev J3~Jo
rr /2 .
sm2
21r ( .
T t -
x)
V dt (85)
108 ELECTROMAGNETIC THEORY [CHAP. IV
In this integration, x is kept constant. The value of the integral
of equation (85) can be readily shown to be 1
T
4
Putting this value of the integral in equation (85) gives
I = __!_ c2 {32 (86)
81r €V o
Remembering that /3 0 is the amplitude of the magnetic vector,
we have, from this equation, the very important relation that
the intensity of an electromagnetic wave is proportional to the square
of the amplitude of either the magnetic or the electric vector, since,
by equation (77), the latter are proportional to each other.
Applying equation (86) to empty space, for which '= = 1, and
v = c, we have
I = !!__ {32
81r O
From definition, I is the energy contai:rred in a train of electro-
magnetic waves comprising a column 1 sq. cm. in cross-section
and c cm. long. The (average) energy 2 per cubic centimeter (or,
more generally, per unit volume) in this wave train is I /c = (3V81r.
This quantity is called the energy density of the radiation. We
shall discuss it later in Chap. VII.
9. The Electromagnetic Theory of Light.-Equation (64)
c
v = y'µ€
makes it possible to compute the velocity v of an electromagnetic
wave in any medium from the constants µ and f of the medium
and the ratio c of the electrostatic to the elect~magnetic system
of electrical units. As stated on page 82, measurements show
that c = 3 X 10 10 cm. per second, very nearly. Considering empty
space, for which bothµ and '= are, by definition, unity, we have
v = c = 3 X 10 10 cm. sec.- 1 (87)
1 This is, at once, seen, since the average value of the integralf sin 2 OdO
from (} = 0 to (} = 1r is H. Or the equation may be integrated directly by
. 21r(
settmg t/; = T t -
x)
v =
21r
vx
21r
T t - </>, where </> = T = const. Then dt =
:: d,J;; and the limits of integration, 0 and~' become, respectively, -<f, and
(1r - </>).
2 It is to be remembered that half of this energy is due to the electric vector
and half to the magnetic vector.
SEC. 9] ELECTROMAGNETIC THEORY OF LIGHT 109
That is, the velocity of an electromagnetic wave in empty space is
equal to the ratio c between the electromagnetic and the electrostatic
systems of units, and has a numerical value of very nearly 3 X 10 10
cm. sec- 1•
Now, numerous measurements have been made of the velocity
of light in empty space:
VELOCITY
CM. sEc.- 1
Romer (1676). . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. 92 X 1010
Fizeau (1849). . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 .13 X 10 10
Foucault (1862) . . . . . . . . . . . . . . . . . . . . . . . . . 2. 986 X 10 10
Cornu (1874) ............................ 3 .004 X 10 10
Michelson (1879). . . . . . . . . . . . . . . . . . . . . . . . . 2. 999 X 10 10
Michelson (1882) ......................... 2. 998 X 10 10
Michelson (1926) 1 . . . . . . . . . . . . . . . . . . . . . . . . 2. 99796 X 10 10
1 Astrophys. Jour., vol. 65, p. 1 (January, 1927).
The velocity of light is, thus, seen to be, within experimental
error, equal to the velocity of electromagnetic waves. Further,
as the phenomenon of polarization shows, light, in common /
with electromagnetic waves, is a transverse phenomenon. These
two coincidences alone gave very strong support to the theory,
first announced by Maxwell, in 1865, that light is an electro-
magnetic phenomenon; and that light consists of electromagnetic
(instead of mechanical) vibrations, or waves, in the "ether"-
a concept which at once removed many of the difficulties of the
mechanical theory of the ether and of light vibrations, while
retaining practically all the advantages of the older wave theory.
If this identity be correct, all of the equations and principles
discussed in this chapter should apply to light. For example,
a beam of plane-polarized light should be composed of electric
and magnetic vibrations at right angles to each other. Further,
since electric and magnetic fields are continuous, the flow of
energy in a light wave should be continuous over the entire wave
front, and the intensity of the light wave should be proportional
to the square of the amplitude (of either the electric or the mag-
netic vector, equation (86)), in agreement with the older theory.
The electromagnetic theory of light predicts, at once, the
electrical nature of matter. For, as we have seen, electro-
magnetic waves originate as a result of the vibrating motion or,
more generally, the periodic motion, of electric charges (e.g., the
alternating current in Fig. 17). If light waves be electromag-
netic, they must have an electromagnetic origin. Now, the
110 ELECTROMAGNETIC THEORY [CHAP. IV
facts of spectroscopy show that each element has a characteristic
line spectrum, which, it must be assumed, originates in the atoms
of the element. The conclusion is very direct, therefore, that
the atoms must contain electric charges the vibrations of which
send out light waves. We shall see, later, that this conclusion
regarding the electrical nature of matter was supported by the
application of the electromagnetic theory to other phenomena.
In the discussion in Sec. 8, we considered, for the sake of
mathematical simplicity, the effect which would be produced by
an alternating current sheet of infinite extent. The general
nature of the phenomena would not, however, be changed if,
instead of the infinite sheet, we had considered a wire, coincident
with the z-axis, through which the alternating current was
flowing. We should, then, have had a cylindrical wave spreading
out from the wire as an axis. (This is the modern wireless
antenna.) Indeed, a single electron, vibrating back and forth
along, say, the z-axis, would send out a train of electromagnetic
waves, i.e., would radiate energy. We are, thus, led to a very
profound and far-reaching generalization of the electromagnetic
theory: namely, that any accelerated charge must radiate energy
in the form of an electromagnetic pulse or wave train. If the charge,
initially at rest, is given a positive acceleration; or if, in motion, it
is given a negative acceleration (i.e., is brought to rest), a single
electromagnetic pulse is sent out-analogous to the sound pulse
or "crack" sent out when a steel ball collides with a massive
steel plate. If the acceleration be such that the charge vibrates
back and forth over a linear path, or moves at uniform speed
around a circular path, a train of waves is given out-analogous
to the sound waves sent out by a vibrating tuning fork. Accord-
ing to the electromagnetic theory, the acceleration of a charge cannot
take place without this radiation of energy. We shall frequently
return to this point later.
10. The Discovery of Electromagnetic Waves.-The electro-
magnetic waves predicted by Maxwell were not demonstrated
experimentally during his lifetime. The concept rested on purely
theoretical foundations and on indirect evidence, until the
work of Hertz in 1887-1888. A description of Hertz's experi-
ments is beyond the scope of this book. The reader is referred
to any one of the numerous accounts. Suffice it to say here, that
Hertz produced electrical oscillations by taking advantage of
the vibratory nature of the electric discharge discovered by
SEC. 11] REFRACTION OF LIGHT 111
Henry. The frequency of the oscillations. in Hertz's apparatus
was of the order of 10 8 vibrations per second, giving rise to waves
a few meters long. He detected the existence of these waves by
means of a resonator consisting of a piece of wire, with polished
brass balls on its ends, bent into the form of a circular loop so as
to leave a small air gap between the knobs. The size of the
loop was so adjusted that the natural period of the electrical
oscillations of the loop would just coincide with the period of
the "sending" apparatus.
Hertz proved conclusively that the oscillating discharge would
produce electromagnetic waves, exactly as predicted by Maxwell,
and that these waves had many, if not all, the properties of light
waves: They could be reflected and refracted; they could be
made to interfere and to produce stationary waves; and they
were polarized.
Later, after the technique of producing and working with
these waves had been developed, mainly by Hertz, their velocity
was measured by various observers, with the following results:
Velocity,
Observers Date centimeter-
seconds-1
Trowbridge and Duane. . . 1895 3.003 x 10 10
Saunders. . . . . . . . . . . . . . . . 1897 2.997 x 1010
lVIacLean..... .. . . . . . . . . 1899 2. 991 x 10 10
This velocity is seen to be identical with the velocity of light and,
as predicted by Maxwell, is exactly equal to the ratio c of the
two systems of electrical units. Maxwell's electromagnetic
theory had become an experimental reality.
11. The Refraction of Light.-Indirect evidence rapidly
accumulated to confirm the electromagnetic theory, and it
proved very fruitful in predicting new phenomena. A few of
these will be mentioned briefly.
Equation (64), v = c/VµE, gives the velocity of an electro-
magnetic wave and, therefore, also, if the electromagnetic theory
of light be correct, the velocity of a light wave, in any medium
for which the permeability µ and the dielectric constant E are
known. Now, the index of refraction n of a medium is given by
Vo
n=- (88)
v
112 ELECTROMAGNETIC THEORY [CHAP. IV
where Vo is the velocity of light in vacuo and v is the velocity in
the medium. For a vacuum µ = E = 1, and, therefore, accord~
ing to equation (64),
Vo= C
For practically all transparent substances, µ = 1, very nearly
Therefore, for such substances,
c
v = ,v;
Substituting these values of and v m equation (88) gives
Vo
n = Ve (89)
That is, the index of refraction of a medium (for which the
permeability is unity) should be equal to the square root of its
dielectric constant. This prediction of the electromagnetic
theory may be tested by direct experiment. The following tables
give comparative data for certain illustrative substances, the
index of refraction being for yellow light, and the dielectric
constant being for room temperature, except as otherwise indi-
cated, and for either direct currents or for alternating currents
of low frequency.
TABLE !.-GASES
Substance n
H2 .................. . 1.000138 1. 000132
Air .................. . 1.000294 1.000295
CO .................. . 1.000346 1.000345
CH 4 • • • . • • • • • • • • • • • • • • 1.00044 1. 00047
C02 ................. . 1.000449 1.000473
N20 ................. . 1.000503 1.000497
CS2 ................. . 1.00150 1. 00145
NHa ................. . 1.00038 1 1. 0039 1
802 ................. . 1.00066 1 1. 0048 1
INote the number of zeros.
TABLE IL-LIQUIDS
Substance n
I
Benzene ............. . 1.50 1.51
Turpen tine ........... . 1.49 1.46-1.51
Toluene .............. . 1.49 1.48-1.55
Petroleum ............ . 1.39 1.39
SEC. 11] REFRACTION OF LIGHT 113
TABLE III.-LIQUEFIED GASES
Substance n
N20 (15°C) .......... . 1.19 1.23
C02 (15°C.) .......... . 1.19 1.23
02 ( -180°C.) ........ . 1.22 1.33
NH3 (14°C.) .......... . 1.32 4.01
TABLE IV.-LIQUIDS
Substance n
Carbon disulphide ..... . 1.64 1.62
Ethyl ether .......... . 1.35 2.17
Phenol ............... . 1.54 3.10
Amyl alcohol. ........ . 1. 95 4.00
Acetone .............. . 1.36 4.55
Methyl alcohol. ...... . 1.33 5.6
Glycerine ............ . 1.47 7.5
Water ............... . 1.33 8.9
TABLE v.-souns
Substance n
Paraffin .............. . 1.42 1.52
Quartz ............... . 1.55 2.12
Rock salt ............ . 1.54 2.42
Glass (crown) . . . . . . . . . 1.53 2.64
Selenium ............. . 2.93 2.73
An examination of these data shows (1) that the agreement
between n and Ve is exceptionally good for most gases (Table
I); (2) that the agreement is fair for certain liquid hydrocarbons
(Table II); (3) but that, for most liquids and solids (Tables
III, IV, V), the law does not hold even approximately, water
being the most notable exception. In general, the index of
refraction is less than the square root of the dielectric constant.
The fact that, in most gases, the predicted equality of n and
VE is observed indicates that the theory leading to this law is,
at least in certain particulars, substantially correct; while
the disagreement found in the case of most liquids and solids
indicates that the theory is incomplete-that certain factors
have not been taken into account.·
114 ELECTROMAGNETIC THEORY [CHAP. IV
We do not have to seek far to find some of these factors. The
data given above for the index of refraction n are for yellow light;
while the values of the dielectric constant e are for room tempera-
tures and for frequencies which are either zero or are very low
compared to the frequencies of light vibrations. Now, it is well
known that the index of refraction depends on the wave length
of the light and that the dielectric constant depends on both the
temperature and the frequency of the alternating current used in
the measurement. Thus, the index of refraction of quartz (ordi-
nary ray) from the extreme ultra-violet to the far infra-red is
shown in Table VI.
TABLE VI.-INDEX OF REFRACTION OF QUARTZ (ORDINARY RAY)
Wave length, Index of
centimeters refraction
0.185 x 10-4 1.676
0.274 1.587
0.396 1.558
0.589 1.544
0.760 1.539
2.33 1.516
3.18 1.494
5.0 1.42
7.0 1.17
The variation of the dielectric constant of glycerin with frequency
is given in Table VII.
TABLE VIL-THE VARIATION OF THE DIELECTRIC CoNsTANT 01r GLYCERIN
WITH WAVE LENGTH
Equivalent
Frequency, wave length Dielectric
sec.- 1 '.1n vacuum, constant
centimeters
0.25 x 10 8 1,200 56.2
1.50' 200 39.1
4.00 7.5 25.4
35.0 8.5 4.4
750 0.4 2.6
5 x 1Q14 (Yell ow light) n 1 = 1.47
1n = index of refraction.
SEC. 11] REFRACTION OF LIGHT 115
The effect of temperature on dielectric constant is illustrated
by the following data (Table VIII) for methyl alcohol.
TABLE VIII. 1-THE VARIATION OF THE DIELECTRIC CONSTANT OF METHYL
ALCOHOL WITH TEMPERATURE
Temperature,
Dielectric
degrees
constant
centigrade
-100 58
50 45
0 35
20 31
1 Data for Tables I to VIII were compiled from (1) LANDOLDT and BERNSTEIN:" Physical-
ische Tabellen" and (2) "The Smithsonian Physical Tables."
Even a casual inspection of Tables VI, VII, and VIII makes it
obvious that one can compare values of the index of refraction
and the square root of the dielectric constant only when the
measurements are made under identical conditions. One can
easil~elimina.te the effect of temperature by making the measure-
ments of the two quantities at identical temperatures. And,
likewise, the value of n is comparable with y; only when the
index of refraction is measured for a beam of radiation the
frequency of which is equal to the frequency of the alternating
current for which y; is observed. An approach to such equality
of conditions is to be found in the measurements on the index of
refraction of water for electromagnetic waves the frequency
of which is very low compared to that for ordinary light. Such
measurements are given in Table IX.
TABLE IX.-INDEX OF REFRACTION OF WATER FOR ELECTROMAGNETIC
WAVES 1
Wave length, Index of
centimeters refraction
65 8.88
8.8 8.89
8.1 8.10
1. 75 7.82
1 Quoted by JEANS: "Electricity and Magnetism," p. 533.
116 ELECTROMAGNETIC THEORY [CHAP. IV
The value of y; for water is 8.94, which checks exceptionally
well with the value of n for long waves. Again, the predictions
of the electromagnetic theory are confirmed. But the question
still remains: How is dispersion, i.e., the variation of index of
refraction with wave length, to be explained on the basis of the
electromagnetic theory?
12. The Dispersion of Light.-The answer to the question
raised at the end of the preceding section is to be found by con-
sidering the electrical nature of matter. It has been pointed out
above that, if light be regarded as electromagnetic, we must look
for the origin of light waves in the vibrations of electrical charges.
Since matter under various conditions of excitation-high tem-
perature, electrical discharge, etc.-sends out light waves, the
implication is obvious that matter contains electrical charges or
electrically charged bodies capable, perhaps, of vibrating with
definite periods about certain fixed positions. The natural
frequencies of these vibrations in unexcited matter, or, at least,
some of them, might be expected to be of the order of magnitude
of frequencies encountered in ordinary light. It might be
expected, therefore, that the passage of light through matter
would be materially influenced by these electrically charged
particles, particularly if the frequency of the light happens to
be near the natural frequency of vibration of the particles about
their several positions of equilibrium. If, however, the frequency
of the light is far removed from, say is very much less than, the
natural frequency of the particles, we should expect the presence
of the electrical charges in the matter to exert little or no effect
on the passage of the light. This is the explanation of the fact,
shown in Table IX, that the index of refraction for electromag-
netic waves several centimeters long agrees so closely with the
square root of the dielectric constant.
It would lead us too far afield to give even an elementary dis-
cussion of the phenomena of dispersion on the basis of the
electromagnetic theory. 1 It may, however, be instructive to
consider the way in which such a discussion may be started, so
as to point out the modifications which have to be made in the
fundamental electromagnetic equations.
In deriving equations (31), we considered the current density
j to be made up solely of a "displacement" current 4: ! due
1 The reader is ref erred to an excellent discussion in HousTouN: "Treatise
on Light," Chap. XXIV.
SEC. 12] DISPERSION OF LIGHT 117
to the change in the electric field F. If, however, the medium in
which the change of field takes place contains electric charges,
there will be an actual displacement of electricity which must
. d"1sp1acement e1r di
b e a dd e d to t h e quasi- dF to get t h e tota1
4
effective current density. Thus, consider a charge e occupying,
in the absence of electric fields, an equilibrium position P 1 (Fig.
20). Imagine an electric field X, directed e
toward the right, to displace the charge to a R-------i1
1
position P2 such that the force Xe is just bal- FIG. 20.
anced by a "return" force, say f · x, where
x = P1P2 is the displacement and f is, for small displacements,
a constant. That is,
Xe= -fx
If X changes at a rate of dX/dt, then the position of the charge
will change at some rate dx/dt, the magnitude of which will be
determined by several factors, such as the mass m of the body
containing the charge e, the value off, etc. Now, let the medium
contain N such /,mobile" charges per unit volume. If, at any
instant, these charges all have a velocity dx/dt, the rate of
transfer of charge through unit area at right angles to dx/dt is
dx
Ne--
dt
The effective current density in. this direction will, then, be given
by
• € ax dx
J x = 41r at + Ne dt (90)
instead of by the simpler equation (27). Introducing this
new value of jx into equation (21a), we have, instead of equation
(31a), the relation
e ax+ 41r Ne dx _ a,y _ a13 (9 l)
c at c dt ay az
This equation can be applied to the case of a light wave
passing through the medium. X then becomes the periodically
varying electric vector in the wave. Let v be the frequency of
the light vibrations, and let vo be the natural frequency of vibra-
tion of the charged particles of the medium about their positions
of equilibrium. It can, then, be shown 1 that the index of refrac-
1 See HousTOUN: "Treatise on Light," Zoe. cit.
118 ELECTROMAGNETIC THEORY [CHAP. IV
tion n of a medium which is transparent to the radiation of
frequency v is given by
n2 = 1 + Ne2 2 (92)
1rm(v~ - v )
N being, as before, the number of charged particles per unit
volume, e the charge, and m the mass of each particle. This
formula shows that the index of refraction should depend on
the frequency of the light v and on the natural frequency of
vibration v0 of the charged particles. These charged particles
are nowadays identified as electrons. There may be different
groups of electrons in the medium. That is, we may have N1
electrons per qnit volume which have a natural frequency
v1; N 2 electrons per unit volume with natural frequency v2; etc.
We should then have
n2 = 1 + N1e2 + N2e2 + . . . (93)
1rm(vi - v2 ) 1rm(v~ - 2
11 )
Or, to put the expression in more compact form, factoring
e2 /1rm, we may write
e2 N
n2 = 1 + - ~ - - (94)
2
1rm 4' ( v~ - 11 )
This formula is similar to the well-known Sellmeier formula for
dispersion, deduced by Sellmeier, about 1870, on the basis of
the elastic solid theory of light.
TABLE X.-COMPARISON OF THE OBSERVED VALUES OF THE INDEX OF
REFRACTION OF WATER THROUGH THE VISIBLE SPECTRUM WITH THOSE
COMPUTED FROM EQUATION (92)
Wave length}.. Index of refraction n
(in vacuo), Frequency v,
sec.- 1
centimeter Observed 1 Computed
0. 7600 x 10-4 0 .3943 x 10 15 1.3293
0.6870 0.4365 1.3309 1.3305
0.6563 0.4567 1. 3317 1.3310
0.5893 0.5085 1.3335 1.3328
0.5270 0.5687 1.3358 1. 3351
0.4861 0.6166 1.3377 1.3371
0.4308 0.6957 1.3412 1.3409
0.3968 0.7553 1.3441
1 Quoted by HousTouN, loc. cit.
This dispersion formula represents, fairly well, the variation
with frequenGV of the index of refraction of transparent bodies in
SEC. 12] DISPERSION OF LIGHT 119
the visible part of the spectrum. For example, Table X shows,
in the third column, the observed values of the index of refraction
of water through the visible spectrum. Equation (92) may be
put in the form
B
n2 = 1 + (95)
v5 - v2
where B = Ne 2 /1rm. Forming a pair of simultaneous equations
by substituting in equation (95) the corresponding values of v
1.34G -----.----,-----.----__,.
l.344~---+----+---+-------4
1.MQt---r---t-----t---if----1
J o:Ohserved Index ofRefraclion
~ - =Index ofRefracf/on computed
c..- I.'3381-----\---f.---LL..JLLLL..=.q,'---'-"'<9.-=-'5).__-+------1
o O
s::
0
!filJ l.336---~-0---------
4=
Q)
t::(.
'bx J . 3 3 4 - - - - - - - - - - - - - -
Q)
0
-t5
~
1 . 3 3 2 - - - - - - -0- - - ~ - -
0
5 6 1 8
Wave length (cm.x 10 5)
FIG. 21.-Comparison of observed values of index of refraction of water with the
predictions of equation (95).
and n from the first and last lines, respectively, of Table X, one
can determine the constants B and vo. One obtains, in this
way, -------
B = 6.48 X 1030
Vo = 2.93 X 10 15
By putting these values of B and vo back in the equation, the
values of n for the other wave lengths are determined, as shown
120 ELECTROMAGNETIC THEORY [CHAP. IV
in the fourth column of Table X. In Fig. 21, the full line gives
the computed values of n, while the points give observed values.
The agreement is seen to be fair.
In passing, a word may be said regarding the physical sig-
nificance of the two constants vo and B. As previously explained,
vo represents the natural frequency of the charged particles.
The frequency v0 = 2.93 X 10 15 corresponds to the wave length
(in vacuo) of 0.102 X 10- 4 cm., which is in the extreme ultra-
violet. If the theory be approximately correct, water must have
particles, i.e., electrons, with a natural frequency corresponding to
a wave length somewhere in the extreme ultra-violet. If radia-
tion of this wave length were to pass through water, the particles
should resonate strongly, i.e., the water should absorb this
wave length. Or, in other words, the vo of equations (92) and
(94) corresponds to absorption bands. Water is known to have
several absorption bands in the infra-red, and it absorbs in the
ultra-violet. The complete dispersion formula should take
account of these several frequencies, as is indicated by equation
(94). A further modification must be introduced in the disper-
sion formula to take account of the effect of absorption in the
medium. For these extensions, the reader is referred to any
standard treatise on the electromagnetic theory of light.
Since B = Ne 2 /1rm, we are enabled to compute N, the number
of "mobile" electrons per cubic centimeter of water, using data
which we shall discuss later. Putting B = 6.48 X 1030 ; f; (the
electronic charge) = 4.774 X 10- 10 e.s.u.; and m (the mass of the
electron) = 0.9 X 10-27 gm., we have
N = 8 X 1022 electrons per cubic centimeter
It can be shown 1 that there are approximately 3.4 X 1022
molecules of water per cubic centimeter. In other words,
approximately 2 electrons per molecule of water (i.e., 8 + 3.4 =
2.3) have a natural period corresponding to a wave length in the
1 The numerical value of Avogadro's constant, i.e., the number of mole-
cules in a gram-molecule, is 6.06 X 10 23 • Since 1 gram-molecule of water
. 6.06
contams 18 grams, there are l8 X 10 23 = 3.37 X 10 22 molecules per gram
of water, or per cubic centimeter, since the density of water is unity.
As water undergoes thermal expansion, of course, the number N of electrons
per cubic centimeter should change. Hence, the index of refraction should
change with density, a conclusion in agreement with experiment.
SEC. 13J SUMMARY 121
ultra-violet, a conclusion which, if it be not rigorous, is, at
least, reasonable.
13. Summary.-In the foregoing, we have attempted to
sketch briefly some of the fundamentals and a few of the applica-
tions of the electromagnetic theory. We have seen that the
electromagnetic waves, predicted by the theory, were observed
experimentally by Hertz and have now passed far beyond the
academic stage of theoretical and experimental science by virtue
of their world-wide application in radio. When applied to
light, we have seen that the electromagnetic theory satisfactorily
accounts for polarization, refraction, and dispersion. Among
the other phenomena of optics to which the electromagnetic
theory has been, at least in part, successfully applied may be
mentioned reflection, total reflection, the Zeeman effect, the
Faraday rotation of the plane of polarization, and the optical
properties of metals. If one is to estimate the value of a theory
by its success in explaining and grouping together previously
observed phenomena and in predicting new ones, there can be
no doubt that the Electromagnetic Theory of Light ranks as
the outstanding development of the nineteenth century.
It may be well to mention again two of the fundamental tenets
of the electromagnetic theory, to which we shall have occasion
to refer later. First, it is an inescapable conclusion from the
theory, a conclusion, in fact, without which the whole theory
falls, that an accelerated charge must send out an electromagnetic
disturbance, i.e., must radiate energy. It will be shown later, that
the rate dW /dt of emission of energy by an accelerated charge is
given by the equation
dW _ 2 Q2g 2
aJ:-37 (96)
where Q is the charge, g its acceleration (positive or negative, g2
being always positive), and c the velocity of light. Second, this
rate of emission of energy, as well as the wave train which carries
it, must be continuous. That is, if we imagine the wave front of
our electromagnetic wave magnified indefinitely, it should still be
continuous, should show no structure. For the wave is composed
of magnetic and electric vectors which, so far as we know, have no
structure and which may assume any values whatever from zero
up. The concept of lines of electric or of magnetic force is, it
must be remembered, simply a geometrical "picture" to help us
keep in mind the various phenomena of electric and magnetic
122 ELECTROMAGNETIC THEORY [CHAP. IV
fields. Lines of force have no physical significance as lines. The
modern Quantum Theory, to the development of which we shall
presently turn, comes into direct conflict with these two indis-
pensable tenets of the electromagnetic theory.
Some References to the Electromagnetic Theory of Light
HousTOUN, R. A.: "A Treatise on Light."
Woon, R. W.: "Physical Optics."
STARLING, S.G.: ''Electricity and Magnetism."
DRUDE, P.: "Theory of Optics," translated by Mann and Millikan.
CHAPTER V
SOME THEOREMS CONCERNING MOVING 1 CHARGES
1. The Magnetic Field Produced by a Moving Charge.-In
his paper entitled "On the Magnetic Effect of Electric Convec-
tion," Prof. H. A. Rowland 2 proved conclusively by experiment
that a moving charge produces a magnetic field. The idea had
been suggeste~d by both Faraday and Maxwell, but Rowland
seems to have been the first to obtain experimental evidence of
-th~ effect. We can use the results of Sec, 6 of the previous
chapter to determine the magnitude of the fiela produced "by
the motion of a charge.
Equation (27) Chap. IV, states that an electric field which is
changing at a rate dF /dt is equivalent to a current of density j
given by
. 1 dF
J =--
41r dt
(We shall restrict the discussion to cases where e = 1.) If the
rate of change of field be uniform over an area S, taken, say per-
pendicular to F which, for simplicity, is assumed uniform over S,
the equivalent current i through that area is given by
i = Sj = __!_ SdF (1)
41r dt
But S · dF. is the change of electric flux through the area, since,
in the terminology of lines of electric flux, dF is the change in
lines per unit area. Representing the total electric flux through
the area S by <P, we may write, 3
1 We shall limit the discussion in this chapter to cases where the velocity
of the charge is small compared to the velocity of light.
2 Amer. Jour. Sci., vol. 15, p. 30 (1878). See also Papers No. 12 and No.
43 in the "Physical Papers of Henry A. Rowland."
3 More generally,
i = JjdS
where dS is an element of the surface S. Therefore,
i = l_ JdF dS = _!_ i JdF · dS = __!_ d<I>
41r ~dt 41r dt 41r dt
since the total change of flux through S is the change of flux dF through unit
area integrated over the whole area.
123
124 CONCERNING MOVING CHARGES (CHAP. v
def!= SdF
. 1 def!
... i = 41r dt (2)
and by computing the rate of change dcf!/dt of the total flux
through an area, we may determine, by equation (2), the equiva-
lent current through the area.
Let a positive charge Q (Fig. 22) be moving with velocity v in
the positive direction of X along the x-axis, the charge coinciding
with the origin O at the instant shown. We wish to find the
magnetic field HP at the point P which is at a distance R from
0. The line OP( = R) makes an angle </> with the direction of
motion of the charge. Describe about O a sphere of radius R
and let PMP' represent a segment of the sphere the plane PNP'
being perpendicular to the direction of motion and distant ON
( = x) 1 from Q. If the velocity v be small compared to the
velocity of light, there will be, symmetrically surrounding Q, 41rQ
lines of force (more correctly, lines of induction).
p
./
I
/ I
./ l
I
I
Frn. 22.
We shall compute the total number of these lines cf!8 which go
through the surf ace P MP' and then the rate of change of cf!8
due to the motion of the charge. The field intensity F 8 every-
where over the area P MP' is
Q
Fs = R2
and the total flux cf!8 through P MP' is
cf!s = Fs X (area PMP')
= 12 X 21rR 2 (l - cos cf,)
= 21rQ (1 - cos </>) (3)
1 Note that xis taken as the distance between Q and N; i.e., x decreases as
Q moves toward the right.
SEC. 1] MAGNETIC FIELD 125
Now,
x
cos <p = - - -
r2 Vx2 +
where r ( =PN) is the radius of the base of the segment and
x = ON. Therefore,
'Ps = 21rQ[ 1 - (x2 +x r2)Y. J (4)
Differentiating both sides of equation (4) with respect to t and
dx
remembering that - dt = v, we have
d:s = 2-n-Q[ (x2 + r2)-Y. - x2(x2 + r2)-H Jv
= 21v (1 - ;:)
by replacing (x2 + r )Y
2 2
by R
Now, I?, = cos <f,, and ( 1 - ~:) = (1 - cos2 q,) = sin 2 <f,
Therefore,
dcI>s 21rQv . 2
- =-- sm cp (5)
dt R
Therefore, by equation (2)
. Qv . 2
is= - sm cp (6)
2R
where is is the equivalent displacement current through the
segment P MP'; incidentally, also, the same displacement current
"flows" through the base of the segment PNP'.
So far as computing magnetic fields is concerned, we may treat
this equivalent current is as if it were an actual current. Since
the phenomena due to the motion of the charge along ON must
be perfectly symmetrical around ON, the magnetic field produced
at all points, such as P or P', on the circumference of the circle
PNP' will ~ everywhere the same. Calling this field HP (it ~s
everywhere tangent to the circumference), the work W done in
carrying a unit magnetic pole once around the circumference is
W = 21rr · Hp (7)
We may, also, compute the work W from the expression
W = 41ris (see equation (11), Chap. IV), viz.,
Q'v
W = 41r- sin 2 cp (8)
2R
126 CONCERNING MOVING CHARGES [CHAP. v
i;
(Note that the quantities and Q' are now expressed in electro-
magnetic units.) Equatmg the right hand sides of equations
(7) and (8) and solving for HP, we have
Q'v sin 2 cp
Hp= R . r
Q'v sin cp r
=R ·-r-·R
. Q'v .
. . HP = R 2 sm </> (9)
since sin </> = r/ R.
Equation (9) is the equation sought. It shows that the
magnetic field HP produced at any instant at a point P by a
moving charge Q is (1) proportional to the velocity of the charge,
(2) inversely proportional to the square of the distance R between
the charge and the point, and (3) depends on the sine of the angle
</> between the direction of motion and a line drawn from the
point to the charge.
A word about units: In equation (8), Q' must be in electro-
magnetic units if W is to be in ergs, to conform to the usual
units in which Wis given in equation (7). That is, in equation
(9), Q' must be in electromagnetic units to give HP in gauss. If,
as is frequently found convenient, one wishes to express the
charge in electrostatic units, one must write,
H 1 Qv .
P c
= R2 sm cp (10)
where c is the velocity of light.
It is to be observed that the field at all points on 0.1l1
(i.e., cp = 0) is zero. The field is a maximum at points given by
</> = 1r/2. Comparing equation (9) with equation (8) Chap. IV,
one observes that the moving charge behaves as if Q'v were a
"current element" i' · ds.
2. The Force Acting on a Charge Moving in a Magnetic Field.
Since HP as determined by equation (9), is the strength of
the magnetic field at P (Fig. 22), a magnetic pole of strength m
placed at P would experience a force f m perpendicular to the plane
of the paper (toward the reader) given by
fm = mHp (11)
By Newton's third law, the moving charge must experience a
force JQ equal in magnitude and opposite in direction to f m; or,
Q'v .
f O = - mHP = - m R 2 sm cp (12)
SEC. 3] ENERGY OF A MOVING CHARGE 127
Now, m/R 2 is the strength of the magnetic field Hat O due to the
pole m which we assumed placed at P. We may write, therefore
(omitting the minus sign),
JQ = HQ'v sin </> (13)
where </> is now to be taken as the angle between the direction of
motion of the charge and the direction of the field H through
which the charge is moving.
There are many cases where the charge moves in a direction
at right angles to the magnetic field. In this case, sin </> = 1,
and we have
JQ = HQ'v (14)
f O being at right angles to both H and v. If the charge moves
parallel to the field, sin </> = 0, and there is no force acting on
the charge.
3. The Energy Contained in the Magnetic Field Surrounding a
Moving Charge.-Consider a charge Q (electrostatic measure)
uniformly distributed over the surface of a small sphere S of
radius a (Fig. 23) and let S be
moving in the direction Ox with
a velocity v. Equation (10)
shows that the magnetic field O--+-+----i--"""-t---"-----.--+--+--X
at some point P, the (polar)
coordinates of which are p, 8,
is given by
Hp -_ Qvsin(}
-- (15)
cp2 Frn. 23.
Now, the energy per unit volume in a magnetic field of strength
H is H 2 /81r. In an elementary volume dV including P, the
energy dW is
dW = _!_ H 2dV (16)
81r p
Since the magnetic field is symmetrical around Ox, we may take as
an elementary volume a ring the cross-section of which is dp·pd(J
and the length of which is 21rp sin 8. We, thus, have
dV = dp · pd(} · 21rp sin (}
= 21rp 2 sin 8dpd8 (17)
Putting into equation (16) the value of HP from equation (15)
and the value of dV from equation (17), we obtain for dW,
dW = - Q
1 _v2 2 • 3 (}
~dpde (18)
4 c2 P2
128 CONCERNING MOVING CHARGES [CHAP. v
We can obtain the total energy W due to the magnetic field in
the space surrounding the moving charge by integrating equation
(18) between the limits 1 0 and 1r for (J and between the limits
a and oo for p, a being the radius of the sphere S. That is,
W -- -1 -Q2v21
4 C2 a
co l1r sin3 2
(J
dpdO (19)
O P
.
.. W-3
- ! Q2v21
2
co dp
2
* (20)
C a p
Integrating equation (20) with respect to p and putting in the
limits gives
1 Q2v2
W = - -2 (21)
3 ac
Thus, W is proportional to the square of the velocity of the
charged body; and the charge behaves as if it had mass, since this
energy W is quite independent of any kinetic energy which
the sphere itself may have. We may write equation (21):
1 2 Q2 2
W = - · - -v
2 3 ac 2
which, by comparison with the well known expression for kinetic
energy, 7~mv\ indicates that the equivalent mass mQ of the
charge Q is given by
2 Q2
mo - 3 ac2 (22)
This equivalent mass is seen to be inversely proportional to the
radius a of the sphere. (Note that Q is measured in e.s.u.)
4. The Energy Radiated by an Accelerated Charge.-Con-
sider a "point" charge initially at rest at O (Fig. 24). Let the
charge be given an acceleration g for a short timer, during which
time the charge reaches P 1, its velocity v at P1 in the direction
Ox being given by
V = g • T, (23)
v being small compared to the velocity of light. After an addi-
tional time t, large compared to the interval r, the charge now
traveling with uniform 2 velocity v will have reached some point
1 This integration gives the energy in a shell of radius p and of thickness dp.
* r,r
Jo
3
sin 3 (}d(} = [cos (} - cos
3
6]7r
= i.
o 3
2
Irrespective of the inertia of any material matter with which the moving
charge may be associated, the charge has inertia by virtue of the magnetic
field set up by its velocity, as explained in the preceding section.
SEC. 4] RADIATION FROM A MOVING CHARGE 129
P 2 , where P 1P 2 = vt. Let ONL represent any line of force from
the charge while stationary at 0. The disturbance due to
starting the charge will travel outward along this line of force
with the velocity of light c. Outside a sphere of radius c(r + t)
concentric about O the field will not have received "news" of the
motion of the charge and will, therefore, be radial from 0, as
shown by the line NL. Inside a sphere of radius ct concentric
about P 2 the field will be radial from P2, as shown by the line
P 2 M. Between these two spheres will be a shell within which
the transition from the one field to the other occurs. The transi-
tion is shown by the line MN, and, at the instant represented, i.e.,
L
Frn. 24.-Diagrammatic representation of the origin of an ether pulse.
(r + t) seconds after starting, the line of force formerly occupying
the position ONL is now given by P2MNL, the "break" MN
traveling outward with velocity c. It will very much simplify
the discussion to make the assumption that the points 0, P1, and
P2 are so close together as, practically, to be coincident. The two
spheres are, then, concentric, and the shell has a uniform thickness
o = er. This is justified if v < < c, and if ! < < t. Then,
OP2 = vt
very nearly ..
This spherical shell is an electromagnetic wave, or pulse,
originating as a result of the acceleration given to the charge.
The line MN represents the direction of the field within the
pulse. We have, for simplicity, represented MN as a straight
line. Its shape will depend on the nature of the acceleration
given to g. The generality of the discussion, however, is not
lessened by assuming that MN is straight, since, if MN be not
straight, we may divide the shell up into a large number of
130 CONCERNING MOVING CHARGES [CHAP. v
elementary shells within each of which the intercepted portion
will be straight. Let Fig. 25 represent, to a larger scale, a portion
of the shell in the neighborhood of MN, in Fig. 24. Let the
direction and intensity of the electric field, within the shell at
this point, be represented by F. We may resolve F into com-
ponents FT tangential to the surfaces of the shell, and FP per-
pendicular to these surfaces. From Fig. 25, it is obvious that
FT MN' (24)
Fp = MM'
But
MM'= o= er, (25)
FIG. 25.
o being the thickness of the shell, and r the duration of the
acceleration given to the charge. Also
MN' = 0 P2 sin (} = vt sin (} (26)
(see Fig. 24), subject to the limitations previously mentioned.
Therefore,
FT _ vt sin fJ
Fp - o (27)
Now, FP, the component of the field within the shell and per-
pendicular to the surfaces of the shell, must be exactly given by
Q
Fp = R2 (28)
where R is the radius of the shell. This is readily seen, since
every line of force such as P 2 M (Fig. 24) has an extension MM'.
Combining equations (27) and (28), we have
Qvt .
FT = oR2 sm O
= Qv sin fJ (29)
coR
since R = ct, approximately.
SEC. 4] RADIATION FROM A MOVING CHARGE 131
We may now regard the field surrounding the moving charge
as practically equivalent to a constant radial field moving at
uniform velocity with the charge, upon which is superimposed a
tangential field FP which moves radially outward with the
velocity of light. This tangential component, therefore, we
may consider as constituting the electric vector in an electro-
magnetic wave, associated with which must be, as we saw in Chap.
IV, a magnetic vector HT (perpendicular to the plane of the paper).
The energy per unit volume, w, within the pulse due to these two
vectors is
Fi Hi
w=81r+81r
or, since we have seen, in Chap. IV, that the energy associated
with the electric vector in an electromagnetic wave is exactly
equal to the energy associated with the magnetic vector, we may
write
w =-
Fi
41r
By equation (29),
- 1 Q2v2 . 2
w - 41r c2 R252 sm fJ
If we assume that w is constant from front to back of the pulse,
the "intensity" I of the pulse, i.e., the energy per unit area of
pulse surface, is
1 Q2v2 .
I = WO = 41r . c2R2o sm2 (} (30)
We, thus, see (1) that the intensity of the pulse is inversely
proportional to R 2 (the well known inverse-square law of radia-
tion); (2) inversely proportional to the thickness of the pulse; and
(3) is zero in the direction of motion of the charg~ (sin () = 0) and
greatest at right angles to this direction. The total energy W
contained in the pulse is determined by integrating this intensity
over the entire surface of the pulse, i.e.,
w = f Ids (31)
where dS is a surface element of the pulse.
This integration is readily performed. The spherical pulse,
radius R, is represented, in Fig. 26, with its center at 0. Since
the phenomenon is symmetrical about Ox, the direction of
motion, we may take as a surface element dS of the sphere a
1
132 CONCERNING MOVING CHARGES [CHAP. v
narrow zone centered at P, the width of the zone being R · do
and its length 21r-R sin e. Therefore,
dS = 21rR sin e · Rde
= 21rR2 sin e de (32)
Using equations (30) and (32), we may write, for equation (31),
W = -1 _v_
Q2 2l1rsin 3 OdO
(33)
2 c 2
o O
. 2 Q2v2
.. w = 3 c2 o (34)
The energy of the pulse is, therefore, inversely proportional to
its thickness.
FIG. 26.
Now, assuming that the acceleration g which the charge
experienced was uniform, v = g · r; also, o = er. We may,
therefore, write equation (34):
2 Q2g2
w= - -7
3 3 c
(35)
This important equation gives the total quantity of energy
radiated as a result of the acceleration given to the charge for
a time r. It leads, at once, to a still more important relation,
namely, the rate of radiation of energy dW /dt as a function of
the acceleration given to the charge. 1 Thus,
dW _ W _ 2 Q2g2
dt - --; - 3 c3 (36)
1 Strictly speaking, W /r gives the average rate of radiation of energy during
the time r. But the same result would be obtained by a more rigorous
discussion in which r is divided up into a large number n of time elements
dr and the shell or pulse into a corresponding number of elements of thickness
do. It can readily be seen that a uniform acceleration g would not result
in a straight line MN (Fig. 24).
SEC. 5] SOME SPEC! AL CASES 133
It is to be noted that, irrespective of whether g is positive or
negative, g2 is always positive; therefore, an accelerated charge
radiates energy whether the acceleration be positive or negative.
As stated at the end of Chap. IV, this is a necessary and an
unambiguous conclusion from the electromagnetic theory. It
explains why a vibrating electron should radiate electromagnetic
waves-the mechanism of the classical theory of light production.
It explains why a swiftly moving electron, when suddenly stopped
by collision, say with the target of an X-ray tube, should send
out an "ether pulse," which was an early theory of the process of
the production of X-rays.
5. Some Special Cases of Radiation by Accelerated Charges.-
It may be of interest to consider one or two applications of the
discussion in the preceding section.
First, to determine the ratio of the energy Wr radiated, when a
charge Q moving with velocity v is brought to rest, to the
energy Wm in the magnetic field initially surrounding the moving
charge: From equation (34),
2 Q2v2
Wr = 3 c20
This applies to the case either of a charge starting from rest and
acquiring a velocity v or to a charge initially moving with velocity
v being brought to rest. From equation (21), the magnetic
energy Wm associated with a charge moving with velocity v is
1 Q2v2
Wm= 3 ac2
Therefore,
2 Q2v2
---
Wr _ 3 c 2o
(37)
wm - 1 Q2v2
(38)
Thus, in the case of a charge in motion being brought to rest,
the fraction of the energy radiated is inversely proportional
to the thickness o of the resulting pulse, which, in turn, is pro-
portional to the time required to stop the charge.
Second; to determine the energy radiated per vibration by
a charge which, having been set into vibration by some disturb ..
ance, IS executing . simple harmonic vibrations about an equi-
134 CONCERNING MOVING CHARGES CHAP. v
librium position: Such a charge wHI gradually be brought to
rest by virtue of the energy radiated on account of the accelera-
tion which the charge experiences at various parts of its path.
At time t let the displacement x of the charge Q from its position
of equilibrium be given by the usual equation of simple harmonic
motion
x = A sin 21rvt (39)
where A is the amplitude of the vibration and v is the frequency.
Differentiating to obtain the velocity v and the acceleration g of
the charge gives
(40)
(41)
From equation (36), we have, for the quantity of energy dW
radiated in time dt,
~ Q 3g
2 2
dW = dt (42)
3 c
2 Q2
=
3 c3 (- 47r v A sin 27rvi) dt
2 2 2
. ·. dW
3271"4 Q2
= - - - 3 v4A 2 (sin 2 27rvt)dt (43)
3 c
To obtain the energy W 1 radiated during a complete vibration,
we must integrate this equation over an entire period, i.e., from
t = 0 tot = l/v. That is,
W1 = -
3271"4 Q2 4 2
3 3C
vA
It=~
t=O
(sin 2 27rvt)dt
The value of the integral is \· 1 Therefore,
2
W1 = 1671"4. v3 Q2A 2 (44)
3 c3
The absolute value of W 1 is not so important as the ratio Pi
which W 1 bears to the total energy of the vibrating system, for
it is this ratio which determines the number of waves in the
electromagnetic wave train which the vibrating charge will
1
f sin 2 q,dq, = q,/2 - 74 sin 2,t,. To integrate, let q, = 21r;t. Then dt
becomes dct> and the limits become O and 21r, respectively.
27r'II
SEC. 5] SOME SPECIAL CASES 135
emit before it is brought to rest by the "damping" due to the
process of radiation. The total energy of the vibrating system
is given by
Wo = }~mv~ (45)
where m is the (electromagnetic) mass of the charge (m is given
by equation (22)) and Vm, is the maximum velocity, i.e., the
velocity at midpoint of the path. From equation (40), v! is
given by
(46)
Therefore,
(47)
This equation shows that the fraction Pi is proportional to the
frequency v of the vibration. That is, longer wave trains will
be emitted when the frequency of vibration is low than when it is
high.
An alternative, and rather suggestive, expression for the
ratio Pi is obtained by introducing into equation (47) the value
of m from equation (22). We, then, have
2 v
Pi = 41r a -
c
= 41r2 --a (48)
A
where X is the wave length of the radiation emitted by the
vibrating charge and a is the ''diameter'' of the charge.
In subsequent chapters, we shall consider the application of
some of the important formulre derived in this chapter.
CHAPTER VI
THE PHOTOELECTRIC EFFECT
An approach to that outstanding problem of modern physics>
namely, the nature of radiant energy, is most directly made by
a study of the experimental facts of the photoelectric effect.
In its broadest meaning, the term '' photoelectric effect'' might
be used to include several phenomena, more or less related, which
involve the interaction between light and electricity, such as
the change in resistance of selenium when illuminated by red or
near infra-red light; the change in e.m.f., when illuminated, of
certain cells containing fluorescent materials, as electrolytes;
or, indeed, the rotation of the plane of polarization of a beam of
light when passing through a medium placed in a strong electric
field. In the broad use of the term, even dispersion is a photo-
electric phenomenon, since it involves a reaction between light
and the electrical charges in matter. That such interactions
should take place is not surprising if light be an electromagnetic
phenomena and if matter be made up of, or contain, charged
particles. In this chapter, we shall restrict the use of the term
"photoelectric effect" to the discharge of negative electricity
from bodies when illuminated by light of appropriate wave
length. We shall make no attempt to give a complete outline 1 of
this interesting subject. Rather we shall discuss photoelectricity
only in so far as may be necessary to point out the embarrassing
situation in which the electromagnetic theory of light finds itself
in attempting to account for the phenomena observed.
1. The Discovery by Hertz.-Reference has already been
made (p. 111) to the experiments of Hertz in which he demon-
strated the existence of the electromagnetic waves predicted by
Maxwell. In these experiments, Hertz was using two spark
gaps, which we may represent diagrammatically by P and S
(Fig. 27). The terminals of P were suitably connected to an
induction coil, and the spark discharge across P was oscillatory,
1 The reader is ref erred to ALLEN, H. STANLEY: "Photo electricity" 2d
ed. (1925).
136
SEC. 2] SOME EARLY EXPERIMENTS 137
g1vmg rise to electromagnetic waves. These waves were
"detected" by a suitably tuned circuit containing the spark
gap S, the length of which could be adjusted by a micrometer.
The maximum length of S which would give continuous sparking
was taken as a measure of the strength of the ''signal'' received by
the secondary. But Hertz observed that this critical length
of the spark gap S seemed to be determined not only by the
strength of the signal but also by other factors as well. For
example, when, to facilitate observation, he enclosed S in a
black box, say of cardboard, the critical length of S was much less.
This disturbing factor might have escaped further study in
the hands of a less observant investigator. But Hertz essayed
to discover its cause. He observed that
the interposition between P and S of any
t
opaque object, such as a piece of board
or black paper, was just as effective in
reducing the critical length of S as if S
were completely surrounded by the box.
Further tests showed that a piece of
glass interposed between P and S acted
as an opaque screen. A plate of quartz,
however, produced no effect. Since glass
I
Frn. 27.
is opaque to ultra-violet light, while
quartz is transparent, and since the primary spark P produced
ultra-violet light, Hertz concluded that the increase in the critical
length of S when not shielded from P was due to the ultra-
violet light from the primary spark gap P falling on the secondary
gap S. He tested this conclusion in various ways, for example
by screening S from P and allowing ultra-violet light from another
source to fall on S. And he found that the light must fall on the
terminals themselves, not simply in the space between. Illumi-
nating the negative terminal was most effective. Further, the
terminals must be "clean and smooth." Recognizing the impor-
tance of the discovery, Hertz published his results 1 in a paper
entitled "An Effect of Ultra-violet Light upon the Electric
Discharge" and then returned to his investigation of electromag-
netic waves.
2. Some Early Experiments.-The extraordinary importance
of Hertz's discovery at once attracted numerous investigators.
1 Ann. Physik, vol. 31, p. 983 (July, 1887). See, also, the English transla-
tion of HERTz's ''Electric Waves."
138 THE PHOTOELECTRIC EFFECT [CHAP. VI
A very fundamental result was soon announced by Hallwachs, 1
who found that a freshly polished zinc plate, insulated and
connected to an electroscope as an indicator, would, when
charged negatively and illuminated by ultra-violet light from
an arc lamp, lose its charge, but that there was no effect if the
charge was positive. 2 Aluminum was found more active than
zinc; iron, less so. Hallwachs concluded that, in some mysteri-
ous manner, when the plate was negatively charged and illumi-
nated, negatively electrified particles were emitted from the
plate. He even observed, by using an electrometer instead of
an electroscope, that a neutral insulated plate, when illuminated,
would acquire a small positive potential, i.e., in present-day
terminology, would lose a negative charge. In other words,
not only did light permit the escape of
p sI negative electricity from a negatively
I
I electrified plate : it even caused the
I
I expulsion of negati"ve electricity from a
I
I neutral plate.
L ____
_.- B l+
)a Stoletow 3 devised an arrangement,
shown diagrammatically in Fig. 28, for
producing a continuous photoelectric
current. P is a photoelectrically sen-
Frn. 28.
sitive plate, say a polished zinc plate,
connected to the negative terminal of a battery B of several cells.
S is a wire grating or gauze connected to the positive terminal
of the battery through a very sensitive galvanometer, or an
electrometer G. When ultra-violet light falls upon P, a continuous
current is observed in G, indicating that a negative charge is flowing
from P to S. No current flows if the battery be reversed. P corre-
sponds to the cathode in an electrolytic cell, and S to the anode.
Elster and Geitel 4 showed that there is a close relation between
the contact potential series of metals and the photoelectric
1 Ann. Physik, vol. 33, p. 301 (January, 1888).
2 This experiment is very easily performed. Take a piece of zinc, scratch
one side with sandpaper so as to expose a fresh surface, suspend it by a wire
or string from an ebonite rod or other good insulating support and connect
the plate to an ordinary electroscope. An ordinary (carbon) arc lamp is a
suitable source of ultra-violet. The difference between the polished and
the unpolished side is at once observable, as is also the effect of glass, quartz
plates, etc., between source and plate.
3 Jour. phys., vol. 9, p. 486 (1890).
4 Ann. Physik, vol. 38, pp. 40 and 497 (1889).
SEC. 3] A PROBLEM 139
effect. The more electropositive the metal the longer the wave
length to which it would respond photoelectrically. The
alkali metals, sodium, potassium, and rubidium, were sensitive
even to light of the visible spectrum.
It was early observed 1 that the photoelectric effect took place
even in the highest attainable vacuum. For exampJe, if the
plates P and S, in Fig. 28, be enclosed in a glass bulb fitted with
a quartz window for admitting ultra-violet light, the photo-
electric current is observed to flow even if the bulb be evacuated
as thoroughly as possible. This shows that the gas surrounding
the plate plays no essential role in the photoelectric effect.
3. A Problem.-An important question now arises. What is
the mechanism by which negative electricity is transferred from
the cathode P to the anode S, Fig. 28? That the charge is carried
by negatively electrified particles was suggested by the experi-
ments of Elster and Geitel, 2 who showed that a transverse mag-
netic field diminishes the photoelectric current if the phenomenon
takes place in a vacuum. But what are the particles?
Negative answers to this question were readily obtained. The
fact that the effect persisted even to the highest attainable
vacuum and was quite independent of the "degree" of the
vacuum after a certain low pressure had been reached seemed
to indicate that the gas molecules themselves in the region
between P and S could not act as carriers of the charge.
The suggestion was made that, perhaps, under the influence of
light, particles of the cathode became detached, and since they
would naturally be charged negatively, they might move under
the influence of the electric field from cathode to anode. This
suggestion was made untenable by an experiment by P. Lenard, 3
in which a clean platinum wire acted as anode, and a sodium
amalgam as cathode, both being in an atmosphere of hydrogen.
The photoelectric current was allowed to flow until about 3 X
10- 5 coulombs had passed through the circuit. If the carriers
of the charge be atoms of sodium, each atom could hardly be
expected to carry a larger charge than it carries in electrolysis.
Taking the electrochemical equivalent of sodium as 0.00024
grams per coulomb, there should have been deposited on the
1 Rrnm, M em. Acad. di Bologna, vol. 9, p. 369 (1888); STOLETow, J our.
phys., vol. 9, p. 468 (1890). LENARD, P., Ann. Physik., vol. 2, p. 359 (1900).
2 Ann. Physik, vol. 41, p. 161 (1890).
3 Ann. Physik, vol. 2, p. 359 (1900).
140 THE PHOTOELECTRIC EFFECT [CHAP. VI
platinum wire at least 0.7 X 10- 6 milligrams of sodium, a quan-
tity sufficient to be detectable by the well-known flame test. 1
On removing the wire from the bulb, no trace of sodium could
be detected. Evidently, particles (molecules) of the cathode
did not act as carriers of the photoelectric current.
If the photoelectric current was carried neither by molecules
of the gas surrounding the cathode nor by molecules of the
cathode itself, what were the carriers? · The answer to this
question came as a result of the convergence of a number of
different lines of evidence which finally culminated in the dis-
covery of the electron by J. J. Thomson and in direct experimental
proof of the electrical nature of matter. We shall interrupt
the discussion of the photoelectric effect proper to consider
briefly some of these developments.
4. The Laws of Electrolysis.-The laws of electrolysis dis-
covered by Faraday suggested that associated with each uni-
valent ion was an elementary charge e which was the same for
all ions. This is readily seen. Faraday's two laws may be
stated
M = k ·i ·t = kQ (1)
A
k=h- (2)
v
where M is the mass, in grams, of one type of ion (say Cu+ in the
electrolysis of Cu80 4 ) deposited by a current i flowing for t
seconds; k is the electrochemical equivalent of the ion; A is
the atomic weight; and V the valency of the ion in the compound;
h is a proportionality constant which is numerically equal to
the electrochemical equivalent of hydrogen for which, in equation
(2), A and V are both unity. 2 Also, by definition, Q = i · t.
Combining equations (1) and (2), we may solve for Q:
Q = t. 1. v (3)
A gram-atom of any element is a quantity of the element, the
mass of which MA in grams is numerically equal to the atomic
weight. That is,
MA= A
for a gram-atom. Thus, a gram-atom of Na is 23.00 grams of
Na; of Cl, 35.46 grams; of Cu, 63.57 grams; etc. When Na and
1 Bunsen had shown that 0.3 X 10-6 mg. of sodium could easily be
detected in the flame.
2
A for hydrogen is really 1.0077.
SEC. 4] LAWS OF ELECTROLYSIS 141
Cl unite, atom for atom, to form NaCl, 23.00 grams of Na com-
bine with 35.46 grams of Cl. That is, 23.00 grams of Na contain
the same number of atoms as 35.46 grams of Cl. Similarly, 1
gram of hydrogen contains the same number of atoms as 35.46
grams of chlorine. This number of atoms N in a gram-atom of
any element is the same for all elements and is called Avogadro's
number; it is one of the fundamentally important constants in
chemistry and physics. We shall discuss its numerical value later.
Similarly, if we define a gram-molecule, or a mole, as a quantity
of the substance the mass of which in grams is numerically
equivalent to the molecular weight (e.g., 159.63 grams 1 of CuS0 4 );
or a gram-ion as a quantity of the ion the mass of which is
numerically equal to the "ionic" weight (e.g., 96.06 grams of
804), it is obvious that the number of molecules in a gram-mole-
cule, or of ions in a gram-ion, is, also, N.
If, by electrolysis, we free a gram-atom, MA, of a univalent
element, equation (3) shows that the quantity of electricity
QA required is given by
1
QA = h (4)
since, then, MA = A and V = 1. This quantity QA is, thus,
seen to be the same for all univalent atoms and is a universal
constant. It is called the "faraday.'' Its numerical value is
readily obtained by taking the reciprocal of the electrochemical
equivalent h of hydrogen. 2 According to the best modern
determinations,
QA = 9,650 electromagnetic units
But a gram-atom of any element contains N atoms; and to
free these N (univalent) atoms, in electrolysis, requires QA units
of electricity, whatever the element. The charge e carried by
each atom (we should, more correctly, say "ion"') is, therefore,
obviously,
(5)
1 Taking atomic weights, viz.: Cu, 63.57; S, 32.06; 0, 16. These are
more correctly called "combining weights." In any event, they do not
give the relative weights of atoms of the several elements. There is no
atom of chlorine, for example, which has a relative weight (mass) of 35.46,
taking oxygen 16. Rather, chlorine has two kinds of atoms: one group with
a relative mass of 35 and the other with a relative mass of 37; and these
two kinds of chlorine atoms (isotopes of chlorine) are mixed in the approxi-
ms.te proportion of 3 of the former to 1 of the latter (see Chap. XIII).
2 See Nate 2, preceding page.
142 THE PHOTOELECTRIC EFFECT [CHAP. VI
and is the same for all univalent ions. The value of QA was early
determined. Had N been known, the value of the elementary
charge e could, also, have been determined. As a matter of
fact, we shall see, later, that the best method of determining N,
the number of atoms in a gram-atom, is by dividing QA by the
independently determined value of e.
Although the facts of electrolysis do not give a means of
determining e, they do give a method of determining the ratio of
the charge e to the mass m of the atom or ion. For, the mass of
an atom is given by 1
A
m=- (6)
N
where A, as before, is the atomic weight.
And accordingly,
(7)
That is, the ratio of the charge e carried by an ion to the mass m
of the ion is obtained by dividing the faraday QA by the atomic
weight A. Since QA is constant, the ratio e/m is inversely pro-
portional to the atomic weight A. The maximum possible value
of e/m for univalent ions in electrolysis is that for hydrogen
for which, since A for hydrogen is unity, (;) H = 9650. Thus,
the facts of electrolysis gave fairly direct experimental evi-
dence of the association with each atom or ion of an elementary
charge e (or a charge 2e with a divalent ion), but they afforded no
means of determining the value of e.
Computations based on the kinetic theory of gases, however,
indicated that N is, at least, of the order of 10 23 • Consequently,
by equation (4), e was known to be of the order of 10- 20 e.m.u., or
10- 10 e.s.u.
5. Dispersion of Light.-We have seen, in Chap. IV, that the
failure of the early form of the electromagnetic theory of light to
predict the phenomenon of dispersion led to the introduction of
the concept, first developed by L. Lorenz,2 of the existence in
refractive media of charges, or charged bodies, with a charge e
and mass m which, when disturbed, could vibrate with a natural
period vo about a fixed, equilibrium position. The extension of
1 Since A grams of the element contain N atoms.
2
Ann. Phy.sik, vol. 11, p. 70 (1880).
SEC. 6] THE 'ZEEMAN EFFECT 143
this theory by H. A. Lorentz 1 and others, and the success with
which the resulting formulre explained the observed facts of
dispersion, both normal and anomalous, gave great weight to
the concept of the electrical constitution of matter.
6. The Zeeman Effect.-In 1862, Faraday, looking for a
possible effect of a magnetic field upon a light source, placed a
sodium flame between the poles of a strong electromagnet and
examined the D lines by a spectroscope. He was unable to
detect any change in the appearance of the lines by turning on
the magnetic field.
Faraday's failure to observe the effect which he expected was
due to the inadequate resolving power of his apparatus. For,
in 1896, Zeeman, 2 repeating Faraday's experiment, with the
improved technique then avail-
able, discovered that spectral
lines are split up into components
when the source emitting the
lines is placed in a very strong R
magnetic field and, further, that
these components are polarized.
The simplest case is shown in
Fm. 29.-Diagrammatic represen-
Fig. 29, where a represents a line tation of the Zeeman effect: a is the
before the magnetic field is original line; R shows the three com-
ponents when viewed at right angles
turned on. If the field be turned to the magnetic field; P shows the two
on and the line be viewed at components when viewed parallel to
the field.
right angles (R) to the direction of
the field, the line is seen to be triple with components ZR, a', and
rR. The central line a' has the same wave length as the original
line a but is plane polarized in a plane at right angles to the mag-
netic field H, the direction of polarization being indicated by the
double arrow above the line. The other two com·ponents ZR (left,
shorter wave length) and rR (right, longer wave length) are plane
polarized in a direction parallel to the ma.gnetic field.
If the pole pieces of the electromagnet be drilled through
longitudinally so that one may view the flame in a direction
parallel to the magnetic field, only two components ZP and rp are
seen, as shown at P. These two lines have the same respective
wave lengths as the outside components in the previous case,
but they are circularly polarized in opposite directions, as shown
1 Lo RENTZ, H. A. : "The Theory of Electrons" (•1909).
2 Phil. Mag. 1 vol. 43, p. 226 (1897).
144 THE PHOTOELECTRIC EFFECT [CHAP. VI
by the arrows above the lines. (The magnetic field, in this
latter case, is directed toward the reader.)
In his original paper announcing his discovery, Zeeman
offered a relatively sjmple, yet at that time adequate, explanation
of the phenomenon on the basis of Lorentz' electrical theory of
matter mentioned in the preceding section. Let it be assumed
that light waves originate in the vibratory motion of electric
charges associated with atoms, and consider the effect of a
magnetic field on such vibrations. In Fig. 30, let O represent
the normal position of equilibrium, within an atom, of a particle
which has a charge 1 e' and a mass m. ,vhen the particle is
displaced, in any direction, by an
amount x from 0, let the return force
/ e' f be given by
..:r ,'
A- - - - - - __..::_: - - -.B
f = -px (8)
0
where p is a constant. When so dis-
turbed, the particle will vibrate about
its position of equilbrium with a period
Fm. 30. T, given by
(9)
/
according to the ordinary laws of simple harmonic motion. Or,
if the particle be set in motion in a circular path of radius r
about point .0 as a center, the motion will be governed by the
equation
mv 2
-=pr (10)
r
Let wo be the angular velocity about 0. Then v = Wor, and,
by equations (9) and (10),
(11)
Now, let a magnetic field of strength H be applied at right
angles to the plane of the paper (Fig. 30) and directed away from
the reader. The charge, in its orbital motion, will experience
an additional force JII, due to its velocity v (at right angles to the
field H), which, according to equation (14), Chap. V, will be
fn = He'v (12)
1
The prime (') denotes that e is in electromagnetic units.
SEC. 6] THE ZEEMAN EFFECT 145
If the charge be positive, and the motion counterclockwise, as
shown, f H will be in the same direction asf, i.e., toward 0, and we
may write, instead of equation (10),
mv 2
- =pr+ He'v (13)
r
Putting v = w 1r, where w1, is the angular velocity after the field
is applied, we have, after rearranging,
2
W1 -
He- ' WI -
- 2
Wo (14)
m
since, by equation (11), p/m = w6. Completing the square
on the left-hand side of equation (14) and extracting the square
root gives
W1 -
He' - ~ w 2 + H-2e'-2 (15)
2m - 0 4m 2
Or, since, as we shall show later, we may neglect the second
term under the radical in equation (15) in comparison with
w~, we have
He'
WI = WO ~ + 2m
(16)
If we had considered a clockwise motion instead of a counter-
clockwise one, we should have obtained
He'
W2 = WO - - (17)
2m
Changing angular velocities to frequencies of revolution by
writing wo = 21rvo, w1 = 21rv1 and w2 = 21rv2, we have, instead of
equations (16) and (17),
He'
v1 = vo + --
41rm
(18a)
He'
V2 = Vo - - (18b)
41rm
Equations (18) show that the frequency of the radiation
emitted by the charge in orbital motion in the magnetic field, as
in Fig. 30, will be either greater, v1, or less, v2, than the frequency
vo emitted when the field is absent, according as the motion is
counterclockwise or clockwise; and, further, that this difference
is proportional to the field strength H.
It can be shown that these principles are equally applicable to
the case of a linear vibratory motion, as, for example, a motion
over the path AOB (Fig. 30). For, we may regard a linear sim-
ple harmonic vibration as made up of two equal and oppositely
146 THE PHOTOELECTRIC EFFECT [CHAP. VI
directed circular motions, each of which, under the influence of
the magnetic field, will behave as specified by equations (18),
the counterclockwise component moving with greater angular
velocity than the clockwise. The effect of this is to cause the
path AOB to rotate, slowly, counterclockwise about 0. Indeed,
this is exactly what would happen to the path of a charge set
vibrating at right angles to a magnetic field. In Fig. 31, let the
magnetic field be, as before, perpendicular to the plane of the
paper and directed away from the reader. A positive charge
moving from A 1 toward B 1 will experience a force, at right
angles to its motion, which will deflect it upward toward B2.
•P
Fm. 31.--Effect of a magnetic field on a charged body vibrating initially with
simple harmonic motion at right angles to the field.
On the return path, the direction of the force due to the magnetic
field is reversed, and the path bends downward toward A 2 , etc.
Thus, the path rotates slowly about 0. When the path is in the
general direction A1B1, the intensity of the radiation sent out
toward point P is a minimum; 1 while, when the path coincides
with A4B4, the radiation toward Pis a maximum. An observer
at P would receive from the vibrating charge a radiation of
periodically varying intensity-which is exactly what would be
observed if the source of radiation were two vibrating systems of
slightly different period, their effects at P being alternately in
phase and out of phase. A spectroscope, receiving this radia-
tion, should resolve it into two components of slightly different
frequency. We are, therefore, justified (1) in regarding the
(linear) vibratory motion as made up of two oppositely directed
1 The apparent amplitude, as seen from the direction of P, is then small.
SEC. 6] THE ZEEMAN EFFECT 147
circular motions, which, under the influence of the magnetic
field, revolve with slightly different frequencies and (2) in apply-
ing to such a vibratory system equations (18).
The observations recorded diagrammatically in Fig. 29 may be
readily explained on the basis of the above principles. Assum-
ing that light waves are the electromagnetic waves emitted by
electric charges vibrating with simple harmonic motion about
fixed equilibrium positions within the several atoms of the light
source, let O (Fig. 32) be the light source which is placed in a
magnetic field of strength H, the direction of the field coinciding
with the x-axis. We may resolve the vibrations of the charges
into components parallel to each of y
the three axes, respectively. Vibra-
tions parallel to Ox will not be
affected by the presence of the field
and will send out toward z radiation
of the same frequency vo, as if there
were no field. This accounts for the H
OJ...-------~.x
presence of the ''undeviated'' line a',
of Fig. 29, R. But this same vibra-
tion will emit no radiation in the
direction Ox, since, to an observer z
FIG. 32.
looking from x toward 0, the vibrat-
ing charge will seem to stand still. This explains the absence of
the undeviated component in Fig. 29, P. Components of vibra-
tions at right angles to Ox will behave exactly as illustrated in Fig.
31, and there will be emitted in the direction Oz, as well as in the
direction Ox, radiations whose frequencies differ from vo by the
He' .
amount + 1rm' as given by equations (18). This accounts
4
for the lines ZR, rR, lP, and rP, in Fig. 29.
If ~v represents the difference in frequency between the
original line and any one of the shifted components, we have
· He'
~v = ---
41rm
(19)
!!_ - 41r~v
m- H
By measuring ~v and the corresponding value of H, it is possible
to compute the ratio e' /m, that is, the ratio of the charge e' of the
vibrating particles to the mass m of the particles. In this way,
148 THE PHOTOELECTRIC EFFECT [CHAP. VI
Zeeman, in his first paper (March, 1897), determined 1 that '' e' / m
is of the order of magnitude of 10 7 electromagnetic units."
In a later experiment, 2 working with much higher resolving
power, Zeeman found a value of e' /m of 1.6 X 10 7• In this
experiment, he used a magnetic field of 22,400 gauss, which
caused a wave-length change, in the case of the D lines of sodium,
of about 1 part in 18,000.
The Zeeman effect and its explanation seemed to give final
confirmation of the electromagnetic origin of light and to adduce
some evidence as to the nature of the charged particles which
take part in the process, for the value of e' /m determined in this
way is in substantial agreement with the value determined by
J. J. Thomson for cathode particles by a method which will be
described in the next section.
7. The Discovery of the Electron by Sir J. J. Thomson.-
Ca) Previous to 1897, many studies had been made of that beauti-
c I)
+
FIG. 33.
ful phenomenon, the discharge of electricity through rarified gases.
Diagrammatically, let the discharge from an induction coil or an
electrostatic machine pass between the negative terminal C
(cathode) and the positive terminal A (anode), sealed into a
glass tube (Fig. 33), which is being exhausted through the side
tube T. At atmospheric pressure, the discharge, if the potential
be high enough, is a spark between the terminals. As the evacu-
ation proceeds, this spark widens out into a glow which fills the
whole tube; then striations appear, which, as the pressure gets
lower, recede toward A. At a very low pressure, there appears
around the cathode a dark space, known as the '' Crookes' dark
space," which, with further decrease in pressure, grows longer
(i.e., extends farther toward A), until, finally, the edge of the
dark space reaches the glass walls of the tube~ The glass is then
1 Zeeman's first apparatus was not of sufficient resolving power to separate
the components. He observed a widening of the line on turning on the field.
2 Phil. Mag., vol. 44, p. 255 (September, 1897).
SEC. 7] DISCOVERY OF THE ELECTRON 149
observed to glow with a color (greenish or bluish), depending on
the kind of glass of which the tube is made. This ''something''
which seems to proceed from the cathode and causes ~he phos--
phorescence of the glass was early called "cathode rays."
There was much controversy as to the nature of these "rays."
The German school of physicists held the opinion that the rays
were due to some process in the ether; while the English physicists
believed that the rays consisted of particles, charged with nega-
tive electricity and shot off at high velocity from the cathode.
In support of the latter view were experimental facts such as
the following:
1. The rays are deflected by a magnetic field. If the cathode
rays be restricted to a narrow pencil by screens 81 and 82
(Fig. 33), a small phosphorescent spot is observed on the
glass wall at D. If the tube be then subjected to a mag-
netic field the direction of which is at right angles to the
plane of the paper and directed away from the reader, the
spot D moves downward-which is exactly what would
take place if the rays consisted of negatively charged
parti~les moving from C toward A.
2. The rays are deflected by an electrostatic field. With screens
81 and 8 2 in place, as shown, an electrostatic field parallel
to the plane of the paper (and at right angles to the
direction CD) will deflect the spot D either upward or
downward, according as the positive direction of the field
is downward or upward.
3. If 81 and 82 be removed, the entire end of the tube at D
will glow. But a small, solid obstacle, placed at 8 3 and
partly filling the tube, will cast a shadow at D, indicating
that "something" is proceeding from C toward D.
4. The rays exert mechanical effects. (a) If the cathode be
suitably curved, the rays may be "focused" at som~ point
within the tube. A piece of platinum foil placed at this
focus may, if the rays be sufficiently intense, be heated to
incandescence. (b) A "paddle" wheel, mounted on suit-
able guides, may be made to revolve by making the rays
strike the paddles.
5. The rays transport negative electricity. This was shown
by Perrin, 1 by allowing the rays to enter an insulated
chamber connected to an electroscope.
1 Compt. rend., vol. 121, p. 1130 (1895).
150 THE PHOTOELECTRIC EFFECT [CHAP. VI
Final confirmation of the correctness of the view that cathode
rays are moving negatively charged particles came from the
classical experiments 1 of J. J. Thomson, whose quantitative
determination of the ratio of the charge e' of the particles to their
mass m carried more weight than did the qualitative evidence
mentioned in the preceding paragraph.
(b) Thomson's experiments are so fundamental in the history
of the electrical theory of matter as to warrant description.
The highly evacuated glass tube (Fig. 34(a)), contains the cathode
C and an anode A, which has a small rectangular slot in it
through which the cathode rays may pass. B is a screen similar
+
--~~~----~~~v-
a
r-<- ,- - - - - -- - - - - - l --------- - ->Je-L
B
(a)
--==--=--=--=--=--_-- - -~ - - - - - - - -?i
---.::t. --
:-:-~~-----------,}]
t. ••
4
l •- - - - --- --~
- - - - • -
--- ~
(b)
Frn. 34.-(a) Thomson's apparatus for determining the ratio e/m for electrons;
(b) Deflection of an electron by the electrostatic field in Thomson's apparatus.
to, and electrically connected with, A. Cathode rays, accelerated
from C toward A, after passing the slot in A move with uniform
velocity and emerge from the slot in B as a small bundle of rec-
tangular cross-section, which, on reaching the far side of the
bulb, causes a small, phosphorescent patch at some such point
as P1. But when a potential difference Vis maintained between
the parallel plates D and E, E being positive, the spot appears
at P2, having been deflected downward by the electrostatic field.
A pair of Helmholtz coils, not shown, whose diameters are equal
to the length of the plates D and E are placed, one in front and
the other behind the tube, so as to produce a magnetic field,
perpendicular to the plane of the paper, the strength of which
can be determined from the dimensions of the coils and the
1 Phil. Mag., vol. 44, p. 293 (October, 1897).
SEC. 7] DISCOVERY OF THE ELECTRON 151
c:urrent through them. If the magnetic field is directed toward the
reader, the spot P 1 will be deflected upward.
Two experiments are now performed:
1. With a given electrostatic field between the plates deflect-
ing the spot downward, the strength of the magnetic field
deflecting the spot upward again is adjusted to such a
value as will cause the spot to return to the original
undeviated position P1.
2. The magnetic field is removed and the deflection P 1P 2
caused by the electrostatic field alone is measured.
From these two experiments, the ratio e' /m may be determined
as follows:
Let e' = charge of moving particle (in e.m.u.).
m = mass of particle.
d = distance between plates D and E.
l = length of plates.
V = potential difference (in e.m.u.) applied between D
and E.
F = V / d = electric field between plates.
H = magnetic field.
v = the (horizontal) velocity 1 with which particles emerge
from B.
In experiment 1, the downward force fF, produced by the electro-
static field and given by
JF = F · e' (20)
is just equal and opposite to the upward force f n, due to the
magnetic field and given by
fn = He'v (21)
Therefore,
He'v = Fe'
or,
F
v=- (22)
H
Experiment 1, therefore, gives a means of measuring the (hori-
zontal) velocity with which the particles pass between the
plates. This velocity is determined, of course, by the potential
1 This horizontal ve1ocity is constant over the entire path B to P 1 since
there are no horizontal forces acting on the particle.
152 THE PHOTOELECTRIC EFFECT [CHAP. VI
difference applied between C and A and, to a certain extent, by
the vacuum.
From the second experiment, it is possible to determine the
downward deflection which the particles experience in passing
between the plates. Let this deflection be S (Fig 34 (b)). The
downward force experienced by the particle while passing between
the plates is given by equation (20), and the downward accelera-
tion a is given by
Fe'
a=-
m
This downward acceleration is constant over the length of
path l. If t is the time required to pass over this path, then, by
the ordinary laws of uniformly accelerated motion,
1
S = - F!'_t 2 (23)
2 m
where t = l/v. All quantities in this equation are known except
e' /m, which may, therefore, be computed.
Thomson found that the value of e' /m determined in this way
was of the order of 10 7 and that it was independent of the kind of
gas in the tube (air, H2, or C0 2) and, likewise, independent of the
material of the electrodes (Al, Fe, or Pt). A later determination
gave
-~' = 1.7 x 10 7 (24)
m
a value numerically identical (almost) with the value of e' /m
determined from the Zeeman effect for the particles taking part in
light emission, which particles, therefore, seemed to be identical
with cathode rays.
(c) This value of e' /mis very much larger (seventeen hundred
times) than the value of e' /m for hydrogen atoms in electrolysis,
as discussed on page 142. This large value of e' /m might result
either from a large value of e' or a small value of m, or both.
It became a matter of much importance to determine the charge
e' carried by these particles.
The charge e' was measured by H. A. Wilson, by a method
based on the following principles: When a gas, saturated with
water vapor, is cooled by expansion, condensation into droplets
tends to take place around any dust particles present. Similar
condensation will take place if the gas contains positive or
negative ions-such as are produced by the passage of X-rays or
SEC. 7] DISCOVERY OF THE ELECTRON 153
gamma rays of radium through the gas-condensation taking
place more easily around the negative ions than around the posi-
tive. By suitably regulating the expansion, it is possible to get a
"cloud" of small, negatively ch~rged droplets. The mass M
of a given droplet can be determined by observing, with a micro-
scope, the velocity v, with which the droplet slowly falls under
the action of gravity, v depending on the viscosity 'YJ of the air,
the radius a of the drop, and the acceleration due to gravity g,
according to the (approximate) equation
v = ~ (J_ a2 (25)
9 'YJ
Observing v, and knowing g and rJ, a may be determined, from
which
M = ;i1ra 3p (26)
where pis the density of water, assumed unity. The downward
force due to gravity, Mg, may now be balanced by an electric
field F causing an upward force F · e', so that
Mg = Fe',
from which e' may be determined. The value of e' was found,
in this way, to be of the order of 1 X 10-20 electromagnetic
units, or 3 X 10- 10 electrostatic units, in substantial agreement
with the approximate value of the charge carried by a univalent
ion in electrolysis, as discussed in Sec. 4 of this chapter.
Clearly, the large value of e' /m for cathode particles, cited in
equation (24), is due to a very small value of m rather than to a
large value of e. Assuming, as is seen to be approximately
justified, that the charge carried by the cathode particle is the
same as that carried by the hydrogen atom in electrolysis, it
follows that the mass of the cathode particles must be of the order of
1)1,700 of the mass of the hydrogen atom, which up to the time of
Thomson's experiments, was the smallest known particle.
(d) Here, then, in the cathode stream, we are dealing with
previously unknown particles, of very small mass and carrying
a negative charge, which Thomson referred to as "corpuscles,"
or "primordial atoms," but which have since become known as
electrons, and which are now regarded as one of the two primordial
substances out of which all matter is composed. Since its
discovery by Thomson, in 1897, the electron has played such
!1:l P.veryday part in modern physics that a detailed discussion of
154 THE PHOTOELECTRIC EFFECT [CHAP. VI
it need not be undertaken here. According to the most recent
data, 1 the values of e, m and e/ m for the electron are
e' = 1.592 X 10-20 e.m.u.
e = 4.774 X 10- 10 e.s.u.
e'
- = 1.769 X 107 e.m.u. per gram
m
e
- = 5.305 X 10 17 e.s.u. per gram
m
m = 8.999 X 10-28 grams
The discovery of the electron by Thomson and the recognition
of the fact that it is one of the constituent parts of all atoms
made possible the explanation of a number of more or less diverse
phenomena on the basis of a single concept. Thus, the electrical
nature of matter was, at once, made apparent. Since atoms in
the normal state are observed to be electrically neutral, if they
contain negatively charged electrons, they must, also, contain
positive electricity in some form. The phenomenon of electric
conduction in metals was ascribed to the existence, in the inter-
atomic spaces of the metal, of "free" electrons, which, drifting
under the influence of an applied electric field (i.e., the potential
difference at the terminals of the conductor), caused the transport
of charge, the actual direction of motion of the electrons being
opposite to the convention concerning the positive direction of
current. The variation of electrical resistance with temperature
also received a ready explanation. The fact that the ratio
e/m for electrons was observed to be the same as the ratio of
e/m for the vibrating particles in the Zeeman effect lead directly
to a confirmation of the previously made assumption that light
1 International Critical Tables, vol. I, p. 12. Mention should be made of
the fact that in the Ii tera ture of physics the charge on the electron is most
frequently expressed in electrostatic units, while the ratio of charge to mass is
usually given in electromagnetic units per gram. This usage has an historical
basis only.
For a fuller account of the determination of the electronic charge e the
reader is referred to MILLIKAN'S "The Electron." The accurate measure-
ment by Millikan of the electronic charge e, although by a method sub-
stantially the same in principle as Wilson's method described above, is one
of the outstanding investigations in modern physics, not only because of
the fundamental importance of the quantity measured, but also because
of the very skillful way in which the many experimental difficulties were
overcome.
SEC. 8] PHOTOELECTRONS 155
waves originate in the vibrations of electrons about fixed positions
1:n the atom, thus, seemingly, completing the chain of evidence in
favor of the electromagnetic theory of light and confirming the
hypothesis resulting from the Lorentz theory of dispersion
(Sec. 5, this chapter).
8. ''Photoelectrons."-We can now return to the question
raised at the end of Sec. 3 of this chapter: What are the carriers
of electricity in the photoelectric current? The discovery of the
electron seemed to furnish an answer: The photoelectric effect is
due to the liberation, from the illuminated metal plate, of electrons
which, under the influence of the electric field, pass from cathode to
anode, thereby causing the photoelectric current. This hypo-
thesis was confirmed by the experiments of Lenard 1 and of Mer-
ter z
c::e4rth :troJile
101..- 1o \ee,
/
.,,,,,,- - --- .......
± c Pz\ To Elecfro-
I \
" I I
13 mefer!
'""' \ I
"'
T "''",, ' ..........
- -- /
/
'
'
Q
"*s
Frn. 35.-Lenard's apparatus for determining the ratio e/m for photoelectrons.
ritt and Stewart,2 who proved independently that the photo-
electric discharge is deflected in a magnetic field, exactly as are
cathode rays. And Lenard, by measuring the deflection of the
'' photoelectric rays'' in a known magnetic field, found a value of
e' /m of (about) 1.2 X 107, in approximate agreement with
Thomson's value of e' /m for electrons. Later work by Alberti 3
gave a value of e'/m for photoelectrons of 1.765 X 10 7 e.m.u.
per gram.
Lenard's method of determining e/m for photoelectrons is
particularly instructive. The apparatus used is shown diagram-
matically in Fig. 35. A glass tube, which may be exhausted to
the highest attainable vacuum through the side tube T, contains
an aluminum cathode C, which may be illuminated by ultra-
violet light from a spark S, the light passing through the quartz
1 Ann. Physik, vol. 2, p. 359 (1900).
2
Phys. Rev., vol. 11, p. 230 (1900).
3 Ann. Physik, vol. 39, p. 1133 (1912).
156 THE PHOTOELECTRIC EFFECT [CHAP. VI
plate Q. C may be charged to any potential, positive or negative.
A screen A, which has a small hole at its center and which js
connected to earth, serves as anode. P1 and P2 are small metal
electrodes connected to electrometers. When C is illuminated
and charged to a negative potential of several volts, photoelec-
trons are liberated and accelerated toward the anode A. A few
electrons pass through the hole in the center of A and proceed,
thereafter at uniform velocity, to the electrode P1, their reception
there being indicated by the electrometer 1. But if, by means of
a pair of Helmholtz coils (represented by the dotted circle), a
magnetic field directed toward the reader is produced in the
region between A and P1, the electrons will be deflected upward
in a circular path and, with a sufficient field strength, will strike
the electrode P2.
l
-----------
5 5 10 15 20
+ Vo Pctential of Ccdhode in Vo!h
Fm. 36.-Variation of photoelectric current with variation of potential of
cathode.
Lenard first investigated the relation between the current
reaching the anode and the potential V applied to C. There was
no photoelectric current when V was several volts positive. But
when V was dropped to about 2 volts positivie, a small current was
observed. This indicated that the photoelectrons were not sim-
ply freed from the cathode but that some of them, at least, were
ejected with sufficient velocity to enable them to overcome the retarding
potential of 2 volts. The current increased when V was reduced
to zero and increased still more rapidly as V was made negative;
but attained a "saturation" value after V had reached some 15
or 20 volts, negative. These data are shown diagrammatically
in Fig. 36. I is the photoelectric current, and Vo is the positive
potential (about 2 volts) which it is necessary to apply to the
cathode to prevent the escape of any electrons.
The determination of e' / m was made essentially as follows:
Let a negative potential V, large compared to Vo, be applied
SEC. 9] PHOTOELECTRIC CURRENT 157
to the cathode. The photoelectron, on reaching the anode, wm
have a kinetic energy given by
Ve' = :,~mv 2 (27)
where e' and mare the charge and mass of the electron, and v its
velocity on reaching A (Fig. 35). Assuming that after leaving
A the electron moves in a uniform magnetic field, the circular
path which the electron follows is determined by the equation
2
He'v = mv (28)
R
where H is the strength of the magnetic field just necessary to
cause the electron to reach P2, and R is the radius of the cor-
responding circular path, determined from the geometry of the
apparatus. Eliminating v between equations (27) and (28), and
solving for e' /m, we have
(29)
The quantities V, H, and R being known, e' /m may be computed;
and v may then be determined from either equation (27) or
equation (28). In this way, Lenard measured the values of e' /m
for a wide range of values of V. Some of his results are shown in
the following table :
e'/m Velocity,
Vin volts e.m.u. centimeters
per gram per second
607 1.17 x 107 12 x 10 8
4,380 1.12 32
12,600 1.18 54
The ratio e' /m is observed 1 to be constant over a wide range
of potentials applied to the cathode. It is, also, to be noted that
the velocity attained by the electrons is, for the higher voltages,
a substantial fraction of the velocity of light.
9. Relation between Photoelectric Current and Intensity of
Illumination of the Cathode.-The experiments of Elster and
Geitel, 2 Lenard, 3 and Ladenburg 4 seemed to show that, so long
1 As pointed out above, more recent determinations give a higher value for
e' / m for photoelectrons.
2 Ann. Physik, vol. 48, p. 625 (1893).
3 Ann. Physik, vol. 8, p. 154 (1902).
4 Ann. Physik, vol. 12, p. 573 (1903).
158 THE PHOTOELECTRIC EFFECT [CHAP. VI
as there was no change in the spectral quality of the light causing
the emission of electrons, the photoelectric current was apparently
proportional to the intensity of illumination on the emitting
surface. But some experiments by Griffith 1 threw doubt on
this conclusion.
In 1909, the author reported measurements 2 on the relation
between the photoelectric current and light intensity, in which a
sodium surface in vacuum (the so-called "photoelectric cell") was
used with an electrometer as the current-measuring instrument
for low intensities (up to 0.5 foot-candles), and a sensitive
d' Arsonval galvanometer for high intensities (up to 600 foot-
candles). The results showed that both for low intensities and for
high intensities the photoelectric current is strictly proportional
to the intensity of illumination. This conclusion was confirmed
by the later experiments of Elster and Geitel, 3 who found the
linear relation to hold over a very wide range of intensities
(50,000,000: 1). Kunz and Stebbins found a similar proportion-
ality over the range of illuminations: 0.018 meter-candle to 1,110
meter-candles. This law of proportionality seems, therefore, to
be well established.
Now, each electron, photoelectrically emitted, carries a charge
e; and if the photoelectric current be 1, the number of electrons
emitted per second is I/ e. Since experiment shows that I is
strictly proportional to the intensity of illumination, it follows
that the rate of emission of photoelectrons is strictly proportional to
the intensity of illumination. 4 This is one of the very important
laws of the photoelectric effect.
10. Velocity Distribution Curves for Photoelectrons.-The
relation between photoelectric current and the potential applied
to the cathode, as in the experiments of Lenard (see Fig. 36),
has been the subject of many investigations 5 and has lead to
results of extraordinary importance. The principle of the
lPhil. Mag., vol. 14, p. 297 (1907).
2 Phys. Rev., vol. 29, pp. 71, 404 (1909).
a Phys. Zeitsch., vol. 14, p. 7 41 (1913); vol. 15, p. 610 (1914); and vol. 17,
p. 268 (1916).
4 The variation in photoelectric current with wave length of the incident
light, for a constant energy flux onto the emitting surface, has been the
subject of many investigations. For a discussion of this subject, the reader
is referred to ALLEN'S '' Photoelectricity."
s For the general method and technique see the experiments of RICHARD-
SON and COMPTON, Phil. Mag., vol. 24, p. 575 (1912).
SEC. 10] VELOCITY DISTRIBUTION CURVES 159
methods employed is illustrated diagrammatically in Fig. 37.
The emitting electrode C and the receiving electrode A are
enclosed in a highly evacuated bulb of glass or of glass with
quartz window, through which C may be illuminated. C is
connected to the slider S of a potentiometer wire, one point P
of which is connected to earth and is, therefore, at zero potential.
The potential of C may, thus, be made either positive or negative.
The receiving electrode A is connected to earth through a
sensitive current-measuring instrument G, usually an electro-
meter. In order to prevent an inverse photoelectric current
from A to C, due to reflected light reaching A when C is positive,
p
+
S -:;=- [c,rfh
Fm. 37.
A is frequently chosen of such material as to be photoelectrically
insensitive to the spectral region used to illuminate C. This
necessitates a correction for the "contact" potential difference
between C and A. In order to eliminate errors due to reflection
of electrons and to unsymmetrical emission of electrons from
C, Richardson and Compton placed Cat the center of a spherical
bulb silvered on the inside except for a small area through which
light might be admitted. This silvered surface served as
receiving electrode A.
After eliminating all spurious effects, and illuminating the
emitting electrode C with strictly monochromatic light, the intensity
E of which can be varied in a known manner, the relation
between the photoelectric current I and the potential V is as
shown in Fig. 38. Thus, let the intensity of illumination be
E1. No photoelectric current is observed when the positive
potential is greater than a critical potential +Va. At potentials
just less than + Vo, a small current is observed, and as the poten-
tial decreases from + Vo to zero, the current rises rapidly to a
160 THE PHOTOELECTRIC EFFECT [CHAP. VI
maximum value I 1, which it reaches when V = 0. No further
increase in I is observed when V becomes negative. I 1 is, therefore,
the maximum current resulting from illumination (at this par-
ticular wave length) of intensity E1. If the intensity of illumi-
nation be doubled, i.e., E2, a curve of the form 2 is obtained.
The maximum current I 2 is exactly double the current I 1, but the
critical potential Vo is exactly the same as for the illumination E 1.
Increasing the intensity of illumination to E3 or E4 increases the
corresponding maximum current proportionately but causes no
change whatever in V 0 • Calling this maximum current Im for a
+ Po+enti cd Von. Emitting Electrode
Frn. 38.-Variation, with potential, of the photoelectric current from a ca tho de
when illuminated by different intensities E1, E2 . . . of monochromatic
radiation. ·
given intensity of illumination E, we have, as shown in Sec. 9 of
this chapter,
(30)
Indeed, if the ordinates of curves 2, 3, and 4 be divided by 2,
3, and 4, respectively, the resulting curves would coincide with
curve 1.
These facts are readily interpreted on the assumption that
the electrons are emitted from the surface of the electrode with
initial velocities v varying from a certain maximum velocity
Vm, corresponding to the critical potential Vo, down to zero veloc-
ity. For, if v is the velocity with which an electron, mass m, is
emitted, its kinetic energy, therefore, being 72 mv 2 , it is obvious
SEC. 101 VELOCITY DISTRIBUTION CURVES 161
that only those electrons will reach the receiving electrode A
(Fig. 37) for which 1
}2 mv2 > Ve (31)
where Vis the potential difference between the plates C and A,
and e is the charge on the electron. Electrons emitted with
smaller velocities do not possess sufficient energy to reach A, and
they, therefore, "fall" back to the plate C, just as a ball, thrown
upward from the surface of the earth, is pulled back to the earth
by the force of gravity. Accordingly, if the current is I 1 (Fig.
39) when the potential difference is V 1, Ii/ e electrons (per second)
I --- k
I Im
r----Ii
I
I
I
I
+ v
Frn. 39.
leave the plate C with velocities in excess of v1, where v1 is deter-
mined by V 1e = ~~mv 1 2 • If the potential be reduced, by a small
amount Ll V, to V2, the current increases, by a small amount Lll,
to 12. Therefore, Lll/e electrons per second leave the plate with
velocities between the limits v1 and v2, which are defined by
Vie = ~~ mvi
V2e = }2 mv~
In other words, the slope of the curve at any point, in Fig. 39,
is proportional to the number of electrons possessing energies
corresponding to the value of V at that point. Since, when the
applied potential V m > Vo, no current is observed, the maximum
velocity Vm possessed by any photoelectron is given by
Y2 mv~ = Voe (32)
1Strictly speaking v in equation (31) is not the actual velocity of the
photoelectron, but is the component of the velocity in the direction of the
electric field. In the above mentioned experiments of Richardson and
Compton (small spherical emittor at the center of a comparatively large
spherical receiving electrode) v is very nearly the actual velocity.
162 THE PHOTOELECTRIC EFFECT [CHAP. VI
When V = 0, any electron freed with any velocity greater than
zero reaches plate A, and, since for small negative (i.e., acceler-
ating) values of V, there is no further appreciable increase in the
photoelectric current, it follows that small electric fields play no
appreciable part in initially freeing the electrons from the plate. 1
And the total number of electrons (per second) freed from the
plate C is given by I m/e, where Im is the photoelectric current for
any negative value of V. Curves of the kind shown in Figs. 38
and 39 are sometimes referred to as "velocity-distribution
curves," although, correctly speaking, the velocity-distribution
curves would be those in which the slopes of these curves are
plotted as ordinates against the velocity v, determined from
Ve = ~~mv 2, as abscissre.
11. Relation between the Velocities of Photoelectrons and the
Frequency of the Light.-If the photoelectric current as a func-
Im
I
:
•t
I
•I
•
I
I
•I
+ T§ v
Fw. 40.-Dependence of photoelectric current on wave lengths of the exciting
light: A1 > A2 > Aa
tion of applied voltage V be determined for several different
monochromatic radiations of wave length X1, X2, Xa, etc., falling
on the emitting electrode, curves of the form shown in Fig. 40
result, in which the intensities of illumination for each wave
length have been so adjusted as to give the same value of Im in
each case. Here X1 > X2 > A3, but V1 < V2 < V3, where V1,
V2, and V 3 are the corresponding critical retarding voltages.
This means that the shorter the wave length the greater is the
1Of course, if Vis positive, it may cause the return to the electrode C of
some or all of the photoelectrons. And if V is negative, it may accelerate
them toward the plate A so that they reach A with a greater velocity than
that with which they left the plate. If V has a large negative value
secondary effects are produced which complicate the observed phenomena.
SEC. 11] VELOCITY AND LIGHT FREQUENCY 163
retarding voltage Vo just necessary to prevent the escape of the
"fastest" electron. Or the maximum kinetic energy 1 of photo-
electrons, given by p~mv2 = Voe, increases with increasing fre-
quency v of the light which causes their emission.
A very simple linear relation has been found to exist between
this maximum energy of emission and the frequency of the light.
If, as shown by Millikan, a curve be plotted between V 0e and the
corresponding frequency v of the light, ~ ~
there results the straight line, shown t
in Fig. 41, which has an intercept vo. "'f~
The experimental meaning of this ~~
intercept is that light of frequency ~
~
less than vo cannot cause the emission i
of photoelectrons from the metal con- _V. . . . --Fi-re_9_u-en_c_:1_J_:v_ _
0
cerned. vo is characteristic of the Fw. 4 1.-Relation of max-
emi tting electrode, but the slope of the imum energy of photoelectrons
to frequency of the exciting
curve is the same for all electrodes. T h e radiation.
equation of the curve may be written:
Voe = }fmv! = h(v - vo) = hv - hvo
or
~2mv! = hv - wo (33)
where his the slope of the curve and wo is written for hvo.
Now, since the dimensions of the left-hand side of equation
(33) are those of energy, both terms on the right-hand side must,
likewise, stand for quantities of energy. Since h, the slope of the
curve, is independent of the nature of the emitting electrode, the
quantity of energy hv depends only on the frequency of the incident
light. On the contrary, wo is a quantity of energy characteristic
of the emittor and is independent of the light. It is, therefore,
natural to interpret equation (33) by saying the, hv is the energy
which the light of frequency v gives to the electron and that wo is
The careful work of Millikan (Phys. Rev., vol. 7, p. 355 (1916)) showed
1
that if the radiation be truly monochromatic, at least if it contained no
wave lengths shorter than the wave length)\ for which a given curve is being
determined, the curve approaches the voltage axis at a finite angle, and not
asymptotically. Failure to exclude stray light of short wave length is
apparently the explanation of the contradictory result of Ramsauer (Ann.
Physik, vol. 45, pp. 961 and 1121 (1914)), who, by deflecting the photo-
electrons in a magnetic field, somewhat after the manner of Lenard's experi-
ments (p. 155), thereby producing a "magnetic spectrum" of velocities,
made a direct determination of the distribution of velocities. The reader
will find Ramsauer's method of procedure instructive.
164 THE PHOTOELECTRIC EFFECT [CHAP. VI
the part of this energy which the electron expends in escaping
from the emittor; the remainder, ~~mv!, is the kinetic energy
with which the electron is actually observed 1 to leave the emittor.
If, in Fig. 41 (and equation (33)), the ordinates be expressed
in ergs and the abscissre in vibrations per second, the numerical
value of his found to be 2 6.56 X 10- 27 • This constant h, called,
for reasons to be explained later, "Planck's constant," is one of
the fundamental constants of nature and has played an extra-
ordinary part in modern physics. The product hv is called a
quantum of energy corresponding to light of frequency v.
At first thought, it might seem that the above discussion
leading to equation (33) applied only to those photoelectrons
which possess the maximum velocity Vm, i.e., to those electrons
corresponding to the critical retarding voltage Vo, and that a
lesser quantity of energy than hv is given to those electrons
which are observed to possess smaller energies. But, as we have
pointed out above, the observed velocities are those possessed by
the electrons after they have left the emittor and, for the swiftest
electrons, it requires an amount of energy wo to" free the electron"
from the surface of the emittor. It is, therefore, just as logical to
assume that those electrons which are observed to possess veloci-
ties smaller than Vm have required energies larger than wo to free
them from the emittor and that each electron initially receives, as
a result of the photoelectric process, an amount of energy hv. Indeed,
this latter view is much the more logical, since we can readily
account for this increased amount of energy required to escape
from the metal by assuming that light penetrates a finite distance
into the metal, through many surface layers of atoms, and that, in
general, the photoelectrons originating beneath the surface lose
some of their energy by collisions before they reach the surface.
Only those which originate right at the surface escape with the
maximum energy hv - w 0• We may assume, therefore, that the
equation
}~mv:! = hv - wo (33)
is generally applicable to all photoelectrons and that, were it not
for these accidental collisions, before reaching the surface of the
emittor, every photoelectron would possess the same velocity Vm.
This equation (33) has had a very interesting history and is
one of the most fundamental equations, or laws, of modern
1 See note page 161.
2 MILLIKAN: "The Electron."
SEC. 12] ORIGIN OF PHOTOELECTRONS 165
physics. It is frequently referred to as "Einstein's photoelectric
equation," since it was first proposed by Einstein,1 in 1905, as a
result of the extension to the photoelectric process of the concept
previously developed by Planck (see Chap. VII), that inter-
changes of energy between radiation and matter take place in
energy quanta hv, where v is the frequency of the radiation
absorbed or emitted and h is a constant. Einstein had at his
disposal only qualitative data to show that his equation gave
results of at least the right order of magnitude. Two years
later, Joffe 2 pointed out that some data, previously reported
by Ladenburg 3 as indicating that the velocity of the photo-
electrons is proportional to frequency, were in at least as good
agreement with Einstein's equation, which states that the square
of the velocity is proportional to (rather, a linear function of)
the frequency. From these data of Ladenburg, Joffe computed
that the constant h in Einstein's equation was about 3 X 10-27
erg X sec. This was one of the very first photoelectric deter-
minations of h. The equation received final and complete exper-
imental verification as a result of the precision experiments of
Millikan, 4 to which reference has been previously made. More
recent work has extended the validity of the equation to the X-ray
region, where frequencies are involved which are several thousand
times the frequencies of the visible and near-ultra-violet light for
which Millikan's verification was made. Indeed, Einstein's
photoelectric equation is, in many respects, to the modern
quantum theory what Newton's second law of motion, i.e.,
F = ma, is to mechanics. But in spite of its generality and of
the many successful applications which have been made of it in
recent theories in physics, the equation is, as we shall see pres-
ently, based on a concept of radiation-the concept of "light
quanta ''-so completely at variance with the most fundamental
postulates and conclusions of the electromagnetic theory of radia-
tion that all attempts at reconciliation between the two theories
have, so far, failed. We shall return to this question in Sec. 15
of this chapter.
12. Origin of the Photoelectrons.-If the phenomenon of the
conduction of electricity in metals be ascribed to the presence in
1 Ann. Physik, vol. 17, p. 132 (1905).
2 Ann. Physik, vol. 24, p. 939 (1907).
3 Phys. Zeitschrift, vol. 8, p. 590 (1907).
4 Phys. Rev., vol. 7, p. 355 (1916).
166 THE PHOTOELECTRIC EFFECT [CHAP. VI
the interatomic spaces of the so-called "free" electrons, then we
may recognize two classes of electrons in metals: (1) these "free"
electrons and (2) the "bound" electrons, i.e., those which are
attached to, or are a part of, the atoms of the metal. The
question then arises: Do the photoelectrons come from the free
electrons or from the bound electrons?
One line of attack on this question is the influence of tempera-
ture on the photoelectric process. As the temperature of a
(solid) substance rises, the energy of agitation of its constituent
molecules or atoms increases proportionally to the absolute tem-
perature; and, correspondingly, the energy of translation of the
electrons in the, '' free electron atmosphere'' increases. This
increase in translational velocity is the basis of the explana-
tion of the increase in electrical resistance of metals with increas-
ing temperature. If the temperature be raised sufficiently high,
the translational energy of some of these free electrons may
become sufficient for them to escape from the surface of the
metal, in much the same way as molecules of a liquid escape
through the surface of the liquid in the process of evaporation;
and we then have the well-known thermionic emission of elec-
trons. Now, if the photoelectrons come from the free electrons
in the metal, we should expect that either the number of the
photoelectrons or their velocities, or both, would depend, in
some way, on temperature. For example, if the kinetic energy
of the free electrons increases with rise of temperature, as both
the variation of resistance with temperature and thermionic
emission seem to indicate, it should be easier for the light to
cause the ejection of photoelectrons at higher temperatures than
at lower, and light of a given intensity might cause the emission
of more electrons at the higher temperature.
On the contrary, if the photoelectrons come from the atoms of
the metal, one might expect that the photoelectric process
should be practically independent of temperature, since there is
no reason for believing that the internal energy of an atom
depends on temperature, at least until comparatively high tem-
peratures be reached.
There have been numerous investigations of the effect of
temperature on photoelectric phenomena, at first with conflicting
results. Thus, Hoor 1 found that the photoelectric effect in zinc
1 Wien Berichte, vol. 97, p. 719 (1888).
SEC. 12] ORIGIN OF PHOTOELECTRONS 167
was less at 55°C. than at 18°C., while Stoletow 1 found that the
activity of platinum increased when it was heated to 200°C.
Elster and Geitel 2 found that the photoelectric discharge from
zinc was independent of temperature but that in the case of
potassium, there was an increase in activity up to 60°C.; while
Zeleny 3 found, with platinum, first a decrease to a minimum and
a subsequent increase as the temperature was raised to several
hundred degrees centigrade, where the true photoelectric effect
began to be masked by thermionic emission.
These early experiments were probably vitiated by the second-
ary effect of gases or of electropositive impurities in changing the
condition of the photosensitive surface. For, contrary to the
above cited results of Elster and Geitel on potassium, Dember 4
and Kunz 5 found no change of the photoelectric effect with the
alkali metals when they were in a very high vacuum. This
result is in entire agreement with the experiments of Ladenburg 6
and of Millikan and Winchester.7 The latter investigators
worked with many different metals as emittors (copper, gold,
iron, lead, aluminum, etc.) over a temperature range from 25°C.
to 125°C. Neither total photoelectric current nor velocity of
emission was found to depend on temperature within the limits
of accuracy of the experiments-I or 2 per cent.
Some recent experiments, particularly at low temperatures,
indicate that the photoelectric current does depend on temper-
ature, 8 but the most reasonable interpretation of these results
seems to be that temperature change produces an alteration in
the surface of the metal so as to change the amount of work wo
necessary for the escape of the electron from the metal but that
there is no change in the amount of energy hv which the electron
initially receives.
This evidence points clearly to the conclusion that, when all
spurious effects are eliminated, the photoelectric process, both as
to number of photoelectrons and their velocities, is independent
1 Comp. rend., vol. 108, p. 1241 (1889).
2
Ann. Physik, vol. 48, p. 634 (1893).
3 Phys. Rev., vol. 12, p. 321 (1901).
4 Ann. der Physik, vol. 23, p. 957 (1907).
5 Phys. Rev., vol. 29, p. 174 (1909).
6 Verhandlungen der Deutsch. Phys. Gesell., vol. 9, p. 165 (1907).
7 Phil. Mag., vol. 14, p. 188 (1907).
8 See the experiments of IvEs, Jour. Opt. Soc. Amer. and Rev. Sci. Inst.,
vol. 8 p. 551 (1924).
168 THE PHOTOELECTRIC EFFECT [CHAP. VI
of temperature and that, therefore, the photoelectrons come from the
atoms of the emittor rather than from the free electrons.
This conclusion receives confirmation in the beautiful experi-
ments of Robinson 1 on the photoelectric effect of X-rays (see
Chap. XII), which cause the ejection from an emittor of several
groups of photoelectrons, each group having a different maximum
velocity Vm- a circumstance explainable only if the electrons
originate at different" depths" within the atom, thereby requiring
different quantities of energy to escape from the atom.
Further, it can be shown that, whatever the nature of the
photoelectric process, the laws of the conservation of energy and
of momentum require that the photoelectron should originate in
an atom, i.e., that a free electron cannot be given a quantum of
radiation.
We must, therefore, base any explanation or theory of the
photoelectric process on the experimental conclusion that the
atom, by a mechanism at present quite unknown, but probably
as a result of absorption of energy from light of frequency v,
causes the emission of one (or more) of its electrons with an
energy hv, part of which may be used to escape from the atom
itself and part from the surface of the emi ttor, the remainder
appearing as kinetic energy of the photoelectron.
13. Source of the Photoelectric Energy.-(a) In the last
paragraph of the preceding section, we stated that probably the
atom absorbs from the incident light the energy used in expelling
a photoelectron. Let us examine this question in more detail.
The following is a summary of the pertinent experimental conclu-
sions, previously discussed:
1. For a given spectral composition of the incident light, the
number of photoelectrons expelled per second is strictly
proportional to the intensity of the light, over a very wide
range of intensities. Thus, let n be the number of elec-
trons emitted per square centimeter per second by light of
intensity I. Then,
n = kl (34)
where k is a constant, which depends on the frequency of
the light and the nature of the emittor.
1 Proc. Roy. Soc. (London), vol. 104, p. 455 (1923); and Phil. Mag., vol. 53.
p. 241 (1925).
SEC. 13] PHOTOELECTRIC ENERGY 169
2. The kinetic energy of the photoelectrons is a linear func-
tion of the frequency of the light, the relation between the
two being given by the Einstein photoelectric equation:
3,~mv 2 = hv - wo
And, with the interpretation of wo as the amount of work
required to escape from the emittor, it follows that each
electron receives, initially, an amount of energy, hv, which
is strictly proportional to the frequency of the incident
light and is independent of the nature of the emittor.
3. The kinetic energy of the photoelectrons, 3,~mv!, emitted
by monochromatic light has been found to be rigorously
independent of the intensity of the light over a very wide
range of intensities.
From experimental conclusion 1, it follows, at once, that the rate
dEP/ dt at which energy is given to photoelectrons per square centi-
meter of emitting surface is strictly proportional to the intensity
I of the light. That is,
dEP = k·I (35)
dt
Now, when light is incident onto a metallic surface, it ~s partly
reflected and partly absorbed; and the rate dE A/dt at which energy
is absorbed, per square centimeter, is strictly proportional to the
intensity of the light, i.e.,
dEA
ex I (36)
dt
and, therefore,
dEP dEA
- a; (37)
dt dt
The presumption is very strong, therefore, that the energy
acquired by the photoelectrons is obtained from the incident
light. In fact, the initial energy acquired by the photoelectrons
seems to be determined solely by the light and not at all by the
emittor.
(b) At first thought, it might seem (1) that this conclusion
is inconsistent with the fact that the kinetic energy of each
photoelectron is proportional to the frequency v and (2) that
therefore the photoelectric process might be a resonance phenom-
170 THE PHOTOELECTRIC EFFECT [CHAP. VI
enon whereby the light functions simply as a "trigger," releas-
ing the photoelectron with an amount of energy which it already
possessed in the atom. To illustrate by analogy:
One can readily imagine an electromechanical device in which
a metal diaphram is set into vibration, when sound of a particular
frequency strikes it, with sufficient amplitude to close an electric
circuit, which, in turn, pulls a trigger which discharges a rifle.
In this case, the energy with which the bullet leaves the rifle
comes from the powder and does not depend on the intensity of
the sound. This coincides with experimental conclusion 3,
above. One can, then, imagine a series of such mechanisms,
each tuned to a different frequency, in which the amount of
powder in each rifle is so adjusted that the energy with which
the bullet leaves the rifle is proportional to the frequency of
the sound which sets it off. This would coincide with experi-
mental conclusion 2, above. Applying this "trigger" theory to
the photoelectric process, we should require that each photo-
electric emittor contained resonating systems of all frequencies,
from the frequency of the long wave length photoelectric limit
to (theoretically) infinite frequencies, and that each of these
resonators, when "set off" by light of appropriate frequency,
should expel an electron with an energy proportional to the
frequency. All this involves quite too much of an "ad hoc''
assumption. And even if the resonance theory were tenable,
we should then have to explain ('see experimental conclusion 1,
above) why the number of the photoelectrons is strictly propor-
tional to the intensity of the light. (In our analogy, it does
not follow that the number of rifles discharged per second is
proportional to the intensity of the sound.)
The facts, therefore, seem to demand the conclusion that the
photoelectric energy comes directly from the light and not from
energy previous stored up in the atom and released by a reso-
nance, or "trigger," effect.
14. What Is the Photoelectric Mechanism?-If the photo-
electrons come from the atom, and if the energy necessary for the
expulsion of the electrons comes from the incident light, it must
be that the atom absorbs the energy from the light and gives this
energy to the photoelectron. But how can we picture an atomic
mechanism which will function in accordance with the experi-
mental facts? The difficulties in the way of postulating such a
mechanism, particularly on the basis of the classical electro-
SEC. 14] THE PHOTOELECTRIC MECHANISM 171
magnetic wave theory of light, can best be shown by some
numerical data:
1. A photoelectric current of 1 X 10-12 e.m.u. per square
centimeter is easily obtained from a sodium surface in vacuo
with very moderate intensity of illumination. Taking the
value of the charge on the electron as 1.592 X 10-20 e.m.u.,
this current corresponds to the emission of 6.28 X 10 7 electrons
per square centimeter per second. These electrons come from
the surface layers of atoms. And we can easily compute the
order of magnitude of the number of atoms involved. For,
calling Avogadro's number N (i.e., the number of atoms per gram-
atom), the atomic weight of the emittor A, and its density o, the
number of atoms per unit volume n is given by
N
n = -o
A
N = 6.06 X 1023 ; and for sodium, A = 23.0 and o = 0.97
grams per cubic centimeter. Therefore, n = 25.5 X 10 21 atoms
per cubic centimeter. Assuming the atoms in a centimeter-cube
of sodium to be arranged in regular rows, columns, and layers, 1
one finds that there are -v1'25.5 X 10 21 = 2.95 X 10 7 atoms
along each edge, or (2.95 X 10 7 ) 2 = 8.7 X 10 14 atoms in a layer
1 sq. cm. in area and 1 atom deep. Assuming, for the sake of
computation, that the photoelectrons are supplied by, say, the
first 10 layers of atoms, we find that 8.7 X 10 15 atoms furnish,
under the conditions of illumination above mentioned, 6.26 X 10 7
. . 8.7 x 10 15 1
e1ectrons per second . Th at 1s, 1 atom m _ X 107 = .38 X
6 28
10 8 , on the average, furnishes 1 photoelectron per second. But
what is it which determines which one of these hundred million
odd atoms shall furnish an electron in any particular second?
According to the wave theory of light, radiant energy should be
distributed continuously over the wave front. Save for differ-
ences in thermal agitation, we know of no reason to expect that
one atom differs from another in any way which might result in
one atom having at any instant a greater absorbing power
than another. And thermal agitation cannot be responsible for
differences among atoms if the photoelectric effect be independ-
ent of temperature. Apparently, we are unable, on the basis of
1 The crystal structure of sodium is really a body-centered cube. See
International Critical Tables, vol. I, p. 340.
172 THE PHOTOELECTRIC EFFECT [CHAP. VI
the classical theory of radiation, to put forward any hypotheses
which should result in singling out the one atom in a hundred
million.
2. But the difficulties in which the classical theory finds itself
in attempting to explain the experimentally well-established
facts of the photoelectric phenomenon are, perhaps, even more
clearly brought out by an attempt to compute how long, on the
classical theory, it should take an atom to absorb enough energy
to emit a photoelectron. An intensity of illumination of 0.1
meter-candle, falling onto a sodium surface, produces an easily
measurable photoelectric current. This illumination corresponds
to an energy flu4 of several ergs (say, ten 1) per square centimeter
per second, depending on the light source. Only a little over 1
per cent of this energy falls within the limit of the visible spec-
trum, and a much smaller proportion, certainly not over 0.05
erg per square centimeter per second, falls within the spectral
range effective in expelling photoelectrons from sodium, which is
sensitive in the blue and near-ultra-violet part of the spectrum.
Sodium has a high reflecting power, and, since only absorbed
energy can possibly be used in expelling electrons, we conclude
that not over 0.001 erg per square centimeter per second is
available for the photoelectric process.
According to the electromagnetic theory, this absorbed energy,
0.001 erg per square centimeter per second, should be equally dis-
tributed among the 8.7 X 10 15 atoms which form the (10, arbi-
trarily assumed) surface layers. That is, each atom should
absorb
X 1015 -- 1. 1 x 10-19 ergs per secon d
_0.001
87
Now, a quantum of light corresponding to X = 0.00004 cm.
(violet light) is
hv = 6.55 X 10-27 X 0.75 X 10 15
= 5 X 10- 12 ergs
This, therefore, is the energy which the atom gives to the electron
as a result of the photoelectric process involving light of this
wave length. If the atom obtains this energy by absorption
from the incident light, the number of seconds required for the
1 Assuming tungsten light and taking the efficiency of the 40-watt vacuum
tungsten lamp as 10 lumens per watt. (See CADY and DATES: "Illuminat-
ing Engineering," John Wiley & Sons, Inc., 1925.)
SEC. 15] CORPUSCULAR THEORi OF LIGHT 173
atom to obtain enough energy to expel an electron would be given
by
5 x 10- 12
l. l X _ 19 = 4.5 X 10 7 seconds
10
This is over 500 days!
In other words, if a sodium surface be illuminated by a tungsten
lamp, the intensity of illumination being 0.1 meter-candle, it
should require several hundred days of continuous illumination
before a photoelectric current starts to flow, if we assume
1. that the photoelectron comes from the atom;
2. that the atom absorbs from the incident light the energy
required to expel the electron;
3. that each atom in the surface layers of the photoactive
material absorbs only it& own proportionate quota of
energy-as seems to be required by the wave theory of
light.
But there has never been observed any time lag between the
beginning of illumination and the starting of the photo-electric
current. Indeed, some recent measurements 1 show that, if there
be such a time lag, it is less than three billionths of a second.
Clearly, one (or more) of these three assumptions just stated
must be wrong, since the experimental facts and the numerical
data on the basis of which the computation is made seem to be
well founded.
The classical electromagnetic theory of light, therefore, not
only gives us no clue as to the mechanism by which photoelectric
emission occurs, but, also, it seems to be in direct conflict with
well-established experimental facts, which facts point unmis-
takably to some kind of discontinuity or "structure" to what.
heretofore, we have called a" wave front."
16. The Photoelectric Effect and the Corpuscular Theory of
Light.-If we could dismiss the phenomenon of interference of
light and the deductions therefrom, both experimental and
theoretical, and if, with Newton, we could return to the cor-
puscular theory, the difficulties raised in the preceding section
could be easily explained, at least qualitatively. Thus, on the
corpuscular theory, we might regard a beam of light as a "rain"
of corpuscles of energy. When such a ''shower" of light falls
on the sodium surface, only here and there an atom is being
1 LAWRENCE and BEAMS: Phys. Rev., vol. 29, p. 903 (1927).
174 THE PHOTOELECTRIC EFFECT [CHAP. VI
struck by a corpuscle at any given instant. If, now, we assume
(1) that each corpuscle, or "quantum" in modern terms, pos-
sesses energy hv and (2) that, by some unknown process, a "colli-
sion" between a quantum and an atom may, under suitable
circumstances likewise quite unknown, result in the absorption
of the whole quantum by the atom and the subsequent emission
of the photoelectron with the initial energy hv, we can, at once,
explain why only one atom in many millions expels an electron
in any particular second and, also, why there is no time lag in
the photoelectric process.
Such a theory would predict, also, that the photoelectric cur-
rent should be proportional to light intensity, as is experimen-
tally observed. For, the intensity of the light should be
proportional to the number of quanta falling per second on each
square centimeter of the phothelectric emittor; and, because of
the large number of quanta and of atoms involved, the proportion
of collisions between quanta and atoms which result in the
emission of an electron should be constant.
Further, we should have a ready explanation of the experi-
mentally observed facts: (1) that the initial velocity of the
photoelectrons is independent of the intensity of the light and (2)
that this velocity depends only on the wave length of the light.
For, keeping the wave length of the incident light constant and
increasing the intensity would increase the number of quanta
striking each square centimeter each second and, proportionately,
the number of collisions resulting in photoelectric emission; but
there would be no change in the nature of each collision: Each
photoelectron would, at the higher intensity of illumination,
receive exactly the same energy, hv, as at the lower intensity.
If the frequency of the incident light be increased, however, the
value of hv is increased, and, accordingly, the energy received by
each photoelectron would be increased. The Einstein photo-
electric equation
~~mv 2 = hv - wo
follows, at once.
The difficulties with such a radical theory of light are many.
For example, if we regard light, incident onto a surface, as a
''shower" of corpuscles or quanta of energy, what can possibly
be the meaning of frequency in connection with such a phenome-
non? There is nothing periodic about a falling raindrop, unless
we think of the drop as in revolution about an axis through its
SEC. 15] CORPUSCULAR THEORY OF LIGHT 175
center of gravity as it falls, the, speed of this angular rotation
determining the frequency, and the distance fallen during 1
revolution being a ''wave length.'' Nevertheless, frequency
plays a very fundamental role in determining the energy hv of the
quantum. It should be pointed out, however, parenthetically,
that we do not measure directly the frequency of a light ray.
We measure (1) the velocity c of the light and (2) its wave length
(X), on the assumption that light is a wave motion, and then we
compute the frequency v by the equation
c = Xv
We have to rely on the wave theory of light to give us the energy value
hv of a quantum!
And there still remains the phenomenon of interference, which,
since its discovery by Young, in 1802, has defied explanation on
any other basis than by assuming light to be a wave motion.
However, the experimental facts of photoelectricity are equally
as cogent as is the phenomenon of interference, and they cannot
be explained on the basis of the present wave theory of light. Here,
then, is the most perplexing question of modern physics, a ques-
tion which physicists have been trying to answer for 23,,2
centuries: Is light undulatory or corpuscular? In spite of the
vast amount of data bearing on the subject, we, today, are
apparently no more able to give a categorical answer to this
question than were Huyghens and Newton.
Several compromise, or combination, theories have been
suggested. For example, J. J. Thomson 1 proposed what has since
come to be called the "ether string" theory, according to which,
radiation, leaving the source, proceeds as a wave motion along
certain localized lines or '' ether strings," much as sound waves
might proceed from a source of sound along speaking tubes
instead of as spherical waves of even growing radii. According
to this concept, a wave front of light, if sufficiently magnified,
would appear as a group of bright specks-the cross-section of
these stream lines in the ether-on a dark background, rather
than as a continuously luminous surface. If it be assumed that
energy is streaming along each string at exactly the same rate,
then the number of such strings per square centimeter of wave
front determines the intensity of the radiation. These strings,
each equally intense, would'' strike'' only here and there an atom,
1 Proc. Cambridge Phil. Soc., vol. 14, p. 417 (1906).
176 THE PHOTOELECTRIC EFFECT [CHAP. VI
and, when photoelectric emission takes place, each electron
would be given the same quantity of energy. Thus, many
of the observed facts of the photoelectric phenomenon could
be accounted for. But the theory was too much of the ad hoc
kind to be either convincing or fruitful in suggesting new
researches, and it is now, probably, of historical interest only.
In short, the situation is about like this: On one side of an
impenetrable barrier, or fence, is to be found a group of phe-
nomena-such as interference, polarization, smaller velocity of
light in optically denser bodies, and, indeed, the whole electro-
magnetic theory and its ramifications-according to which we
should say, without the slightest doubt, that light must be a wave
motion. On the other side of the fence is to be found another
group-the photoelectric effect and other phenomena which we
shall consider in subsequent chapters-according to which we
should say, again without the slightest hesitation, if we did not
know what was on the first side of the fence, that light must be
corpuscular. Just as the discovery of polarization, in 1811, led
to a radical modification of the wave theory-namely, the sub-
stitution of the concept of transverse for longitudinal vibrations
-so, today, the discovery of some new experimental fact or
phenomenon may lead to a fusion of the undulatory and the
corpuscular theories in some quite unexpected way or, indeed,
may result in an entirely new theory. However that may be,
the concept of wave motion, as a description of certain of the
phenomena of radiation, has become an indispensible and a very
real part of modern physics. And, likewise, the concept of quanta
of radiant energy, as a description of certain other phenomena of
radiation, has become equally real and indispensible. Either
concept is used at will wheneyer the particular phenomenon at
hand happens to yield to discussion or treatment more readily in
terms of one concept than of the other. Ample confirmation of
this will be found in the following pages.
CHAPTER VII
THE ORIGIN OF THE QUANTUM THEORY
The quantum theory, first proposed by Planck, in 1900, arose
out of the inability of the classical physics to explain the experi-
mentally observed distribution of energy in the spectrum of a
black body. We have seen, in Chap. VI, how, likewise, the older
theories of radiation could not explain the experimentally
observed facts of the photoelectric effect. In the present chap-
ter, we shall first discuss certain observed phenomena of radiation
which any theory must explain. We shall then attempt to show
(1) just how far the problem of black-body radiation can be
solved by classical methods based first upon thermodynamics
and then upon statistical mechanics and (2) at precisely what
point each of these classical methods failed and the introduction
of the concept of radiation quanta seemed to offer the only
solution to the problem.
1. Thermal Radiation.-It is a matter of common observation
that bodies when heated emit radiant energy-or, more simply,
radiation-the quantity and quality of which depend, for any
given body, on the temperature of that body. Thus, the rate at
whiflh an incandescent lamp filament emits radiation increases
rapidly with increase in the temperature of the filament; and the
quality of the radiation, as observed visually, changes markedly
as the temperature increases-the emitted light being "whiter"
at higher temperatures. Radiation, the quantity and quality
of which, emitted by any given body, depends solely on the
temperature of that body, is called thermal radiation. It is a
characteristic of thermal radiation that, when dispersed by a
prism or other similar means, a continuous spectrum is formed.
In order that thermal radiation may become visually observ-
able, it is necessary that the temperature of the radiator should
be (about) 500 to 550°C. or above. But that bodies radiate at a
much lower temperature is easily observed by the feeling of
warmth experienced when holding near the skin a heated body,
say an iron ball heated a comparatively few degrees above body
177
178 QUANTUM THEORY [CHAP. VII
temperature. Indeed, as early as 1792, Prevost proposed his
famous "theory of exchanges," which states that there is a con-
tinuous interchange of energy between bodies as a result of the
reciprocal processes of radiation and absorption. Thus, if we
have a system of bodies in an enclosure, all at uniform temper-
ature, each body receives (i.e., absorbs) from its surroundings as
much energy as it radiates. According to this concept, which
agrees well with observation, any given body would cease to emit
thermal radiation only when its temperature has been reduced to
absolute zero.
Thermal radiation is emitted, ordinarily, only by solids or
liquids.
There are many types of radiation other than thermal radi-
ation. Thus, gases and vapors, when suitably "excited," emit
a characteristic radiation, which, when dispersed, results in a line,
or discontinuous, spectrum, the wave length of the several lines
being characteristic of the emitting substance. The excitation
may result from thermal agitation, electric discharge, bombard-
ment by electrons, or absorption of incident radiation of suitable
wave length. As examples may be mentioned the sodium
flame, the mercury arc, and the luminescence of sodium vapor
when illuminated by light of the D lines.
Characteristic radiation, in the X-ray region of the spectrum,
is emitted also by solids and liquids when bombarded by elec-
trons of suitable speeds. Such bombardment results, in addition,
in the emission of a continuous X-ray spectrum.
Certain solids and liquids emit a characteristic radiation when
illuminated (or excited) by light of suitable wave length, even
though the intensity of the incident light is not sufficient to pro-
duce a perceptible rise of temperature of the emittor. This
phenomenon is called fluorescence if the emission ceases as soon as
the exciting light is removed and phosphorescence if the emission
persists an apprecible time after removing the excitation. There
is no sharp dividing line between fluorescence and phosphor-
escence. In general, the exciting light must be of shorter wave
length than the fluorescent (or phosphorescent) radiation. 1
Radiation of very high frequency, known as "gamma rays,"
is emitted during the process of the radioactive disintegration of
the heavier elements.
1 For a description of the phenomena of fluorescence and phosphorescence,
see R. W. Woon: "Physical Optics."
SEC. 2] SOME FUNDAMENTAL CONCEPTS 179
The classification of radiation into the several types-thermal,
characteristic, fluorescent, etc.-is based, of course, upon super-
ficial peculiarities in laboratory methods of production and study,
rather than upon any real dijferences in the nature of the ultimate
emitting mechanisms involved in the several cases. What these
differences are, if any, we are not now in a position to state. As a
matter off act, considerable progress in the study of the problems
of radiation can be made without inquiring either into the atomic
mechanism involved or into the nature of radiation itself.
Indeed, in this chapter, we shall see that the classical methods
were generally success!ul in discussions of thermal radiation which
did not involve assumptions as to the nature of light but which were,
rather, concerned with certain energy relations. We have had
occasion to point out a similar case in connection with Carnot's
cycle, which was found to be correct irrespective of the particular
theory of heat adopted.
2. Some Fundamental· Concepts and Definitions~ (a) Total
Emissive Power.-The rate at which a given body emits radiation 1
depends upon the temperature of the body and on the nature of
its surface. We may define the total emissive power, symbol Es,
of a body S as the "total radiant energy emitted per unit time
per unit area of surface of the radiating body." The total
emissive power Es increases rapidly with increasing temperatures
and may be conveniently expressed in ergs per square centimeter
per second. Thus, the total emissive power of cast iron at a tem-
perature of 1600°K. 2 is about 1.1 X 10 8 ergs per square centimeter
per second; that of tungsten at 2450°K. is about 5 X 10 8 ergs per
square centimeter per second (or 50 watts per square centimeter).
(b) Monochromatic Emissive Power.-If the radiation from a
heated body, such as an incandescent lamp filament at a given
temperature, be dispersed into a spectrum by a prism or other
suitable device, it will be found, by means of a sensitive thermo-
pile, that the energy in the spectrum is distributed among the
various wave lengths in a regular manner, as is shown by the curve
in Fig. 42, which shows the distribution of energy in the spectrum
of tungsten at a temperature of 2450°K. We may define the
monochromatic emissive power, e-x, at any given wave length X,
1
In this chapter, unless otherwise stated, we shall use the term" radiation"
to refer to thermal radiation, the adjective being omitted to save repetition.
2
"K" refers to the absolute Kelvin scale of temperature, in centigrade
degrees.
180 QUANTUM THEORY [CHAP. VII
by saying that the radiant energy emitted in the spectral range
X to X + dX, per unit area per unit time, is given by e""dX. Thus,
the ordinates of the curve in Fig. 42 are ex, and the area of the
shaded strip is e""dX. Since the total emissive power Es refers to
all wave lengths combined, we obviously have the relation
between Es and e"" given by
Es = Jo"'e>-dh (1)
That is, the total emissive power is proportional to the area
between the curve and the wave-length axis. We shall
40 consider in detail, later, the way
in which Es and ex vary with
temperature.
The terms '' total emissive
:n
en power" and "monochromatic
'-
Q)
c emissive power'' are not to be
L.Ll
~ 2Q1--~-+--+-~b1-\----t--~~--i confused, respectively, with
>
+
~
Q)
0:: N
d)..· -
/ .
.,::
·6J / ~ dB
01----~--1-~~~~~__, -~ /
\ //
0 2 3.]Jw \ #
Woivelen9th -->dS
0
Frn. 42.-Distribution of energy in Frn. 43.
the spectrum of a tungsten lamp.
"total emissivity" and "monochromatic emissivity," which we
shall define later (Sec. 4).
(c) Intensity of Radiation from a Surface.-It will be convenient,
to avoid confusion, to define the term intensity of radiation from a
surface. Let dS (Fig. 43) be a small surface element of a radi-
ating body. Describe about dS a hemisphere of radius p, and
let dB, located at P, be a small element of the surface of this
hemisphere, the radius OP to this element making an angle (}
with the radius ON which is normal to dS. The rate dQ/dt at
which radiant energy is incident upon dB is easily seen to be pro-
portional to (1) the area dB; (2) the apparent area 1 dS cos fJ of
1 This is equivalent to saying that the rate at which a surface radiates
energy in a direction () from the normal to the surface is proportional to the
cosine of e. This well-known "cosine law" is obeyed only approximately,
the deviations from the law being greater for large values of O.
SEC. 2] SOME FUNDAMENTAL CONCEPTS 181
the radiator, as seen from P; (3) 1/p 2 , on account of the inverse-
square law; and (4) to a quantity is which depends upon the
temperature and character of the radiator, and which is called the
intensity of the radiation of dS. That is, by proper choice of units,
dQ . dB
- = isdS cos o · - 2 (2)
dt p
But dB/ p2 is the solid angle, call it def>, which the area dB sub-
tends at 0. For purposes of defining is we may, therefore, write
. dQ/dt
is = dS ·def>· cos e (3)
Or, in words, the intensity of radiation is from the surface dS at any
given temperature is equal to the rate at which dS radiates energy per
unit of (its) area per unit solid angle in a direction normal to its
surface (i.e., for cos (} = 1). This definition applies to the total
radiation from the surface. Just as in the case of emissive power,
we may designate by "monochromatic intensity of radiation,"
symbol ix, the intensity of radiation in a wave-length range d'J...
at wave length A.
There is a very simple relation between the total emissive power
Es and the intensity of radiation is. Referring to Fig. 43, if we
assume (1) that the radiation from the surface dS is perfectly
symmetrical about the normal ON and (2) that the surface obeys
the cosine law, we may choose as an element of area of the hemi-
sphere a ring of width pd() and of length 21rp sin(} the area of the
ring being, thus, 21rp sin(} · pd8. The flux of energy dQ/dt
through this ring is given, as in equation (2), by
dQ . dS 21rp(sin 0) · pd(}
-
dt
= is cos 8 · p2
(4a)
= 21risdS cos e sin Odo (4b)
If this equation be integrated from (} = 0 to (). = 1r/2, we shall
have the total rate of flux of energy dQs/dt from dS. We have
designated by Es the rate of flux of energy per unit area from a
surface. Therefore,
dQs/ dt r1r1 2
Es = dS = 27r'is J cos Osin odo
: . Es = 1ris
O
= 'Tris sin2 (J r
2
(5)
(6)
And, similarly, we have
(7)
182 QUANTUM THEORY [CHAP. VII
(d) Absorptivity.-In general, radiation, falling upon a surface,
is partly absorbed, partly reflected, and, unless the body be very
thick or very opaque, partly transmitted. We shall define the
absorptivity of a surface, symbol A, as the fraction of the radiant
energy, incident on the surface, which is absorbed. Absorptivity
is (1) a pure numeric; (2) for any actual body, less than unity;
and (3) varies greatly with wave length of the incident radiation
and, to a lesser extent, with the temperature of the absorber.
As we shall see in Sec. 4, there is a very simple relation between
the total emissive power Es of a surface and its absorptivity A.
(e) Reflectivity.-We may define the reflectivity of a surface,
symbol R, as the fraction of the radiant energy incident upon the
surface, which is reflected. R is a pure numeric.
(f) Transmissivity.-Likewise, we may define the transmis-
sivity, symbol T, of a body as the fraction of the radiant energy,
incident on the surface of the body, which is transmitted. We
shall, in this chapter, confine our attention to cases where the
body is so thick or so opaque as to transmit no energy, i.e., to
bodies for which T = 0. In these cases, since all the incident
radiation is, then, either reflected or absorbed, we may write
A+ R = 1 (8)
(g) The Density of Radiation.-The radiant energy per unit
volume in a stream of radiation is spoken of as the "energy
density of the radiation," symbol if;. Thus, the solar constant 1
is about 1.3 X 10 6 ergs per square centimeter per second. This
is the amount of energy contained in a column 2 of solar radiation
1 sq. cm. in cross-section and 3 X 10 10 cm. long, i.e., in 3 X 10 10
cc. The energy density i/; 8 of the sun's radiation in the neigh-
borhood of the earth is, therefore,
Yls = 1.3Xx1010
106 = 4 . 3 X 10- 5 ergs per cub"IC centimeter
.
3
In the case of the sun's rays, just considered, the radiation is
streaming in parallel directions. A different condition exists in
the interior of a hollow heated enclosure. Here radiation is
streaming back and forth in all possible directions. And the
term "energy density" then refers to the total quantity if; of this
1 The solar constant is the amount of the sun's radiation received on unit
area in unit time, the receiving area being perpendicular to the sun's rays
and at a distance from the sun equal to the mean radius of the earth's orbit.
2 The column being chosen in the neighborhood of the earth's orbit, and
the sun's rays being considered parallel throughout the length of the column.
SEC. 3] THE BLACK BODY 183
radiation in unit volume of the interior. Assuming symmetrical,
equilibrium conditions within the interior, we may, for simplicity,
regard this radiation the same as if an amount l/; /3 were streaming
in parallel streams in each of three mutually perpendicular
directions.
3. The "Black Body" and Its Properties.-We have, in nature,
no bodies which absorb all the radiation falling upon them, i.e., we
have no bodies for which the absorptivity is one. Some bodies,
however, such as black velvet or lampblack, reflect only a very
small fraction of the incident radiation, and in common parlance
we speak of such bodies as "black." Borrowing the term for
scientific phraseology, we may define the perfect or ideal black
body as a body the surface of which
absorbs all the radiant energy inci-
dent upon it. For such a body,
A= 1 and R = 0. 0
Although no such body actually
exists, we may approximate it as
closely as we please for both the-
oretical and experimental pur-
poses by making use of a small
hole in the side of a hollow enclo-
sure. Thus, in Fig. 44, let a ray FIG. 44.~The absorption of energy
of ra;diant energy PR enter the by a black body.
small hole O in the side of a hollow sphere, say of iron. A large
part of the radiation is absorbed as it strikes the inside of the
sphere at R, the remainder being diffusely reflected. Only a
very small fraction of this diffusely reflected energy goes out
through the hole, the rest being finally completely absorbed by
successive reflections. Theoretically, perfect absorption would
be reached only when the area of the hole is infinitely small com-
pared to the area of the hollow interior. Practically, an approxi-
mation sufficient for experimental purposes is obtained by using
a hole 1 or 2 cm. in diameter in the end of a hollow cylindrical
tube some 20 cm. long and 4 or 5 cm. in diameter.
When such an enclosure is heated, the inside walls radiate,
and some of this radiation passes out through the hole. That is
to say, the "black body" may radiate, as well as absorb, energy.
Such radiation is called "black body" radiation. Its properties
are of great importance, for both theoretical and experimental
purposes, in studying the phenomena of radiation. We shall
184 QUANTUM THEORY [CHAP. VII
designate by E (without subscript) the total emissive power of
the black body. As we shall see in Sec. 5 of this chapter, E
depends only on the temperature, and not on the material, of the
hollow enclosure.
4. Relation between Absorptivity and Emissive Power.-In
1833, Ritchie 1 performed an experiment which is very instructive
in connection with theories of radiation. A and B (Fig. 45),
are two hollow cubes containing air and connected by a glass tube
TT, in which is a small drop of fluid I to serve as an index, the
combination forming a differential air thermometer. Between
A and B is placed a third cubical vessel G, which may be filled
with hot water. , The faces 2, 2 of B and G, respectively, are
covered with a thin coating of lamp-
A1~:r B black. The faces 1, 1 of A a.nd C,
respectively, are of, say, silver and are
highly polished. The position of the
index I is noted, and then hot water is
I poured into C. Whereupon, the com-
FIG. 45.
partments A and B are heated by
radiation from C. Bis heated by virtue of the fact that its sur-
face 2 absorbs a certain fraction of the energy radiated by surface
1 of C; and vice versa for A. Since the respective emitting and
absorbing surfaces are inverted for the two compartments, it
might be expected, on account of this dissymmetry, that one of
the compartments would absorb heat at a greater rate than the
other. But on performing the experiment, it is observed that the
index I does not move when C is filled with the hot water. A and B
each absorb energy at the same rate. This experiment leads to a
law of extraordinary importance, which may be derived as follows:
Let E1 and A1 represent, respectively, the total emissive power
and the absorptivity of surface 1; and E2 and A2, the correspond-
ing quantities for surface 2. Assuming the emitting and absorb-
ing faces each to have unit area, the rate of emission from C
toward B is given by E1. A fraction A2 of this energy, is
absorbed by surface 2 and enters B. That is, B is receiving
energy at a rate A2E1. Similarly, A is receiving energy at a
rate A1E 2• Now, Ritchie's experiment shows that these two
rates are equal. Therefore,
A2E1 = A1E2
1 Ann. der Physik und Chemie, vol. 28 (Poggendorf), p. 378 (1833).
SEC. 4] ABSORPTIVITY AND EMISSIVE POWER 185
or,
(9)
In this same manner, we may compare other surfaces-3, 4, 5 ·
. . -in turn with surface 1, and we may include, in this com-
parison, a black body the emissive power of which we designate by
E and the absorptivity of which is unity. We may, then, write,
E1 = E2 = E3 = . . . = E = E (10)
A1 A2 A3 1
In words: At a given temperature, the quotient obtained by dividing
the emissive power by the absorptivity of any body is the same for all
bodies and is equal to the emissive power of a black body at the same
temperature. This is known as Kirchhoff's law, derived by
Kirchhoff, in 1859.
The derivation of Kirchhoff's law on the basis of Ritchie's
experiment is open to several objections among which is the fact
I
that the two surfaces of the same kind
(i.e., the two lampblack surfaces, for
example) are not at the same temperature.
An alternative derivation which avoids this
difficulty is as follows: Consider a uni-
formly heated enclosure (Fig. 46), say the
inside of a hollow iron sphere, in which
are two very small spherical bodies B 1 and
B2 of respective superficial areas 81 and 82. Frn. 46.
After a time, these bodies will reach the same temperature as that
of the interior, and, thereafter, each will radiate as much energy as
it absorbs. (We assume the interior of the sphere to be evacu-
ated, so as to avoid convection currents.) Within the enclosure,
radiation will be streaming uniformly in all direc.tions, and there
will fall upon the two bodies in unit time exactly the same inci-
dent radiation per unit area, call it I. Let E1, A1 and E2, A 2 be
the respective emissive powers and absorptivities of the two
bodies. Then B1 will emit radiation at a rate E1S1 and will
absorb radiation at a rate AiIS1. Since the temperature of B 1
remains constant, we have
E1S1 = AiIS.1
And for body B2, we have, similarly,
E2S2 = A2IS2
186 QUANTUM THEORY [CHAP. VII
From these two equations, we at once obtain, as before,
E1 E2
(9)
A1 - A2
The same argument could be extended to any number of bodies
within the enclosure.
A similar, though somewhat more rigorous, discussion would
show that Kirchhoff's law applies to the ratio of the monochro-
matic emissive power e.,._ to the monochromatic absorptivity A.,.. at
the same wave length. That is,
e.,..
A.,.. = constant (11)
the constant being the monochromatic emissive power of a black
body at the same temperature.
N ote.-There is considerable confusion in the terminology of the emission
and absorption of radiation. Thus, for the quantity which we have herein
defined as "total emissive power," one finds used, variously, such terms as
"radiating power," "emission," and "emissivity." And for the quantity
which we have called "absorptivity," 1 one finds "absorbing power" and
"absorptive power." The use of the word "power" in "total emissive
power" is obviously justified. For, "power" in the technical sense, in physics
and engineering, means a "rate of doing work" or of delivering energy; and a
radiating surface is emitting energy at a definite rate. But "power" used
in connection with absorption, as in the term "absorbing power," is clearly
an incorrect use of the word; for the quantity defined does not involve a rate
of absorbing energy.
One must distinguish between "absorptivity," as herein defined, and
'' coefficient of absorption," which latter is an entirely different term, refer-
ring to the absorption of radiation in its passage through matter. Thus, if
a beam of radiation of intensity I O is incident upon and passes through a slab
of absorbing material (such as light through smoked glass), the thickness
of which is d, the intensity I of the emerging beam is given by the equation
I = Ioe-ii.d, where ''e" is the Naperian base of logarithms andµ is the coeffi-
cient of absorption of the material. ,
It is essential, also, to distinguish between the terms "total emissive
power" and "emissivity." If E 1 is the total emissive power of a body at
temperature T, and Eis the total emissive power of a black body at the same
temperature, we have seen (equation (10)) that
E1 = E
A1
where A1 is the absorptivity of the substance. Now, the emissivity of a
surface is defined as "the ratio of its total emissive power to the total emis-
1 So far as possible, the terminology defined in the International'Critical
Tables, vol. 1, pp. 34-42, has been adopted in this chapter.
SEC. 5] EMISSIVE POWER OF A BLACK BODY 187
sive power of a black body at the same temperature;" i.e., emissivity equals
Ei/E. It is, thus, seen that the emissivity of a substance is a pure numeric
and is equal to its absorptivity. The same relations hold between mono-
chromatic emissive power and the corresponding absorptivity.
One or two comments may be made in passing: No body can
emit (thermal) radiation at a greater rate than can a black body
at the same temperature. For, the quotient Es/ As is constant
for all bodies, including the black body. The value of A for the
black body is the maximum possible, namely unity. There-
fore the emissive power of a black body is also the maximum
possible. For this reason, since the black body is called a perfect
absorber, it may, also, be called a perfect radiator.
Actual bodies are inferior, as radiators, to the experimental
black body. Thus, tungsten, at 2750°K., radiates, per unit
area, only about 25 per cent as much energy as a black body at
the same temperature. For iron at 1600°K., the corresponding
fraction is about 30 per cent.
5. The Emissive Power of a Black Body.-It will be instruc-
tive to examine a little more closely the factors which determine
the rate of emission of energy from a black body.
-· -h --
Equal areas of actual bodies-silver, tungsten,
carbon, iron-maintained at the same tempera-
ture will emit quite different amounts of energy.
That is to say, the radiation from an actual c
body depends both on the nature of the body
and on its temperature. Is the radiation from
a black body, likewise, dependent both on the
temperature and the material of the hollow
enclosure, the hole in the side of which forms the FIG. 47.
black body?
To simplify the discussion of this question, of course at the
expense of generality, consider the two dissimilar but parallel
surfaces 81 and 82, which form the opposite faces of a flat, disk-
shaped enclosure, the distance h between 81 and 82 being small
compared to their lateral dimensions. Sections of these surfaces
are shown in Fig. 47. Let the exterior faces of the enclosure be
impervious to heat, and let the temperature within the enclosure
have reached a constant value T. Each surface is then emitting
as much as it absorbs. Although, actually, radiation is stream-
ing in every direction within the enclosureJ on account of the
small distance between the faces 81 and 82 compared to their
188 QUANTUM. THEORY [CHAP. VII
dimensions, we may consider the phenomenon the same as if
all the radiation were streaming back and forth at right angles
to the two surfaces. Let E1, A1 and E2, A2 be the respective
total emissive powers and absorptivities of the two surfaces.
Further, we shall assume that the radiation travels with the
velocity of light c.
Consider, now, the "history" of the radiation emitted by unit
area (1 sq. cm.) of 81 during a time dt just sufficient for the
radiation, streaming toward 82, to fill a column 1 sq. cm. in cross-
section and of length h equal to the distance between the surfaces.
The amount of this radiation contained in the column is E 1dt,
the relation between h and dt being given by
h = c dt (12)
This energy falls upon 82, where a fraction ·A2 is absorbed and
the remaining fraction (1 - A 2) is reflected. After reflection
from 82 we shall have a column streaming toward 81, and the
energy contained in this column will be E1(l - A2)dt. After
this, in turn, has been reflected from A 1 and is streaming toward
the right again, we shall have E1(l - A1)(l - A2)dt. And so
on, for successive reflections. These successive values, together
with arrows representing the respective directions in which the
radiation is moving, are shown in column 1 of the following table:
RADIATION INITIALLY EMITTED RADIATION INITIALLY EMITTED
BY 81 BY 82
+-E2dt
~Eidt ~E2(l - A1)dt
+-E1(l - A2)dt +-E2(l - A 1) (1 - A2)dt
~E1(1 - A1)(l - A2)dt ~E,;.(1 - A1) 2 (l - A2)dt
+-E1(l - A1)(l - A2) 2dt +-E2(l - A1) 2 (1 - A2) 2dt
~E1(l - A1) 2 (l - A2) 2dt ~E2(l - A1) 3 (1 - A2) 2dt
etc. etc.
In column 2 are shown values, after successive reflections, of
energy E2dt, initially emitted by 8 2 during the same time interval
dt.
It is easy to see that, when a steady state has been reached,
there exist simultaneously columns of radiation, streaming right
and left, as shown in the table. Suppose we now wish to com-
pute the radiation Wr which, at any given instant, is streaming
toward the right. We should simply have to add those terms in
the above table with arrows pointing toward the right.
SEC. 5] EMISSIVE POWER OF A BLACK BODY 189
That is,
Wr = [Eidt+ E1(l - A1)(l - A2)dt +
E1(l - A1) 2 (l - A2) 2 dt + . . ] (13)
+ [E2(l - A1)dt + E2(l - A1) 2(l - A2)dt +
E2(l - A1) 3 (1 - A2) 2dt +
This readily reduces to 1
W _ E1 + E2(l - A1) dt (14)
r - A 1 + A 2 - A 1A 2
By Kirchhoff's law (equation (9)),
A2
E2 = -E1
A1
Putting this value of E2 in equation (14), we readily find, after
simplifying,
(15)
where, as before, E ( = E i/A 1) is the total emissive power of a
black body. That is to say, the radiant energy streaming to the
right within a column h = cdt cm. long and 1 sq. cm. cross-section is
Edt, or E/c units of energy per cubic centimeter.
If, now, a small hole, say of unit area, be opened up in the
right-hand face, this radiation, streaming toward the right
opposite the hole will pass out through the hole, the density of
the radiation, i.e., the radiant energy per unit volume emerging
from the hole, being E / c; or, in 1 unit of time an amount of
energy equal to E per unit area will be "emitted" by the hole.
We thus see (1) that a hole {n the side of the hollow, heated enclo-
sure emits radiation as if the hole were a perfect black body and (2)
that this rate of emission is quite independent of the nature of the
interior surfaces.
Thus, as previously pointed out, a hole in the side of a hollow
enclosure not only absorbs as if it were a perfect black body, but it
also emits radiation at the same rate as would a perfect black
body. Since this rate E is independent of the material of the
enclosure, it can depend only on the temperature T of the enclo-
sure; i.e., Eis some function of T, or,
E = f(T) (16)
1
Hint: These sums are of the form 1 + x + x 2 + , where x =
(1 - A1) (1 - A2). Since both A1, and A2 are always positive and less than
unity, 0 < x < 1, and 1 + x + x 2 + . = 1/(1 - x).
190 QUANTUM THEORY [CHAP. VII
It was one of the important successes of classical physics to find
the exact form of this function. As we shall see in subsequent
sections, the solution of this problem was reached without the
necessity of making any postulates as to the nature of radiant energy.
6. Pressure of Radiation.-That radiation falling upon a body
exerts a pressure in the direction of motion of the radiation is by
no means a concept peculiar to modern physics. Over 300 years
ago, Kepler suggested that the curvature of comets' tails away
from the sun might be due to light pressure. Newton recognized
that Kepler's suggestion was in harmony with the corpuscular
theory of light. 1 And later, Euler showed that such a pressure
should follow on the wave theory of light.
Experimentally, the subject was attacked by Du Fay, about
1750; by Bennet, in 1792; by Fresnel, in 1825; by Crookes, in
1873; and by Zollner, in 1877-all with inconclusive or conflicting
results, due for the most part to the disturbing action of the
gas surrounding the illuminated surface.
Meanwhile, Maxwell, in 1871, proved conclusively that a
radiation pressure is to be expected on the basis of the electro-
magnetic theory of light. Qualitatively, the line of argument is
somewhat as follows: (1) Lines of either electric or magnetic
force may be thought of as being under stress longitudinally and
as repelling each other laterally, this repulsion causing a lateral
pressure in the field. (2) According to the electromagnetic
theory, light consists of electric and magnetic vibrations at right
angles to the directions of propagations of the light. The lateral
pressure in these vibrating electric . and magnetic fields is, there-
fore, parallel to the direction of propagation of the light. (3)
When electromagnetic radiation is absorbed by a surface, these
fields are destroyed at the surface and the lateral pressure of the
fields in the oncoming waves produces a pressure on the surface.
Quantitatively, it may be shown 2 that the lateral pressure in
an electric field For in a magnetic field His, respectively, F 2 /81r
or H 2 /81r. When a train of electromagnetic waves falls normally
onto a surface, both the electric field and the magnetic field cause
periodically varying pressures on the surface, the average values of
the pressure being F~/l61r and H~/l61r, where F o and Ho are the
1 Just as the pressure produced by a gas is due to the impact of its
molecules.
2 See any treatise on electricity and magnetism, such as STARLING: "Elec-
tricity and Magnetism."
SEC. 6] PRESSURE OF RADIATION 191
respective maximum values of the two vectors. (We assume, for
simplicity, a plane-polarized wave.) The average value of the
total pressure onto the surface is the sum of these two. But it
has been shown (Chap. IV, Sec. 8) that the average energy per
unit volume in an electromagnetic wave train is F~/l61r +
H~/l61r, which is identical with the expression for the pressure.
Therefore, the very far-reaching conclusion: The pressure p
exerted by radiation onto a perfectly absorbing surface is numerically
equal to the energy per unit volume in the beam of radiation incident
onto the surface. This latter quantity we have called (see Sec.
2(g), this chapter) the "radiation density" VI· Accordingly, for
a parallel beam of radiation incident normally onto a perfectly
absorbing surface, we have
P = VI .(17)
If we are dealing with the radiation of density VI in a uniformly
heated enclosure, such as, for simplicity, a hollow cube in which
radiation is incident onto the inner surfaces at all possible angles
from O to 1r /2, we may conclude, from the fact that the pressures
on the three pairs of opposite faces must be equal, that the radi-
ation may be thought of as divided into three equal parallel
beams each of density 7'SVI, each beam being reflected back and
forth normally between a pair of parallel faces. In this case, the
pressure on each face would be >'SVI, where VI is the total radiation
density within the enclosure. That is, the pressure p produced
on the walls of a hollow heated enclosure, by the radiation within
the enclosure, is, for an enclosure of any form, given by
p = ~iVI (18)
where VI is the actual density of the radiation.
It can be shown that the pressure due to radiation is a
consequence of the law of the conservation of energy and is,
therefore, independent of, or more fundamental than, any theory
of radiation. Thus, consider a parallel stream of radiation, of
density VI, falling normally onto a black body of unit area (1 sq.
cm.). Energy is entering the body at a rate c · VI, where c is the
velocity of light. Now let the black body be displaced a distance
h opposite to the direction of the incident radiation. As a result
of this displacement more energy will have entered the black
body than would have entered had the body remained stationary,
by the amount of the radiation contained in a column of the inci-
dent stream 1 sq. cm. in cross-section and h cm. long, i.e., by the
192 QUANTUM THEORY [CHAP. VII
additional amount of energy hf. To introduce this additional
energy into the black body requires that a force f shall have acted
on the body to produce the displacement h such that the work
f · h done by the force shall equal f h. This force f must be
balanced by an opposing force by which the radiation must act
on the black body; and, since the black body has unit area, this
force f is numerically equal to the pressure p of the radiation.
That is,
p. h = fh
or
P = l/1 (19)
in agreement with equation (17). In this derivation of equation
(19), nothing was said about the nature of radiant energy.
If the radiation is totally reflected, instead of totally absorbed,
the energy density in front of the reflector is doubled, and the
pressure on the reflector is, therefore, double that which would
be experienced by a black body which has the same incident
radiation.
Under ordinary terrestrial conditions, this pressure p due to
radiation is a very small quantity. For example, as pointed out
in Sec. 2(g) of this chapter, the energy density of solar radiation
incident upon the earth is 4.3 X 10- 5 ergs per cubic centimeter.
Accordingly, the pressure which the sun's radiation exerts on the
earth's surface normal to the direction of the rays is 4.3 X 10- 5
dynes per square centimeter. It is, therefore, not to be wondered
at that Drude, as late as 1900, 1 remarked that "this pressure
. . . is so small that it cannot be detected experimentally."
Nevertheless, almost simultaneous with Drucie's statement came
the announcements of two independent researches-one in
Europe, by Lebedew, 2 and the other in America, by Nichols and
Hull 3-in which the pressure due to radiation was actually
measured with acceptable precision and the measured value was
found to agree almost perfectly with that predicted by theory!
For an account of this extremely important investigation the
reader is referred to the complete paper by Nichols and Hull, 4 who
1 DRUDE: "Theory of Optics" (English translation by Mann and Milli-
kan), p. 491.
2 Rapports presentes au Congres International de Physique, Paris, 1900.
3 Science (October, 1901).
4 Phys. Rev., vol. XVII (First Series), pp. 26 and 91 (July and August,
1903).
SEC. 6] PRESSURE OF RADIATION 193
measured the intensity of a beam of radiation by allowing it to
fall onto a blackened silver disk of known mass, the rate of rise of
temperature of the disk being observed. This same beam was
then allowed to fall on to a mirror system com prising a torsion
balance suspended by a quartz fiber in an evacuated enclosure,
the deflection produced by the radiation being observed. Know-
ing the constants of the torsion balance, it was possible to com-
pute the pressure produced by the radiation. The agreement
between theory and experiment for three different beams of
radiation is shown by the following table taken from the paper
by Nichols and Hull:
Observed pressure, Observed energy
Beam filtered through dynes per densities, ergs per
square centimeter cubic centimeter
Air .......................... . 7 .01 x 10-5 7.05 X 10-5
Red glass .................... . 6.94 6.86
Water cell .................... . 6.52 6.48
The agreement is seen to be excellent.
The experimental demonstration of the existence of radiation
pressure points to a very fundamental similarity which exists
between :radiation and a perfect gas. A gas confined within an
enclosure, at a given temperature, consists of molecules moving
in all possible directions with an average velocity dependent on
the temperature. By virtue of the kinetic energy of these
molecules, each unit volume of the gas contains a definite amount
of energy, which is determined by the temperature. These
molecules impinge on and rebound from the walls of the contain-
ing vessel at all angles of incidence from Oto 1r /2, producing there-
by a pressure which depends on the temperature of the gas. This
pressure is independent of the nature of the walls of the container.
Likewise, if we have a hollow enclosure completely empty of all
gas and at a definite temperature, there will be radiation stream-
ing back and forth in all directions. Irrespective of whether
radiation be corpuscular or undulatory, each unit volume of the
enclosure contains a definite amount of energy if; which is deter-
mined by the temperature. 1 This radiation is incident onto the
1 In the case of the gas, the energy per unit volume is determined by the
velocity (linear and angular) of the molecules. In the case of radiation, the
l'Plocity of the radiation remains fixed independent of temperature. On
194 QUANTUM THEORY [CHAP. VII
walls of the enclosure at all angles from O to 1r /2. Part of the
radiation is reflected, and part is absorbed and reemitted, the
whole process producing a pressure which depends on the tem-
perature and is independent of the nature of the enclosing walls.
This similarity between radiation and a perfect gas makes it
possible to apply to radiation the laws and methods of thermody-
namics, such as are applied, with far-reaching results, to gases.
In the next section, we shall discuss the application of Carnot's
cycle to the problem of determining the relation between the
emissive power of a black body and its temperature, using the
concept of radiation pressure as just discussed.
7. The Stefan-Boltzmann Law.-(a) In 1879, Stefan,1 on the
basis of data observed by Tyndall, was led to suggest that the
total emissive power of any body is proportional to the fourth
power of its absolute temperature. Tyndall had shown that the
ratio of the radiation from a platinum wire at 1200°C. to the
radiation from the same wire at 525°C. was as 11.7: 1. Stefan
pointed out that the ratio
(1,200 + 273)4
11.6
(525 + 273)4 -
thus indicating that
(20)
where Ept is the emissive poweif of the platinum at absolute
temperature T and <lpt is some constant. Subsequently, Boltz-
mann2 deduced the same law from theoretical considerations by
applying the laws of Carnot's cycle to the radiation in an "ether
engine," the radiation playing the part of the working substance.
The ideal Carnot engine consists of a cylinder with walls imper-
vious to heat; a piston likewise impervious to heat and moving
without friction; and a base (end opposite to piston) which is a
perfect (thermal) conductor. It may assist the reader to call
the corpuscular theory, the increase in energy with temperature comes from
an increase in the number of corpuscles. On the undulatory theory, the
increase comes from an increase in the amplitude of the light vector. In
the case of the gas, we have a distribution of velocities (more correctly, of
energies) among the molecules, a particular velocity being the most prob-
able. In the case of radiation, we have a distribution of energy among the
various frequencies, a particular frequency possessing a maximum energy.
1 Wien Akad. Sitzungsberichte, vol. 79, p. 391 (1879).
2 Wiedemann's Annalen, vol. 22, pp. 31, 291 (1884).
SEC. 7] TllE STEFAN-BOLTZMANN LAW 195
attention to the usual cycle of events, in the case where an ideal
gas is used as a working substance:
(a) Initially, the cylinder, containing a volume of perfect V1
gas at a pressure p1 and a temperature T1, is placed with
its perfectly conducting base on a source of heat at a
temperature T1, as shown at (a) in Fig. 48. The initial
position of the piston is indicated at P1 corresponding to
the state of the gas p1, v1, T1. The gas is allowed to
expand slowly, pushing the piston upward until the pres-
sure and volume are p2 and V2. The cooling which the
gas would experience as a result of doing work against
the external pressure is prevented by the conduction of
P.c VE 7} / P.J V,171 (/];, '!;, 'l}
:' ---- -.;, 1 - - - - - 1
~-- ~ 1-.....__--t
J 1-~--t
Perf'f!cf Perfecf
.7nsulcdor Jnsulcdor
(a) (b) (c) (d)
Frn. 48.-Carnot's engine.
heat, in total amount H 1, through the perfectly conduct-
ing base from the heat reservoir at tern pera ture T 1. The
expansion takes place so slowly that the temperature of
the gas remains constant during the expansion and the
process is said to be an isothermal process. It is repre-
sented on the p-v (pressure-volume) diagram, in Fig. 49,
by the line between the points P1 and P2.
(b) The cylinder is then transferred to the perfect insulating
stand, as shown in (b) (Fig. 48), and further expansion is
allowed to take place until the pressure and volume are
p3 and V3, respectively. Since no heat is allowed to enter
during this process, the external work done by virtue of
the expansion must cause a drop in temperature of the
gas to temperature T2. This process is an adiabatic proc-
ess and is represented in Fig. 49 by the line between the
points P2 and P3.
(c) The cylinder is now transferred to the heat "sink" at
temperature T2 (Fig. 48 (c)), and the gas is compressed
slowly to volume V4 and pressure p 4 , the heat H 2, pro-
196 QUANTUM THEORY [CHAP. VII
duced in the gas by the compression, flowing through
the conducting base to the heat reservoir, and the tem-
perature of the gas remaining constant. This is an iso-
thermal process and is shown by the line between P3 and
P 4 of Fig. 49.
(P) Finally, the cylinder is again placed on the insulating
stand (Fig. 48(d)), and the gas is further compressed,
along an adiabatic, P4P1 of Fig. 49, until the initial
pressure, volume and temperature are reached.
v
FIG. 49.-Carnot's cycle for a perfect gas.
We are concerned, in what follows in this article, chiefly with
the efficiency of the cycle. This we may express either as 1
. Area of cycle, P1P?aP4
Effi mency = Hi
or as
. T1 - T2
Efficiency = ~ --
1
Therefore,
T1 - T2 Area of cycle
__T_1_ - H1 (21)
If the adiabatic expansion, P2 to P3, is (relatively) very small, the
temperature difference between the upper and lower isothermal
is so small that we may write
T1 - T2 = dT
and under these restrictions, we may write equation (21)
dT _ Area of cycle
(22)
T - H1
(b) The fundamental principles involved in using radiation
as the working substance in the Carnot cycle differ in no way
from the above. Only the details of the ideal operating mecha-
nism are somewhat different from those employed in the case of
1 For derivation, see any standard text.
SEC. 7] THE STEFAN-BOLTZMANN LAW 197
the perfect gas engine. We may picture Boltzmann's "ether
engine" as made up of a hollow cylinder and a frictionless piston.
The entire cylinder walls, piston and base are impervious to heat;
have vanishingly small heat capacity; and are placed in a perfect
vacuum. In the base is a small opening 0, serving as a black
body, through which radiation may either enter or leave, and
which may be covered a~will by a perfectly reflecting cover
impervious to heat. The cylinder, whatever its initial temper-
ature, is placed with the opening O uncovered and opposite the
black body B 1 which is maintained at a temperature T1 (Fig.
50(a)). Radiation from B1 enters the opening 0, and, presently,
the interior of the cylinder, whatever the nature of its inner walls,
0 0 _o 0
<a> t b)
CD{ c) (d)
Fro. 50.-Boltzmann's ''ether" engine.
reaches the temperature T1, at which time the rate of radiation
from O to B1 is the same as the rate of radiation from B 1 to 0.
Under these conditions, the radiation density VI within the cylin-
der has a certain definite value determined only by T1. We
require to find the general relation between y; and T; for if we
know VI, the radiation density within the cylinder, we may, at
once, determine the emissive power of the opening in the cylin-
der-i.e., the emissive power of a black body-as explained in
Sec. 5 of this chapter.
Remembering that when the energy density within the cylinder
is VI, the pressure on the cylinder walls of }if (see equation (18)),
we may now consider the following cycle of events:
(a) Starting with the piston in the initial position P 1, (Fig.
50(a) )-the initial temperature being T1; the initial vol-
ume v1; and the initial pressure p1 = }13V11-we allow
the piston to move upward, slowly, as a result of the
pressure, until position P2 is reached, the volume
increasing to v2. If, during this process, the tempera-
198 QUANTUM THEORY [CHAP. VII
ture within the cylinder is to remain constant at T1,
radiation in amount H 1 must enter the opening O from
the black body B 1, for two reasons:
1. External work We is done by the piston. If T1
remains constant, i/;i and, therefore, p1 likewise remain
constant, and
We = }itJ;1(V2 - Vi) (23)
2. The interior of the cylinder has increased in volume
(v 2 - v1), which requires an additional influx of energy
equal to t/;1(v2 - v1).
To maintain the temperature constant, the influx H 1 of
radiation from B1 must be given by
H1 = }il/11(V2 - V1) + t/;1(V2 - V1)
= 'J,'$t/;1(V2 - V1) (24)
This isothermal process is represented on the p-v dia-
gram (Fig. 51) by the line P1P2. If we make the expan-
sion during this isothermal process very small, we may
write (v2 - v1) = dv1, and the corresponding small amount
of radiation dH flowing in to keep the temperature
constant at T1 is given by
dH 1 = ')il/11dv (25)
(b) When the piston has reached P 2, the perfectly reflecting
cover is placed over the opening O (Fig. 50(b)), thereby
effecting complete thermal isolation of the interior of
the cylinder, and a further expansion to position P 3 is
allowed to take place. External work is done, . as
before, by the piston. The energy required for this
external work must be supplied by the radjation within
the cylinder. Partly because of this external work,
partly because of the increase in volume during this
process, the energy density of the ra&iation within the
cylinder must decrease from t/;1 to t/;2, and, therefore,
the temperature within the cylinder must decrease 1
1 That the temperature within the cylinder has decreased can be shown by
again placing the engine, after the piston has reached P3 in the position
shown in Fig. 50(a), opposite the black body B1. On removing the slide
from the opening 0, radiation will stream from Oto B1 at a lesser rate than
at the beginning of process (a). Since B1 is at the same temperature T 1
as formerly, more radiation will enter O than leaves. The temperature
within the cylinder, therefore, will increase until the former temperature T 1
is reached and equilibrium between O and B 1 is again established.
SEC. 7] THE STEFAN-BOLTZMANN LAW 199
from Ti to T 2 • The pressure, likewise, has decreased,
since 'if; has decreased. This is, obviously, an adiabatic
process and is represented in Fig. 51 by the line P2Ps.
If we make the expansion during this process very
small, we may represent the change in temperature
(Ti - T 2 ) by dT and the corresponding change in
energy density (t/;i - y;2) by di/;. Since p = }if, we have
(26)
where dp represents the ~hange in pressure during the
process.
(c) The engine, now at the lower temperature T2, is placed
opposite the black body B2 (Fig. 50(c)), at temperature
T2, the slide removed from the opening 0, and the
piston is moved, by the
application of suitable ex-
ternal force, from Ps to P4.
On account of this com-
pression, there is a tendency '--~-'-~~~~~-v
for the temperature within Frn. 51.-Carnot's cycle for
the cylinder to rise and for black-body radiation in the
''ether'' engine.
the rate of emission of rad-
iation from O to B2 to increase. The compression is
supposed to take place so slowly, however, that the
temperature remains constant at a value only infinitesi-
mally in excess of T 2. During this second isothermal
process, PsP4 of Fig. 51, radiant energy, in amount H 2,
leaves the engine. The decrease in volume, dv, is as in
(a), assumed to be very small.
(d) Having reached a suitable point P4, the opening O is
closed, and the radiation is further compressed adiabati-
cally until the initial position Pi is reached correspond-
ing to the initial volume, energy density, and temperature.
The net external work done _during this cycle is represented by
the area P1P2PsP 4 of Fig. 51. If we assume the changes of
volume, of pressure, and of temperature to have been very small,
this · area equals dp·dv. (The figure is practically a small
parallelogram of length dv and of altitude dp.) Calling the net
external work dW, we, therefore, have
dW = dp · dv
= }i di/;dv (27)
200 QUANTUM THEORY [CHAP. VII
by equation (26). Remembering that the heat taken in during
the isothermal process P1P2 is 'JiVldv, according to equation (25),
we may now write the equation, corresponding to equation (22),
dT }i dVldv
(28)
T ~'3 Vldv
dT ldVI
.. T - (29)
4,f;
a relation between VI and T which, at once, integrates to
4 log T = log VI + constant (30)
or
VI = aT 4 (31)
where a is some constant.
Equation (31) states that the energy density of the radiation
within an enclosure is proportional to the fourth power of the
absolute temperature T. Since the emissive power E of a small
opening in the side of the enclosure is proportional to the energy
density within the enclosure, i.e., E ex VI, it follows at once that
the total emissive power of a black body is proport1:onal to the fourth
power of its absolute temperature. We may write, therefore,
E = aT 4 (32)
where u is a constant. This equation is of the same form as the
empirical equation (20) and is known as the "Stefan-Boltzmann
law." The constant a, known as "Stefan's constant," or as the
"Stefan-Boltzmann constant," has a numerical value,1 deter-
mined by experiment (see next section), of 5.709 X 10- 5 ergs
per square centimeter per second per deg. 4, if T is the absolute
temperature on the Kelvin scale.
In passing, attention may be called to an important relation
which exists between the constant a in equation (31) and the
constant u in equation (32). It may be readily shown that
4
a = - u
c
and therefore,
Vl=!E (33)
c
1 From the International Critical Tables.
SEC. 8] THE STEFAN-BOLTZMANN LAW 201
where c is the velocity of light. The proof of this relation is left
to the reader. 1 The numerical value of a, computed from the
above value of u, is
a = 7.617 X 10- 15 erg cm.-3 deg.- 4
8. Experimental Verification of the Stefan-Boltzmann Law.
The importance of the Stefan-Boltzmann law of total radiation
has inspired many investigators to check its validity. Among
these may be mentioned Lummer and Pringsheim, 2 who studied
4 80
w
401---~~~lf----~~+-~~~-t-~~~,
0
400 800 1200 1600°K
Temperature
Frn. 52.~Verification of the Stefan-Boltzmann total radiation law.
the total radiation from an experimental black body over a
temperature range from 373 to 1535° Abs. Their data are shown
1 Hint: Since the radiation within an enclosure is streaming equally in all
directions, it is readily seen that a fraction dw / 41r of the en tire radiation
within an element of volume dv is streaming in directions which lie inside a
small cone of solid angle dw described about dv as a center. Accordingly,
the flow of radiation through an element dS of the surface of a black body,
within a cone of solid angle dw, the axis of which makes an angle Owith the
normal to dS, is given by
dw
cfdS cos o 1r
4
where c is the velocity of light. From this, the intensity of the radiation i in
the direction fJ may be computed (see Sec. 2(c), this chapter), and equation
(33) follows at once by use of equation (6).
2 Ann. Physik, vol. 63, p. 395 (1897). These experiments are described
in PRESTON'S "Theory of Heat."
202 QUANTUM THEORY [CHAP. VII
graphically in Fig. 52, in which the abscissre are absolute tempera-
tures of the black body as measured with a thermo-element which,
in turn, was calibrated (indirectly) against a nitrogen gas
thermometer. The ordinates are the fourth roots of the total
emissive power E (in relative units) of the black body. Except
for the two lowest points (which involved troublesome correc-
tions), the relation between {/E and Tis seen to be in excellent
agreement with the Stefan-Boltzmann law.
From these measurements, Lummer and Pringsheim obtained
a value of u of 5.32 X 10- 5 erg cm.- 2 sec.- 1 deg.- 4 Numerous
recent investigations 1 have shown that the value of u is consider-
ably higher than that found by Lummer and Pringsheim, the
present accepted value, as quoted above, being u = 5.709 X 10- 5
erg cm.- 2 sec.- 1 deg.- 4
From these two constants u and a, one may compute the total
emissive power E of a black body at any temperature, and the
energy density VI of the radiation within an enclosure, by the
formulre
E = 5.709 · 10- 5 T 4 erg cm.-2 sec.- 1
VI = 7.617 · 10- 15 T 4 erg cm.-3
Thus, the total emissive power of a black body at a temperature
of 1000° Abs. is
E1,ooo = 5.71 X 107 erg cm.-2 sec.- 1
The energy density of the radiation within an enclosure at the
same temperature is
=7.62 X 10-3 ergs per cubic centimeter.
The pressure due to radiation on the walls of such an enclosure is
2.54 X 10- 3 dynes per square centimeter.
Note.-A word about temperature measurement: The universally
accepted scale of temperature is the Kelvin scale. It cannot be utilized
directly but is approximated with sufficient accuracy for experiment by a gas
thermometer (hydrogen or, for higher temperatures, nitrogen). Due to
obvious limitations, it is impossible to use gas thermometers above tempera-
tures of 1300 or 1400°C. It is, therefore, impossible to extend the Kelvin
scale, by means of the gas thermometer, above this temperature. The several
radiation laws-Stefan's and others-are checked experimentally up to 1300
or 1400°C. by use of the gas thermometer. Beyond these temperatures, the
1 For a summary of experimental values of u and other radiation constants,
see W. W. COBLENTZ, J our. Opt. Soc. Amer., vol. 5, p. 131 (1921) and vol. 8,
p. 11 (1924).
SEC. 9] BLACK-BODY RADIATION 203
laws themselves are used to establish a temperature scale. This procedure is
perfectly logical, particularly in the case of Stefan's law, since that law itself
is derived by use of the Carnot's cycle, which forms the basis of the Kelvin
scale. In other words, excepting as one method may be superior to another
experimentally, it should be a matter of indifference, in realizing the Kelvin
scale, whether we use a gas thermometer (making corrections for the behav-
ior of the actual gas used so as to make the measurements conform to a
perfect gas scale, which, in turn, is identical with the Kelvin scale) or whether
we use a radiation law, such as Stefan's, which is based upon Carnot's
cycle, and which, likewise, should provide a satisfactory means of realizing
the Kelvin scale.
9. The Spectral Distribution of Black-body Radiation.-In
the discussion in the preceding section on the Stefan-Boltzmann
law, it was not necessary to introduce any theories as to the
nature of radiation. And the experimental results were found to
be in complete agreement with theory. When, however. we
M
p
Frn. 53.-The spectroradiometer (Coblentz).
come to a discussion of the spectral distribution of the radiation
from a black body, we introduce, at once, the concept that
radiation consists of a wave motion. Here, classical theories
made some progress but could find only a partial solution of the
problem.
There have been many investigations to determine experi-
mentally how the distribution of energy in the spectrum of a
black body depends both on wave length and temperature.
As typical of the apparatus used in making measurements of
this kind may be mentioned the spectroradiometer, described by
Coblentz 1 and shown diagrammatically in Fig. 53. Radiation
from an electrically heated black body B enters the slit 8 1 , an
image of which is thrown on the slit 82 by means of the concave
1
GLAZEBROOK: '' Dictionary of Applied Physics," vol. IV, p. 554. This
contains (pp. 541-572) an excellent account of the determination of the
several radiation constants, the verification of the radiation laws and an
abstract of radiation theory.
204 QUANTUM THEORY [CHAP. VII
mirror M 1 and the plane mirror M 2. 1 Slit S2 is placed at the
principal focus of the concave mirror M 3, so that the rays reflected
from M 3 are parallel. After reflection from the plane mirror M 4,
the rays pass through the quartz or fluorite prism P, by which
they are dispersed into a spectrum, which is focused at S3 by the
concave mirror M 5 • By suitable angular adjustments, radiation
of any desired wave length may, thus, be caused to enter S3 and
fall upon the receiving element r of a
140 - - - - - - - - - - . - - - bolometer, contained in the enclosure
D, from which wires ww lead to the
measuring apparatus. The temper-
110---------
----------!
100
ature of the black body B is measured
by a thermo-junction.
After making necessary corrections
~ 90 t-----H---+---\++--+--+----1
+ for absorption in various parts of the
~ 80-----+----+--+----t apparatus, curves of the type shown
:n
~ 10 -----#4---!,-J..1--+--+----t in Fig. 54 are obtained for the energy
-+-
~
......_
<:C
60----.!---1.-4-.1.~-+-- distribution in the spectrum of a
black body at various temperatures .
~ 50 t---+--+---+--++---""4---+----t
Q) These curves are from the much
Q 401---+-t--+---+-~>--+-~ quoted measurements of Lummer
30--------~~~ and Pringsheim. 2 The wave lengths
20l-----+---#----4---...j...--
A are in microns (lQ- 4 cm., symbol
- - -
µ), the visible region of the spectrum
IOt---++-+-7""'=~~;:--P....,~
lying at the left of the curves,
00 I 2 3 4 5 6 between 0.4 and 0:8µ. The ordi-
Wavelength (in cm.x10 4) nates are monochromatic emissive
Ft o. 5 4. -- Distribution of powers ex (see Sec. 2(b)) in arbitrary
energy in the spectrum of a black units. The temperatures are in
body at various temperatures.
degrees (centigrade) absolute. In
the curve for the temperature 1646°, the circles represent actual
observations. (These are omitted in the remaining curves.) It
is to be noted: (1) that the energy at each wave length rises rap-
idly with increase in temperature and (2) that the higher the
temperature T the shorter the wave length Am at which maximum
energy occurs, the relation between Am and T being found
experimentally to be
Am T = constant = w
1
All mirrors are silvered on the front face.
2
Verhandlungen der Deutschen Physikalischen Gesellschaft, vol. 1, pp. 23
and 215 ( 1899) and vol. 2, p. 163 (1900).
SEC. 10] CLASSICAL THEORIES FAIL 205
This law is known as "Wien's displacement law." We shall
discuss it later.
At the time Lummer and Pringsheim reported the measure-
ments quoted in the preceding paragraph, several formula were,
or had been, proposed to express e as a function of X and T.
Among these were
1. Wien's formula: 1
c2
_L
,).
ex = c1x- 5e-XT (34)
2. Thiessen's formula: 2
c2
ex = c1X- 5 (XT)~e-xT (35)
3. Rayleigh's formula: 3
(36)
In these formulro, "e" is the natural base of logarithms, and c 1
and c2 are constants, not necessarily the same for the several
formulro. The measurements of Lummer and Pringsheim, and
the subsequent measurements of Rubens and Kurlbaum 4 at very
long wave lengths, proved that none of these formulro gave values
for the energy distribution throughout the entire spectrum which
agreed with experiment. For example, although Wien's formula
agreed with experiment at short wave lengths, there were sys-
tematic deviations on the long-wave-length side of the point of
maximum energy, the formula predicting a value too low, as is
shown by the dotted line in Fig. 54 for the temperature 1646°.
Just how far formulro such as these, based on classical methods
of deduction, were successful, we shall discuss in the next section.
10. The Successes and the Failure of Classical Thermody-
namics.-The fact that Wien's law predicts a spectral energy
distribution which agrees with experiment over part of the wave-
length range suggests that it is, in part at least, based upon
fundamentally correct principles. In this section, we shall
consider how far these principles succeeded and where they failed.
In this discussion, we shall resort to highly idealized experiments
with imaginary apparatus quite impossible of realization. Such
1 Ann. Physik, vol. 58, p. 662 (1896).
2 Verhandlungen der Deutschen Physikalischen Gesellschaft, vol. 2, p. 37
(1900).
3 Phil. Mag., vol. 49, p. 539 (1900).
4 Ann. Physik, vol. 4, p. 649 (1901).
206 QUANTUM THEORY [CHAP. VII
procedure is, of course, quite permissible in any discussion,
provided we test the final conclusions by a real experiment.
(a) The Wave-length-Temperature Displacement Law.-Consider
a hollow sphere (Fig. 55) radius r1, with walls which (1) are
perfect specular reflectors and (2) are capable of expansion so as
to increase the radius, say to r 2 • Let this sphere be entirely
empty initially of all radiation. At the center, we shall imagine
a very small spherical radiator b, which is at any specified tem-
perature T and which initially is covered so as to prevent the
escape of radiation into the interior of the sphere.
At a given instant, the radiator b is very suddenly uncovered,
and radiation begins to stream radially, in an expanding spherical
~------ .........
.......
wave train, toward the walls of the
' " sphere. We shall allow b to remain
/
I I
/
' \ in the sphere until the first wave of
I
I
I
, ,rz-,
-~
\\ the train, after specular reflection from
b ._------I'i ---- I the walls of the sphere, has returned
... , ,
I
\ I to b. At this instant, we shall sud-
\ I
\ I denly remove b from the enclosure.
\ I
We shall then have a train of (spher-
' '........
-
'-
- -- /
/
/
ical) waves (of course, of all wave
lengths) extending from the center to the
FIG. 55.
walls and back again.
These waves, after specular reflection from the walls, will
return toward the center, pass to the opposite side, where they
will again be reflected, and so will travel back and forth, remain-
ing spherical waves.
If we fix our attention on any one particular wave length X1,
the number n of those waves between the center and the walls of
the sphere is obviously given by
r1
n=- (37)
X1
Now, imagine the sphere to expand slowly until its radius has
reached r 2 • Because of the change in wave length, by Doppler's
principle, when waves are reflected from a receding mirror, we
shall have after this expansion exactly the same number of waves n
in the new radius r2, the new wave length X2 being given by
the relation
(38)
SEC. 10] CLASSICAL THEORIES FAIL 207
Therefore, combining equations (37) and (38), we have
(39)
or,
(40)
which shows that the increase in wave length is proportional to
the increase in radius, and this proportionality holds for any
wave length within the sphere. Since r2 was any new radius, we
may generalize equation (39) by writing,
r
-X = P>.. (41)
where P>.. is some constant.
This highly artificial limitation of the radiation within the
enclosure to spherical waves is, of course, unnecessary, as is also
the limitation of the walls to specular reflectors. (The walls must,
however, be perfect reflectors.) Thus, we might imagine a
hollow sphere, with perfect reflecting walls in which is a small
opening which can be placed opposite a black body at temperature
T. Presently, the sphere will be filled with radiation of density
if;, appropriate to the temperature T. Now close the opening
with a perfect reflector, thus imprisoning the radiation, which will
travel back and forth in all directions, in general, along chords
of the sphere. Because of Doppler's principle, and because the
proportionate increase in length of any cord of the sphere, if
the sphere be allowed to expand, is the same as the proportionate
increase in radius, it is obvious that any expansion will result in
an increase in wave length, of any given wave train, in the pro-
portion specified by equation (41). The radiation, both before
and after expansion, contains a whole spectrum of wave lengths,
but any particular wave length A1 before expansion corresponds
to a particular wave length A2 after expansion.
Consider, now, such a sphere with perfectly reflecting, exten-
sible walls, filled with (black-body) radiation at temperature T
and of density VI, such that
VI = aT 4
If the radius of the sphere be r, and its volume, accordingly,
;3,rr 3 , the total energy content U of the sphere is
(42)
208 QUANTUM THEORY [CHAP. VII
The pressure against the walls of the sphere is y,; /3. If the
sphere expands adiabatically, this pressure does external work
dW, given by 1
dW = }sf · 41rr2dr
= '},imfr2dr (43)
where dr is the increase in radjus. This external work will be
done at the expense of the internal energy U, which will, there-
fore, decrease by an amount dU obtained by differentiating equa-
tion (42),
(44)
since, in equation (42), U is seen to be a function of both r and y,;.
Since the expansion is adiabatic, we may write, with proper
regard for signs,
dU + dW = 0 (45)
Or, substituting for dU and dW from equations (44) and (43),
respectively, we have
41rr 2 fdr + ~i1rr 3dy,; + ~fn-fr 2dr = 0
: . 4fdr + rdf = 0 (46)
and therefore
dy,; + 4dr = 0 (47)
lf r
Integrating, we have
log y,; + 4 log r = constant
or, y,;r 4 = r, (48)
where r, is some constant.
Equation (48) shows that as the sphere expands adiabatically
the energy density y,; decreases inversely as r 4• This decrease in
l/; means a decrease in the temperature of the radiation. For, if
the slide over the opening in the sphere be removed after the
expansion has taken place, the radiation within the enclosure
will be in equilibrium with a black body at a lower temperature.
Now, from equation (31), y,; = aT 4 and, focusing our attention on
any wave length X, we have, from equation (41), a relation
between rand that wave length given by r = P>..A. Putting these
values of y,; and r in equation (48) gives
aT 4 pfA 4 = r, (49)
:. "X.4T4 = ~ = constant
llp>..
1Since there is an increase dv in volume of 41rr 2dr, and we may app]y the
general law dW = pdv.
~EC. 10] CLASSICAL THEORIES FAIL 209
or
A·T=G\ (50)
where CA is a constant for the particular wave length under
consideration.
Equation (50) is a very important equation. It states that if
black-body radiation at a particular temperature T1 be altered by
an adiabatic change to some other temperature T2, then a wave
length X1 in the radiation at T1 corresponds to a wave length X2
in the radiation at temperature T2, the relation between the two
wave lengths being given by X1T1 = X2T2, according to equation
(50). We may speak of this law as the ''wave-length temper-
ature" displacement law.
(b) The Energy-temperature Displacement Law.-N ot only
does a given wave length change according to equation (50)
when the temperature of the radi-
ation changes adiabatically, but also ~
the energy associated with that ~ 1 - - - - - dJ,. 1
particular wave length must change 1t
TAz
because of the necessity of doing
external work.
If, in the derivation of equation
(48), we had filled the sphere not v11
with the total radiation from the FIG. 56.
black body but with a narrow
spectral range of radiation comprised between the wave lengths
X and X +
dX, the monochromatic energy density of this radiation
being lf>.., we should have been led by exactly the same kind of
1
reasoning to an equation, analogous to equation (48), of the form
(51)
where "f/-x is a constant for the particular wave length under con-
sideration. Accordingly, when the expanding sphere has a radius
r, let us fix our attention on a narrow spectral range dX 1, at Ai
(Fig. 56). Let the energy density of this radiation be initially
f-x 1dA 1 and its temperature T1. Now, let the sphere expand
1 t/;-x is related to t/; in the same way that e>,. is related to E in equation (I);
i.e.,,/; = .( "',t,vA and ,t,-,..dA in equation (51) obviously corresponds to ,f; in
equation (48).
210 QUANTUM THEORY [CHAP. VII
adiabatically to radius r2. By this process, X1 will be displaced to
X2 by the relation (equation (40))
X1 r1
- (52)
X2 r2
The wave-length range dX1 will change to dX2, given by
2
dX2 = dXt (53)
r1
The temperature will drop from T1 to T2 (equations (50) and
(52)), given by
(54)
T2 = X1 = r1
And, as explained previously, the energy density VJ>.. 2dX2 will be
less than Vi)qdX1. We wish to find a relation between Vi>.. and T.
From equation (51),
i/;>,. 1dX 1r~ = i/;>,. 2dX2 r! (55)
By equations (53) and (54), we have
T1
dX2 = dX1T
2
and by equation (54),
(56)
Therefore,
1/1>..1 - f>..2 (57)
T51 - T52
That is, the value of the monochromatic energy density VI>.. is, at
corresponding wavelengths, directly proportional to the fifth power
of the absolute temperature. In general, omitting subscripts,
VJ>..
T5 = F>.. (58)
where F>.. is some constant. Since the monochromatic energy
density Vi>.. within an enclosure at temperature Tis proportional to
the monochromatic emissive power e>.. at the same temperature
(see equation (33) 1), we may write, instead of equation (58),
e>,. =
-T5 B>.. ( 59)
1 Equation (33) applies to the relation between total emissi ve power and
total energy density, but the same relation obviously holds between mono-
chromatic emissive power and monochromatic energy density.
='EC. 10] CLASSICAL THEORIES FAIL 211
where Bx is some constant. This relation is the energy-tempera-
ture displacement law sought.
(c) Experimental Confirmation of the Displacement Laws.-
These two displacement laws expressed in equations (50) and
(59), namely, X · T = C-x and e-xT- 5 = B-x, give important rela-
tions among e-x, X, and T, which have been verified by experiment.
Thus, to illustrate the applicability of the relations, let the curve
in Fig. 57 represent the experimentally determined energy distribu-
tion in the spectrum of a black body at a temperature T1.
Equations (50) and (59) enable us to compute what the energy
distribution ought to be at some other (say higher) temperature
"12~ ~
Frn. 57.
T2, as follows: Choose any wave length X1, for which the mono-
chromatic emissive power, at temperature T1, is e1, correspond-
ing to point P1 on the curve. The wave length X2, which, at
temperature T2, corresponds to X1, is given by equation (50),
T1
X2 = X1-
T2
And the emissive power~2, which corresponds to this new wave
length at temperature T2, is given, by equation (59), as
5
e2 = T 2)
e1 ( Ti
These two "displacements" shift point P1 to P2, which is a point
on the energy-distribution curve for temperature T 2 • In this
same way, every point on the curve at T1 could be transferred
and the energy-distribution curve for T2 could, thus, be mapped
out. It is found that if we start with a known distribution curve,
such as that shown for Ti, the predicted curve for temperature T 2
agrees with the experimentally observed curve for that temperature.
212 QUANTUM THEORY [CHAP. VII
Thus, turning to Fig. 54, point a on the isotherm for 1259°
corresponds to a wave length of 4.1µ and an energy eA (i.e., emis-
sive power) of 20. The wave length at temperature 1449° which
corresponds to 4.1µ at 1259° is given by
1,259
A = 4.1 · l = 3.55µ
449
'
This shifts point a to point b. The energy e~ at the new wave
length and temperature, which corresponds to eA = 20 a~t 1,259°,
is given by '-
, (1,449\ 5 40 5
e>. = 20 · 1,259} = ·
This shifts point b to point a', which falls on the experimentally
observed isotherm at 1449°. Points a and a' are, therefore,
corresponding points on the two isotherms.
(d) Wien:s DisplacementLaw.-It follows, from the above, that
the maximum points of the several isotherms are corresponding
points related to each other by the equations
e1n = B (60)
T5
and
Am· T = A (61)
where A and B are constants; em is the maximum (monochro-
matic) emissive power at temperature T; and Am is the wave
length at which this maximum emission occurs. Equation (61)
is known as "Wien's displacement law," and the constant A is
known as "Wein's displacement constant.'" Its value is
obtained by observing the wave length Am at which maximum
energy occurs in the spectrum of a black body at temperature T.
Thus, in Fig. 54, it is seen that maximum energy occurs in the
isotherm at 1449° at a wave length of about 2µ (0.0002 cm.), and
the product AmT for this isotherm is about 0.29, where Am is
expressed in centimeters.
Extensive investigations have shown that Wien's displacement
law is confirmed experimentally, the present accepted value 1 of
A being
A = 0.2885 cm. deg.
As an illustration of the constancy of A over a wide temperature
range may be quoted some measurements by Coblentz :2
1 International Critical Tables.
2 Bureau of Standards, Bull., vol. 13, p. 459 (1916).
SEC. 10] CLASSICAL THEORIES FAIL 213
TEMPERATURE, XmT = A,
DEGREES K. CENTIMETER-DEGREES
879 0.2888
953 0.2896
1,077 0.2898
1,175 0.2874
1,277 0.2882
1,353 0.2891
1,452 0.2884
1,540 0.2881
1,663 0.2887
(e) Another experimental test for the correctness of the two
displacement laws follows from the conjugate relation which
• =T=/646 °K
x ='11 =144.9 °.K
O:= 'F=IZS9°K
~T
Frn. 58.-Experimental verification of the displacement laws of black-body
radiation.
exists between points a and a', in Fig. 54. From the steps taken
to determine the position of point a' from the coordinates of point
a, it follows that if the ordinates e>.. of any isotherm at temperature
T be divided by T 5, and the corresponding abscissre X be multi-
plied by T, then a curve plotted with e>../T 5 as ordinates against
X · T as *1bscissre should be the same for all isotherms. There
is plotted, in Fig. 58, such a composite curve from data taken
from the three isotherms 1259, 1449, and 1646°, of Fig. 54, the
solid circles corresponding to 1646°K.; the crosses, to 1449°K.;
214 QUANTUM THEORY [CHAP. VII
and the hollow circles, to 1259°K. It is seen that a (reasonably 1)
smooth curve results.
(f) The curve shown in Fig. 58 is extremely significant, for it
represents the limit of success of classical theories in explaining the
spectral distribution of energy in the spectrum of a black body. We
have seen, in Sec. (c) of this article, that, given an experimentally
observed isotherm at one temperature, the isotherm at any other
temperature can be successfully predicted. This fact is, in
reality, also contained in Fig. 58. But the classical theories have
failed to predict the relation between e-x and X in the original isotherm
from which others may be computed. Or, to put the same idea in
different form, it is obvious, from Fig. 58, that e-x/T 5 is some
function (alike for all wave lengths and temperatures) of (X · T).
That is to say,
Te">,..5 = f ( X · T) (62)
wheref is the unknown function. In view of the relation between
X and T given by equation (50),
X • T = C-x,
it is possible to transform equation (62) into the alternative form
cA = cx- • f(x · T)
5
(63)
where C is a constant.
Wien, Rayleigh, 2 and others attempted to evaluate the function
f(X · T) and obtained the formulre previously quoted (equations
(34) and (35)). As pointed out previously, these formulre
failed to agree with experiment. The case for classical physics,
as based on thermodynamical grounds, therefore, rests with the
experimentally correct, but incomplete, equation (63). We
shall consider, in Secs. 11 to 13, a method of approach to the
radiation problem based upon the concepts of statistical
mechanics and the law of the equipartition of energy.
11. Degrees of Freedom and the Equipartition of Energy.-
(a) A gas molecule within an enclosure is free to move in a variety
of ways: It may, in general, have a motion of translation; a
1
The original data of Lummer and Pringsheim were given in graphical
form only. In obtaining data for Fig. 58 the coordinates e>.. and A were
read directly from the curves of Fig. 54.. Since only approximate values
were thus obtainable, some sea ttering of the points in Fig. 58 is to be
expected.
2
For an outline of the arguments of Wien and of Rayleigh see PRESTON'S
" Theory of Rea t."
SEC. 11] DEGREES OF FREEDOM 215
motion of rotation; and, if polyatomic, the atoms may vibrate
with respect to each other. A billiard ball upon the table does
not have so great freedom of motion as the gas molecule, for,
although the ball has as much freedom of rotation as does the
molecule, its motion of translation is confined to a plane. A flat
disk sliding upon the ice is not so free as is the billiard ball, for the
disk is capable of rotation about only one axis, namely, the
vertical axis. A block sliding in a closely fitting groove is still
more restricted in its freedom of motion.
Now, it is obvious that the relative freedom of motion of these
several bodies can be expressed by giving the number of inde-
pendent quantities which need be known to express completely
the position and orientation of the bodies. Thus, in the case of
the block sliding in the groove, its position and orientation is
completely specified by one quantity-its distance from some
or1gm. For the disk on the ice, three quantities are required-
two to express the position of some point, say the center of
gravity, of the disk with respect to some coordinate system on the
ice, and one more to specify the orientation of the disk with
respect to some fixed line. For the billiard ball, five such inde-
pendent quantities are required-two for position on the tab.le,
and three for orientation. These separate and independent
quantities which need be known to specify completely the position
and configuration of the body are called the '' degrees offreed om'' of
the body. The sliding block has 1 degree of freedom; the disk
has 3; and the billiard ball has 5.
This concept of degrees of freedom is readily extended to a
system of bodies. If we have three billiard balls on the table,
fifteen independent quantities are necessary to express the "posi-
tion and configuration of the system." Degrees of freedom are
additive-the total number of degrees of freedom of a system of
bodies is the sum of the number of degrees of freedom posessed by
the several bodies which make up the system. If, in a certain
volume of gas, we have a mixture of Ni molecules each with n 1
degrees of freedom, and N2 molecules each with n2 degrees of
freedom, the total number of degrees of freedom of the system is
n1N1 + n2N2.
Now, corresponding to each degree of freedom of a system is a
possible motion with which is associated kinetic energy. With
sufficient preciseness for our purpose, we may say that the num-
ber of degrees offreedom of a system of bodies is equal to the number
216 QUANTUM THEORY [CHAP. VII
of independent terms which are necessary to express the kinetic
energy of the system as a function of its coordinates. A monatomic
gas molecule behaves as if it had a motion of translation only. 1
Accordingly; its kinetic energy is completely expressed with
respect to an x-y-z coordinate system by the terms
}2mv; +~2mv; +
}2mv;
where Vx, Vy, and Vz refer to the component velocities in the three
coordinate directions and mis the mass of the molecule. If there
are N monatomic molecules in a given mass of gas, 3N such
independent terms are necessary to express the kinetic energy of
the gas at any given instant. On the contrary, a diatomic gas
molecule has 6 degrees of freedom, which we may count up as
follows: three for translation (of the center of gravity); two for
rotation of the "dumb-bell" shaped molecule about each of a
pair of axes at right angles both to each other and to the line
joining the atoms ;2 and one for the vibration of the atoms with
respect to each other. It would require 6N terms to express the
kinetic energy of N diatomic molecules. For a triatomic
molecule, we should have 9 degrees of freedom: 3 degrees of
translation; 3 degrees of rotation; and 3 degrees of vibration (on the
assumption that each of the three modes of vibration is possible).
(b) Let us consider, say, a monatomic gas in an enclosure at a
uniform temperature T. According to the kinetic theory of
gases, the molecules of the gas are in motion in all possible direc-
tions with a wide range of velocities, the average kinetic energies
of the molecules being a function of the temperature. The
degrees of freedom of such a system may be divided into three
groups, corresponding to the three coordinate axes, namely, an
x group, a y group, and a z group. The theorem of the equiparti-
tion of energy now says that each degree of freedom in one group
has, on the average, exactly the same amount of kinetic energy as has,
on the average, a degree of freedom in any other. group. For
example, let us take a census, or "snapshot," of 1,000 of the gas
molecules. Each molecule has a component of velocity in each
of the coordinate directions x, y, z. Let V'x, Vy, and Vz be these
components, and }2mv;, ~2mv!, }2mv; the corresponding kinetic
1 As will be seen later, this 1s probably due to the fact that the moment
of inertia of a monatomic gas molecule is negligibly small, so that, even if it
does rotate, its rotational kinetic energy may be negligibly small.
2 Rotation about the axis joining the atoms involves, apparently, no
kinetic energy (see preceding note).
SEC. 11] DEGREES OF FREEDOM 217
energies. The average kinetic energy per molecule Ex in the x
direction for the 1,000 molecules is obtained by adding the terms
}2mv; for all the molecules and dividing by 1,000. And,
similarly, for the y and the z components. The theorem of the
equipartition of energy states that
Ex= Ey = Ez
Let us designate by Ethe common value of Ex, EY, and Ez. Then,
if in the gas under consideration there are N molecules each with
3 degrees of freedom, the total kinetic energy of the system, due to
the random motion of its molecules, is 3NE. The same prin-
ciples hold for a diatomic gas, the total kinetic energy, in this case,
being 6NE. It is to be noted that in the case of the diatomic gas
the average rotational energy about each of the two axes is the
same as the average translational energy in each of the three
directions x, y, and z.
If we have an enclosure containing N 1 molecules of a mona-
tomic gas, say argon, and N 2 molecules of a diatomic gas, say
hydrogen, each hydrogen molecule should, 1 on the average, possess
twice as much kinetic energy as each argon molecule, since the
hydrogen molecule has 6 degrees of freedom while the argon
molecule has only 3. If, as before, the average energy per degree
of freedom be E, the total kinetic energy of the system, due to the
random motion of the molecules, should be (3N1 + 6N2)E.
In this latter case, the average energy E associated with the
"vibratory" degrees of freedom of the diatomic molecules gives
the average kinetic energy of vibration. This vibratory motion
becomes possible only if we assume that the 2 atoms which make
up the molecule are bound together by forces which, normally,
hold the 2 atoms in certain fixed equilibrium positions with
respect to some fixed point within the molecule. Vibrations
take place when this equilibrium is disturbed. Now, the energy
of any vibrating system is, in general, part kinetic and part
potential, the vibration really consisting of a cyclic interchange of
energy from the one form to the other, as, for example, in the
simple pendulum. For the simple pendulum it is easily shown
that the time average of its kinetic energy is exactly equal to the
time average of its potential energy. Applying this same
principle to the vibrations of the diatomic molecule, we see that
the average amount of potential energy associated with each
1
That is, if we assume that all of the degrees of freedom are "active.''
See, however, Chap. VIII, Sec. 10.
218 QUANTUM THEORY [CHAP. VII
"vibratory" degree of freedom is the same as its average kinetic
energy E. The total amount of energy associated with the
vibratory motion is, therefore, 2E.
If a small crystalline solid be suspended in the enclosure with
the gas, the solid soon reaches the temperature of the gas, the
molecules of the solid, by virtue of thermal agitation, executing
vibratory motions about positions of equilibrium. Each mole-
cule possesses, in respect of these vibrations, 3 degrees of freedom.
And if the average kinetic energy per degree of freedom of the
molecules of the gas in which the solid is suspended be E, the law
of the equipartition of energy requires that the average kinetic
energy per degree of freedom in the solid shall, also, be E and that
the average kinetic energy of each molecule of the solid shall be
3E. But the average potential energy is the same as the average
kinetic energy. The total (average) energy of each molecule of
the solid is, therefore, 6E.
This quantity E depends only on temperature and is independ-
ent of the nature or state of the substance.
12. Relation between Energy per Degree of Freedom and
Temperature.-The average energy per degree of freedom, as
discussed in the preceding section, is a very simple function of
temperature. This may be shown as follows:
According to the kinetic theory of gases, the pressure P which
a gas exerts on an enclosure is given by 1
P = Yspv 2 (64)
v
where p is the density of the gas and is the "mean square"
velocity of its molecules. Since p = nm, where n is the number
of molecules per cubic centimeter in the gas and mis the mass of
each molecule, we have
p = ~inmv 2 (65)
Assume that we are dealing with a monatomic gas at temperature
T. Then, the average kinetic energy EK of each molecule of the
gas is given by
EK=~~ mv 2
and we may write, instead of equation (65),
P = '3,f,nEx (66)
EK depends on the temperature.
Now,
p. VM = RT (67)
1 See any textbook on general physics.
SEC. 12) ENERGY PER DEGREE OF FREEDOM 219
where V M is the volume of a gram-molecule of the gas at tem-
perature T, and R is a universal constant known as the "gas
constant," the numerical value of which is 8.315 X 10 7 ergs per
mole per degree, or 1.987 calories per mole per degree. Putting
the value of P from equation (66) into equation (67) gives
(68)
The product of n, the number of molecules per cubic centimeter
of gas, by V M, the volume of the gram-molecule, is equal to No, the
number of molecules in a gram-molecule. That is,
nV m = No (69)
No, known as Avogadro's number, is constant for all substances.
Its numerical value is 6.061 X 10 23 molecules per gram mole-
cule. Putting this value of No from equation (69) in equation
(68) and solving for EK, we have
- 3 R
EK=2NoT
= J1 koT . (70)
where
R
ko = No
ko being another universal constant, the numerical value of which
is given by
k o -- 8.315
. X 101 -- 1.372 X 10-16 ergs per mo Iecu1e per degree
X 1023
6 061
The constant ko is known as "Boltzmann's constant," or as the
"molecular gas constant."
Equation (70) gives the average kinetic energy of the mona-
tomic molecule which we are considering. As pointed out in the
preceding section, such a molecule has 3 degrees of freedom.
The average energy per degree of freedom, therefore, i.e., the
quantity which we have designated by E, is obviously
E = }1koT (71)
which is the desired relation between E and T.
The theorem of the equipartition of energy thus leads to a
very simple way of computing the energy of thermal agitation,
kinetic or kinetic plus potential, of a system, provided (1) that
the system has a large enough number of degrees of freedom and
(2) that temperature equilibrium has been reached. We do not
need to specify, at least within certain limits, the various com-
ponents of the system. We need to know only its temperature
220 QUANTUM THEORY [CHAP. VII
and the number of degrees of freedom involved. Thus, the
average energy per degree of freedom in a gas in statistical
equilibrium at 300° Abs. is
E300 = }~ X 1.372 X 10- 16 X 300
= 2.06 X 10- 13 ergs
And the average kinetic energy of translation of a molecule of the
gas (3 degrees of freedom) is 6.18 X 10- 13 ergs, irrespective of
the nature of the molecules. The total kinetic energy of transla-
tion is obtained by multiplying this quantity by the number of
molecules involved.
(c) These deductions from the theorem of the equipartition of
energy have been verified in a very striking manner, both qualita-
tively and quantitatively, by observations on the Brownian
movements. One hundred years ago, Robert Brown, an English
botanist, noted that minute particles (small enough to be visible
only through a high-power microscope) in suspension in liquids
undergo continuously random motions, almost as if they were alive.
After the development of the kinetic theory of matter, these ran-
dom motions were explained as due to the continuous, though irreg-
ular, bombardment of the particles by the molecules of the liquid.
In 1905, Einstein developed a theory of the motion of these
particles based upon the concept of the equipartition of energy,
according to which their mean kinetic energy of agitation should
be exactly the same as the mean kinetic energy of agitation of a
molecule of the liquid at the same temperature, which, per
degree of freedom, is given by equation (71). Einstein's equation
is as follows: 1
(72)
where ~S is a small distance moved by the particle in a time r
-2
(several seconds) and flS is the average square of a large number
of such observed values of flS; K is a constant depending on the
viscosity of the medium in which the particle is suspended and
on the size of the particle; R, No, and Tare, respectively, the
-2
gas constant, Avogadro's number, and the temperature. flS
can be observed directly by watching the particle for a long time
in the field of the microscope and observing its position at fixed
1 For a derivation of this equation, and a discussion of its experimental
verification, see LOEB: "Kinetic Theory of Gases" (1927) or MILLIKAN:
"The Electron."
SEC. 13] THE RAYLEIGH-JEANS LAW 221
intervals of time; K can be computed; and it is, thus, possible to
determine experimentally R/No, which should be numerically
equal to the constant ko of equation (71). Actually, the value
of R/No, as determined experimentally by use of equation (72),
is combined with the known value of R and a value of N O is thus
determined from the Brownian movements. The value of N O sc)
obtained agrees, almost within the limits of error of measurement,
with that obtained by dividing the faraday by the charge on
the electron, thus proving that the law of the equipartition of
energy, upon which equation (72) is based, is valid and inciden-
tally verifying, beyond doubt, the kinetic theory of matter.
The law of the equipartition of energy is, thus, proved to have
a wide validity. And it was suggested by Lord Rayleigh 1 and
by Jeans 2 that this law might be applied to the radiation problem
by computing the number of modes of free vibrations in the
ether in an enclosure and by assuming that with each mode of
vibration, or degree of freedom, there is associated, on the average,
~~koT of kinetic energy and a like amount of potential energy, or
a total energy of koT per degree of freedom.
13. The Rayleigh-Jeans Radiation Law. (a) Degrees of
Freedom Associated with Ether Vibrations.-Any vibrating system
is, in general, capable of a great many modes of vibration. Thus,
a violin string or an organ pipe may vibrate not only in its funda-
mental mode but also in a great many overtones. We may
think of these several overtones as the various " degrees of free-
dom" of the vibrating system.
We may_ regard an organ pipe as a one-dimensional system, so
far as its standing sound waves are concerned, since the vibra-
tions take place longitudinally and in one direction only-
namely, parallel to the axis of the pipe. Standing waves may be
set up, also, in a "three-dimensional organ pipe," i.e., in a hollow
enclosure such as a large concrete-lined room. In this case, too,
we may have a great many possible overtones, or modes of vibra-
tion. We shall compute the number of such modes of vibration
within a given frequency range, per unit volume of the enclosure
for sound waves, and shall then apply the results to the analogous
case-the standing waves in the ether within an enclosure- and
shall, thus, determine the number of degrees of freedom per unit
volume of the ether, for a given frequency range. This number
1 Phil. Mag., vol. 49, p. 539 (1900).
2 Phil. Mag., vol. 10, p. 91 (1905).
222 QUANTUM THEORY [CHAP. VII
multiplied by the energy per degree of freedom, assuming the law
of equipartition, should give us the energy density of the radiation
of that particular frequency.
Since we are not so familiar, in our ordinary experience, with a
three-dimensional system of standing waves, it will be instructive
to apply the method of computation first to the case of a one-
dimensional system, say a stretched string, in which we may
have standing waves of a great range of frequencies. Let a
string of very great length L (Fig. 59) be stretched between two
~ ]~~
B
-··-··-------~-i.l
Fm. 59.
supports A and B. The condition that standing waves may be
set up requires (1) that A and B (i.e., the ends of the string) shall
be nodes of motion and (2J that there shall be an integral number
of equal loops between A and B. Since each loop is equivalent
to one half of a wave length, we may say that standing waves may
occur in the string only for those wave lengths X defined by
• L . 2L
X/2 = i = T (73)
where i is some (positive, of course) integer.
For a string of given length and tension, i represents the
number of possible modes of vibration of the string for wave
lengths X or greater, up to 2£. As i increases, X decreases. We
now wish to inquire: if i (when i is very large) increases from
i to i + Lii, what is the corresponding change of wave length
Li'X. in the system of standing waves? Or, to put the question in
somewhat more direct form: how many modes of vibration Lii
are possible in the wave-length range 'X. to X + LiX? The answer
is easily obtained: From equation (73) we have
2
;: = i (74)
When i changes to i + Lii, 'X. changes to X - Li'X., and
2L . +A. (75)
X - Li'X. = i /..l.i
SEC. 13] THE RAYLEIGH-JEANS LAW 223
Subtracting equation (74) from equation (75) and neglecting
small quantities (Lli is always taken small compared to i) gives 1
Lli = 2L ~X (76)
x2
If we divide Lli by L, we have the increase in i per unit length of
string, for a given change of wave length LlX, i.e.,
Lli = 2Ll~ (77)
L X2
Now, associated with each of these modes of vibration i, as
here defined, are 2 degrees of freedom, since the vibration is
transverse and any segment of the string is free to move in a
plane at right angles to the string. Representing degrees of
freedom by f, we have, therefore,
llf = 4~A2 (78)
L X
The number of degrees of freedom per unit length of string in the
wave-length range A to A - ~A is 4~;.
The situation is somewhat more complex but involves
identically similar principles, when we consider the possible
systems of standing waves in a hollow enclosure, say, a cubical
box. The essential difference in the two cases is most clearly
brought out by discussing first the systems of sound waves in a
flat box, a sort of two-dimensional organ pipe. Let ABCD (Fig.
60) represent such a rectangular flat box, and consider a system
of plane-parallel sound waves moving parallel to the top and
bottom of the box and being reflected, without absorption, from
its sides, according to the ordinary law of reflection. A set of
waves moving initially in the direction OM1 (Fig. 61) will, after
reflection at the face CD move in the direction M1M2. After
reflection at M 2, the direction M 2M 3 will be parallel to OM 1, etc.
For this group of waves, only four directions of motion are pos-
sible (i.e., +OM1 and + M1M2).
Let w1w1, w2w2, W3W3 • • • in Fig. 60, represent a series of
alternate wave crests and troughs moving initially in the direction
1 This equation may be derived readily from equation (74) by differentia-
tion: di = - 2L:2d>.. Strictly speaking, this is not permissible, since i is
an integer. When, however, i is very large and Lli is very small (e.g., when i
= 700 and Lli = 2), the algebraic result is the same by differentiation as by
the longer process.
224 QUANTUM THEORY [CHAP. VII
AP. After reflection from BC or CD, the direction of motion
will be parallel to +BP'. If, now, the direction of motion of
the waves and the wave length X be so chosen that the lines W1W1
W2W2, W3W3 • • • , when projected to the sides, divide the side
AB into an integral number of parts i1 and the side AD into
another integral number of parts i2 (i2 may or may not be equal
to i1), we shall have within the box a system of standing waves,
similar to those found in an organ pipe.
'' /
/
'' /
'~x
,.
,.
\. p' p /
/
/
/
' \. \. /
/
Frn. 60.-Standing sound waves in a "two-dimensional" enclosure.
Let the length of the sides of the box be d1 and d2; LPAB = (Ji
and LPAD = 82 • Then we shall have standing waves when,
simultaneously,
d1 cos 01 - . d d2 cos 02 - . (79-)
X/ 2 - ii an X/ 2 - i 2
where i1, and likewise i2, may be
any integer. The conditions for
standing waves in the three-
0 dimensional cases, i.e., the hollow
A-----~~"-~~---..J cubical box each edge of which is of
Ms length d is, by obvious extensions
FIG. 61.
of equations (79),
d cos 81
(80)
X/2
where 01, 82, and 83 are the angles which the direction of the wave
motion makes with the three edges of the box. The necessary
relation among the integers i1, i2, and i 3 is readily obtained by
making use of the relation that
cos 2 81 + cos 2
82 + cos 2
03 = 1
SEC. 13] THE RAYLEIGH-JEANS LAW 225
This gives
·2
i1
+~+
i2
·2 -
i3 -
4 2
d
A2 (81)
That is, the number of possible modes of vibration within the box
corresponding to waves the wave length of which is Am or greater,
i8 given by the number of possible combinations of integers i1, i2, i3,
the sum of the squares of which is 4d 2 /A~ or less.
By making use of a simple geometrical construction, we can
readily compute to a very close approximation the total number
'
i3
FIG. 62.
of such modes of vibration for wave length Am or greater, if we
confine ourselves to values of Am which are small compared to
the dimensions of the box, i.e., to large values of the sums of
the squares of the integers. Assuming for the wave length the
minimum value Am, we may write equation (81),
·2
i1
+~+
i2
·2 - 4d2
i3 - A2 (82)
m
which is the equation of a sphere if we regard i1, i2, i3 as the
variables of a rectangular coordinate system, the radius of the
sphere being 2d/Am. The maximum possible value im of each of
the three groups of integers will be that corresponding to a zero
value for each of the other two, i.e.,
iJ, +0 +0 = t~
m
2
226 QUANTUM THEORY [CHAP. VII
Now, lay off, on each axis of a rectangular coordinate system
(Fig. 62), equal unit increments of length to represent the integers
1, 2, 3 . . . im; 1, 2, 3 . . . im; and 1, 2, 3 . . . im, respec-
tively. The total possible number n of combinations of the three
integers giving sums of squares equal to or less than 4d 2 /X! is,
obviously, numerically equal to the volume (in terms of the
arbitrary unit of length) of the octant of the sphere of radius
2d/Xm. This number n is, therefore,
n = 7i · %1r(~~)3 (83)
Each one of these possible n combinations of integers corresponds
to one mode of vibration within the box, i.e., to 1 degree of freedom.
Since d 3 is the volume of the box, the number of degrees of free-
dom per unit volume of the box is 1
n 41r 1
cf> = da = 3 X3 (84)
where cf> is the number of degrees of freedom per unit volume
corresponding to vibrations of wave length X or greater, up to 2d.
Now, just as in the case of the string, we are interested in the
number of degrees of freedom (per unit volume) in a wave-
length range X to X +
dX. This is readily determined by differ-
entiation2 of equation (84), viz.,
d<t, = -41rdX (85)
X4
which is analogous to equation (76) for the string.
In this discussion, we have been dealing with the longitudinal
vibrations of sound waves. An identical analysis will hold for
the transverse vibrations of the ether waves, excepting that, in
the latter case, because ether waves are transverse, the number of
degrees of freedom is twice as great as for longitudinal waves.
For the ether, therefore, we may write,3
dX
d<t,}.. = 81r X4 (86)
where d<h is the number of degrees of freedom per unit volume of
the ether in the wave-length range X to X + dX.
1 We may now drop the subscript in "Xm to make the formula general.
2 See note, p. 223.
3 In writing equation (86) we drop the minus sign of equation (85),
since, in what follows, we are interested in d<t> only as a positive number.
SEC. 13] THE RAYLEIGH-JEANS LAW 227
(b) The Rayleigh-Jeans Formula.-We are now in position to
apply to the radiation problem the law of the equipartition of
energy. Consider the radiation within an enclosure the walls of
which are at a temperature T, equilibrium having been reached
between the radiation and the walls. As previously explained,
the radiation then corresponds to black-body radiation at that
temperature. Whatever the nature of the radiating and absorb-
ing mechanism of the walls, each degree of freedom of the en(ire
system, walls and ether included, should have associated with it an
average amount of kinetic energy Y2koT, as given by equation
(71). Each "vibratory" degree of freedom has an equal amount
of potential energy. The total amount of energy associated with
each degree of freedom in the ether should be, therefore, k 0 T.
By equation (86), the number of degrees of freedom per unit
volume in the ether is 81rA- 4d}.. for the wave-length range }.. to
A + dA. The radiant energy per unit volume in the ether in the
wave length }.. to }.. + d}.., therefore, should be the product of koT
by 81r}..- 4d}... This product gives the monochromatic energy
density fxdA of the radiation. That is,
fxdA = 81rkoTA- 4dA (87)
is the equation for the spectral energy distribution in black-body
radiation as deduced from the theorem of the equipartition of energy.
This is the famous Rayleigh-Jeans formula.
(c) Test of the Rayleigh-Jeans Formula.-It is obvious, at once,
that equation (87) cannot represent the distribution of energy
throughout the spectrum of a black body. For a given temper-
ature T, the value of Vix, according to the equation, should vary
inversely as the fourth power of the wave length, increasing
rapidly to infinity as the wave length approaches zero; whereas the
actual curves shown in Fig. 54 rise to a maximum with decreasing
wave length and then decrease to zero. In only one particular
does the Rayleigh-Jeans equation agree with the experimental
facts: At long wave lengths it is found experimentally that the
energy is inversely proportional to the fourth power of the wave
length. For a limited range, the Rayleigh-Jeans expressjon
seems to be in accord with experiment-a circumstance which
suggests that, in some particulars, at least, the fundamental
principles upon which the deduction of the formula is based must
be correct.
But a little consideration will show that the law of the equipar-
tition of energy, as stated above, and the computation of the
228 QUANTUM THEORY [CHAP. VII
number of degrees of freedom in the ether, according to equation
(86),
dX
def, = 81r-
X4
cannot simultaneously be correct. For, if we assume that the
ether is perfectly continous, i.e., has no structure such as have
material bodies, it should be capable of transmitting 1 waves of
wave length varying from infinity to zero. Hence, by integration
of equation (86) from X = oo to X = 0, we should find that the
number of degrees of freedom in the ether should be infinite.
And if the law of the equipartition of energy be correct, we should
find that the energy density of black-body radiation, when in
equilibrium with the walls of an enclosure at a given temperature,
should be vastly greater, really infinitely greater, than the energy
density in the walls themselves. ,vhereas, we find that quite
the opposite is the case. Thus, as stated on page 202, the radi-
ation density in the ether in equilibrium with an enclosure at
1000°C., is of the order of 8 x 10-3 ergs per cubic centimeter,
while the energy density of the thermal agitation of the molecules
of the walls (say of iron) is of the order of 10 10 ergs per cubic
centimeter. Actually, the energy density of the radiation is
only an almost imperceptible fraction, 10- 12 , of the energy den-
sity in the walls with which the radiation is in equilibrium; while,
according to the Rayleigh-Jeans formula, the reverse ought to
be the case.
Apparently, then, classical statistical mechanics has failed to
yield a law for the spectral distribution of energy in black-body
radiation, just as classical thermodynamics failed. Wien's
formula (equation (34))
c2
ex = c1X- 5e -xr
based on the thermodynamic arguments, predicts correct values
of ex for short wave lengths but gives too low values at long wave
lengths. The Rayleigh-Jeans formula (equation (87)), which
may be put in the form
ex = CkoTx- 4
where C is a constant, agrees with observed values of ex at long
wave lengths but gives too high values at short wave lengths.
1 It is obvious that the minimum wave length of the elastic vibrations
which such a body as a crystalline solid can transmit must be at least
greater than the average distance between the molecules of the body.
SEC. 14] PLANCK'S RADIATION LAW 229
The problem was attacked in various ways by numerous invest.i-
gators1 during the first decade of the twentieth century. All
deductions based on classical statistical mechanics led to the
same incorrect result-the Rayleigh-Jeans formula. Arguments
based on classical physics could get no farther than equation (63).
Meanwhile, Planck, in 1900,2 by introducing a radical innova-
tion quite at variance with previous concepts, discovered a
function of XT which gave a formula in complete agreement with
experiment. This was the birth of the quantum theory. We shall
devote the remaining sections of this chapter to a deduction of
Planck's formula.
14. Planck's Radiation Law: The Birth of the Quantum
Theory. (a) The Statistical Distribution of Energy among
Degrees of Freedom.-In the preceding sections, we have con-
sidered only the average energy per degree of freedom without
paying attention to the actual amount of energy possessed by
any particular degree of freedom. Let us consider, again, the
case of a gas in an enclosure and at uniform temperature. For
a monatomic gas, each molecule has 3 degrees of freedom, one
corresponding to each of the three coordinate axes, x, y, and z.
We may, as before, divide the total number of degrees of freedom
into three groups-an x group, a y group, and a z group-the
number of degrees of freedom in each group being equal to the
total number of molecules in the enclosure.
Of these molecules, some have a very high velocity; some have a
very low velocity, theoretically approaching zero; while the great
majority have intermediate velocities grouping about a "most
probable" velocity, which depends, for a given gas, on its tem-
perature. The law governing the distribution of velocities was
deduced by Maxwell and is found in current treatises on the
kinetic theory of gases. 3 It would take us too far afield to
deduce this law, but the following may help to make its meaning
clear: Any very small volume of the gas contains a very large
number of molecules. Let us represent the velocity of each
molecule of this small volume at a particular instant by a vector
1 For a summary see F. REICHE: "The Quantum Theory," p. 15 (English
Translation by Hatfield and Brose, E. P. Dutton and Co.).
2 M. PLANCK: Verhandlungen der Deutschen Phys. Gess., vol. 2, p. 237,
1900. Also Ann. Physik, vol. 4, p. 553 (1901).
3 JEANS: "The Dynamical Theory of Gases"; LOEB: "The Kinetic Theory
of Gases''; LEWIS, W. M. C.: "Kinetic Theory."
230 QUANTUM THEORY [CHAP. VII
drawn from the origin of an x-y-z coordinate system. The
totality of such vectors, one for each molecule, will result in a
perfectly symmetrical system of vectors, or lines, extending
radially from the origin, in all directions, some of the vectors
being long, some short, and the majority of them being of inter-
g mediate length. Imagine each vector
1 1111
1
to terminate in a minute sphere, such
I 1111 1
as s, in Fig. 63 (the z-ax1s is assumed
1, 1, 11
1 1 11 1
1 perpendicular to the plane of the
11 1
---t'+'+-++14-H-+-J-l-+----.x figure). Now divide the whole con-
1 1111
1
111 I i , struction into a series of very thin
11111 I 11 I I 11
1 1 1 11 1 1 1 1 1 1 1 slices parallel to the yz plane as shown
:::l:: ::l ::l by the dotted lines, the slices being all
Frn. 63. of the same thickness. And let us
count the number of the small spheres
(ends of vectors) which we find in each slice. A little reflection
will show that the nearer a slice is to the yz plane the more
spheres the slice will contain. Representing by nx the number
of spheres found in a particular shce and by Px the ordinal number
of the slice, counting from the one next to the yz plane as p = 1
FIG. 64.
(i.e., p = 11, say, represents the eleventh slice from the yz
plane), a curve plotted between nx and Px/qo has the form shown
in Fig. 64, 1 q0 being the number of slices in unit length along the
1 Only one-half of the curve is shown, namely, that in the positive direc-
tion of x of Fig. 63. The curve is obviously symmetrical about the y-axis.
SEC. 14] PLANCK'S RADIATION LAW 231
x-axis of Fig. 63. Applying the laws of probability, it is deduced
in connection with the kinetic theory of gases that this curve has
an equation of the form
(88)
where x = Px/'l_o, e is the Naperian base of logarithms, and no is
a constant the significance of which is readily seen by letting
x = 0. Then, nx = no; i.e., no is the value to which n approaches
as x ..:. 0. To a very close approximation, no may be thought of
as the number of spheres in the first slice next to the yz plane,
provided qo is large.
Since the velocity vectors are distributed perfectly sym-
metrically about the origin of coordinates in Fig. 63, both as to
direction and magnitude, it is obvious that curves identical with
Fig. 64 would have been obtained if we had taken the slices in
Fig. 63 parallel to the xz plane or to the xy plane.
Now, each little sphere counted in Fig. 63 represents 1 degree
of freedom-in this case a degree of freedom of the x kind, since
the distance of the slices from the yz plane is proportional to the
x component of the velocity. And, therefore, nx is equal to the
number of x degrees of freedom associated with each slice.
Further Px, the ordinal number of a given slice, is proportional
to the x component of those velocity vectors (all assumed equal)
whose end points are found in that particular slice. Likewise,
qo must be proportional to a unit velocity arbitrarily assumed as a
unit in drawing the vectors in Fig. 63. The square of Px is
proportional to the kinetic energy Ex in the x direction 1 of the
particular group of molecules whose velocity vectors end in slice
Px· And the square of qo is proportional, likewise, to an arbitrary
unit of kinetic energy E 0 • Remembering that x 2 = Pf = EEx
qo o
we may, therefore, write, instead of equation (88),
(89)
where nx is now the number of degrees of freedom possessing
energy2 Ex. Because of the symmetrical distribution of veloci-
ties, we have identical expressions for the y group and for the z
group of degrees of freedom.
1 More correctly, "the kinetic energy due to the x-component of the
velocity," since energy itself is not a directed quantity.
2
As noted below, Ex is assumed to be made up of small, but finite, incre-
ments.
232 QUANTUM THEORY [CHAP. VII
The total number of degrees of freedom is the sum of the three
groups. Letting n (without subscript) stand for degrees of
freedom, irrespective of direction, and, similarly, replacing Ex
by E, we have, with suitable corresponding change in the mean-
ing1 of no,
n = noe-E/Eo (90)
In the kinetic theory of gases, Ea is shown to be identical with
the quantity which, in preceding paragraphs, we have called the
average energy per degree of freedom; i.e., Ea = }~koT, where, as
in Sec. 12 (equation (71)), ko is the Boltzmann constant and Tis
the temperature. We may, therefore, write
- E
n = noe "H,koT (91)
as the equation which represents the distribution of energy2
among the various degrees of freedom in a gas in statistical
equilibrium at temperature T.
(b) Mean Energy per Degree of Freedom.-Let us consider a
system containing a very large number N of degrees of freedom in
the condition of statistical equilibrium to which equation (90)
n = noe -E/Eo
applies. Let the energy be distributed among the several degrees
of freedom in multiples of a small unit of energy e, which, later,
we may allow to approach zero, so that E = pe. where pis some
integer. We then have
np = noe -pe/Eo (92)
where np is the number of degrees of freedom possessing energy
p<:, and no is the number of degrees of freedom possessing zero
energy. We may then write,
no degrees of freedom possess energy O
n 0e -E/Eo degrees of freedom possess energy e
n 0e- 2e/Eo degrees of freedom possess energy 2e
n 0e - 3e!Eo degrees of freedom possess energy 3e
and so forth.
1 That is, in equation (90), n 0 refers to the total numbe1 of degrees of
freedom possessing zero energy, all three groups combined.
2 Here, Eis to be taken as a multiple of some small quantity of energy
E; i.e., E = pE, where p is an integer; n is the number of degrees of freedom
possessing p such units of energy. In the limit, of course, e should approach
zero.
SEC. 14] PLANCK'S RADIATION LAW 233
The total number N of degrees of freedom in the system is the
sum
That is,
N = no(l + e-e/Eo + e-2E/Eo + e-3E/Eo ... ) (93)
no
= 1 _ e-e/Eo (94)
since the series in the bracket is of the form
1
(1 + x + x2 + x3 . . . ) = 1- x
where x < 1. Here x = e-e/Eo, which must always be less than
unity.
The total energy W of these N degrees of freedom is obtained
by multiplying the number of degrees in each group by the energy
per degree possessed by that group and adding the products;
thus,
W = (no X 0) + (noe-e/Eo X e) + (noe- 2e/Eo X 2e)
+ (nae - 3e/Eo X 3e) +
= noee-e/Eo (1 + 2e-e/Eo + 3e-2e/Eo + ... ) (95)
noee-e/Eo
(1 _ e-e/Eo)2 (96)
since the bracket, in equation (95), is of the form
1
(1 + 2x + 3x + 4x
2 3
• • • ) = (l _ x) 2
when x < 1
The average energy U per degree of freedom of these N degrees
of freedom is
-U=-w (97)
N
Putting in equation (97) the value of W from equation (96) and
the value of N from equation (94), we have
_ ee-e/Eo
U = 1 - e-e/Eo (98)
- €
ee/ EO -1
.·. U = - -
(99)
by dividing both numerator and denominator of equation (98)
by e-e/Eo.
234 QUANTUM THEORY [CHAP. VII
If the energy unit e be assumed to have a small, but still finite,
value, equation (99) gives the average energy per degree of
freedom. But if, as we would ordinarily assume in the kinetic
theory of gases, the value of U is obtained only when e approaches
zero, i.e., if we assume that there is no finite lower limit to the
possible dijference in energy between molecules, we find, from
equation (99), when 1 e = 0,
U = Ea
in agreement with the previous statement with regard to the
meaning of Ea, namely, that Ea is the mean energy per degree of
freedom.
(c) Planck's Formula.-We shall assume that the equation (99)
U = __e_ _
eE/Eo _ 1
is of general validity in giving the average energy per degree of
freedom in any system. Let us consider once more a hollow
enclosure which contains radiation in equilibrium with the walls
of the enclosure, all at a temperature T. The radiation com-
prises a continuous spectrum of frequencies covering a wide
range. The walls are absorbing and reemitting this radiation by
means of an atomic mechanism the nature of which we need not
specify in detail but which, if radiation be electromagnetic,
probably consists of vibrating electric systems of some kind.
The frequencies of these "atomic oscillators," as we may call
them, each of which has a particular natural frequency, must,
all together extend continuously over a range of frequencies at
least as wide as does the radiation with which the oscillators are
in equilibrium. And it is reasonable to assume that the radiation
within a frequency range v to v +
dv must be in equilibrium, irre-
1 If Eis put equal to zero in the right-hand side of equation (99), the value
of U is 0/0. If
f(x) 0
F(x) =0
when x = O, the value of the indeterminate may be found by putting
x = 0 in
f'(x)
F'(x)
where f'(x) and F'(x) are the derivatives with respect to x of the functions
f(x) and F(x).
SEC. 14] PLANCK'S RADIATION LAW 235
spective of the presence in the system of other frequencies, with
the group of oscillators whose frequencies cover the same range
v to v + dv.
We may think of these oscillators as possessing 1 degree of
freedom each, or we may think of them as possessing several
degrees of freedom each. In any event, let us assume that
equation (99)
- €
U = ee/Eo - 1
gives the average energy U per degree of freedom of the oscil-
lators, where E 0 , taking into account both kinetic and potential
energy, has the value
Ea = koT
That is, let
- €
Uv = ee/koT _ 1 (100)
where now Up is the average energy per degree of freedom of the
group of oscillators found in the frequency range v to v + dv.
Because of the assumed equilibrium between the radiation
and the oscillators, Up must also be the average energy per degree
of freedom in the group of radiation frequencies lying between v
and v + dv. Corresponding to this frequency range dv is a wave-
length range dX, and, for convenience, we shall change the sub-
script of U and let Ux equal the average energy per degree of
freedom in the ether vibrations in the wave-length range ""A to X
+ d""A..
In Sec. 13 (a), we have computed the number of degrees of
freedom per unit volume d</Jx in the wave-length range X to
X + dX, viz. (equation (86)),
dX
dcpx = 81r X4
If each degree of freedom has energy Ux, the total radiant energy
per unit volume in the range X to X dX is given by +
i/;)..d""A. = Ux • dcp).. (101)
Or, putting into equation (101) the values of Ux and dcf>x, we have
81r €
Vl'J..d""A. = A. 4 • ee/koT - 1dX (102)
an equation which should represent the relation among Vix, X, and
T in black-body radiation.
236 QUANTUM THEORY [CHAP. VII
It now remains to interpret the small, arbitrary element of
e,
energy which we used in deriving equation (99). According to
classical mechanics and the law of the equipartition of energy, E
should be an infinitesimal quantity, i.e.; as stated on page 232,
there should be no finite lower limit to the possible difference in
energy between the several degrees of freedom. The value of
ih. dX, in equation (102), should, therefore, be obtained by letting
E approach zero. It is readily seen that when Eis zero, equation
(102) reduces to
(103)
a result identical with equation (83), the Rayleigh-Jeans formula,
which, as we know, is in complete disagreement with experiment
except for very long waves. Something wrong somewhere!
When theory and unimpeachable experimental evidence disa-
gree, theory must be modified. Accordingly, let us compare the
incomplete formula, equation (63), 1 which, so far as it goes, does
agree with experiment, with equation (102), viz.,
i/;)..dX = C1X- 5f(X · T)dX (63')
V/"AdA = 81rA-\.;k) _ 1dA (102)
The two formulre differ in two respects: (1) Equation (63') has
x- 5 , while equation (102) has x- 4 ; (2) the fraction
€
ee/koT - 1
in equation (102), as it now stands, is a function of T and not of
"'A • T, as required by equation (63'). Both of these differences
would at once disappear, if, in equation (102), instead of allowing
e to approach zero, we were to disregard classical principles
and were to put E proportional to 1/X. This requires that E shall
be a finite quantity. We therefore put
1
€ ex: -
"'A
or,
1
e = a- (104)
"'A
1 In rewriting equation (63) here, to compare it with equation (102), we
have written th in place of e'J.., with a suitable implied change in the constant
c1. This is permissible, since t/;).. ex: e'J.., as pointed out by equations (33).
SEC. 14] PLANCK'S RADIATION LAW 237
where a is a proportionality constant. In terms of frequency v,
instead of wave length X, this is equivalent to assuming, since
X = c/ v (c =
velocity of light),
a
€ = c-v = hv (105)
where h = a/ c is another constant.
Putting the value of€, from equation (104), into equation (102),
we have
(106)
which is of exactly the form required by equation (63), since, in
equation (106), we have x- 5 , and the fraction is now a function
of X · T. The equation is more usually written in the form,
since a = ch,
(107)
This is one form of the famous Planck equat1:on for the spectral
energy distribution in black-body radiation. An alternative form
of this equation is obtained by replacing wave length X by fre-
quency v as follows:
c c
}... = - and dX = - -dv
v 712
This gives
(108)
Before proceeding to discuss these equivalent formulre (equations
(107)) and (108), it is pertinent to emphasize the radical nature of
the assumption e = hv (equation (105)), which marks the birth
of the quantum theory. According to the older concepts, including
the electromagnetic theory of radiation, the interchange of energy
between two "systems," as, for example, between one gas mole-
cule and another or between radiation and matter, should be a
perfectly continuous process. Thus, if one could follow the
''history'' of a particular gas molecule and observe its kinetic
energy at short intervals of time over a very long period, one
would find, of course, that its energy varied over enormously
wide limits. But if the numerical values of the energy observed
at the various instants were to be arranged in a sequence as to
order of magnitude, this sequence, on the concept of the equiparti-
238 QUANTUM THEORY [CHAP. VII
tion of energy, would become more and more continuous as the
number of observations becomes greater. Or, if, instead of
observing a single molecule over a long period, one were to
observe, at a given instant, the kinetic energy of a very large
number of gas molecules or of the above-mentioned oscillators,
chosen at random, one ought to find that a similar sequence of
values of energy would approach more and more nearly to continu-
ity as the number of observations becomes larger. This continu-
ity of these energy values is not only in accord with, but is also
imperatively demanded by, classical physics. For example, the
electric and magnetic vectors in a light wave may have any values
whatsoever, from zero up; and, accordingly, the wave may have
any intensity, from zero up. Further, the emission and absorp-
tion of this energy by the walls of an enclosure should, likewise,
be a perfectly continuous process.
But the assumption contained in equation (105)
e = hv
is completely at variance with this older doctrine of continuity.
According to this assumption, the numerical sequence of the
values of the energy of a large number of the oscillators, of, say, a
particular frequency v, is a discontinuous series increasing by
equal jumps e, from zero up. Some oscillators would be found to
have zero energy; others would have energy e; still others, energy
2e, 3e, etc. Not a single oscillator would be found which had
energy 1.75e or 3.94e. These oscillators are in equilibrium with
radiation of the same frequency in the enclosure, the several
degrees of freedom of which, therefore, must, likewise, possess
energy 0, le, 2e, 3e, or 4e, etc. The interchange of energy between
the oscillators and the radiation must take place in multiples of
e, i.e., of hv, whatever the nature of radiation, of the oscillators, or
of the mechanism of the interchange.
The difference between the two views is rather crudely illus-
trated by the difference between putting very fine sand and mar-
bles all of_ equal mass into a bag. The mass of sand in the bag
may be adjusted to any desired value whatever, at least so far as
ordinary weighing apparatus goes; while the mass of marbles in
the bag must always be an integer multiple of the mass of one
marble. The analogy is worth carrying further: Even in the
case of the sand, if we were limited either in filling or in emptying
the bag to the use of a cup of certain size and to transferring the
SEC. 14] PLANCK'S RADIATION LAW 239
sand, in or out, only by cupfuls, we should find that the mass of
sand in the bag was an integer multiple of the mass of a cupful.
The process of filling, rather than the material involved, may be
responsible for this integer relation. But we must remember
that, even in the case of the marbles, an ingenious small boy with
a hammer might get 3.94 marbles into the bag!
That Planck's postulate is in direct opposition to the law of
the equipartition of energy is clearly brought out by considering
the equation (100) for the average energy U of a single oscillator,
11
or of a degree of freedom of ether vibration of frequency v,
U = _ _e_ _
,, eE/koT _ 1
which, by replacing e by hv, becomes
- hv
Uv = -eh_v_/k-oT-_ 1 (lOOa)
In the oscillator-ether system are degrees of freedom of various
kinds: oscillators and ether vibrations of many different fre-
quencies. According to the classical theories, i.e., equipartition
of energy and continuity, each degree of freedom of one kind
should have exactly the same energy, on the average, as a degree
of freedom of any other kind. That is, the vibrations of high
frequency within a given frequency range should have the same
average energy as vibrations of a lower, or any other, frequency.
This at once follows from equation (lOOa), by introducing the
principle of continuity, which, as above stated, requires that
e = hv = 0
Putting hv = 0 in equation (lOOa) gives, as we have seen,
Uv/:= koT
0
which shows that, on the classical theory, the energy per degree
of freedom is independent of frequency and is the same for all
frequencies.
But if hv cannot be zero but is, as Planck assumed, a finite
quantity, then Uv varies with frequency, and there is no longer
equipartition of energy among the degrees of freedom of different
frequencies. For very high frequencies, i.e., for short wave
lengths, U v becomes very small: in the limit, when v = oo,
Up = 0, as may be shown by putting v = oo in equation (lOOa).
Accordingly, in spite of the fact that the number of degrees of
240 QUANTUM THEORY [CHAP. VII
freedom def>>.. for ~ given wave-length range dX increases very
rapidly as X becomes smaller, (equation (86)), the total
energy included in that wave-length range decreases, as is
required by experiment. On the other hand, for small values of
v, i.e., for large values of X, Uv approaches koT. The energy per
degree of freedom, therefore, ranges all the way from zero for X =
Oto koT at very long wave lengths. This assumes, of course, the
validity of the computation for the number of degrees of freedom
per unit volume
dX
d<t>x = 81r X4
an assumption which seems justified in view of the fact that it is
one of the important factors leading to Planck's equation.
In passing, it may be remarked that the principle of equiparti-
tion is retained in a restricted sense. It is assumed that the
average energy U of an oscillator of frequency vis the same as the
11
average energy of ether vibrations of the same frequency. In
other words, there is equipartition of energy, of a kind, among
degrees of freedom of the same frequency, although, even here,
the distribution of energy among the oscillators is discontinuous.
It is worth commenting on the fact that Planck's revolutionary
assumption e = hv was not based upon logical reasoning follow-
ing the ordinary lines of classical physics. Quite the contrary:
the assumption was an attempt, entirely empirical, to bring the
deductions of classical physics into harmony with experiment. And
the attempt was most astonishingly successful.
(d) Verification of Planck's Law and Evaluation of the Con-
stants h and ko.-Equation (107)
l/;xdX = 81rchX- 5 (ech/koXT - 1)- 1dx
contains three constants c, the velocity of light; ko, Boltz-
mann's constant, the numerical value of which is (p. 219) 1.372 X
10- 16 (in c.g.s. units); and a third constant h, which has come to
be known as "Planck's constant." If this equation for the
relation among YI>.., X, and T be the correct one, it should be in
agreement with the following experimentally correct relations:
1. Wien's law at short wave lengths.
2. The Rayleigh-Jeans law at long wave lengths.
3. Wien's displacement law: Xm · T = constant.
4. The Stefan-Boltzmann law of total radiation: y; = aT 4
SEC. 14] PLANCK'S RADIATION LAW 241
That Planck's equation coincides with Wien's law at short
wave lengths is readily seen. For a given value of the tempera-
ch
ture T, the exponent of e, namely koXT' can be made to assume
any positive value by choosing a sufficiently small value of X,
since c, h, and ko are constants. The term
ech/koXT
may, therefore, be made very large compared to unity. For such
small values of X, we may write,
i/1>-..dX = 81rchx- 5e -ch/koXT dX (109)
which is identical with Wien's equation
i/1>-..dX = C1X-5e-c /XT dX 2
provided 1 we put c1 = 81rch
and
ch
C2 = ko
Considering both X and T to vary, Planck's equation is practically
identical with Wien's for values of X · T, which are small compared
to ch/ko.
At sufficiently large values of X, or, more generally, of X · T,
the exponent of e, namely,
ch
koXT
becomes small. Expanding2 ech/ko>..T and dropping terms involving
the square and higher powers of the small quantity k;:T' we have
1 koXT
-ec-h/-k-oX-T - l - cJ:
Putting this value in Planck's equation gives, for long wave
lengths,
i/1>-..dX = 81rkoTX- 4dX
which is identical with the Rayleigh-Jeansformula, the latter, as has
been pointed out, giving correct values of fx for long wave lengths.
1 Note that, in writing Wien's equation here, we are again using i/;x and ~
interchangeably, with the usual implied change in the constant c1•
x2 xa
2
ex=l+x+-+-+
I~ l~
242 QUANTUM THEORY [CHAP. VII
That Planck's formula yields Wien's displacement law
Am · T = constant
where Am is the wave length at which the energy density y;)\
is a maximum, follows, at once, from differentiating YI>-. with
respect to A and putting the derivative equal to zero. Writing
Planck's equation,
lf')\ = C1A-5(ech/ko)\T - 1)-l
(where c1 = 81rch), we have
ch
+ C1A-5(ech/ko)\T _ 1)-2ech/ko)\T koTA-2
Putting this derivative equal to zero, and designating by Am the
value of the wave length for which lh is a maximum, we have,
after simplifying,
1_ ch ) e ch/ko)\mT = 1 (110)
( 5koAmT
Equation {110) is of the form
(111)
where
ch
x=--
- koAmT
Equation (111) has two roots 1 : x = O; and x = 4.965. In the
former root (x = 0)- we are not interested. From the latter,
we have
ch
4 955
koAmT = ·
or,
ch
Am. T = ko X 4.965 (112)
which, since the right-hand side is constant, is in complete agree-
ment with Wien's displacement law.
1 The value of x in equation (111) may be determined by Newton's
method. SILBERSTEIN: "Synopsis of Applicable Mathematics," p. 85.
SEC. 14] PLANCK'S RADIATION LAW 243
The experimental value of Xm • Tis 0.2885 cm.-degree, (see p.
212). Putting this value into equation (112) and treating both
constants h and ko as unknowns, we may solve for a numerical
value of h/ko, V1,°z., (c = 2.998 X 10 10 ),
h _ 4.965 X 0.2885 _ 4 778 10-11 d d (113)
ko - 2.998 X 1010 - . X egree-secon
To compare Planck's formula with the Stefan-Boltzmann law
of total radiation, it is necessary to integrate Yl'XdX over the range
of wave lengths from A = 0 to X = oo; i.e.,
i/; = aT 4 = f 0
"' lf>..dX. (114)
The integration is somewhat more readily performed by using
equation (108), it being obvious that
f: i/; .dv = J: 00
lf>..dA
We write, therefore,
o 8~hfo
YI =
J co Vlvdv = c3
To integrate this expression, first expand the bracket under the
co v3(ehv/koT - 1)-1 dv (115)
integral, as follows:
1 -~ -2 .!!!!__ -3 .!!!__
ehv/koT - 1 =e koT +e koT +e koT
Equation (115) then becomes
YI =
8~
c3
hf O
CX> v3(e
- !!!!_
koT + e- 2 !!!__
koT + e - 3.l!!_
koT + ... )dv (116)
This can be integrated term by term. Take, for example, the
first term (omitting the constant 8~h/c3 ),
O - hv/lcoT
J CX>
v 3e dv.
Letting -hv/k 0 T = z, a new variable, we can write this integral
in the form,
(
k
~ T)4JO"'z edz 3
Using the standard form,1
f xme"dx = xme"' - m xm-le"'dx f
1 See any table of integrals.
244 QUANTUM THEORY [CHAP. VII
one finds, by straightforward calculation, after putting in the
limits, 1 that
enJz ezdz = 3
Similarly, for the second term of equation (116), one finds
6en 4
and for the third,
kT)
6( h
0
4
_!_
34
and so on.
The complete integral of equation (116) thus becomes
YI = 81rh.
c3
6(koT)
h
1+ _!_2 + 3_!_ + 4_!_ +
4
(
4 4 4
.. ·)
=[
4
~1r • :! · 1.082 T 4 J (117)
Since the square bracket of equation (117) contains only constant
terms, we see that equation (117) is identical in Jorm with the
Stefan-Boltzman law for total radiation,
YI = aT4
and, that, therefore,
a = 4~t . tta . 1.082
The experimentally determined value of a (see p. 201) 1s
7.617 X 10- 15 ergs-cm.- 3-deg.-4. Therefore, in c.g.s. units,
:: = 7.948 x 10- 16 (118)
0
These two numerical relations between h and k, viz.,
; = 4. 778 X 10- 11 (113)
0
~: = 7.948
0
x 10- 16 (118)
1 In evaluating the several terms resulting from this integration, it will
be necessary to obtain the value of the indeterminate oo / oo by successive
differentiation of numerator and denominator.
SEC. 14] PLANCK'S RADIATION LAW 245
when solved simultaneously for hand ko, give
ko = 1.344 X 10- 16 ergs per degree,
h = 6.423 X 10- 27 erg-seconds (119)
The value of ko as determined by dividing the gas constant R by
the number of atoms in a gram-atom No is (seep. 219) 1.372 X
10- 1 6, a value which differs by only 2 per cent 1 from the value com-
puted from radiation data on the basis of Planck's formula. This
numerical agreement alone, although not quite within the limits
of experimental error, is so significant as to give very great weight
to Planck's formula and its general underlying principles. To
this confirming evidence must be added the fact that Planck's
formula yields Wien's law for short wave lengths, the Rayleigh-
J eans law for long wave lengths, the Wien's displacement law, and
the Stefan-Boltzman total radiation law. And this confirmation
includes the confirmation of the revolutionary postulate which,
alone, distinguishes Planck's result from its predecessors, namely,
e = hv
But what is the nature of the constant h which is contained in
Planck's assumption and the numerical value of which is found
above, to be 6.423 X 10- 21 ? That is a question which modern
physics is trying to answer. We know a great deal more about
the numerical value of h than of its physical significance. With
regard to the latter, although various statements about h take on
a great variety of forms, they boil down to about this: e = hv!
Or, in words, "h is the constant, dimensions erg-second, which
multiplied by frequency v gives a quantum of energy of that
frequency." Not a very "satisfying physical definition."
1 It is interesting to see whether Wien's formula in the form given in
equation (109) (i.e., Planck's formula for short waves) would give values of
h and ko in agreement with experiment. Deriving from equation (109)
(instead of equation (108)) the expression for XmT corresponding to equation
(112), one finds
ch
XmT = -
5ko
Similarly, instead of equation (117), one finds
f =481r ~ T4
c3 h3
The numerical values of h and ko determined from these are
ko = 1.526 X 10-rn ergs per degree
h = 7 .25 X 10-27 erg-seconds
Wien's formula is seen to give a value of k 0 about 11 per cent too high.
246 QUANTUM THEORY [CHAP. VII
As to numerical value, if we combine the ratio h/ko = 4.778 X
10- 11 with the best experimental value of ko, as determined from
the gas constant R, namely, ko = 1.372 X 10- 16 , we find, for h,
h = 6.555 X 10- 27 erg-second
Birge 1 gives a summary of the values of h determined in various
ways and concludes that the most probable value 2 is
h = (6.5543 + 0.0025) X 10- 27 erg-second
~
\<.t:_
\~
\ -·
~
\ \
''--
\ ('\)
\~
\CJ\-
' \Q
r
,-f:..
\
''
''
''
' ....
...........
0 2 J 4 5 6
Wave Length (cm.x 104 )
F10. 65.-Comparison of the three radiation laws with experiment, for the
radiation from a black body at 1600°K.
Finally, in Fig. 65 is shown a comparison between the several
spectral energy distribution formulre and experimental data.
1 Phys. Rev., vol. 14, p. 361 (October, 1919).
2 It is a curious instance of the mutual cancellation of inaccuracies in
numerical data that Planck in his original paper in which he computed the
value of h (Ann. Physik, vol. 4, p. 563 (1901)) obtained the value 6.55 X
10-21 • He used for the value of "'AmT that found by Lummer and Pringsheim
a few months before, namely "'AmT = 0.294 cm deg.; and for the value of a
in the Stefan-Boltzmann equation i/; = aT4, a =7.061 X 10-15 erg. cm.-a deg.- 4 •
According to the best of recent measurements this value of "'AmT is too high
by some 2 per cent; while the value of a is too low by about 7 per cent. But
from these data one obtains
h = 6.55 X 10-27 erg-second
ko = 1.346 X 10-16 erg-deg. - 1
SEO. 14] PLANCK'S RADIATION LAW 247
The circles show observations by Coblentz 1 on the energy distri-
bution in the spectrum of a black body at 1600°K. The full
line shows the distribution predicted by Planck's formula. The
lower dotted line, which coincides with the full line from short
wave lengths up to about A = 2.2µ, corresponds to Wien's
formula. The upper dotted line is from the Rayleigh-Jeans
formula. The superiority of Planck's formula is, at once, evident.
1 Bull. Bureau of Standards.
CHAPTER VIII
THE QUANTUM THEORY OF SPECIFIC HEATS
We saw in the preceding chapter that the quantum theory
originated in an empirical attempt to bring the then current
theories of black-body radiation into line with experiment.
The attempt was highly successful. But far more striking, and
of vastly greater significance, are the successes which the
postulate
E = hv
has had in other branches of physics.· Indeed, but for these
later successes, it is doubtful whether the theory would have
been of much more than secondary importance. For after all,
from the standpoint of agreement with experiment, the difference
between Planck's formula and that of Wien is not so very great,
as judged from the graphs in Fig. 65 and the footnote on page 246.
We have already mentioned, in Chap. VI, how the concept of
energy quanta seemed to offer the only explanation of the photo-
electric effect, the value of the constant h as determined in this
way being identical with that obtained in Chap. VII from radi-
ation data. At first glance, there does not seem to be a very
close relation between the observed phenomena of photoelectricity
and those of black-body radiation. But that there must be a
very fundamental connection seems to be demonstrated by the
identical numerical values of h. Likewise, neither of these phe-
nomena appears to be very closely related to that branch of
physics which deals with specific heats. Yet in the field of the
specific heats of bodies, the quantum theory had one of its earliest
and, in some respects, most astonishing successes. In this chap-
ter, we shall deal very briefly with this aspect of the quantum
theory, mainly with the purpose of pointing out, as in Chap. VII,
how the newer theory succeeded where the old one partially
failed.
1. The Empirical Law of Dulong and Petit.-Over 100 years
ago, Dulong and Petit,1 as a result of the determination of the
specific heats of a number of elements such as iron, lead, silver,
1 Ann. Chemie et de Physique, vol. 10, p. 395 (1819).
248
SEC. 1] LAW OF DULONG AND PETIT 249
and gold, concluded that "the product of the atomic weight and
the specific heat is the same for all elementary (solid) substances.''
This law is known as the "law of Dulong and Petit." The prod-
uct, atomic weight by specific heat, is known as the atomic heat.
A few years later, Regnault made an extensive investigation
of the specific heats of numerous substances. The mean atomic
heat of some 30 solids-excluding some near their melting points
-was found to be 6.38 calories per gram-atom per degree.
In 1848, Woestyn 1 extended the law of Dulong and Petit to
include molecular heats (i.e., product of molecular weight by
specific heat), on the supposition that each atom in a compound
has the same thermal capacity as when uncombined.
That the law of Dulong and Petit is approximately true for
the atomic heats of many elements in solid form is shown in
Table I, compiled from data given in the International Critical
Tables:
TABLE !.-ATOMIC HEATS OF VARIOUS ELEMENTS IN THE SOLID STATE
Melting Temper-
Atomic point, Atomic ature,
Element
weight degrees heat degrees
Centigrade Centigrade
H ................... 1.008 -259 0.57* -260.6
Li ................... 6.94 186 5.5 0
Be .................. 9.02 1,350 3.85* 0-100
Bo .................. 10.82 2,300 3.35* 0-100
C ................... 12.00 ......... 1.45* 20
N ................... 14.01 -210 5.5 -212
Na .................. 23.00 97.5 6.8 20
Al ................... 26.96 660 5.75 20
Si ................... 28.06 1,420 4.95* 20
Cl ................... 35.46 -101 6.7 -113
Ar .................. 39.91 -189 6.2 -223
Fe .................. 55.84 1,535 5.95 20
Zn .................. 65.38 419 6.0 0-100
Br .................. 79.29 -7.2 5.6 -150
Rb .................. 85.44 38.5 6.85 0
Mo .................. 96.0 2:620 6.2 20-100
Sn .................. 118. 7 232 6.4 18
w................... 184.0 3,370 6.2 20-100
Au .................. 197.2 1,063 6.15 18
Pb .................. 207.2 327 6.35 20
U ................... 238.2 1,850(?) 6.7 0-100
1 Ann. Chemie et de Physique, vol. 23, p. 295 (1848).
250 THEORY OF SPECIFIC HEATS [CHAP. VIII
The atomic heat, in calories per gram-atom per degree, is shown
in the fourth column. The fifth column gives the temperature
at which the measurement was made. Excluding those marked
with an asterisk(*), the mean value of the atomic heats
given in the table is 6.3 calories per gram-atom per degree, the
majority of values being included in the range 5.8 to 6.8. This
grouping of atomic heats around the mean value 6.3 is all the
more remarkable since the elements shown in the table cover a
wide range of properties; e.g., the melting points vary from that of
argon ( -189°C.) to that of tungsten ( +3370°C).
The law of Dulong and Petit can at best, however, be only an
approximation, since the departures from the mean value 6.3
are considerably larger than experimental error. Further, the
values marked with an asterisk-namely, those for hydrogen,
beryllium, boron, carbon, and silicon-are in complete disagree-
ment with the law. These elements, with the exception of
hydrogen, are light elements with hjgh melting points.
2. Variation of Atomic Heats of Solids with Temperature.-
These exceptions to the law of Dulong and Petit were the subject
of extended research, and it was early discovered that at least one
reason for the failure of the law in these several cases was the
fact that the specific heat varies with temperature. This
variation was, indeed, noted by Dulong and Petit themselves as
early as 1819. In 1872, Weber 1 observed that the specific heat
of diamond (i.e., carbon) increases threefold between O and
200°C.
The study of specific heats of substances as a function of
temperature has been greatly facilitated in recent years by the
availability of liquified gases-liquid air, liquid hydrogen, and
liquid helium-for making measurements at low temperatures.
It has been found that below a certain temperature characteristic
of each substance the specific heats of all solids decrease rapidly
with decreasing temperature, reaching a value almost zero even
before the temperature of absolute zero is reached. Figure 66
shows the variation with temperature of the atomic heats at
constant volume of four characteristic substances: diamond,
silicon, aluminum and lead. The data for diamond and silicon
are from a paper by A. Magnus; 2 for aluminum and lead, from
1 Phil. Mag., vol. 44, p. 251 (1872).
2 A. MAGNus: Ann. Physik, vol. 70, p. 303 (1923).
SEC. 2] ATOMIC HEATS AND TEMPERATURE 251
the work of Nernst and Lindemann, quoted by W. M. C. Lewis. 1
The atomic heat of lead at ordinary (room) temperatures is
observed to approximate the value required by the law of Dulong
and Petit. But below a temperature of about 100° Abs., the
atomic heat of lead drops very rapidly toward zero. The curves
for the other three substances have the same general form as that
for lead. They differ from lead mainly (1) in the temperature
above which they seem to obey the law of Dulong and Petit
and (2) in the rate at which the atomic heat decreases below that
00~--=-~-,~oo~-----i-----~-A-~~--J"""------10~00------~
400 600 800 1200
Absolute Temperature (Kelvin)
FIG. 66.-Variation of atomic heat at constant volume with temperature.
t~mperature. In the case of the diamond, the value of (about)
6 calories per gram-atom per degree would not be reached until
the temperature exceeded 2000° Abs.
The curves for nearly all substances are similar to those shown
in Fig. 66, and they lie between the curves for lead on the one
side and for diamond on the other. Indeed, the curves for a large
number of elements lie between those for aluminum and for lead.
Since most of the measurements on which the law of Dulong and
Petit was based were made at room temperature, it is obvious,
from Fig. 66, why the law is so nearly obeyed.
1 LEw1s, W. M. C.: "Quantum Theory," 3d ed., p. 68, (1924).
252 THEORY OF SPECIFIC HEATS [CHAP. VIII
That the curves not only are similar in appearance but also
are very closely related to each other by some fundamental
underlying law is clearly indicated by the fact that they can be
brought (almost) into coincidence by suitable change in the
temperature scale for each curve. Thus, if the abscissre of
the curve for aluminum be multiplied by a factor of 4.6, the
curve for aluminum will nearly coincide with that for diamond.
The crosses ( X) on the curve for diamond represent the points
from the aluminum curve shifted in this way. What is this
fundamental law which governs the variation of specific heat
with temperature and which, at sufficiently high temperatures,
must give, approximately, the law of Dulong and Petit? Classi-
cal physics failed to find the law. The quantum theory was more
successful.
3. The Classical Theory of Specific Heats of Solids.-Accord-
ing to the kinetic theory of matter, the atoms of solids execute
vibrations about positions of equilibrium with amplitudes which
depend on the temperature of the body. With each vibratory
atom is associated energy, part kinetic and part potential, the
two parts being, on the average, equal, provided the restoring
forces are proportional to the displacement so that the vibration
of the atom is simple harmonic. If we were to follow the history
of any particular atom, we should find that its energy varies
greatly from time to time. But its average energy over a long
period of time would be the same as that of any other atom.
Let us consider an enclosure containing a gas and a small
crystalline monatomic solid, all at a temperature T. According
to the principle of the equipartition of energy, discussed in Sec.
11 of Chap. VII, each degree of freedom of the solid should have,
on the average, as much kinetjc energy E as each degree of freedom
of the gas, namely, an amount given by
E = ~2koT
where ko is Boltzmann's constant and T is the absolute tempera-
ture. Each atom of the solid has 3 degrees of freedom by virtue
of its vibratory motion. It therefore possesses kinetic energy
equal to ~1k 0 T. Since with this kinetic energy is associated an
equal amount of potential energy, we see that the total average
energy per atom of the solid is
3koT
SEC. 3] CLASSICAL THEORY 253
Since a gram-atom of the solid contains No atoms, No being
Avogadro's number, it follows that the total thermal energy of
agitation WA of a gram-atom of the solid is
WA= 3koNoT
= 3RT (1)
since koN o = R, the gas constant.
Now, the specific heat of a substance is equal to the energy
required to raise the temperature of 1 gram 1 degree; and, like-
wise, the atomic heat is equal to the energy required to raise the
temperature of 1 gram-atom 1 degree. If, during this rise in
temperature, the substance is not allowed to expand, according
to the kinetic theory of matter all of this added energy goes to
increase the energy of vibration of the atoms. Calling the value
of the atomic heat at constant volume Cv, it follows that
c. = ~~A = 3R (2)
According to classical theory, then, the atomic heat at constant
volume of a monatomic solid should be (1) constant and inde-
pendent of temperature and (2) equal to three times the gas
constant R. Since the numerical value of R is 1.987 calories
per gram-atom per degree (seep. 219), it follows that the numer-
ical value of Cv, on the classical theory of the equipartition of
energy, should be
Cv = 5.961 calories per gram-atom per degree
This is, to a first approxjmation, the numerical value of atomic
heat required by the law of Dulong and Petit. It is close to the
value to which the curves in Fig. 66 approach asymptotically
as the temperature increases. But equation (2) predicts no
variation in atomic heat with temperature, as is required by these
experimental curves.
Suggestions have been made that, perhaps, with decreasing
temperature, the number of degrees of freedom get rapidly less,
due to groups of atoms becoming "frozen" together. Were this
true, the variation of compressibility with temperature should
be other than is observed. It appears, therefore, that the clas-
sical theory, although predicting approximately the numerical
value required by the law of Dulong and Petit, offers no solution
to the variation of atomic heat with temperature. 1
1
See, however, Compton's theory, p. 275.
254 THEORY OF SPECIFIC HEATS [CHAP. VIII
Note.-Relation between GP and Cv.-In general, when bodies,
either solids or fluids, are heated, they tend to expand. If, as
in the case of a gas, the expansion be prevented by the applica-
tion of suitable pressure, the energy given to the gas goes to
increase the kinetic energy of the molecules. We then have to
deal with the atomic heat at constant volume Gv. If, however,
the pressure be maintained constant, additional energy is required
on account of the work, external or internal, or both, which the
expanding body does. We then are concerned with the atomic
heat at constant pressure Gp· As regards these two specific
heats, a solid differs in no way from a gas, except that, in general,
the magnitude of the expansion, for a given temperature change,
of a solid is much less than the corresponding expansion of a gas.
The specific heat of solids must, in general, be measured at
constant pressure. The discussion, however, which led to
equation (2), namely,
Cv = 3R
considered only the increase in the thermal energy of agitation
of the atoms of the solid, no mention being made of external or
internal work resulting from heating. Since the experimental
values of atomic heats of solids are the atomic heats at constant
pressure GP, to compare with experiment the values of Gv, as
predicted, either on the classical theory given above or on the
quantum theory given below, requires that the observed values
of atomic heat, i.e., GP, be reduced to atomic heat at constant
volume, i.e., to Gv. This reduction is accomplished by computing
the work required to compress the solid isothermally back to its
original volume at temperature T, after it has been heated from
T1 to T2. The equation to express the required relation between
Gp and G11 takes the form 1
(3)
where a is the linear coefficient of expansion; Mis the molecular
(or the atomic) weight; K is the compressibility; and pis the
density. Thus, Magnus 2 finds, for diamond,
Gp - Gv = 8.00 X 10-6 XT
1 G. N. LEw1s: (Journ. Amer. Chem. Soc., vol. 29, p. 1165 (1907)),
derives a similar equation.
2 Loe. cit., p. 250.
SEC. 3] CLASSICAL THEORY 255
where the constants in the neighborhood of 50°C. are
a = 1.32 X 10- 6 deg.- 1
K = 0.16 X 10- 12 cm. 2 dyne- 1
p = 3.51 grams cm.- 3
For many substances, these constants a, K, and p, are not
known over the temperature range for which values of CP are
measured. In these cases, use is made of the relation, due to
Griineisen, 1 that at high temperatures the coefficient of expansion
a is proportional to CP, while the product Kp remains practically
constant. Putting
a= aCP
and
Kp = B
where a and B are experimental constants determined at a tem-
perature at which a, K, and pare known, we may write equation
(3),
9a 2M
Cp - C~, =
13 c~T
= AC;T
where
9a 2M
A= 13 (4)
For diamond, A = 2.64 X 10- 6, according to Magnus. The
constant A is, also, a function of the melting point Tm:
1
A= Ao-
Tm
where Ao, as given by Nernst and Lindemann, 2 is a constant and
has the numerical value 0.0214. Table II gives some numerical
values of A.
1Ann. Physik, vol. 26, p. 393 (1908).
2See NERNsT: ''The New Heat Theorem,'' English translation by G.
Barr, p. 65 (1926).
256 THEORY OF SPECIFIC HEATS [CHAP. VIII
TABLE n.-NuMERICAL vALuEs oF A rn cp - c'V = AC;T
Substance A Source
Diamond. . . . . . . . . . . . . . . . . . . . . . . . 0. 264 X 10-5 Magnus and Hodler 1
Boron........................... 0. 75 Magnus and Danz 2
Aluminum. . . . . . . . . . . . . . . . . . . . . . . 2. 2 W. M. C. Lewis 3
Copper .......................... 1.3 W. M. C. Lewis 3
Lead ............................ 3.0 W. M. C. Lewis 3
Platinum ........................ 1.0 W. M. C. Lewis 3
1 Ann. Physik, vol. 80, p. 818 (1926).
2
Ann. Physi'k, vol. 81, p. 422 (1926).
3 Quoted by W. M. C. LEWIS m "Quantum Theory," 3d ed., p. 67.
In Table III is shown the order of magnitude of the difference
between CP and Cv. The difference is small in the case of the
diamond but becomes very important for other substances.
TABLE III.-SHOWING THE ORDER OF MAGNITUDE OF THE DIFFERENCE
BETWEEN Cp AND Cv FOR SEVERAL SUBSTANCES AT DIFFERENT
TEMPERATURES
T,
Substance degrees,
absolute
300 1.522 1.520
Diamond 1 . • • • • . • . • • • • . • • • • • • • • . . • . . . . • 900 4.890 4.833
1,100 5.137 5.060
473 3.782 3.731
Boron 2 .••••••••••••••••••••••••••••••. 873 5.351 5.164
1,173 5.514 5.252
87 3.37 3.35
Copper3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290 5.75 5.60
450 6.03 5.81
23 2.96 2.95
Lead3................................. 200 6.12 5.90
409 6.40 5.94
1 MAG~us and HoDLER, loc. cit.
2 MAGNUS and DANZ, loc. cit.
3 W. M. C. LEWIS, loc. cit.
4. Einstein's Theory of the Atomic Heats of Solids.-(a) An
important advance in the theory of the specific heats of solids
was made by Einstein, 1 who applied the quantum theory to the
vibrations of atoms about their positions of equilibrium. Let
1 Ann. Physik, vol. 22, p. 180 (1907).
SEC. 4] EINSTEIN'S THEORY 257
it be assumed: (1) that the atoms of a monatomic solid vibrate
with a frequency v, the same for all atoms, which depends on the
mass of the atom and on the restoring forces brought into play
when the atom is displaced; and (2) that these atoms, like Planck's
oscillators discussed in Chap. VII, are in equilibrium with ether
Yibrations of the same frequency and that the average energy
[T v (kinetic plus potential) of each degree of freedom of such
vibrations is given by equation (lOOa), Chap VII.
-
uv = ehv/koThv _ 1 [(lOOa), Chap. VII]
Since each atom, by virtue of its vibrations, has 3 degrees of
freedom, the average energy per atom Ea is
3hv
Ea = ehv/koT _ 1 (5)
and the total energy WA of the No atoms in a gram-atom is
- 3Nohv
WA = NoEa = ehv/koT _ 1 (6)
As before (equation (2)), the atomic heat at constant volume is
obtained by differentiating WA with respect to T. Therefore,
dWA 1 hv_
Cv -- dT -- 3N 0hv (ehv/koT - 1)2ehv/koT _
koT2
Multiplying numerator and denominator by ko, we have, after
rearranging and remembering that N oko = R,
Cv = 3R[ (ehv~::tT 1)2 · (!;YJ (7)
Equation (7) is Einstein's equation, based upon the quantum
theory, for the atomic heat of a solid at constant volume. It
differs from the equation (2), given by classical theory,
Cv = 3R (2)
in that the term 3R is multiplied by the quantity within the
square bracket, which quantity is a function of temperature .
.According to the quantum theory, therefore, C'l) is a function of
temperature.
(b) Test of Einstein's Equation.-The experimental curves
shown in Fig. 66 show three main characteristics: (1) At low tem-
peratures, the atomic heat approaches zero; (2) at high temper-
atures, i.e., sufficiently high for each substance, the atomic heat
258 THEORY OF SPECIFIC HEATS [CHAP. VIII
approaches 6.0 calories per gram-atom per degree; (3) between
these limits, the curves rise rapidly with increasing temperature,
approaching the maximum value asymptotically.
Einstein's equation, (7), agrees with experiment as regards the
first two of these characteristics: (1) If T = 0, the quantity
hv/koT becomes infinite, and it may be readily shown that the
square bracket in equation (7) becomes zero. Einstein's equa-
tion, therefore, predicts correctly that the atomic heat should
become zero at absolute zero. (2) If T becomes very large with
6.--------------.---------------.----r------...
3i-----i----+++----+----+-A I um in u m - - - - - - - -
Cv Upper Curve: Debye ~ Formu let
Lower Curve: E/nsfein:S Formula
Circles: Oboerved Poinf.s '
00~.,.__--s~o-----,o~o-----,s~o-----20~0--~-2s~o~---~o~o----~35~0~~40~0
Absolute Temperature
FIG. 67.-Comparison of the specific-heat formulre of Debye and of Einstein
with experiment.
respect to the product hv, hv/koT approaches zero, and the
square bracket approaches unity. At high temperatures, there-
fore, the atomic heat should approach 3R in agreement with
experiment and with the law of Dulong and Petit.
For intermediate temperatures, equation (7) predicts qualita-
tively the variation of atomic heat with temperature but yields
too low values at low temperatures. This is shown in Fig. 67,
in which the lower curve is for the atomic heat of aluminum as
given by Einstein's formula,1 while the circles are the observed
To graph Einstein's formula, it is necessary to determine v. the charac-
1
teristic frequency (see Sec. 6, page 260).
8EC. 5] CHARACTERISTIC TEMPERATURES 259
values of Cv used in plotting the curve for aluminum, in Fig. 66.
In spite of the disagreement, really comparatively small, between
theory and experiment, it is quite remarkable that Einstein's
theory, based upon the quantum hypothesis, should predict
values of Cv so nearly correct. Before considering an improved
formula, due to Debye, it will be desirable to comment upon some
other consequences of Einstein's equation.
5. Characteristic Temperatures.-If, in the equation (7),
Cv = 3R[ (ehv~:::koT I){k:;)2] (7)
such a value of T be chosen as to make koT = hv, h being Planck's
constant and v being constant, presumably, for a particular sub-
stance, the exponent of e becomes unity, and the numerical
value of the square bracket becomes 0.921. For this particular
value of T, call it Tc, the value of Cv becomes 5.49 (i.e., 5.96 X
0.921). Tc, as so defined, is a temperature characteristic of the
substance and, of course, varies from one substance to another.
According to Einstein's theory, we may say that the character-
istic temperature Tc of a substance is determined, for any
particular substance, by the temperature at which the atomic heat
reaches the value 5.49. Referring to Fig. 66, the characteristic
temperatures may be read directly by observing the points
where the horizontal dotted line at the ordinate 5.49 crosses
the several curves. These temperatures as read from the curves
in Fig. 66 are given in Table IV.
TABLE IV.-CHARACTERISTIC TEMPERATURES AND FREQUENCIES (EINSTEIN)
Tc,
v
Substance degrees
sec.- 1
Kelvin
Lead............ 78 1. 61 x 10 12
Aluminum. . . . . . . 307 6.42
Silicon....... . . . . 567 11.9
Diamond 1 . . . • • • • • 1475 30.8
1 Extrapolated.
If, for each substance, we now define a new temperature scale e
in which the value of Tc for that substance is the unit temper-
ature, the values of the atomic heats of all substances would be
the same at equivalent points on these scales. Thus, the atomic
260 THEORY OF SPECIFIC HEATS [CHAP. VIII
heat of lead at ePb = 0.5 (i.e., at 39°K.) should be the same as
the atomic heat of silicon at its temperature 8si = 0.5 (i.e., at
283°K.).
6. Characteristic Frequencies. (a) From Specific Heats.-
From the definition of the characteristic temperature Tc of a
substance, we have
ko
:. v = hTc (8)
Knowing the constants hand ko, and determining Tc, as explained
in Sec. 5, we can compute the value of v, the characteristic f re-
quency with wh1:ch the atoms of the solid vibrate, assuming for the
moment the correctness of Einstein's equation and its funda-
mental underlying concepts. The numerical value of the ratio
ko/h is 2.09 X 10 10 • The values of v computed in this way are
given in the third column of Table IV.
One now raises the question: Are these values of v, computed
from the variation of atomic heat with temperature, values which we
may reasonably expect for the frequencies of (thermal) vibration
of the atoms of crystalline solids? As to magnitude, they are
seen to be of the order of one one-hundredth of the frequencies of
light vibrations in the visible part of the spectrum. The follow-
ing independent methods for determining these frequencies in.
solids lead to values in substantial agreement with those computed
from specific-heat data.
(b) From the '' Reststrahlen.''-One of the most instructive,
and one of the most convincing, of these methods is that due to
Rubens 1 and collaborators, in which they used the phenomenon
of selective reflection for isolating long waves. The reflectivity 2
R of a substance for radiation of wave length X depends on both
the coefficient of absorption 3 of the substance and its refractive
index n at that wave length, the value of R being given by
(1 - n) 2 + n 2K 2
R = (9)
(1 + n) 2 + n 2K 2
1 RUBENS and NICHOLS: Ann. Physik, vol. 60, p. 418 (1897); RUBENS and
KuRLBAUM: Ann. Physik, vol. 4, p. 649 (1901); RUBENS and HoLLNAGEL:
Phil. Mag., vol. 19, p. 761 (1910); RuBENS and VON WARTENBERG: Sitzb. d.
konigl. preuss. Akad., p. 169, 1914.
2 For definition, see Chap. VIII, Sec. 2 (e).
3 Not "absorptivity" (see p. 186).
SEC. 6] CHARACTERISTIC FREQUENCIES 261
where K is a quantity which is known as the "extinction coeffi-
cient" and which is closely related 1 to the coefficient of absorption
µ, defined on page 186. Any substance which exhibits selective
absorption in any spectral region presumably has resonating
mechanisms whose natural periods coincide with the central part
of the absorption band. In those regions, the value of K and,
therefore, of nK may become large enough compared to (n - 1)
to raise the value of the reflectivity R considerably above its
value for neighboring spectral regions. The substance then
exhibits selective reflection. Confirmation of this relation
between selective absorption and selective reflection is found in the
work of Nichols and of Rubens on quartz. Quartz is transparent,
or nearly so, up to about 7.5µ but has a very strong absorption
band (really, two bands near together) at about 8.5µ. At 8.5µ,
quartz is found to reflect 80 per cent, while, at 4µ, where it is
quite transparent, it reflects only a few per cent.
If a beam of light of wave length X and intensity Io passes through a thin
1
slab of absorbing material of thickness d, the intensity I on emerging 1s
given by
-4,rK d
I = Ioe A
where K is the extinction coefficient. Let the thickness of an absorbing layer
be equal to X so that d/X = 1. Then
I1 = Ioe- 4,r"
where I 1 is the emergent intensity after passing a thickness equal to one
wave length. We may then write
or
1 Io
" = 41r loge Ii
Unless the substance is so strongly absorbent that the ratio Io/ I 1 is appreci-
able, K is a very small quantity, and may be neglected in computing R by
equation (9). Thus for glass for which nK < < (1 - n), at least in the visible
. (1 - n) 2
reg10n R = · But for most metals in the visible region of the
' (1 + n) 2
spectrum " may be considerably larger than n. Thus Drude (" Theory of
Optics") gives the product nK for platinum as 4.26, while the value of n is
2.06. These values, put in equation (9), give for the reflecting power of
platinum about 0.70. Hagen and Rubens (Ann. Physik, vol. 1, p. 352
(1900)) found the reflecting power of platinum, at X = 0.65µ, to be 0.66.
A derivation of equation (9) may be found in Woon's "Physical Optics"
or in any textbook on the electromagnetic theory of light.
262 THEORY OF SPECIFIC HEATS [CHAP. VIII
Conversely, if a substance is found to exhibit selective reflection
in any region, the substance must have resonators whose natural
frequency coincides nearly 1 with the frequencies which are
selectively reflected. If a continuous spectrum be reflected
from a substance which shows selective reflection at a wave
length X0 , the reflected beam will be relatively much richer in
radiation of that wave length than was the incident beam. After
several such reflections, the radiation may be comprised almost
entirely of the wave lengths which are selectively reflected, the
remainder having been almost completely absorbed. The
radiation remaining after several such reflections is called ''residual
rays" or "reststrahlen."
By this method of'' reststrahlen,'' Rubens and his collaborators
succeeded in isolating the residual rays from a number of sub-
stances. Some of their results are shown in the second and third
columns of Table V. For purposes of comparison, there is
TABLE V.-RESIDUAL RAYS FROM CERTAIN SOLIDS
Computed from Computed from
Einstein's formula the Nernst-
\Vave Frequency, Lindemann for-
Substance length sec.- 1 mula, fre-
Frequency, quency,
sec.- 1 sec.- 1
NaCl. ......... . 52. Oµ 5. 77 X 10 12 220 4.60 x 10 1 2 5.55 x 10 12
KCl ........... . 63.4 4. 73 175 3. 66 4.55
KBr .......... . 82.6 3.64 155 3.24 3.87
KI ............ . 91.4 3.28
CaC03 . . . . . . . . . 98. 7 3.04
AgBr ...... .' ... . U2.7 2.66
given, in the fifth column, for the three substances NaCl, KCl,
and KBr, the frequency determined by the specific-heat method
(Einstein's formula), m·z.: From daia quoted by Lewis 2 the atomic
heat was plotted as a function of temperature, and the value
of Tc was determined in the same way as in Fig. 66 and Table
IV. These values of Tc are given in the fourth column of Table
V. The frequency v was then computed by equation (8).
1 "Nearly," because R depends not only on K but on n, and n varies rapidly
in the neighborhood of an absorption band.
2 "Quantum Theory,'' 3d. edition, p. 70.
SEC. 6] CHARACTERISTIC FREQUENCIES 263
While the agreement in these three cases is not quantitative
within the limit of error of experiment, it is excellent as far as
concerns order of magnitude and relative values from one sub-
stance to another. The corresponding frequencies, computed
from the Nernst-Lindemann formula, to which reference will be
made later, are given in the last column. These are seen to be
in excellent agreement with the values obtained by the method of
reststrahlen, which method comes as near as one could wish to a
direct method of observing the frequencies of atomic vibrations
in solids.
(c) From Compressibilities.-Besides this method of deter-
mining the frequency, based on Einstein's formula, or on the
Nernst-Lindemann formula, there are several indirect methods
which give reasonably concordant results. Einstein 1 has
derived a relation which makes it possible to compute the fre-
quency of the atomic vibrations from the compressibility K,
the atomic weight A, and the atomic volume VA of the substance.
The following are the underlying principles of this method:
Solids are, with few exceptions, crystalline--a concept abundantly
confirmed by X-ray analysis. In each crystal, the atoms are
arranged in a regular "lattice," each atom being held in a position
of equilibrium by forces of some kind due to the presence of
neighboring atoms. These forces are the same on each atom
throughout the entire crystal. They must, also, be intimately
connected with the compressibility of the crystal; the greater
these forces the less the conwressibility. Let the mass of an
atom be m ai;id the force per unit displacement acting on it when
displaced 'from its equilibrium position be fa. Then, according
to the ordinary laws of mechanics, we should have the period
of vibration T of the atom given by
T =27!" ~R (10)
if the return force be proportional to the displacement. The
frequency of vibration is
v = __!__ /Jo (11)
21r"J m
The compressibility is defined as the ''fractional decrease in
volume per unit increase in (applied) pressure." The compres-
sibility should be (1) greater the less f o and (2) greater the greater
1 Ann. Physik, vol. 34, pp. 170, 590 (1911).
264 THEORY OF SPECIFIC HEATS [CHAP. VIII
the average distance d between the atoms of the crystal. It is
reasonable to aRsume-although a rigorous proof would have to
justify the assumptions-that
1
1. K a: fo
2. K a: d
or,
d
K = C- (12)
fo
where C is a proportionality constant. If VA is the volume of a
gram-atom, and if we assume a regular arrangement of the atoms
of the crystal, we may determine d by
d = /VA
3
(13)
~ No
when N O is Avogadro's number. Putting this value of d into
equation (12) and solving for Jo, we have
Jo = C 3/VA (14)
K~No
We then have, for the frequency v, by equatjons (11) and (14),
(15)
(16)
by multiplying numerator and denominator under the radical by
N5 and remembering that A = mNo where A is the atomic
weight. Einstein evaluates the constant C by considering the
lattice array of atoms. He found the numerical value of C' to
be 2.8 X 10 7 • This gives
v = 2.8 X 10 7 · ~ (17)
which is one form of Einstein's equation for the atomic fre-
quencies based on compressibilities. Some values of frequency
~etermined by Einstein in this way are given in Table VI.
SEC. 6] CHARACTERISTIC FREQUENCIES 265
TABLE Vl.-SoME OF EINSTEIN's VALUES OF v FROM CoMPRESSIBILITIES
Frequency,
Substance
sec.- 1
Al ............... . 6.6 x 10 12
Cu .............. . 5.7
Pt ............... . 4.6
Au .............. . 3.8
Pb .............. . 2.2
Bi ............... . 1.8
The values for aluminum and lead are seen to compare favorably
with those given in Table IV.
(d) From Melting Points.-This method of computing fre-
quencies is due to Lindemann. 1 It is based on the assumption
that the melting point Tm of a solid is the temperature at which
the amplitude of vibration of the atoms is equal to (" of the same
order of magnitude as" would be preferable) the average distance
apart d of the atoms. We can determine the way in which the
several quantities enter into Lindemann's formula as follows:
As before (equation (11)), let
(11)
where vis the frequency, Jo is the restoring force per unit displace-
ment of the atom, and m is its mass. If the amplitude of vibra-
tion is d, the potential energy at maximum displacement is
~2 Jod 2• This is, also, the (time) average Ed of the energy of the
atom for the amplitude d. Therefore,
Ed = ~2 Jod 2
.. . Jo=~
2Ed (
18
)
Putting this value of Jo into equation (11) and solving for Ed gives
(19)
Now, the melting point Tm is usually much higher than the
characteristic temperature Tc, so that the energy of an atom due
to its vibratory motion at temperature Tm is given by the classical
value 3koT m· Therefore,
21r 2 v2md 2 = 3koT m (20)
1 Phys. Zeitsch., vol. 11, p. 609 (1910).
266 THEORY OF SPECIFIC HEATS [CHAP. VIII
As before (equation (13)), let
3
d = ~ [13]
\)~
where VA is the volume of a gram-atom at temperature Tm. We
now have a value of v from equation (20), given by
v = I [3ko 31 T~ NO
1r\)2 \Jrn VA 72
Since A = rnN 0 (A = atomic weight), we have, finally, for v,
V = [I1r\){3k;N%]
2·
• / Tm
\)AV~
0
(21)
The quantities within the square bracket are all constant, so that
the equation for v is of the form
v = C~A 1;7] (22)
C is, however, determined from comparison with experiment
rather than from ko and No. The equation (22) is, therefore, of
value only in determining the way in which the frequency depends
on the melting point Tm, the atomic weight A, and the atomic
volume VA at the melting point. C is found to be of the order
of magnitude of 2.8 X 10 12 • Some values of v computed by
Nernst and Lindemann from this formula are given in the
following table:
FREQUENCY BY LINDEMANN's "MELTING-POINT" FORMULA
Tm,
v,
Substance degrees
sec. - 1
Kelvin
Al.............. 930 7. 6 x 10 12
Cu............. 1,357 6.8
Diamond. . . . . . . 3 , 600 ( ?) 32.5
Pb............. 600 1.8
(e) These several methods of determining the frequency of
vibration of the atoms of solids, although differing widely, yield
values so nearly in accord with each other as to leave little doubt
that the general underlying principles are, in each case, substan-
tially correct. The first method (using Einstein's specific-heat
formula) is based on the quantum theory as introduced by
SEC. 7] THE NERNST-LINDEMANN FORMULA 267
Planck. The others (using, respectively, the reststrahlen, com-
pressibilities, and melting points) are quite independent of the
quantum theory.
Biltz, 1 by use of the melting-point formula of Lindemann, has
computed the frequencies of a large number of the elements in
the solid state and finds that the frequencies are periodic func-
tions of atomic weight.
7. The N ernst-Lindemann Formula for Atomic Heats.-It
was pointed out in Sec. 4 (Fig. 67) that Einstein's formula (equa-
tion (7)) predicts values of specific heat which are considerably
too low at low temperatures. Nernst and Lindemann 2 proposed
an empirical modification of Einstein's expression, which was
based upon the assumptions (1) that there are two, instead of one,
characteristic frequencies in a solid, namely, a frequency v and
its first lower octave v/2; and (2) that one half of the atoms
vibrate with the one frequency and the other half with the other
frequency. Applying Einstein's formula to each group of atoms
the total atomic heat Cv becomes
c,, =
3 [
2~
ehv/koT
(ehv/koT_
( hv
1)2 koT
)2 + ehv/2koT
(ehv/2koT -
( hv
1)2 2koT
)2] (23)
The assumptions on which the formula is based can hardly be
justified, yet the formula agrees remarkably well with experi-
ment. In Table VII are shown the observed values of the atomic
heat of aluminum (quoted by LEWIS, Quantum Theory, p. 68)
TABLE VIL-COMPARISON OF THE NERNST-LINDEMANN AND OF THE
EINSTEIN FORMULAS WITH EXPERIMENT. Cv OF ALUMINUM.
Cv computed
Temperature,
Cv observed
degrees Kelvin Nernst-
Einstein
Lindemann
32.4 0.25 0.22 0.03
35.1 0.33 0.32 0.05
83.0 2.41 2.40 2.12
86.0 2.51 2.51 2.26
137 3.91 3.99 3.98
235 5.17 5.14 5.17
331 5.58 5.51 5.54
555 5.98 5.79 5.81
1
Quoted in LEWIS: "Quantum Theory," 3d ed., p. 63.
2 Sitz. d. konigl. preuss Akad., p. 494 (1911).
268 THEORY OF SPECIFIC HEATS [CHAP. VIII
and those computed by the Nernst-Lindemann formula, together
with, for comparison, the values predicted by Einstein's formula.
In spite of the superiority of the Nernst-Lindemann formula,
its empirical nature robs it of much of its importance. It does,
however, call attention to a very important point in connection
with the quantum theory of specific heats. Einstein's equation
presupposes a single frequency characteristic of each solid, an
assumption which is hardly tenable, even in the case of mona-
tomic solids, and is inadmissable in the case of such diatomic
compounds as NaCl, KBr, etc. That the Nernst-Lindemann
formula is more successful by assuming two frequencies suggests
the direction in which improvement in theory may be expected.
This extension was made by Debye.
8. Debye's Theory of Atomic Heats. 1-(a) Let us postulate,
for the moment, a solid the molecules of which have no thermal
vibrations but are at rest in their respective positions of equilib-
rium. Now, let a system of standing waves, say longitudinal,
be set up in the solid. Each atom is now vibrating with an
amplitude which depends on the position of the atom with respect
to the nodes and loops of the wave system. If we superpose
more and more standing-wave systems of both the same fre-
quency as the original and of different frequencies, the vibrations
of any particular atom become more and more complex, until,
finally, we may approach a condition of atomic agitation similar
to the temperature vibrations. Conversely, we may think of the
temperature vibrations of the atoms of a solid as being the equiva-
lent of a vast complex of standing waves of a great range of
frequencies. The range of frequencies, however, does not extend
from zero to infinity. The lower limit to the frequency is the
"fundamental" frequency of the solid-just as in the case of a
violin string or an organ pipe. The highest frequency is the
frequency of the shortest waves which the solid can transmit.
These latter wave lengths are of the order of magnitude of the
distance between atoms.
In elastic solids, two distinct systems of waves are recognized:
(1) longitudinal waves, velocity VL and (2) transverse waves,
velocity vr. These velocities are determined by the elastic
constants of the material and its density.
We may think of these standing-wave systems, just as in the
case of ether waves, as making up a great number of degrees of
1 Ann. Physik, vol. 39, p. 789 (1912).
SEC. 8] DEBYE'S THEORY 269
freedom. In Chap. VII, Sec. 13, we discussed the number of
such degrees of freedom per unit volume. We showed that the
number of degrees of freedom per unit volume d<h of longitudinal
waves in the wave-length range X to X + dX is (see equation
(85), Chap. VII)
dX
d<h = 41r X4
and for transverse waves (see equation (86), Chap. VII),
dX
dcpT = 81rX 4
Changing wave length X to frequency v by the relations X = v/v
and dX = (v/v 2 )dv, these equations become
v2
dcpL = 41r-
3 dv
VL
v2
dcpT = 81r-
3 dv
VT
The total number of degrees of freedom dcp in unit volumes of the
solid in the frequency range v to v + dv is the sum (d<PL + d¢T),
since both systems of waves are simultaneously present.
Therefore,
dcp = dcpL + d¢T = 41r(~
VL
+ ~)v dv
VT
2
(24)
(b) Now, let us assume that with each of these degrees of
freedom of frequency v is associated an average amount of
energy Up (see equation (100), Chap. VII, where e = hv) equal to
- hv
u p
=--
ehv/koT _ 1
The energy dW A in all the degrees of freedom in this frequency
range v to v + dv in a volume VA equal to the volume of a
gram-atom is
dW A = VAUpdcp
(1 2)
hv 2
= 4,r VA vi+ v~ ehv/koT - 1 v dv (25)
The total energy WA of the gram-atom, taking into account all
frequencies, is
W A -_ fdW A -_ 4,rVA (1vi+ v~2)f Pm
0
hv
ehv/koT_
2
Iv dv (26)
270 THEORY OF SPECIFIC HEATS [CHAP. VIII
In this integration, the lower limit is zero, since the fundamental
frequency of any solid of experimental dimensions is very small,
practically zero. The upper limit vm is the maximum frequency
of vibration which the solid is capable of transmitting. This
frequency is assumed to be of the order of magnitude of the
natural frequency of the atoms as determined by, say, the
Lindemann melting-point formula.
The total number of degrees of freedom </> in the gram-atom
must be the same whether we compute </> by integration, within
proper limits, of equation (24) or assign to each of the No atoms 3
degrees of freedom. Therefore,
o VA(\+ 3) Jo
3N = 41r
VL
2
VT
4
(1'm v2dv = 1r 3
VA(\+ 3)v!
2
(27)
VL VT
:. 9~o = 41rV.-1(\ + VT23)
Vm VL
(28)
Since the velocities are computable from the elastic constants,
equation (28) makes it possible to compute vm. Better values of
Cv, however, are obtained 1 in the formula for Cv derived below
(i.e., equation (35)) by computing vm from Lindemann's melting-
point formula (equation (22) ).
By use of equations (27) and (28), we may now write for WA,
in equation (26),
_ 9Nolvm hv 2
(29)
WA - 7 ehv/koT - 111 dv
m O
(This equation in Debye's theory corresponds to equation (6)
in Einstein's theory.)
To obtain the atomic heat at constant volume Cv, we wish to
compute from equation (29) the value of dWA/dT. This
derivative is usually, and more easily, obtained by a change of
variables, viz.: Let us define the characteristic temperature Tc
in Debye's theory in a manner similar to that followed in Ein-
stein's theory (sec. 5) by the relation
T = hvm (30)
c ko
vm being the maximum frequency. Also, put
hv
x koT =
koT
:. dv = hdx (31)
1 LEw1s: "Quantum Theory," 3d ed., p. 330.
:'EC. S] DEBYE'S THEORY 271
The necessary changes in limits are
when v = 0, x = 0
hvm Tc
wh en v = vm; X = Xm = k/I' = T (32)
\\~ith these changes in the variable from v to x, we rewrite
equation (29),
T4lTc/T x3
WA = 9Nokt-- dx (33)
T~ o e:z:- 1
,Ve may now compute Cv:
dWA [ T3 rTc/T x3
C. = dT = 9Noko 4T~ e"' _ 1dx Jo
T4f Tc/T( 3x2
- Tt O
x3ex ) x
e"' - 1 - (e"' - 1) 2 Tdx (34)
J
The second integral at once becomes, after integration,
T3[ x4 ]Tc/T
T~ ex - 1 0
After substituting the limits in this integral, we finally have,
for Cv,
T 3 Tc/T
c'V = 9Noko[4(-T) X
3
c
I
dx - Tc _ 1 _ J
O
(35)
ex - 1 T eTc/T - 1
This is De bye's equation for the specific heat of a solid at constant
volume.
(c) Test of Debye's Equation.-(1) At high temperatures
Debye's equation should yield the classical value of Cv, namely,
3Noko or 3R. For large values of T, x becomes small, and
Tc/T becomes small. Since ex = 1 + x + · · · , in which
expansion for small values of x we may drop x 2 and higher terms,
we have, for high temperatures,
1. ~c •,eTc// - 1 = 1
2.
(T)3iTc/T 3 dx = 4(T
4 -
Tc ex - 1
O
-
Tc
X) f Tc/Tx dx = -4
3
3
O
2
(36)
:. Cv = 9Noko(~~ - 1)
= 3Noko = 3R
This is in agreement with the law of Dulong and Petit and with
experiment.
272 THEORY OF SPECIFIC HEATS [CHAP. VIII
(2) At low temperatures, x and Tc/T are both large. The
second term in the square bracket of equation (35) vanishes.
The integral, now taken between the limits O and oo, becomes 1
6(1 + l_24 +- _!_34 + · · · ) .
Therefore, since
(1 + ;4 + 3\ + · · · ) = 1.082
Cv = 9Nok{ 4(f )3 X 6·1.082 J
= 234NokoT 3
T~
= (constant) X T 3 (37)
Cv
0.4 t - - - - - - + - - - + - - - + - - - 4 - - - , 6 - + - - - " 7 " " ' - - - - t - - - - t - - - - - - + - - - - ;
4-0 60 80 100 l'ZO 140 160 180
T 3xl0-3(Kelv·,n sca\e)
FIG. 68.-At low temperatures, the molecular heat at constant volume is pro-
portional to the cube of the absolute temperature.
At low temperatures, therefore, the atomic heat of a solid should
be proportional to the cube of the absolute temperature. This
is confirmed by experiment over a considerable range of temper-
atures. Schroedinger 2 quotes extensive data bearing on this
point, curves for several of the substances given in his table being
shown graphically in Fig. 68, where Cv is plotted against the cube
of the absolute temperature for CaF2 (17 to 40°K.), FeS2 (22 to
1 See a similar evaluation on p. 244.
2 Phys. Zeitsch., vol. 20, p. 498 (1919).
~EC. 9] FURTHER CONSIDERATIONS 273
.j·t)K.), and Si (20 to 53°K.). The proportionality is seen to
hold from almost vanishingly small values of Cv, well up toward
C_. = 1. Beyond this point, as theory requires, the curves
become increasingly convex toward the T 3-axis.
(3) For intermediate temperatures, it is necessary to evaluate
Debye's formula by the summation of series. 1 These computa-
tions are now available in the form of tables 2 which give Cv as a
function of Tc/T. The upper curve of Fig. 67 is the graph of
Debye's formula for aluminum, taking Tc = 398. The agree-
ment between the graph and the observations is seen to be
excellent. 3
(4) From equation (28), the maximum frequency vm should be
given by
9No
3
vm= (1-+-2)
41rVA v3 vs (38)
L T
It is possible to compute the velocities VL and vr from the elastic
constants of the solid. H. S. Allen 4 has calculated vm in this
way for a number of elements, several of which are given here:
ELEMENT Pm
Al .................................. . 8. 26 X 10 12 sec. - 1
Cu .................................. . 6.81
Ag .................................. . 4.39
Au ........................... ········ 3.44
Pb ..................... ·.············ 1.49
There is reasonably good agreement between these values and
those found by other methods.
In short, Debye's theory of atomic heats is in excellent agree-
ment with experiment over a very wide temperature range.
9. Further Considerations.-Excellent as it is in so far as
agreement with experiment is concerned, De bye's theory is by no
means the "last word" in the theory of atomic heats. For a
more detailed discussion, the reader is referred to more exten-
sive treatises on the quantum theory. Here we can make brief
reference only to two or three further considerations.
1 For details see De bye's paper Zoe. cit.: Ann. Physik, vol. 39, p. 789
(1912).
2 NERNST: "The New Heat Theorem," pp. 246-254.
3 LEw1s' "Quantum Theory" gives numerous tables showing agreement
between observed values of Cv and those computed from Debye's formula.
4
Proc. Roy. Soc., vol. 94, p. 100 (1917).
274 THEORY OF SPECIFIC HEATS [CHAP. VIII
(a) The Theory of Born and Karman. 1-This theory, developed
by Born and Karman, simultaneously with, but independent of,
the theory of Debye, differs from Debye's theory in that the
number of free vibrations in a crystalline solid is computed by
considering the arrangement of the atoms of the crystal in a space
lattice or in an interpenetrating series of such lattices. Although
the theory meets several formal objections inherent in Debye's
theory, the resu~ting formulre and their interpretation are much
more complex than those of Debye and are, therefore, quite
beyond the scope of an elementary text.
(b) Specific Heats of Very Low Temperatures.-Schaefer 2
points out that below a certain frequency vp it is inadmissible
to compute frequencies by integration from zero as a lower
limit, as was done in equation (27), since in this region of very
low frequencies the "distribution" of frequencies is not suffi-
ciently continuous. Rather, from the fundamental frequency
(or from zero frequency) up to some frequency vp, one must use a
summation rather than an integration. From vp up to vrn., the
integral is applicable. The result of this formal change is of
importance only at very low temperatures. Schaefer concludes
that there is for each substance a very low temperature To
below which Debye's T 3 law does not hold. From absolute
zero up to To, the variation of Cv with temperature should follow
an exponential law.
(c) Specific Heats at Very High Temperatures.-The several
formulre discussed above approach asymptotically at high tem-
peratures the classical value Cv = 3R = 5.961 calories
per gram-atom per degree. It has long been known, however,
that at very high temperatures the atomic heat considerably
exceeds this value. The atomic heat of Fe at 1200°C. is 9.6;
of Sn at 1100°C., 9.2; of copper at 900°C., 7.1. Magnus and
Danz 3 find that the atomic heat of tungsten increases linearly
from 7.092 at 400°C. to 7.506 at 900°C. These high values are
explained by assuming that with higher temperatures an increas-
ing number of degrees of freedom come into play. If any atomic
vibrating mechanism in any system has a natural frequency
vi, the quantum of energy for which is hvi, that mechanism
will, in general, not possess its full average quota of energy
1 REICHE: "The Quantum Theory," English translation, pp. 42-58.
2 Zeit. fur. Physik, vol. 7, p. 287 (1921).
3 Ann. Physik, vol. 81, p. 407 (1926).
SEC. 9] FURTHER CONSIDERATIONS 275
unless the average energy per degree of freedom of the system
with which the mechanism is associated equals or exceeds hvi.
Taking the classical value of koT as the energy per vibration
(the quantum theory is in agreement with this at high temper-
atures), the temperature of the system below which the frequency
vi will be '' partially suppressed'' is determined by
koT = hvi
When the temperature approaches or exceeds this limit the
degrees of freedom of this frequency vi become increasingly
active; the total number of degrees of freedom of the system
increase; and, accordingly, the specific heat increases. It is
reasonable to assume that these additional degrees of freedom
which become active at high temperatures in solids are due to
electrons, which, in part because of their small mass, have higher
frequencies than the atoms.
(d) Extension of the Classical Theory.-In Sec. 3, it was shown
that the value of the atomic heat of solids demanded by the
classical theory is 3R, independent of temperature. It has,
however, been shown by A.H. Compton,1 for the region of lower
temperatures, and by Adams, 2 for the region of higher tempera-
tures, that formulre can be developed without the aid of the
quantum theory and by use of very reasonable assumptions,
which are in good agreement with experiment.
Compton assumes that
. . . if the relative energy between two neighboring atoms in a
solid falls below a certain critical value, the two atoms become" agglom-
erated'' so that the degree of freedom between them vanishes; but as
soon as the energy increases again above the critical value the degree of
freedom reappears.
This means that as, with decreasing temperature, the thermal
agitation becomes less, more and more degrees of freedom vanish
and the atomic heat becomes less. From this postulate, Compton
deduces an equation for C11:
Cv = 3Nokoe -- T,/T(~ + 1) (39)
1 Phys. Rev., vol. 6, p. 377 (1915).
2
ADAMS, E. P.: "The Quantum Theory," Nat. Research Council, Bull.,
vol. 7, part 3 (November, 1923).
276 THEORY OF SPECIFIC HEATS [CHAP. VIII
Tc is, as in the other theories, a "characteristic" temperature
defined by the equation
e
Tc = 2ko
where e is "the critical value of the energy below which a degree
of freedom remains agglomerated.'' This characteristic tem-
perature in Compton's theory is the temperature at which 1
2
Ci = 3Noko _
2 71
= 4.40
Several such temperatures, determined by Compton are as follows:
Tc,
SUBSTANCE DEGREES KELVIN
Diamond ............................. . 794
Cu ................................. ··· 133
Al .......................... ·.········ 170
Pb ............................. ······· 38.4
Compton's formula agrees well with experiment, as Table VIII,
taken from his paper, shows.
TABLE VIII.-ATOMIC HEAT BY COMPTON'S FORMULA
T, Cv, T, Cv,
Cv, Ci,,
(degrees (calcu- (degrees (calcu-
(observed) (observed)
Kelvin) lated) Kelvin) lated)
Diamond Aluminum
II
88 0.03 0.01 32.4 0.25 0.19
205 0.62 0.63 83 2.40 2.40
222 0.76 0.80 137 3.91 3.94
306 1.58 1.60 235 5.17 5.09
413 2.64 2.62 331 5.58 5.55
1169 5.24 5.18 555 5.98 5.85
Copper Lead
II
23.5 0.22 0.14 23.0 2.95 3.04
33.4 0.54 0.56 28.3 3.91 3.69
88 3.37 3.36 36.8 4.38 4.38
137 4.53 4.54 38.1 4.43 4.45
290 5.66 5.60 90.2 5.63 5.67
450 5.87 5.86 200 5.91 5.98
1 Put T = Tc in equation (39).
SEC. 9] FURTHER CONSIDERATIONS 277
In a critical comparison of the predictions of Compton's
formula with experimental results, however, F. Schwers 1 con-
cludes that at very low temperatures Compton's formula gives
too low values of Cv; that De bye's formula is much superior.
This is shown by Table IX, for the specific heat of aluminum,
taken from Schwers' paper.
TABLE IX.-COMPARISON OF THE FORMULAE OF DEBYE AND OF COMPTON
WITH EXPERIMENT
Cv computed
Tempera t ure, Cv observed
degrees Kelvin Debye: Compton:
Tc = 382 Tc = 164
19.1 0.066 0.058 0.010
23.6 0.110 0.109 0.045
27.2 0.162 0.166 0.101
33.5 0.301 0.310 0.260
37.1 0.396 0.409 0.390
41.9 0.597 0.613 0.586
49.6 0.901 0.907 0.939
53.4 1.105 1.084 1.126
62.4 1.543 1.520 1.561
73.4 2.070 2.060 2.077
79.l 2.345 2.362 2.303
Adams 2 , by the methods of classical statistical mechanics,
finds that at high temperatures Cv should be a linear function of
T, viz.,
Cv = 3N oko(l + akoT) (40)
where a is a constant (independent of the quantum constant hand
not empirical). Adams shows that this same formula may also
be derived from quantum-theory considerations. Equation
(40) is seen to be in agreement with the data on tungsten given
by Magnus and Danz 3 , in so far as the linear variation of Cv
with temperature is concerned.
That equation (40) is obtainable on the basis of either the clas-
sical or the quantum theory is, perhaps, only an extension of the
agreement between the two theories in predicting the value 3R
1 Phys. Rev., vol. 8, p. 117 (1916).
2 Loe. cit.
3 Loe. cit., p. 274.
278 THEORY OF SPECIFIC HEATS [CHAP. VIII
for the specific- heat at (moderately) high temperatures. The
difference between the two methods of deriving the equation lies
in the fundamental concepts involved. Perhaps we shall ulti-
mately come to understand that even these fundamental con-
cepts are in harmony with each other.
10. The Atomic Heat of Gases. (a) Classical Theory.-The
problem of the specific heats of gases is somewhat more com-
plex than the corresponding problem for solids. The atoms of
the latter have freedom of motion in respect of vibratjon only.
The atoms of a gas, however, have, in general, at least three
types of motion-translation, rotation, and, for polyatomic gases,
vibration of the atoms of the molecule with respect to each other.
(In addition, we have, of course, electronic vibrations, which, for
the present, we shall not consider.) With each one of these
types of motion we should expect to find a quantity of energy
depending on the number of degrees of freedom of each type.
The total energy WM of a gram-molecule at temperature T
should, therefore, be
(41)
where Wt, Wr, and Wv stand, respectively, for the energy per
gram-molecule associated with translation, rotation, and
vibration.
As regards translation, there should be 3 degrees of freedom
per molecule, whether the molecule be monatomjc or polyatomic.
Assuming }2koT of kinetic energy per degree of freedom, we
should have
Wt = No · 3 · ~koT = ~2RT (42)
As regards rotation, we might expect to find, for a monatomic
molecule, 3 degrees of freedom, since such a molecule should be
capable of rotation about each of three mutually perpendicular
coordinate axes. This should add, per molecule, Wr = %RT of
kinetic energy. Since the single atom of a monatomic molecule
of a gas can have no vibration with respect to any other molecule
Wv should be zero and, therefore, WM, for a monatomic molecule,
should be
WM= Wt+ Wr = ~'2RT + ~'2RT = 3RT
and, for the specific heat at constant volume, we should have
C,v = 3R = 5.96 calories per mole
SBC. 10] ATOMIC HEAT OF GASES 279
The values of Cv for the monatomic gases He and Ar at 15°C.
are,1 however,
He · · · Cv = 2.96 calories per mole
Ar · · · Cv = 2.90 calories per mole
These values are seen to be almost exactly half the value predicted
above. Since we know that the molecules of a gas have transla-
tion, it is usually assumed that, for some reason, the rotation of
the molecule, if rotation exists, does not contribute to the energy
of the molecule and, therefore, that the total kinetic energy is
that due to translation only. For a monatomic molecule,
therefore,
w M = Wt = ~f2RT = 2.98 . T (43)
from which Cv = 2.98, in agreement with experiment.
We shall discuss, later, the apparent reasons for the "suppres-
sion" of these rotational degrees of freedom. For the present,
suffice it to remark that this absence, at ordinary temperatures,
of the rotational energy of the molecule is in harmony with
Rutherford's nuclear type of atom, according to which almost
the entire mass of the atom resides in the nucleus, or central
"sun," the dimensions of which are very small compared to the
dimensions of an atom. The moment of inertia of the atom
about any axis through its center of mass must, then, be due, in
very large part, to the moment of inertia of the electrons. For
argon, for example, which has 18 electrons at a mean distance
from the nucleus of the order of, roughly, 10-s cm., the moment
of inertia I would be (I = "2;mr 2 ) of the order of 10- 42 , the mass
of the electron being 9 X 10- 28 grams. The energy of a degree of
freedom at 300°K. is (E = }1koT), about 2.5 X 10- 14 ergs.
For an argon atom to possess this kinetic energy due to rotation
would require that its speed of rotation should be of the order of
10 13 revolutions per second. This is well toward the frequency
of visible light. According to the classical theory, then, argon at
300°K. should emit radiation, due to the rotation of the atom,
of frequency 10 13 , which is in the comparatively near infra-red
just beyond the visible spectrum. No such radiation from
argon is observed. The absence of this radiation does not explain
why the argon atom does not rotate. It only confirms the specific
heat observation.
1 International Critical Tables.
280 THEORY OF SPECIFIC HEATS [CHAP. VIII
For a diatomic molecule, we have, for translation, Wt =
~~NokoT per gram-molecule. Assuming the molecule to be
dumb-bell shaped, rotation about the line joining the two nuclei
should contribute no energy, for the same reason as in a mona-
tomic molecule. Rotation about each of two axes passing
through the center of gravity of the dumb-bell and at right angles
to each other and to the axis of the molecule involves, however, a
much larger moment of inertia. Thus, for 02, atomic mass
26 X 10- 24 grams, distance between atoms of the order of 10-s
cm., the moment of inertia is of the order of l0- 39 • The fre-
quency of rotation required for a kinetic energy equal to that of a
degree of freedom at 300°K. is of the order of 10 12 revolutions per
second, which, for radiation, corresponds to a wave length of 0.1
mm. We shall see, from later considerations based on the quan-
tum theory, that this frequency of revolution is permissible. A
diatomic molecule may, therefore, have 2 degrees of freedom as
regards rotation.
As regards vibration, a diatomic molecule should have 1
degree of freedom. With the kinetic energy ~2koT, due to vibra-
tion, is associated an equal amount of potential energy.
The total energy WM of a gram-molecule of a diatomic gas
should be given by
WM = J~RT + %RT + 7~RT = Y2RT
Therefore, Cv should be 7AR, or 6.95 calories per mole per
degree. The following data, however, give the molecular heat
of four common, diatomic gases at 15°C.:
CALORIES PER
MOLE PER
GAS DEGREE
H2 ........... . 4.92
N2 ........... . 4.94
02 ........... . 4.94
CO ........... . 5.02
These observed values are very nearly FJ2R (%R = 4.97) instead
of ~2R. From this, we may conclude that one of these predicted
groups of degrees of freedom is absent. According to the
quantum theory, the energy per degree of freedom at 300°K. is
not sufficient to "activate" the vibratory degrees of freedom.
W v is, therefore, for ordinary temperatures, zero. We have,
accordingly, for a diatomic molecule,
Wv =Wt+ Wr = J~RT + '7iRT = %RT (44)
SEC. 10] ATOMIC HEAT OF GASES 281
and for the molecular heat Cv,
%R Cv =
For a triatomic molecule, such as H20, C02, or S02, we should
have 3 degrees of freedom in respect of translation; 3 in respect of
rotation, if we assume that the nuclei do not make a linear
molecule (i.e., all three in a straight line); and none for
vibration, assuming the "vibratory" degrees of freedom to be
suppressed, as in the case of a diatomic molecule. We should,
therefore, have, for a triatomic molecule,
Cv = %R + 3,,2R = 3R = 5.96 calories per mole per degree (45)
8
9 g'
f... ,,,,---- - z
d - e -a..: _.,,.)
IJl{T -
~ ~·:aho: _z
Y"'
_e~ sR
a/..-
~-- - C,..,o ,v /
- -- - --- -
Rofahon
----- --~ - -- c' ff
I
I
Trr.ms/afion~
I II
I
I
~
50 100 150 500 150 IOOO 'Z.500 5000
Absolu+e Temperafore (°K}
Fw. 69.-Showing the variation with temperature of the contribution to molecu-
lar heat of energy due to translation, rotation, and vibration.
Actually, the molecular heat of water vapor at 50°C. is 5.96
calories per mole per degree. That for C0 2 at 20°C. is about
6.2. These values are in reasonable agreement with theory.
This agreement between classical theory and experiment is,
however, to be regarded as fortuitous. We find quite another
story when we come to consider the variation of the specific heat
of gases with temperature.
(b) Variation of the Molecular Heat of Gases with Temperature.-
This variation is best illustrated by data on hydrogen, which,
because of its peculiar importance, has been extensively studied.
The data, compiled from several sources, 1 are given in Table X
and are shown graphically in Fig. 69, in which temperature is
plotted to a logarithmic scale. In this curve, the observed data
1
LEWIS: "Quantum Theory;" KEMBLE and VAN VLECK, Phys. Rev.,
vol. 21, p. 653 (1923).
282 THEORY OF SPECIFIC HEATS [CHAP. VIII
are shown by the small circles. The data in the square brackets
at the beginning of the table are not plotted, since, except for the
temperature 29.6°K., they were not corrected 1 to constant
volume. These data for the three temperatures 18.6, 21.2, and
24.0°K., are, therefore, represented qualitatively by the full line
ab of the graph.
TABLE X.-VARIATION OF THE MOLECULAR HEAT OF HYDROGEN WITH
TEMPERATURE
T, Cv, T, Cv,
degrees calories per mole degrees calories per mole
Kelvin per degree Kelvin per degree
18.6
21.2
24.0 l
,2.70
2.78
3 .02
lJ 85
90
100
3.21
3.26
3.42
29.6 3.20 § 2.98 110 3.62
35 2.98 196 4.39
40 2.98 273 4.84
50 3.01 333 4.91
60 2.99 625 5.08
65 3.04 1,000 5.34
70 3.10 1,399 5.62
80 3.14 2,000 6.34
Figure 69 is very instructive. The experimental values show
an increase in Cv from point a, corresponding to some 2 calories per
mole per degree at 20°K. to over 6 calories per mole per degree
at 2000°K. (The dotted lines at either end of the graph are
hypothetical, and qualitative, extensions.) Taking the graph
as a whole, the increase from a to f is seen to take place in three
steps:
1. From the lowest temperatures, Cv rises rapidly to a
maximum at point b corresponding to 2.98 ( = %R)
calories per mole per degree at about 30°K. From b to c,
30 to 60°K., Cv remains constant at J~R, and in this region
H2 behaves as a monatomic gas. Now, J~R is the energy
Wt, which we associate with translation. The graph
abcc', therefore, represents the variation, with tempera-
ture, of Cv, due to the "translatory" degrees of freedom
of the molecule. Were there no other kinds of degrees of
1
LEw1s: "Quantum Theory," p. 106.
SEC. 10] ATOMIC HEAT OF GASES 283
freedom, the value of Cv would remain constant with temper-
ature above 60°K., as is indicated by the dotted line cc'.
2. At point c, corresponding to about 60°K., a second
increase begins. Cv rises until point d is reached, cor-
responding to 4.97 calories per mole per degree ( = ~~R) at
a temperature in the neighborhood of 350°K. For a short
range d to e, Cv remains constant at this value, which is
the value predicted by classical theory (equation (44))
after deleting the vibratory degrees of freedom and the
rotation about the molecular axis. We may, therefore,
conclude that this second rise c to d is due to the gradual
bringing into play of the two "rotational" degrees of
freedom. The curve cdee', therefore, represents the
variation, with temperature, of that part of the specific
heat which is due to molecular rotations. We thus see
that below 60°K. either the hydrogen molecule has no
rotational energy or, rather, that such rotational energy,
if rotation exists, does not vary with temperature.
3. Beginning at point e, a third rise in molecular heat is
observed. Data in the region e to f (500 to 2000°K.) can
hardly be regarded as more than qualitative, but there is
little doubt that an actual increase in Cv does take place
above 500°K. We may ascribe this increase to the
activation of the "vibratoryu degrees of freedom. The
part of the curve efgg', largely qualitative, represents
the variation, with temperature, of that part of Cv which
is due to vibratory motion of the two atoms of the hydro-
gen molecule with respect to each other.
Each of these three increases in Cv with increasing temperature,
namely, a to b, c to d, and e tog, are suggestive of the increase with
temperature in the atomic heat of solids, as shown in the curves
in Fig. 66. We have seen that an approximate solution of the
problem of the variation of the specific heat of solids with temper-
ature was made by introducing the quantum postulate, that the
energy per degree of freedom is given by
hv
ehv/koT _ 1
Einstein and Stern 1 applied this same principle to the rotation
of the molecules and computed the variation of rotational specific
1
Ann. Physik, vol. 40, p. 551 (1913).
284 THEORY OF SPECIFIC HEATS [CHAP. VIII
heat with temperature. Here, v is taken as the (constant)
frequency of rotation of the molecule. Of course, this frequency
must vary with the rotational energy of the molecule. Further
attempts to explain the variation of rotational specific heats with
temperature have been made by Ehrenfest, Kemble, Van Vleck,
and others. 1 The reader is referred to the original articles.
Bjerrum 2 has applied Einstein's theory of specific heats to the
atomic vibrations of polyatomic molecules of a gas in order to
compute specific heats at high temperatures-the region repre-
sented by e to g, in Fig. 69. For C02, for example, in the
equation
where Ct, Cr, Cv are the respective contributions to Cv, due to
translation, rotation, and vibration, Bjerrum assumes
Ct= %R
Cr = ~2R
in accordance with the discussion above. (The temperature is
supposed to exceed that corresponding to point d of Fig. 69; and,
since the molecule C02 is triatomic, all three of the rotational
degrees of freedom are active.) For Cv, Bjerrum uses a formula
of the Nernst-Lindemann type for the specific heat of solids, in
which two infra-red frequencies are inserted: v1 = 0.60 X 10 14 ,
and v2 = 0.37 X 10 14 • Reasonably good agreement with experi-
ment is obtained. Carbon dioxide is known to possess three
absorption bands in the infra-red, as follows:
v: 0.204 X 10
= 14
sec.- 1
v; = 0.70
v~ = 1.1
If these three observed values of frequency for C02 be used instead
of v1 and v2, the agreement with experiment is found to be almost
as good. 3
1
P. EHRENFEST: Verh. d. deusch. Phys. Ges., vol. 15, p. 451 (1913);
KRUGER: Ann. Physik, vol. 50, p. 396; vol. 51, p. 450 (1916); REICHE:
Ann. Physik., vol. 58, p. 657 (1919); MAcDouGAL: Journ. Amer. Chem.
Soc., vol. 43, p. 23 (1921): KEMBLE and VAN VLECK: Phys. Rev., vol. 21, p.
653 (1923). See also a summary by REICHE, "The Quantum Theory"
(English translation, Chap. V).
2
LEwrn: "Quantum Theory," 3d ed., Chap. IV.
3 LEw1s: "Quantum Theory," p. 84.
SEC. 11] SUPPRESSION OF DEGREES OF FREEDOM 285
These infra-red frequencies of polyatomic molecules due to
molecular rotations and vibrations play a very important part
in the general scheme of spectra. 1
We can see, in a general way, how the expression
hv
ehv/koT _ 1
can be applied to the rotations or to the vibrations of polyatomic
molecules in order to explain the corresponding variations of
specific heat with temperature. Two of the three parts of the
curve in Fig. 69 are, thus, qualitatively accounted for. But the
third part of the curve, a to b, exhibits a like variation with
temperature. Do we have here, also, the same type of explana-
tion of the rise of Cv with temperature? It is difficult to see how
we can apply to the translatory motion of gas molecules any
formula involving a frequency, since, for such motion, frequency
can have, ordinarily, no meaning. Various attempts have been
made to apply the quantum theory to translations. Thus,
Nernst postulates that at very low temperatures the translatory
motion of the molecules "degenerates" into a kind of circular
motion (revolution around some fixed point not within the
molecule) and that the energy U of a molecule in this degenerate
state is given by
- 3 hv
u = 2 koT ehv/koT 1
where v may now have a physical meaning: namely, the frequency
of revolution of the molecule in its orbit. This "picture" is
exceedingly artificial. The variation of the "translatory"
specific heat of gases with temperature can hardly be said to have
received even a qualitatively satisfactory explanation. 2 One
might, perhaps, apply Compton's method (see p. 275) of the
"agglomeration" of degrees of freedom by assuming that when
the relative energy with which 2 molecules collide falls below a
certain minimum value e, the molecules coalesce, with a conse-
quent reduction in the number of degrees of freedom and,
therefore, of Cv.
11. The Suppression of Degrees of Freedom.-The quantum
theory, granted its premises, offers a ready "explanation" of the
suppression (to which attention was called in the preceding
1 "Molecular Spectra in Gases," Nat. Research Council, Bull. 57 (1926).
2
See LINDEMANN: Phil. Mag., vol. 39, p. 21 (1920).
286 THEORY OF SPECIFIC HEATS [CHAP. VIII
section) at sufficiently low temperatures, of the rotational and
the vibrational degrees of freedom. According to the quantum
theory, a vibratory (atomic or electronic) mechanism or oscillator
of frequency v may have 0, 1, 2, 3 . . . quanta of energy hv.
If we know, or can compute, the frequency v of an oscillator,
we may, at once, calculate the least amount of energy, i.e., 1
quantum, which the oscillator may have, if it possesses any
energy at all. If the temperature T of the system with which the
oscillator is associated is such that hv is considerably in excess of
the average energy per degree of freedom of the system 1/'2koT,
i.e., if
hv >> }~koT (45)
only very rarely will it happen, according to Maxwell's distribu-
tion law, that the oscillator, as a result of the statistical exchanges
of energy in the system, will be given a ''chance'' to acquire as
much energy as hv. Or, in other words, of a large number of
such oscillators, only very rarely will one be found with any
energy. The relative values, therefore, of hv and koT determine
the "activity" of any type of degree of freedom.
(a) Rotation.-We shall see, later (Chap. X), that the quantum
conditions applied to a rotating system, such as a molecule
(monatomic or polyatomic), require that the angular momentum
M O of the system be restricted to integer multiples of h/21r, h
being Planck's constant, i.e.,
h
Mo,T = r. 21r (46)
where r may be any integer-0, 1, 2, 3 .-and M 0 7 is the
angular momentum corresponding to a particular valtie of T.
Since
and
M0 = I· 21rv l (47)
Ek = }2 · I · (21r v) 2
J
where I is the moment of inertia, v is the frequency of rotation,
and Ek is the kinetic energy, we have, by combination of equa-
tions (46) and (47),
(48)
Ek,T is the kinetic energy for the particular value of r.
If we
know the moment of inertia I of the rotating system, we can
SEC. 11] SUPPRESSION OF DEGREES OF FREEDOM 287
compute, at once, the series of values of kinetic energy which,
under the quantum rules, the system may possess.
For a monatomic molecule, such as argon, we saw (p. 279) that
the moment of inertia was of the order of 10- 42 gm.-cm. 2 The
smallest possible quantity, therefore (other than zero), of kinetic
energy of rotation which the argon atom could possess is found
by putting r = 1 and I = 10- 42 in the equation (48). This giveR
(6.5 x 10- 21 ) 2
Ek,1 = l2 1r • _ = 5.5 X 10- 13 ergs
8 2 10 42
The average energy per degree of freedom at 300°K. is about
2.5 X 10- 14 ergs (i.e., }1 X 1.37 · 10- 16 X 300). We see, there-
fore, that the smallest rotational energy which an argon atom
may possess is some twenty times the average energy per degree
of freedom at 300°K. Clearly, then, the rotational degrees of
freedom of argon are not active at 300°C. The atomic heat
should be due solely to translation. This, as mentioned above,
is found to be the case.
On the other hand, for a diatomic molecule, such as H2, rotating
about an axis perpendicular to a:qd bisecting the line joining the
two atoms, we have a much larger moment of inertia. Taking
the mass of a hydrogen atom as 1.6 X 10- 24 grams and the dis-
tance between the 2 atoms of the H2 molecule as of the order of
10-s cm., the moment of inertia In 2 of the molecule is
In 2 = 7~ X 2 · 1.6 · 10- 24 X (l0- 8 ) 2 = 3 X 10- 40 gram-cm. 2
Inserting this value in equation (48) gives, for r = 1, about
0.2 X 10- 14 erg as the minimum rotational energy permissible
on the quantum hypothesis. We see that this is much smaller
than the average energy per degree of freedom at 300°K. Con-
sequently, the 2 rotational degrees of freedom of H2 should be
active at 300°K. (The third, namely, that about the axis join-
ing the 2 atoms, should be inactive, as in the case of the mona-
tomic molecule, since the moment of inertia of the H2 molecule
about this axis is of the order of l0- 44 .) The molecular heat of
H 2 at 300°K. should, therefore, be FJ,1R, which is the experimental
value.
(b) V ibration.-We know, from the phenomena of band spectra
(see Chap. X, Sec. 15), that the frequency of vibration, with
respect to each other, of the atoms of a polyatomic molecule of a
gas corresponds to infra-red frequencies of the order of 10 14 sec.- 1•
A quantum of this frequency is of the order of 10- 13 ergs. This is
288 THEORY OF SPECIFIC HEATS [CHAP. VIII
considerably more than the average energy per degree of freedom
at 300°K. We should, therefore, not expect the vibrational
degrees of freedom to be active at that temperature.
The computations in this section are, of course, only roughly
quantitative. They should be taken as illustrative rather than
as of numerical significance.
CHAPTER IX
SERIES RELATIONS IN LINE SPECTRA
We saw, in Chap. VII, that attempts to explain the experi-
mentally observed laws of the distribution of energy in the con-
tinuous spectrum emitted by a black body were unsuccessful
until Planck introduced the revolutionary concept of radiation
quanta. Planck's hypothesis became both possible and necessary,
because the very careful experiments of Lummer and Pringsheim
had proved that Wien's law of temperature radiation was
untenable.
A somewhat similar sequence of events is to be found in the
development of our present rapidly growing knowledge of the
characteristic line spectra of atoms and molecules. Correspond-
ing to the empirical laws of temperature radiation, there was
accumulated in the field of line spectra a vast array of very
accurate measurements of the wave lengths of lines in the spectra
of various substances, considerable impetus being given to this
work because of the rigorous demands of spectroscopy as a
method of chemical analysis. From these data, certain relations
were empirically discovered between the frequencies of various
lines in the spectra of certain of the elements. Spectral lines the
frequencies of which are related to each other in a regular
sequence are said to form a series. Similarities among such
series in the spectra of the several elements or substances pointed
to some fundamental mechanism, common to all atoms, as the
origin of characteristic line spectra. The gradual accumulation
of evidence bearing on the problem of atomic structure, on the
one hand, and the increasing importance of these spectral series
relationships, on the other, culminated, about 1913, in the pro-
posal, by Bohr, of the now famous theory of atomic structure and
the origin of spectra which bears his name. The concept of radi-
ation quanta, introduced by Planck to explain the laws of the
distribution of energy in temperature radiation, played a very
fundamental role in Bohr's theory of the laws of the distribution
offrequencies in characteristic radiation.
289
290 SERIES IN LINE SPECTRA [CHAP. IX
In this chapter, we shall consider, briefly, the development of
the empirical laws of spectral series. One can fully appreciate
our present knowledge of this subject and its bearing on the
problems of atomic structure and the nature of radiant energy
only by understanding how, out of an apparently almost hopeless
mass of tangled numerical data, these spectral-series relation-
ships were sifted. No subject illustrates better the patient,
persistent effort necessary to the advancement of scientific
knowledge.
1. Units and Methods of Measurement. (a) Wave Lengths.
In spectral measurements we are concerned, sometimes with the
wave lengths of lines; sometimes with frequencies. The meter, or
the centimeter, is the basis of practically all wave-length measure-
ments. To express the wave length of a line in, or near, the
visible region of the spectrum in meters is, however, awkward
because of the very small decimal required. Various submul-
tiples of the meter are, therefore, employed as units of length in
different parts of the spectrum, viz.,
The micron,1 symbolµ - 10- 4 cm. (or 10- 6 meters)
The millimicron, symbol mµ
0 0
= 10- 7 cm.
The Angstrom, symbol A - 10-s cm.
The X-unit, symbol X.U. - 10- 11 cm.
The last named is used in the X-ray region of the spectrum.
To illustrate these several units, the wave length of the red
cadmium line may be expressed, variously, as
0.000064384696 cm.
0.64384696 µ (microns)
643.84696 mµ (millimicrons)
0 0
6438.4696A (Angstroms) 2
1 The prefix "micro" connotes "million th;" the prefix "milli" connotes
"thousandth," as in "milliampere" or "microampere."
2
Strictly speaking, the .Angstrom is not defined from the meter as a primary
standard of length. Michelson and Benoist in 1895 and, later (1907),
Fabry, Perot, and Benoist measured the wave length of the red cadmium
line in terms of the standard meter. The two measure men ts were almost
exactly in agreement, the wave length according to the latter measurement
being
6438.4696 Angstroms
The International Union for Solar Research, in 1907, adopted this value of
the wave length of the red cadmium line as the primary standard of wave
SEC. I] UNITS AND MEASUREMENTS 291
(b) Frequency and Wave Number.-Wave lengths of lines are
of fundamental importance in the technique of spectroscopy and
in spectroscopic analysis. In physical theory, on the contrary,
frequency v is more fundamental than wave length. We do not,
however, measure frequencies directly. Laboratory measure-
ments yield wave lengths. Frequencies are computed from these
measured values of wave length A and from the velocity of light
c by the relation
c= v·A
Frequencies may be expressed in vibrations per second. This
involves very large numbers, of the order of 10 14 for the visible
region. For many purposes, it is more convenient to use, as
proportional to frequency, the number of waves per centimeter,
symbol n. Thus,
n = 1/A cm.- 1
where A is expressed in centimeters. Thus, the wave number
n of the red cadmium line is 15,531.64 cm.- 1 • Both wave number
and wave length vary with the medium in which the radiation
is propagated. Corrections to standard conditions, usually to
vacuum, are made. When so corrected, wave numbers are
strictly proportional to frequency. Frequency, in vibrations per
second, is, of course, dependent only on the source and not on the
medium through which the radiation is propagated.
(c) Methods.-lt is quite beyond the scope of this book to discuss
the technique of spectroscopic measurements. The student is
referred to the many excellent treatises on the subject. 1 Direct
measurements of wave lengths of light are based on the phenom-
enon of interference. Thomas Young, in 1801, was the first to
make such estimates of wave lengths (see Chap. III) from data
given by Newton on Newton's rings. Twenty years later,
Fraunhofer developed the diffraction grating as a method of
measuring wave lengths. At first, he used equally spaced wires
wound on a frame. Later, he ruled lines on a glass-of the order
of 8,000 per inch. The next important step in the improvement
length on the basis of which all other wave lengths were to be expressed.
Formally, this amounts to a new definition of the Angstrom in terms of the
wave length of the cadmium line such that this wave length is exactly
6438.4696 Angstroms. Other wave lengths are expressed in terms of the
Angstrom so defined.
1
BALY, E. C. C.: "Spectroscopy;" H1cKs, W. M.: "The Analysis of
Spectra;" FOWLER, A.: "Report on Series in Line_ Spectra."
292 SERIES IN LINE SPECTRA [CHAP. IX
in the technique of wave-length measurements was made by
0
Angstrom, who, in 1868, published an elaborate table of wave
lengths of lines in the solar spectrum, made by three carefully
ruled gratings. Measurements with gratings, approaching
modern accuracy, were made by Rowland, who, about 1885,
introduced the concave reflecting grating and greatly improved
the technique of ruling gratings and of using them. Michelson's
introduction of the interferometer and his use of that instrument,
in 1895, for measuring the number of wave lengths of the red cad-
mium line in the standard meter 1 marks the latest important
step in precision measurements of wave lengths. Hicks 2 gives the
following values of measurements of the wave length of the D1
line of sodium as evidence of improvement in spectroscopic
measurements:
Fraunhofer (1822). . . . . . . . . . 5887. 7 .Angstroms
.Angstrom (1864-9) ........ 5895 .13
Rowland (1887-1893). . . . . 5896 .156
Michelson 1 (1887-1893). . . . . 5895. 932
1 Computed by Fabry and Perot.
Michelson's measurement of the wave length of the cadmium
0
line as 6,438.4696 Angstroms 3 and the subsequent adoption of
this as a primary standard, were succeeded by the adoption of a
group of secondary standard lines based on this primary standard
and extending at conveniently spaced intervals throughout the
spectrum. From these, in turn, a group of more closely spaced
tert-iary standards has been prepared. Measurements of wave
lengths in the visible and near-visible region of the spectrum are
now possible with an absolute accuracy (in terms of the centi-
meter) of considerably better than 1 part in a million. The 0
relative accuracy (in terms of the Angstrom) is considerably
higher. .
2. Early Search for Series Relations in Spectra.-As soon as
dependable wave-length measurements became available, numer-
ous investigators, reasoning from the analogy of overtones in
acoustics, sought for harmonic relations in the lines found in the
spectrum of a given element. This search was the more promis-
ing, since certain regularities had already been noted. Thus,
For description of the method used by Michelson, see his "Studies in
1
Optics." University of Chicago Press (1927).
2
"Analysis of Spectra," p. 4.
3 See note, p. 290.
SEC. 2] EARLY INVESTIGATIONS 293
Mascart, in 1869, pointed out that there were several pairs of
lines in the spectrum of sodium.
In 1871, Stoney 1 discovered a very suggestive integer relation
between certain of the lines of the hydrogen spectrum. If a
certain number, 131,277.14, be divided, in turn, by 20, 27, and 32,
the quotients gave, respectively, the wave lengths of first (red),
second (blue), and fourth (violet) lines of hydrogen. Thus:
Observed
Divisor Quo ti en t wave length,
Angstroms
20 6563.86 6563.96
27 4862 .12 4862 .11
32 4102 .41 4102.37
Further search for such integer relations proved fruitless, and, m
1881, Schuster 2 showed that the number of cases found where
such integer relations between lines exist is not greater than might
be expected from probability considerations.
The papers of Liveing and Dewar,3 in the early eighties, empha-
sized the physical similarities occurring in the spectra of such
elements as the alkali metals. They called attention to the
successive pairs of lines in the spectrum of sodium and pointed
out that these pairs were alternately "sharp" and "diffuse"
and that they were more closely crowded together toward the
short-wave-length end of the spectrum, suggesting some kind of
series relation, which, however, they were unable to discover.
A little later, Hartley 4 discovered an important numerical
relationship between the components of doublets or triplets in
the spectrum of a given element: If frequencies, instead of wave
lengths, be used, Hartley found that the difference in frequency
between the components of a multiplet (i.e., doublet or triplet) in a
particular spectrum is the same for all similar multiplets in that
spectrum. Hartley's law is illustrated by Table I, which shows,
in modern measurements, the wave-number differences D..n
between the components of a series of triplets in the spectrum of
1 Phil. Mag., vol. 41, p. 291 (1871).
2 Proc. Roy. Soc., vol. 31, p. 337 (1881).
3
Proc. Roy. Soc., vol. 29, p. 398 (1879); vol. 30, p. 93 (1880).
4
Jour. Chem. Soc., vol. 43, p. 390 (1883). ·
294 SERIES IN LINE SPECTRA [CHAP. IX
TABLE !.-COMPONENT DIFFERENCES t:i.n IN A SERIES OF TRIPLETS IN THE
SPECTRUM OF ZN
(Wave lengths X and wave numbers n, taken from Hicks' '' Analysis of
Spectra," p. 280)
X
n, D.n,
centi- centi-
(Angstroms)
meter- 1 meter- 1
I 4,810.54 20,781.2
388.9
I 4,722.16 21, 170.1
l 4,680.14 21,360.2
190.1
I 3,072.19 32,540.8
388.7
2 3,035.93 32,929.5
l 3,018.50 33, 119. 6
190.1
I 2, 712.60 36,854.3
388.7
3 2,684.29 37,243.0
l 2,670.67 37,432.9
189.9
I 2,567.99 38,929.6
389.7
4 2,542.53 39,319.3
l 2 ,530.34 39,508.7
189.4
I 2,493.67 40,089.6
388.9
5 2,469.72 40,478.5
l 2 ,457. 72 40,676.0
187.5
zinc. The difference in wave numbers between the first and
second component of the series of triplets is very close to 389;
between the second and third, 189. 1 The agreement among these
1
Wave lengths are usually given for air at a standard pressure and
temperature. Values of wave numbers, however, are expressed as the
reciprocal of the wave length reduced to vacuum. These reductions are
made by means of empirical formulre or by tables based on those formulre
(see Hicks, Zoe. cit.). This accounts for the fact that the values of n in the
above table are not quite equal to the reciprocal of X.
SEC. 3] BALMER'S FORMULA FOR HYDROGEN 295
differences is so striking as to point, almost with certainty, to
some fundamental underlying law governing the emission of
these lines by the zinc atom. The early workers were, thus,
apparently justified in the belief that numerical relations among
the multiplets themselves awaited discovery.
Hartley's law made it possible to isolate from the large number
of lines in any given spectrum those groups of lines which were
undoubtedly related. But the task was not an easy one. Figure
70 shows the lines in the zinc spectrum from about 2500 to
2800 Angstroms which region includes the third and the fourth
triplets shown in Table I. It is by no means obvious which of these
many lines are associated by Hartley's law. The two triplets
given in Table I are marked by crosses (X) in Fig. 70. This
same spectral region, however, contains another overlapping
series of triplets, which are designated by circles (o). Clearly,
, r 1111 r r 11m Ir
I I
r i r I rn r 1 1
I I
2500 2600 1100 2800
Wavelen9+h (in An9s1roms)
Frn. 70.-Some lines in the spectrum of zinc in a limited wave-length range
in the ultra-violet. The lines marked with a cross (X) are the triplets shown in
Table I. The lines marked (o) belong to another series of triplets.
the sorting out of these related lines required a great deal of
diligent and patient study.
3. Balmer's Formula for the Hydrogen Spectrum.-The real
beginning of our knowledge of spectral-series formulre dates
from the discovery by Balmer,1 in 1885, that the wave lengths
of the nine then known lines in the spectrum of hydrogen could
be expressed by the very simple formula
m2
"- = hm 2 - 4 (1)
where h is a constant the numerical value 2 of which, to give "- in
0
Angstroms, is 3645.6 and m is a variable integer which takes on
the successive values 3, 4, 5 . . . for, respectively, the first
(beginning at the red) second, third . . . line in the spectrum.
Balmer discovered first, apparently in the course of his search
for integer relations between the wave lengths of lines, that the
1 Ann. Physik, vol. 25, p. 80 (1885).
2
Balmer used also the value 3645.0 as an approximate mean.
296 SERIES IN LINE SPECTRA [CHAP. IX
0
wave lengths (in Angstroms) of the four hydrogen lines in the
visible spectrum, which lines had been very carefully measured
0
by Angstrom, were given, within experimental error, by multi-
plying the number 3645.6 by 7'5, fi, 2~'21, and 7'8- He observed
that these four fractions made no regular series. If, however,
the numerator and denominator of the second and of the fourth
be multiplied by four the series becomes 7'5, lJ{2, 2~'2l, 3Ji 2.
The numerators of these fractions make the series 32, 42, 5 2 , 6 2
. . . ; while the denominators are these same squares decreased
by four. The series is thus given by putting into the fraction
m2
m2 - a2
the value 2 for a, and the successive values 3, 4, 5 . . . for m.
Balmer com pared the predictions of this formula (equation (1))
with the best available values then known for the wave lengths of
the hydrogen lines. The four lines in the visible region had been
0
measured by Angstrom, Mendenhall, Mascart, and Ditscheiner.
Five ultra-violet lines in the spectrum of white stars had been
measured by Huggins. Table II, taken from Balmer's paper,
shows the comparison of the formula with the measurements of
0
Angstrom and of Huggins. The agreement is seen to be excellent
TABLE II.-W AVE LENGTHS OF THE FIRST NINE HYDROGEN LINES
m2
COMPUTED BY BALMER FROM HIS FORMULA X = 3645.6 2
m -a
/a = 2)
X (computed), X (observed)
Line m 0 0
Angstroms Angstroms
0
Ha .................... 3 6562.08 6562.10 (Angstrom)
Ht3 .................... 4 4860.80 4860.74 (Angstrom)
H-y .................... 5 4340.0 4340.10 (Angstrom)
H()···················· 6 4101.3 4101. 2 (Angstrom)
H-E .................... 7 3969.7 3968.1 (Huggins)
H~ .................... 8 3888.6 3887.5 (Huggins)
H77 .................... 9 3835.0 3834.0 (Huggins)
Ho .................... 10 3797.5 3795.0 (Huggins)
HL····················· 11 3770.2 3767.5 (Huggins)
in the visible spectrum. The discrepancy between Balmer's
computed values and the measurements of Huggins increases to
nearly 1 part in 1,000 for HL. Balmer questioned whether this
discrepancy indicated that the formula was only an approxima-
SEC. 4] RYDBERG'S FORMULA 297
tion or whether the data were in error. Recent measurements
have considerably revised Huggins' data but have also revealed
the need for a slight correction to Balmer's formula. We shall
return to this point later.
Balmer correctly predicted that in this series of lines in hydro-
gen no lines of longer wave length than Ha would be found and
0
that the series should "converge" at A = 3645.6 Angstroms,
2
since the fraction m approaches unity as m becomes larger.
m -a 2
2
The impetus which Balmer's discovery gave to work in spectral
series is another illustration of the highly convincing nature of
relations which are expressible in quantitative form. Soon after
the publication of Balmer's work, intensive investigations in
spectral series were initiated by Kayser and Runge and by
Rydberg. We shall discuss Rydberg's work because of its impor-
tant bearing on subsequent developments.
4. Rydberg's Formula for Spectral Series. (a) Some Types of
Series.-Balmer, in the article announcing his discovery, raised
the question whether his formula might not be a special case
of a more general formula applicable to other series of lines in
other elements. Rydberg 1 succeeded in finding this more
general formula. Using the comparatively large mass of wave-
length data then available and starting from the above-
mentioned work of Liveing and Dewar, Rydberg isolated
still other series of doublets and triplets of constant frequency
difference, according to Hartley's law of constant wave-number
separation. He then had at hand a reasonably large number
of series of lines from different elements. The lines in any
given series tended to crowd together toward the ultra-violet,
and Rydberg found that this spacing followed some kind of
regular law. For, if either the wave lengths or the wave numbers
of the lines in a given series (the lines belonging to the series, be
it recalled, being identified by Hartley's law of constant frequency
difference in doublets or triplets) were plotted as ordinates against
the number of the line in the series as abscissre, smooth
curves resulted. Thus, in the spectrum of sodium are found
numerous doublets of which the well-known yellow D
lines are the most intense. Table III2 shows the first
1 A brief, though not very satisfactory, account of Rydberg's work is
given by him in Phil. Mag., vol. 29, p. 331 (1890). A fuller account is
given in Baly' s "Spectroscopy."
2 Taken from FowLER' s "Report on Series in Line Spectra."
298 SERIES IN LINE SPECTRA [CHAP. IX
TABLE III.-DouBLETs IN THE SPECTRUM OF SonrnM
Ordinal Series number
Wave
Wave length, Ordinal number
0 number,
Angstroms number omitting
cm.- 1 Sharp Diffuse
X = 5,889.9
I
11,404.2 8,766 1 1 1
8, 194.8 12,199 2 2 1
6,160.7 16,227 3 3 2
5,889.9 16,973 4
5,688.2 17,575 5 4 2
5, 153 .6 19,398 6 5 3
4,982.9 20,063 7 6 3
4, 751. 9 21,038 8 7 4
4,668.6 21,414 9 8 4
4,545.2 21, 995 10 9 5
4,497.7 22,227 11 10 5
4,423.3 22,601 12 11 6
4,393.4 22,755 13 12 6
4,344.8 23,010 14 13 7
14 of these doublets, only the component of shorter wave length
being given. (The wave lengths of the two D lines are 5,889.963
0
and 5,895.930 Angstroms, respectively.) The spacing of these
lines on a wave-number scale is shown in the strip AB at the left
edge of Fig. 71. The spacing between the lines decreases toward
higher wave numbers (shorter wave lengths) but the spacing is not
regular. If the wave numbers of these lines be plotted as
ordinates against the ordinal number of the line (the numbers
shown in the third column of Table III) as abscissre, the graph
abed (Fig. 71) results, with a break at be, the point represented
by the double circle being the D line. From c to d it is seen that
alternate points fall on a smooth curve, the other points being
below the curve. If, however, we discard from the series the
line A = 5,889.9 Angstroms and plot wave numbers against the
ordinal numbers given in the fourth column, the smooth curve
abef through the solid circles results, the alternate points shown by
the hollow circles falling regularly below the curve. We evidently
have here two nearly parallel series of lines (really, four series,
since each of the plotted points represents a doublet)-one series
shown by the solid circles; the other by the hollow circles. The
two series, as thus isolated, have the distinguishing physical
characteristics that the former consists of sharp lines and is,
SEC. 4] RYDBERG'S FORMULA 299
therefore, called the sharp series; the latter, of diffuse lines and
is called the diffuse series. The ordinal numbers of the lines in
each of these two series are shown in columns five and six of table
III. (The discarded line A = 5,889.9 belongs to still another
series, the principal series, the remaining lines of which are
in the ultra-violet, the second line of this series being at
A = 3,302.34.)
(b) Graphical Relations .-That these two series are similar, so
far as concerns the relation of the frequencies of the several lines
B JO
~
,---
22 ~ ~
..,~ .... -d
k?: ,.,,.. ..... '. )
/("
/, ~/
20 /,
e//)
/~
~18
VI/
t..
Q)
llV' C)
..n lf~C
§16 b i
zI
Q)
l'
:>
0
3= 14
l/
II • Sharp Series - -
-
12
/
I' o Diffuse Ser/es- -
@ The ''.D"Une
-
-
10 I
I
oa
A
2 4 6 8 10 12
Ordinal Number of Line
Frn. 71.-The sharp and the diffuse series of sodium.
to their respective ordinal numbers, is qualitatively evident from
Fig. 71. Rydberg observed, in this way, that such a similarity
existed not only among the several series of a given element but
also among series of different elements. He then showed that this
similarity was quantitative. His method is illustrated by Table IV
and Fig. 72. The difference in wave numbers between successive
lines in the two series in the sodium spectrum given in Table III
are computed and recorded in columns two and three of Table IV.
300 SERIES IN LINE SPECTRA [CHAP. IX
TABLE IV.-WAvE-NUMBER DIFFERENCE tin BETWEEN SuccESSIVE LINES
IN THE SHARP AND THE DIFFUSE SERIES OF SODIUM
tin, in wave numbers
Ordi:p.al number
Difference
of difference
Sharp Diffuse
1 to 2 7461 5376 1
2 to 3 3171 2488 2
3 to 4 1640 1351 3
4 to 5 957 813 4
5 to 6 606 528 5
6 to 7 409 6
4
-3
4nxl0
2 3 4 5 6
Ordinal Number of Difference
Frn. 72.-Wave-number differences between successive lines in the sharp and
the diffuse series of sodium.
These differences are then plotted as ordinates against the ordinal
number of the difference (column four) as abscissre, giving the
two curves shown in Fig. 72. These curves are exactly parallel,
the horizontal lines aa and bb being exactly equal. The two
curves would, thus, coincide, if the one for the diffuse series were
displaced horizontally toward the right by a distance µ equal to
SEC. 4] RYDBERG'S FORMULA 301
the length of the line aa. In short, the values ~ns of the succes-
sive differences for the sharp series are the same function f of
(m + µ) as the corresponding differences ~nd for the diffuse
series are of m, where m is the ordinal number of the difference.
(Note thatµ is less then 1 ordinal unit.)
Rydberg found: (1) that frequency-difference curves plotted
in this way for a large number of series for various elements were
parallel to each other; and (2) that, by a proper choice of the
ordinal number of the differences, it was always possible to
make less than unity the value of the horizontal displacement µ
required to shift any curve into coincidence with one chosen as a
standard. By giving µ a suitable value, therefore, for each
series, the curves were identical in form, a fact which can be
expressed by the equation
~ni = f(m + µi) (2)
when f is some function, to be determined, of (m + µi); µi is the
''shift,'' characteristic of a particular series i the wa ve-n umber
differences of which are ~ni; and m is the ordinal number so
chosen as to make µi less than unity.
(c) A General Formula.-Equation (2), since it contains ~n, is
somewhat analogous to a differential equation. We are, how-
ever, more interested in wave numbers n than in wave-number
differences ~n. By a procedure, partially empirical, we can
determine a relation between n and (m + µ), as follows: Let
the mth ~n be the difference in wave number between the mth
line nm and the line of next higher frequency; i.e., let
~n = nm+l - nm (3)
Since ~n = f(m + µ), we may write the series of equations
nm+l - nm = f(m + µ)
nm+2 - nm+l = f (m + 1 + µ)
nm+3 - nm+2 = f(m + 2 + µ) (4)
noo - noo-1 = f( 00 - 1 + µ)
Adding all of these equations (4), we have
co
n 00 - nm= !J(m + µ), (5)
m
where nrx, is the wave number for very large values of m. We
can see from inspection of the two curves in Fig. 71 that, as m
302 SERIES IN LINE SPECTRA [CHAP. IX
becomes larger, n apparently approaches a limiting value which
we may designate, as above, by n Further, an inspection of
00
•
Fig. 72 shows that, for large values of m, ~n seems asymptotically
co
to approach zero, and i1cm +µ)is, therefore, probably finite, so
m
that we may put
co
if(m + µ) = F(m + µ) (6)
m
where F(m + µ) is some other function of (m + µ) to be deter-
mined. We may, therefore, write equation (5) in the form
nm = noo - F(m + µ) (7)
which is a relation, involving wave numbers, in which n and µ 00
are constants characteristic of the particular series under con-
sideration, n being the limiting or convergent wave number for
00
the series, and µ being the "shift" constant. The function
F(m + µ) can be determined only by a judicious guess.
Rydberg finally concluded that F(m + µ) must be of the form
N 00
(m + µ)- 2 , so that equation (7) becomes
•
nm = n"' - (m ~ µ) 2 (8)
where N was found to be a uni·versal constant for all series (of this
00
type) for all substances, n and µ being characteristic of each
00
substance. The value of N may be determined from Balmer's
00
equation, (1),
2
X=h m
m2 - a2
which-in confirmation of Balmer's surmise-is a special case of
equation (8). Thus, from equation (1),
{ = n= ~ (m a2) 2
;
1 a 2 /h
= h- m2 (9)
which agrees with Rydberg's more general equation if we put
and n°" ={ ) (10)
=~ ~
2
N °" =
since, in Balmer's equation, a 2 = 4.
SEC. 4] RYDBERG'S FORMULA 303
Now, Balmer found that, approximately, h = 3,645.6 Ang-
stroms, or 0.000036456 cm. In wave numbers and to the same
approximation, therefore,
4 '
N CJ = 0.000036456 = 109,720 cm.-1
N is called the "Rydberg (wave number) constant." Because
CJ
of its great theoretical, as well as numerical, importance, its
exact value has been the subject of considerable study. It has
been found to vary 1 slightly, from one element to another, the
values for hydrogen and helium being, according to Hicks, 2
For hydrogen .... N CJ = 109,679.2
For helium ...... N 00
= 109,723.6
in international units.
(d) Numerical Values.-Rydberg found that the formula
NCJ
nm. = nCJ - (m + µ)2
predicted wave lengths very close to the observed ones, although
there were small systematic errors. The constants µ and nCJ for
a particular series were determined empirically by the use of
tables (see Hicks' "Analysis of Spectra" or Fowler's "Report on
Spectral Series"). Thus, for the wave numbers nd of the succes-
sive lines in the diffuse series in sodium shown in Table III, we
may use the approximate equation
109,675
nd = 24,470 - (m + 0 _987 ) 2
The series is seen to "converge" for m = oo at the wave number
24,470.
It has been found that a formula of the Rydberg type but with
three characteristic constants n µ, and a, viz.,
00
,
(11)
1 Throughout this chapter, except as otherwise noted, we shall use the
symbol N co for the Rydberg constant. We shall see, later (Chap. X), that
there are theoretical reasons for the variation of N co with the atomic number
of the atom emitting the spectrum. N co will then be used to refer to the
value of the Rydberg constant for an atom of theoretically infinite mass
compared to the mass of an electron.
2
"Analysis of Spectra," p. 83.
304 SERIES IN LINE SPECTRA [CHAP. IX
agrees more closely with experiment. Hicks' "Analysis of
Spectra" gives for the long-wave-length component of the
diffuse series of doublets of sodium (the same series as is shown in
Table III) the formula
109,675
nm. = 24,475.66 - ( 0 013266)2
m + 0.982390+ -·--
m
This formula predicts the values of the wave lengths of this series
0
within 0.1 Angstrom.
The student will find in such excellent books as Hicks' "The
Analysis of Spectra" or Fowler's "Report on Series in Line Spec-
tra" a large amount of tabular data on spectral series of the
various elements, giving wave lengths, wave numbers, series
formulre, and constants. Such data is well worth careful study,
graphical and otherwise, since the interpretation of these series
and their relations to the problems of atomic structure are
becoming increasingly important.
5. Relations between Series.-It would not serve our purpose
to go into great detail with regard to all the various laws of
spectral series. We shall call attention to only a few significant
relations. In so doing, we shall use the notation given by Fowler, 1
since, for a brief discussion, his notation is rnore suitable than
more recent ones, which can be more profitably employed after
we have discussed Bohr's theory of the atom.
(a) Symbols and N otation.-Rydberg, as we have already men-
tioned, recognized three types of spectral series, distinguishable
from each other by certain physical characteristics, viz., principal,
sharp, diffuse. To these were added, later, by the observations
of Bergmann and others ~bout 1908, a fourth series which
Hicks named the "fundamental series." We may designate
these series by the abbreviations, respectively, P, S, 'n, and
F. To each of these series a formula of the Rydberg type
applies, at least to a sufficiently close approximation for our
discussion. 2
1 '' Report on Spectral Series.''
2 See, however, equation (11).
SEC. 5] RELATIONS BETWEEN SERIES 305
We may represent these four series by the respective formulre
P(m) = P 00 - (m !""
P) 2 (12a)
S(m) = 8 00 - (m !"" S) 2 (12b)
D(m) = D 00 - (m: ""15) 2 (12c)
F(m) = F 00 - (m : ""F) 2 (12d)
The meaning of the symbols is obvious: For example, in the
P series (i.e., the principal series), P(m), P and P stand, 00 ,
respectively, for nm, n and µ of Rydberg's general formula.
00 ,
Each of these series may consist of singlets (i.e., single lines) or of
doublets or of triplets. For the singlet series, Fowler uses the
letters P, S, D, and F, as in equations (12). For the doublet
series, he uses the lower-case Greek letters 1r, <1, o, and ¢, with
subscripts 1 and 2, to designate the two components of the
doublet, subscript 1 being for the stronger (or more intense)
component. For triplet systems, he uses the lower-case letters
p, s, d, and f, with subscripts 1, 2, and 3. Thus, the formula
p1(m) = P1,ro- (m:""p )2 (13)
1
represent the series made up of the first (and strongest) com-
ponents of a series of triplets. And
02(m) = 02,ro - (m !"" 02)2 (14)
represents the second (and weaker) component of a series of
doublets.
A further abbreviation ("shorthand" notation) is effected by
writing the terms containing N as follows: 00
mo2 = N 00
(m + 02) 2
or,
mp1 = (m + P1)2
so that, in this abbreviated notation, the equations (13) and (14)
become
P1(m) = P1,oo - mp1 (13a)
02(m) = 02,00 - mo2 (14a)
306 SERIES IN LINE SPECTRA [CHAP. IX
Note that one must distinguish carefully between p1(m) and mp.
The former represents, for the particular values of m, the num-
bers for the various lines of the series; while the latter is the
value of the term N /(m + p1) 2• Thus p1(2) stands for a
00
particular line; it is the strongest line (subscript 1) in the triplet
(lower-case letter) the value of which is obtained by putting
m = 2 in the equation for the series. On the contrary, lp1 is the
value of the term Noo/(m +
p1) 2 which is obtained by letting
m = 1.
(b) Interrelations between Series.-The "cross" relations
which have been found to exist among the several series in the
spectrum of a given element effect considerable simplification in
the series notation and are indicative of fundamental laws com-
mon to the series. These relations are best illustrated by
numerical examples:
1. Convergence frequency of the S (sharp) and the D (diffuse)
series. Fowler gives, for the P, S, and D series 1 of lithium (quoted
from Rydberg's work),
6 09 72
P( ) 1. 2' (m = 1' 2 ' 3 · · · ) (a)
m = 43 ' 488 - (m l+ '0.9596)
S( m ) = 28 ' 601 - (m 109,0.5951)
721.6
+ (
2' m = 2' 3 ' 4 · · ·
) (b)
(lS)
D(m) = 28,599 - (m ! i!~1 )2'
1 9
4 (m = 2, 3, 4 · · · ) (c)
The convergent wave numbers of the Sand the D series are seen
to be practically identical. That is, in terms of the notation of
equations (12),
(16)
2. Convergence frequency of the P and the S series. A relation
between the P and the S series is as follows: The value of the
variable term for m = 1 in either series is, numerically, very
nearly the convergence frequency of the other series. Thus,
(1 +~;596)2 = 28,573
1 These series for lithium are really close doublets, the separation between
the components being of the order of 0.3 of a wave number unit. The lines
are so close together that, for our purpose, they may be treated as singlet
systems; hence the capital letters P, Sand D.
SEC. 5] RELATIONS BETWEEN SERIES 307
and
(1 + ~.5951)2 = 43,124
Now, 28,573 is (nearly) equal to 28,601, the convergence fre-
quency of the S series; and 43, 124 is (nearly) equal to 43,488.
This agreement is sufficiently good to warrant the law, justified
by later work, that
p = N oo (a)
00
(1 + 8) 2
and (17)
8 Noo (b)
oo = (1 + P)2
Using these relations, equations (16) and (17), we rnay now
rewrite equations (12) as follows:
P( ) N oo N oo (18a)
m = (1 + 8) 2 - (m + P) 2
S( ) N oo N oo (18b)
m = (1 + P) 2 - (m + 8) 2
D( ) N oo N oo
m = (1 + P) 2 - (m + D) 2 (18c)
Or, in the abbreviated notation of equations (13a) and (14a), we
may write, instead,
P(m) = 18 - mP (19a)
S(m) = IP - m8 (19b)
D(m) = lP - mD (19c)
3. Convergency frequency of the F series. For the fundamental
series in lithium, Fowler gives
F(m) = 12,203.1 - mF
If, in the variable term in the D series, we put m = 2, we have
109, 721.6 12 212
(2 +
0.9974) 2 = '
which is very nearly equal to 12,203.1, the convergence wave
number of the F series. We may, therefore, add to the three
equations (18) a fourth,
F( ) N oo N oo (18d)
m = (2 + D) 2 - (m + F) 2
and to equation (19),
F(m) = 2D - mF (19d)
308 SERIES IN LINE SPECTRA [CHAP. IX
4. Terms.-Tbe quantities such asN /(l + S) 2 or N /(m + P) 2
00 00
on the right-hand side of equations (18) are called the terms
of the spectral-series formulre. In the "short hand" nota-
tion of equations (19), these terms are referred to, respectively,
as the "IS" term or the "mP" terms. Since the latter have a
multiplicity of values depending on the successive integer values
of m, these latter terms are frequently referred to as "sequences."
Thus, we have the "mS" sequences given by the successive
quantities
N oo , ____
____ Noo ,
(1 + 8) 2
(2 + 8) 2
Each series of lines is given by the differences of two kinds of
terms. For example, the principal series is given by the differ-
ence between an S term (IS) and the several P terms. Instead
of designating a series by a single letter such as P, S, etc., we may,
as an alternative symbolism, employ the letters in the two terms
which make up the series. For example, the principal series
may be called either the P series or the S-P series; similarly, we
may designate the sharp, the diffuse, and the fundamental series
by, respectively, P-S, P-D, and D-F.
6. The Rydberg-Schuster Law.-From equation (18a), we
may write, for the first line of the principal series,
P(l) = (1 : °"s)2 (1 :'°
P)2 (20)
This equation states the Rydberg-Schuster law, announced inde-
pendently by Rydberg and by Schuster, in 1896, that the wave
number of the first line in the principal series is equal to the differ-
ence between the convergence wave number of that series and the
common convergence wave number of the S and the D series.
7. Relations between Doublet Series.-The spectra of the
alkali metals, as well as those of several other elements, are
characterized mainly by series of doublets. The separations of
the components of these doublets follow certain regular rules.
Table V gives illustrative data for the first five doublets in the
P, the S, and the D series in the spectrum of sodium. An inspec-
tion of this table confirms the following conclusions: (1) The
differences in wave number between the components of the
doublets of both the sharp and the diffuse series, as well as of
their respective convergence wave numbers, are constant
throughout and are equal to the wave-number difference between
SEC. 7] RELATIONS BETWEEN DOUBLET SERIES 309
the first pair of lines of the principal series. (2) The respective
components, first and second, of the sharp and the diffuse series
have the same convergence wave numbers. (3) The differences
in wave numbers between the components in the principal series
decrease rapidly for the higher members of the series (i.e., for
larger values of m), so that the two principal series have the
same convergence frequency.
TABLE V.-DOUBLET SEPARATIONS IN THE SPECTRUM OF SODIUM
Principal series Sharp series I: Diffuse series
II
Wave Wave Wave
m ~n1r m ~nu m ~n5
number number number
1 16,973.35 2 8,766.34 2 12,199.48
17.18 16.79 17 .16
16,956.17 8,783.13 12,216.64
2 30,272.86 3 16,227.37 3 17,575.30
5.49 17.17 17 .17
30,267.37 16,244.54 17 ,592 .47
3 35,042.66 4 19,398.34 4 20,063.20
2.49 17 .17 17 .14
35,040.17 19,415.51 20,080.34
4 37,297.70 5 21,038.37 5 21,413.73
1.50 17 .18 17 .16
37,296.20 21,055.55 21,430.89
5 38,541.54 6 21, 995 .00 6 22 ,227 .11
1.47 17.18 17.14
38,540.07 22,012.18 22,244.25
00 41,449.00 1 ..... 00 24,475.65 2 00 24,475.652
17 .18 17.18
24,492.833 24,492.833
2 3 24,492.83 = N oo
1
41,449.00 - + u) 2
- (1~ 24,475.65 = (1 ~ : 1)2 (1 +1r2)2
The order of the components of the doublets in the principal
series is inverted with respect to the corresponding order for the
sharp and the diffuse series. One reason for this is seen by
recalling the Rydberg-Schuster law (equation (20)) that the first
line of the P series is obtained by subtracting from the con-
vergence wave number of the P series the common convergence
wave number of the Sand the D series. Since, for doublet series,
the two principal series have the same convergence frequency, it
follows that the component of greater wave number in the first P-
series doublet will be given by subtracting from the P-series
310 SERIES IN LINE SPECTRA [CHAP. IX
limit the S and the D limits of lesser wave number; and vice
versa. This inversion of order receives physical confirmation
from the fact that, in the S and the D doublets, the component
of lower wave number (longer wave length) is the stronger;
while, in the P doublets, the reverse is the case.
Recalling the cross relations between the P, the S, and the D
series given by the equations (18), and remembering that the
lower-case Greek letters with suitable subscripts refer to doublet
series, we may write the six series for sodium as follows:
First principal series (m)
Noo Noo (21a)
1r1
= (1 + u) 2 - (m + 1r1) 2
Second principal series 1r2 ( m)
Noo Noo (21b)
- (1 + u) 2 - (m + 1r2) 2
First sharp series u 1 ( m)
Noo Noo (21c)
- (1 + 1r1) 2 - (m + u) 2
Second sharp series u2(m) = [ (1 ~"'1r1)2 + ~u] (m!"'u/21d)
Nco Noo (21d')
= (1 + 1r2) 2 - (m + u) 2
First diffuse series 01(m) = (1 ~:1)2 - (m ~ 0)2 (21e)
Second diffuse series 02(m) = [ (1 ~"'1r1)2+ ~u ] - (m ~ o/21!)
= (1 ~"'1r2)2 - (m ~ 0)2 (21!')
In equations (21d) and (21f), ~u refers to the constant difference
in frequency between the components of the sharp and the
diffuse doublets.
8. Relations between Triplet Series.-Triplet systems of lines
are found in the spectra of the alkaline earth metals-magne-
sium, calcium, strontium, and barium-and in the spectra of
such other elements as zinc, cadmium, and mercury. Singlet
systems of lines are, also, found in these spectra. As in the
case of doublet series, the systems in each element include
the principal, the sharp, and the diffuse groups.
The interrelations among these triplet series are very similar
to those found in doublet series: (1) The three principal series
converge to the same wave number, which, as in equation (20),
is the lS term of the sharp series. (2) The wave-number separa-
SEC. 9] SATELLITES 311
tions of the components of the sharp and the diffuse triplets is the
same for corresponding components throughout the series.
(3) The first sharp and the first diffuse series con verge to the
wave number given by putting m = 1 in the term N (m + p1) of 00
/
tlie first principal series. Similar relations hold for the second
and the third sharp and the diffuse series. (4) The component
of shortest wave length in the principal-series triplets is the
stronger. The reverse is the case with the sharp and the diffuse
series. These laws are summed up in the following equations,
which give, in abbreviated notation, the formulre for the several
series. (Lower-case letters are used for triplets.)
First principal series ........................ p1(m) = ls - mp1 (22a)
Second principal series.. . . . . . . . . . . . . . . . . . . . . p2(m) = ls - mp2 (22b)
Third principal series ....................... ps(m) = Is - mp3 (22c)
First sharp series ........................... s1(m) = lp1 - ms (22d)
Second sharp series ......................... s2(m) = lp2 - ms (22e)
Third sharp series .......................... S3(m) = lp3 - ms (22/)
First diffuse series .......................... d1(m) = lp1 - md (22g)
Second diffuse series ........................ d2(m) = lp2 - md (22h)
Third diffuse series ......................... ds(m) = lp3 - md (22i)
9. Satellites.-The doublets and triplets in the diffuse series
are, in many instances, accompanied by "satellites," i.e., faint
companion lines. In doublets, one satellite accompanies the
component of longer wave length, the satellite being on the long-
wa ve-length side, and the line itself being displaced slightly
toward shorter wave lengths, which displacement, however,
decreases with higher orders of the line (i.e., larger values of m).
In triplets, the long-wave-length component has two satellites;
the middle component has one; and the short-wave-length
component, none. These satellites make the lines appear diffuse
at low dispersion. We shall see, later, that these satellites play
a very fundamental role in the theory of spectra.
10. Combination Lines.-The formulre, such as equations (21)
and (22), which have been found to express the wave numbers of
series of lines are all seen to give the wave number of any given.
line by means of the difference between two terms. These terms
are, for a given series, a fixed term, called the "limit of the series"
and a variable term the values of which depend on the integer
values of m. The fixed term, or limit, of a given series is obtained
by putting m = 1 (or m = 2 for the fundamental F series) in the
312 SERIES IN LINE SPECTRA [CHAP. IX
variable term of some other series. Thus, the second principal
series, say of sodium, is given by (equation (2lb))
1T2(m) = (1 !rou)2 (m !ro1r2)2 (21b)
The first term on the right is the variable term of the sharp
series with m = 1.
It occurred to Rydberg, the actual discovery being made later 1
by Ritz, that other combinations of terms than those giving the
four chief series might correspond to spectral lines observed to
be present in spectra but not belonging to the series. Thus, if
the difference in the two terms
Nm Nm
(1 + cr) 2
(1 + 1r2)
2
gives the first line of the second principal series, one might pre-
dict a line by taking the difference between such terms as
Nm Nm
(2 + cr) 2
- (3 + 1r2)
2
Indeed, one might get a whole series of such lines with the first
term N m/(2 + cr) 2 as the convergence limit and the second term
as the variable term of such a series, replacing the integer 3 by
4, 5, 6 . . . in turn. Such lines, or series, are actually found
to exist and are called combination lines or series. Thus, in the
infra-red spectrum of sodium is found a line X = 3.418µ, wave
number 2,925. The wave number of this line is given by the
difference between the terms
Nm Nm
(2 + 1r1)
2
(3 + cr) 2
(11,175 - 8,248 = 2,927)
Many other such "combination" lines are found in spectra.
11. The Significance of Spectral-series Terms.-These com-
bination lines serve to emphasize the fact that in spectral series
the terms are the fundamental quantities; that to any given atom
belongs a multiplicity of such terms, the wave numbers of the lines
emitted by the atom being given by the differences between pairs
of terms. Now, multiplying wave number n by c, the velocity
of light, gives frequency v, i.e., nc = v; and, therefore,
h · nc = hv
1 Astrophys. Jour., vol. 28, p. 237 (1908).
SEC. 11] SPECTRAL-SERIES TERMS 313
where h is Planck's constant. The product hv, as we have seen,
in Chap. VII, represents a quantity of energy, or a quantum.
Accordingly, if a formula for a spectral line, such as
1r2(3) = (1:rocr)2 - (3 : :2)2 (22)
r
be multiplied through by ch, viz.,
ch. 1r2(3) = c\1:rocr)2 - c\3 :2)2 (23)
the quantity on the left-hand side becomes a quantum of energy
corresponding to the frequency v of the emitted line, in this par-
ticular illustration the long wave-length component of the third
doublet in the principal series, say, of sodium, the wave number
of which (see Table V) is 35,040.17. This wave number, multi-
plied by ch gives
35,040.17 X 2.998 · 10 10 X 6.554 · 10- 27 = 6.885 · 10- 12 ergs
Whatever the actual numerical magnitudes of the two terms on
the right-hand side of equation (23), they must be quantities of
energy, say W1 and W2, respectively, such that their d(fference is,
for this particular line, equal to 6.885 X 10- 12 ergs. Or, to
generalize, we may write equations of the type of equation (23) in
the form
(24)
This equation is a generalization from the empirical data on line
spectra extended by aid of Planck's quantum hypothesis. These
data, however, give us no clue whatever to the interpretation
of the equation in terms of more fundamental relations. It
remained for Bohr to give an interpretation of this equation in
his theory of atomic structure and the origin of spectra, to which
we shall turn in Chap. X.
Independently of Bohr's theory, however, we may remark
that the numerical significance of equation (24) is quite unam-
biguous: The quantum of energy hv corresponding to a spectral
line of frequency v is equal to the dijference between two energy
terms W 1 and W 2 the numerical values of which are computable
from the terms in the spectral-series equations, by multiplying these
terms by c·h. Indeed, we may speculate a little farther into
the deeper meaning of the quantities W1 and W2 on the basis
of the law of the conservation of energy. When an atom radiates
it emits energy. This energy must come from energy previously
314 SERIES IN LINE SPECTRA [CHAP. IX
stored in the atom. We may, therefore, think of W 1 as a certain
amount of energy so stored. As a result of some atomic '' catas-
trophy," the energy of the atom "drops" to W 2, and an amount
of energy hv equal to the difference (W 1- W 2) is emitted. We
reach these conclusions without the necessity of making further
postulates as to the nature of radiation or as to the atomic mechanism
involved in the process.
12. Spectral Series and Atomic Properties.-(a) Numerous
attempts have been made to establish numerical relations
between the various spectral series and the several properties
so~~--~~---~~---~~--.-~~-r~~-y-~~---r~~--,
10 20 30 40 50 60 10 80
Atomic Number
FrG. 73.-Relation between convergence wave numbers of several series and
atomic number.
of the respective atoms in which they originate, as, for example,
atomic weight, atomic volume, atomic number, etc. Evidence
of the probable existence of such relations comes from several
sources: (1) The fact that the Rydberg constant N co appears,
with only slight variation in value, in all spectral series of the
type discussed above is evidence of something common to all
series of this type, irrespective of the atoms from which they
originate; i.e., N must be very intimately connected with the
00
energy quantities W 1 and W 2 of equation (24). (2) The spectra
of elements of a given group in the periodic table are made up of
similar multiplets. For example, the spectra of the alkali
metals consists of doublets, comparatively close together; of the
alkaline earths, triplets and singlets. (3) The very fact that
SEC. 12] ATOMIC PROPERTIES 315
spectral-series formulre of the same type are applicable to the
spectra of many different elements is itself suggestive.
(b) Some examples of the way in which atomic properties
affect spectral series are given in Table VI and are shown graph-
ically in Figs. 73, 74, and 75. Figures 73 and 74 show the
convergence wave numbers of the principal and of the first sharp
series as functions, respectively, of the atomic number and the
atomic volume. In Fig. 73, it is observed that, although there is,
in general, a decrease in the series limits of any one group of
elements with increasing atomic number, with the possible
exception of the principal series of the group Zn, Cd, and Hg, this
decrease follows no regular law.
TABLE Vl.-SERIES LIMITS AS FUNCTIONS OF ATOMIC NUMBER AND ATOMIC
VOLUME
Series limits
Atomic Atomic Doublet
Element
number volume 1 Princi- First separation
pal sharp
Li .............. 3 13.1 43,486 28,582 0.3
Na ............. 11 23.7 41,449 24,476 17.2
K .............. 19 45.5 35,006 21,963 57.9
Rb ............. 37 55.8 33,689 20,872 236.0
Cs ............. 55 70.0 31,405 19,672 553.9
Triplet separations
1~n2 2~n3
I
Mg ............. 12 14.0 20,474 39,760 40.5 20
Ca ............. 20 25.9 17,765 33,989 105 52.5
Sr .............. 38 33.7 16,898 31,038 394 187
Ba ............. 56 39.2 15,869 28,515 878 370
Zn ............. 30 3.76 22,094 42,876
Cd ............. 48 13.1 21,054 40, 711
Hg ............. 80 14.2 21, 830 40, 138
1 Cubic centimeters per gram-atom. Data from Internat10nal Critical Tables.
It was suggested by Halm 1 that there might be a possible
relation between series limits and atomic volume. Figure 74
shows the same series limits as Fig. 73 but plotted against atomic
volume instead of against atomic number, as in the latter figure.
1 Trans. Roy. Soc. Edinburgh, vol. 41, p. 593 (1905).
316 SERIES IN LINE SPECTRA [CHAP. IX
A comparison of the two figures shows that, so far as regularity
is concerned, there is little to choose between the two methods
of plotting.
50
r") /:;;.S'/ ~I Li
zo-_ -...
<..>~c,,~Cr[- - - Na A ·
r--
I
0 i".1. •
n . -o._--'9 --o.... - - /?r_/Pq/
x 40
~ .... , . . . ~ri.r.JL
Alg ---
L.
(1)
_o
E
r"
',(f~
r--
- K -- ---o..Rb__
""()..
..... Q'~
I
:)
2 . . 'o...Sr
Q) 30 '.L',~........ . . . 'O Ba
>
~ . . ... .. ' , ~_:sf Shor
vL)
c
Q)
'°:: >:._cJJc,lcd
Zn
1
Dg .1 ~
- - ~ Pr//Jc .
K
- - - --e- - ---<>--- Rb
---o-- - - c......,
f 20 Mg - - :?_c,/ Ca .,£
-
>
c:
0
-o.. - - - -t:. ---a
u
Ba
10
0 10 '20 30 40 50 60 10
Atom.1c Volume
FIG. 74.-Relation between convergence wave number of several series and
atomic volume.
1
dn2 =Separcd/on bef'ween Fi'rsf Md
Second Comp0!7el7fs of' Trlp/efs
i 25 1-------____:-------r-~-C:;;-s'
~ ~ z'1n3 =Separcd/on befweenJ'econ
i ~ and Th/rd Components
Z8 '20 ~-____:o~f'--=li...:....r:...;:./:_:_/. :. . .:ef.--=-s--~&;;-n-n--;-~-
1 1, ~n =s~parcd/on between °.lia
g; 0- Ooub!ef.s
c; +
~-s
~~
15~----l---+---:H----:.r--t7---;------,
+-
0
4-
0
&. g
~
::,
i L-
IO 1-------4----=-~~~"'-t------t----r-----,
o- 0
(/') 0..
Q)
<f)
0 ~~~\0~~2~0---=-~~0---4~0:----.s~o----;;-:60
0
Atomic Number
Frn. 75.-Multiplet separation as a function of atomic number.
(c) The separations of the components of doublets and triplets
follows a much more regular law. Data for the difference in
wave numbers between the doublets in the sharp and the diffuse
SEC. 12] ATOMIC PROPERTIES 317
series of the alkali metals and between the first and second and
the second and third components of the triplets of the alkaline
earth metals are included in Table VI. These relations are
shown graphically in Fig. 75, in which V ~n is plotted as a func-
tion of atomic number. The linear relation between V ~n and
atomic number is seen to be quite exact in the case of the alkali
metals and only slightly less so for the triplet separations in the
other group of elements. Hence, we have the approximate law
that the square roots of the separations between the corresponding
components of doublets or triplets is proportional to the atomic
number.
(d) Comparison of Spectral-series Formulre with the Formula for
the Hydrogen Spectrum.-Let us return, for a moment, to the
formula for the Balmer's series in the spectrum of hydrogen.
Equation (9), viz.,
1 a 2 /h
n=---- [9]
h m2
may be put in the form
(27)
remembering that N = a 2 /h. For the Balmer's series, a has the
00
value 2, and m takes on the successive values 3, 4, 5 . . .
This equation is seen to be a special case of the Rydberg type
of formula, such as, for example (equation (18b)),
S( ) N N co [18b]
+ P) + 8)
00
m = (1 2 - (m 2'
since, in equation (27), the denominators of the terms are integers,
instead of integers plus decimal fractions, such as P or S, in
equation (18b ). As a matter of fact, equation (27) is only a very
close approximation, since more accurate observations since
Balmer's time have shown that the Balmer's series in hydrogen
is more accurately represented 1 by the equation
n = (2 - 0.~000383) 2 (m + o.:000210)2 <28)
where
N co = 109,678.28
Neglecting these small decimals, however, we may say that the
denominators of both terms in Balmer's equation are integers,
1
FoWLER's Report, p. 89.
318 SERIES IN LINE SPECTRA [CHAP. IX
in contrast to Rydberg's general type of equation. We may now,
in passing, point out two important relations between Balmer's
equation for hydrogen and the more general equations:
1. For large values of m (terms of high order), the terms of
the Rydberg type of formula approach the value of the
corresponding term (i.e., for the same value of m) of the
Balmer formula for hydrogen, since for such large values
of m the decimal fractions, such as P and S, become
relatively less important. This means that for large
values of m spectral terms in general become more
'' hydrogen like.''
2. Corresponding terms in the respective series become,
in general, more "hydrogen like" in the order S, P, D, F.
This is illustrated by the following data 1 for the terms for
m = 3 in the several series in lithium:
S series:
NCO -
NCO
(3 + cr) 2
(3.5976) 2
P series:
Noo - NCO
(3 + 1r1)
2 - (3.9540) 2
D series:
Nw -
NCO
(3 + 0) 2
(3.9978) 2
F series:
Noo -
NCO
+
----
(3 ¢)2 (3.9998) 2
The denominators under N are seen to approach more
00
nearly the integer 4 in the order S, P, D, F.
We shall have occasion to refer to these relations later.
13. Enhanced or Spark Spectra.-The series of lines considered
in the preceding sections are made up, for the most part, of the
lines appearing in the arc or flame spectra of the elements.
When the excitation is more intense, as, for example, when the
excitation is produced by a high-voltage condenser discharge,
additional lines appear in the spectrum of many elements, such
as helium and the alkaline earths. Because of the method of
production, these lines were called "spark" lines or "enhanced"
lines.
Series relationships have, in many cases, been established
among these spark lines. Such series are similar to those dis-
cussed above, except that, instead of the Rydberg constant N 00
,
1 Computed from tables in HrcKs' "Analysis of Spectra," p. 316.
SEC. 13] SPARK SPECTRA 319
the constant (very nearly) 4N is used. Thus, in the spark
00
spectrum of helium are three series of lines represented by the
formula (very similar to Balmer's formula for hydrogen, equation
(27))
1 ·1 \
n = 4NHe ( a2 - m2) (29)
in which, according to Fowler, N He = 109, 723.22. The first
four lines of each of these three series are given in Table VII. 1
TABLE VIL-SERIES OF ENHANCED LINES IN THE SPARK SPECTRUM OF
HELIUM
Lyman series: n = 4Nm(:2 - ~ 2) where a= 2
m = 3 X = 1,640.2 A n = 60,968
4 1,216.0 82,237
5 1,084.9 92,174
6 1,026.0 97,466
"4,686" series: n = 4N11e(:2 - ~ 2) where a = 3
m = 4 }.. = 4,685.6 A n = 21,334
5 3,203.3 31,209
6 2,733.3 36,575
7 2,511.3 39,808
Pickering series: n = 4N u.(~ - ~ 2) where a= 4
m = 5 }.. = n=
6 6,560.4 15,239
7 5,411. 7 18,474
8 4,859.5 25,573
In the alkaline earth group are found series of spark lines
involving the constant 4N which correspond to the four main
00
,
series, P, S, D, and F, of arc spectra. The arc spectra of these
elements are triplets and singlets. Their spark spectra, how-
ever, are made up of doublets, similar to the spectra of the alkali
metals which constitute the preceding group in the periodic table.
1
Data taken from FOWLER'S "Report on Series in Line Spectra."
320 SERIES IN LINE SPECTRA [CHAP. IX
As an illustration, Fowler gives for the longer-wave-length
component of the sharp series of doublets in the spark spectrum
of calcium a formula of the Hicks type, viz.,
ITJ (m) = 70,325.29 - ( 4N ro 0.064899)2
m + 1.205543 - ----
m
Doublet differences in these spark spectra show the same charac-
teristics as in arc spectra; i.e., the doublet separations in sharp
and diffuse doublets are constant for a given element and are
equal to the doublet separation in the first pair of the principal
sen es.
We shall discover a ready explanation of these series of spark
lines in the discussion of Bohr's theory, and we shall see that,
by very intense spark discharges, series involving still higher
multiples of N can be produced.
00
14. Band Spectra.-If one looks at the spectrum of the carbon
arc with a spectroscope of moderate resolving power, one will
observe, at the extreme (violet) edge of the visible part of the
spectrum, "bands," or very broad lines, which are sharply
defined and brightest on the long-wave-length edge, and which
fade out gradually toward shorter wave lengths. With higher
resolving power, these ''bands'' are seen to be composed of a
large number of lines which are crowded together at the long-
wave-length edge, called the "head" of the band, and are sepa-
rated farther and farther toward the short-wave-length side, the
lines, however, being so close together as to appear, under low
resolving power, like a continuous spectrum.
It will not be profitable to go into detail at this point with
regard to the large amount of very complex empirical data on
the band spectra of various substances. In general, several
band edges or heads are found grouped together in a regular
sequence, as in the violet bands in the carbon arc, and these
constitute a group of bands. Several related groups form a sys-
tem of bands. And a given spectrum may contain several such
systems. Three classes of bands are now recognized: (a) bands
in the extreme infra-red part of the spectrum, called, from the
present theory of their production, "rotation" bands; (b)
another class, usually found in the near infra-red, called'' rotation-
vibration" bands; and a third class (c) found in the ultra-violet
or, in some cases, in the visible, called "electronic" bands.
SEC. 15} EXTERNAL PHYSICAL CONDITIONS 321
The series formulre for band spectra are quite different, alge-
braically, from the formulre of the Balmer-Rydberg type which
we have discussed in Sec. 4. For example, Deslandres 1 found
that a formula of the type
n = A + bm 2
(30)
(n= wave number; b = a constant characteristic of the band;
A = the wave number of the "head" of the band; and m = 1,
2, 3 . . . ) gives an approximation to the wave numbers of
ljnes of certain bands. Thus, in the spectrum of nitrogen is a
band with a head at A = 3914.6A the lines of which, according
to Baly, 2 can be represented by the formula
n = 2554.54 + 0.015335(m - 1) 2 (31)
as is shown by Table VIII.
TABLE VIII.-THE "3914.6" BAND IN NITROGEN; WAVE NUMBERS BOTH
AS OBSERVED AND AS COMPUTED FROM EQUATION (31)
Wave number, n
m
Observed I Computed
1 2554.54 2554.54
10 2555.78 2555.78
23 2561.97 2561.96
39 2576.72 2576.68
55 2599.16 2599.26
63 2613.08 2613.49
We shall refer to band spectra again, very briefly, at the end
of Chap. X. The mechanism involved in the production of the
band spectrum of a substance is not so different from that giving
rjse to the spectra of the Balmer-Rydberg type as the difference
in the formulre might suggest. Band spectra arise from mole-
cules; the Balmer-Rydberg spectra, from atoms.
15. Effect of External Physical Conditions on Spectral Lines.-
The frequencies of spectral lines are only slightly affected by
external physical influences. Such effects, though small, are of
considerable importance. We shall make very brief mention of
a few of these effects, some of which we shall consider later.
1
Compt. rend., 1885 to 1904.
2
BALY, E. C. C.: "Spectroscopy."
322 SERIES IN LINE SPECTRA [CHAP. IX
(a) The Zeeman Ejfect.-We have already discussed, in Chap.
VI, the Zeeman effect, namely, the splitting up of a spectral line
into components when the source is placed in a strong magnetic
field. In the so-called normal Zeeman effect, the line is split up
into a triplet. The Zeeman effect in most spectral lines, how-
ever, is abnormal, in that more than three components are
observed. Although there are great differences among series, the
lines of a given series usually behave in a similar way as regards
their Zeeman patterns. This fact is of some little assistance
in identifying lines in series.
We saw, in Chap. VI (equation (19)) that the classical theory
of the Zeeman effect gave the change in frequency 11v when the
emitting source is in a field of strength H, as (e is expressed here
in electrostatic units)
(32)
If 11nH denotes the wave-number (instead of frequency) shift in
a field of strength H, we may write,
11nn 11v e
-=--= =a (33)
H H ·c 41rmc 2
where a is a universal constant.
Accordingly, 11nn/H, the Zeeman shift in wave number per
gauss, should be constant for all values of H. This constancy
has been confirmed by experiment, and the value of a, i.e., the
shift per gauss for the so-called "normal" triplet, to which equa-
tion (32) applies, has been found experimentally to be
a = 4.692 X 10- 5 wave number per gauss
Theory gives 4.70 X 10- 5, by putting the numerical values of
e, c and m in equation (33).
The classical theory, however, fails completely to account for
the abnormal Zeeman effect, in spite of the fact that, as we saw in
Chap. VI, the success of classical theory in explaining the normal
Zeeman effect and in deriving therefrom the correct value of e/m was
one of the very strong supports for the existence of the electron. The
main experimental facts of the abnormal Zeeman effects are
simply stated: (1) The number of components, instead of being
three, as in the normal effect, may be four or more depending on
the line. (2) The shift of any component is proportional to H
and is a simple submultiple of the quantity a of equation (33).
(3) The "pattern" is symmetrical both as to number, distribu-
SEC. 15] EXTERNAL PHYSICAL CONDITIONS 323
tion, and intensity of lines, about the position of the line in
zero field.
As an example of an abnormal Zeeman pattern may be taken
the neon line 1 X = 6678A. This is a "nonet" (nine components
when viewed transtersely to the field). The distribution of these
nine components is shown on a wave-number scale in Fig. 76(a),
zero of the scale corresponding to the unshif ted position of the
line. The relative intensities of the respective components are
given by the numbers immediately above the lines. The three
lines comprising the central group are plane polarized in a direc-
tion parallel to the magnetic field and are spoken of as 2 1r compo-
nents. The groups of three at the ends are plane polarized in a
direction at right angles to the direction of the field. These are
called <J' components. Figure 76(b) shows a conventional diagram
6 4 2 10
1b--io= 4.692-----1 4 6
i v! ~It i z 1, ! 11 i ~
Zeeman Sepor~+ion in Wo.ve Numbers.
1 I? I~ J
+
(a)
I I I . I
'if (II)
o'(-1.i ( b)
Fm. 76.-The Zeeman pattern of the neon line, A = 667·8A.
for representing both the polarization and intensity of the lines.
In this diagram, the lengths of the lines are proportional to their
respective intensities, and the 1r components are plotted above
the horizontal line; the <J' components, below.
The normal separation a = 4.692 wave numbers, as required
by classical theory, is shown above (Fig. 76(a)). The outside
lines of the central group are displaced exactly Jia from zero.
The end components are displaced + a, + Jia, and + ~ia from
zero. In other words, the separations are all multiples of a/4.
All these facts are conventionally included in the mnemonic
notation:
-1 · 0 · 1
---6-·---5---_-4-:4--.-5-.-6 4
1 Quoted by H1cKs, "Analysis of Spectra," p. 92.
2
and u, as here used, are not to be confused with 1r and u as used to
1r
designate series of doublet lines in the Fowler notation for spectral series.
The new series notation does not use Greek letters. Hence, the use of 1r and
u, as applied to components in the Zeeman effect, is permissible.
324 SERIES IN LINE SPECTRA [CHAP. IX
the "-1 · 0 · 1" above the line indicating that there are three 1r
components, and" -6 · -5 · -4: 4 · 5 · 6" below the line indicat-
ing that there are six <J' components, the numbers, in each case,
givjng the displacements of the lines from the zero position in
~ia, the "4" at the end of the line indicating the fraction of
"a." In a similar notation for intensities, we may write,
10 · 10 · 10
6·4·2·2·4·6
Since the pattern is perfectly symmetrical, we may omit the nega-
tive displacements and the corresponding intensities and write,
for the Zeeman pattern,
0 · 1 4 ( 10 · 10) (34)
4 · 5 · 6 (2 · 4 · 6)
the numbers within the bracket standing for intensities. There
are numerous variations of this notation. Thus, we may write
for this pattern,
o · 1 / 4 · 5 · 6 / 4 (10 · 10 I 2 · 4 · 6) (35)
Table IX, compiled from Hicks' "Analysis of Spectra," gives
the Zeeman patterns of a number of representative lines. Note
that the lines given in the table for Na, K, Cu, and Mg, although,
in each case, components of multiplets, have different Zeeman
patterns.
TABLE IX.-ZEEMAN PATTERNS OF SoME REPRESENTATIVE LINES
(Compiled from Hicks' "Analysis of Spectra")
Wave
Element length of Zeeman pattern
"line"
He ........... 5015. 73A 0/1/1 Normal triplet
He ........... 4437.72 0/1/1 Normal triplet
Li ........... 6707.84 0/1/1 Normal triplet
Na ..........
f 5889. 96 1/5 · 3/3 Sextet
l5895.93 2/4/3 Quadruplet
{ 4044.14 1/5 · 3/3 Sextet
l( . . . . . . . . . . .
4047.20 2/4/3 Quadruplet
3247.65 1/5 · 3/3 Sextet
Cu ...........
3274.06 2/4/3 Quadruplet
3838.28 0 · 1/1 · 2 · 3/2 Nonet
Mg .......... 3832.31 1 · 2/0 · 1 · 3/2 Nonet
l3829.36 0 · 2/0 · 2 · 4/2 Octet
Hg .......... 5460.74 0 · 1/2 · 3 · 4/2 Nonet
Hg .......... 3663.27 3 · 6 I 4 · 7 · 10 · 13 /7 Twelve components
Cr ........... 5208 0 · 1 · 2/3 · 4 · 5 · 6 · 7 /6 Fifteen components
SEC. 151 EXTERNAL PHYSICAL CONDITIONS 325
(b) Stark Ejfect.-An effect of an electric field, somewhat
analogous to the Zeeman effect, was discovered by Stark, in 1913,
and is known as the "Stark effect." It is observed in the well-
known "canal rays," when these are moving in an electric field of
several thousand volts per centimeter. The arrangement for
producing the Stark effect is shown diagrammatjcally in Fig. 77.
A and C are, respectively, the anode and cathode in a glass tube,
the gas in which may be maintained at such a pressure that the
Crookes' dark space in front of C is several centimeters long. C
is perforated with small holes through which pass in cylindrical
bundles luminous streams of atoms of the gas which have acquired
a positive charge immediately in front of C. These streams of
atoms are the canal rays. A third terminal S is placed immedi-
To Speclromefer
t
A
1------+
('
To Pump
Frn. 77.-Schematic arrangement for producing the Stark effect.
ately behind C, at a distance of a few millimeters, and an electric
field of several thousand (20,000 and up) volts per centimeter is
maintained between S and C. For the "transverse" Stark
effect, light from the canal rays enters the spectroscope in the
direction of the arrow. Stark observed that the lines in the
spectrum emitted by the canal rays when no field exists between
S and C, are split up, when the field is applied, into numerous
components somewhat after the manner of the Zeeman effect,
the lines being polarized, some of them parallel to the field (p
components), others perpendicular (s components).
Using the same type of notation as in the Zeeman effect, but
expressing the displacements from the zero positjon of the line
directly in wave numbers (instead of in rational submultiples of
some such "normal" displacement as applies to the Zeeman
effect), the Stark pattern for the hydrogen line Hfl (A = 4861.33A)
326 SERIES IN LINE SPECTRA [CHAP. IX
in a field of some 30,000 volts per centimeter is made up of six
components, as follows:
-4.01 · +0.89 · +4.18 (p) ((7 · 1 · 10))
-2.14 · +0.89 · +2.23 (s) (6 · 1 · 8)
The three numbers above the line indicate the position of the
three components polarized parallel (p) to the field, and the three
numbers below the line indicate perpendicularly (s) polarized
components. The numbers within brackets indicate relative
intensities. It is observed that the pattern is not symmetrical
about the zero position. In very strong fields (75,000 volts per
centimeter), the pattern changes completely: the number of
components is greatly increased, and there is symmetry about the
zero position both as to the position of the components and their
intensities. Hicks (" Analysis of Spectra," pp. 102-109) gives
examples of various Stark patterns.
Classical theory fails to account for the. Stark effect. The
quantum theory has been more successful.
(c) Temperature.-The Zeeman effect and the Stark effect are
changes produced in the actual frequency of the atomic radiating
mechanism itself. The observed frequency of a spectral line,
however, may be slightly changed by the motion of the radiating
atom in the line of sight, on account of Doppler's principle: the
apparent frequency increasirig if the motion is toward the observer
and decreasing if the motion is away from the observer. As is well
known, the measurement of the radial velocity of stars is based
on this principle.
In a luminous gas, such as the mercury vapor in the mercury
arc lamp, the atoms are moving with a Maxwellian distribution
of velocity, the average velocity increasing with temperature.
The apparent frequency of the radiation emitted by a given atom
at a particular instant when it has a component of velocity, due
to temperature agitation, away from the observer, will be slightly
less than the actual frequency, the difference between the actual
frequency, and the observed frequency increasing with increase
in this component velocity. Only for those atoms which have no
component of velocity in the direction of the observer will the
observed frequency of the emitted light be equal to the actual
frequency.
It is readily seen, therefore, that a spectral line emitted by a
gas comprises, as observed, a range of frequencies symmetrically
SEC. 15] EXTERNAL PHYSICAL CONDITIONS 327
distributed about the true frequency of vibration for the atom at
rest; and, further, that this range should increase with increasing
temperature. In short, the breadth of a spectral line should
depend on the temperature of the source: the higher the tem-
perature the broader the line. The distribution of intensity
throughout the width of the (observed) line is determined by
Maxwell's distribution of velocities. According to Rayleigh,1
the brightness of the line at a distance + ~n wave numbers from
the center is proportional to
2
e -k · An
where k is a constant which depends on the temperature and on
the mass of atom, and e is the Naperian base of logarithms.
Michelson 2 has studied the width of spectral lines by the
interferometer method. He found that at low gas pressures the
breadth of the line, as observed experimentally, agrees with
that computed from Doppler's principle and Maxwell's dis-
tribution law. Subsequently, Fabry and Buisson, 3 using the
interferometer method, investigated the breadth of lines in the
spectra of the rare gases. They confirmed the formula that the
breadth ~ of a line of wave length X should be
~ = 0.82 x 10-6 >... ~:
where T is the absolute temperature and m is the atomic weight
or molecular weight of the "luminous particle." At liquid-air
temperatures, the breadth of the krypton line X = 5570 A was
found to be only 0.006 A, of which breadth practically all could
be ascribed to the Doppler effect resulting from thermal agitation
at that temperature. When we realize that a very large number
of atoms "cooperate" in the production of any single line as we
see it (or photograph it) in the spectroscope and that, neverthe-
less, except for the Doppler effect, the frequencies emitted by
these several atoms are identical within the very high limits of
resolution of spectroscopic apparatus, we see that the atom must
be a thing of great definiteness, at least so far as concerns the
frequency of the characteristic radiation which it emits
1 Phil. Mag., vol. 27, p. 298 (1889).
2
Phil. Mag., vol. 34, p. 280 (1892); Astrophys. J our., vol. 3, p. 251 (1896).
See, also, DRUDE: "Theory of Optics" (translation by Mann and Millikan),
p. 537.
3 Jour. phys., vol. 2, p. 442 (1912).
328 SERIES IN LINE SPECTRA [CHAP. IX
(d) Pressure.-From the fact that, under favorable conditions,
interference fringes can be produced in an interferometer when
the difference in the path of the two beams is as much as several
hundred thousand wave lengths, it is inferred that the wave
train sent out by any particular atom is continuous, i.e., is without
change of phase, for at least that number of vibrations. In order
that the atom may emit wave trains of this length, it must be
"free from interruptions" for a corresponding period of time.
In the terminology of the kinetic theory of gases, this means that
the mean free time between collisions with other atoms must,
on the average, exceed the time required to emit a complete
wave train, since it may be assumed that a collision with another
atom would cause either a change of phase or excessive damping
or other disturbance.
Now, collisions between atoms become more frequent the
higher the pressure, for a given temperature, of the gas. The
higher the pressure the shorter, therefore, should be the wave
trains, and the more frequent the abrupt changes of phase.
Houstoun 1 has shown that "a train of sine waves with a constant
period but irregular changes of phase is equivalent to the super-·
position of a number of perfectly regular trains, the periods of
which differ slightly from the period of the irregular train.''
In short, at higher gas pressure, not only should there be a broad-
ening of the line due to the Doppler effect but also an additional
broadening due to the increasing frequency of phase changes
resulting from collisions. Michelson has confirmed this by
showing, by measurements with the interferometer, that below a
pressure of the order of a millimeter, the breadth of the hydrogen
line A = 6563A is almost entirely due to the Doppler effect;
but that at higher pressures, the line becomes considerably
broader.
The picture of the atom and its vibrating mechanism, which
one has in mind in considering, as above, the effect of temperature
and pressure on the width of the spectral lines, is more or less
definitely based upon the classical concept of the emission of radiation
in long wave trains. Indeed, we might regard the explanation of
these comparatively minute effects as a final triumph of classical
theory. These effects can be explained by the quantum theory,
1 "A Treatise on Light," Chap. XXL
SEC. 15] EXTERNAL PHYSICAL CONDITIONS 329
but the quantum theory suggests no picture of the radiating
mechanism involved.
In this chapter, we have presented a brief outline of some of
the experimental facts connected with the characteristic line
spectra of the atoms and molecules. We have limited the dis-
cussion to the visible and the near-visible region of the spectrum,
omitting, for example, all reference to X-ray spectra, since the
latter can best be presented in connection with a consecutive
discussion of X-ray phenomena (Chap. XII). We have seen
that the data are very complex: that a given atom may emit
hundreds or even thousands of lines, each line involving a per-
fectly definite, sharply defined, frequency. We have seen that
the relation between the frequencies of the various lines in the
spectrum of a given element are such as to preclude the possi-
bility of their arising from a fundamental frequency and its
overtones. Classical physics could picture no atomic mechan-
ism composed of electrons and positive charges and capable of
vibrating in so many different modes. In characteristic line
spectra, as in temperature radiation, the quantum theory suc-
ceeded, when the classical theory failed.
CHAPTER X
THE NUCLEAR ATOM AND THE ORIGIN OF SPECTRAL
LINES
The atomic model proposed by Bohr, in 1913, and now known
as the "Rutherford-Bohr atom model" or, more briefly, as the
"nuclear atom," was the result of a gradual evolution of ideas,
which extended over several decades and to which many different
investigators contributed. A bird's-eye view of the speculations
on atomic structure preceding the work of Rutherford and of
Bohr will assist the reader in appreciating fully the revolutionary
nature of the concepts introduced by these two scientists.
1. Early Views on Atomic Structure.-Speculations as to the
structure of the atom probably date from the work of Dalton, in
the early years of the nineteenth century. Dalton showed that
chemical combinations between substances always take place in
definite proportions-a fact which could be accounted for by
assuming that each elementary substance is composed of atoms.
Once the existence of atoms was rendered probable, it was natural
to speculate about their structure.
In 1815, Prout proposed the hypothesis which bears his name
and which asserts that all elements are made up of the atoms of
hydrogen as a primordial substance, since the atomic weights of a
large number of the elements are very nearly simple multiples of
that of hydrogen. Prout had no further data to support his
views other than the circumstantial evidence coming from atomic
weights. Accordingly, when more accurate determinations
showed that, in general, atomic weights were not exact integer
multiples of the atomic weight of hydrogen and that there were
such notable exceptions as chlorine, atomic weight 35.5, Prout's
hypothesis was abandoned-only to be revived again decades
later in a new form and, of course, on the basis of newly discovered
experimental evidence (see Chap. XIII).
The first real foundations upon which our modern theories of
atomic structure are being built were laid by Faraday, about
1835. when he enunciated the laws of electrolysis and thereby
330
SEC. 2] THE THOMSON ATOM 331
made it clear that with each univalent atom taking part in elec-
trolysis is associated a definite electric charge, or multiple thereof
for a multivalent atom, which is the same for all atoms. We
have previously discussed (Chap. VI) the development of the
concept of the electrical nature of matter, which culminated in
the discovery of the electron by Sir J. J. Thomson, in 1897.
Therewith, theories of atomic structure began to assume a more
definite form, since it then became obvious that the atom must
be made up of numerically equal quantities of negative and of
positive charges and that practically the entire mass of the atom
must be associated with the positive charge.
2. The Thomson Atom. 1 (a) The Number of Electrons per
Atom.-When, following the discovery of the electron, it became
evident that electrons were constituent parts of all atoms, there
arose two questions of importance: (1) How many electrons are
there in atoms? and (2) How are the electrons and the positive
charges in the atom arranged? Three more or less independent
lines of evidence gave an answer to the first of these two questions:
1. On the basis of classical theory, Sir J. J. Thomson showed
that, when a beam of X-rays passes through matter, it should
be scattered, the scattering coefficient <1 being given by 2
81r e4
(j = -3 24N
me
where e and mare, respectively, the charge and mass of the elec-
tron, c is the velocity of light, and N is the number of electrons
per unit volume. Knowing N from the experimentally deter-
mined value of <1 and knowing the number of atoms per unit
volume of the scattering material, the number of electrons per
atom could be computed at once. From the early measurements
by Barkla on the magnitude of the scattering coefficient, Thom-
son3 computed that the number of electrons per atom was of the
order of the atomic weight, taking the atomic weight of oxygen as
16. But from later measurements, 4 Barkla concluded that the
number of electrons per atom is, for the lighter atoms at least
(except hydrogen), more nearly one-half the atomic weight. The
carbon atom, for example, atomic weight 12, has, according to
1 See THOMSON, J. J.: "The Corpuscular Theory of Matter," Chaps. VI
and VII (1907).
2
Chap. XII, p. 462.
3
"Electron Theory of Matter," p. 145.
4
Phil. Mag., vol. 21, p. 648 (1911).
332 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
Barkla's measurements, 6 electrons. It was later shown that the
scattering coefficient for hydrogen is such as to indicate that it
has only 1 electron per atom.
2. A stream of electrons traveling with a very high velocity
is able to pass through thin sheets of matter but, in so doing, is
diffused, the electrons being deflected, more or less, from their
initial direction, as a result of the electrostatic forces acting on
the moving electrons when they pass near the electrons in the
atoms of the thin sheet. By comparing the computed value of
the diffusion with that observed experimentally, Thomson found 1
that the number of electrons per atom should be of the order of
the atomic weight, taking the atomic weight of hydrogen as unity.
3. Thomson showed 2 that the dispersion of light by a mona-
tomic gas should be determined by the number of electrons per
atom. From data on the dispersion of hydrogen, he concluded
that the number of dispersion electrons per atom of hydrogen
cannot differ much from unity. Drude found, from similar data
on dispersion, that, in general, the number of electrons per unit
volume taking part in refraction is somewhat greater than the
number of atoms per unit volume, but not much greater, and that
the greater the chemical valency of the atom the greater is the
number of dispersion electrons per atom. Even in the heavy
elements, the number of these '' refracting systems'' in unit
volume was found to be only three or four times the number of
atoms. Thomson concluded, however, that it was very probable
that, in the heavier atoms, at least, only a small part of the total
number of electrons in the atom takes part in dispersion and
refraction and, therefore, that there are more electrons per atom
than data on dispersion alone indicated.
These several lines of evidence did not give an entirely unam-
biguous answer to the question of the number of electrons per
atom, but they indicated, at least, the probable order of magni-
tude. They made it certain, for example, that the atom of
carbon contains not more than a dozen or so electrons. Since
the normal atom is electrically neutral, the quantity of positive
electricity per atom was thus roughly determined. Now,
Thomson had shown that the mass of the electron is of the order
of one two-thousandth of the mass of the hydrogen atom. It was
obvious, therefore, that the total mass of the atom is vastly
1 "Electron Theory of Matter," p. 151.
2 Phil. Mag., vol. 11, p. 769 (1906).
SEC. 2] THE THOMSON ATOM 333
greater than the total mass of the comparatively few electrons
which it contains; and it was logical to assume that practically
the entire mass of the atom is associated with its positive charge.
The problem of atom-model building now became somewhat
more definite: (1) The atom contains electrons, the number per
atom being of the order of magnitude of, but probably smaller
than, the atomic weight. (2) The neutral atom must contain as
much positive electricity as there is negative electricity associated
with its negative electrons. (3) The mass of the atom is associated
with the positive charge which it contains. (4) The ensemble of
positive charge and negative electrons which make up the atom
must be stable; the electrons, for example, must be held by
(electrostatic?) forces in fixed positions of equilibrium about which
they may vibrate, when disturbed, with the definite frequencies
required to explain the characteristic line spectra of the elements.
(5) Except when so disturbed, the electrons must be at rest, since,
otherwise, they would emit radiation as required by the elec-
tromagnetic theory. The much greater mass of the positive
charge made it reasonable to assume that it is the electrons,
rather than the positive charges, which vibrate in the process
of emitting radiation.
(b) The Distribution of Positive and Negative Charge in the
Atom.-To meet the requirements of stability and "in default
of exact knowledge of the nature of the way in which positive
electricity occurs in the atom," Thomson considered "a case in
which the positive electricity is distributed in the way most
amenable to mathematical calculation, i.e., when it occurs as a
sphere of uniform density, throughout which the corpuscles 1 are
distributed." 2 According to Thomson's proposal, the atom
consists of a sphere of radius a, throughout which there is a
uniform distribution of positive electricity of density p given by
E
p=-- (1)
''3'7ra3
where E is the total positive charge contained by the sphere.
Stability can be secured by assuming that the electrons are so
arranged inside the sphere that their mutual repulsions are
exactly balanced by the attraction toward the center of the sphere
1 The modern term is "electrons."
2
Quotations from THOMSON'S "The Corpuscular Theory of Matter,"
p. 103.
334 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
due to its positive charge. Consider a single electron, charge - e,
inside the sphere and at a distance r from its center (Fig. 78).
Assuming the ordinary laws of electro-
statics to apply, it is obvious that the
,,,- --,-e net force exerted on the electron by that
I/ ;:f,\
I r ' part of the positive charge outside the
I , \
I ~- - - - a-1- - - sphere of radius r is zero. The total
\ I
\ I force on the electron is, therefore, due to
'\. /
'--- ..,.../ that part of the positive charge, call it
Er, inside the sphere of radius r, and this
FIG. 78.
positive charge acts as if it were concen-
trated at the center of the sphere. Since,
numerically, E = e if the atom is neutral, by use of equation
(1) we obtain
r3
E r = e-
a3 (2)
The force Fe toward the center, acting on the electron, is, there-
fore,
eEr e2
F c = -r2 = as-r (3)
Fe is, of course, zero when r is zero. The position of equilibrium
with 1 electron is, therefore, the center of the sphere, and the
return force acting on this electron when displaced should be
proportional tor. The vibration should be simple harmonic, and
an atom so constituted should be capable of emitting only a
single frequency of the order of 10 15 vibrations per second (taking
e = 4.77 X 10- 10 e.s.u, a = 10-s cm., and the mass of the elec-
tron 9 X 10- 28 gram). It is, perhaps, suggestive that the order
of magnitude of this frequency corresponds roughly to that
observed in spectra. But, of course, no atom is known which
emits only a single frequency.
In an atom which contains 2 electrons, the positions of equilib-
rium are symmetrically situated on opposite sides of the center
and are at a distance apart equal to the radius a of the sphere. 1
1 This is readily shown to be the case, since, remembering that now
E = 2e, and calling r the distance of the equilibrium position from the
center, the force of attraction Fe on each electron toward the center iR
!..
Fe = 2e 2aa
SEC. 2] THE THOMSON ATOM 335
Three electrons inside the sphere of charge E = 3e are in equilib-
rium when at the corners of a symmetrically placed equilateral
triangle the side of which is equal to the radius a. For 4 elec-
trons, the positions of equilibrium are the corners of a regular
tetrahedron the edge of which is equal in length to the radius a.
Continuing in this way, Thomson determined the conditions of
equilibrium for larger numbers of electrons. He found, in gen-
eral, that more than 5 electrons could not exist in stable equilib-
rium in a single ring, unless there were other electrons within
the ring. Thus, 6 electrons at the corners of a hexagon are
unstable, but stable equilibrium is obtained if 1 of the electrons is
put at the center and the remaining 5 at the corners of a regular
FIG. 79.-The Thomson atom model, showing the successive rings of electrons
within the positive sphere.
pentagon. With 22 electrons, 12 are in an outside ring, 8 in a
middle ring, and 2 in an inner ring, as shown diagrammatically in
Fig. 79. The way in which the successive rings are formed with
the addition of more electrons is suggestive, in a rough, qualita-
tive way, of the repetition of properties of elements in the several
columns of the periodic table.
Although now of historical interest only, a table is given 1
by Thomson showing the arrangement of electrons for numbers
from 1 to 100 inside the positive sphere. This table is repro-
and the repulsion Fr, between the electrons is
e2
Fr = 4r2
For equilibrium, Fe = Fr. The proof that this is a position of stable equilib ...
rium is left to the reader.
1 ''The Corpuscular Theory of Matter," p. 109.
336 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
duced, in part, in Table I, in order to emphasize the fact that the
prirne considerations in Thomson's atom were the requirements of
stability, on the one hand, and the demands of the electromagnetic
theory, on the other, that the electrons in the normal, neutral atom
must be at rest, otherwise they would radiate.
TABLE !.-ARRANGEMENT OF ELECTRONS IN THE THOMSON ATOM
Total number .. · 111 21 31 41 51 I I I I I I I I I I I I I
Ring 1........ . 1/ 21 31 41 51 I I I I I I I I I I I I I
Total number... 61 7j 8j 9l10j11j12j13jl4jl5j16I I I I I I I I I I
:~:: t ·.:::::: ~I ~I ~I ~I ~I !I !lt~lt!I I I I I I I I I I
Total number ... 17l 18l 19l2oj21 j22!23j24j25l26j27l28l29j30l31 j I I I I I
Ring 1 ......... 11 11 11 12 12 12 13 13 13 13 13 14 14 15 15
Ring 2 ......... 5 6 7 7 8 8 8 8 910 10 10 10 10 11
Ring 3 ......... 1 1 1 1 1 2 3 3 3 3 4 4 5 5 5
Total number ... 32l33l34l35l36!37j38j39l40l4ll42j43j44l45j46j47j48I I I I
Ring 1 .......... 15 15 15 16 16 16 16 16 16 16 17 17 17 17 17 17 17
Ring 2 ......... 11 11 11 11 12 12 12 13 13 13 13 13 13 14 14 15 15
Ring 3 ......... 5 6 7 7 7 8 8 8 8 9 910 10 10 10 10 11
Ring 4 ......... 1 1 1 1 1 1 2 2 3 3 3 3 4 4 5 5 5
Total number ... 49j 5ol 5l l52 l53l 54l 55l56j57l 58l59/6oj6 l j62j 63l 64l 65[ 66j67[68j69
Ring 1 ......... 17 18 18 18 18 18 19 19 19 19 20 20 20 20 20 20 20 20 20 21 21
Ring 2 ......... 15 15 '15 15 16 16 16 16 16 16 16 16 16 17 17 17 17 17 17 17 17
Ring 3 ......... 11 11 11 11 11 12 12 12 12 13 13 13 13 13 13 13 14 14 15 15 15
Ring 4 ......... 5 5 6 7 7 7 7 8 8 8 8 8 9 9 10 10 10 10 10 10 11
Ring 5 ......... 1 1 1 1 1 1 1 1 2 2 2 3 3 3 3 4 4 5 5 5 5
1 The rings are numbered starting with the outermost and going inward.
Thomson pointed out that such an arrangement of electrons as
is shown in Fig. 79 might, when disturbed, give rise to radiations
of many different frequencies, due partly to the vibrations of
each electron about its equilibrium position and partly to the
orbital revolutions of the rings around the center of the atom.
Thomson's model, however, gave no quantitative data on spectra.
And, although it conformed to the requirements of stability and
the electromagnetic theory, it came ultimately into conflict
with the experiments of Rutherford and his collaborators on the
scattering of swiftly moving alpha particles, as they pass through
thin sheets of matter. These experiments led to a very different
atom model-one in which, in order to meet the conditions of
stability, it became necessary, ultimately, to discard the fundamental
SEC. 3] SCATTERING OF ALPHA PARTICLES 337
deduction from the electromagnetic theory that an accelerated charge
rnust radiate.
3. The Scattering of Alpha Particles in Passing through
Matter.-(a) Radioactive materials in the process of decay
emit three kinds of rays, or radiations, known, respectively, as
the "alpha (a)" the "beta ({3)," and the "gamma (,y) '' rays.
The ')'-rays have been shown to be extremely short electromag-
netic rays the wave length of which is of the order of 0.001 to
0.1 A. The {3-rays have been proven to be negatively charged
particles with mass and charge identical with the mass and charge
of the electron; i.e., the {3-rays are electrons ejected from the
radioactive atom with very high velocities, approaching, for
some kinds of rays, the velocity of light. The a-rays have been
shown to be positively charged particles which have (1) a mass
exactly equal to that of the helium atom and (2) a positive charge
numerically equal to twice the charge on the electron. Since
helium is known to result from radioactive decay, the a particles
are identified with helium atoms which, since the a particles
contain a twofold positive charge, have lost 2 electrons. The
initial velocity of the a particles, although depending somewhat
on the radioactive material from which they originate, is of the
order of 2· 10 9 cm. per second.
(b) These radioactive rays may be studied by means of the
various effects which they produce. Among these effects are
the flashes of light or scintillations which the particles produce
on striking a zinc-sulphide screen, the impact of a single particle
producing a single flash. These flashes are readily observed by a
low-power microscope.
If a stream of a particles, limited by means of suitable
diaphrams to a narrow cylindrical pencil, be allowed to strike
a zinc-sulphide screen placed at right angles to the path of
the particles, the scintillations will occur over a well-defined
circular area equal to the cross-section of the pencil. If,
now, a thin film of matter, such as gold or silver foil, be
interposed in the path of the rays, it is found that they pass
quite readiJy through the foil but that the area over which the
scintillations occur becomes larger and loses its definite boundary,
indicating that some of the particles have been deflected from
their original direction. This spreading out of the stream of
particles on passing through thin layers of matter, solid or
gaseous, is called "scattering."
338 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
Qualitatively, it is easy to explain the origin of the forces
which cause the deflection of the a particle. The particle itself
has a twofold positive charge. The atoms of the scattering
material contain charges, both positive and negative. In its
passage through the scattering material, the particle experiences
electrostatic forces the magnitude and direction of which depend
on how near the particle happens to approach to the centers of
the atoms past which or through which it moves.
If we assume the Thomson model of the atom, with its sphere
of positive electrification, inside which are electrons, the path of
an a particle in passing through such an atom might be as indi-
cated in Fig. 80, the net result of the passage being to deflect the
path of the particle through a
a Pcrr-lJ'cle
+-2 - ---...c::::::_ small angle 0. The major part of
this deflection arises from the
Thomson
+ electrostatic repulsion on the a
Afom particle due to the charge on the
positive sphere, which, for the
heavier atoms at least, has a
Frn. 80.-The deflection of an a mass many times that of the a
particle in passing through the Thom- particle. The electrons within
son atom.
the positive sphere, being capa-
ble of motion about their respective positions of equilibrium and
possessing a mass which is very small compared to the mass of
the a particle, will produce no appreciable deflection of the latter.
Rather, the electrons themselves would ·be pulled from their
positions of equilibrium and set vibrating by the passage of
the a particle. The total deflection of any given particle in
passing through or past a number of such atoms in a thin layer of
scattering material will be governed by the laws of probability.
Thomson showed 1 that the mean deflection <f>m of a particle in
passing through a thin plate of thickness t should be
<l>m = £JVN 1ra t
2 (4)
where O is the average deflection due to a single atom; N is the
number of atoms per unit volume; and a is the radius of the
positive sphere.
This process of scattering of the a particles as a result of a
large number of small deflections produced by the action of a
large number of atoms of the scattering material on a single a
1 Proc. Cambridge Phil. Soc., vol. 15, p. 465 (1910).
SEC. 3] SCATTERING OF ALPHA PARTICLES 339
particle is called multiple, or compound, scattering. It is readily
seen that the structure of the atom assumed by Thomson would
not result in a large deflection due to any single encounter. If,
for example, the a particle were
to pass through the atom in such
a way as to pass near its center
(Fig. 81), the forces acting near
the center being very small, the
resulting deflection would not be
large. According to Rutherford, 1 Frn. 81.-The deflection of an a par...
the number of a particles N <1> ticle which passes very close to the
. h , as a result of multiple center of a Thomson atom.
whIC
scattering, should be scattered through an angle ¢ or greater is
given by
2
N <1> = Noe - <<1>!<1>m) (5)
where No is the number of particles for ¢ = 0, and </Jm is the
average deflection after passing through the scattering material.
(c) Now, Geiger had shown 2 experimentally that the most pro ba..
ble angle of deflection of a pencil of a particles in passing through
gold foil 1/2,000 mm. thick is of the order of 1 degree. It is evident,
therefore, from equation (5), that the probability for scattering
through large angles becomes vanishingly small. For 30 degrees,
for example, the probability would be of the order of 10- 13 •
Geiger showed that the observed scattering obeyed this probability
law for small angles of scattering but that the number of particles
scattered through large angles was much greater than this theory of
multiple scattering predicted. Indeed, Geiger and Marsden
showed 3 that 1 in 8,000 a particles was turned through an angle of
more than 90 degrees by a
thin film of platinum, i.e., were, in
effect, diffusely reflected. This so-called "reflection," however,
was shown to be not a surface phenomenon but rather a volume
effect, since the number of particles turned through more than
90 degrees increased, up to a certain point, with increasing thick-
ness of the scattering foil. It was also found that the proportion
of particles diffusely reflected increased approximately as the ~~
power of the atomic weight of the foil.
It was impossible to explain this excess scattering of a particles
through large angles on the basis of the multiple scattering to be
1 Phil. Mag., vol. 21, p. 679 (1911).
2 Proc. Roy. Soc., London, vol. 83, p. 492 (1910).
a Proc. Roy. Soc., vol. 82, p. 495 (1909).
340 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
expected from the Thomson atom and the laws of probability.
There must be something wrong with Thomson's picture of the
atom.
4. Rutherford's Nuclear Atom. (a) Rutherford's Hypothesis.
Accordingly, Rutherford, in a classic article 1 which may be
regarded as the starting point of our modern ideas on atomic
structure, proposed a new type of atom model capable of giving
to an a particle a large deflection as a result of a single encounter.
This required that an a particle, to be so deflected, must, at some
point in its path past or through an atom, experience much
larger forces than the Thomson atom was capable of producing.
These large forces could be produced only by assuming that the
e'
........
"' \
/
/
a a' h 11 c.'I
I
--;----,
t-._
I
- - - "::{'
\ ,,;""\
....
p
\ 1+ I• .. -"'
' __
\ I
.........
//
(a) The Thomson A+om (b) The Rutherford Afom
Frn. 82.-Comparison of deflections produced on an a particle by (a) the
Thomson atom and (b) the Rutherford nuclear atom for similar conditions of
incidence.
positive charge of the atom, instead of being uniformly distributed
throughout a sphere of atomic dimensions, as Thomson had assumed,
is concentrated in a very small region less than 10- 12 cm. in diameter
at the center of the sphere. This concentrated charge, later called
the "nucleus," is (in Rutherford's original article) assumed to
be surrounded by a numerically equal negative charge distributed,
for sake of computation, "uniformly throughout a sphere of
radius R." This negative charge is, of course, identified with
the charge on the electrons which, in Rutherford's atom, surrmmd
the nucleus in some sort of configuration.
The difference in the effect which the two atoms, Thomson's
and Rutherford's, have on an a particle passing in their neighbor-
hood, is shown in Fig. 82. In (a) is represented diagrammati-
cally the path of an a particle through Thomson's atom, the initial
path of the atom being so directed that, ·were the particle not
1Phil. Mag., vol. 21, p. 669 (1911). Every student of modern physics
should read, and thoroughly digest, this article.
SEC. 4] RUTHERFORD'S NUCLEAR ATOM 341
deflected, it would pass through the atom at a distance p from
its center. As is shown by equation (3), the force which the
particle experiences as it penetrates more deeply into the (Thom-
son) atom becomes less and less. At point c, the force, although
at right angles to the path, is a minimum. The deflection pro-
duced is small and is equal to the angle between ab and de.
In Rutherford's model (Fig. 82(b)), we have a different state
of affairs. Here, an a particle is assumed with an initial path
which, were the particle not deflected, would pass the center of
the atom at a distance p, the same as in Fig. 82(a). Over the
path a'b', the forces which the particle experiences are the same
as for the corresponding path in (a). After passing point b',
however, the forces continue to increase according to the inverse
square of the distance between the particle and the nucleus,
rather than decrease as was the case in (a). It can be shown,
readily, that at a given distance r from the center of the Ruther-
ford atom (r being less than the radius a of the Thomson atom),
the force experienced by the a particle is a3 /r 3 times as great as
for the corresponding position in the Thomson atom, assuming,
in each case, the inverse-square law to hold. The difference in
the forces experienced by the a particle in the two cases becomes
very great as the particle approaches the center. It is readily
seen that the net result of all this is that the Rutherford model
gives to the particle a much greater deflection than does the
Thomson atom.
Now, the chance that an a particle shall approach the atom
along a path which, extended, is distant p from the center is just
as great for the Thomson atom as for Rutherford's. But we see
that, in the latter case, the deflection is large, as a result of a
single encounter. In the former case, the deflection is small.
It is obvious, therefore, that applying the same laws of probability
to the two cases, the proportion of a particles deflected through
large angles would be much greater for the Rutherford atom than
for Thomson's. Rutherford refers to this phenomenon as single
scattering as distinguished from multiple scattering. It is to be
pointed out that multiple scattering is still present, of course, with
the Rutherford model but becomes of importance only for very
small angles of scattering.
The scattering is usually measured by allowing the particles
after passing through the foil F (Fig. 83) to fall on the zinc-
sulphide screen S placed normal to the initial path of the par-
342 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
ticles. By means of a low-power microscope, the number of
S particles per unit area striking
the screen in the neighborhood of
, P Pis observed for various angles of
Sfream of , ~ -:::.:'..: 'j_ _
scattering <f.,. Rutherford, in the
0
alpha parh'cles article referred to, shows that on
the basis of the nuclear type of
atom the number of particles
per unit area striking the screen
FIG. 83.
should be proportional to
(1) 1/sin4~ •
2'
(2) the thickness t of the scattering material, for small values
of t;
(3) (Ze) 2 where e is the charge on the electron and Z is a mul-
tiple depending on the atom;
(4) inversely proportional to the square of the initial kinetic
energy of the particles.
(b) Experimental Confirmation.-These predictions were com-
pletely verified by the experiments of Geiger and Marsden. 1
Table II, taken from their paper, shows the number of scintilla-
tions observed per unit area on the zinc-sulphide screen, for the
various angles of scattering, for gold and for silver foils. The data
TABLE IL-ANGULAR DISTRIBUTION OF SCATTERED a PARTICLES 1
Number of scintillations per unit
1
area
Angle, degrees • 4 </>
sm -
2
Silver Gold
150 1.15 22.2 33.1
135 1.38 27.4 43.0
120 1. 79 33.0 51.9
105 2.53 47.3 69.5
75 7.25 136 211
60 16.0 320 477
45 46.6 989 1,435
37.5 93.7 1,760 3,300
30 223 5,260 7,800
22.5 690 20,300 27,300
15 3,445 105,400 132,000
1 From GEIGER and MARSDEN: Phil. Mag., vol. 25, p. 610 (1913).
1 Phil. Mag., vol. 25, p. 605 (1913).
SEC. 4] RUTHERFORD'S NUCLEAR ATOM 343
are shown graphically in Fig. 84, in which the logarithms of the
number of scintillations per unit area are plotted as abscissre
against the logarithms of 1/sin 4 !
4·-----------"T----r- ----------,.----.
oSilver
• Gold
-!~21------+----+ ~'-4--+---~----'
c:,,
0
..J
4 5 6
Log of Sc"mti\\Qtion5 per Unit Area at Ar3lepfromlnc.iden+ Direction
Frn. 84.-The law of the scattering of a particles: </> is the angle of scattering.
Q)
-+-
:::,
c
'-
Q)
o.zo1---J.~1--~~· ~~----;;,------t--~ --~-;
1,/)
c
0
+
~
+c:
·~
V)
Frn. 85.-Showing that at a given angle of scattering, the number of scattered
particles per unit area is proportional to the thickness of the scatterer.
as ordinates. If these two quantities are proportional to each
other, the points, for each substance, should lie on a straight
line inclined at 45 degrees with the axes. The two lines are
344 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
drawn at exactly 45 degrees with the axes, and the observed
points are seen to agree well with the predictions. This is
the more remarkable since, as the table shows, the number of
scintillations per unit area cover a very wide range of values.
In Fig. 85 (taken from the paper by Geiger and Marsden) is
shown the number of scintillations per minute as a function of
the thickness of the scattering material for several different
metals. The prediction of Rutherford that the scattering should,
for small thicknesses, be proportional to thickness is seen to be
confirmed. It will be recalled (see equation (4)) that, on the
theory of multiple scattering by the Thomson atom, the square
root of the thickness is involved. The decision is, therefore,
quite emphatically in favor of the nuclear type of atom.
Geiger and Marsden also showed, in confirmation of the fourth
of Rutherford's predictions, that "the amount of scattering by a
given foil is approximately proportional to the inverse fourth
power of the velocity (inverse square of the energy) of the incident
a particles," over a range of velocities such that the number of
scattered particles varied as 1: 10.
(c) Atomic Number.-Geiger and Marsden further concluded,
from a study of the variation of scattering with atomic weight
and of the fraction of the total number of incident particles
scattered through a given angle, (1) that the scattering is approx-
imately proportional to the atomic weight of the scatterer over
the range of elements from carbon to gold and (2) 1 that "the
number of elementary charges composing the center of the atoms is
equal to half the atomic weight." This second conclusion was in
agreement with Barkla's experiments on the scattering of X-rays
previously mentioned (p. 331), which determined the number of
electrons associated with an atom. According to these results,
carbon, nitrogen, and oxygen should have, respectively, 6, 7, and 8
electrons. Now, these elements are, respectively, the sixth,
seventh, and eighth elements in the periodic table. The hypoth-
esis was natural, therefore, that the number of electrons in the
atom is numerically equal to the ordinal number which the atom
occupies in the series of the elements, counting hydrogen as the first.
This assumption gives to hydrogen 1 electron, in agreement with
the data on the scattering of X-rays by hydrogen. Helium would
then have 2 electrons and a twofold positive charge on the
nucleus; and we see, therefore, that the a particles, which have
1 For details see the original paper.
SEC. 4] RUTHERFORD'S NUCLEAR ATOM 345
been shown to be helium atoms with a twofold positive charge,
are helium nuclei. Lithium, the third element, should have 3
electrons and a threefold positive charge on the nucleus. Neon,
the tenth element, should have 10 electrons; and so on. Thus
originated the concept of atomic number, the importance of which,
in determining atomic properties, was soon to be emphasized
by the pioneer work of Mosley in the X-ray spectra of the ele-
ments. The atomic number, symbol Z, of an element we may
think of variously as (1) the ordinal number of the element in
the series of the elements starting with Z = 1 for hydrogen; or
(2) the (net) 1 positive charge carried by the nucleus of the atom,
in terms of the electronic charge e as a unit; or (3) the number of
electrons surrounding the nucleus in the normal, neutral atom.
These experiments of Geiger and Marsden so completely
confirmed the conclusions which Rutherford had reached by
-e +2e -e
• CBk-. -r - -~>I!
FIG. 86. FIG. 87.
postulating the nuclear type of atom that, in spite of very weighty
objections, the Rutherford atom model was, at once, universally
adopted.
(d) Some Dijficulties.-The objections to the nuclear type of
atom, in which the positive charge occupies a negligibly small
volume at the center of the atom and is surrounded by negatively
charged electrons in sufficient number to make the atom electri-
cally neutral, are, in large part, concerned with questions of
stability, which conditions formed the very foundation on which
the Thomson model was built. It is obvious that, in the Ruther-
ford atom model, equilibrium cannot be secured, if the electrons
are at rest, by the operation of electrostatic forces alone. For,
consider (Fig. 86) a nucleus with a double positive charge +2e
and with 2 electrons symmetrically placed at a distance r from
the nucleus. Assuming the inverse-square law, the attraction
of the nucleus for each electron is 2e 2 /r2, while the electrons repel
1 We shall see, later (Chap. XIII), why we say "net" charge.
346 THE NUCLEAR ATVM AND SPECTRA [CHAP. x
each other with a force equal to only one eighth of this, namely,
e2/ (2r) 2• The electrons will, therefore, "fall into" the nucleus.
Nothing is gained by giving to the electron an orbital motion,
as that of the earth round the sun. Let the nucleus (Fig. 87)
have a charge + E and a mass M. Let a single electron, charge
- e, mass m very small compared to M, revolve around the
nucleus in a circular orbit of radius a at such velocity v that
Ee mv 2
(6)
r2-r
We should then have mechanical equilibrium, the centripetal
force being supplied by the electrostatic attraction. But
according to the fundamental laws of the electromagnetic theory,
such a revolving electron, since it is subject to a constant accelera-
tion toward the center of its orbit, should radiate energy. This
energy can come only from the energy, part potential, part
kinetic, of the system. The system will, therefore, '' run down''
the electron approaching the nucleus by a spiral path and giving
out a radiation of constantly increasing frequency. This, however,
is contrary to observation. From the standpoint of stability,
the Thomson model is much to be preferred. Yet the experi-
ments of Geiger and Marsden and their interpretation by Ruther-
ford on the basis of the nuclear model of the atom were not to be
denied. It was at this point that Bohr introduced his epoch-
making theory of the structure of the atom and of the origin of
spectra.
5. The Phase Integral.-Bohr's celebrated theory of atomic
structure is, in effect, an extension of Planck's theory of quanta
applied to Rutherford's nuclear atom in an attempt, extraor-
dinarily successful, to explain the origin of the characteristic
spectra of the elements. It is indicative of something very
fundamental to both temperature radiation and characteristic
spectra that, although these two types of radiation differ from
each other in important respects-since the former depends only
on the temperature of the source and is quite independent of the
kind of matter ·acting as emittor, while the frequency of the
latter depends only on the kind of atom and is quite independent
of the temperature-the same revolutionary concepts of quanta
should be successfully applied to both, classical methods having
in each case failed.
Let us return, for a moment, to Planck's linear oscillator,
discussed in Chap. VIII, and let us, for simplicity, think of it as
SEC. 5] THE PHASE INTEGRAL 347
having only mechanical properties; the fundamental principles
involved are not changed thereby. In Fig. 88, let a particle
of mass m vibrate with simple harmonic (linear) motion about a
~ - -- --- -A - - -----~
Fm. 88.
center O with an amplitude A and a frequency v. The displace-
ment x at any instant is given by
x = A sin 21rvt (7)
where t is the time measured from an instant when the particle is
at the center 0. The energy W of the system oscillates between
all kinetic at O and all potential at maximum displacement.
Planck's assumption was that the oscillator could have energy
only in multiples of hv, i.e., that
W = r · hv (8)
where r is an integer and h is Planck's constant, the numerical
value of which is 6.55 X 10- 27 erg. sec.
The total energy of the system is obtained by computing the
kinetic energy W = }imv:ax at midposition. Since
dx
v = dt = 21rvA cos 21rvt
(9)
and
Vmax = 21rvA
.·. W =71L2mv2 = 21r2v2A 2m (10)
/ max
As the particle oscillates, its momentum Px changes from zero at
maximum displacement P to a maximum at midpoint 0. The
momentum at any point in the path is given by
(11)
Now, plot a curve (Fig. 89) with Px as ordinates and the dis-
placement x as abscissm. The figure can readily be shown to be
an ellipse. (The proof of this is left to the student.) The area
S of the figure is obviously given by
S = J Pxdx (12)
348 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
taken over the complete cycle. Since
dx = 21rvA ( cos 21rvt)dt (13)
we have, using equations (11) and (12),
t= 1/v
S = 41r 2 v2mA 2
L=0
(cos 2 21rvt)dt (14)
(15)
If we write equation (15) in the form
21r 2 v 2mA 2
s = ----
v
(16)
FIG. 89.
the numerator, according to equation (10), is equal to W, the
energy of the system; and, by equation (8), W = r · hv. Equa-
tion (16), therefore, becomes, using the right-hand side of equa-
tion (12) for S,
t.
'Yp.,dx =
T • hv
-v- = r · h (17)
where f means "integration over a complete cycle."
That is to say: On the basis of Planck's hypothesis, as given by
equation (8), the integral f p,,dx, taken over a complete cycle, can
take on only a series of values obtained by multiplying the universal
constant h by the integers 0, 1, 2, 3 . . . In terms of Fig. 89, this
means that out of the infinite number of ellipses geometrically pos-
sible, we have a limited number: one ellipse for r = 1, another
for r = 2, another for r = 3, etc. The area between successive
ellipses is numerically equal to h.
fl/v 1
* Jo (cos 2 21rvt)dt = 2 v·
SEC. 5] THE PHASE INTEGRAL 349
This integral in equation (17) is spoken of as the phase integral.
Its value is derived here for the case of an harmonic oscillator to
which Planck's quantum conditions are assumed to apply.
We may assume, however, in accordance with the heuristic
methods of applying the quantum conditions, that an equation
of the type of equation (17) is generally applicable wherever with
any coordinate q (corresponding to x) there is associated a momen-
tum p (corresponding to Px). So that, in general, we may apply
to the coordinate q and the corresponding momentum p the
integral
f pdq = rh (18)
To illustrate, let us apply this equation (18) to the case of a
body moving in uniform circular motion, as, for example (since
we shall consider this case a little later), the
motion of an electron around a nucleus (Fig.
90) in the Rutherford type of atom, dis-
regarding for the moment electromagnetic
radiation. The coordinate in terms of which
the motion is described is an angle ¢ measured
from some arbitrary starting point. Cor-
responding to this (angular) coordinate ¢ is
the angular momentum p</J of the electron Frn. 90.
around the nucleus, p</J being equal to the
product of the moment of inertia of the electron taken about Oby
the angular velocity about the same point. As applied to this
case, equation (18) may be written,
f pq,d<P = rh (19)
The angular velocity being assumed constant, p</J is constant, and,
since f dcp = 2-ir, we have
21rp</J = rh
or
h
P</J = T-
21r
(20)
This is an important equation to which we shall refer presently.
It states that, under the restrictions of Planck's quantum condi-
tions, the angular momentum P</J of an electron moving in a cir-
350 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
cular orbit around a (massive) positively charged nucleus may
take on only those values which are integer multiples (r) of h/21r.
It should be emphasized that this equation follows, more or less
logically, from Planck's quantum hypothesis W = r · hv.
We shall now return to Bohr's development.
6. Bohr's Extension of the Nuclear Atom Model.-(a) We are
now in position (1) to apply to the Rutherford atom model the
principle, derived in the preceding section (equation (20)), that
the angular momentum of an electron in orbital (circular) 1
motion around a nucleus may take on only a series of discrete
values given by r ;1r and (2) by aid of two new postulates intro-
duced by Bohr, to derive an equation for
-~m
, ,, the Balmer's series of lines in the spectrum
,' a' of hydrogen.
•' Let Fig. 91 represent a nuclear atom with
+.lil,.M a nucleus of charge + E and mass M around
which moves in a circular orbit a single
electron of charge - e and mass m, the
Fm. 91.-The Bohr radius of the orbit being a. M is so large
atom model for a nucleus compared to m that, for the present, we
with a single electron. shall assume the nucleus to remain at rest.
Assuming the inverse-square law of electrostatic attraction to
apply, the system will be in equilibrium, according to the ordi-
nary laws of mechanics, when
Ee mv 2
(21)
a2 a
where v is the linear velocity of the electron in its orbit. This
equation at once becomes
Ee 1
- =-mv 2 = T (22)
2a 2 '
T being the kinetic energy 2 of the electron. Now, the potential
V at a distance a from the (point) charge +E is, assuming the
zero of potential to be at an infinite distance from E,
V = E (23)
a
1 The path of an electron around a massive central nucleus is, in general,
an ellipse, to which Kepler's laws of planetary motion apply. The circular
orbit is a special case. We shall discuss the more general case later.
2
We assume, for simplicity, that m does not vary with velocity (see
Sec. 8, infra).
SEC. 6] BOHR'S ATOM MODEL 351
and the potential energy U of the system when a charge - e is at
the distance a from E is
E
U= -e-
a
(24)
The total energy W of the system is equal to the sum of the
kinetic and the potential energy. Therefore,
W = T +U = Ee _ Ee = _ Ee (25)
2a a 2a
A comment, at this point, regarding equations (22) and (24)
may be instructive. An electron, coming from infinity to a
distance a from the charge E "does" work in amount Ee/ a,
according to equation (24). Of this work, only one half, namely,
Ee/2a, by equation (22), remains in the system as kinetic energy.
What has become of the other half? The answer, on the classical
theory, is obvious: In coming from infinity to the orbit in ql)estion,
by whatever path, the electron has experienced acceleration, as a
result of which it must have radiated. Without a knowledge of
the acceleration experienced at each point in the path, we cannot
compute the amount of energy so radiated. But we may apply
the law of the conservation of energy and conclude with confi-
dence that the missing energy has been radiated. If the path
from infinity to the orbit of radius a has been a spiral, the fre-
quencies in this radiated energy would range from zero to w/21r
where w is the angular velocity in the orbit of radius a.
(b) On the classical theory, this circular orbit of radius a could
not be a permanent orbit. It should be, rather, a part of the
spiral path by which the electron reaches the nucleus. Con-
sequently, a could assume any value whatever. But now Bohr
introduces his first hypothesis:
that the values of the angular momentum and, therefore, of the
radius a are restricted to certain discrete values (as required by
equation (20)) and that, therefore, only a limited number of
orbits are possible; further, that these orbits are stable orbits
and that the motion of the electron in them is governed by
the ordinary laws of mechanics, and electrostatics, but is not
subject to the requirements of the electromagnetic theory.
In other words, Bohr assumes that, while moving in these
orbits, the electron does not radiate.
352 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
How are these privileged orbits chosen; what are the various
possible values of the radjus a? This question is answered at
once by introducing the equation (20)
h
P¢ = r - (20)
21r
Since P¢ = ma 2·w (i.e , moment of inertia times angular
velocjty), we have the equation
h
ma 2w = r - (26)
21r
Introducing v = wa into equation (22), we have
_!__m(wa)2 = Ee (27)
2 2a
Eliminating w between equations (26) and (27) and solving for a,
we have
h2
a= r 2 • - - - (28)
41r2mEe
That is, the radii a of the successive privileged orbits are proportional
to the squares of the integers 1, 2, 3 . . . These integers are
called the quantum numbers of the respective orbits. Applying
equation (28) to the hydrogen atom for which E = e = 4.774 X
10- 10 e.s.u., and taking m = 9 X 10- 28 gm. and h = 6 55 X 10- 27
erg. sec., we find that the successive radii aH of the permissible
orbits around the hydrogen nucleus are given by
aH = r 2 • 0.53 X 10- 3 cm.
Since r cannot be less than unity, the radius of the smallest
possible orbit in the hydrogen atom is 0.53 X 10- 3 cm.; or the
diameter of the orbit is of the order of 10- 3 cm., which is exactly
the order of magnitude of the diameters of atoms computed from the
kinetic theory of gases. The diameters of the other orbits will be,
approximately, 4 X 10-8 , 9 X 10-8 , 16 X 10- 8 , etc. Corre-
sponding to each of these orbits, the system has a definite
amount of energy W which is at once obtained by putting into
equation (25)
Ee
W= (25)
2a
the value of a from equation (28). This gives
(29)
SEC. 6] BOHR'S ATOM MODEL 353
where WT stands for the energy of the system when the electron is
in the rth orbit. We see that the larger the value of r the smaller
numerically but the larger in absolute value (because of the
minus sign) is the energy of the system. For hydrogen, the
value of WT for the first orbit, taking the values of m, e, and h
as before, is -2.15 X 10- 11 ergs; for the second, -0.535 X 10- 11
ergs; for the third, -0.238 X 10- 11 ergs; and so on. The lowest
value of WT is that corresponding to the first orbit. Accordingly,
this orbit should be the most stable and is the one which the electron
finally occupies in the normal hydrogen atom.
(c) If we grant the quantum postulates, however much they may
conflict with the older classical ideas, we see that the problem of
stability in the nuclear atom is now solved, but at the expense of
throwing away the only picture which we had of the mechanism by
means of which the atom could radiate energy: for we now assert
that an electron in revolution in one of the privileged orbits
does not obey the electromagnetic laws; and Bohr's postulates
provide no picture of the sequence of events during transitions
between orbits.
It is an experimental fact, however, that atoms radiate. We have
seen (equation (29)) that the larger the quantum number r of a
given orbit (and, therefore, by equation (28), the larger its radius)
the greater is the energy WT of the system. If, now, an electron,
initially revolving in an outer orbit of quantum number r 2 ,
"drops," as a result of some kind of atomic catastrophy, to an
inner orbit of quantum number r1, energy, in amount Wn given
by (remember that r2 > r1)
(30)
has disappeared somewhere. By analogy with the classical
picture and again placing implicit confidence in the law of the
conservation of energy, we may guess that this energy WR has
been radiated. How? That question we cannot answer. Let us
take the guess as similar to an axiom based on observation, just
as was Newton's inverse-square law of gravitational attraction
or Faraday's laws of electromagnetic induction, neither of which
could be explained in terms of more fundamental concepts.
Now, equation (30) is so similar to equation (24), Chap. IX,
namely,
[(24), Chap. IX]
354 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
-which latter equation, be it remembered, was set up by intro-
ducing Planck's quantum postulate into the empirical data on
spectral series-that Bohr was led to his second postulate, namely,
that, when an electron drops from an outer to an inner orbit,
an amount of energy TVR equal to the difference (W 72 - W 71 )
between the energy in the respective orbits is radiated as a quan-
tum of energy offrequency v given by the equation
(31)
(d) The quantities Wr and WT now give us an interpretation
2 1
of the terms of the formulre for spectral series discussed in Chap.
IX: Interpreted as in equation (24) of that chapter, these terms
are simply the amounts of energy possessed by the atom in its
various stationary states, using the expression "stationary state"
to mean the condition or state of the atom when its electrons
are revolving in their several privileged, or stationary, orbits.
We are thus led at once to a formula for spectral series by
putting into equation (31) the respective values of WT given by
equation (29), viz.,
21r 2mE 2e2 1
-----·-
h2 T2
1
21r 2mE 2e 2 1
-----·-
h2 T2
2
and we have
(32)
(33)
Or, in terms of wave number n, instead of frequency v,
2 2 2
n_-21r
-- mE-- e(-1 --
1) (33')
ch3 rj r~
This equation suggests a very simple picture of the origin of
spectral lines: A spectral line originates when an electron initially
in revolution in an outer orbit of quantum number 72 "drops"
to an inner orbit of quantum number 71, the frequency of the line
being given by equation (33). A whole series of lines corresponds
to the dropping of electrons from various outer orbits (i.e., various
SEC. 6] BOHR'S ATOM MODEL 355
values of r 2) into a given inner orbit, designated by r1. Indeed,
equation (33) is identical in form with the empirical formula for
the Balmer's series of lines in the spectrum of hydrogen (equation
(27), Chap. IX), viz.,
[(27), Chap. IX]
in which 1 a = 2 and m = 3 4 5 · · ·
' ' .
Remembering that for hydrogen E = e, we have Bohr's formula
for the hydrogen spectrum:
(34)
These formulre (equation (27), Chap. IX, and equation (34)) are
identical by set ting:
T1 a = 2
=
T2 = m = 3, 4, 5,
21r 2me 4
NH = chs (35)
A crucial test of equation (34) is the value which equation (35)
predicts for the Rydberg constant NH. Putting into equation
(35) the values2 of m, e, c, and h, we have
21r 2 x 9.00 . 10- 28 x (4. 774 . 10-10)4
NH = 2.998 . 10 10 x (6.554 . 10-27) 3 = 109 ,3oo
which differs by only a fraction of 1 per cent from the value of NH,
109,679, 'quoted on page 303. Bohr's theory, therefore, not only
leads to an expression (equation (34)) for the Balmer's series in
hydrogen, which is of the same forrn as that found empirically,
but also, what is quite astonishing, it actually predicts the
numerical values of the frequencies of the lines almost within
the limits of experimental error involved in the determinations of the
constants m, e, c, and hf
(e) A picture, based on the quantum theory, of the origin of
the Balmer' s series of lines in hydrogen is, thus, reasonably com-
plete. The various lines originate when electrons fall from out-
side orbits, the quantum numbers of which are given by the
various values of r2, into the second orbit of the hydrogen atom,
1 We shall now use NH, instead of N oo, as the Rydberg constant for
hydrogen.
2 Taken from the International Critical Tables.
356 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
i.e., the one for which r1 = 2. Thus, the first line Ha of the series
is produced when an electron falls from the third orbit (r 2 = 3) to
the second orbit (r1 = 2); the second line Hf3, by a transition from
the fourth orbit (r2 = 4) to the second orbit; and so on. Of
course, a given atom can, at a particular instant, emit only 1
frequency, namely that corresponding to the particular transition
which it is experiencing at that instant. In the actual production
of the hydrogen, or any other, spectrum in the laboratory, a very
large number of atoms participate. In some atoms, transitions
giving rise to Ha occur. In others, Hf3, H'Y, H 0 • • • are pro-
Z=6
Z:=S
t=4-
FIG. 92.-" Privileged" circular orbits surrounding the hydrogen nucleus,
showing the interorbital transitions which give rise to the various lines in the
hydrogen spectrum.
duced. The number of atoms involved is so large and there are
so many transitions of a given kind occurring each second, that,
to an observer, the lines appear to be produced continuously and
to be of constant relative intensity. If we could observe the
spectrum of hydrogen as produced by only a few hundred atoms,
the lines would appear in flashes, first one line, then another, the
relative frequency of the flashes in the several lines being propor-
tional to their relative intensities as actually observed.
7. Further Successes of the Rutherford-Bohr Atom Model.
(a) Other Series in the Spectrum of Hydrogen.-The Balmer series
of lines in hydrogen is represented diagrammatically in Fig. 92,
by the transitions of electrons between orbits 3 ~ 2, 4 ~ 2,
5 ~ 2 · · · · It is obvious that one may, also, expect transitions
SEC. 7] FURTHER SUCCESSES 357
between outer orbits and orbit number 1 and also between outer
orbits and orbit number 3. The first of these two groups of
transitions should give rise to the series of lines given by the
formula
where r2 = 2, 3, 4 · · This series, as computation from the
known value of NH readily shows, should lie in the ultra-violet.
Actually, such a series of lines, known as the "Lyman series," was
found by Lyman in the ultra-violet. Four of these lines are
shown in Table III. The orbital transitions giving rise to their
origin are represented diagrammatically in Fig. 92.
TABLE III.-THE LYMAN SERIES IN THE SPECTRUM OF HYDROGEN:
n = NH(J2 - ~ ) , T2 = 2, 3, 4, · · ·
2
A
2 n
T .Angstroms
2 1,216.0 82,260
3 1,026.0 97,480
4 972.7 102,820
5 949.7 1 105,300
00 .......... 109,678
1 LYMAN: Nature, vol. 118, p. 156 (1926).
Transitions to the third orbit should give rise to a series in the
infra-red. Several lines of such a series, called the "Paschen
series,'' are known. They are shown in Table IV.
TABLE IV.-THE PASCHEN SERIES IN THE SPECTRUM OF HYDROGEN:
n = NH(~ - ::), T2 = 4, 5, 6,
A
n
.Angstroms
4 18,756 1 5,331.5
5 12,8211 7,799.7
6 10,900 2 9, 180
7 10, 100 2 9,090
8 9,500 2 10,500
00 ....... 12,186
1 Observed by Paschen.
2 Observed by Brackett, Nature, vol. 109, p. 209 (1922).
358 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
Brackett has also observed two lines in the far infra-red spec-
trum of hydrogen which correspond to transitions from the fifth
and the sixth orbits, respectively, to the fourth. These are given
in Table V.
TABLE V.-THE BRACKETT SERIES IN THE SPECTRUM OF HYDROGEN:
n = NHU 2 - ~ ) , r2 = 5, 6,
A
n
(cm.)
5 4.05 x 10- 4 2470
6 2.63 3800
6855
With these several series of lines as actually observed in the
laboratory, the Bohr theory is in excellent agreement, not only
qualitatively but also quantitatively.
(b) The Spectrum of H elium.-Mention was made in Chap.
IX (p. 319) of three series of lines in the spark spectrum of
helium given by the equation (29), Chap. IX:
n = 4NHe(_!_ __!_)
a2 - m2 [(29), Cha.p. IX]
where for each series a takes on the respective values 2, 3, or 4,
and m has corresponding values 3, 4, 5 . . . ; 4, 5, 6 . . . ;
5, 6, 7 . . . These spark lines occur when the conditions
of excitation are very intense.
Equation (33')
n = 21r2mE2e2(l -
ch3 ,2 r2
1°) (33')
1 2
gives the general series formula for a nucleus of charge +E
around which revolves a single electron. (The formula does not
apply to more complicated systems.) Now, helium, the second
element in the periodic table, has a nucleus with a twofold positive
charge E = 2e and, normally, 2 electrons. With intense excita-
tion, however, it may be supposed that, occasionally, both these
electrons are removed and the bare helium nucleus remains.
This nucleus, with its twofold positive charge, should then
behave, in "capturing" an electron, in much the same way as
the hydrogen nucleus. Equation (33') should, accordingly,
SEC. 7] FURTHER SUCCESSES 359
predict the resulting spectral series in helium, if we put E = 2e.
We should, therefore, expect series of lines in the spark spectrum
of helium the wave numbers nHe of which are given by
nHe
= 4 . 21r2me4(_! -
h3 2
_!)
2 (36)
C Tl '2
where r 1 and r 2 take on appropriate values. The constant
multiplying the bracket is seen to be 4NH if we set NH equal to
21r 2 me 4 /ch 3 , according to equation (35). Equation (36) is very
nearly in agreement with the empirical equation (29), Chap. IX,
for the several series in the spark spectrum of helium, given on
page 319. There is, however, a slight numerical discrepancy,
which becomes significant when measurements of spectroscopic
precision are considered; if we determine the Rydberg constant
NH from the hydrogen spectrum, we find (see p. 303)
Nu = 109,679.2
The experimental value of 4 · NHe in equation (29), Chap. IX
(see page 319) gives for NHe,
NHe = 109,723.22
(c) The Spectroscopic Value of the Ratio of the Mass of the
Electron to That of the Hydrogen Atom.-This slight discrepancy,
mentioned in the preceding paragraph, in the two values of the
Rydberg constant, as determined from hydrogen and from the
spark spectrum of helium, although they differ by only a few
parts in 10,000, has led to a very important deduction from, and,
incidentally, to a very surprising confirmation of, the Bohr theory.
In setting up the fundamental equa-
tions (21) and (22)
(21)
1+.lJ~,
(22) ----r ----
----a·-··
as the sta1ting point of Bohr's theory,
we assumed that the mass M of the
Frn. 93.-The revolution of
nucleus is so large in comparison with the nucleus and the electron of
the mass m of the electron that the the Bohr atom about their com-
mon center of gravity.
nucleus remains at rest with respect to
(and, of course at) the center around which the electron revolves.
Strictly speaking, this is not true; for both masses revolve about
360 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
their common center of gravity 0, the nucleus in a small circle
of radius R (Fig. 93) and the electron in a larger circle of radius
r. As before (p. 350), letting a represent the distance between
nucleus and electron, we have, from ordinary mechanics, the
relations
m
R = aM+m (37a)
M
r = aM +m (37b)
We must now write, instead of the single equation (21), two
equations, viz.,
Ee MV 2
a2 - ~
(38a)
Ee mv 2
(38b)
a2 r
where V = wR is the velocity of the nucleus and v = wr is the
velocity of the electron, w being the common angular velocity of
the system about its center of gravity. The total kinetic energy
T of the system is now made up of two parts: that of the nucleus
TM and that of the electron Tm· Introducing the values of R
and r from equations (37), we may, analogously with equation
(22), change equations (38) to
2
Ee m _ 1 2 ( m ) _
2a · M + m - 2M(wa) M + m. - TM (39a)
!:·M!m =~m(wa) 2(M!mY = Tm (39b)
Since the total kinetic energy T of the system is
T=TM+Tm
we have, by adding equations (39),
T = :: = HmM ! m](wa)2 (40)
We thus obtain the same value, namely Ee/2a, for the total kine-·
tic energy of the system as when we considered M infinite (equa-
tion (22)).
To compare equation (40) with the corresponding equation
(22), we may write the latter
T = ~im(wa) 2 (23')
SEC. 7] FURTHER SUCCESSES 361
We may think of these two expressions for T, namely, equations
(23') and (40) as differing from each other only in the fact that in
equation (40) the mass rn of the electron is multiplied by the fac-
tor M/(M + m). Without following through the entire deduc-
tion, we can proceed to the final result simply by replacing min
equation (33') for the spectral series by m · M/(M + m). We
then have for the general equation of a spectral series in the
nuclear type of atom, in which the mass of the nucleus is not
(taken as) infinite with respect to the mass of the electron,
n = 21r2mE2e2 • M
ch 3 M +m
[l'i -- ,~lJ (41)
We now see that the factor multiplying the bracket, which factor
contains the Rydberg constant, depends on the ratio of the mass of
the electron to that of the nucleus. This gives a clue to the observed
difference in the values of the Rydberg constant as computed
from hydrogen and from ionized helium.
For, letting MH and MHe stand for the respective masses of the
hydrogen and the helium nuclei (MH and MHe are practically the
masses of the two atoms, since the masses of their electrons are
negligible) and remembering that for helium E = 2e, we should
have for the Rydberg constant for He and for H, respectively,
21r 2me 4 MHe
NHe = ·---- (42a)
ch 3 MHe+m
21r 2me 4 · - -MH
NH= -~ (42b)
ch 3 MH+m
The atomic weight of hydrogen is 1.0077; that of helium, 4.00.
From these values,
MHe = 3.969MH
Making this substitution for MHe in equations (42), we have, by
division,
m
1 +
MH
---- (43)
m
l + 3.969MH
According to Sommerfeld, 1 the best values of N He and Nu are,
respectively, 109,722.14 and 109,677.69. Putting these values in
1
"Atomic Structure and Spectral Lines," English translation, p. 224.
362 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
equation (43) and solving for m/ MH gives, for the ratio of the
mass of the electron to that of the hydrogen atom,
m 1
MH = 1,846
This value of the ratio of these two masses is practically identical
with that determined by other methods. For example, Sharp, 1
from measurements on the Compton effect in scattered X-rays,
finds 2 that the mass of the electron is 8.99 X 10- 28 grams. The
mass Mn of the hydrogen·atom, determined from Faraday's con-
stant of electrolysis F, the atomic weight of hydrogen A., and the
charge on the electron e, by the relation
A.
Mn= F/e
is 1.663 X 10- 24 grams. Hence, the ratio
m 8.99 X 10- 28 1
MH - 1.663 x 10- - 1,850
24
Other methods for determining the ratio m/ Mn give identical
results within the limits of precision of the measurements.
(d) Variation in the Rydberg Constant.-This agreement
between the value of m/ Mn computed from the slight difference
in the Rydberg constant for hydrogen and for helium and that
determined by other methods is added confirmation of Bohr's
theory. Further, equation (41) provides a method of computing
the Rydberg constant for any element. Putting E = Ze, where Z
is the atomic number of an element, we may write equation (41)
in the form
n = z2 . 21r2me4.
ch 3
___!!_~ .
Mz +m
(J_Ti - __!_)
Ti
(44)
where Mz stands for the mass of the nucleus of the element of
atomic number Z. Evidently, the Rydberg constant Nz for this
element is
21r 2me 4 Mz
Nz = ch3 . M z m +-
1 Phys. Rev., vol. 26, p. 691 (1925).
2 The usual methods of determining the mass of the electron are based
upon a measurement of e/m by any of the various methods (magnetic deflec-
tion, Zeeman effect, etc.) and a computation of m from the independently
measured value of e. Sharp's method is independent of e/m but depends
upon e, since the latter constant is used in measuring the wave length of the
X-rays (see Chap. XII).
SEC. 7] FURTHER SUCCESSES 363
or,
1
(45)
We may write this in the form
(46)
where N co, given by
(47)
is the value of the Rydberg constant for very large (practically
infinite) values of Mz. The constant N co can, of course, be com-
puted from equation (47); but the quantities involved are not
known with spectroscopic precision. Rather, N co is computed
from equation (46) by using the observed value of Nz, say, for
hydrogen. From the observed value of Nm namely 109,677.69,
and the known ratio m/MH, it is found that
N 00 = 109,733.11
Using this value of N in equation (46), one can compute Nz
00
for any element. For example, taking the atomic weight of
lithium (Z = 3) as 6.94, one finds that
NLi = 109,724.51
For the heavier elements, Nz approaches N co within the limits of
error of experiment.
(e) The Rydberg-Schuster Law; Combination Lines.-In Chap.
IX, p. 308, we referred to the Rydberg-Schuster law that
. . . the wave number of the first line of the principal series is equal
to the difference between the convergence wave number of that series
and the common convergence wave number of the sharp and the diffuse
series.
Without, at this point, raising the question as to which of the
series of lines predicted by the Bohr model of the atom is the
principal, the sharp or the diffuse series, or how doublets and
triplets arise in spectra, we see that the Bohr model at once sug-
gests a law very similar to the Rydberg-Schuster law. For
example, the Balmer series in hydrogen (see Fig. 92) originates
from electrons dropping from outside orbits into orbit number 2.
The convergence frequency vB of this series is given by electrons
364 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
dropping from infinity (i.e., ,2 = oo) to the second orbit, i.e.,
by (see equation (31))
hvB = Q- lf7'2
where W2 stands for the energy of the system for r = 2 (i.e., the
second orbit). In like manner, the convergence frequency vL
of the Lyman series is given by
hvL = 0 - W1
Now the first line of the Lyman series is given by
hv1L = W2 - W1
= -hvB + hvL
That is, the first line of the Lyman series is given by the difference
between the convergence frequency vL of that series and the con-
vergence frequency vB of the Balmer series. A similar relation
exists between the first line of the Balmer series and difference
between the convergence frequency of that series and the Paschen
series. This result, extended, is the Rydberg-Schuster law.
In Sec. 6 of Chap. IX, mention was made of the existence of
"combination" lines, which, while belonging to none of the
recognized series, were characterized by the fact that the fre-
quency of such lines was found equal to the difference in fre-
quency of two other lines. Bohr's theory forms a ready picture
of the production of such lines. For example, the frequency
V4B of the fourth line of the Balmer series is given by
hv4B = w6 - W2
and the frequency of the second line by
hv2B = w4 - W2
Their difference in frequency is
h(v4B - V2B) = w6 - w4
This frequency difference should, therefore, be the frequency of
the second line of the Brackett series (p. 258), the wave number
of which is 3,800. The wave number of the fourth line of the
Balmer series is 24,372 and of the second line, 20,564; and their
difference is 3,808-equal, within the limit of measurement, to
the wave number of the second line of the Brackett series.
8. Elliptical Orbits in Bohr's Theory.-(a) In the general case
of motion under the action of a central force obeying the inverse
I
SEC. 8] ELLIPTICAL ORBITS 365
square law, such as the motion of a planet around a massive
central sun or of an electron around a massive nucleus, the path
is, as was shown by Newton, an ellipse with the central body at
one of the foci. We have simplified the preceding discussion by
considering only circular orbits. In this section, we shall discuss
the more general case of elliptical orbits, partly to emphasize the
physical principles involved and partly to extend the concept of
quantum numbers. (The treatment is partly adapted from
Sommerfeld's '' Atomic Structure and Spectral Lines.'')
In Fig. 94, let the path of the electron, charge e, mass m,
around the nucleus, charge +E and mass M, be the ellipse,
the semimajor axis of which is a; the semiminor axis, b; and
bI
I
I
Frn. 94.-The motion of an electron in an elliptical orbit about a nucleus.
the eccentricity, f = c/a; 0 being the center of the ellipse, and the
nucleus_ being at one focus which is distant c from 0. Let the
motion be described in terms of the two polar coordinates r and ¢.
At any instant when the Langential velocity is v the momentum is
mv. This velocity v has components: (1) vr, a radial component
equal to dr/dt (or, more conveniently, r), and (2) a component at
right angles thereto, equal tor :t (or ref>). Likewise, correspond-
ing to the two coordinates we have two momenta: (1) a radial
momentum Pr given by
Pr= mr
. (48a)
and (2) an angular momentum p</J given by
P</J = mr2¢ (48b)
The radial momentum Pr varies throughout the cycle: Pr is zero
at point P1; will then reach a maximum; and will become zero
again at point P2. If, however, we grant (1) that the elec-
tron does not radiate as it traverses the orbit and (2) that the
mass m of the electron remains constant throughout the orbit
366 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
(i.e., does not vary with velocity-we shall consider the genera.I
case later), it is obvious that P<t> must remain constant, since the
force on the electron is always directed toward M, and, therefore,
there is no component offorce at right angles to r
(b) Now, Sommerfeld assumes that to each of these momenta
Pr and P<t> we may apply the phase integral (equation (18))
1 pdq = rh (18)
as follows:
j pq,d</> = T q,h (49a)
f p,dr = Trh ( 49b)
where T<t> is an angular or azimuthal quantum number and Tr is
a radial quantum number. If the state of the atom is defined by
a total quantum number T (identical with the quantum number
which we have considered in previous sections for circular
orbits, for which, of course, Tr = 0, since the radial momentum is
zero for circular orbits), these three quantum numbers are related
by the equation
T=T</>+Tr (50)
For a given state corresponding to total quantum number T,
r <t> and Tr may have various (integer) values, but their sum must
always be equal to r. The significance of this will appear
presently.
Introducing the values of Pr and P¢, from equations (48), into
equations (49), we have
f mr2 <fad<t> = r 1 h (51a)
t mr'dr = Trh (51b)
Since, as we have just seen, mr 2 ¢ is constant, we can, at once,
integrate equation (51a). We then have
21rp</> = Tcph
h
: • P<t> = T </> . 21r (52)
(Compare with equation (20).)
SEC. 8] ELLIPTICAL ORBITS 367
Equation (51b) is not so easily integrated, since it contains
two variables rand r. We shall use the properties of the ellipse
and the equation (52) to express both of these variables in terms
of the coordinate ¢. A relation between r and </> for the ellipse
is given by
1 1 + t: cos</>
- 2
(53)
r a(l - t: )
where a is the semimajor axis and t: is the eccentricity. The
semiminor axis b is
b = a'Vl - t: 2
We shall, also, need to use the value of l/r · dr/d<f>. From
equation (53), we find
dr a(l - t: 2)t:sin </>
(54)
d <P - ( 1 + € cos <P) 2
Multiplying this by 1/r from equation (53), we have
1 dr sin </>
t:
(55)
r d<t> 1 + t: cos </>
From equations (48) we have a relation between Pr and P<t>,
given by
r 1 dr
(56)
Pr = P<t> • r2¢ = P<t, • r2 d<f>
For dr we may write,
dr
dr = -d<t> (57)
d<t>
Inserting into equation (49b) these values of Pr and dr from
equations (56) and (57) gives
~ 1 dr
dr
dcp . dicp
'j'P<J> • r2 = r,h
.·.P<1>f(t :~ydcp = r,h (58)
By use of equations (52) and (55), we have, from equation (58),
2
-€
21r
f (1 +
2
sin <P
€ COS <P) 2
d<t> = -Tr
T <l>
(59)
368 THE NUCLEAR ATOM AND SPECTRA [CHAP.. x
We now have a single variable</> under the integral. Integrating
this equation by parts, 1 we obtain, first,
T
Tr
q, =
[ €
271" 1 +
sin
E
<P
COS cjJ
]
2
O
,r
-
€
271"
1•21r
O 1 +
COS
E
<P
COS cjJ dcjJ.
(60)
The term in brackets is zero for both limits, and there remains
Tr = _ ~ { 2 ,r COS <P d<p (61)
T </> 21r JO 1 + ECOS <p
Equation (61) readily takes the form
Tr -
T </> -
_!_ r21r(
21r JO 1 + € COS
1 - 1) dq, <p
(62)
After integration2
Tr 1 l
T<t, = Vl - €2 -
Therefore,
1- €2 = ( T<t, )2 (63)
T<t, + Tr
.".E = ~1-(Tq,T~TJ (64)
Equation (64) shows us that, for a given value of the total quan-
tum number T = T<t> + Tr, the number of possible ellipses is not
infinite but is limited by the number of possible values of T </J·
If T = T </J + Tr = 4, for example, 4 different ellipses are possible.
1 for each of the values of T <t> = 1, 2, 3, or 4. For T </J = 0, the
eccentricity would be 1, and the path of the electron would be a
straight line through the nucleus. This value of T </J is, therefore,
ruled out. We can more readily interpret equation (64) further
after computing the energy W of the system corresponding to
the various elliptical orbits.
(c) The energy W of the system when the electron is at some
such point in its orbit as is represented in Fig. 94, is partly
potential U, and partly kinetic T. The total energy W is
W = T + U (65)
1 In the standard equation udv = uv - f
vdu, let f
. sin </>
u = sm </> and dv = (l + E COS </> )2d<1>.
2 See any standard table of integrals.
5EC. 8] ELLIPTICAL ORBITS 369
The potential energy when the electron is at a distance r from
the nucleus is, as in equation (24),
U = _ Ee = _ Ee . 1 + e cos </) (
66 )
r a ·1 - E: 2
by equation (53). The kinetic energy is
T = }~mr 2 + }~m(r¢) 2
(67)
\Ye wish to express T in terms of </) only. By multiplying and
dividing the first term on the right-hand side of equation (67)
by m and the second term by mr 2 , and factoring, we have
2
2 r21 P<t>2)
T = 1m ( Pr+ (68)
Substituting for Pn by equation (56), givecs
T = Pl .
2m r 2
_![(! dcp
r
dr)
2
+ iJ' (69)
By use of equations (53) and (55), equation (69) readily reduces
to
T = J!L € 2 + 2€ cos ¢ + 1
(70)
2ma 2 (1 - E: 2 ) 2
We can now write the expression for the total energy of the
system,
W = T + U = pJ . t
2
+
2t cos </> +
1 _ Ee . 1 + t cos<!> (7l)
2ma 2 (1 - E: 2 ) 2 a 1 - E: 2
(d) Now, granting that there is no radiation, the energy of the
system is constant for all values of </). Therefore,
dW = 0
d<t>
Differentiating equation (71), we have
dW = 0 = _ pJ . 2t sin </> + Ee . t sin </> (
72 )
d<t> 2ma 2 (1 - e2 ) 2 a 1 - e2
Solving this equation for a gives
a = P<t> (73)
mEe(l - E: 2 )
Inserting the values of P<1> and of (1 - E: 2) from equations (52)
and (63), respectively, we finally have for a, the semimajor axis
of the ellipse,
(73')
370 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
This equation, comparable with equation (28), shows that the
semimajor axis of the ellipse depends only on the sum of the two
quantum numbers and is, therefore, the same for all ellipses
(including the circle). of the same total quantum number r,
irrespective of the particular values of r </> and r r·
The semiminor axis b, however, depends on the azimuthal
quantum number r <1>· For, since
b = aV1 - e2
we find
h2
b = Tq,(T<t, + Tr) 4 2
1r m e
E (74)
For a given total quantum number r, therefore, the semiminor
axis b is proportional to the azimuthal quantum number r <1>·
The perihelion distance d = (a - c) (Fig. 94) is given by
d = a(l - e)
= (rq, + Tr)\1r2~Ee( 1 - ~1 - (rq, ~ r,) 2) (75)
For a given total quantum number (r<I> + rr), the perihelion
distance d becomes smaller the smaller the value of r <1>·
(e) Returning to equation (71) for W, inserting therein the
p2
value of_± as determined from equation (73), we obtain
ma
W = a(l Ee
- e2)
[e 2
+ 2t: cos¢+
2
1 _ (l +
e cos ¢
)]
Ee t: 2 - 1
= a(l - t: 2) • - ~
Ee
(76)
2a
In other words, the energy W of the system depends only on the
major axis 2a of the ellipse, and since, as we have seen, this major
axis depends only on the total quantum number, the energy of the
system is the same for all ellipses having the same total quantum
number. By putting into equation (76) the value of a determined
by equation (73'), we finally have, for W,
21r 2mE 2e2 1
W = - - -2 - · (77)
h (r<t, + r,) 2
Introducing, as before, Bohr's postulate that, when the system
changes from a state designated by the total quantum number
SEO. SJ ELLIPTICAL ORBITS 371
r 2 to a state of lower energy designated by r1, (r2 > r1), a quantum
of energy hv is radiated the magnitude of which is given by
hv = W T2 - W Tl
we obtain, for the frequency of radiation (we change to wave
number n),
n = 21r 2mE 2e2[ _ _
ch 3
1____1_]
( T <I> + Tr )I (T <I> + Tr)~
(78)
where the subscripts refer to the two different states .. The value
of r <I> and Tr in (r <!> + rrh may assume any integer values so long as
(r <I> + Tr)i = r1; and so for (r <!> + rr)2
For given values of the total quantum numbers r1 and r2, this
equation is numerically identical with equation (41),
n = 21r2mE2e2(_! -
chs T2
1
l)
r2
2
( 41)
derived from a consideration of circular orbits only. The exten-
sion of the discussion to elliptical orbits, therefore, has apparently
introduced no new frequencies. The significance of equation (78),
however, lies rather in the fact that it emphasizes the importance
of the two quantum number ref> and Tr introduced to describe in
quantum terms the more general case of elliptical orbits. The
various possible orbits for total quantum numbers 1, 2, 3, and 4,
respectively, are shown in Table VI. The column headed
TABLE VI.-PossrnLE ORBITS FOR VARIOUS AzIMUTHAL AND RADIAL
QUANTUM NUMBERS
Orbit and Major Minor
T Eccentricity
designation axis axis
1 1 0 Circle 11 a a 0
2 2 0 Circle 22 4a 4a 0
2 1 1 Ellipse 21 4a 2a 72y'3
3 3 0 Circle 33 9a 9a 0
3 2 1 Ellipse 32 9a 6a V3v'5
3 1 2 Ellipse 31 9a 3a V3v's
4 4 0 Circle 44 l6a 16a 0
4 3 1 Ellipse 43 16a 12a ~iv'7
4 2 2 Ellipse 42 l6a Sa ~4 v'12
4 1 3 Ellipse 41 16a 4a ~v'15
372 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
"Designation" is in a terminology, now largely superseded but
useful for our purpose, in which the azimuthal quantum number
r <1> appears as a subscript to the total quantum number r. Thus,
the orbit 42 represents the orbit with total quantum number 4 and
azimuthal quantum number 2. 1
When the atom is in the state designated by r = 4, the electron
may occupy any one of the four orbits. For r = 2, there are two
orbits. So far as anything which we have said up to the present
is concerned, transitions from the state r = 4 to the state r = 2
could take place in any one of eight different ways: from each
one of the four orbits for r = 4 to each one of the two orbits for
r = 2. According to our assumptions, namely, (1) that the
atom does not radiate while the electron is revolving in a privi-
leged orbit and (2) that the mass of the electron is constant and
independent of velocity, all orbits for a given quantum number
should have the same energy (equation (77)). And, therefore,
these eight transfers should, as mentioned in the preceding para-
graph, result in the emission of identical frequencies. We shall
see in the next section, however, that the second of these two
assumptions is not justified.
9. Fine Structure of Spectral Lines.-(a) It has long been
known, both from experimental results and from theoretical con-
siderations, that the mass of bodies should depend on their velocity.
Many years ago, Sir J. J. Thomson showed that, on the hypoth-
esis that the mass of the electron is of electromagnetic origin,
the mass should increase with velocity. The variation of mass
computed by Thomson 2 on this hypothesis was in qualitative
agreement with the experiments of Kauf mann,3 who measured
the values of e/m for the high-velocity electrons ejected by radio-
active substances. Not until the velocity becomes an appreci-
able fraction of the velocity of light, however, is there any
measureable increase in the mass.
Later, it was shown that, on the principle of relativity, an
increase in mass with velocity is to be expected, and Lorentz
1 The reader will find it instructive to sketch these families of ellipses to
scale and in correct relative positions around a central nucleus. The
absolute value of a, say for the hydrogen atom, is given by putting the known
quantities in equation (73').
2 See, for a brief account, THOMSON: "Corpuscular Theory of Matter."
3 Ann. Physik, vol. 29, p. 487 (1906).
SEC. 9J FINE STRUCTURE OF LINES 373
computed that the mass m of an electron moving with velocity v
should be
mo
(79)
m=~-~
1--
c2
where mo is the "rest" mass, i.e., the mass for very small veloci-
ties, and c is the velocity of light. This expression for the vari-
ation of mass with velocity is much simpler than the one deduced
from the electromagnetic hypothesis of the variation of mass and
leads to slightly different results. There has been much con-
troversy and much weighing of both theoretical and experimental
evidence to decide whether the relativity law of variation of mass
is the correct one. For our purpose, the important point is that,
qualitatively, a variation of mass with velocity is to be expected,
irrespective of the principle of relativity. The expression '' rela-
tivity change of mass" should, therefore, refer to the particular
formulre used to compute masses at various velocities rather than
to the origin of the phenomenon.
(b) Now, in the elliptical orbits to which we have referred in
the preceding section, the velocity of the electron varies consider-
ably at different points in the orbit: the velocity varies from a
maximum when the electron is nearest the nucleus, i.e., when the
radius vector r is a minimum, to a minimum when r is a maximum.
If we admit the constancy of angular momentum;
mr 2 ~! = constant
Kepler's law of equal areas swept over by the radius vector in
equal times follows at once. For, the area dA swept over m a
time dt during which the radius advances d<t> is
dA = ~~r 2d<J>
and it follows that
dA
2m dt = constant
as long as m is constant. If, however, m varies from point to
point in the path on account of the variation of velocity, condi-
tions are quite different, and it turns out that the path is no
longer an ellipse; is, indeed, no longer a closed figure but is
somewhat as shown in Fig. 95, which represents diagrammatically
an elliptical path referred to a major axis which rotates in the
374 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
plane of the figure about one of the foci as a center at such an
angular velocity as to cause an advance (of the axis) between
successive maximum values of the radius vector r through an
angle P1MP2, the advance being in the same angular direction
as the motion of the electron around the nucleus. We might
describe the path by saying that it is equivalent to an ellipse in
which the radius vector between its successive maximum values,
instead of sweeping through an angle 21r, as in the ellipse proper,
sweeps through an angle (21r LP1MP2) or, more gener-+
ally, through the angle 21r/'Y where 'Y is a quantity less than unity.
The equations of the ellipse, such as those given on page 367
,,,....-- - -.... -......
/ ............
/
/
/
I
I
I
I
I
I
\
\
\
\
'' '
-- ---
Frn. 95.-The effect of variation of mass with velocity on the motion of an
electron in an eccentric orbit around a nucleus.
may be retained, if, in place of </>, is written 'Y<P· Sommerfeld
shows 1 that 'Y is given by
r;-E2e2
'Y = ~ J. - - pic 2
the quantities having the meanings previously assigned to them.
(c) We see that the motion of the electron now involves two
periodicities: one, the period of the variation of the radius vector
between successive maxima; the other, the period of rotation
of the axis around Mas a center. Since the quantity 'Y differs by
only a small amount from unity, the second of these two periods
is much longer than the first. We, therefore, have a motion of the
1 "Atomic Structure and Spectral Lines," Chap. VIII.
~EC. 9] FINE STRUCTURE OF LINES 375
electron somewhat similar to that discussed in connection with
the Zeeman effect, on page 322. And if the electron, in describ-
ing this path, were to radiate according to classical laws, we should
expect to find in the emitted radiation two slightly different
frequencies. The quantum theory predicts somewhat similar
results but gives a very different "picture" of the phenomenon.
The results of an analysis due to Sommerfeld 1 are, briefly, as
follows:
Starting with the equations of the ellipse' similar to those
given on page 367 but with 'Y<P substituted for </>, and following a
line of development similar to that used above for elliptical
orbits, Sommerfeld first deduces a value for the energy W of the
system for a given quantum state r, assuming the mass of the
electron to vary with velocity according to equation (79). His
final equation for W (analogous to equation (77) but very
different in form) is
W = -moc 2 (az)2
+ moc 2[ 1 + [rr + v\; J-H (80)
- (aZ) ]2 2
where m 0 is the "rest mass" of the electron (previously called m),
a =
21re 2/ ch, Z is the nuclear charge, and the other quantities
have meanings as previously assigned to them.
For comparison 2 with the equation (77),
21r 2mE 2e2 1
W = - - - - -2 [77]
h2 (rr + 7¢)
which is the equation for W derived by disregarding variation of
mass with velocity, we may change equation (80) as follows:
For simplicity in writing, put
(aZ) = p
Tr+ Vr_!_--(aZ) 2 =S
Equation (80) then takes the form 3
W = - moc 2 + moc 2
[ + ~2]-~
1 2
= -moc2 + moc2 [ 1 -
1
2 s2
p2 3 p4
+ 8 84 - ... ]
1 Loe. cit.
2 Remember that E = Ze.
3 (a + x)n = an + nan-Ix •
376 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
2
. wh'1c h we may neg1ec t t erms of h'1gh er m·der, smce
. P Pt
m 82 ~ .
Since
1
. ·s -
1
.. 32 -
and
1 1
84 - (Tr+ T<t,) 4
Putting into the above equation for W these values of 1/8 2
and 1/8 4 (in all of which we neglect terms of higher order than
p 4 / 8 4 ) and replacing p and 8 by their equivalents, we have, after
sim plifying1
w = - 2
21r mZ e
2 4
[ __ 1 + 41r Z e (Tr + Tq,
2 2 4
- 3) 1 l
h2 (Tr+ T<t,)
2
c 2h 2 T<t, 4 (Tr+ T<t,) 4 J
(80')
This equation is seen to differ from equation (77) in that it
contains a second term within the square bracket, which term
involves the ratio of the total quantum number (Tr + T <t>) to T <t>·
W, therefore, depends not simply on the total quantum number
(Tr + r <t>), as in equation (77), but also on how the total quantum
number is distributed between T <t> and Tr. Thus, for total quan-
tum number 3, the parenthesis within the square bracket may
have any one of the three following values:
Tr + Tq, 3 3 3 1
For Tq, = 3 and Tr = 0: Tq, - 4 =
3- 4= 4
Tr+ Tq, 3 3 3 3
For T"' = 2 and Tr = 1: T"' - 4 - 2 - 4 = 4
For TA.
'+'
= 1 and T
r
= 2 · Tr
•
+T
Tq,
</> - ~4 = ~1 - ~4 - 49
1 We drop the subscript of m in equation (80'); i.e., m in equation (80')
is the "rest mass" of the electron.
SEC. 9] FINE STRUCTURE OF LINES 377
W is (algebraically) smaller the smaller r <1>, for a given value of
(rr + r<t>).
In other words, an ellipse of given total quantum
number r represents an orbit of lower energy level than the cor-
responding circle; and the greater the eccentricity of the ellipse
the lower the level.
Consequently, an electron dropping from some given outside
orbit into the ellipse r <t> = 1, Tr = 1 will emit a frequency
slightly different from that emitted by an electron in dropping from
the same outside orbit into the circle r <t> = 2, Tr = 0. The differ-
ence in these two frequencies is determined by the difference
in the energies for the two states of quantum number 2.
More specifically, let W 22 (the subscript 22 designates the cir-
cular orbit, as in the terminology for orbits given in Table VI)
be the energy corresponding to the circular orbit 22; and W 21 the
same for the elliptical orbit 21; and let WT be the energy cor-
responding to some given outside orbit r. Then, the frequency
v2 2 corresponding to the transition r --? 22 is given by Bohr's
quantum postulate
(8la)
and, similarly,
(81b)
The difference in the two frequencies is given by
(82)
To illustrate: The Balmer series in hydrogen originates in transi-
tions from outside orbits into either of the two orbits correspond-
ing to r = 2. The frequency of a given line, therefore, should
be slightly different according as the electron falls into the ellipse
or the circle. In other' words, the lines of the Balmer series should
be doubtlets, close doublets, since the difference in energy of the
two orbits is not very great. Spectral lines made up of such
components of slightly different frequency are said to have a
"fine structure."
Sommerfeld shows that, neglecting small, second-order quan-
tities (which, however, are of importance in the heavier elements),
the difference in wave number ~n arising from the difference in
energy between two such orbits (more correctly, "states") of
the same total quantum number r but of different azimuthal
378 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
quantum numbers r ¢ and radial quantum numbers Tr should be
given by 1
24
a Z
74
[(Tr) 1
T¢
(Tr) J
T¢ 2
(83)
2 7T"e 2
. t he R y dberg constant 2 ;
wh ere Roo 1s a = cf:; · t h e atomic
Z 1s ·
number of the element; Tis the total quantum number; and the
subscripts refer to the different distributions of Tr and r ¢,
Subject to T = T r + 7¢•
(d) Let us apply equation (83) to the Balmer series in hydrogen
(Z = 1) which originates in the "dropping" of electrons from
outside orbits into either the circle of quantum number 2 or the
ellipse of quantum number 2. For the circle r ¢ = 2 and Tr =
0; for the ellipse, T ¢ = 1 and Tr = 1. We, therefore, have
a2
= R 24 = 0.365 cm. - 1
00
(83')
takingR = 109,700;e = 4.77· I0- 10 e.s.u.;c = 3 · 10 10 cm.sec.-·1 ;
h = 6.55 · 10- 21 erg. sec. This gives the difference in wave
number ~nH to be expected between the two components of
the several lines of the Balmer series if these lines are actually
doublets. This difference in the case of Ha amounts to only
about 3 parts in 150,000-about one-fiftieth of the distance
between the sodium D lines. Nevertheless, in spite of very great
experimental difficulties numerous investigations have shown
that the lines of the Balmer's series are actually close doublets
and that the difference in wave number between the two compo-
nents is of the order of that predicted by Sommerfeld's theory.
Sommerfeld (Zoe. cit., p. 482) quotes the following measurements
for the difference in wave number between the components of
the several lines:
1 The derivation of this equation is left to the student. Starting with
equation (80') (analogous to equation (29), obtain an equation for n (analo-
gous to equation (33')). Equation (83) follows at once.
2 See equation (47).
SEC. 9] FINE STRUCTURE OF LINES 379
CENTIMETERs-1
Michelson ..................... Ha: ~nH = 0.32
Michelson ..................... Hs: = 0.33
Fabry and Buisson . . . . . . . . . . . . . Ha: = 0. 306
Meissner and Paschen .......... Ha: = 0.288
Ha: = 0.272
Gehrcke and Lau............ Hf3: = 0.283
l H'Y:
To these may be added the more recent measurements of Shrum :1
= 0.271
CENTIMETERs-1
Ha: ~nH = 0.33
Hf3: = 0.36
H'Y: = 0.37
H0 : = 0.36
HE: = 0.35
Still more recently, W. V. Houstoun, 2 working with hydrogen
at liquid-air temperature finds
CENTIM.ETERs-1
Ha: ~nH = 0.315
Hf3: = 0.331
H'Y: = 0.353
There can, thus, be no doubt as to the existence of the effect.
As to its magnitude, there has been considerable controversy,
arising, in large part, from the theoretical importance attached to
Sommerfeld's formula, coupled with the fact that many of the
measurements quoted above tend to somewhat lower values than
those predicted by equation (83'). In view of subsequent
developments, such as the "spinning'' electron of Uhlenbeck and
Goudsmit and the wave theory of Schrodinger-which provide
alternative explanations of fine structure and of many other
spectral and atomic phenomena-it is pertinent to emphasize
the experimental facts. Theories regarding those facts will
ultimately become stabilized.
(e) One further comment, however, regarding the implications
of Sommerfeld's equation (83) for fine structure may be in order.
One of the experimental difficulties in making accurate measure-
ments of ~n lies in the fact that the actual positions of the com-
ponents of the several lines is made somewhat uncertain by
the broadening of the spectral lines due to the Doppler effect.
The magnitude of this effect depends on the mean velocities of
1 Proc. Roy. Soc., vol. 105, p. 259 (1924).
2 Astrophy. Jour., vol. 64, p. 81 (September, 1926).
=380 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
the molecules or atoms of the gas. Even at liquid-air tempera-
ture, the theoretical width of the lines of the hydrogen spectrum
• 0
1s of the order of 0.05 Angstrom, or one-seventh of the actual
expected separation of the fine-structure components. The
heavier the gas molecule, however, the less its velocity at a given
temperature. And, further, equation (83) shows that ~n increases
as the fourth power of the atomic number. The fine structures in
the lines of helium should be subject to greater precision of
measurement than corresponding lines in hydrogen, since Z = 2
for helium, which means a gain of 2 4 in separation (for the same
quantum transfers as hydrogen); and the mass of the helium
atom is four times that of the hydrogen atom, which means that,
because of the smaller velocity of the helium atom at a given
temperature, the Doppler broadening in helium should be
approximately one-half that in hydrogen. Paschen 1 measured
the ~n's for the components of the "4686" line of helium (of
which there are five, see below) and, computing from these
measurements the value of ~nH (for hydrogen), found
~nH = 0.364 + 0.004 cm.- 1
in complete agreement with Sommerfeld's prediction.
10. The Selection Principle.-The hydrogen line Ha arises
from transitions from the quantum state r = 3 (three possible
orbits) to the quantum stater = 2
A B C (two possible orbits). Six differ-
T=J ~,O '-.. 2,1 J,2 ent transfers from r = 3 to r = 2
' ....... ..,..I
I
I might, therefore, be expected.
I ....._
We have seen that transfers from
any outside orbit to the orbitr = 2
O b should result in a doublet. We
Frn. 96.-The origin of the six might expect that Ha should be
mathematically possible components made up of six components or of
of the hydrogen line Ha.
three doublets, one doublet corre-.
sponding to each of the three orbits r = 3. We can represent
these mathematically possible components of Ha in the schematic
way shown in Fig. 96. The numbers are the quantum num-
bers of the respective orbits-the boldface numbers being
the azimuthal quantum numbers, and the others the radial
quantum numbers. Thus, the orbit "2, 1" is the orbit T<f> = 2
and Tr = 1. For brevity in writing, we shall designate the orbits
1 Ann. Physik, vol. 50, p. 901 (1916).
SEC. 10] THE SELECTION PRINCIPLE 381
by the letters A,.B, C, a, b; and any given transfer by an obvious
combination of the two letters; e.g., "A b" represents the trans-
fers from the orbit 3, 0 to the orbit 1, 1. The arrows represent
the six possible transfers. We shall see, presently, the reason for
dotting some of the arrows.
From an equation for the wave number of a spectral line based
on equations (80) and (81) we might compute the wave numbers
of each of these six components. It will be simpler, however, to
compute differences in wave numbers and to represent the location
of each of the six components graphically from some one com-
ponent as an arbitrary starting point. Using the equation (83)
(83)
we can compute the difference in wave number ~n for any pair of
transfers represented in Fig. 96. Thus, ~nib, the difference in
wave number between the lines Aa and Ab, is given by
.6.nlt = Rooa 2Z4. ;4 (% - H) = -Rooa 2 Z4. H6 = - 0.365 cm.- 1
and, similarly,
.6.n~: = R 00 a 2Z4 · ii{i - ~) =R 00 a 2 Z4. .
2 81
3
= 0.110 cm.- 1
In this way, we obtain data for Table VII, which shows the wave-
number differences for a sufficient number of pairs to locate the
components of Ha.
TABLE VIL-WAVE-NUMBER DIFFERENCES BETWEEN THE FINE-STRUCTURE
COMPONENTS OF Ha
Multiplier ~'1, in
of R 00 a 2Z 4 cm.- 1
~nit 1~G-D -0.365
~nit 11/~-D -0.365
~ng: i6G-D -0.365
~nAa
Ba s\G-i) 0.036
~nBa
Ca li(I-D 0.108
382 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
From these wave-number differences, and starting with, say,
the line Aa as an arbitrary origin, the position, on a wave-num-
ber scale, of the six components of Ha may be located as shown in
Fig. 97. This is the mathematically possible fine structure of Ha.
(The student will find it instructive to locate, in this way, the
corresponding 8 components of HfJ or the possible 12 components
of the helium line "4686," which is made up of transfers from
T = 4 to T = 3.)
Now, experimental evidence based upon observation on lines
where one can obtain greater resolution than is possible in Ha
1
- -- --0.36S - -· -->J
t I
shows that only a limited number of
these mathematically possible lines are
r
I I I
I
I
I
I
I
I
actually present. Rubinowicz, in 1918,
t
I I
I
I
I from a consideration of the conservation
Ca BaAa Cb BbAb of angular momentum as applied to the
FIG. 97 .-The observed interchange of angular momentum be-
components of the hydrogen
line Ha. The dotted compo- tween the atom and the emitted radia-
nents are ruled out by the tion during any interorbital transition,
Eelection principle. The two
left-hand components are so showed that in such a transition the
close together as to be ob- azimuthal quantum number could
served spectroscopically as a
single line. change by only + 1, no restriction, how-
ever, being made to the change in the
radial quantum number. An identical conclusion has been reached
by Bohr on quite different grounds. This limitation of the change
in the azimuthal quantum number makes it possible, by ruling out
the other changes, to'' select'' for any given line the components
which ought to be present; hence, the name "selection principle,''
which states that only those interorbital transfers are possible for
which the change Llr</> of the azimuthal quantum number is given by
Llr </> = +1 (84)
The prohibited lines are those which are dotted in Figs. 96 and
97. For example, the line Ab (Fig. 96) is ruled out, since for it
Ar</> = 2. The line Ca is permitted, since for it /j.7 = -1. The
lines of the Balmer's series, therefore, should really be made of
three components instead of only two, as discussed in the preced-
ing section. Both experiment and theory show that not all of
these three components have the same intensity, the component
Ca (Fig. 97), for example, being much weaker than Aa and suffi-
ciently close to Aa to prohibit separation spectroscopically.
Measurements, therefore, give the center of gravity of these two
SEC. 11] SYSTEMS WITH SEVERAL ELECTRONS 383
components, and the separation between this center and the com-
ponent Bb is the ''doublet'' difference for the line Ha, as actually
measured. For the higher components, HfJ, H'Y, H 0 , etc., the
separation between Ca and Aa become less, and the doublet differ-
ence as observed becomes more nearly that predicted, a conclu-
sion in agreement with the measurements of Houstoun, quoted
on page 379.
The reader, by application of the selection rule, can readily
determine which components ought to be expected in other lines,
say the helium line "4686," 1 and can locate their relative
positions.
The selection principle, which we have here applied to the
azimuthal quantum number, has been extended to other quantum
numbers to which we shall make reference in subsequent chapters.
11. Systems with More than 1 Electrono-Bohr's theory, in
the simple elementary form, as presented in this chapter, is not
applicable to atoms containing more than 1 electron revolving
around the nucleus, since the presence of other electrons very
materially affects the energy W corresponding to a particular
orbit, or state. The mathematical difficulties of the famous
'' three body'' problem prevent a rigorous extension of the Bohr
equations for spectral series. Nevertheless, the Bohr theory has
shown the way, partly by analogy, partly by empirical methods,
for notable advances in the interpretation of spectral phenomena
and their relation to atomic structure. Some of these we shall
mention in the remaining articles of this chapter. Others we
shall leave until later chapters, after we have considered some
developments along other lines, particularly the evidence which
led to the hypotheses concerning the distribution of electrons in
atoms.
12. The Absorption of Energy by Atoms. (a) Excitation and
I onization.-In the preceding articles of this chapter, we have
discussed the emission of characteristic radiation by atoms. In
terms of Bohr's "picture" of the atomic mechanism, we have
discussed the phenomena occurring when electrons '' drop from
outside orbits into inner orbits," the various orbits or, more
generally, stationary states of the atom, being characterized by
certain quantum numbers, which, together with certain constants,
determine the energy WT corresponding to a given stater. We
1For the actual measure men ts on this helium line, see P ASCHEN: Ann.
Physik, vol. 50, p. 901 (1916).
384 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
have pointed out that, by analogy with mechanical systems with
which we are familiar, we should expect the normal state of the
atom to be the state having the lowest energy. For the simple
atom (i.e., nucleus plus 1 electron), this is the state for which
T = 1; or, again in terms of the "picture," the normal state of
1
the simple atom is the state in which its electron is revolving in
the innermost privileged orbit, the circle.
When, still speaking of the simple atom, the electron is revolv-
ing in orbits of higher quantum numbers than the normal orbit,
the atom is said to be in an excited state, and the process of trans-
ferring the atom from the normal state to an excited state is
referred to as excitation. Jt may happen that the excitation is so
intense that the electron is completely removed from the nucleus.
The atom is then left with a net positive charge and, if free, will
tend to move in an electric field, like an ion in an electrolytic
solution. Such an atom which has lost 1 electron (or more) is
said to be in an ioni.zed state, and the process of ''raising'' the a tom
from the normal state to the ionized state is called ionization.
1 It should be emphasized that we should not take this picture of the atom
with its several electrons revolving in the various "privileged" orbits too
seriously, at least so far as concerns the actual physical makeup of an atom.
The concept of orbits and of the "dropping of electrons from one orbit to
another" is of value largely in helping us to keep in mind the observed phenom-
ena and in suggesting other previously unknown phenomena which, in turn,
suggest new experiments. It is probable that the picture of an atom with
a planetary system of electrons is as far from the real architecture of the
atom as are "lines of force" from the actual structure of the magnetic field.
We find the concept of lines of force very useful in helping us to visualize
certain phenomena. We compute electromotive forces by ascertaining
the number of lines cut per second. Yet we know that, physically, there are
no such lines in a magnetic field as we find it convenient to visualize. We do
know, however, that irrespective of any concrete picture of its structure, a
magnetic field involves energy, the amount of which per unit volume at any
point we can compute by dividing a certain quantity H2, characteristic of
the field at the particular point, by 81r.
In a similar way, the concept of orbits, electron transfers, and the like
is a very concrete and useful geometrical "picture" to help us keep in mind
the various states of the atom. But whether the picture be correct or not, we
are reasonably sure that the energy associated with these various stationary
states is real. Sometimes it is more convenient to speak in terms of orbits;
at other times, the more general phrase "energy corresponding to a given
state" will be much more appropriate. We should always keep in mind,
however, that the latter phraseology is physically justified; the former
is quite artificial.
SEC. 12] ABSORPTION OF ENERGY 385
If, now, an atom is to radiate by a transfer from an excited (or
an ionized) state to one of lower energy, the atom must first, by
some means, be put in that excited state. That is, the atom
must absorb energy sufficient to raise it from the normal to the
excited state. This process of absorption is the converse of the
process of radiation.
A full discussion of the entire subject of the excitation and
ionization of atoms, and of the far-reaching conclusions drawn
from experiments in this field, is quite beyond the scope of this
book. The reader is referred to more exhaustive or special
treatises on the subject. We may, however, make brief mention
of some of the ways by which experiment shows that atoms may
become excited or ionized and of the interpretation of some of
these experiments.
(b) ColUsions between Atoms and Electrons or between Atoms and
Other Atoms.-The atoms of a gas (or vapor) may become excited
To fhe C--------
Specfrometer Sr-r-v-.....,,...,...~---..___,
Ear fl,
Frn. 98.-The apparatus of Foote, Meggers, and Mohler for studying excitation
potentials.
or ionized by bombarding them with electrons possessed of
sufficient kinetic energy. If the energy given to an atom by
such a collision is sufficient to expel an electron from its normal
orbit in the atom to infinity, the atom may become ionized.
Otherwise, it becomes simply "excited." As one illustration of
the mode of procedure followed in studying phenomena of this
kind may be mentioned the experiments of Foote, Meggers, and
Mohler. 1 Their apparatus is shown diagrammatically in Fig. 98.
A filament F of tungsten or lime-coated platinum is heated, by a
battery B1, to such temperature that it emits electrons. Around
the filament is a spiral grid S which, by means of a battery B 2 or
potentiometric source of pd, is maintained at any desired positive
potential with respect to the filament. Around S and electrically
connected thereto is a metal cylinder CC; S and CC are thus at
the same potential. Inside CC is the gas or vapor under study,
maintained at a suitable pressure.
1 See FooTE and MoHLER: "The Origin of Spectra."
386 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
Electrons, emitted by the filament, are accelerated toward the
grid S which is so close to F that with proper regulation of the
gas pressure comparatively few atoms of the gas are struck by
the electrons in their passage from filament to grid. After
leaving the grid, the electrons move in the (electrical) force-free
space between S and CC, in which space, because of the greater
distance S to CC, the electrons collide with the gas molecules,
causing the excitation of the latter, provided the velocity (energy)
of the electrons is sufficiently great. vVith sodium vapor inside
the cylinder, no luminosity is observed in the vapor until the
potential difference between the grid and the filament reaches
2.09 volts. For voltages slightly above this value, the spectro-
graph shows that the sodium vapor emits the well-known D lines,
and those only. The mean wave length of these lines is 5,893
0
Angstroms, corresponding to a frequency v of 0.509 · 10 15 sec.- 1 ;
the quantum energy hv is 6.55 · 10- 21 X 0.509 · 10 15 = 3.33 ·
10- 12 ergs. According to the quantum picture of the origin of
spectra, there must have been a change of state of the sodium
atom, in emitting this line, corresponding to an energy drop of
3.33 · 10- 12 ergs-whatever may have been the absolute values
of the energies in the initial and the final state. Now, the elec-
trons, which, after emission from the filament, have dropped
through 2.09 volts, have acquired a kinetic energy given by
e v = 4. 77 . 10- 10 x 3 ~ 0io2 = 3.32 . 10- 12 ergs
We see, therefore, that the kinetic energy possessed by the elec-
trons as they pass through the space between the grid and the
cylinder is exactly equal to the quantum energy of the radiation
emitted by the sodium vapor. The presumption is very strong,
therefore, that the electrons have, by collision, transferred to
the sodium atoms sufficient energy to raise the latter from their
normal state to an excited state 3.33 · 10- 12 ergs above the
normal, so that the atoms in returning from that excited state to
the normal state emit the line (doublet) A = U;~ig .Angstroms
which, be it noted, is the first line of the principal series of sodium
(seep. 299). We shall return to this latter point presently.
As the voltage between filament and grid is raised above 2.09
volts, the D lines continue to appear, but no others appear until
the voltage reaches 5.12 volts, beyond which many more lines--in
SEC. 12] ABSORPTION OF ENERGY 387
fact, the complete arc spectrum-are produced. Beyond 35
volts, the so-called "spark" spectrum is seen. With these latter
observations, important as they are, we are not now concerned.
Let us consider a little further the significance of the observation
(1) that up _to 2.09 volts no emission of radiation occurs and (2)
that above this voltage only the single pair of lines, the D lines, is
produced until the voltage becomes considerably higher.
Electrons possessing energy equivalent to, say, 3.0 volts should,
so far '8iS energy is concerned, be able to communicate to sodium
atoms sufficient energy to generate lines of wave length 4,100
0
Angstroms, or longer. 1 In the spectrum of sodium is found a
large number of lines of longer wave length than this, for example
practically the entire sharp and the diffuse series, the convergence
0
wave length of which is of the order of 4,100 Angstroms. Why
do not '3-volt electrons excite these lines? Or, the wave length
of the second line of the sharp series is 6,161 Angstroms, corre-
sponding to about 2.01-volt electrons. Why is this line not pro-
duced at lower voltages than the D lines which correspond to
2.09 volts?
We can more adequately answer these questions after consider-
ing, further, the various excited states of the sodium atom and
their relation to the several known spectral series, as viewed in
the light of the arrangement, in the sodium atom, of the 11
electrons which it is known to containo A clue to the answers to
the above questions, however, is furnished by considering the
stationary states of the simple (hydrogen) atom (nucleus plus 1
electron), as represented by the circular orbits in Fig. 92.
It may be assumed that, normally, the electron of the hydrogen
atom will be in the lowest level, the circler = 1. In order that
the atom may be excited so as to radiate, it must be given at least
Assuming that the quantum energy hv of the emitted lines is equal to the
1
energy acquired by the electron in falling through V volts, we have
Ve = hv = h~
A
A= ch
Ve
In this formula, Vis to be expressed in electrostatic units of potential differ-
ence, which are given by vil~~ 2 • H Vis kept in volts and the numerical
3
values of c, h, and e are inserted, we find for A the expression, convenient to
remember,
, 12,345 Ao t ..
I\ = -V- ngs roms.
388 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
a sufficient quantity of energy so that the electron will be raised to
the second orbit r = 2. The electron may then drop back to the
first orbit and, in so doing, emit the first line of the Lyman series
(X = 1,216Angstr6ms), which, computing as above, should require
about 10-volt electrons. Electrons possessing less ~nergy than
that corresponding to 10 volts, if they were to collide with normal
atoms, could not raise the atom from the normal to the first
excited state, and they could, therefore, excite no radiation. To
such collisions, in which the atom receives no energy from the
electron, is applied the term "elastic," by analogy with mechan-
ical systems; while collisions in which the atom is raised from
one state, say the normal, to another, are called inelastic collisions.
To raise the electron from the first orbit to the third (Fig. 92)
should require some twelve volts. (The wave length of the second
0
line of the Lyman series is 1,026 Angstroms.) Once the electron
is in the third orbit, it could return to the normal state either by
dropping to the second orbit (thereby producing the first line of
the Balmer series) and thence to the normal orbit, or it might go
di"rectly from the third orbit to the first. We see then, that,
although the wave length of the first line of the Balmer series
0
(X = 6,563 Angstroms) is such as to correspond to an energy
transfer in the atom equivalent to only about two volts, this line
could not be produced un]ess electrons corresponding to at least
12 volts were to collide with the normal atom.
Returning to the experiments with sodium vapor and assuming
that the sodium atoms before collision with the 2.09-volt elec-
trons were in the normal state, we conclude that the impact with
the 2.09-volt electrons was sufficient to raise the sodium atom
from the normal state to the first excited state above normal.
On returning from that excited state the atoms emit the D line-
the first line of the principal series. Continuing the analogy with
hydrogen, before the sodium atom can emit any other line than
the first line of the principal series it (the atom) must be raised
to some higher stationary state than the second. This requires
more than 2.09 volts.
In passing, we may call attention to the fact that the excitation
of the D lines does not cease when the voltage exceeds 2.09 volts,
but continues at higher voltages.
W c are now also in position to point out the significance of the
observation that with sodium vapor the entire arc spectrum of
the element is emitted when the potential difference between
SEC. 12] ABSORPTION OF ENERGY 389
filament and grid exceeds 5.12 volts. The convergence wave
number of the principal series of sodium is 41,450, wave length
2,410 Angstroms. Electrons falling from infinity to the normal
state should generate this frequency, the quantum energy value
of which is
hv = 6.55 · 10- 27 X 41,450 X 3 · 10 10 = 8.15 · 10- 12 ergs.
Conversely, electrons possessing this amount of kinetic energy,
which energy should result from their falling through 5.12 volts
(between filament and grid), should, by collision with sodium
atoms, impart to the atoms sufficient energy to raise an electron
in the sodium atom to infinity: i.e., to ionize the atom. In
returning to its normal orbit this electron might "drop" into
any of the intervening privileged orbits and therefore generate
the entire (arc) spectrum of sodium. In other words 5.12 volts
is the ionization potential of the sodium atom.
(c) Excitation by Radiation.-On the classical theory, the char-
acteristic frequencies emitted by an atom should be identical
with, or overtones of, the natural frequencies of the vibrating
"systems" of which the atoms are composed. These "systems"
should, according to the familiar principle of resonance, absorb
(i.e., resonate with) radiation of identical frequencies; and, thus
being set into vibration, the atom should re-emit these same
frequencies. Confirmation of this view of resonance absorption
was apparently to be found in the well known phenomenon of the
''reversals'' of spectrum lines, of which phenomena the most
conspicuous are the dark lines in the solar spectrum. The
"reversal" of the D lines of sodium is a familiar laboratory or
lecture demonstration.
R. W. Wood showed 1 many years ago that a bulb containing
sodium vapor at very low pressure would, when irradiated by
light from an intense sodium flame, emit the D lines, and those
only. To this phenomenon Wood gave the name resonance
radiation.
The resonance is very sharp. An examination of the reemitted
D lines showed that they were very narrow, their width corre-
sponding almost exactly to the width predicted by the Doppler
effect at the temperature of the sodium vapor. The width of
the lines of the exciting source was much greater, since the tem-
1 See Woon, R. W.: "Physical Optics." This book gives an account of
these beautiful experiments on resonance radiation.
390 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
perature of the source was much higher than that of the vapor.
On analyzing the light which had passed through the vapor, it
was found that there was a narrow absorption line at the centers
of the broad D lines. So far, the classical theory seems to agree
with experiment.
But (the p'resent) Lord Rayleigh showed 1 that if sodium vapor
be illuminated by the second line of the principal serias, X = 3,303
Angstroms, both that line and the D line were emitted by the
vapor. This fact cannot be explained by any application of the
principle of overtones, since the frequency of the second line of
the principal series bears no simple relation to the first line, the
. of f requencies
ratio . 30 ,212 = 1. 7784 .
. b emg
16 973
'
The quantum theory offers a simple explanation to observa-
tions such as those of Wood and of Rayleigh. The sodium atom
in the vapor is in the normal state. To raise the atom from this
normal state to an excited state, the return from which (to
normal) emits the D lines, requires that the atom must be given
energy exactly equal to hvD, where VD is the frequency of the D
line. The quantum energy of the incident D lines is exactly
equal to this, and hence the sodium atom is excited by the absorp-
tion of this quantum. This explains Wood's observations. To 0
excite the second line of the principal series, X = 3,303 Angstroms,
requires that the atom should be raised to an excited state higher
than normal by exactly a quantum corresponding to X = 3,303
0
Angstroms. When once in this latter excited state, the atom
could return to the normal level, emitting the line X = 3,303, or
it could return to the intermediate level next above normal
and by dropping from there to normal would emit the D lines.
Absorption of the quantum corresponding to X = 3,303 Ang-
stroms, therefore, should result in the emission both of that line
and of the D line, exactly as found by Rayleigh.
By following the same line of reasoning as that employed in
considering excitation by electron impact, we see why it is that
the normal sodium atom does not exhibit the phenomena of
resonance radiation and absorption when radiated by some such
• • 0
lme as, say, the second line of the sharp series, X = 6,161 Ang-
stroms. The quantum energy corresponding to this line is not
sufficient to raise the normal atom from its energy level even to
1 STRUTT, R. J.: Bakerian Lecture, Proc. Roy. Soc., vol. 96, p. 272 (1916).
SEC 12] ABSORPTION OF ENERGY 391
the first level above normal which, as we have seen, requires a
quantum corresponding to the D lines, X = 5,896 and 5,890
0
Angstroms.
Excitation by the absorption of quanta, therefore, is quite
analogous to exc~tation by electron impact except for this one
very important difference. The atom may become excited
when colliding with an electron the energy of which equals or
exceeds that required for the increase in energy-level; while
resonance excitation takes place only when the energy of the
incident quantum is exactly equal to that required to produce the
particular change of state. 1
(d) Excitation by Collision with Other Atoms.-The atoms of a
gas are continually interchanging energy by collisions resulting
from thermal agitation. We have seen in Chap. VIII that the
mean energy of thermal agitation of a monatomic gas molecule
at 300°C. is of the order of 0.078 X 10- 12 ergs (i.e., ;kT).
This we see is only a fraction of the energy, 3.33 X 10- 12 ergs
(page 386) required to raise the normal sodium atom from its
normal state to the next higher energy level. Very rarely could
it happen that a molecule of sodium vapor at room temperature
would, as a result of a collision with another molecule, acquire
sufficient energy to excite it. If, however, the temperature is
raised, as by putting sodium in the Bunsen flame, the average
kinetic energy of translation may become comparable with 3.33
X 10- 12 ergs; collisions between atoms may then result in raising
an atom to an excited state, and radiation may result. With
further increases in temperature the higher members of the
principal series, and also members of other series should appear.
This is confirmed by experiment.
There is ample evidence to confirm the view that an atom A in
an excited state may lose its energy of excitation by collision with
another atom B, the potential energy of A being transferred to B.
This potential energy may appear in B either (1) as energy of
excitation, in which case B may later radiate its own characteris-
1
If the energy hv of the incident quantum equals (or slightly exceeds)
the energy Ei required to remove the electron from the atom, the
quantum may be absorbed, and ionization of the atom results. Data on
this point are not very extensive. See MOHLER, FooTE, and CHENAULT:
Phys. Rev., vol. 27, p. 37 (January, 1926); FooTE, Phys. Rev., vol. 29, p.
609 (1927).
392 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
tic frequencies; or (2), the kinetic energy of B may be increased,
in which case A, initially excited, has returned to a lower, perhaps
the normal, state with no emission of radiation. The latter kind
of collision, in which the potential energy of an atom in the excited
state appears as kinetic energy after collision, .. is known as a
collision of the second kind-a very unsatisfactory name, since the
name is not at all descriptive of the phenomenon.
As illustrative of these phenomena involving the transfer of
the potential energy of an excited state from one atom to another
may be mentioned the experiments on mercury vapor 1 which is
excited to resonance radiation by the absorption of its own line
X = 2,536 Angstroms (equivalent to about4.9 volts). If with the
mercury vapor be mixed the vapor of thallium, the characteristic
lines of the latter element appear in addition to the mercury
resonance radiation when the mixture is radiated by the mercury
line X = 2,536, the mercury resonance radiation being then weaker
than when no thallium is present. The thallium vapor alone is
not excited to resonance by the mercury line. The phenomenon
is explained by assuming that the mercury atoms are first raised
to an excited state by resonance absorption of the line X = 2,536;
some of these excited atoms by collision with thallium atoms
transfer this potential energy of excitation to the latter, which,
then being excited, radiate their characteristic lines by returning
to the normal state. The thallium lines so produced correspond
to energy transfers in the thallium atom less than 4.9 volts.
Cario showed that the presence of argon in mercury vapor
very materially reduces the intensity of the resonance radiation
of the mercury vapor without, however, exciting argon radiation
when the mixture is radiated by the X = 2,536 line, although an
examination of the light transmitted through the mixture shows
that there is no diminution in absorption as a result of the presence
of the argon. This effect is explained by assuming that collisions
of the second kind take place between the excited mercury
atoms and the argon atoms and that the potential energy of the
former is transferred to kinetic energy of the latter.
(e) Electrical Methods of Observing the Excitation of Atoms.-
Referring to Fig. 98, it is readily seen that if the potential differ-
ence between grid and filament be sufficiently high, the electrons
1 LORIA, S.: Phys. Rev., vol. 26, p. 573 (1925); CARIO, G.: Zeit. fur
Physik, vol. 10, p. 185 (1922); and CARIO and FRANCK, Zeit. fur Physik,
vol. 17, p. 202 (1923).
SEC. 12] ABSORPTION OF ENERGY 393
may be given sufficient energy to ionize by collision the molecules
of the vapor in the space between the cylinder and the grid. In
Fig. 99, the filament Fis so connected to the battery B2 that the
potential difference between F and the grid S is less than, and in
the opposite direction to, the potential difference between CC and
S. Electrons accelerated from F toward S can therefore not
reach CC. The cylinder CC is connected to earth through the
sensitive electrometer E and the grounding key k. So long as
no positive ions are produced in the space between S and CC,
the cylinder should acquire no charge. But when the velocity
of the electrons accelerated from F to S becomes sufficient to
ionize the vapor, the positive ions should be attracted to CC and
a so-called ionization current should be set up which is measurable
by the rate at which the electrometer acquires a charge when the
grounding key is open.
Jc
c----------~~~~
C----------------- Earfh
Frn. 99.
An effect of exactly this kind was observed by Lenard many
years ago. He found that with such gases as air, hydrogen and
C02, a current which he thought to be a true ionization current,
began to flow when the exciting electrons had fallen through a
potential difference of some eleven volts.
We have seen however from the experiments of Meggers,
Foote and Mohler, that in the case of sodium vapor resonance
radiation is produced by 2.09-volt electrons; and in the case of
mercury vapor by 4.3-volt electrons. This resonance radiation,
proceeding in all directions, falls on the inside walls of the cylin-
der, and, if the radiation be of sufficiently short wave length,
photoelectrons will be expelled from the cylinder. Because of the
direction of the field between CC and S these photoelectrons will
flow toward the grid and we shall have in the circuit CB 2S a
photoelectric current in exactly the same direction as if there were
an ionization current due to the ionization of the vapor. This
394 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
"false" ionization current will begin abruptly as soon as the reso-
nance potential is reached.
To separate the true ionization current from the photoelectric
current Davis and Goucher 1 introduced a coarse wire gauze G
surrounding the grid and just inside the plate CC. By means of a
battery B 3 , 2 (Fig. 100), this gauze may be maintained at either a
positive or a negative potential with respect to CC, or G may be
connected directly to CC without a battery. When G is sufficiently
negative with respect to CC, photoelectrons, even if ejected from
CC by the action of resonance radiation, will not reach G. Photo-
electrons if ejected from G, however may reach CC and a photoelec-
tric current may result. If now B3 be reversed, so that G is
positive with respect to CC, a photoelectric current in the opposite
so®®@]]f:
p.,,·
G-------------- Earfh
C------------------
FIG. 100.-The apparatus of Davis and Goucher for studying ionization
potentials.
direction should result. In either case the potential difference
between CC and G is small compared to that between G and S.
The true ionization current cannot be reversed by reversing the
direction of B 3 , since the positive ions produced between S and G
acquire sufficient velocity so that, passing through the gauze,
they will reach CC in spite of any small opposing field between
CC and G.
With this type of apparatus Davis and Goucher obtained, for
mercury vapor, curves of the type shown in Fig. 101, in which
the abscissre are the accelerating potentials between filament
and grid, and the ordinates are the currents to the cylinder CC
as measured by the rate of charge of the electrometer. Curve A
was obtained when the gauze G was positive with respect to the
1 Phys. Rev., vol. 10, p. 101 (1917).
2 Of course, these "batteries" are really variable potentiometric sources of
pd.
SEC. 12] ABSORPTION OF ENERGY 395
cylinder; curve B when the gauze was negative. No current was
observed until the accelerating voltage reached 4.9 volts.
Beyond that voltage a current was observed (a to c in curve B,
negative charging of cylinder; a to c' in curve A, positive charg-
ing of cylinder), which could be reversed by reversing the battery
B3, indicating that the current was a photoelectric current. At
point c corresponding to 10.3 volts a sudden reversal of the
direction of change of current took place, indicating that the
20------------,~~--t-~--,~~~
c..>
J-
a.>
~
.~ 4~--h--~+--.>t---t-~---t-'""l--t-----1
::n I
'-' + I I
__g 6- I 8 IQ
1: 0 1-----+-,:,,~E::::-+.A;--cc+.ele-rcx--.+:T'in9:........P~o-.-te-n+..t1a-';-l-t;(,--vo:-:-lts71)
a.> I I
t- I
3 41--~-+-~~-t-b~-t------i-t-'r---1
Fm. 101.-The resonance and ionization potentials of mercury vapor, as observed
by Davis and Goucher.
cylinder was beginning to collect positive ions. Up to point c,
curve A is qualitatively the mirror image of curve B.
The beginning of the photoelectric current at 4.9 volts is due to
the excitation by electron impacts of the mercury line X = 2,536,
which then acts photoelectrically on the cylinder and the gauze.
When the cylinder is positive with respect to the gauze,
the photoelectrons move from gauze to cylinder; and vice versa.
The rapid increase in photoelectric current which occurs when the
voltage reaches 6. 7 volts is attributed to the generation, at that
voltage, of the mercury line X = 1,849; while the sudden break in
both curves at 10.3 volts is indicative of the beginning of the
396 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
real ionization current. 4.9 volts and 6.7 volts are spoken of as
the resonance potentials of mercury vapor, and 10.3 volts is its
ionization potential.
Many beautiful experiments have been devised to extend the
data and theories regarding the excitation of atoms by impact or
resonance. The reader will find further study in this field, from
original sources, very fascinating. The literature is extensive.
13. Energy-level Diagrams. (a) Hydrogen.-We have seen
that spectral lines originate when an atom changes from one
state to another of lower energy. In Fig. 92 these states were
represented pictorially in terms of Bohr's privileged orbits.
Much more fundamental than this geometrical concept of the
various states is the energy associated therewith. We can
compute the energy of the atom in its various states from the
known value of the spectral terms, the differences between which
give the spectral lines. From these values of the energy, we
can show by a diagram, known as the energy-level diagram or as
the diagram of the stationary states, the origin of the various
spectral lines and their interrelations, in a manner which is
quite independent of any theory of atomic structure or of the
mechanism of radiation, but which is, however, suggested by
the geometrical diagram of Fig. 92. We shall iJlustrate the
construction of the energy-level diagram by considering in turn
the stationary states of hydrogen and of sodium.
Table VIII shows the first four lines in each of the four series
in the spectrum of hydrogen, column one giving the wave length
and column two the wave number. In column three are given,
in wave numbers, the term values, the differences between which
give the wave numbers of the lines. The first term value for
each series is the convergence wave number for that series. The
last column gives the quantum number of the term.
In these series of the hydrogen spectrum there are of course
many repetitions of term values, the table containing only
eight different terms. The wave numbers given for these terms
are proportional to the energy values of the several states. We
can, if we wish, compute the energy in, say, ergs by multiplying
each term value m wave numbers by ch, i.e., (see equation
(23) Chap. IX).
W = chn
where c is the velocity of hght, h is Planck's constant and n is
the wave-number value of the term; or we may represent the
SEC. 13] ENERGY-LEVEL DIAGRAMS 397
TABLE VIIL-TERM VALUES OF THE STATIONARY STATES OF 'THE HYDROGEN
ATOM
A, n, Terms, Quantum
0
Angstroms centimeters- 1 cen ti meters- 1 number, T
Lyman series
109,678 1
1,216.0 82,258 27,420 2
1,025.8 97,481 12, 186 3
972.5 102,823 6,855 4
949.5 105,291 4,387 5
Balmer series
27,420 2
6,562.8 15,233 12, 186 3
4,861.3 20,565 6,855 4
4,340.5 23,032 4,387 5
4, 101. 7 24,373 3,047 6
Paschen series
12, 186 3
18,756 5,331 6,855 4
12,821 7,799 4,387 5
10,939 9, 139 3,047 6
10,052 9,948 2,238 7
Brackett series
6,855 4
4.05µ 2,468 4,387 5
2.63 3,808 3,047 6
2.16 4,617 2,238 7
1.94 5,141 1, 714 8
energy of a particular state by its term value in wave numbers.
It is convenient to treat the energy values of the various states
as negative quantities, so that the state which has the highest
numerical term value (the state r = 1 in hydrogen) has the lowest
energy. In Bohr's theory this is equivalent to assuming that
the energy of the atom jg zero when the electron is at infinity.
398 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
We now arrange these eight terms in a table in order of descend-
ing energy values. This arrangement is shown in Table IX, in
which, for comparison, is also shown the energy in ergs.
TABLE IX.-ENERGY vALUES OF THE STATES OF THE HYDROGEN ATOM
ARRANGED IN DESCENDING ORDER OF MAGNITUDE
Energy in
State Logarithm n
Wave-number
Ergs
units, n
r=8 1, 714 3 .37 x 10- 13 3.234
7 2,238 4.39 3.350
6 3,047 5.99· 3.484
5 4,387 8.61 3.642
4 6,855 13.45 3.836
3 12, 186 24.0 4.086
2 27,420 53.9 4.438
1 -109,678 -215.5 5.040
The relative values of these states, or of these levels of energy,
are shown by the horizontal lines in Fig. 102. (To secure a
more open scale at the top, logarithms of wave numbers are
plotted.) The transitions which give rise to the several lines
are represented in an obvious manner by the arrows. The
diagram brings out clearly the fact that each series "ends" on a
particular energy level.
Actually, as we have seen in the discussion of the fine structure
of the hydrogen lines, each of these levels has a '' fine structure.''
Thus the level r = 2, is double (one energy value for the ellipse
and one for the circle); the level r = 3 is triple, etc. These
different levels for any given value of r are too close together to
be shown to scale on the diagram. The multiple character of
the three levels r = 2, 3, 4 is shown diagrammatically by the
light lines. And we recall (see p. 382) that the line Ha should
be a triplet, according to the selection principle.
(b) Sodium.-By an exactly analogous procedure we construct
the energy level diagram for sodium. We first form a table
showing the term values, the differences between which give the
wave numbers of the lines for the various series. These data 1
are shown in Table X. The arrangement is identical with that of
1 Taken from Fowler's Report.
SEC. 13] ENERGY-LEVEL DIAGRAMS 399
Table VIII for hydrogen, except as to the designation of the
terms, which follows that used in the spectral series formulre
discussed in Chap. IX. These sodium series are in reality
doublets (hence the Greek letters). To avoid confusion we omit
this feature of the series, and consider them as singlet terms.
-oo
T=B
,=r ,_
3.4 .... -~
~
T=6 Hrf ti) I~
.\lJ
~
~
'i,.,,
3.& I-
T=S Hr 'i,.,,
().,
"
~ I"
0
~
u ~
t:> L...
tl (()
cl)
3.8 '"""i-=4, .M Cl...
E
I-
~ ~
........
(.,.
't 4.0 .... Cl)
c.. L:>
<D 7=3 Ha ,, ,
..0
E
:J
~
!> 4.2 - ~
~
.....
~
..........
ts
3...
~ ~
l.,
(I) Cl'.:}
0
O') 4.4- .... 'T=2 I,
0
..J t;:
~
t
4.6 ~
~
'-I
4.8 -
5.0 ~i=1 ' I, It
Frn. 102.-Energy-level diagram for the hydrogen atom.
We now sort out these terms and arrange them in a descending
series of energy values, just as we did in the case of hydrogen.
We note that there are four kinds of terms: u, 1r, o, and </> terms.
We arrange like terms in vertical columns. The order of the
columns is determined by the highest wave-number value (i.e.,
lowest energy value) in the column, the u column thus being first
and the </> column last. We then have Table XI.
400 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
TABLE X.-TERM VALUES OF THE STATIONARY STATES OF SomuM
0
x, n, Terms,
Term symbol
Angstroms centimeters-1 cen timeters- 1
Principal series: lu - m1r
41,449 lu
5,890.0 16,973 24,476 l1r
3,302.3 30,273 11, 176 21r
2,852.8 35,043 6,406 31r
2,680.3 37,298 4, 153 41r
Diffuse series : h - mo
24,476 l1r
8, 194. 8 12, 199 12,276 20
5,688.2 17,575 6,900 3o
4,982.9 20,063 4,412 4o
4,668.6 21,414 3,062 5o
Sharp series : l1r - mu
24,476 l1r
11,404 8,766 15, 710 2u
6, 161 16,227 8,248 3u
5, i54 19,398 5,077 4u
4,752 21,038 3,437 5u
Fundamental: 2o - 3c/>
12,276 20
18,459 5,416 6,860 3cp
12,678 7,886 4,390 4cp
3,041 5cp
It will be observed that the wave numbers of the lines of a
given spectral series are obtained by taking the differences
between terms in adjacent columns: for example, lu - 27r gives
the second line of the principal series; or l 1r - 48 gives the third
line of the diffuse series. This limitation of combinations to
adjacent columns is suggestive, though perhaps vaguely, of the
selection principle enunciated in Sec. 9, according to which the
azimuthal quantum number of a given state can change by only
+ 1 in any transition between states which involves the emission
of radiation. If we assign to the columns from left to right,
beginning with the <1 column as 1, a number k as shown (note
SEC. 13] ENERGY-LEVEL DIAGRAMS 401
that the columns are arranged from left to right in ascending
algebraic order of the lowest term in the column), the observed
limitation of inter-state transfers to those in adjacent columns
takes the form: jj_k = + 1. The numeral k js identified with the
azimuthal quantum number which we have previously called
TABLE XL-ENERGY VALUES OF THE STATES OF THE SODIUM
ATOM ARRANGED IN DESCENDING ORDER OF MAGNITUDE
Wave number, n, centimeters-1
Loga-
Term Energy
rithm
symbol u 7r (ergs)
n
terms terms terms terms
51r ........ .. .. 2,907 ... . . . .. . . 5.71 x 10-13 3.463
5</> ... . . . . . ...... . ..... ...... 3,041 5.98 3.483
5o .......... ...... ...... 3,062 ..... 6.02 3.486
5u ......... 3,437 ...... ..... . . . . .. 6.75 3.536
41r ......... ...... 4, 153 ...... . .... 8.16 3.618
4</> ........ ...... ...... ...... 4,390 8.63 3.643
4o .......... . . . . . . ...... 4,412 ..... 8.67 3.645
4u ....... . . . 5,077 . ..... . ..... . .... 9.98 3.706
31r ......... ...... 6,406 . ..... ..... 12.60 3.807
3</>. .. . . . . . . . . . . . . ..... . ..... 6,860 13.48 3.836
3o ....... . . ...... . ..... 6,900 . .... 13.55 3.839
3u ....... . . 8,248 ...... ...... ..... 16.20 3.916
21r ......... ...... 11, 176 ...... ..... 21.9 4.048
20 ....... . . . ... . . ...... 12,276 . .... 24.1 4.089
2u ....... . . 15, 710 ...... . . . . .. . .... 30.9 4.196
l1r ........ ...... 24,476 ...... . .... 48.2 4.389
lu ....... . . 41,449 ..... ...... .. . . . 81.5 4.617
. ·········I k=l I k=2 I k=3 lk=41
r<1>, and the statement D,.k = + 1 is identical with the selection
principle previously stated.
The horizontal lines in Fig. 103 show in relative positions the
energy levels for the sodium atom plotted to a logarithmic scale.
The transitions giving rise to the several lines are shown by the
arrows.
(c) Energy level diagrams of this kind are of great assistance
in understanding and interpreting the interrelations among
spectral lines. The reader will find it very instructive to prepare
such a diagram starting with the spectral series data of, say,
402 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
potassium and including in the diagram the doublet levels and
the resulting lines.
,.
iM'
3.6 ....
[""-..
-~
3.8 ...... ~
v:)
-. , __
~ I"-
<.:, ~
(lJ (l)
'-
Q) c::: ~ ·-
_n Cl) ts
E t-:> ~
~
:J ")
z 4.0 ,_ (U
..... t ~
Q) l. ~
> ~ -.,t::
0 ~ V) \ ~ '
3 2
m ~
0
...J ll)
4.2 -
,:::,
~.....
z
ll)
(l> ()
.....
t...
~
V)
I , . II ii'' II'. II' ,i,
4.4 - 1
~
-~
'-l
. .....
.c::
ct
4.6,_ 'ii I,, ·~ I,,
1
}-,IG. 103.-Energy-level diagram for the sodium atom, showing the origin of
the several series of lines.
14. Notation; Inner Quantum Numbers.-A word about the
notation for designating spectral terms: vVe have deliberately
retained the simple notation of Fowler, as introduced in Chap.
IX. This notation is easy to understand and is, therefore,
preferable for a brief survey. The student who goes more
deeply into the subject can familiarize himself with the particular
notation which happens to be prevalent at the time! The
essential facts are not changed by the notation in use. Notation
depends on theory and the interpretation of the facts.
SEC. 14] INNER QUANTUM NUMBERS 403
The need for a more comprehensive notation than Fowler's,
as given in Chap. IX, can be illustrated by reference to Table XI.
The term 31r (6,406) is a doublet. In the Fowler notation, the
two terms of this doublet are indicated by subscripts, thus:
31r 1 and 31r 2 • In general, many such "multiplicities" are found
in spectra. In the spectrum of magnesium, as previously men-
tioned, are found two complete groups of terms-singlets and
triplets. In Fowler's notation, we have represented the singlet
terms by capital letters (S, P, D, and F) and the triplet terms by
lower-case letters (s, p, d, and f) with subscripts 1, 2, and 3.
Actually, however, the s term of the triplet series is a singlet
term, while p, d, and f terms are triplets. The triplet nature of
the lines of the spectrum results from the fact that the lines
arise from the combination of the terms. Thus, the principal
series p(m) = ls - mp arises from combinations of the ls term
with each in turn, of the three p terms. The sharp series
s(m) = lp-ms arises from combinations of each of the three p
terms with the singlet s term. Thus, the components s 1 (1), s2 (1),
and 83 ( 1) of the first ''line'' in the sharp (triplet) series of
calcium are given by the following combinations:
s1(l) = lp1 - ls = 33,989 - 17, 765 = 16,226
s2(l) = lp2 - ls = 34,095 - 17,765 = 16,330
S3(l) = lps - ls = 34,147 - 17,765 = 16,382
Since, therefore, the s terms are single and the p terms are
triple, we should expect the sharp series of lines and the principal
series of lines to consist of triplets, as is actually observed to
be the case.
The diffuse series, however, arises from combinations of the
p terms with the d terms, thus: d(m) = lp - md. Since, in
calcium, both the p terms and the d terms are triple, we might
expect the lines of the diffuse series to consist of nine components,
instead of three. Actually, including the satellites mentioned on
page 311, the diffuse series of lines in calcium contains six com-
ponents, which we tabulate in Table XII. In this table, the
wave numbers of the p terms are written above the vertical
columns, the designation of the terms in the Fowler notation
being written above the wave number. Similarly, the wave
numbers of the d terms are written at the left of the horizontal
rows. The wave numbers (observed) of the six components,
together with the combinations of terms which give rise to the
404 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
TABLE XII.-COMPONENTS OF THE FIRST "LINE" IN THE DIFFUSE (TRIPLET)
SERIES OF CALCIUM
(
I
lp1 lp2 lp3 ~old notation
Terms~ 33, 988. 7 cm.-1 34,094. 6 cm.- 1 34, 146 . 9 cm. -1
l l I3P2 l3P1 13Po ~new notation
ld3 lp1 - Id3 lp2 - ld3 lp3 - ld3
28, 968. 8 cm.-1 5,019.6 5,125.2 5, 177 .3
l3D1 l3P2 - l3D1 l3P1 - l3D1 l3Po - l3D1
Id2 lp1 - Id2 lp2 - ld2 lp3 - Id2
28, 955. 2 cm.- 1 5,032.9 5,139.5 missing
l3D2 l3P2 - l3D2 l3P1 - l3D2 l3Po - l3D2
ld1 lp1 - ld1 Ip2 - ld1 lp3 - ld1
28,933. 5 cm.- 1 5,055.1 missing missing
13D3 l3P2 - l3D3 l3P1 - l3D3 l3Po - l3D3
components, are given in the squares. It is observed that the
three components are missing corresponding to the combinations
lp2 - ld1; lp3 - ld2; and lp3 - ld1. There is evidently in
operation here a selection principle somewhat similar to that
discussed in connection with the fine structure of the hydrogen
lines. We shall return to this point presently.
Multiplicities much higher than triplet terms occur in spectra.
Indeed, multiplicities as high as octets have been found. These
multiplicities exhibit a beautiful regularity not only among the
several groups of terms of the spectrum of a given element but
also, as we shall mention in Chap. XI, from element to element
in the periodic table of the elements. In a singlet system of
terms-as in the singlet system of calcium-all the terms are
singlets. In a doublet system, the S terms 1 are singlets and the
P, D, F . . . terms are doublets. In a triplet system, the 8
terms are singlets and the remainder are triplets. In a septet
system, the S terms are singlets, the P terms are triplets, the D
terms are quintets, and the F terms are septets. The multi-
plicity of a system of terms is designated by the highest multi-
plicity found in the system. The S terms are always singlets; the
P terms are never more than triplets; and so on.
Table XIII shows the multiplicities of the terms (S,P, D . . . )
of the several systems (singlets, doublets, triplets . . . ).
1We here use the capital letters to designate terms, irrespective of
multiplicity.
SEC. 14] INNER QUANTUM NUMBERS 405
The P terms in a sextet system are triplets. The F terms in
an octet system are septets; and so on
With this complexity of systems and of terms, it is obvious
that designation of terms by capital letters, Greek letters and
lower-case letters for singlet, doublet, and triplet systems, is
TABLE XIIL-SHOWING THE MULTIPLICITY OF TERMS IN SPECTRAL
SYSTEMS
Doub- Quar- Quin-
System~ Singlet Triplet Sextet Septet Octet
let tet tet
Terms lk
s 1 1 1 1 1 1 1 1 1
p 2 1 2 3 3 3 3 3 3
D 3 1 2 3 4 5 5 5 5
F 4 1 2 3 4 5 6 7 7
G 5 1 2 3 4 5 6 7 8
no longer adequate. Many alternative systems for the desig-
nation of terms have been proposed. The one most generally
adopted is that proposed by Russell and Saunders 1 and referred
to as the "new notation" in Table XII. In this notation, the
capital letters S, P, D, F, G . . . are used for the sequences
of terms. Since the azimuthal quantum numbers of these
sequences are, respectively, 1, 2, 3, . . . , we may think of S,
P, D . . . as not only designating a sequence but also as standing
for 1, 2, 3 . . . , the azimuthal quantum number of that
sequence. Instead of using different letters for the different
multiplicities, as 1r for doublets, p for triplets, etc., Russell and
Saunders prefix as a superscript to the letter a number, as 2P, to
indicate the multiplicity of the system (not the term) to which the
term belongs. Thus, 2P refers to the P terms of a doublet
system, 5P refers to the P terms of a quintet system, etc.
To designate the several components of a sequence, suitably
chosen subscrjpts are used. Thus, in a quintet system, the P
terms have a threefold multiplicity- we require three subscripts.
The D terms have a fivefold multiplicity-we require five sub-
scripts. These subscripts are the so-called inner quantum
numbers, symbol J (see below). Their selection, based on con-
siderations partly empirical, partly theoretical, is by no means
an arbitrary matter. Empirically, we may illustrate the choice
1 Astrophys. Jour., vol. 61, p. 64 (January, 1925).
406 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
of these inner quantum numbers (and, incidentally, the Russell-
Saunders notation) by reference to Table XII. The three P
terms, the first in the P sequences in the triplet system of calcium,
are designated, respectively, by l 3P2, l 3P1, 13Po, the subscripts
2, 1, 0 being the inner quantum numbers of these P terms. Like-
wise, the three D terms, which combine with these P terms to
give the first "line" of the diffuse series, are represented by l3D 1,
l3D2, l3D3, the inner quantum numbers here being 1, 2, 3. With
this selection of inner quantum numbers (" mnemonic sub-
scripts," one may prefer to call them) it is noted (see Table XII)
that only those combinations of terms take place for which the
change in the inner quantum number is + 1 or 0. All other com-
binations are "ruled out;" e.g., there is no component correspond-
ing to l3P 1 - l3D3, since, for this combination, the change in j
is +2.
Observations on more complex systems, quartets, sextets,
etc., have led to a selection rule for this inner quantum
number, which rule, when the inner quantum number is properly
chosen, states that the permissible change 11j, in j, is given by
11j = + 1 or 0, (85)
the change 11j = 0 for zero values of j being also ruled out. Table
XIV gives the inner quantum numbers for the S, P, D . . .
sequences of singlet, triplet, quintet, and septet systems, these
numbers being so chosen that the selection rule (equation (85))
predicts correctly the component lines actually observed in the
spectra, as shown in Table XII. Conversely, from the inner
quantum numbers, as assigned in this table, together with the
two selection rules (equations (84) and (85)) it is possible to
11k = + 1
flj = + 1 or O
predict the ''permitted'' combinations between terms.
TABLE XIV.-INNER QUANTUM NUMBERS, j
Singlet Triplet Quintet Septet
Sequence
system system system system
S (k = 1) .......... 0 1 2 3
P (k = 2) .......... 1 2 1 0 3 2 1 4 3 2
D (k = 3) .......... 2 3 2 1 4 3 2 1 0 5 4 3 2 1
F (k = 4) .......... 3 4 3 2 5 4 3 2 I 6 5 4,3 2 1 0
G (k = 5) .......... 4 5 4 3 6 5 4 3 2 7 6 5 4 3 2 1
SEC. 15] MOLECULAR SPECTRA 407
On theoretical grounds, the inner quantum number is associ-
ated with the angular momentum of the atom as a whole; i.e.,
the inner quantum number is determined by the vector sum of
the angular momenta of the several electrons of the atom in their
orbital motion about the nucleus. We shall say a word further
about this in the next chapter. Mention may also be made of the
fact that both theory and observation demand the use of half-
quantum numbers, such as j = ~~' ~% . . . It will not be
profitable for us to enter into a further discussion of this point.
15. Molecular Spectra.-In the preceding sections of this
chapter, we have considered only those spectra which are emitted
by the atom as a result of energy changes in the electron system
surrounding the atom. We may think of this type of spectra as
having its origin in a periodic phenomenon associated with the
atom, namely, the revolution of the electrons around the atomic
nuclei, although, according to the quantum theory, the emitted
frequencies are not the same as the frequencies of these orbital
motions but are determined by the quantum condition
hv = W1 - W2
The orbital motions of the electrons around nuclei constitute
only one of several types of periodic phenomena which, on the
basis of our present picture of molecular structure, should be
exhibited by matter. We saw, in Sec. 10 of Chap. VIII, that the
variation of the molecular heat of gases with temperature could
be explained by assuming that, in addition to translation, the
molecules of a polyatomic gas have degrees of freedom as regards
both rotation and vibration. It may be expected that these
types of periodic motions would also result in spectral phenomena
and that to the energy changes associated with these motions we
might apply the quantum laws introduced by Bohr. Such proves
to be the case. In addition to the Balmer type of spectra dis-
cussed in Chap. IX there are found both emission and absorption
lines comprising, in general, the type of spectra known as band
spectra (see Chap. IX, Sec. 10). These can be definitely ascribed
to changes in molecular energy, which may be thought of as made
up of three parts: (1) rotational energy of the molecule, say a
diatomic molecufo, about an axis through its center of mass and
at right angles to the line joining the two nuclei; (2) vibrational
energy due to the relative vibratory motion of the two atoms with
respect to each other; and (3) energy of the electronic system
408 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
associated with the molecule. Corresponding to each of these
three types of energy there is a quantum number, mr, mv, and me,
respectively, changes in which play the same role in band spectra
as does the orbital quantum number in the Balmer type of
spectra. For a full account of band spectra the reader is referred
to "Molecular Spectra in Gases," 1 by the Committee on Radi-
ation in Gases of the National Research Council. In this sec-
tion, we shall discuss only two simple cases illustraitive of the
fundamental concepts involved.
Consider the rotation of a dumb-bell-shaped (i.e., diatomic)
gaseous molecule about an axis passing through its center of
mass and perpendicular to the line joining the two nuclei. Let
the moment of inertia of the molecule about this axis be I,
assumed constant and independent of the angular velocity w.
The angular momentum Mis given by
M = lw (86)
and the rotational kinetic energy Tis
T = 3,1Iw 2 (87)
Let it be assumed that we can apply to this angular momentum
Bohr's quantum condition (equation (18))
fpdq = rh
Inserting in this general equation the angular momentum M
and the corresponding angular coordinate e, we have, for the
rotating molecule,
f MdO = 21rM = mrh (88)
where mr, which replaces r, is the molecular rotational quantum
number. Eliminating M and w from equations (86), (87), and
(88) gives
(89)
This is the kinetic energy of the rotating molecule for a given
quantum "state of rotation'' designated by the integer mr.
Applying Bohr's quantum condition (equation (31)), when the
1 Nat. Research Council, Bull. 57 (December, 1926).
SEC. 15] MOLECULAR SPECTRA 409
energy changes from T1 corresponding to mr,1 to T2 correspond-
ing to mr, 2, a quantum of energy hv is emitted, given by
h2
hv = T1 - T2 = 1r (m;, 1 - m;, 2) 90)
8 21
from which we obtain, for the frequency v,
h
v = 81r2J(m;,1 - m;,2) (91)
Now, since the momentum M involves only angular motion,
we may regard the quantum number mr as similar to the azi-
muthal quantum number k for electron orbits, which, according
to equation (84), may change by only + 1. We may, therefore,
apply to mr the selection rule for azimuthal quantum numbers,
~mr = +1
from which we see that i.e. rnr,1 can differ from mr,2 in equation
(91) by only 1 unit. For the emission of radiation, therefore,
and, dropping the subscripts, equation (91) becomes
h
v = 81r2J (2mr + 1) (92)
This equation predicts a series of lines which, for the several
integer values of mr, should occur at constant frequency intervals
~v given by
h
ilv = 41r2J (93)
Such a spectrum is called a pure rotation spectrum.
From a knowledge of the approximate value of the moment of
inertia I of molecules (of the order of 10- 40 gram. cm. 2 ), these
"rotation" lines should be in the far infra-red. Czerny 1 finds
in the absorption spectrum of HCl a group of lines in the neigh-
borhood of 50 to 100µ, the frequencies of which are given by the
relation
v = (20.8 X c)m (94)
where c is the velocity of light and m stands for successive
integers. Comparing this empirical result with equation (93),
1 Zeit. Physik, vol. 34, p. 227 (1925).
410 THE NUCLEAR ATOM AND SPECTRA [CHAP. x
we see that the difference in frequency ~v between successive lines
should be
h
dv = 1r 2I = 20.8 X c
4
From this, one computes that the moment of inertia I of the
HCI molecule is 2.7 X 10- 40 gram. cm. 2-which corresponds
to a distance apart of the two atoms of HCl of, approximately,
1.2 X 10-s cm., a result which is certainly of the correct order
of magnitude.
In general, a diatomic molecule may rotate and vibrate at the
same time. Its total energy is then partly rotational, partly
vibrational, with the latter of which is associated a quantum
number mv analogous to mr for rotation. Any given change of
state may involve simultaneous changes in both the quantum
numbers mr and mv-in somewhat the same way as the ellip-
tical electronic orbits involve changes in both azimuthal and
radial quantum numbers. That is to say, the energy hv of a
given emitted quantum may come partly from the rotational
energy, partly from the vibrational energy of the molecule. In
that event, it can be shown 1 that the frequency should be given
by an equation of the form
h
V = Vv + S1r 2J ( +2mr + 1) (95)
where vv is the contribution to the frequency v due to the vibration
and the second term on the right is the contribution due to
rotation. It turns out that vv is, in general, much the larger of
the two terms. For a given vv, the second term has a series of
values for the several integer values of mr. (That is, associated
with a given quantum jump in vibratory energy there may be
any one of a large number of jumps in rotational energy.) Equa-
tion (95), therefore, predicts a series of lines differing from each
other by frequency differences dv given by
h
Liv = 41r2J
the same as for jumps involving rotational energy only, but
coming at higher frequencies. This type of spectrum is known as
a rotation-vibration spectrum. Spectra of this type occur in the
near infra-red. For example, in the absorption spectrum of
HCl occurs a series of equally spaced bands, or lines, in the
1 See Nat. Res. Council, Bull. 57, loc. cit.
SEC. 15] MOLECULAR SPECTRA 411
neighborhood of 3.6µ, the frequency separation of which agrees
with that computed from the pure rotation spectrum of HCI.
A still further complication results when changes of energy
of the molecular electron system is superimposed on the rotation-
vibration energy changes. This produces the so-called electronic
band spectra-of which the cyanogen bands referred to in Chap.
IX; Sec. 10, are examples.
The solution, even though still incomplete, of the very complex
problem of molecular spectra is one of the most astonishing
successes of the quantum theory.
CHAPTER XI
THE ARRANGEMENT OF ELECTRONS IN ATOMS
The Rutherford-Bohr atom model was built out of raw material
which grew in the field of physics. The atom itself, originally a
child of Chemistry, was soon adopted by physics and has long
claimed this joint parentage. In all of our theories and theoriz-
ing, we must not fail to distinguish between the real atom and the
atom model. The ultimate atom model-if such perfection be
ever reached-must explain and agree with the facts of chemistry,
which facts are just as cogent as are the facts of physics. The
task of constructing an atom model which shall do all the things
that atoms are made to do by both chemists and physicists was
(and is!) gigantic. A model, built and trained to act as mario-
nette for a physicist, could not be expected to perform well for
the chemist.
Out of the material furnished by chemistry have been built
various atom models, which, in the main, go by the generic name
of static atom models, as contrasted with a model of the Ruther-
ford-Bohr type, which is a dynamic model. The origin of the
two terms is significant. The dynamic model has been con-
structed by the phyaicist largely from information acquired in
watching atoms in action or, at least, in watching the results of
their activity, such as the emission of radiation, characteristic
or black body, or the scattering of a particles, and the like. It is
natural, therefore, that the physicist should picture an atom as
full of moving mechanisms. His one static atom, the Thomson
model, proved inadequate. The chemist, on the other hand, is
somewhat more concerned with the atom at rest. He weighs it
in finding what combinations it makes with other atoms; he
observes it apparently at rest in crystals, except for possible
thermal agitation; he thinks of it as occupying a definite position
as one of the constituent atoms of various complex molecules,
being held in place by certain interatomic forces, perhaps of elec-
trostatic origin. It was difficult to see how a thing so full of
whirling mechanisms as the physicist pictured the atom to be
412
SEC. 1] THE INERT GASES 413
could keep as quiet as the chemist found it when he looked at it.
Hence, the static atom.
PART I. THE STATIC ATOM
1. The Inert Gases: Atomic Numbers and Properties.-The
suggestion that the electrons of the nuclear atom occupy fixed
positions in concentric shells was first made by G. N. Lewis 1 and
was, in a sense, a revival of the Thomson atom, modified, as a
result of Rutherford's experiments on the scattering of a par-
ticles, by putting the p_ositive charge, the nucleus, inside the
concentric shells and at their center. This suggestion was
greatly extended by Langmuir,2 who discussed in detail the
relation of this model to many chemical and physical properties
of atoms and molecules. Mention was made, in the preceding
chapter, of the origin of the concept of atomic number, a concept
which was firmly established by Moseley (see Chap. XII). As a
result of Moseley's work, the total number of elements up to and
including uranium was fixed at 92, and the atomic number of
each element was unambiguously assigned. The atomic numbers
of the elements are shown in Appendices I(a) and I(b ), the latter
giving the elements as arranged in the well-known periodic table.
Strictly speaking, the atomic number of an atom is the ratio of
the (positive) charge on its nucleus to the (numerical value of
the) charge on the electron. This ratio is numerically equal
to the number of electrons which surround the nucleus of the atom
in its normal condition.
N ote.-When, as a result of the action of some physical agency, an atom
loses 1 or more electrons, the a tom is said to be ionized-singly ionized if it
has lost 1 electron, doubly ionized if it has lost 2 electrons, etc. The loss
of 1 electron leaves the atom with an excess of positive charge equal numer-
ically to the charge on the electron. We may designate such an atom by
writing "+"as a superscript (or subscript) following the chemical symbol
of the atom. Thus, singly ionized magnesium is written Mg+; doubly
ionized, Mg++. Or, Roman numerals may be used as subscripts; thus:
Mg1, Mgn, . . . Mgv1, etc.
All theories of the spatial distribution of the extranuclear 3
electrons in atoms start from the experimentally observed fact
1 Jour. Amer. Chem. Soc., vol. 38, p. 762 (1916).
2
Jour. Amer. Chem. Soc., vol. 41, p. 868 (1919); Jour. Ind. Chem. Eng.
(April, 1920); Nature, vol. 105, p. 261 (April, 1920).
3
"Extranuclear" since, as we shall see later, the nucleus contains
electrons.
414 ARRANGEMENT OF ELECTRONS [CHAP. XI
that as we proceed along the sequence of the elements from hydro-
gen to uranium we come periodically upon monatomic elements
which form no chemical compounds with atoms of other elements.
These elements are the so-called inert gases, helium, neon, argon,
krypton, xenon, and niton. In addition to their chemical inac-
tivity, these elements, as is shown by Table I, have very low
melting and boiling points. If we make the very reasonable
TABLE !.-MELTING POINTS AND BOILING POINTS OF THE INERT GASES
Melting point, Boiling point,
Atomic number Element degrees degrees
Centigrade Centigrade
2 He < -272.2 -268. 9
IO Ne -248. 7 -245.9
18 Ar -189.2 -187.0
36 Kr -169 -151. 8
54 Xe -140 -109.1
86 Nt - 71
1 hydrogen -259 .1 -252. 7
assumption that the chemical activity of an atom is conditioned
somehow upon the magnitude of its external electric (or perhaps
magnetic) field, which field would also determine the melting point
and the boiling point, we conclude that the atoms of the inert
gases have very weak external fields. Or, in different words, the
arrangements of the extranuclear electrons in the atoms of the inert
gases, whatever that arrangement is, is such as to give to those atoms a
very stable structure. Now, when atoms combine with atoms of
the same kind, or of other kinds, to form molecules, the electron
systems of the several atoms unite to form some kind of spatial
distribution of electrons characteristic of the molecule. All
atom-model builders, whether chemists or physicists, make as
a fundamental postulate-it amounts almost to an axiom-that
chemical combinations between atoms in forming molecules tend to
take place in such a way as to imitate, by the combination, the
arrangement of electrons in the atoms of the inert gases-again,
whatever the latter arrangement may be. The steps, then, in
building up theories of the spatial distribution of electrons in
atoms are (1) to make some sort of reasonable guess as to the
distribution in the inert gases and then (2), guided, in part, by
data on chemical combinations which atoms are observed to
SEC. 1] THE INERT GASES 415
make with each other and, in part, by spectroscopic data from
the visible, the near-visible, and the X-ray regions of the spec-
trum, so to arrange the electrons in the other atoms as to be
able to predict, from the proposed arrangement, the observed
physical and chemical facts.
This "reasonable guess" concerning the arrangement of elec-
trons in the atoms of the inert gases originally grew out of an
observation by Rydberg,1 made soon after Moseley's work had
assigned atomic numbers to the elements, that the atomic num-
bers Z of the inert gases can be expressed by a simple numerical
series, as follows:
z = 2 (l2+2 2 +2 2 +3 2 +3 2 + 4 2 ••• ) (1)
~He-j'
+-Ne-t
---Ar--t
~---Kr---t
~----Xe -t
~ Nt ~
For example, the atomic number of argon is 18. And
18 = 2(1 2 +2 +2
2 2
)
This numerical series suggests that the electrons in these inert
gases may be arranged in "layers" or shells: that the argon
atom, for instance, contains three shells, the first shell containing
2 · 12 electrons; the second, 2 · 2 2 ; and the third 2 · 22; and so
on, for the other inert gases. The factor 2 suggests symmetry
of some kind.
Of course, this suggestion that the electrons may be arranged
in layers contains no hint as to the state of the electrons in those
layers. The electrons may be in orbital motion around the
nucleus, or they may be either at rest or in vibration about some
fixed position of equilibrium determined by the force field of the
nucleus. If we assume the latter, as is apparently demanded by
the static atom of the chemist, we must invoke the aid of some
unknown force or law of force to produce equilibrium, since, as
we saw in discussing the problems of equilibrium raised by
Rutherford's nuclear atom, electrons outside the nucleus cannot
be in equilibrium under the action of the inverse-square law of
electrostatic force, alone. ~ If, however, with Bohr, we assume
1 Phil. Mag., vol. 28, p. 144 (1914).
416 ARRANGEMENT OF ELECTRONS [CHAP. XI
f
the former, we must invoke the ajd of some unknown law or
principle which makes it possible for an electron to execute
orbital motion and, in so doing, to evade the inexorable laws of
the electromagnetic theory. Formally, there is nothing to
choose between the two horns of the dilemma. It will be
instructive, therefore, for the time being, to follow the Lewis-
Langmuir concept of stationary electrons, in discussing the
evolution of the atom model, 1 especially since, in the initial
formulating of ideas, it is easier to visualize the static than the
dynamic atom.
Helium has 2 electrons. Since its inertness indicates that the
structure of the helium atom is stable, the general law is sug-
gested that, in combination with a nucleus (or, perhaps, nuclei),
2 electrons tend to make a stable, a very stable, combination.
Why, we cannot say, until we know much more about the force
fields surrounding nuclei and electrons. We take the stability
of such a pair of electrons as an empfrical fact.
Neon has 10 electrons: 10 = 2(l2 + 2 2 ). We may guess that,
since 2 electrons make a very stable combination, (a) 2 of the 10
electrons in neon are, somehow(!), combined in a stable pair
around the neon nucleus (charge + 10) and (b) the remaining
8 electrons (2 X 2 2 ) form an outside layer. Again, since the
structure of neon is stable, it follows that, whatever the reason
1 There has been much criticism of the "static" atom model. Thus, one
eminent writer states: "It is true that if the eleven postulates, on which
Langmuir's theory is founded, are granted, and a certain freedom of inter-
pretation be allowed, a great body of chemical observation is covered by
this theory, but it must be observed that the postulates have very much an
ad hoc character, often paying little attention to any previously established
laws of electromagnetism." As if Bohr's postulates or Planck's quantum
hypothesis paid any attention to the "previously established laws of electro-
magnetism!" It is the essence of the whole quantum theory that "ad hoc"
assumptions are made, ad libitum, where such assumptions serve a useful
purpose. Of course, the fewer such assumptions the better. But, logically,
we may as well make a dozen assumptions contravening classical theory,
as one. If atom models growing out of the concept of the orbital motions
of electrons, i.e., the so-called "dynamic" a tom models, have been more
successful than static atom models, it is because the former have been more
prolific in explaining and interpreting physical phenomena, rather than
because of any superiority in fundamental tenets. The great number of
qualitative deductions from static atom models, deductions in agreement
with chemically observed facts, is very suggestive. But the computation of
spectral frequencies, even for 1 atom, from fundamental constants of nature
is a quantitative achievement that is all but convincing.
SEC. 1] THE INERT GASES 417
may be, a group of 8 electrons, inside which are (one or more)
stable groups, tends to make a stable combination. Since we
know of no reason why these 8 electrons should not be disposed
symmetrically about the nucleus, it is natural to assume that
they are located at the corners of a cube-although we must
admit that this argument is much like that by which Pythagoras
proved the world to be round: since the sphere is the most
perfect figure! On account of this suggestion of the "cubical"
arrangement of these 8 electrons, this picture of the atom is
frequently referred to as the cubical atom. The group of 8
electrons is called an octet.
Argon has 18 electrons, 2(1 2 22+ +2 2 ). It is reasonable to
picture the first 10 of these as occupying positions similar to
those in the neon atom and the remaining 8 as forming another
layer outside this neon-like structure. We shall designate the
distribution of electrons in groups surrounding the nucleus by
an obvious abbreviation which gives the number of electrons in
successive layers: For helium, the distribution is (2); for neon,
(2, 8); for argon, (2, 8, 8). Langmuir assumed that the electrons
of this second group of 8 in argon form eight stable pairs with the
first group of 8, and that these eight pairs occupy eight cells in
a layer around the central pair and the nucleus. Bury,1 on the
contrary, subsequently showed that this latter assumption is
unnecessary and that the properties of the elements could be
more readily explained by the assumption that the outside stable
group is always a group of 8 electrons. Before taking up the
distribution of electrons in the atoms of the inert gases of higher
atomic number than argon, we shall pause to consider the appli-
cation, to chemical combinations of the lighter elements, of the
following two general principles:
1. Electrons around, or in combination with, nuclei tend to
form either groups of 2 (pafrs) or groups of 8 (octets).
(We shall have to extend this to groups of 18 and of 36 for
the heavier atoms.)
2. Such combinations of atoms to form molecules tend to
take place as make it possible for the electron systems of the
combining atoms to imitate so far as possible the configura-
tion of the electrons in one or the other of the inert gases.
1 Jour. Amer. Chem. Soc., vol. 43, p. 1602 (1921).
418 ARRANGEMENT OF ELECTRONS [CHAP. XI
2. Some Chemical Properties of the Lighter Elements. (a)
Hydrogen (Z = 1).-This element has 1 electron associated
with a single-charged nucleus. Since we postulate that electrons
tend to form pairs or octets, it may be expected that (atomic)
hydrogen would be very active chemjcally either in combining its
single electron with an electron from another atom to form a
helium-like pair or in joining irs electron with a group of elec-
trons in another atom to assist in the completion of an octet.
A pair could also be formed by the union of 1 hydrogen atom
with another to form molecular hydrogen H2. H2 should be a
helium-like structure (in so far as concerns the electrons) and
should, therefore, resemble helium in its physical and chemical
properties. Hydrogen has a very low boiling point (see Table I)
and in its molecular state is comparatively inactive.
(b) Helium (Z = 2).-0f course, we cannot explain the
properties of helium, since those properties are taken as a stand-
ard of comparison (more correctly, of imitation).
(c) Lithium (Z = 3).-We may expect 2 of the 3 electrons of
lithium to form a stable pair around the +3e nucleus, the com-
bination (nucleus plus 2 electrons) being called the kernel of the
atom. This kernel has a net unit (i.e., +e) positive charge.
The other electron is positioned at a greater distance from the
kernel. The distribution of the electrons in lithium is (2, 1).
On account of this distribution of charge, the lithium atom
should have an electric moment, and, accordingly, lithium
atoms should "line up" like iron filings in a magnetic field, as is
shown diagrammatically in Fig. 104(a), where the dots represent
electrons, each with a negative charge - e; and the circles, the
lithium nuclei with their threefold positive charge +3e, close to
which are the 2 electrons which make the stable pair, the com-
bination ( +3 - 2)e giving to the kernel a net charge +e. This
lattice array of lithium atoms, of course built up in three dimen-
sions, constitutes a crystal, in which are numerous comparatively
free electrons. Lithium then, should be a crystalline solid and
should be a good conductor of electricity. Lithium is a crystalline,
conducting (metallic) solid with a melting point at 186°C.-
much higher than the melting points of hydrogen and helium,
because of the greater electric force holding the constituent parts
of the crystal, the ions, in place.
It is to be observed that in the crystal of lithium as shown in
Fig. 104(a), there is no particular kernel and electron which we
SEC. 2] THE LIGHTER ELEMENTS 419
can picture as belonging together to form a molecule, since each
kernel is surrounded by a number of electrons, and vice versa.
A crystal of this type is, therefore, composed of ions but not of
molecules. Even in the liquid phase, the ionic condition persists.
In the solid and the liquid phase, therefore, there are no lithium
molecules as separate entities. Only in the vapor phase may we
speak of "lithium molecules."
When a lithium atom and a hydrogen atom are brought
together, the 4 electrons and the 2 nuclei combine in such a way
• • •
@
• • @
•
• @
• •
• • •
• @
•
• @)
• • ©
•
• • •
0
•
• ®
•
• @
• •
(Ol) Lithium
I @
/.
,_
+1
-,
\
)
-1
/.,...-'
~· ® ·~
\ • _.,... I
' /
(,b) A Single 11 Molecu\e"of LiH
Frn. 104.-(a) Schematic representation of the crystalline structure of metallic
lithium. The dots represent electrons. The circle is the positively charged
lithium nucleus; (b) the molecule of lithium hydride, showing its ionic com•
position.
as to form two stable pairs of electrons, as is shown djagrammati-
cally, in Fig. 104(b). The one group, A, has a net charge +e,
and the other, B, a :net charge -e. We should, therefore, pre-
dict the compound LiH. For the same reasons as in the case of
metallic lithium, these combinations of pairs should build up into
a three-dimensional crystalline lattice. Since there are now no
free electrons, this crystal should be a poor conductor of elec-
tricity. The crystal LiH is known, and it has a high melting point
~680°C.), which could be predicted from the comparatively
strong forces between the ions. When melted however, LiH
420 ARRANGEMENT OF ELECTRONS [CHAP. XI
should be made up of Li+ and H- ions; it should, therefore, be
an electrolytic conductor-which agrees with the facts.
(d) Fluorine (Z = 9) and Sodium (Z = 11).-(We pass, for a
moment, the intervening elements.) Between fluorine and
sodium comes the stable inert gas neon (Z = 10), which has its
outer shell completed or "satisfied" with 8 electrons. Fluorine
has only 7 electrons in its outer shell; sodium, with a total of
11 electrons, has its outer shell of 8 completed and 1 electron to
spare. The distribution of electrons in fluorine, therefore, is
(2, 7); and in sodium, (2, 8, 1). A sodium and a fluorine atom
by combining could make two octets, the extra electron from
sodium filling the vacant place in the outer shell of fluorine.
We should, therefore, predict the compound NaFl which should
be a crystalline solid composed of Na+ and FI- ions. In the
crystalline form it should be a non-conductor. NaFI is a crys-
Na(+I) Fl(-!}
I
~-·-
I
I
._ _ _J _ _
I
I
Frn. 105.-A "molecule" of N aFl, illustrating the tendency to form octets.
talline, non-conducting solid with a high meltjng point (980°C.).
We might picture it diagrammatically by Fig. 105, in which the
Na+ ion is made up of a + lle nucleus with its pair of electrons,
around which, at the corners of the cube, are the remaining 8
electrons. The fluorine ion is simHar, except, perhaps, as to size,
but has a ninefold plus charge for its nucleus. The substance
N aFI should conduct electrolytically when melted.
Sodium, having 1 electron outside the stable group of 8 which
completes the neon structure, should have chemical properties
similar to lithium. Both should be monovalent, since they
each have a single electron to share. Sodium, like lithium, should
be a crystalline, conducting solid. (Its melting point is 97.5°C.)
Fluorine has only 7 electrons in its outside shell. But 2 fluorine
atoms, by uniting along an edge so as to hold 2 electrons in
common, could form two octets, as shown in Fig. 106. This
structure should approach in stability the neon structure which
it imitates. Thus, fluorine should be a diatomic gas with low
SEC. 2] THE LIGHTER ELEMENTS 421
boiling point. This is confirmed, the actual boiling point of
Fb being -189°C.
These combinations of Na and Fl illustrate the tendency to
form octets which imitate the stable gas neon.
(e) Carbon (Z = 6); Nitrogen (Z = 7); Oxygen (Z = 8).-
0xygen has 6 electrons surrounding its kernel. To make an
I
I •@• I I •@•
/
l--- ~---
I I /
I I
FIG. 106.-The F2 molecule. FIG. 107 .-The molecule of O 2.
octet 2 more are needed. Oxygen should, therefore, be divalent,
and we should expect such compounds as H20, Li20, Na20, etc.,
since each of the atoms H, Li, and Na have 1 extra electron.
Further, 2 atoms of oxygen could combine to form a structure
resembling a double octet, as shown in Fig. 107. This combina-
tion imitates neon. We should, therefore, expect that oxygen
should be a diatomic gas, 02, and should have a low boiling point.
I
._ ___
•@•_
I
I I I
I •@• 1 •@• I
f---- +- - - --·-------
1 /
I I
Frn. 108.~The mobcule of Oa.
The boiling point of 0 2 is actually -183°C. By joining another
cube to the edge of the double combination shown in Fig. 107,
we should predict the gas 03, as shown in Fig. 108. 03 has a boil-
ing point of -112°C.
The electron from a single hydrogen atom by combining with
a single oxygen atom, making the combination OH, would fill
seven of the eight places in the outside layer of the oxygen atom,
422 ARRANGEMENT OF ELECTRONS [CHAP. XI
one place still remaining vacant. The combination OH, there-
fore, should behave somewhat like fluorine, in that it should be
univalent. We would, thus, predict the compound N aOH,
which should be made up of Na+ ions and OH- ions, as repre-
sented in Fig. 109. Actually, N aOH is a solid with a melting
point at 318.4°C.
The unique properties of nitrogen are not so easily accounted
for. Nitrogen, in molecular form N 2, is a very inactive gas with
a very low boiling point ( -195°C.). It resembles, somewhat,
the inert gases, particularly argon (Z = 18). Langmuir assumes
that 2 nitrogen atoms combine their 14 electrons as follows: The
two kernels account for 4 electrons. Around these two kernels
is formed a stable octet, similar to the octet of neon. There
are then 2 electrons left over. Langmuir places these with the
+
: •®• : •®• ~-fl-1-f
I
t---- I
,----- ,- --t- - -
I
I I I
Fw. 109.-The "molecule" of NaOH, showing Frn. 110.-Langmuir'ssugges-
the ionic, or polar, structure. tion for the structure of N 2.
two kernels in the inside of the octet. The whole structure
(Fig. 110) externally resembles neon.
The carbon atom with 4 electrons surrounding its kernel is
midway between helium and neon. In making pairs or octets,
the carbon atom can either share its 4 electrons with other
atoms or take on 4 more electrons to complete its octet. We
should, therefore, expect the combination CH4 (methane, boiling
point -161 °C.), in which the 4 extra electrons necessary to
complete the octet surrounding the carbon atom are supplied
by the 4 hydrogen atoms, the nuclei of which become attached to
four of the corners of the octet. Following Pythagoras(!), we
assume that these 4 hydrogen nuclei are symmetrically distrib-
uted about the octet. The combination CH2 would be impos-
sible, or at least unstable, since no octet would thereby be
formed; but two CH 2 groups might unite, somewhat after the
manner of 2 oxygen atoms, to form the gas C2H4 (ethylene,
boiling point -104 °C.).
r
SEC. 2] THE LIGHTER ELEMENTS 423
The oxides of carbon and nitrogen are interesting. For
example, we have the well-known combinations C02 and N20.
In each of these, after taking out the 6 electrons associated with
the three kernels, we have 16 electrons left over to make up the
outside structure. Quite irrespective of the particular picture
which we draw to represent the 2 molecules, we should expect
that the 16 electrons would, in each case, be arranged about the
respective three kernels-which are of similar magnitudes as t6
net positive charge-in such a way as to result in similar struc-
tures. From this similarity alone, we should expect the 2 mole-
cules C02 and N 20 to have similar properties. We may draw
I I
I O c I O
I
)- - - -}- - -
I
- +----
/
I I I
(a) The C03 Molecule
I I I
IN N I O I
)-----/---- ---}----
/ I I
I I I
(h) TheN"zO Molecule
Frn. 111.-Similarity in molecular structure of C02 (a) and of N 20 (b).
pictures (Fig. 111, (a) and (b)) to help us keep jn mind this fact-
but the pictures probably bear no relation to the actual appear-
ance of the molecules. From the appearance of the schematic
molecules in Fig. 111, we should expect the two substances to
be gases not unlike oxygen in thermal properties.
Somewhat similar remarks as to comparative properties may
be made regarding the similarity of N 2 and CO, each of which,
after accounting for the 4 electrons belonging to the kernels,
have 10 electrons grouped around the two kernels. Whether
they both have the structure assumed by Langmuir for nitrogen
is immaterial-they should have similar properties. A compari-
son of the properties of these pairs of molecules, C0 2 with N 20 and
CO with N 2, is shown in Table II.
424 ARRANGEMENT OF ELECTRONS [CHAP. XI
TABLE II.-COMPARISON OF THE PROPERTIES OF MOLECULES WITH SIMILAR
STRUCTURES
C02 N20 co N2 02 Fb
Freezing point, de-
grees Centigrade .. ........ -102.4 -207 -209.8 -218.4 -223
Boiling point, de-
grees Centigrade .. -78.5 - 89.5 -192 -195.8 -183.0 -187
Critical pressure, at-
-, mospheres ........ 77 75 35 33 50
Critical temperature,
degrees Centigrade 32 35 -151 -146 -118
Critical volume, cm.3
per gram ........ 2.2 2.4 5.05 5.17 1.5
Viscosity at 20° Cen-
tigrade .......... 148 . 10-6 148. 10-6 163. 10- 6 166. 10-6 187. 10-6
Density of liquid at
boiling point ...... . . . . . . . .
I
......... 0.793 0.7961 1.14 1.11
Index of refraction
(liquid) .......... 1.190 1.193J .......... 1.197 1.221
It may be pointed out that, while the theory predicts that
these substances ought to have simnar properties, there is no
suggestion whatever of a method of computing those properties
from fundamental constants or even of making a quantitative
comparison with the properties of the inert gases. Further,
the similarity loses some of its significance when it is observed
that the properties of 0 2 and of Fl 2 are quite similar, although
these two molecules are assumed to have dissimilar structure
(see Figs. 106 and 107).
Nitrogen, with a total of 7 electrons, 3 short of enough to
complete an octet surrounding the kernel, could combine with
3 hydrogen atoms, making the molecule NH3 (ammonia) which
should have a neon-like structure, except for the 3 hydrogen
nuclei attached to three corners. Ammonia should have a
low boiling point. Actually, its boiling point is - 33, and its
freezing point is-78°C.
One would not expect that 4 hydrogen atoms could combine
with a nitrogen atom to make a stable combination, since there is
1 more electron than enough to make an octet. We might
expect, however, that 4 hydrogen atoms together with a nitro-
gen atom might behave much like sodium toward combining
with either an atom or a combination of atoms which lacks 1
electron to complete an octet. Thus, we should predict the
well-known ammonium (NH4) compounds, the ammonium form-
ing the positive ion, as in (NH4)+ + (OH)-.
SEC. 3] THE HEAVIER ELEMENTS 425
The student will find it interesting and, up to a certain point,
instructive to "build" other molecular models by use of this
general principle that combinations tend to take place in such a
way as to form pairs or octets. The illustrations which have
been given are sufficient to indicate the importance of this
principle in any theory which has to do with the distribution of
electrons in atoms.
(f) Magnesium (Z = 12); Silicon (Z = 14); Phosphorus (Z =
15).-The properties of magnesium are determined largely by
the fact that it has 2 electrons more than enough to complete
the octet. We may assume that its structure is represented by
(2, 8, 2). Hence, we should expect such stable combinations
as MgO-periclase (melting point 2800°C.)-and MgF 2 (melting
point 1400°C.).
Silicon, with 4 electrons jn addition to the underlying octet
and with a probable structure (2, 8, 4), should behave much like
carbon. Thus, we have Si0 2 (quartz), SiH4 (gas, melting point
-185; boiling point -112°C.), and SiF 4 (silicon tetrafluoride,
melting point -77; boiling point -65°C. under 2}~ atmospheres).
Phosphorus (assumed structure (2, 8, 5)) like nitrogen lacks
3 electrons of filling an octet (the second octet for phosphorus).
The properties of phosphorus should, therefore, be somewhat
similar to those of nitrogen in the formation of compounds.
Thus, we have the well-known gas PH3-phosphine-analogous
to NH3. Because of the greater charge on the nucleus, however,
and the greater number of electrons involved, one would not
expect 2 phosphorus atoms to unite to form a neon-like struc-
ture as was postulated for N2 (Fig. 110). Langmuir 1 discusses
many of the compounds in which phosphorus enters.
3. The Heavier Elements. (a) General Factors Determining
Chemical and Spectroscopic Properties.-Passing by, for the
moment., the question of how the electrons are distributed in
the atoms of the heavier inert gases-Ar(18), Kr(36), Xe(54),
and Nt(86)-but taking it as an experimental fact that their
structures are stable, one readily sees that the next element
beyond any one of the permanent gases should have 1 more than
enough electrons to form a stable structure. These elements,
together with lithium and sodium, are the metals of the alkali
group-K(39), Rb(37), Cs(55)-the chemical properties of which
are strikingly similar. The second element beyond a permanent
1
J our. Amer. Chem. Soc., vol. 38, p. 762 (1916).
426 ARRANGEMENT OF ELECTRONS [CHAP. XI
gas should have 2 superfluous electrons. Hence, we have the
series of similar elements, commonly divalent-Mg(l2), Ca(20),
Sr(38), Ba(56), Ra(88). Going in the other direction, the next
element before a perfect gas lacks 1 electron of enough to make a
stable structure. These elements are the halogens-F(9),
Cl(17), Br(35), 1(53), and perhaps the undiscovered element
(85). The series of elements next preceding the halogens are
0(8), 8(16), Se(34), Te(52). These, likewise, are similar.
The prediction of the similarities of these series of elements
is quite independent of any assumption concerning the structure
of the atom of the neighboring inert gas. Calcium and barium,
for example, are similar, whatever their internal electron struc-
tures may be, simply because they have 2 electrons outside the
stable shells corresponding to the "su;face" layer of the
preceding inert gases. These gases themselves are inert, so we
believe, because, in addition to their stable internal structures,
they possess completed, or satisfied, superficial groups of
electrons. A comparison of the series systems of spectra in
these groups of chemically similar elements shows that they
have similar spectra. The spectral series (arc spectra) of
the alkali elements consist of doublets; in the arc spectra of
the elements of the Mg-Ca-Sr-Ba group are found two sys-
tems of lines: a complete singlet system containing S, P, D,
and F terms and a similarly complete triplet system. Other
groups of elements in the periodic table show like regularities
as regards the multiplicity of their spectra. Also, the spark
spectra of the alkali metals (i.e., the additional lines brought out
by the more intense spark discharge) are very similar to the arc
spectra, respectively, of the previous inert gases; and the spark
spectra of the elements of the Mg-Ba group are similar, in turn,
to the arc spectra of the previous alkali metal. We may ascribe
spark spectra as originating in atoms which, on account of the
intensity of the excitation to which they have been subjected,
have lost 2 electrons, instead of only 1 as in arc spectra. This
similarity between the spark spectrum of an element and the
arc spectrum of the element of next lowest atomic number is,
thus, quite understandable, since, for example, a Sr(38) atom
which has lost 2 electrons should be very similar to a Rb(37) atom
which has lost only 1 electron, the two structures Srrr and Rb 1
having the same total number of extranuclear electrons and
differing mainly in the magnitude of the nuclear charges.
SEC. 3] THE HEAVIER ELEMENTS 427
All these similarities, both chemical and spectroscopic, suggest
a very fundamental principle, which we shall find useful in dis-
cussing further the arrangement of electrons in atoms, namely,
that the chemical and the spectroscopic 1 properties of an atom are
determined much more by the number and the arrangement of the
electrons in the outside shell than by the total number of electrons in
the atom. Conversely, by studying the chemical and spectro-
scopic properties of atoms, we may obtain information as to
the number of electrons in the surface layer, but not the number
or arrangement in inside layers. This latter information regard-
ing the more deeply lying electrons comes largely from X-ray
phenomena, which we shall discuss in the next chapter.
In support of this general principle, one may compare one by
one the elements Li to Ne inclusive with the corresponding ele-
ments in the sequence, Na to A inclusive. Li and Na are similar
because they each have 1 more electron than enough to complete
a stable structure; Be and Mg each have 2 "superfluous" elec-
trons; Band Al have 3; C and Si have 4; etc.
For making this comparison clear, the periodic arrangement of
the elem en ts as proposed by Bohr (Appendix 1 (c)) is useful.
In this diagram, the elements are arranged in periods, a period
constituting the sequence of elements from one inert gas to and
including the next. The first period contains H and He.
Period II, the first "short period," extends from Li to Ne; and
Period III, the second short period, from Na to Ar. Then
come two long periods of 18 elements each, K(19) to Kr(36)
and Rb(37) to Xe(54). Period VI, sometimes called the "rare-
earth period," contains 36 elements, Cs(55) to Nt(86). The last
period, VII, is incomplete (at least among terrestrial elements).
(b) The Elements of the Fourth Period.-The elements of Period
II-Li(3), Be( 4) . . . N e(lO)-contain, respectively, in the
outside shell 1, 2 . . . 8 electrons. Similarly, the outside
shells of the corresponding elements of Period III contain 1,
2 . . . 8 electrons. Comparing those elements which have
identical numbers of electrons in the outside shell, we find that
Li is similar to Na; Be is similar to Mg . . . N is similar to
P; and so on. These similarities are indicated by the hori-
zontal lines between the respective members of the two periods
in Appendix I(c).
1 "Spectroscopic," as here used, is limited to that part of the spectrum in
the neighborhood of the visible region.
428 ARRANGEMENT OF ELECTRONS [CHAP. XI
Comparing Period III with Period IV, we find that Na is
similar to K; Mg is similar to Ca. Beginning with Sc(21), in
Period IV, however, the similarity with the corresponding ele-
ment in group III becomes less. Scandium (21), with 3 electrons
more than enough to complete the second stable group of 8, is
somewhat similar to Al(13) and Boron (5). All are, commonly,
trivalent. Thus, we have the oxides B203, Ah03, Sc203 and the
chlorides BC13, AlCh, ScCl3. But a further comparison of
properties shows that the similarity in chemical behavior is not
so great as might be expected if the extra 3 electrons were iden-
tically arranged in the three elements. They differ in their
hydrates and in the solubilities of some of their salts. Differ-
ences of this sort increase as one goes farther down the two
periods. In general, the elements of Period IV, between Sc(21)
and Co(27), show a greater tendency to variable valency than
corresponding elements in Period III. Furthermore, the mag-
netic elements iron, cob9Jt, and nickel have no analogues in
Period III. In addition to their being magnetic, these three
magnetic elements are similar to each other in other respects.
They have, in addition to variable valency, many similar
hydrates of their various salts; thus:
FeS04 · 7H20 (NH4)2S04 · FeS04 · 6H20
NiS04 · 7H20 (NH4)2S04 · NiS04 · 6H20
CoS04 · 7H20 (NH4)2S04 · CoS04 · 6H20
Beginning at the other end of the two periods and working
backward, the similarity extends for several elements. Preced-
ing the inert gases argon(18) and krypton(36) come the halogens
chlorine(l 7) and bromine(35). Preceding these in turn are the
similar elements sulphur (16) and selenium (34). Then come
phosphorus (15) and arsenic(33); silicon(14) and germanium (32);
aluminum(l3) and gallium (31). Magnesium (12), although
very similar to calcium (20), the second member of group IV, is
also very similar to zinc(30). Magnesium and zinc are both
divalent and form very similar compounds. The arc spectrum of
zinc contains a complete system of triplets and another complete
system of singlets, just as do the corresponding spectra of the
members of the Mg-Ba group of elements. Both chemical
and spectroscopic evidence indicates that the external electron
structure of the zinc atom should be similar to the structures of
magnesium and calcium. These latter elements, being, in each
SEC. 3] THE HEAVIER ELEMENTS 429
case, the second beyond an inert gas, probably have 2 electrons
outside of the outermost stable group of 8. The conclusion is
obvious 1 therefore, that zinc, although in the middle of the
period, contains 2 electrons outside of a stable group. Now,
zinc has 30 extranuclear electrons. If we represent the structure
of Mg(12) by (2, 8, 2) and of Ca(20) by (2, 8, 8, 2), Zn(30) should,
similarly, be (2, 8, 18, 2), the third shell containing 18 electrons.
Copper, immediately preceding zinc, has a spectral-series
system containing doublets, just as have the alkali elements.
This, together with the fact that in the well-known series of
cuprous compounds copper is monovalent, suggests the structure
for copper (2, 8, 18, 1), with the alternative for divalent copper
(2, 8, 17, 2). Comparing either of these with the structure of
calcium (2, 8, 8, 2), we see that in Period IV, after passing cal-
cium, electrons in the succeeding elements, instead of being added
to the fourth shell, have been added to the third shell, increasing
the latter from 8, in the case of calcium, to 18, in the case of
copper and magnesium. After the third shell has been built up
to 18 electrons, there seems to be no further tendency for it to
increase, since the elements of Period IV, following zinc, have
analogues in Period III in the elements following magnesium.
Taking the structure of silicon (14) to be (2, 8, 4), that of the
analogous element in Period IV, namely, germanium (32),
should be (2, 8, 18, 4). Similarly, we should expect for selenium
(34) the structure (2, 8, 18, 6); and, finally, for krypton (2, 8, 18,
8). Thus, in krypton, as in argon, the outer shell contains 8
electrons. The sequence of numbers as suggested by Rydberg
(p. 415) still holds, only we must invert the order of the last two,
viz., 2(1 2 + 2 2 + 3 2 + 2 2 ).
It appears, therefore, that the elements scandium to nickel
inclusive, in Period IV, form a transition series during which the
number of electrons in the third shell is being built up from 8 to
18. (These are the elements of Period IV enclosed in the rec-
tangle in Appendix I(c).) Bury 1 makes the following postulate:
The maximum number of electrons in the outer layer of an atom
is 8; more than 8 electrons can exist in a shell only when there is an
accumulation of electrons in an outer layer. During the change of an
inner layer from a stable group of 8 to one of 18 (as from scandium to
nickel) or from 18 to 32, there occurs a transition series of elements
which can have more than one structure.
1 J our. Amer. Chem. Soc., vol. 43, p. 1602 (1921).
430 ARRANGEMENT OF ELECTRONS [CHAP. XI
That is to say, the elements in this transition series have variable
valencies, as is illustrated by the various oxides of manganese.
Bury proposes the following as possible structures for some of the
elements of this transition series:
Cr(24): (2, 8, 8, 6); (2, 8, 11, 3); (2, 8, 12, 2)
Mn(25): (2, 8, 8, 7); (2, 8, 11, 4); (2, 8, 12, 3); (2, 8, 13, 2)
Fe(26): (2, 8, 10, 6); . . . ; (2, 8, 12, 4); (2, 8, 13, 3)
(c) Period V, the Second "Long Period."-This period contains
exactly the same number of elements, namely 18, as the pre-
ceding period, Period IV. Arguing from the similarities of the
respective elements in Periods II and III, each of which contains
eight elements, we may infer that each element of group V should
have a homologous element in Period IV. This proves to be the
case. At the beginning of the periods, we have the alkali ele-
ments K and Rb. Next come the alkaline earths Ca and Sr.
Chromium (24) and molybdenum (42) both form very similar
compounds. Corresponding to the elements Fe, Co, Ni, in
Period IV, are the elements Ru, Rh, Pd, in Period V. Although
the latter are not ferromagnetic, they have many chemical simi-
larities to the magnetic elements Fe, Co, Ni. Ag (4 7) is similar
to Cu (29) as regards both chemical and spectroscopic properties.
The similarity continues, growing more striking, until we reach
the end of the period.
This similarity of the elements of Period V with corresponding
elements of Period IV leads to the belief that we have, in Period
V, transitions in electron arrangements similar to those found
in Period IV. Accordingly, Rb(37) should have the structure
(2, 8, 18, 8, 1); Sr (38), the structure (2, 8, 18, 8, 2). Then comes
the "transition series," Y(39) to Pd(46), in which the fourth
shell is being built up from 8 to 18 electrons. Ag(47) should
have a structure (2, 8, 18, 18, 1); Cd( 48), (2, 8, 18, 18, 2); . . .
1(53), (2, 8, 18, 18, 7); and Xe(54), (2, 8, 18, 18, 8). Again we
find that the stable outside layer contains 8 electrons. \Ve may
write the Rydberg series of numbers for xenon: 2(1 2 + 2 2 + 3 2
+ 32 + 22).
(d) Period VI, the Rare-earth Period.-The first two elements
of this period, namely, cesium (55) and barium (56), are exactly
analogous to the first two elements of each of the preceding
periods, and we may predict the respective structures (2, 8, 18,
SEC. 3] THE HEAVIER ELEMENTS 431
18, 8, 1) and (2, 8, 18, 18, 8, 2). Beginning with lanthanum
(57) and extending through (and including) platinum(78), two
important transition series are found: (1) the first transition
series1 in which the fourth shell, containing 18 electrons in
lanthanum, is built up to 32 electrons from lanthanum to-
lutecium (71); (2) the second transition series, hafnium (72)
to platinum (78), in which the fifth layer, containing probably 9
electrons in lutecium, is built up to 18 in platinum, which number,
by comparison with xenon at the end of Period V, should give a
stable combination. With gold (79) begins the building up of
the sixth, or outside, layer of the elements of Period VI, to 8
electrons, a process which is complete with niton (86). The
structure of Nt should, thus, be (2, 8, 18, 32, 18, 8). This
distribution of electrons among the several layers is again repre-
sented by Rydberg's series of squared integers, by rearrange-
ment, viz., 2(l2 + 2 2 + 3 2 + 4 2 + 3 2 + 22 ).
(e) The Seventh Period.-This period contains only three well-
known elements-radium (88), thorium (90), and uranium (92).
The unknown element (87), should be similar to the alkali ele-
ments. Radium is similar to the elements of the Mg-Ba group.
With actinium (89) probably begins the building up of the sixth
shell from 8 electrons in radium to 12 electrons in uranium, the
last of the known elements. The structure of radium should be
(2, 8, 18, 32, 18, 8, 2); of uranium, (2, 8, 18, 32, 18, 12, 2).
(f) Table III, reproduced, in part, from similar tables based
on spectroscopic data, shows the distribution of electrons in the
various levels for all the atoms of the periodic table. The
arrangement is similar to Bohr's table (Appendix I(c)), in that all
the elements of a given period are arranged in one column. The
several groups of electrons in each element are designated by a
number n = 1, 2, 3 . . . , as shown at the top of each column.
This number n may be identified with the total quantum
number i' of the successive orbits in Bohr's scheme as discussed
in Sec. 5, below.
We shall see in Sec. 5 that it is necessary still further to sub-
divide the groups of electrons which we have here designated
by n.
1 The elements of this transition series, the rare earths, are similar to
the transition series Se to Ni of Period IV but, according to Bury, "do
not show the variety of structure" characteristic of the latter series.
~
Ci,j
~
TABLE IIL-THE DISTRIBUTION OF ELECTRONS IN ORBITS CHARACTERIZED BY TOTAL QUANTUM NuMBER n
Period VI Period VII
n=12 3 4 5 6 n = 1 2 3 4 5 6 7
Cs 2 8 18 18 8 1 2 8 18 32 18 8 1
Ba 2 8 18 18 8 2 Ra 2 8 18 32 18 8 2 ~
La 2 8 18 18 9 2 Ac 2 8 18 32 18 9 2 ~
Ce 2 8 18 19 9 2 Th 2 8 18 32 18 10 2 ~
Period IV Period V Pr 2 8 18 20 9 2 Pa 2 8 18 32 18 11 2 ~
Nd 2 8 18 21 9 2 U 2 8 18 32 18 12 2
~~
n = 1 2 3 4 n = 1 2 3 4 5 II 2 8 18 22 9 2
K 2 8 8 1 Rb 2 8 18 8 1 Sa 2 8 18 23 9 2
Period II Ca 2 8 8 2 Sr 2 8 18 8 2 Eu 2 8 18 24 9 2
Period III Sc 2 8 9 2 y 2 8 18 9 2 Gd 2 8 18 25 9 2
n = 1 2 n=l23 Ti 2 8 10 2 Zr 2 8 18 10 2 • Tb 2 8 18 26 9 2 tlj
Period I Li 2 1 Na 2 8 1
v 2 8 11 2 Nb 2 8 18 12 1 Dy 2 8 18 27 9 2 ~
Be 2 2 Mg Cr 2 8 13 1 Mo 2 8 18 13 1 Ho 2 8 18 28 9 2
2 8 2 Mn 2 8 13 2 Ma 2 8 18 14 1 Er ~
n=l Bo 2 3 Al 2 8 3 2 8 18 29 9 2
Hl Fe 2 8 14 2 Ru 2 8 18 15 I Tu 2 8 18 30 9 2
He2
c
N
2
2
4
5
81
p
2 8 4
2 8 5
Co 2
Ni 2
8
8
15
16
2
2
Rh
Pd
2 8 18 16 1 Yb 2 8 18 31 9 2 ~
0 2 6 s 286 2 8 18 18 Lu 2 8 18 32 9 2
F
Ne
2 7
2 8
Cl 2 8 7
Ar 2 8 8
Cu 2
Zn 2
8
8
18
18
1
2
Ag
Cd
2
2
8
8
18
18
18
18
1
2
Hf
Ta
2
2
8
8
18
18
32
32
10
11
2
2 ~
tlj
Ga 2 8 18 3 In 2 8 18 18 3 w 2 8 18 32 12 2 ("')
Ge 2 8 18 4 Sn 2 8 18 18 4 Re 2 8 18 32 13 2
As
Se
2
2
8
8
18
18
5
6
Sb
Te
2
2
8
8
18
18
18
18
5
6
Os
Ir
2
2
8
8
18
18
32
32
14
15
2
2 ~
Br 2 8 18 7 I 2 8 18 18 7 Pt 2 8 18 32 16 2 0
Kr 2 8 18 8 Xe 2 8 18 18 8 Au 2 8 18 32 18 1 ~
Hg 2 8 18 32 18 2 ~
Tl 2 8 18 32 18 3
Pb 2 8 18 32 18 4
Bi 2 8 18 32 18 5
Po 2 8 18 32 18 6
2 8 18 32 18 7
Nt 2 8 18 32 18 8 Q
p:j
>
~
~
~
SEC. 4] ELECTRONS: DISTRIBUTION IN ORBITS 433
p ART II. THE DYNAMIC ATOM MODEL
4. The Problem of the Distribution of Electrons in Orbits.-
By considerations based largely on optical spectra, Bohr 1 pro-
posed, regarding the distribution of electrons in the various
possible orbits surrounding a nucleus, a system which, in its
general features, is somewhat similar to that proposed by Bury
for the distribution of electrons in various levels or shells. Bohr
assumes that around a nucleus of atomic number Z there are, in
the normal, neutral atom Z electrons revolving jn orbits which
are characterized by radial and azimuthal quantum numbers.
Using the terminology suggested on page 372, there are around
each nucleus various "privileged" orbits designated, respectively,
by 11, 21, 22, 31, 32, 3s, 41 . . . (The number stands for the
total quantum number n = r, and the subscript for the azimuthal
quantum number k =
r <t>· Thus, a 32 orbit is an orbit of
total quantum number 3, azimuthal quantum number 2, and
radial quantum number 1. In general, an orbit is designated
by nk, where n is the total quantum number and k the azimuthal
quantum number.)
We saw, in Chap. X, equation (80'), that the energy Wn,k of
an nk orbit around a nucleus of charge Z is given by the equation
Wn,k
= _
2
21r mZ e I!_
h2 Ln2
2 4
+ !_ . 41r Z e
n4
2
c2h2
2 4
(!!k _ ~)]
4 (2)
This equation is equivalent to equation (80'), Chap. X. The
changes in symbols are as follows:
(rr + r</>) = n; r</> = k.
For a given nuclear charge Z, the larger the total quantum num-
ber n the greater (algebraically) is the energy of the orbit; and for
a given value of n, the energy is greater the greater k; i.e., orbits
of a given quantum number n are at a higher energy level the less
the eccentricity. This, however, applies only to the case of a
nucleus and a single electron in an nk orbit. The situation is
much more complex jf the nucleus is surrounded by other elec-
trons; as prevjously pointed out, it is then impossible to obtain
a rigorous expression for W.
The method, initiated by Bohr and developed by him and
others, of assigning electrons to various orbits on the basis of
1 BoHR, N.: "The Theory of Spectra and Atomic Constitution," Chap.
III.
434 ARRANGEMENT OF ELECTRONS [CHAP. XI
spectroscopic data consists in picturing the sequence of events in
a hypothetical experiment in which a nucleus of charge + Ze,
initially stripped of all its electrons, acquires one by one the Z
electrons necessary to make a neutral atom. Bohr assumes that
in this process each added electron goes into such a vacant permis-
sible orbit as makes the energy of the system, after that electron
has been added, a minimum, a condition which is assumed to
give greatest stability. In deciding which orbit the next "cap-
tured" electron will permanently occupy, Bohr is guided, in part,
by equation (2) and, in part, by the terms in the optical spectra,
arc and spark, of the element concerned. We shall discuss very
sketchily in the next section some of the factors which are
consjdered in ascertaining which orbits the various electrons in
the normal atoms occupy.
5. Some Examples Illustrative of the Spectroscopic Method.
(a) Hydrogen.-The orbit of lowest level is, of course, the 1 1
orbit, i.e., the circle of total quantum number 1. Hence, the
electron captured by a hydrogen nucleus ultimately reaches and,
in the normal atom, continues to move in the Ii orbit.
(b) Lithium; Effective Quantum Numbers.-Passing by the
arguments by which Bohr concludes that both of the electrons
. . . -, in helium move in Ii orbits, we come to lith-
-{ @ )- 0
_ ium. The first two electrons captured by
a, . . _/ b c the lithium nucleus will, according to Bohr,
occupy the two possible 11 orbits as in the
Frn. 112.~The ker-
nel of the lithium atom case of helium. We know from chemical
with its net charge of data that the third electron of lithium is
+e should behave as a
~ingle-charged nucleus much less closely bound than are the first two
toward the third more -a fact which, in the orbit scheme, suggests
distant electron c.
that the third electron revolves in an orbit
of total quantum number larger than 1, probably in a 21 orbit,
since this would give to the lithium atom a lower energy than
if the third electron were to occupy a 2 2 orbit. This surmise is
strengthened by considering the spectrum of lithium.
Lithium, in common with the other alkali elements, has four
sequences of terms-S, P, D, and F terms-the combinations of
which give rise to the four main series-principal, sharp, diffuse,
and fundamental. Although we cal).not apply .to the lithium atom
any equation of the type of equation (2) for the determination of
the terms from fundamental constants, we can make some quali-
tative inferences based on the equation. The lithium nucleus,
SEC. 5] THE SPECTROSCOPIC METHOD 435
with its two close-in electrons a and b, (Fig. 112) revolving in 1i
orbits, will behave toward a third electron c at a greater distance
from the nucleus almost as if the nucleus together with a and b
constituted a single positive charge +e. Or, we may say that a
and b partially screen the nucleus, so that its effective charge is
less than +3, so far, at least, as concerns points at considerably
greater distances from the nucleus than the radius of the l 1 orbits.
Accordingly, if the lithium atom be ionized by the removal of
electron c, we should expect that the resulting spectrum when
the electron is recaptured would be "hydrogen-like," at least
for interorbital transfers involving quantum numbers of, say, 3
or larger.
Now, we obtain the terms of the hydrogen spectrum by dividing
the Rydberg constant N by the square of the total quantum
00
numbers n of the orbits-speaking, for the moment, only of
circular orbits-thus the expression:
where n stands for the successive integers 1, 2, 3 , gives
the terms of the hydrogen spectrum. In spectra other than
hydrogen, the terms are obtained by dividing N by such 00
quantities as (m + µ) 2 , where mis an integer andµ is a constant.
Thus, the 1r terms in the spectra of the alkali elements are given
by (see Chap. IX)
,r(m) = (m !"' ,r)2
By comparison with the hydrogen spectrum, we may think of
the quantity (m + 1r) as the effective quantum number ne of the
term. Knowing the values of the terms, say in wave numbers,
from spectral-series data, the effective quantum number can be
obtained at once by the relation
Noo
Term va1ue = -
ne2
The wave numbers of the S, P, D, and F terms in the spectrum
of lithium 1 are given in Table IV, together with the computed
effective quantum numbers.
1 Data taken from Fowler's report.
436 ARRANGEMENT OF ELECTRONS [CHAP. XI
TABLE IV.-EFFECTIVE QUANTUM NUMBERS ne IN LITHIUM. N co = 109,720
S terms P terms D terms F terms
Wave
I Wave
Wave Wave
ne ne ne ne
number number number number
I
43,486 1.588 28,582 1.959 12,203 2.998 6,856 4.000
16,280 2.595 12,560 2.955 6.863 3.998 4,382 5.003
8,475 3.598 7,018 3.953 4,390 4.999
5,187 4.599 4,474 4.952 3,047 6.000
The effective quantum numbers, computed directly from
spectral data, for the two F terms are seen to be almost exactly the
integers 4 and 5. These two terms correspond very closely to
terms which would be computed for a hydrogen nucleus for the
quantum numbers 4 and 5. The inferences, therefore, seem
justified (1) that these two F terms correspond to actual quantum
numbers 4 and 5, the lowest F term having quantum number 4;
and (2) that for the corresponding orbits we may regard the
core ( +3 nucleus and 2 electrons) of the lithium atom as equiva-
lent to a single-charged, hydrogen-like nucleus. For these orbits,
the inverse-square law, therefore, holds_, since the inverse-square
law was assumed for the hydrogen orbits.
Likewise, the effective quantum numbers of the D terms are
very nearly the respective integers 3, 4, 5, 6; from which we may
infer that the lowest D term has quantum number 3 and cor-
responds to a circular orbit n = 3.
The effective quantum numbers for the P terms are a little
less than the respective integers 2, 3, 4, 5. This corresponds to
a stronger ''binding'' of the electron in the respective orbits
than if the numbers were exactly 2, 3, 4, 5. We may think of
this increased binding in these orbits, particularly orbit number
2, as resulting from an increase in the effective nuclear charge,
which, in turn, is due to a decrease in the screening effect of the
2 electrons in l 1 orbits.
The S terms have effective quantum numbers which differ
considerably from integers. To what orbit may we assign the
lowest S term, for which the effective quantum number is 1.588?
Not to a 2-quantum orbit, of the same kind as the lowest P term,
since the binding is considerably stronger than that of the lowest
P term; not to a 11 orbit, since the binding is not so strong as it
SEC. 5] THE SPECTROSCOPIC METHOD 437
is for the first 2 electrons. The next orbit above the Ii orbit is
the 21 orbit, the ellipse. We may guess that the lowest S term,
therefore, corresponds to a 21 orbit and that the lowest P term
corresponds to a 22 orbit.
The difference in energy between the lowest S term and the
lowest P term, however, -43,486 as against 28,582-is greater
than equation (2) (see, also, equation (83)) leads us to expect for
the difference in energy between the 21 and the 22 orbit. If these
two terms correspond to the suggested orbits, how can the great
difference in energy be explained? The explanation involves a
very fundamental principle. The 22 orbit is a circle of quantum
number 2. It is symmetrically situated with respect to the core
of the atom and is in a field which departs only a little from the
inverse-square law, since its effective quantum number is very
nearly 2. On the contrary, the 21 orbit is an ellipse. At perihe-
lion, the electron is much closer to the nucleus than is any part of
the 22 orbit. We see that even with the 22 orbit the screening
effect of the two core electrons becomes less than with orbits of
higher quantum number, and, accordingly, we should expect the
perihelion part of the 21 orbit to be much less screened by the first
2 electrons than is any part of the 22 orbit. An electron in the
21 orbit should, therefore, be much more strongly bound than
one in the 22 orbit.
This penetration of elliptical orbits of higher quantum number
into the orbits of lower quantum number and of lower eccen-
tricity results in a much greater binding energy for the more
eccentric orbits than for the lesser eccentric orbits of the same
total quantum number. It may even happen in elements of
higher atomic number that the binding energy correspondjng to
an eccentric orbit of given quantum number, such as a 51 orbit,
may become greater than the binding energy corresponding to
a lesser eccentric orbit, say a 43 orbit, of lower quantum number.
Now, since the first 8 term is the lowest observed term in the arc
spectrum of lithium, we may assume that the orbit corresponding
to this term-the 21 orbit-is the lowest orbit reached by the third
electron of Whium. Or, we might say that the 21 orbit is the basic
orbit for the third electron. We conclude, therefore, that the
first 2 electrons in lithium normally occupy 1i orbits; and the
third, a 21 orbit.
Returning to the lowest D and F terms, we conclude that
since their effective quantum numbers are integers, the orbits
438 ARRANGEMENT OF ELECTRONS [CHAP. XI
lie entirely in an inverse-square field and either are circular or
are ellipses of low eccentricity. For otherwise we should expect
a marked departure of the effective quantum numbers from
integer values, as occurs in the case of the 21 orbit. We may
guess, provisionally, that the lowest D orbit, therefore, is a 3 3
orbit; and the lowest F orbit, a 4 4 orbit.
The orbits corresponding to the lowest terms in each of the
four sequences S, P, D, and Fare, thus, respectively, 21, 22, 33, 44.
The azimuthal quantum numbers (i.e., subscripts) of these orbits
are seen to be, respectively, 1, 2, 3, 4. There is evidence to
support the view that the higher terms of a sequence have the
same azimuthal quantum number as the lowest term of the
sequence but that the total quantum number presumably increases
by unity from one term to the next. Thus, the orbits correspond-
ing to the S sequence in lithium should be 21, 31, 41, . . . There
is considerable uncertainty in the assignment of the total quantum
number to the spectral terms in the several elements. There is
less ambiguity regarding the azimuthal quantum numbers.
(c) Sodium, Potassium, Rubidium, and Cesium.-We pass on
to the other alkali elements. From the similarity of their
spectra with that of lithium, which we ascribed to the binding
of the third electron, we may conclude that the spectra of sodium,
potassium, rubidium, and cesium arise, also, from the process of
binding a single electron, namely, the eleventh, nineteenth,
thirty-seventh, and fifty-fifth, respectively. We saw from the
discussion of the lithium spectrum that much information may
be gained concerning orbits from a consideration of the effective
quantum numbers of the terms of the several sequences. When
an atom has lost 1 electron, the remainder of the atom-the core
-has a net positive charge +e, exactly the same as has the hydro-
gen nucleus. The hydrogen nucleus is surrounded by an inverse-
square field-sometimes called a "coulombian field," from
Coulomb's law-and the terms in the hydrogen spectrum have
integer quantum numbers. When we find that the terms of a
sequence in the spectrum of some other element have effective ·
quantum numbers which are successive integers (as the F terms
in lithium), we conclude that the corresponding orbits (1) lie
in an inverse-square, or coulombian, field and (2) do not pene-
trate the more deeply lying orbits. Conversely, a material
departure of the effective quantum number from an expected
integer value indicates a stronger binding of an electron in that
SEO. 51 THE SPECTROSCOPIC METHOD 439
orbit due either to interpenetration of the orbit or to some other
departure from the inverse-square law.
Table V gives the effective quantum numbers of the first four
terms in the S, P, D, and F sequences in the alkali elements, com-
puted from term values in the same way as for lithium in Table
IV.
TABLE V.-EFFECTIVE QUANTUM NUMBERS AND CORRESPONDING ORBITS
FOR THE TERMS IN THE SPECTRA OF THE ALKALI ELEMENTS
Effective quantum numbers I Corresponding orbits
Term
number
s I p I D I F I s I p I D I F
Na
1 1.63 2.12 2.99 4.00 31 32 33 44
2 2.64 3.13 3.99 5.00 41 42 43 54
3 3.65 4.14 4.99 6.00 51 52 53 64
4 4.66 5.14 5.99 .... 61 62 6a 74
1 1. 77 2.24 2.85 3.99 41 42 33 44
2 2.80 3.26 3.79 4.99 51 52 43 54
3 3.81 4.27 4.76 5.99 61 62 53 64
4 4.81 5.27 5.75 6.99 71 72 63 74
Rb
1 1.80 2.29 2.76 3.99 51 52 33 44
2 2.84 3.32 3.70 4.98 61 62 43 54
3 3.85 4.34 4.68 5.97 71 72 53 64
4 4.85 5.35 5.67 6.97 81 82 63 74
Cs
1 1.87 2.36 2.55 3.97 61 62 33 44
2 2.91 3.40 3.52 4.97 71 72 43 54
3 3.93 4.42 4.52 5.97 81 82 53 64
4 4.94 5.42 5.52 6.97 91 92 63 74
In sodium, the F terms have the integer values 4, 5, 6
As in lithium, we may assume that these orbits are 4 4, 5 4, 6 4
. . . and that they do not penetrate the sodium core, which
presumably differs from the structure of the previous inert gas
(neon) only in having a nuclear charge of + lle. The D terms
440 ARRANGEMENT OF ELECTRONS [CHAP. XI
differ only slightly from the integers expected for "hydrogen-
like" terms, and we conclude that the D orbits are similar to those
in lithium, namely 33, 43, 53 . . .
The effective quantum number of the lowest P term exceeds 2.
Since departure from the inverse-square law should make the
effective quantum number less, instead of greater, than that of
the corresponding hydrogen term, we assume that the quantum
number of the first P term is 3 in sodium. (It was 2 in lithium.)
Since the binding of this first P orbit is greater than that of the
first D orbit, which is a 3s orbit, the first P orbit must be a 3 2
orbit, and the succeeding P terms correspond to 42, 52, 6 2 • • •
orbits.
The effective quantum number of first S term, the lowest in the
arc spectrum of sodium, departs less from the integer 2 than did
the first term of the S sequence in lithium, although, because of
the much larger core of sodium and the greater nuclear charge,
there should be more deviation from the corresponding hydrogen
quantum number than in lithium. If, however, we assume that
the first S term of sodium corresponds to an orbit of quantum
number 3, instead of to an orbit of the quantum number 2, as
in lithium, we provide for the expected greater departure. If
this be correct, the first S orbit in sodium-the basic orbit, since
it is the lowest-must be a 31 orbit. This is the orbit, then, into
which the eleventh electron of sodium finally "settles."
It is pertinent to inquire at this point, In what orbits are the
other 10 electrons of sodium to be found? The first 2 occupy
11 orbits, and the third a 21 orbit, as in lithium. The remaining
7 presumably are distributed among 21 and 22 orbits, since they
must all be more closely bound than is the eleventh electron,
which occupies a 3 1 orbit. We shall say a word, presently, about
the distribution of these 7 electrons in the 21 and 22 orbits.
From a similar study of the effective quantum numbers in
the other alkali elements, we may conclude that the basic orbits
of the nineteenth electron in potassium, the thirty-seventh in
rubidium, and the fifty-fifth in cesium are, respectively, 41, 51,
and 61.
(d) Magnesium and Aluminum.-Magnesium has 12 electrons
- 1 more than sodium. The spectrum of Mg1 1 consists of a
1 That is, the arc spectrum of magnesium. This is the spectrum produced
when Mg1, which is the magnesium "core" remaining after neutral mag-
nesium has lost 1 electron, recaptures the twelfth electron.
SEC. 5] THE SPECTROSCOPIC METHOD 441
complete system of triplets and a complete system of singlets.
The three lowest terms are (in wave numbers)
IS (singlet): 61,672.1
lp3 (triplet): 39,821.3
ls (triplet): 20,474.5
The basic term is thus seen to be the lS term, which, as we
explained in the case of sodium, is a 31 orbit. This is, therefore,
the orbit occupied by the twelfth electron.
The spectrum of ionized magnesium, Mgu, consists of a com-
plete system of doublets, similar to the sodium spectrum. The
three lowest terms are
lcr: 121,267
hr: 85,506
20: 49, 777
The basic term is, again, the 1cr (or the lS, if we disregard the
doublet nature of the spectrum). This spectrum is produced
by the binding of the eleventh electron, which, therefore, finally
reaches a 31 orbit.
From a study of the spectrum of Mg1 and of Mgrr, we conclude,
therefore, that the eleventh and the twelfth electrons of mag-
nesium revolve in 31 orbits.
Aluminum (Z = 13) yields an arc spectrum of doublets.
Here the basic term is l 1r:
l1r1: 48,168
1cr: 22,933
Just as the lowest 1r term in sodium (N a1 ) corresponds to a 3 2
orbit, so the lowest 1r term in Al1 corresponds to a 32 orbit. But
this orbit has a lower energy value (larger wave number) than
the 1cr term. We conclude, therefore, that the thirteenth
electron in Al revolves in the basic orbit 32. The lS orbit,
however, is the basic orbit in Alm, as is shown by the following
data from Bowen and Millikan. 1
Alm ls: = 229,455
lp1: = 175,536
This corresponds to a 31 orbit. Similarly, the basic term in
Alu is a ls term, and the basic orbit a 31 orbit. In aluminum,
1
Phys. Rev., vol. 25, p. 295 (1925).
442 ARRANGEMENT OF ELECTRONS [CHAP. XI
therefore, the eleventh and twelfth electrons occupy 3 1 orbits,
while the thirteenth electron occupies a 3 2 orbit.
Boron (Z = 5), in Period II of the periodic table, corresponds
to aluminum (Z = 13), in Period III. The thirteenth electron
of aluminum, which, as we have seen, occupies a 32 orbit, is
analogous to the fifth electron of boron. By analogy, we infer
that the fifth electron of boron occupies a 22 orbit. Similarly, we
may infer that the fourth electron of beryllium (analogous to the
twelfth of magnesium) occupies a 21 orbit. We may guess,
subject to check by spectroscopic data, that the last-added
electron in each of the remaining elements of Period II of the
periodic table, namely, C, N, 0, F, and Ne, occupies a 22 orbit,
since we have seen that the last-added electron, i.e., the eleventh,
in Na, the first element in Period III, occupies a 31 orbit.
The above example will serve to illustrate very briefly the
methods and some of the fundamental principles employed in
ascertaining the distribution of electrons in orbits.
6. Inner Quantum N umbers.-As was pointed out in Chap.
X, Sec. 13, the characterization of spectral terms by the two
quantum numbers n (total) and k (azimuthal) has been found
insufficient to account for the observed spectroscopic facts.
It was found necessary to introduce a third quantum number-
the inner quantum number j. In the Russell-Saunders notation
of spectroscopy (see Chap. X, Sec. 13), the azimuthal quantum
n urn her is irn plied in the (capital) letter used to designate the
term, and the inner quantum number is appended as a sub-
script, thus: P3 designates a term with azimuthal quantum num-
ber k equal to n (P = 2) and inner quantum number 3.
Irrespective of any atom model, the spectroscopic terms are
observed physical quantities. Until we attempt to interpret the
terms in the light of some atomic mechanism, we may, if we
choose, regard the various quantum numbers n, j, and k as
playing the part of a mnemonic notation to assist in classifying
and in correlating the observed data. When, however, we set
up an atom model and attempt to describe the observed phe-
nomena in terms of its properties, these quantum numbers must
take on a physical meaning. In terms of the Bohr atom model,
the total quantum number n and the azimuthal quantum
number k are definitely related to the orbits. The diameter of
the orbit-major axis in case of an elliptical orbit-is determined
by n. The shape, i.e., eccentricity for a given value of n, is
SEC. 6] INNER QUANTUM NUMBERS 443
determined by k. We may think of the inner quantum number
j as determining the orientation of an orbit with respect to the
orbits occupied by the other electrons of the atom. This is
equivalent to assuming that there is some "preferential" axis
associated with an atom about which its total angular momentum
may be measured and that the inner quantum number of an
orbit determines the inclination of the orbit to this axis. This
inclination is determined by the rules of quantization in much
the same way as we have quantized angular momentum.
We have seen that the angular momentum P<1> possessed by an
electron in its orbital motion around a nucleus is an integer multi-
ple k of the quantity h/21r, where k is the azimuthal quantum
number, i.e. (see equation (52), Chap. X),
h
P<1> = k 1r [ (52), Chap. X]
2
Since P<1> is a vector, we may represent it by a line of length k
(which is porportional to P<1>) drawn perpendicular to the plane
of the orbit. If there are 2 electrons in orbital motion around a
nucleus, the resultant angular momentum of the system is the
vector sum of the two individual momenta. If, for example,
these two orbits have the same value of k and are coplanar but
oppositely directed, this vector sum is zero; it is, however, 2k · h/21r
if the 2 electrons revolve in the same direction;itisv'2k · h/21rif
the orbits are inclined at 1r /2 to each other. In general, if there
are n electrons in orbital motion around a nucleus, the total angu-
lar momentum-call it P <!>-of the entire system is the vector sum
of the individual momenta P<1>· These n orbits mfJiY be inclined
to each other in various ways, so that, even though the individual
momenta P<1> do not change, P <!> may have different values for the
various possible configurations. According to the quantum
rules, however, only those values of P </> are permissible for which
. h
P <p = J · -
21r
where j is an integer (or a half-integer) and is the inner quantum
number.
Applying this concept to an orbit exterior to the core of an
atom, which, we will assume, has a fixed angular momentum, we,
in effect, quantize the inclination of the plane of the orbit with
respect to the axis about which the angular momentum of the
core is measured. Just as the angular momentum P<1> is restricted
444 ARRANGEMENT OF ELECTRONS {CHAP. XI
to a few values determined by the values of k, so the inclination
of the orbit to the axis of the core is restricted to a few angles
determined by the values of j. It is generally assumed that the
total angular momentum of the electrons which form a closed
group, such as the 2 electrons in the 21 orbits of lithium, is zero.
This means that the total angular momentum of the atoms of
the inert gases, which, presumably, are made up entirely of closed
groups, 1s zero.
The effect of the introduction of the inner quantum numbers,
which play a very important role in the quantum theory of the
Zeeman and the Stark effects, is to subdivide still further the
types of orbits normally occupied by electrons in atoms. In
what has been said above, we have considered only subdivisions
designated by the azimuthal quantum number k, that is, 3 2, 4 1
. . . or, in general, nk orbits. Orbits of a given nk are subdi-
vided, except when k = 1, into two groups by the use of the inner
quantu1:1 number j, which is added as a second subscript, thus: nki
and nki'. There are two kinds of 22 orbits-221 and 22 2-the
second subscripts 1 and 2 being the inner quantum numbers j and
j' of the orbits. Of orbits which have total quantum number 2
there are three kinds-211, 221, and 222. For total quantum num-
ber 3, there are five kinds-311, 321, 322, 332, 333. This subdivision
apparently finds confirmation in X-ray spectra.
Table VI shows the number and the designation of orbits for
various values of the total quantum number n. (In X-ray
terminology, orbits characterized by the total quantum num-
bers 1, 2, 3 . . . are designated by the respective letters, K, L,
M . . . These are included in the table for comparison and
for future use.)
TABLE VI
K L M N 0
I I
n I
1 2 3 4
I 5
k 1 122! 122331 12233441 1 2 2 3 3 4 .
I
J 1 !112!112231 11223341 1 1 2 2 3 3
I
It would take us too far afield to discuss the methods used in
subdividing the electrons of an nk group into the two groups nki
SEC. 7] COMPLETE DISTRIBUTION SCHEMES 445
and n1ci'· The results of such an analysis will be referred to in
the next section. Reference, however, may be made to the rule
that the total (maximum) number of electrons in a completed or
filled nki group is 2j. The maximum number of electrons which
can occupy the several orbits given in Table VI is found by
multiplying by two the values of j given in the bottom row of the
same table. The total number of electrons in complete groups
characterized by a quantum number n is the sum of the numbers
in each of the subgroups. Thus, the total number of electrons
required to complete the orbits for n = 2 is 2 + 2 + 4 = 8,
which is the number of electrons in the outer shell of neon, in
agreement with the conclusions from chemical data. For n = 3,
the number of electrons required is 2 + 2 + 4 + 4 + 6 = 18.
7. Complete Distribution Schemes for All Elements.-Guided,
in part, by spectroscopic data of the kind above mentioned and,
in part, by further generalizations based on theoretical studies,
complete orbital schemes for the various elements have been
worked out. For details, we must refer the reader to original
articles or to more extensive treatises. McLennan, Lay, and
Smith 1 give a set of "rules" on the basis of which they have
prepared a table showing the proposed distribution in orbits of
the electrons of each of the 92 elements. Among these rules, the
following are readily comprehended from what we have said
above:
"If the total maximum number of electrons that can occupy
an orbit type characterized by the same total and azimuthal
quantum numbers n and k, be N, the number N is given by N =
2(2k - 1)." For 11, 21, 31 . . . orbits, k = 1 and, therefore,
N = 2. This agrees with the conclusion mentioned above that
11 orbits of lithium and of all following elements contain 2
electrons. We saw, also, that there are 2 electrons in 21
orbits of beryllium and succeeding elements. Similarly, in
aluminum, the 31 orbits have 2 electrons. For 22, 32, 42 . . .
orbits, k = 2, and, hence, N = 6. We saw that with boron
(Z = 5) the fifth electron occupies a 22 orbit and that succeed-
ing elements of Period II have all but their first 4 electrons in
22 orbits. Neon (Z = 10) has, therefore, 6 electrons in 22 orbits.
For 33, 43 . . . orbits, k = 3 and N = 10.
"No two electrons can occupy identical orbits." There may
be, for example, a maximum of 6 electrons in 22 orbits, but
1 Proc. Roy. Soc. (London), A, vol. 112, p. 77 (1926).
446 ARRANGEMENT OF ELECTRONS [CHAP. XI
these six 22 orbits have different space orientations within the
atom.
There is, also, a "rule" that, in the process of capturing suc-
cessive electrons to make a neutral atom, electrons already
captured are not shifted into new orbits by succeeding captures.
A table has been prepared by P. D. Foote based on the work
of Bohr,1 Stoner,2 Main Smith,3 Pauli,4 Heisenberg,5 Sommer-
feld,6 Hund,7 and others, which shows the proposed distribution
of electrons in the various (nki) groups. This table is reproduced
as Appendix II.
At first glance, such tables as these which purport to show
the distribution of electrons in orbits may seem very speculative
and highly artificial. The fact, however, that the proposed
distribution receives confirmation from quite diverse fields in
physics and chemistry, and, more important, the usefulness of
these tables and the general underlying theories in predicting
new phenomena 8 are indicative of a fundamental "something,"
which-though our concrete picture of the atom may be, and
probably is, quite incorrect-is of far-reaching importance.
1 "Theory of Spectra and Atomic Constitution," Cambridge University
Press (1922).
2
Phil. Mag., vol. 48, p. 719 (1924).
3 "Chemistry and Atomic Structure," D. Van Nostrand Company (1924).
4 Zeit. fur Physik, vol. 31, p. 765 (1925).
5 Zeit. Jar Physik, vol. 32, p. 841 (1925).
6 "Atombau und Spektrallinien," Friedr. Vieweg & Sohn (1924).
1 Zeit. fur Physik, vol. 33, p. 335 (1925).
8 Witness the discovery of the element hafnium (Z = 72) by Coster and
Hevesy (Nature, vol. 111, p. 79 et seq, 1923) as a result of the prediction by
Bohr that this missing element should be analogous to zirconium, instead
of belonging to the rare-earth group.
CHAPTER XII
X-RAYS
There is probably no subject in all science which better illus-
trates the importance to the entire world of research in pure science,
than do X-rays. Within 3 months after Roentgen's fortuitous
discovery, X-rays were being put to practical use in a hospital
in Vienna in connection with surgical operations. The use of
this new aid to surgery soon spread rapidly. Since Roentgen's
time, X-rays have completely revolutionized certain phases of
medical practice. Had Roentgen deliberately set about to
discover some means of assisting surgeons in reducing fractures,
it is almost certain that he would never have been working with the
evacuated tubes, induction coils, and the like, which led to his famous
discovery.
In other fields of applied science, both biological and physical,
uses have been found for X-rays, which approximate in impor-
tance their use in medicine. One may mention, for example, the
study of the crystal structure of materials; "industrial diagnosis,"
such as the search for possible defects in the materials of engineer-
ing; the analysis of coal; the detection of artificial gems; the
study of old paintings; the study of the structure of rubber;
and many other uses.
But transcending these uses in applied science are the applica-
tions which are made of X-rays to such problems as the atomic
and the molecular structure of matter and the mechanism of
the interaction of radiation with matter. X-rays provide us
with a kind of supermicroscope, by means of which we can "see"
not only atoms and their arrangement in crystals but also even
the interior of the atom itself. Roentgen's discovery must be
ranked with the most important scientific discoveries of all time.
In this chapter, we shall give a brief account of the develop-
ment and the present status of X-rays, with particular reference
to their application to some of the fundamental problems of
physics.
1. Roentgen's Discovery.-In the autumn of 1895, Wilhelm
Konrad Roentgen, professor of physics at Wurzburg, was study-
447
448 X-RAYS [CHAP. XII
ing that fascinating phenomenon, the discharge of electricity
through rarefied gases. A large induction coil was connected
to a highly evacuated tube (Fig. 113) the cathode C being at
one end and the anode A at the side. The tube was covered
'' with a somewhat closely fitting mantle of thin black cardboard.'' 1
Wjth the apparatus in a completely darkened room, he made the
quite accidental observation that "a paper screen washed with
barium-platino-cyanide lights up brilliantly and fluoresceses
equally well whether the treated side or the other be turned
toward the discharge tube." 1 The fluorescence was observable
2 meters away from the apparatus. Roentgen soon convinced
himself that the agency which caused the fluorescence originated
at that point in the discharge tube where the glass walls were
struck by the cathode stream in the tube.
Toinducfion Co,!
+
Frn. 113.-Schematic representation of the tube with which Roentgen discoverPd
X-rays.
Realizing the importance of his discovery, Roentgen at once
proceeded to study the properties of these new rays-the unknown
nature of which he indicated by calling them "X-rays." In
his first communications, 2 he recorded, among others, the follow-
ing observations:
1. All substances are more or less transparent to X-rays. For
example, the fluorescent screen Hghted up when a bound volume
of 1,000 pages was held between it and the tube. Wood 2 to 3
cm. thick was very transparent. Aluminum 15 mm. thick
"weakens the effect considerably, though it does not entirely
destroy the fluorescence." Lead glass was quite opaque. Other
glass of the same thickness was much more transparent. "If
1 Quotations from papers by RoENTGEN: Electrician (London), vol. 36,
p. 415 (Jan. 24, 1896).
2 In addition to the paper mentioned above, see Roentgen's second com-
munication: Electrician, vol. 36, p. 850 (Apr. 24, 1896).
SEC. I] ROENTGEN'S DISCOVERY 449
the hand is held between the discharge tube and the screen the
dark shadow of the bones is visible within the slightly dark
shadow of the hand." The opacity of a substance depends not
only upon its density but also upon some other property. For
example, a sheet of platinum 0.018 mrn. thick had the same
opacity as 3.5 mm. of aluminum. The ratio of the thickness of
these two sheets is about 1: 200. The ratio of the densities of
the materials is about 8: 1.
2. Many other substances besides barium-platino-cyanide
fluoresce-phosphorus, calcium compounds, uranium glass, rock
salt, etc.
3. Photographic dry plates and films "show themselves sus-
ceptible to X-rays." Hence, photography provided a valuable
method of studying the effects of X-rays.
4. The rays are neither reflected nor refracted. Hence, "X-
rays cannot be concentrated by lenses."
5. X-rays, unlike cathode rays, are not deflected by a magnetic
field. They travel in straight lines, as Roentgen showed by
"pinhole" photographs.
6. X-rays discharge electrified bodies,1 whether the electri-
fication be positive or negative.
7. X-rays are generated when the cathode rays of the dis-
charge tube strike any solid body, such as aluminum or platinum.
A heavier element, such as platinum, however, is much more
efficient as a generator of X-rays than is a lighter element, such
as aluminum. Roentgen, in the second paper, describes
. . . a discharge apparatus in which a concave mirror of aluminum
acted as ca tho de and a sheet of platinum as anode, the platinum being
at an angle of 45 degrees to the axis of the mirror and at the center of
curvature.
This form of X-ray tube was the prototype of practically alJ
X-ray tubes until the introduction of the Coolidge tubes, about
1913.
In his first communication, Roentgen, arguing from the fact
that these new rays "behave in quite a different manner to any
infra-red, visible or ultra-violet rays hitherto known," asked the
1 This property of X-rays was independently discovered by Sir J. J. Thom-
son (see letter to Electrician (Feb. 4, 1896)). Thomson pointed out that this
phenomenon provides a method of studying X-rays much more delicate and
expeditious than the photographic plate or the fluorescent screen and, fur-
ther, that it yields quantitative measurements.
450 X-RAYS [CHAP. XII
question, "May not these new rays be due to longitudinal
vibrations in the ether?''
2. Some Early Experiments and Theories.-Roentgen's dis-
covery excited intense interest throughout the entire scientific
world. His experiments were repeated, and extended, in very
many laboratories in both America and Europe. The scientific
journals, during the year 1896, were filled with letters and short
articles describing new experiments or confirmations of observa-
tions previously made. 1 These experiments were facilitated by
the fact that during the years immediately preceding Roentge:.~'s
discovery the discharge of electricity through rarefied gases had
been a very popular topic for study by physicists. Equipment
similar to Roentgen's was, therefore, in operation in many
research laboratories. Further, considerable impetus was given
to research in this field because of its important practical applica-
tions in surgery and elsewhere.
As was to be expected, many of these early experiments
brought forth conflicting evidence, and most of them were the
result of gropings in the dark in the absence of any guiding
theory. Nevertheless, particularly from a historical standpoint,
these early experiments are not without interest. By way of
illustration, we shall summarize a few of them-making no
attempt, however, to give a complete survey.
(a) Sources of X-rays.-Tubes for the production of X-rays
very soon became more or less standardized along lines suggested
by Roentgen, as mentioned in the previous section. A tube
similar to Roentgen's, in which the cathode rays from a curved
cathode were brought to a focus onto a piece of platinum inclined
at 30 to 40 degrees to the direction of the rays, was described by
S. P. Thompson. 2 Figure 114(a) shows, schematically, such a
tube. C is the curved cathode, the cathode rays from which
converge onto, or are "focused" onto, the anode A, which is a
sheet of metal such as platinum, inclined at 45 degrees to the
axis of the cathode stream. The potential is supplied to the tube
by an induction coil, electrostatic machine, or other source of
sufficiently high potential. In this type of tube, known as the
"gas" tube, it is essential that the gas pressure within the tube
1 See, for example, Science, vol. 3 (1896), which contains translations of
Roentgen's papers and numerous short articles with X-ray photographs.
The Beiblatter zu der Ann. Physik for 1896 contains 400 titles on X-rays!
2 Compt. rend., vol. 122, p. 807 (1896).
SEC. 21 EARLY EXPERIMENTS AND THEORIES 451
be maintained at the desired value. Various ingenious devices
were introduced for accomplishing this. These gas tubes were
practically the only source of X-rays up to the introduction of
the Coolidge tube.
The Coolidge tube,1 introduced by Dr. W. D. Coolidge, in
1913, differs from the gas tube in two very important particulars:
(1) In the gas tube, the electrons of the cathode stream are
supplied as a result of the ionization of the residual gas in the
tube. In the Coolidge tube (Fig. 114(b)), the electrons are
supplied by the thermionic emission from a flat, spirally wound
tungsten filament F heated by a small battery B or by a low-
+
c _____
(a)
'I' F
+ -
---- .. · --- - -- --- -High Vo!fctge r5ource ·--- .. - - -- ·- • ... -
(b)
j l•I•
B
. rrn. 114.-(a) An early form of X-ray tube; (b) the Coolidge X-ray tube.
voltage transformer. (2) It is, therefore, possible to dispense
entirely with the gas in the tube-the Coolidge tube is evacu-
ated to the highest a.ttainable vacuum. One very important
advantage of the Coolidge X-ray tube over the gas tube is the
possibility of controlling the current through the tube inde-
pendently of the voltage applied to the tube. In the Coolidge
tube, the current can be adjusted to any desired value, irrespec-
tive, within limits, of the applied voltage. In the gas tube,
current, applied voltage, and gas pressure are more or less inter-
dependent. The controllable features of the Coolidge tube have
greatly facilitated research in X-rays, particularly in those
1 Phys. Rev., vol. 2, p. 409 (1913).
452 X-RAYS [CHAP. XII
investigations which require high precision in the measurement
of the energy in an X-ray beam.
(b) Improvements in Photographic Methods.-The use of X-ray
photography in surgical work made it desirable to shorten expo-
sures as much as possible. The application of fluorescent-or,
as we now say, intensifying-screens to increase the effective
speed of photographic emulsions was independently made by
numerous investigators. That the intensity of the photographic
image was greatly increased by painting the glass side of the
photographic plate with phosphorescent zinc sulphide was dis-
covered by C. Henry. 1 Basilewski 2 modified this procedure by
coating a sheet of paper with the fluorescent material and laying
the sheet face down on the photographic emulsion. Buguet 3
used a sheet of lead instead of fluorescing material. Edison 4
discovered that calcium tungstate properly crystallized fluoresces
brilliantly. This was later used for intensifying screens. Elihu
Thompson 5 made the very important discovery that two X-ray
photographs of an object taken at slightly different directions
and viewed by aid of a stereoscope would stand out in high
relief.
(c) Ionization Measurernents.-Use was at once made of the
discovery independently announced by Thomson and by Roent-
gen, that X-rays discharge charged bodies and that the rate of
discharge is a measure of the intensity of the rays. Perrin 6
showed that this discharging action of the rays is apparently due
to the fact that the rays make the air surrounding the charged
body conducting. It was suggested that the air molecules are
broken up by the rays into positively and negatively charged
ions. This suggestion was in harmony with the observation of
Roentgen 7 that under the action of X-rays charged bodies placed
in a vacuum lose their charge very much more slowly than they
do in air. Further, Villari 8 showed that "Roentgenized" air
preserves its property of discharging an electroscope even after
passing through a tube 10 meters long. He also showed that the
1 Compt. rend., vol. 122, p. 312 (1896).
2 Compt. rend., vol. 122, p. 720 (1896).
3 Compt. rend., vol. 122, p. 702 (1896).
4 Electrician, vol. 36, p. 702 (1896).
5 Electrician, vol. 36, p. 661 (1896).
6 Compt. rend., vol. 122, p. 186 (1896).
7 Electrician, vol. 36, p. 850 (1896).
8 Compt. rend., vol. 123, pp. 418, 446 (1896).
SEC. 2] EARLY EXPERIMENTS AND THEORIES 453
discharging action in a gas at a given pressure depends on the
nature of the gas. The following were found increasingly active
in the order given: H, CO, air, C02, ether vapor, CS2. Benoist
and Hurmuzescu 1 showed that, for a given gas, the discharging
action increases rapidly with density.
This property of X-rays, in making a gas through which they
pass electrically conducting, has been used from the very first
as a quantitative means of measuring the intensity of an X-ray
beam. At first, the rate of discharge of a charged electroscope
was used, the motion of the leaves being observed by a low-power
microscope. Later, an auxiliary device, known as an "ioniza-
tion chamber," was introduced. This is shown schematically in
Fig. 115. C is a metal tube several centimeters in diameter,
from 20 to 100 cm. long, and closed at both ends except for an
k
G
X-Rays w\ r
c
Frn. 115.-The ionization chamber used for measuring electrically the intensity
of a beam of X-rays.
opening or "window" W, over which may be placed a thin sheet
of aluminum for admitting the X-rays. A rod rr suitably
supported by good insulating material, such as amber or quartz,
is connected to an electrometer. An electric field is maintained
between the rod rr and the cylinder C by a battery B, of several
hundred volts, one end of which is connected to the ground wire
G. An earthed guard ring g prevents leakage from the cylinder
to the rod rr. In accordance with the observations of Villari, the
cylinder may be filled with a heavy gas to make the arrangement
more sensitive.
When X-rays enter the window W, the gas within the cylinder
is made conducting, and on account of the electric field between
the cylinder and the rod the latter acquires a charge at a rate
which can be measured by the electrometer, which rate is deter-
mined by, and is, therefore, a measure of, the intensity of the
X-ray beam.
1
Compt. rend., vol. 122, p. 926 (1896).
454 X-RAYS [CHAP. XII
(d) The Absorption of X-rays.-Roentgen's initial observation,
that the relative opacity of materials to X-rays depends not alone
upon density, was early confirmed. Bleunard and Lablesse, 1
Novak and Sule,2 Benoist,3 and Marangoni 4 all came, independ-
ently, to the conclusion that the opacity of a material to X-rays
depends upon its atomic weight. Buguet 5 showed that the beam
of rays from an X-ray tube is not homogeneous, since the opacity
of a given material to the rays depends upon the thickness of
material previously traversed by the beam. This showed that
the quality of a beam of X-rays could be altered by varying the
thickness of the material through which the beam is made to pass.
For many years, the absorbability of X-rays
r-x- in some standard material such as aluminum
p was the only measure of the quality of the beam.
:: Rays were called "hard rays" or "soft rays,"
Io ·1 I
_ _...,...~I~
1
according as they more readily passed through,
il or were more readily absorbed by, aluminum.
11 We shall have frequent occasion to make
1
,,-~ - quantitative referenc(, to the absorption of
dx X-rays. It will be desirable, therefore, to derive
FIG. 116. the general formula for the absorption of any
type of radiation as it passes through matter.
Let a beam of radiation of intensity Io be normally incident onto
a slab (Fig. 116) of absorbing material of thickness x; and let I be
the intensity of the emergent beam. Let i be the intensity of
the beam after it has penetrated into the slab as far as the thin
layer of thickness dx; and let di be the diminution in intensity
in passing through this thin layer dx. Then, so long as di/i is
very small,
di
-;- = -µdx (1)
i
whereµ is a proportionality factor called the coefficient of absorp-
tion. Integrating equation (1) from one face of the slab to the
other, we have
1d·
i
_; = -
I
µdx
i
ix
O
o I
:. log Io = -µx
1 Compt. rend., vol. 122, p. 723 (1896).
2 Zeit. filr Phys. Chem., vol. 19, p. 489 (1896).
3 Compt. rend., vol. 122, p. 146 (1896).
4 Rend. Acc. Linc., vol. 5, p. 403 (1896).
5 Compt. rend., vol. 125, p. 398 (1897).
SEC. 2] EARLY EXPERIMENTS AND THEORIES 455
or
(2)
By measuring I, Io, and x, one may computeµ. If we are meas-
uring the coefficient of absorption of X-rays, Io and I may be
measured by the ionization chamber mentioned above.
In deriving equation (2), we are not concerned with what ulti-
mately becomes of the radiation which is removed from the inci-
dent beam. We are concerned only with the fact that as a result
of the passage through the slab of material the intensity of the
radiation has been reduced from Io to I.
In discussions of the absorption of X-rays, a quantity known
as the mass absorption coefficient is frequently used. Let the
exponent in equation (2) be both multiplied and divided by p,
the density of the slab. We then have
_#!..p:r,
I = Iae P (3)
The quantity µ/ p is called the mass absorption coefficient; px is,
obviously, the mass of a portion of the slab 1 sq. cm. in cross-
section and x cm. thick.
(e) Scattering.-That X-rays are scattered or diffused, as is
light by fog particles, was early recognized. In an attempt to dis-
S
A Vll lZl 222 222 22222 22 22 2222 22 222 221A M
FIG. 117.-The arrangement of Imbert and Bertin-Sans showing "pseudo"
reflection (scattering) from the mirror M.
cover the reflection of X-rays, Imbert and Bertin-Sans 1 arranged
apparatus, as shown diagrammatically in Fig. 117. Between the
source S of the rays and the photographic plate P was placed a
thick copper screen AA. A plane mirror M was so positioned
that a beam of rays, if reflected, would pass to the photographic
plate P, on which would appear an image or shadow of an obstacle
B. Such a shadow was obtained irrespective of the angular posi-
1 Compt. rend., vol. 122, p. 524 (1896).
456 X-RAYS [CHAP. XII
tion of the mirror M. Indeed, a plate of paraffin was just as effec-
tive as the mirror M, from which fact Imbert and Bertin-Sans
concluded that the rays were diffused or scattered from M, rather
than reflected, and that probably the rays were waves of very small
wave length, too small to be reflected from (artificially) polished
surfaces.
Righi 1 showed that an electroscope E, although placed in the
''shadow'' of a lead screen Pb (Fig. 118) would be discharged by
the action of X-rays emitted by the source S. This effect was
later shown to be due to the scattering effect of the air in the
neighborhood of the screen.
An important observation was made by Winkelmann and
Straubel. 2 A beam of X-rays was passed through a photographic
plate P (Fig. 119), the emulsion being on the rear side. Behind
A
I
pl
I
I
Ph X-Raf/6' • :[r
E I
s I
* A A
I
I
I
Frn. 118. rm. 119.
a part of the plate was placed a piece of fluorspar F. On develop-
ing the plate, it was found that the film was much denser in the
neighborhood of F, as if F had reflected the rays. An observa-
tion of this kind had been made by Roentgen. But Winkelmann
and Straubel showed that the phenomenon was not one of true
reflection, for they repeated the experiment with a thin sheet of
paper AA between F and P and found that the intensifying action
of F was almost entirely destroyed, although the paper was very
transparent to the incident beam. From this, they concluded
that the quality of the rays had been altered by the spar in such a
way as to make the beam of rays returned by F more absorbable
in paper than was the original beam. In other words, the
primary rays, incident onto the spar, had been transformed into
characteristic "spar" rays.
1 Rend. Acc. Linc. (1896).
2 Zeit. fur Naturw., vol. 30 (1896).
I
SEC. 2] EARLY EXPERIMENTS AND THEORIES 457
(f) Refraction and Dijfraction.-Roentgen, in his first papers,
reported that he had failed to find either refraction or diffraction
of X-rays. Because of the importance of this observation, more
diligent search for evidences of these phenomena was under-
taken. Gouy 1 had observed that X-rays emitted in different
angular directions from a plane platinum target were of nearly
the same intensity down almost to grazing incidence. By taking
rays, from a plane target, at nearly grazing incidence, a very
narrow, and quite intense, line source of X-rays is thus available.
Such a line source was used to get a shadow of a fine platinum
wire on a photographic plate, a prism being placed in the path of
part of the rays in such a way that if the rays were refracted by the
prism, the shadow of the wire would be ''broken.'' The shadow
was sharp and continuous, showing no measurable refraction.
Gouy estimated that the index of refraction could .not be greater
than 1.000005.
Gouy 2 and, independently, Forum 3 searched for evidences of
diffraction, with negative results. They concluded that if
X-rays are waves, subject to the same laws as light, their wave
length cannot be greater than about 10 X 10-7 cm.
(g) Early Theories.-The experimental conclusions from these
early experiments were, substantially, as follows:
1. X-rays are produced when the cathode stream in a highly
evacuated discharge tube falls upon a solid obstacle.
2. These rays cause fluorescence, affect a photographic plate,
and ionize air-just as does ultra-violet light.
3. But, unlike ultra-violet light, X-rays are neither reflected,
refracted, nor diffracted; and they readily penetrate
matter opaque to ultra-violet light.
4. Air and other gases, however, as well as solid bodies,
scatter X-rays, much after the manner of the action of
fog particles upon light.
What were these new rays?
Roentgen's early suggestion that perhaps X-rays might be
longitudinal ether waves was not well received. The weight of
opinion inclined to the view that X-rays are very short ultra-
violet waves, since X-rays exhibit so many of the properties of
1
Compt. rend, vol. 122, p. 1197 (1896).
2
Compt. rend., vol. 123, p. 43 (1896).
3 Ann. Physik, vol. 59, p. 350 (1896).
458 X-RAYS lCHAP. XII
ultra-violet light. It was pointed out by Maltezos 1 that the
Helmholtz dispersion formula
n2 = 1 + GX + 2
where n is the index of refraction and C is a constant, predicts
that n ...:. 1 as X ...:. O, which is in agreement with the assumption
that X-rays are very short waves, of an order of magnitude
probably not exceeding 10-7 cm. This same theory would
explain the absence of reflection and diffraction.
A little later, Haga and Wind 2 carried out a series of experi-
ments in an attempt to detect the diffraction of X-rays. With
wedge-shaped slits only a few thousandths of a millimeter wide,
they observed evidences of a slight widening of the image on the
photographic plate, from which they deduced that the probable
wave length of the rays must be of the order of 10-s cm. This
result-although confirmed qualitatively by experiments over
2 decades later-was not given much weight, partly because of
the smallness of the effect, partly because, in the meantime, the
classical "ether pulse" theory of X-rays had been developed by
Schuster, 3 Wiechert, 4 Stokes, 5 Sir J. J. Thomson, 6 and others.
We shall discuss this theory briefly in the next section.
3. The Ether-pulse Theory of X-rays. (a) The Production
of X-rays.--Theories concerning the mechanism of the cathode-
ray discharge were beginning to take definite form as a result
of experiments in this field covering a period of many years.
These experiments culminated in the discovery of the electron, by
Sir J. J. Thomson, in 1897. It then became clear that the cathode
stream in a discharge tube consists of swiftly moving electrons.
If these electrons, moving with high velocity and carrying, as
they do, a negative charge, be suddenly brought to rest by col-
liding with a solid obstacle-the glass walls of the discharge
tube or a piece of metal deliberately put in the path of the rays-
they must experience, for a brief interval, a very large negative
acceleration. According to the electromagnetic theory, such an
accelerated body must radiate energy. The sudden stopping
1 Compt. rend., vol. 122, pp. 1115, 1474, 1533 (1896).
2 Ann. Physik, vol. 68, p. 884 (1899).
3 Nature, p. 268 (Jan. 23, 1896).
4 Ann. Physik, vol. 59, p. 321 (1896).
5 STOKES: The Wilde Lecture (1897).
6 Phil. Mag., vol. 45, p. 172 (1898).
SEC. 3J THE ETHER-PULSE THEORY 459
of each electron as it collides with the target of the X-ray tube
results in the emission of an electromagnetic disturbance or pulse
frequently called an "ether pulse." On this theory, the X-rays
emitted by a target consist of a very rapid succession of such
pulses, coming at random intervals. It was shown by Stoney 1
that such a stream of pulses could be analyzed into wave trains,
the components of shorter wave length being the more abundant
the greater the velocity of the electrons which are brought to
rest by the target. Stoney showed that matter should, in
general, be more transparent to the shorter waves than to the
longer waves. The hard, or penetrating, X-rays should, there-
fore, be produced by high-velocity electrons, which, in turn, are
produced by high voltages applied to the X-ray tube. Qualita-
tively, this picture of the mechanism of the production of X-rays
seemed to b~ in agreement with the experimental facts.
In Chap. V, we discussed some fundamental principles con-
cerning accelerated charges, which are directly applicable to the
ether-pulse theory of X-rays. It was shown that when a charge
e moving with velocity vis brought to rest in such a time that a
pulse of thickness o is emitted, the total energy W radiated is
given by (see equation (34), Chap. V).
[ (34), Chap. V]
This energy W which is radiated by the stopping of the electron
was shown to be a fraction
2a
0
of the total kinetic energy of the moving charge (see equation
(37), Chap. V), where a is the diameter of the electron. The
fraction of its energy which is radiated when the electron is
stopped is, therefore, inversely proportional to o, the thickness
of the pulse. Now, the thickness of the pulse is directly pro-
portional to the time required to stop the electron (see equation
(25), Chap. V). And it is reasonable to assume that an electron
moving with given velocity will be stopped more suddenly when
colliding with a heavy atom than with a lighter one. Other
things being equal, therefore, a more intense X-ray beam should
be produced when a heavy material such as platinum is used as
the target in an X-ray tube than when a lighter metal such as
1 Phil. Mag., vol. 45, p. 532; vol. 46, p. 253 (1898).
460 X-RAYS [CHAP. XII
aluminum is employed. This is in agreement with Roentgen's
observation, noted above. In other words, heavy targets are
more efficient sources of X-rays than are lighter ones.
(b) The Scattering of X-rays.-Mention was made above
(see Sec. 2(e)) of the experiments of Imbert and Bertin-Sans, who
showed that X-rays are scattered when they pass through matter.
This phenomenon of the scattering of X-rays has played a very
important part in the theories of modern physics and has been
the object of many researches, both theoretical and experi-
mental. The ether-pulse theory of X-rays, together with the
concept of the electrical nature of matter, which was firmly
established by the discovery of the electron, formed the basis
of a theory of sea ttering proposed by Sir J. J. Thomson. 1
Let us consider that a swiftly moving electron in an X-ray
tube is suddenly brought to rest by colliding with a target at
d'=c1'
F
D!recf!on of'
O
morion of' Pulse
Frn. 120.~The force acting on an electron during the passage of an ether pulse.
point O (Fig. 120). As a result of this collision, a. spherical pulse
of thickness o = er, where r is the time required to stop the
electron, will spread out radially from 0. This pulse, as was
shown in Chap. V, consists of an electric vector and a magnetic
vector at right angles both to each other and to the direction of
propagation. Let us consider a point, in the x direction from
0, at which is located an electron of charge -e and mass m.
As the pulse passes over the electron, the latter will experience a
force F given by
F = Y·e
where Y is the electric vector in the pulse. For simplicity, let
us consider that Y is constant throughout the thickness of the
pulse. The electron will then be given a constant acceleration g
given by
Ye
g=- (4)
m
1 "Conduction of Electricity through Gases/' 2d. ed., p. 321.
SEC. 3] THE ETHER-PULSE THEORY 461
which acceleration will last for a time r required for the pulse to
pass. As a result of this acceleration, the electron will radiate
energy at a rate given by equation (36) of Chap. V, namely,
dW 2 e2 g 2
dt - 3 ---
c3
[ (36), Chap. V]
Putting into this equation the value of g from equation (4),
we have
2 e4 y2
dW = - --3 - dt
3 c m2
The total energy W radiated by the electron as a result of the
passage of the pulse over it will be
(5)
(6)
if, as mentioned above, we consider Y constant. 1
This energy TV is radiated as a secondary pulse as a result of
the passage over the electron of the primary pulse. The electron
has, therefore, scattered some of the energy of the primary pulse.
We shall consider presently the direction in which this energy is
scattered.
1 This limitation of Y to a constant value throughout the pulse is quite
• unnecessary. In equation (5), we are concerned with the integral J:r Y dt.
2
Since the energy per cubic centimeter from point to point in the pulse is
proportional to Y 2, the intensity of the pulse I is given by
I = _!_ f 0 Y2do
41rJo
where do is a thin slice of the pulse taken parallel to the surfaces of the
pulse. Since o = er and, therefore, do = cdt, it readily follows that
I = ~ frY2dt
41rJo
We then have for W, from equation (5),
W :! S1r · ___!!!___ I
3 m 2c 4
from which equation (9) may be readily deduced.
462 X-RAYS [CHAP. XII
Let us consider that this scattering electron is simply one of a
large number of electrons in a slab of material of thickness dx
(Fig. 121) through which the primary pulse passes. And let
us assume that every electron in the slab scat-
··d.x ters exactly the same amount of energy W. Let
I n be the total number of electrons per cubic cen-
timeter in the slab. Then, ndx is the number of
electrons in a portion of the slab 1 sq. cm. in
cross-section and dx cm. thick. Each of these
FIG. 121.
ndx electrons removes from the primary beam
by the process of scattering an amount of energy W. The
total amount of energy dl removed from the beam by 1 sq. cm.
of the slab is, obviously,
2 e4
dl = W · ndx = - -2-3 n Y 2rdx (7)
3 m c
The energy per cubic centimeter in the pulse is given by
y2
41r
and the intensity I of the pulse is
y2 y2
l=-o=-cr (8)
41r 41r
where o = er is the thickness of the pulse. The fractional
diminution of energy resulting from the scattering process when
the pulse passes through the slab is, therefore, given by
dl = - [ 1r · ~
8
n]dx (9)
I 3 m 2c4
By analogy with equation (2) for the absorption coefficient µ,
we may write equation (9) in the form
dJ = -<r. dx
where <J' is the scattering coefficient and is given by
81r e4
u=--- n (10)
3 m 2c 4
As a result of the scattering process, therefore, an incident beam
of X-rays of intensity I O should, in passing through a slab of
material of thickness x containing n (scattering) electrons per
cubic centimeter, be reduced in intensity to I according to the
equation
I = Ioe-ux (11)
SEC. 3] THE ETHER-PULSE THEORY 463
If, now, we assume that it is only by the scattering process
that energy is removed from the incident beam, we can measure
u by measuring I 0 , I, and x. In equation (10) for u all quantities
are known except n, the number of scattering electrons per
cubic centimeter. By measuring a:, therefore, we may compute n.
Barkla 1 was the first to determine, by this method, the number
of electrons per cubic centimeter of a scattering material and,
from this number, the number of electrons per atom. Barkla
found from observations of scattering that for carbon the value
of a/ p, the mass scattering coefficient, is approximately 0.2. In
the equation
u 87r e 4 n
p = 3 m 2c 4 p (12)
we may set <1/p = 0.2 from Barkla's observation; e = 4.77 X
10- 10 ; m = 9 X 10- 28 ; c = 3 X 10 10 • We thus find that n/p,
the number of scattering electrons per gram of carbon, is about
3 X 10 23 • The number of atoms of carbon per gram is given by
1= 6.06 ~ lQ23 = 5.05 x 1022
where N is Avogadro's number and A is the atomic weight. The
number of electrons per atom of carbon is, therefore,
n/ p _ 3 · 10 23 •
NI A - . . 1022 = 6 (approximately)
5 05
(13)
This was one of the first determinations of the number of electrons
per atom, a result which was reached by a straightforward appli-
cation of classical theory.
(c) Angular Distribution of Scattered X-rays.-We saw, in
Chap. V (equation (30)), that the distribution of intensity over
the surface of a pulse emitted when an electron is accelerated is
not uniform but is proportional to sin 2 e, where 8 is the angle
between the direction of the acceleration and the direction in
which the intensity is measured; i.e., the intensity 1 0 at the angle
() is given by (see equation (30), Chap. V),
- 1 e2v2 . 2
Io - 47T' c2R20 sm () [ (30), Chap. V]
1 Phil. Mag., vol. 28, p. 648 (1911).
464 X-RAYS [CHAP. XII
where v is the initial velocity of the charge e which is brought to
rest, ois the thickness of the resulting pulse, and R is the distance
of the point of observation from the center of the pulse (see Fig.
24). Taking v = gr, where g is the uniform acceleration which
brings the charge to rest in time r, and o = er, we can transform
equation (30), Chan. V, into
1 e2g2 •
I e = 47r c 3R 2 T sm 2 () (14)
which gives 10 in terms of the acceleration g.
y p Now, let a plane-polarized primary
pulse of intensity I and thickness o = er
proceeding in the x direction pass
over an electron at O (Fig. 122), the
electric vector in the pulse being Y.
- - ~ ~------.:» The acceleration Oe experienced by the
0
electron will be
Ye
FIG. 122. Oe = -
m
The intensity 1; of the secondary pul~e in a direction e = LPOY
and at a distance R from O will be, according to equation (14),
' - 1 e2 y2e2 . 2
Io - - 3R 2 • - -2 T sm ()
(15)
41T' c m
Since er = o, and since the intensity of the primary pulse I is
I = _!_ Y2 0 (16)
411"
(assuming Y constant throughout the pulse), we have, from
equation (15),
I '0 = I · m2c4 1m
-e4- -R2
s . 2 () (17)
That is to say, the intensity I' of the secondary pulse is propor-
tional to the intensity of the primary pulse and to sin2 8. The
intensity I' is greatest in the xz plane; it is zero in they direction.
Let us now consider the more general case where the X-ray
beam consists not of a single plane-polarized pulse but of a rapid
succession of such pulses, the electric vectors of which are dis-
tributed at random in the yz plane so that the time average of the
components of the electric vector in they direction is Y, and in the
z direction is Z. Since the distribution is random
Y=Z
SEC. 3] THE ETHER-PULSE THEORY 465
Applying an equation of the type of equation (17) to determine
the intensity of the scattered beam at some point Pin the xy plane
(Fig. 123), due to each component Y and Z separately, we have
p , - I e4 1 . 2
y I e,y- y·m2c4R2s1n ()
' e4 1
I m2c4 R2 cos cp
2 (18a)
=
~
Y • -- -
''
0 I 'o,z -- I z • m2c4
e4 1 . 2
R2 sm
7r
2
z e4 1
= lz· - -R2
m2c4 - (18b)
Frn. 123.
since LPOz = 1r /2. Remembering that the intensity I of the
primary beam is given by
I= Jy + lz
and that the intensity of the secondary beam is
I 0' = '
Ie,Y + Io,z
'
we have for I~, by addition of equations (18),
I~ = I
me
e24 4. R\(1 + cos2 cp) (19)
where cp is the angle which the direction OP makes with the direc-
tion of propagation of the x . . ray beam. According to equation
(19), the intensity of the scattered X-ray beam should be
symmetrical about the direction of propagation of the primary
beam and, also, forward and backward, about a plane (the yz
plane in the figure) passing through the scatterer and perpendi-
cular to the direction of propagation. We shall see, later, that
there are found, experimentally, radical departures from this
second condition of symmetry. In a later section, we shall return
to the question of the scattering of X-rays.
(d) Polarization of X-rays.-Measurements of the scattering
of X-rays yield information concerning polarization. Let a
stream of cathode rays, proceeding in the direction zO (Fig. 124)
impinge on a target at 0. Consider the beam of X-rays proceed-
ing from this target in the direction Ox1. We shall call this the
"primary beam." If we assume that the electrons of the cathode
stream are all brought to rest by accelerations in the direction Oz,
this primary beam should be plane polarized with the electric
vector in the z direction, as indicated by the vector Z. Let a
466 X-RAYS [CHAP. XII
scattering material-a small piece of carbon, say-be located at
P1. As explained in the preceding section, the electrons of this
scatterer should experience accelerations in the direction P 1z 1 as a
result of the passage of the primary beam, and the scatterer
should emit a secondary or scattered beam. The intensity of this
scattered beam should be a maximum in the plane yOx1, and
should be zero in the direction z1z:. If this analysis be correct,
the intensity of the X-rays scattered by a substance at P1 should
vary from zero in a direction P1z1 to a maximum in the direction
P1P2.
Experiments testing this conclusion were performed by
Barkla. 1 He found that the intensity of the scattered radiation
/
/
/
/
Z.1
z
]'IG. 124.
in the direction of P 1z1, while not zero, was considerably less than
the intensity in the direction P1P2. This indicates that the
primary beam is at least partially polarized, though not com-
pletely polarized. A little further consideration indicates that
this is just what we should expect. We assumed, at the begin-
ning of the above discussion, that the electrons of the cathode
stream are brought to rest by accelerations directed in the direc-
tion Oz. If we attempt to picture the sequence of events by
means of which a swiftly moving electron is brought to rest by
collision with the atoms of the target at 0, we should conclude
that rarely will an electron be stopped by a single "head on"
collision with an atom. In the general case, an electron will
1 Proc. Roy. Soc. (London), vol. 77, p. 247 (1906).
SEtC. 4] SECONDARY RADIATION 467
pursue a zigzag course and will collide with many atoms before
being brought to rest. While the preponderance of accelerations
may be in the general direction Oz, accelerations in quite different
directions are to be expected. In the primary beam, therefore,
the Z components of the electr:c vector should predominate, but
Y components are also to be expected. This means that the
intensity of the X-rays scattered from P1 in the direction of P 1P 2
should be greater than in the direction P1z1-which is what.
Barkla observed .
. With the secondary beam of X-rays proceeding in the direction
P1P2, however, the situation is different. This secondary beam
is produced by the acceleration of electrons at P1 due to the
passage of the primary beam. On account of the transverse
nature of electromagnetic radiation, this primary beam, regard-
less of its state of polarization, can accelerate electrons at P 1
only in directions lying in the plane defined by P 2P 1z1, i.e., in
directions at right angles to P1x1. Consequently, the electric
vector of the secondary beam proceeding in the direction P 1P 2
must lie entirely in the plane defined by P?1z1; i.e., this secondary
beam must be completely plane polarized. If, then, this second-
ary beam be allowed to pass over a second scatterer placed at
P2, the intensity of the tertiary radiation sent out from P 2 should
vary from zero in the direction P2z2 to a maximum in the direc-
tion P 2 x 2 • Compton and Hagenow 1 have shown that, after
eliminating various errors, the intensity of the tertiary radiation
seattered in the direction P 2Z2 is zero within the limit of error of
measurement, and, therefore, as electromagnetic theory predicts,
the secondary beam is completely polarized.
The ether-pulse theory was, thus, very successful in explaining
many of the observed phenomena of X-rays; indeed, in common
with other phases of classical theory, the ether-pulse theory is
still an important factor. But there was gradually accumulated
a mass of data which the ether-pulse theory could not explain.
Again, as in other branches of physics, the quantum theory
came, more or less successfully, to the rescue. We shall return
to this point again after we have discussed the work of Laue,
Friedrjch and Knipping, Bragg, Moseley, and others, which led
to a better understanding of the origin and nature of X-rays.
4. Characteristic Secondary Radiation.-It will not serve our
purpose to go into a detailed discussion of the development of
1 Jour. Optical Soc. Amer. and Rev. Sci. Inst., vol. 8, p. 487 (1924).
468 X-RAYS [CHAP. XII
the subject of X-rays from the introduction of the ether-pulse
theory, in the closing years of the njneteenth century, to the
discovery by Laue and Friedrich and Knipping, in 1912, of the
action of a crystal grating in analyzing a beam of X-rays into
its component wave lengths. But no discussion of X-rays,
however brief, would be complete without at least reference to
the classical work of Barkla,1 in the decade preceding 1912,
which gave a deeper insight into X-ray phenomena and, in large
measure, prepared the way for the very remarkable series of
advances which began in 1912.
Mention was made, on page 456, of the observation of Winkel-
mann and Straubel that a beam of X-rays incident onto a piece
of Iceland spar was transformed into characteristic ''spar''
rays the quality of which, as measured by absorption in paper,
was quite different from that of the incident beam. Barkla and
his collaborators made a thorough and systematic study of this
phenomenon of the emission of characteristic secondary radia-
tion by a substance when a primary beam is incident onto the
substance.
We may illustrate the fundamental principles involved by
describing a simple but instructive experiment with visible light.
Prepare a finely divided suspension of some such material as
gum mastic in water. 2 Place the suspension in a rectangular
glass box C (Fig. 125) and illuminate it by a beam of white light,
as shown. The suspension will appear white when viewed from
some such position as point O; that is to say, the suspension emits
1 Among the papers by Barkla and his collaborators may be mentioned
the following: BARKLA and SADLER: "Homogeneous Secondary Roentgen
Radiation," Phil. Mag., vol. 16, p. 550 (1908); BARKLA and SADLER: "The
Absorption of Roentgen Rays," Nature, vol. 80, p. 37 (1909); BARKLA:
"Phenomena of Roentgen-ray Transmission," Proc. Cambridge Phil. Soc.,
vol. 15, p. 257 (1909); BARKLA and N1coL: "X-ray Spectra," Nature,
vol. 84, p. 139 (1910); BARKLA: "Spectra of Fluorescent Roentgen Radia-
tions," Phil. Mag., vol. 22, p. 396 (1911); Barkla: "Absorption of X-rays and
Fluorescent X-ray Spectra," Phil. Mag., vol. 23, p. 987 (1912).
The student is urged to read these papers, not only because of their very
great historical importance but also because they constitute a background
on the basis of which subsequent advances may be better understood
and appreciated.
2 A suitable suspension is readily prepared by dissolving a small quantity
of mastic in alcohol and then pouring a few drops of this solution into a
booker of water.
SEC. 4] SECONDARY RADIATION 469
or scatters in the direction of O a radiation of the same qualit.y 1 as
is incident onto it. Consequently, a piece of absorbing glass,
say blue, will produce the same reduction in the apparent bright-
ness of the box as viewed from point 0, whether the glass be put
in the primary beam at A or in the secondary beam at B. Con-
versely, if an absorbing screen is observed to produce the same
reduction in the brightness of the box whether the screen be
placed at A or at B, we may conclude that the secondary radia-
A
tion emitted by the material in the box
![]
is of the same quality as the primary in-
WhJfe cident radiation.
Llghf
Now replace the suspension in the glass
- - - - -B box by a solution of some fluorescent
~1 -~
~
~
~
~
.ts
~
material, such as fluoresin. The glass
box, when illuminated by white light, will
nolongerappearwhite but will emit a fluor-
~ ~
c.,s •O
escent radiation, blue, green, or red, char-
FIG. 125.
acteristic of the solution. 2 An absorbing
screen now will not in general absorb the same proportion of the
primary beam when placed at A as of the secondary beam when
placed at B. Conversely, if we observe that a given absorbing
screen does not absorb the same fraction of the secondary beam
when placed at Bas of the primary beam when placed at A, we
may conclude that the quality of the secondary ra<liation re-
emitted by the contents of the box is different from the quality
of the primary beam. Incidentally, if the glass box contains a
mixture of the scattering suspension and the fluorescent solution,
the secondary radiation emitted from the box will contain partly
scattered, partly fluorescent, light.
Barkla showed that the secondary X-ray radiation emitted by
a substance (which we shall refer to as the "secondary emittor")
irradiated by a primary beam of X-rays is made up partly of
scattered X-rays, the quality (hardness) of which is (practically)
identical with the hardness of the primary beam, and partly of
1 The particles of the suspension are assumed large enough so that the well-
known Tyndall effect plays no role in the sea ttering process.
2 The quality of this fluorescent radiation is independent of the quality
of the primary radiation. According to Stokes' law, however, the exciting
primary radiation must be of shorter wave length, or must contain compo-
nents of shorter wave length, than the wave length of the fluorescent radia-
tion. Red light cannot excite blue-green fluorescence.
470 X-RAYS [CHAP. XII
fluorescent rays, the quality of which 1s characteristic of the
secondary emittor. Barkla's conclusions were based upon a
study of the mass absorption coefficient, µ/ p (see equation (3)) of
both the primary and the secondary beam in some standard sub-
stance, such as aluminum, as compared with the absorption in
other substances. Schematically, let a primary beam of hard
X-rays from a target T (Fig. 126), after passing through holes
in lead screens SvS 1 , fall upon the secondary emittor E. Let the
secondary beam, taken off, say, at right angles to the primary
beam, after passing screens 8 282, enter the ionization chamber C
by means of which the intensity of the secondary beam can be
measured either with or without slabs of absorbing material
placed at A or at B.
S.,A
I : E
I l
- --)J
-~
- -S2
rfL}
Frn. 126.-Schematic arrangement for studying secondary radiation.
When the secondary emittor is of some light material, such as
carbon, an aluminum absorbing screen placed at B absorbs
(nearly) the same fraction of the secondary beam as it does of the
primary beam when placed at .A.. This shows that the quality of
the secondary beam as measured by jts absorption in aluminum
is (nearly) the same as that of the primary beam. The primary
beam has been scattered by the secondary emittor.
If, however, a heavier material, such as silver, be substituted
for the carbon, the absorption coefficient of the secondary beam
in aluminum is no longer the same as that of the primary beam.
The mass absorption coefficient of the secondary beam is greatet
than that of the primary beam-which indicates that the quality
(hardness) of the secondary beam is no longer the same as that of
the primary. Barkla found that the hardness of the secondary
SEC. 4] SECONDARY RADIATION 471
beam, as measured by its coefficient of absorption in aluminum,
is characteristic of the material used as secondary emittor. If,
for a given (hard) incident primary beam, the mass absorption
coefficient in aluminum of the secondary radiation emitted by
various substances be plotted as a function of the atomic weight
of the emittor, a smooth curve results, as is shown in Fig. 127.
The heavier the secondary emittor the more penetrating its sec-
ondary radiation This secondary radjation is, thus, seen to be
characteristic of the emittor. By using various emittors, we can
obtain secondary radiations of various qualities as measured by
absorption coefficients in aluminum.
'e'
::::,
·; 500
<t
g
t: <\Ca
0
i 400
\
""t5
'5
a::
t:s
21
-g 300
~
V)
<+-
i-~
-
t;::.
2.00
\ . \
~
QI
0
v
c
0
100
t-
o
~
_n
<( ~ -s-__ A[
~ 0 ~
~ 4-0 50 60 10 80 90 100 110
Atq_mic Weig ht
Frn. 127.-The quality of the secondary radiation from various elements
as a functiqn of atomic number. The absorption of the radiation by aluminum
is taken as an index of quality.
If the absorption coefficients of such a series of secondary
radiations obtained from a series of emittors of different atomic
weights be measured in some absorber such as iron or copper and
these values be compared with the absorption coefficients of these
same radiations in aluminum, a curve of the type shown in Fig.
128 results. Point a corresponds to a "soft" radiation which
has a large coefficient of absorption in aluminum. For more
penetrating radiations, the coefficients of absorption in both
aluminum and iron decrease along the line ab. At b, which
corresponds to a radiation the absorption coefficient of which in
aluminum is about the same as that of the secondary radiation
from an iron emittor, the absorption coefficient of the radiation in
472 X-RAYS [CHAP. XII
the iron absorber suddenly increases to point c. Thereafter, for
more penetrating radiations, the absorption in iron decreases
again toward point d. Radiations in the range a to bare emitted
by elements of lower atomic weight than iron; in the range c to d,
by elements of higher atomic weight. The position of the discon-
tinuity be depends, in a perfectly regular way, on the atomic
weight of the absorber: the higher the atomic weight of the
absorber the greater the penetration (in aluminum) of the radi-
ation at which the discontinuity occurs.
These facts pointed unambiguously to the emission by a second-
ary emittor of a fluorescent radiation characteristic of the emittor,
when the emittor is radiated by a primary beam of sufficient
hardness. Barkla called this radiation the K fluorescent radia-
c
0
L..
r-f
.~ t
+c: I
Cl.> c
·0
ti:
CY
0
u
c: a
0
+
t d ib
_2 I
~.__~~~~~~I~~~~~
Absorption Coeffi'cient in Atum.1nurn
FIG. 128.
tion of the emittor. For the heavier absorbers, a second discon-
tinuity was observed beyond point a of Fig. 128. This indicated
that along with the K fluorescent radiation was another and
a softer fluorescent radiation, which Barkla designated by L.
Barkla recognjzed that these Kand L radiations constitute lines
(or groups of lines) in the fluorescent X-ray spectra of the several
secondary emittors. In these X-ray spectra, the mass absorp-
tion coefficients in aluminum were somewhat analogous to wave
lengths in visible light. The discovery of the action of the
crystal grating showed that this analogy has a very real origin.
6. The Crystal Diffraction Grating. (a) Previous Estimates
of X-ray Wave Lengths.-The researches of Barkla in isolating
the K and the L lines in the fluorescent spectra of the elements
pointed unmistakably to something akin to definite wave lengths
in X-rays, particularly when the close analogy with the phenom-
ena of scattering and fluorescence in the visible region of the
SEC. 5] THE CRY ST AL GRATING 473
spectrum was considered. Evidences that X-rays are waves
similar to, but shorter than, ordinary light waves were also
accumulating from attempts to detect diffraction. We have
already referred to the experiments of Haga and Wind (p. 458),
who, from the slight evidences of diffraction which they obtained,
concluded that X-rays were very short ultra-violet waves, the
wave length of which was of the order of 10-s cm. Further
attempts to detect diffraction were made by Walter and Pohl,1
who obtained negative results, from which they concluded that if
X-rays were short ultra-violet waves, the wave length could not
exceed 10- 9 cm. Sommerfeld,2 from a study of the diffraction
experiments of Haga and Wind and of Walter and Pohl, con-
cluded that the wave length of X-rays might be of the order of
magnitude of 3 X 10-9 cm. And W. Wien, 3 from a measurement
of the velocity of the secondary cathode rays (i.e., photoelectrons)
emitted when X-rays are incident onto a secondary emittor,
concluded, by application of Planck's quantum hypothesis,
that the wave length of the X-rays must be about 6.7 X 10-9
cm. By 1912, therefore, it was well established, at least as a
good working hypothesis, that X-rays are very short electro-
magnetic waves, the order of magnitude of the wave length being
in the neighborhood of 10-s to 10-9 cm. There was great need
for a direct experiment to confirm this hypothesis. This experi-
ment was supplied by Friedrich and Knipping as a result of a
brilliant suggestion by Laue, 4 that a crystal, with its regular,
three-dimensional array of atoms, should behave toward a
beam of X-rays in somewhat the same way as does a ruled
diffraction grating toward a beam of light. This experiment
very definitely marks the beginning of a new era in the technique
of X-ray measurements and in X-ray theory.
(b) Elementary Theory of the Crystal Grating. 5-Let us consider
a train of plane parallel waves (water, air, or light waves), the
crests of which are represented by the straight lines a, b, c, d . . .
(Fig. 129) proceeding in the direction of incidence indicated by
the arrow I. Let these waves pass over a series of small obstacles
1 Ann. Physik, vol. 29, p. 331 (1909).
2 Ann. Physik, vol. 38, p. 473 (1912).
3 Gesel. Wiss. Gottingen, N ach1 ., vol. 5, p. 598 (1907).
4 FRIEDRICH, KNIPPING, and LAUE: Le Radium, vol. 10, p. 47 (1913);
Ber. Konigl. Bayer, Akad., Mfinchen, p. 303 (1912).
5
See BRAGG, W. L.: Proc. Cambridge Phil. Soc., vol. 17, p. 43 (1912).
474 X-RAYS [CHAP. XII
1, 2, 3 . . . arranged in a straight line. These obstacles
become secondary sources of disturbance (i.e., they act as scat-
terers), and a system of secondary circular (or spherical, as the
case may be) wavelets surrounds each obstacle, as shown, the
circles being drawn to represent correct phase relations among
the systems of wavelets. It is observed that these secondary
wavelets have, in general, a random distribution with respect to
each other, except in the one direction indicated by the arrow
R where there is a tendency for the wavelets to combine into a
series of envelopes which gives rise to a reflected wave train pro--
a b
m
n
FIG. 129.-Schematic representation of the secondary rays from centers of
disturbance 1, 2, 3 . . . when set into forced vibrations by the incident primary
waves a, b, c . . .
ceeding in the direction R. This series of obstacles then causes
a small fraction of the incident beam to be, in effect, reflected
according to the well-known law of reflection that the angle of
incidence equals the angle of reflection.
Next, let us consider the effect on an incident beam of a three-
dimensional array of such "diffracting centers," arranged in
regular rows, columns, and layers. Figure 130 shows three
rows AA, BB, CC of such centers, in one plane only of course, the
distance between the rows being d. (We imagine the array to
be extended in three dimensions.) Let an incident beam of
wave length X be incident onto this array at a "glancing angle"
(} of 30 degrees. As shown in Fig. 129, each of these rows of
SEC. 5] THE CRY ST AL GRATING 475
diffracting centers will give rise to a reflected beam, and we shall
have the reflected beams Oil, 022, 033, etc. As drawn in the
figure, the phase relations among these reflected beams are such
that the crests of the waves in the train 033 exactly coincide with
the troughs of the train 011. These two beams will, therefore,
1
I
I
I
JOO
I
A ' A
ct 3
I
I
B -B
c c
FIG. 130.-The destructive interference of secondary wave trains.
destructively interjere and w1:ll cancel each other. In like manner,
022 will interfere with O44; and so on, for all other pairs of reflected
waves. The net result is that there is no wave train reflected
from the three-dimensional array of diffracting centers. In
general, for any combination of 'A, d, and 8, it will be possible to
FIG. 131.-The reinforcement of secondary wave trains.
find pairs of reflected beams, which, in like manner, will destruc-
tively interfere with each other, and no reflected beam will
result.
It is possible, however, to choose particular sets of values of
'A and () for a given d such that the reflected waves are all in
phase, as is shown in Fig. 131, where, for the particular values of
476 X-RAYS [CHAP. XII
X and d, for which the construction was made, the glancing angle
(} is 16.6 degrees. The crests of the trains Oil, 022, 033 . . . are
seen to coincide; the reflected wave trains reinforce each other;
and a reflected beam results. The condition that these indi-
vidual reflected trains shall be in phase is, obviously, that the
reflected train 022 shall be exactly 1 wave length, or an integer
multiple of wave lengths, behind the train Oil. The simple
relation among X, d and 8 to which this requirement leads is at
once seen from Fig. 132, which shows the incident beam I reflected
at 01 and 02, as in Fig. 131. The line ab is perpendicular to the
reflected rays Oil and 0 22. The length of path 0102b is greater
than the length of the path 01a by the length of the broken line
c02b, the line Oc being perpendicular to 0102. The length of the
.t
J
I
e
I
02
FIG. 132.
broken line c02b is, obviously, 2d sin 8. The condition that there
should be a reflected beam is, therefore, that
nX = 2d sin(} (20)
where n is an integer. This is known as Bragg's law.
Figure 133 represents one horizontal plane of a three-dimen-
mensional array of diffraction centers. There are numerous
possible combinations of parallel layers (the planes of which are
perpendicular to the plane of the paper) 1, 2, 3, 4 . . . with
different distances d1, d2, d3 . . . between planes. Now,
imagine a parallel wave train, containing a continuous spectrum
of wave lengths, to be incident onto this group of diffracting
centers as shown by the parallel arrows a, b, c, d . . . If, in this
incident heterochromatic wave train, there is a wave length A2
such that
SEC. 5] THE CRYSTAL GRATING 477
where n is an integer, d2 is the distance between the set of planes
numbered ".re," and 02 is the glancing angle of incidence between
the direction of the incident radiation and the planes, there will
be reflected from this group of planes a beam A, of wave length
A2 , which will proceed in the direction of the arrow A. Similarly,
we may have reflected beams B, C, D in different direc-
tions. In order that a beam shall be reflected through a measur-
/
O e O e O e O e O e O e e O e O
e O • 0 e O e O e O e e O e O e
O e O e O e O e O e e O e O e O
e O e O • 0 O e e O e O e O e
O e O e O e e e O e O e O • 0
.A.J
• . 0 e O e O e O e O e O e O
o_ •. o • o • e O e O e O e
A
• 0 • • 0
(:):.
O e O e J e O e O
• o· • e O e O ff3
'3
0 •
•
• • 0
• 0 e O
e O • 0 e O e O •
• 0 • 0 • 0 • 0 e O e O e Q
Frn. 133.-Schematic representation of the reflection of monochromatic
wave trains of X-rays by a crystal of NaCl when radiated by a heterochromatic
incident beam.
able angle, however, it is necessary that d shall be not too great
compared to A.
Now, (1) if a beam of X-rays is comprised of waves the wave
length of which is of the order of magnitude of 10-s or 10- 9 cm.,
and (2) if a crystal is composed of a regular array of atoms
which, it might be expected, act as diffraction centers for X-rays,
then, on passing a beam of X-rays through a crystal somewhat
478 X-RAYS [CHAP. XII
after the manner of the diagrammatic arrangement of Fig. 133,
beams of X-rays corresponding to A, B, C . . . should be
reflected in various directions from the crystal. It was this
experiment, suggested by Laue, that Friedrich and Knipping
tried.
6. The Experiment of Friedrich, Knipping, and Laue.-By
means of suitable screens S1S 2 (Fig. 134), a narrow pencil of
X-rays from the target Twas allowed to pass through a crystal
C beyond which was a photographic plate PP. After an exposure
of many hours, it was found on developing the plate that, in
addition to the interior central image at 0, where the direct
beam struck the plate, there were present on the plate many
regularly arranged, but fainter, spots, indicating that the
incident X-ray beam had been reflected from the various crystal
planes, exactly in accordance with the discussion of the preceding
Frn. 134.-The arrangement by which Friedrich and Knipping discovered the
action of a crystal on a beam of X-rays.
section. Figure 135 shows such a photograph, taken by Dr.
George L. Clark, 1 of an iron crystal. The perfect symmetry of
the spots about the central image is clearly brought out. In
their original paper, 2 Friedrich, Knipping, and Laue, from an
analysis of a series of photographs of a crystal of zinc blend
oriented at various angles with respect to the incident pencil,
concluded that there were present in the X-ray beam wave
lengths varying between 1.27 · and 4.83 · 10-9 cm.-values in
excellent agreement with previous estimates. This positive
result at once proved the correctness of the two postulates under-
lying the experiment: (1) that X-rays are electromagnetic waves
of definite wave lengths and (2) that the atoms of a crystal are
1 The author is indebted to Dr. Clark and to the McGraw-Hill Book
Company, Inc., for permission to use this photograph.
2 Loe. cit. p. 473.
SEC. 6] A FUNDAMENTAL EXPERIMENT 479
arranged in a regular three-dimensional lattice, as may be
inferred from the external symmetry of crystals.
As a result of this famous experiment of Friedrich, Knipping,
and Laue, two new and very important fields of investigation
were at once opened up: (1) the study of X-ray spectra and the
determination of wave lengths and (2) the study of crystals
with particular reference to the arrangement of atoms in crystals.
These fields are, to a certain extent, mutually interdependent,
Frn. 135.,-,Laue photograph of an iron crystal (taken by Dr. George L. Clark).
since the crystal grating is such an indispensable part of any
X-ray spectrometer. In the following sections, we shall discuss
some of the more important aspects of X-ray spectra and their
interpretation. For more detailed presentations of X-ray
phenomena and their applications in chemistry, crystallography,
biology, medicine, and industry, we refer the student to the
numerous standard treatises. 1
1 COMPTON, A. H.: "X-rays and Electrons" (1926); DE BROGLIE, M.:
"X-rays" (translated by Clark) (1924); SIEGBAHN, M.: "The Spectroscopy
of X-rays" (translated by Lindsay) (1925); CLARK, GEORGE L.: "Applied
X-rays" (1927).
480 X-RAYS [CHAP. XII
7. The X-ray Spectrometer.-Immediately following the
announcement by Friedrich, Knipping, and Laue of their
successful experiment, many investigators took up a study of
the new phenomenon. Among these were W. H. and W. L.
Bragg/ to whom we are chiefly indebted for the early develop-
ment of the X-ray spectrometer.
A spectrometer of the Bragg type is shown diagrammatically
in Fig. 136 (a). _X-rays from the target T of an X-ray tube pass
through two narrow slits 81 and S2, a few hundredths or tenths
of a millimeter wide, the edges of which are made of some
T
----Jj ---~
I
I
----S2 ---S2
_,\\ -(:;) ->\ ·(:)
D
I
I I I
I I I
I I f
r-20' k,c>c9*·za·
/ I
I I
0 .L p
( b)
~ ca)
Frn. 136.-(a) The X-ray spectrometer, ionization method; (b) the X-ray
spectrometer, photographic method.
material, such as lead or gold, which is very opaque to X-rays.
This ribbon-shaped incident beam of X-rays I falls at a glancing
angle () onto the cleavage face of a crystal K-rock salt, calcite,
mica, gypsum, quartz, etc.-which is mounted on a table D
the angular position of which can be accurately read by verniers
or micrometer microscopes. The reflected beam of X-rays,
which makes an angle 20 with the incident beam, enters, through
the "window" w, an ionization chamber C, by means of which
1 BRAGG, W. L.: "Specular Reflection of X-rays," Nature, vol. 90, p. 410
(December, 1912); BRAGG, W. H., and BRAGG, W. L.: "Reflection of X-rays
by Crystals," Proc. Roy. Soc., London vol. 88, p. 428 (July, 1913); BRAGG,
W. H.: "X-rays and Crystals," Nature, vol. 91, p. 477 (July, 1913).
SEC. 7] THE X-RAY SPECTROMETER 481
(see Fig. 115) the intensity of the reflected beam may be measured.
By suitably turning the table D about the axis A, the incident
beam may be made to strike the face of the crystal at any glanc-
ing angle 0. The ionization chamber C is mounted on an arm
(not shown) by means of which the chamber can be rotated about
the axis A so as to admit the reflected beam through the window
W. For protection against stray, scattered radiation, a third slit
S 3 is attached to the chamber.
For photographic registration, the ionization chamber may
be replaced by a photographic plate PP (Fig. 136(b)). With
the crystal set at a glancing angle fJ, the reflected beam will
strike the plate at L (or at L', if the crystal be "reversed").
From the position O at which the direct beam strikes the plate,
the distances OL and OA and, hence, the angles 28 and fJ may be
determined. The wave length >.. is then obtained from the
formula
n>.. = 2d sin fJ
The distance d between the reflecting planes of a crystal
such as NaCl is determined as follows: From his investigations,
Bragg showed 1 that, in the rock-salt crystal, the Na and the Cl
atoms (inore accurately, the Na and the Cl ions. See Chap.
XI, Sec. 2) occupy alternate positions at the corners of elemen-
tary cubes in the cubic lattice characteristic of the crystal, the
arrangement being similar to that shown in Fig. 133, which
represents one plane of atoms. Taking the atomic weight of
chlorine as 35.46 and of sodium as 23.00, the molecular weight
of NaCl is 58.46. Therefore, 58.46 grams of the NaCl contain
2No atoms, i.e., No atoms of Na and No atoms of Cl, where
No = 6.06 X 10 23 is Avogadro's number. The number of
atoms n in 1 cm.-cube of rock salt is
2 x 6.061 . 10 23
(21)
n= 58.46 X P
where p = 2.163 is the density of NaCL If d is the distance
between the center of one atom and the next along the edge of
the cube, 1/d is the number of atoms in a row of atoms 1 cm.
long, and the number of atoms in the centimeter-cube is
1
n = d3 (22)
1 See BRAGG: "X-Rays and Crystal Structure."
482 X-RAYS [CHAP. XII
Equating equations (21) and (22) and solving ford, which is the
desired distance between the (cleavage) planes in NaCl, we find
d = 2.81 X 10-s cm.
The value of d is seen to be dependent on M, the molecular
weight of rock salt; on p, the density of rock salt; and on Avo-
gadro's number No. The values of No and p are not known to
much better than 0.1 per cent. There is, hence, a corresponding
uncertainty in the value of d. At present, measurements of
X-ray wave lengths can be made with a precision many fold
greater than the best measurements of M, No, or p. It is, there-
fore, customary to assign an arbitrary value, the best obtainable
by experiment, to the grating space of some crystals, as, for
example, the value of 2.81400 X 10-s cm. for NaCl, and to
express X-ray wave lengths in terms of this value. The grating
space of other crystals are then compared to this arbitrary value
for NaCl as a standard. 1
Table I gives some of the crystals used in X-ray spectros-
copy and the corresponding grating spaces.
TABLE I.-SOME CRYSTALS USED IN X-RAY SPECTROSCOPY
Grating space,
Crystal centimeters times
108
Rocksalt, NaCl. ..................... . 2.814
Calcite, CaC0 3 (18°C.1) ............... . 3.02904
Quartz, 8102 ........................ . 4.247
Gypsum, CaS04 · 2H20 ............... . 7.578
Potassium ferrocyanide, K4Fe (CN) 6 ••••. 8.408
Mica, .............................. . 9.93
Sugar .............................. . 10.57
Palmitic acid ........................ . 35.59
Stearic acid ......................... . 38.7
1 Of course, the grating space of any crystal depends upon its temperature. The grating
space of calcite increases 0.00003 X 10- 8 cm. per degree (centigrade) increase of tempera-
ture in the neighborhood of 20°C. For NaCl, the corresponding change is approximately
0.00011 X 10- 8 cm. per degree.
There are numerous alternative arrangements for using a
crystal in connection with an X-ray spectrometer. Among these
may be mentioned that of Seemann, 2 in which the two colli-
mating slits 8 1 and 8 2 , of Fig. 136, are omitted and in their place
1 The International Critical Tables give 3.028 X 10-s cm. as the grating
space of calcite.
2 Ann. Physik, vol. 49, p. 470 (191~"
SEC. 8] BRAGG'S DISCOVERY 483
is used a wedge W (Fig. 137) the edge of which is placed close
to and parallel to the cleavage face of the crystal. For more
penetrating radiation, Siegbahn 1 has used the arrangement shown
in Fig. 138, in which the incident beam passes through the crystal
and is reflected from the inner planes. The second slit 82 is
placed immediately behind the crystal and is rigidly connected
to the table D on which the crystal is mounkd. The table can
be rotated about an axis passing through 82.
I
I
p 0 L p p 0 L p
FIG. 137.-The Seemann "slit." FIG. 138.-A slit system employed
by Siegbahn.
8. Bragg's Discovery of Monochromatic Characteristic Radi-
ations.-With a beam of X-rays incident onto the cleavage face
of a (mica) crystal, as shown in Fig. 136(b), W. H. Bragg found 2
that there was produced on the photographic plate PP animage
or line at Lin the exact position to be expected if the beam were
reflected from the crystal in accordance with the theory outlined
in Sec. 5. He then replaced the photographic plate by an ioniza-
tion chamber and found that when the chamber was so placed
that the reflected beam entered the chamber, an ionization cur-
rent was produced, which current he took as a measure of the
intensity of the reflected beam. Then, using a rock-salt crystal
and an X-ray tube with a platinum target, he investigated the
intensity of the reflected beam for various glancing angles e.
He found that the intensity did not vary uniformly with angle
but that at certain angles the intensity rose to a sharp maximum.
When intensity was plotted as a function of angle, a curve
1 "The Spectroscopy of X-rays" (English translation), p. 57.
2 Nature (Jan. 23, 1913).
484 X-RAYS [CHAP.XII
similar to that shown in Fig. 139 was obtained. A group of
three maxima, a1, b1, and c1, was observed at the respective
angles (} of 9.9, 11.6, and 13.6 degrees. A second group of three
maxima, a2, b2, and c2, was observed at approximately double
these angles. This second group is similar, as regards relative
intensities of the maxima, to the first group. Bragg interpreted
the maxima a1, b1 and c1 as three monochromatic lines; and the
second group of maxima a2, b2, and c2 as second-order reflections
Wavelength (Angstroms)
0 0.4- O.B J.2 I.6
\
\
\
+i::
Q)
L.
'-
::,
u
c:
0
+0
N
. i:
0
H
0 5 \0 15 20 25 30 35
Glancing Angle (in degrees)
F1G. 139.-Bragg's curve for the energy distribution in an X-ray spectrum, show-
ing the characteristic lines a, b, c.
of the lines a1, b1, and c1. He computed their wave lengths by
the formula
n'A = 2d sin (}
where n = 1 for the ''lines" a1, b1, and c1; and n = 2 for the
second-order lines a2, b2, and c2. A third-order peak b3 was
observed at 36.6 degrees. Using the present accepted value of
d = 2.814 Angstroms for rock salt, the wave lengths correspond-
ing to these three lines were as follows (Table II):
TABLE IL-FIRST MEASUREMENTS OF X-RAY WAVE LENGTHS BY BRAGG
(PLATINUM TARGET)
Line O, degrees Sino n I A, Angstroms
a1 9.9 0.172 1 0.9,
b1 11.6 0.200 1 1.13
b2 23.6 0.400 2 1.13
ba 36.6 0.597 3 1.12
C1 13.6 0.235 1 1.32
SEC. 9] MOSELEY'S LAW 485
Curves similar to Fig. 139 were obtained with other crystals-
calcite, iron pyrites, zinc blend, etc.-the only difference being
that the maxima occurred at different glancing angles, indicating
that each crystal had a characteristic grating space d. Bragg
convinced himself, however,_ that these respective maxima for
different crystals always represented the same monochromatic
radiation, since, for example, the absorption in aluminum of peak
b1 was always the same, whatever the crystal used. In short, the
interpretation of the curve in Fig. 139 became clear: the peaks
are spectral lines the wave lengths of which, as Bragg showed
later,1 are characteristic of the target emitting the rays. These
monochromatic lines are superimposed on a continuous spectrum 2
represented by the partially dotted line in Fig. 139. Curves of
the type shown in Fig. 139, therefore, represent (subject to certain
corrections to be mentioned later) the distribution of energy in
the X-ray spectrum, continuous and characteristic combined, of
an element.
9. Moseley's Law.-In two classic papers, 3 Moseley presented
a systematic study of the characteristic radiations emitted by
various targets, using a photographic method similar in principle
to that shown in Fig. 136(b). He found a larger number of
characteristic lines than did Bragg and, also, that these lines
could, in general, be classified into two groups: (1) a group of
shorter wave lengths, which, by means of the value of absorption
coefficients in aluminum, he identified with Barkla's K character-
istic secondary radiations and (2) a group of lines of longer wave
length, similarly identified with Barkla's L radiation. Unlike
the optical spectra, the X-ray characteristjc spectra of the ele-
ments were found to be similar from element to e!ement,
homologous lines occurring, in general, at shorter wave lengths
the greater the atomic weight of the element in which the lines
originate.
Table III, taken from Moseley's paper, shows the wave lengths
found by him for the two strongest lines of the K series, which
lines Moseley named Ka and K{3. In searching for a relation
between the frequency of a given line, say the Ka line, and some
property of the atom in which the line originated, Moseley first
1 Proc. Roy. Soc., London, vol. 89, pp. 246, 430 (1913).
2 See also MosELEY and DARWIN: Phil. Mag., vol. 26, p. 210 (1913).
3 Phil. Mag., voL 26, p. 1024 (1913) and vol. 27, p. 703 (1914). The stu ..
dent should consult the original papers.
486 X-RAYS [CHAP. XII
observed that the frequency did not vary regularly with the
atomic weight, as is shown by curve A (Fig. 140) in which the
square root of the frequency is plotted against the atomic weight.
TABLE III.-MosELEY's VALUES FOR THE WAVE LENGTHS OF (HIS) a and {3
LINES OF THE K SERIES
Wave len~th,
0
Angstroms Atomic Atomic
Element
weight number
a {3
Al ....................... 8.364 7.912 26.96 13
Si ....................... 7.142 6.729 28.06 14
Cl ....................... 4.750 35.46 17
I( . . . . . . . . . . . . . . . . . . . . . . . 3.759 3.463 39.09 19
Ca ...................... 3.368 3.094 40.07 20
1"i ....................... 2.758 2.524 47.9 22
Va ...................... 2.519 2.297 50.96 23
Cr ...................... 2.301 2.093 52.01 24
Mn ...................... 2 .111 1.818 54.93 25
Fe ...................... 1.946 1.765 55.84 26
Co ...................... 1.798 1.629 58.97 27
Ni ...................... 1.662 1.506 58.69 28
Cu ...................... 1.549 1.402 63.57 29
Zn ...................... 1.445 1.306 65.38 30
Y ....................... 0.838 89.0 39
Zr ....................... 0.794 91 40
Nb ...................... 0.750 93.1 41
Mo ...................... 0.721 96.0 42
Ru ...................... 0.638 101.7 44
Pd ...................... 0.584 106.7 46
Ag ...................... 0.560 107.88 47
Bohr had recently proposed his theory of the origin of spectra,
in which theory the charge Z on the nucleus played a fundamental
role. According to Bohr's theory, the frequency v of a spectral
line is given by (see equation (33), Chap. X)
]) = z2 . 21r2me4 (_!_ - _!_)
h3 Ti T~
from which
y'; a: Z.
Rutherford had shown, from his experiments on the scattering
of a particles, that the value of the nuclear charge, for a given
atom, is very approximately one-half the atomic weight; and
SEC. 9] MOSELEY'S LAW 487
Barkla had shown, from experiments on the scattering of X-rays,
that the number of electrons surrounding the nucleus is, also,
approximately one-half the atomic weight. So, silicon, for
example, atomic weight 28.06, should have 14 extranuclear
electrons. It is the fourteenth element in the periodic system.
The ordinal number of an element should, therefore, be identical
with Z, the charge (taking e = 1) on the nucleus. Moseley,
accordingly, assigned atomic numbers Z (column five of Table
III) to the elements which he had investigated, and, in accord-
Atomic Weight
O JO 20 :!>O 40 50 60 10 80 90
24.--~--r--~~~~~~---~~~-.-~----~~....,c.-~~~~100
co
I
~
x
a1l6
c
Q)
:J
er
~
l.L.
12--~--~---~---------
'+-
0
-+-
~g 81-----+---+----+-__,.~~---t---~--1-~--t------+--~
er.,
L.
0
::, .
.$ 4 - - - - - - - - - - - - - - - - - - - - - - - - - - - -
-b
00 4 12 16 20 '24 28 32 36 40
Atomi'c. Number, Z
Fm. 140.-Moseley's curve showing the relation between the frequencies of
X-ray lines and atomic number.
ance with the relation suggested by Bohr's equation, he plotted
a curve between V v and the atomic number Z. Such a plot of
Moseley's data for the Ka line is shown in Fig. 140, curve B.
The graph is seen to be a straight line, with a small intercept b = 1
on the Z-axis. It is obvious from a comparison of the two
curves A and B, of Fig. 140, that, so far as concerns the determi-
nation of the frequency of characteristic lines, atomic number is a
much more fundamental quantity than atomic weight.
Empirically, the relation between the frequency v of the Ka
line and Z, as determined from Fig. 140, curve B, is
v = 0.248 X 10 16 (Z - 1) 2 (23)
488 X-RAYS [CHAP. XII
In Bohr's equation (33), Chap. X, for the frequency v of a
spectral line, if we set r1 = 1 and r2 = 2, we obtain for v
v = o.328 x 101 6 • z2 c71 2 - }'22) (24)
= 0.246 x 10 z16
•
2
(24')
by inserting the numerical values of m, e, and h. Except for
the slight correction to Z, equations (23) and (24') are seen to
be almost identical. We may, therefore, conclude that equation
(24), in which Z is to be replaced by (Z - 1), gives to a first
approximation the frequency of the Ka line of the elements.
This leads to a very far-reaching conclusion as to the originof
the Ka line: From Bohr's interpretation of the meaning of r 1 = 1
and r2 = 2 in equation (33), Chap. X, we conclude that the
Ka line originates when an electron transfers (or jumps) from an
orbit corresponding tor = 2 to an orbit corresponding tor = 1. We
can thus form a reasonably concrete picture of the atomic
processes (in terms of Bohr's theory) which lead to the production
of the Ka line in the X-ray spectrum of an element. In a normal
atom for which Z > 10, we have seen that there are 2, and only 2,
electrons in orbits corresponding to quantum number 1; there are
8 electrons, and onlY. 8, in orbits corresponding to quantum num-
ber 2. The effect of the bombardment of the atoms in the target
of the X-ray tube by the swiftly moving electrons of the cathode
stream is, if these electrons possess sufficient energy, to "knock"
from (some of) the atoms in the target electrons which normally
occupy orbits of quantum number 1. To fill the orbits thus
made vacant, electrons "drop" from orbits of quantum number 2
to the vacant orbits of quantum number 1 and thereby emit
monochromatic radiation-the Ka line. In a similar way, the
K{3 line may be shown to originate when electrons drop from
orbits of quantum number 3 into vacant orbits of quantum num-
ber 1. Because of tpe part which they thus play in the produc-
tion of the K lines, these 2 electrons, normally occupying orbits
for which r = 1, are called K electrons; and the orbits are called
K orbits. Similarly, the electrons normally occupying orbits for
which r = 2 are called L electrons, and the corresponding orbits
are L orbits.
This picture affords, also, an explanation, at least qualitative,
of the appearance of the factor (Z - 1) instead of Zin equation
(23). When, as a result of bombardment by the cathode stream,
one of the K electrons is removed from an atom, there is 1 K
SEC. 10] THE CONTINUOUS SPECTRUM 489
electron left near the nucleus. This electron "screens" the
nucleus and makes its effective nuclear charge 1 unit less; hence,
the factor (Z - 1).
Following Moseley's work, the technique of X-ray spectros-
copy developed rapidly. New lines were discovered and
classified, and their origin in terms of electron transfers were
systematically worked out. We shall postpone further discussion
of characteristic X-ray spectra until after we have considered the
continuous spectrum and related phenomena.
10. The Continuous X-ray Spectrum. (a) Limiting Minimum
Wave Lengths and the Determination of h.-As has been pointed
out above, the X-ray spectrum, in general, consists of a charac-
teristic, or line, spectrum superimposed on a continuous spectrum.
The positions, i.e., wave lengths, of the lines are determined
solely by the material of the target; their intensity is deter-
mined, for a given target material and tube current, by the
voltage applied to the tube. On the contrary, the wave-length
characteristics of the continuous spectrum are quite independent
of the material of the target but are determined by the voltage
applied to the tube. The intensity of the continuous spectrum,
for a given tube current, is dependent both on the target material
and on the applied voltage as well as on the thickness of the
target. In this section, we shall give a brief survey of some of
the more important facts concerning the continuous spectrum.
If, with the ionization spectrometer shown diagrammatically
in Fig. 136(a), the crystal be set at a series of angles (} and the
ionization current at each angle be measured, then a curve,
plotted between these several ionization currents as ordinates
and the values of the corresponding wave lengths determined
from the formula nX = 2d sin 8, as abscissre, is called a spectral
energy-distribution curve. A part of such a curve was shown in
Fig. 139. A series of four such curves for the radiation from a
tungsten target and for applied voltages of 20,000, 30,000,
40,000, and 50,000 volts, respectively, is shown in Fig. 141. 1
Starting at the long-wave-length side, the curves rise to a maxi-
mum and then drop rapidly toward zero, the position of the
maximum depending on the applied voltage. The curves are
seen to be somewhat similar to the black-body curves of Fig.
54, voltage in X-ray radiation playing a role in a rough way
1 These curves are from the measurements of ULREY, Phys. Rev., vol. 11,
p. 401 (1918). They are subject to cert~in corrections to be mentioned later.
490 X-RAYS [CHAP. XII
analogous to temperature in black-body radiation. But there
is this very important difference between the black-body and
the X-ray curves: Whereas the black-body curves, on the short-
wave-length side, approach the wave-length axis asymptotically,
the X-ray curves meet that axis at finite angles, as is shown at
the intersections a, b, c, and d, respectively, of Fig. 141. These
intersections can be very accurately measured by making read-
ings nearer the axis 1 than the readings shown in Fig. 141. It
is seen that the intersections come at shorter wave lengths the
higher the voltage. In Table IV are shown the limiting wave
lengths-the intersections a, b, c, and d, of Fig. 141-and the
TABLE IV.-LIMITING WAVE LENGTHS AND FREQUENCIES IN THE
CONTINUOUS X-RAY SPECTRUM AT VARIOUS VOLTAGES
(From Ulrey's data, Fig. 141)
Limiting Limiting fre-
Potential,
wave length, quency, times
kilovolts 0
Angstroms 10-18 cm.- 1
20 0.615 4.87
30 0.405 7.40
40 0.310 9.76
50 0.250 12.0
corresponding limiting (or maximum) frequencies for the four volt-
ages. A very simple relation exists between these limiting frequen-
cies and the applied voltage. This relation is shown graphically
in Fjg. 142, in which the limiting frequency is plotted as ordinates
against the applied voltage as abscissm. The graph is a straight
line passing through the origin; the limiting frequency is strictly
proportional to the applied voltage, the empirical equatjon of the
curve being
14 1
Vm = 2.43 X 10 V (25)
where vm is the maximum frequency which an applied voltage
V' (in volts) can generate.
The explanation of this sharp "cutoff," at a limiting minimum
wave length, shown by the curves in Fig. 141, and of the simple
relation contained in equation (25) illustrates, in a very striking
manner, the inadequacy of the classical theory. According to
the ether-pulse, i.e., the classical, theory 2 of the production of
1 See DuANE and HuNT: Phys. Rev., vol. 6, p. 166 (1915).
2 See Sec. 3, this chapter.
SEC. 10] THE CONTINUOUS SPECTRUM 491
:n
-+-
.in 6
r:
CL>
-t-
c
H
CV
>
+Ci 4
Cl)
~
'2.
Fm. 141.-Ulrey's curves for the distribution of energy in the continuous X-ray
spectrum of tungsten at various voltages.
12.5---------.-----.-------,------.-----,
TQ 10.0
x
:n
0
@
:::,
g- 1.5 1 - - - - - + - - - - - + - - - - - , . . , - . . - - - + - - - - - - t
it:
E
:J
E
·x:
~ s.o~---+----~---t------+------i
t:n
c
+
.E
:J 2.5L--~--""~--'---+----l-----+-------I
OIC--~~_.._---~~~-1--~~---~--'
o 10,000 w)ooo 30,000 40,000 so,ooo
Voltage Applied~ X-Ra!:j Tube
Fm. 142.-The relation between the maximum or limiting frequency and applied
voltage.
492 X-RAYS [CHAP. XII
X-rays, the Fourier analysis of the ether pulses into wave
trains should result in an asymptotic approach of the X-ray
energy-distribution curves to the wave-length axis on the short-
wave-length side; whereas experiment requires a sharp cutoff.
There have been various attempts to postulate a type of ether
pulse which should yield such a sha.rp cutoff, or a limiting
maximum frequency, for a given applied voltage. These
attempts have been, in general, unsuccessful. In short, the
classical theory offers no explanation for this phenomenon.
Again, the quantum theory is more successful. From the
standpoint of the sequence of processes involved, the production
of X-rays by the bombardment of a target with electrons is the
converse of the photoelectric effect. In the latter, monochro-
matic radiation of frequency v falling upon a photoelectrically
sensitive substance causes the emission of electrons the (kinetic)
energies of which, as they leave the suface of the emittor, vary
all the way from a maximum Em, given by
down to zero. In the production of X-rays, a stream of elec-
trons, all possessing the same kinetic energy E, determined by
the potential V applied to the X-ray tube and given by
E = Ve
where e is the charge of the electron, are incident onto the target
of the X-ray tube and cause the emission of radiation (X-rays)
the frequencies of which vary all the way from a maximum
frequency vm down to zero. Because of the similarity of these
two phenomena, it may be expected that Einstein's photoelectric
equation (equation (33), Chap. VI)
Ve = hv - wo
which was successful in "explaining" the photoelectric effect
should, when "inverted," apply, also, to the phenomenon of
X-ray production. In equation (33), Chap. VI, wo was identified
with the work required for the photoelectron to escape from the
surface of the emittor. Since this work was, at most, equiva-
lent to only a volt or so, it is obvious that in applying equation
(33), Chap. VI, to X-ray phenomena we may neglect w0 with
respect to the other two terms of the equation, since in the pro-
SEC. 10] THE CONTINUOUS SPECTRUM 493
duction of X-rays we are dealing with many thousands of volts
and, therefore, Ve > > wo. We may, therefore, write, as the
equation which expresses the converse of the photoelectric process,
hvm = Ve (26)
where Ve is the energy which is acquired by the electrons in
the X-ray tube in "falling" from the cathode of the tube through
the potential difference V to the target, and vm is the maximum
frequency of the X-rays produced when these electrons are
brought to rest by impact with the atoms of the target. In
other words, when an electron possessing kinetic energy Ve
collides with, and is brought to rest by, an atom, a quantum of
(X-ray) energy hvm is emitted, and so long as the kinetic energy
of the incident electrons does not exceed Ve, no frequencies
greater than vm can be emitted. We should, however, expect
in addition to vm a whole spectrum of lower frequencies in the
emitted X-ray beam, since, as was pointed out above, in only
a very small proportion of cases would it happen that an electron
incident upon the target would be brought to rest by a single
encounter; and it is only these single encounters which give rise to
the frequency vm. In the vast majority of cases, an electron
would experience many collisions before being brought to rest.
At each of these collisions, some of the energy Ve is dissipated.
Occasionally, an electron may convert a part or all of its reduced
energy into a quantum of frequency less than vm. If the target
be sufficiently thin-say of very thin gold foil-only a few of
the electrons of the incident cathode stream will collide with
the atoms of the target, the majority of the electrons passing
through the target substantially undeviated. (This phenome-
non is utilized in the well known cathode-ray tube developed by
Dr. W. D. Coolidge.) It is obvious (1) that only those electrons
which collide with atoms can produce X-rays; and (2) that,
because of the thinness of the target, the chances that a given
electron will dissipate some of its energy before a collision are
much less than when thick targets are used. Accordingly, we
should expect that a greater proportion of the energy in the
continuous spectrum from thin targets should be nearer the vm
limit, than from thick targets. This is in agreement with
experiment.
We thus have a qualitative explanation, on the basis of the
quantum theory, of some of the pertinent features of the energy-
494 X-RAYS [CHAP. XII
distribution curves shown in Fig. 141. The maximum fre-
quencies vm, corresponding to the points a, b, c, . . . for the
respective voltages are emitted as a result of a few of the electrons
in the cathode stream, being brought to rest by a single collision
with an atom of the target, the relation between vm and V being
given by
hvm = Ve (26)
Frequencies lower than vm in the X-ray spectrum at a given volt-
age are emitted as a result of collisions in which only a part of
the initial energy Ve of the electron is converted into radiation.
In the sum total of collisions between the electrons and the atoms
of the target, it might be expected that there would be a "most
probable'' type of collision which would correspond to the peak or
maximum of the energy-distribution curve.
With the shape of the energy-distribution curve we are not, for
the moment, concerned. Rather, let us return to the empirical
equation (25) and to the curve (Fig. 142) which showed, in com-
plete agreement with the equation (26), that Va: vm. Writing
equation (26) in the form
e e V'
Vm = hV = h 300 (27)
(where V is in electrostatic units and V' is in volts), we see that
the numerical constant in equation (25) should be equal to the
multiplier of V' in equation (27). That is, we should have
3;0h = 2.43 x 1Q14
from which, taking e = 4.774 · 10- 10 e.s.u.,
h = 6.55 . 10- 21
in exact agreement with the value of Planck's constant h determined
by other methods.
This identity in numerical values of h points to the general
correctness of the explanation, on the basis of the quantum
theory, of the cutoff points a, b, c . . . of the energy-distribu-
tion curves in Fig. 141 as well as of equation (26). In fact,
measurements of the maximum frequency vm for given voltages
V applied to the X-ray tube have yielded some of the most
dependable values of h. Some of these values determined in
this way are given in Table V.
SEC. 10] THE CONTINUOUS SPECTRUM 495
TABLE V.-VALUES OF PLANCK'S CONSTANT h DETERMINED FROM THE
LIMITING wAVE LENGTH IN THE CONTINUOUS X-RAY SPECTRUM
Target h,
Investigators
used erg-sec.
Duane and Hunt1........ Tungsten 6.50 x 10-27
W ebster 2 . . . . • . • • • • • • • • • • Rhodium 6.55
Blake and Duane 3 • • • • • • • • , Tungsten 6.556
Ulrey4. . . . . . . . . . . . . . . . . . Tungsten 6.54
Wagner 5 • • . • • • • • . • • • . • • • Various 6.52
1 Phys. Rev., vol. 6, p. 166 (1915).
2 Phys. Rev., vol. 7, p. 599 (1916).
3 Phys. Rev., vol. 10, pp. 93, 624 (1917).
4 Phys. Rev., vol. 11, p. 401 (1918).
s Physik. Z., vol. 21, p. 621 (1920).
(b) Total Intensity and Distribution of Intensity .-Determina-
tions of the total intensity of, and of the distribution of intensity
in, the continuous X-ray spectrum are byno means so satisfactory
as are the measurements of minimum wave lengths. This is
due to the fact that numerous troublesome corrections have to
be made to the observed energy-distribution curves, such as
those shown in Fig. 141, before the true curve can be obtained. 1
Among these necessary corrections may be mentioned the
following:
1. The variations with wave length of the absorption of the radia-
tion in its path from the atom of the target in which the radia-
tion originates to the ionization chamber where the intensity
of the radiation is measured. This path includes the
material of the target itself, since the X-rays originate at
finite (though very small) depths below the surface; the
walls of the X-ray tube and the window of the ionization
chamber; and the air between the tube and the chamber.
2. The reflecting power of the crystal grating. This varies with
wave length.
3. Incomplete absorption in the ionization chamber. Either an
ionization chamber of sufficient length to absorb the entire
X-ray beam must be used, or else corrections must be
made for variation with wave length of the absorption
within the chamber, since the observed ionization current is
proportional to the energy absorbed in the chamber.
1
These corrections, however, in no wise affect the determination of mini..
mum wave lengths.
496 X-RAYS [CHAP. XII
4. "Second order" reflection. According to equation (20),
which may take the form
A = 2d sin o,
n
when the crystal in the spectrometer (Fig. 136) is set at a
given angle 8, a series of wave lengths A will be reflected
for different values 1, 2, 3, . . . of n, provided those
several wave lengths are present in the incident radiation.
Suppose, for example, that when the voltage applied to the
X-ray tube is 50 kilovolts, the crystal is set at such an
angle 80.6 as to give a first order reflection (i.e., n = 1) of
0
X = 0.6 Angstrom. According to Fig. 141, the minimum
wave length present in the X-ray spectrum generated by
0
50 kilovolts is 0.250 Angstrom. Under these conditions,
there will be reflected into the ionization chamber the
0
second order wave length A = 0.3 Angstrom, in addition to
0
the first-order wave length X = 0.6 Angstrom. The
observed intensity when the spectrometer is set at A = 0.6
0 •
Angstrom 1s, therefore, due to the 2 wave lengths. For
an applied voltage of 30 kilovolts, only first-order reflec-
tion is present at 80.6, and no correction is necessary.
Without discussing the methods of making these several
corrections, 1 the final results may be summarized as follows:
(a) The total radiation, i.e., the area under the (corrected)
energy-distribution curves, is, for a given target material
and tube current, very nearly proportional to the square
of the applied voltage.
(b) At a given voltage and tube current, the total radiation
varies as the first power of the atomic number of the
target.
(c) The wave length Am at which a given energy-distribution
curve attains its maximum energy is given, to a rough
approximation, by the relation
Am = 1.5Ao
where "Xo is the wave length at which the same curve
meets tne wave-\engtn axis.
1 See KIRKPATRICK, P.: Phys. Rev., vol. 22, p. 414 (1923).
SEC. 111 LAWS OF ABSORPTION 497
(d) Wagner and Kulenkampff 1 give the following empirical
formula for the intensity Iv of the X-ray energy at
frequency v from a target of atomic number Z when
bombarded by electrons which have fallen through a
potential V:
Iv = CZ(vo - v) + BZ2
where v0 is the maximum frequency determined by the
quantum relation Ve = hvo, and C and B are constants
independent of voltage and target material. This
formula is in approximate agreement with a formula
deduced by Kramers 2 en theoretical grounds.
Except for determinations of minimum wave lengths, how-
ever, and the resulting law
hvm = Ve
our knowledge of the continuous X-ray spectrum is in a very
unsatisfactory state. The problem is a difficult one, whether
approached from the theoretical or the experimental side, and
much work remains to be done to bring this phase of X-rays up
to the level of other branches of the subject.
11. The Empirical Laws of the Absorption of X-rays.-In
contrast with the apparently chaotic state of affairs attending
the absorption of light of the visible or near-visible portions of
the spectrum, one finds comparative simplicity in the empirical
laws of the absorption of X-rays. Measurements of absorption
coefficients are readily made by use of the ionization spectrom-
eter (Fig. 136(a) ). For a given crystal angle () and, therefore,
wave length A, the ionization current is measured both without
and with a sheet of absorbing material of known density p and
thickness x placed in the path of the beam, say between the two
slits 81 and 82. These measurements give, respectively, Io and I
of equation (3)
_!!:.px
I = Ioe P (3)
from which the mass absorption coefficientµ/ p may be computed.
In this way, being careful to use such a voltage applied to the
X-ray tube as to eliminate ''second order'' reflections, 3 one may
1
Ann. Physik, vol. 68, p. 369; and vol. 69, p. 548 (1922).
2
Phil. Mag., vol. 46, p. 836 (1923).
3 Seep. 496.
498 X-RAYS [CHAP. XII
obtain values of µ/ p for var10us wave lengths and various
substances.
Figure 143 shows, to approximate scale,1 the mass absorption
coefficients of lead in the wave-length range 0.1 < X < 1.2
0
Angstroms. Beginning at point o, in the neighborhood of 0.1
0
Angstrom, µ/ p rises rapidly with increasing wave length until
0
point a, corresponding to X = 0.138 Angstrom and µ/ p = 8
1:
<l)
120,~~-+-~~--+-~-4--+~-H-+--+-~~+-1
·;:;
~
~
L)
100~~-+--~~+-~-+-+-~~~-+-,e--+----t
c
0
i_
6 ao~~-+--~~-+-~~~~-+-~--+-,~~---t
cJ)
..n
<(
cJ)
~ d'
~ 601--~-+-~~--+--l--+~~-+-~..;._,f-=-~-1
40 1 - - - - - - - ~ - - - + - - K = O.IJ8A
Iq= a780
.L11 =0.8/J
w~~-+--+-+-~-4-~~r=-~-+-~~
Lllr0.950
a o'
O C!A O.&
O K0.2 0'4- 0.6 . LILll LJJi
IO
Waveleng+h in angstroms
Frn. 143.-K and L absorption limits.
(about) is reached, at which occurs a sudden drop in the value
of µ/ p to point a'. With further increase in wave length, the
absorption increases rapidly again, until point b, corresponding
to A = 0.780 Angstrom, is reached, at which occurs another
drop in the absorption coefficient to point b'. Similar "drops" 0
or "breaks" occur at cc' corresponding to X = 0.813 Angstrom,
1
The abscissre represent wave lengths accurately, but the ordinate scale
is only approximate.
SEC. 11] LAWS OF ABSORPTION 499
0
and at dd', corresponding to A = 0.950 Angstrom. Beyond
point d', the absorption again increases rapidly. If we could
follow, by direct measurement, the absorption beyond point 0
e, we should find that in the region 3.2 < X < 5.0 Angstroms, a
group of five such "breaks" occur. Still beyond these, beginning
at about 14 Angstroms, comes another group of seven "breaks."
These breaks, or discontinuities, are called, respectively,
K, L, M . . . discontinuities, because of their intimate con-
nection with the K, L, M . . . series of characteristic X-ray
lines. There is one K discontinuity-aa'. There are three L
IOO r - - - - - r - - - - - - . - - - - , - - - - , - - - . - - - - . . - - - - - - - - - , . - - - -
so~~-t-~-+-~---,r--~-+-~-+-~---,~~-t-~--+-~---t
vf
40i--~--t~~---t-~~-+~~-t-~"'---t-~~-+-~~-+-~~+-~--J
10 20 30 40 50 60 10 80 90
Atomic Number
Frn. 144.~The Moseley diagram for the K and L absorption limits.
discontinuities-L1 , Lu, and Lm at, respectively, bb', cc', dd';
there are five M discontinuities; and so on.
Curves exactly similar to Fig. 143 are obtained for the absorp-
tion of X-rays in other substances, the respective discontinuities
or ''limits'' occurring at longer wave lengths the lower the atomic
number of the absorber. In fact, a curve plotted between the
square root of the frequency of a given limit and atomic number
is nearly a straight line similar to Moseley's curve (Fig. 140) for
line spectra. Such a curve for the K limit, from Mg (Z = 12)
to Th (Z = 90) is shown 1 in Fig. 144, in which, as ordi-
nates, are plotted V v/R, where R is the Rydberg constant.
1
Data from SIEGBAHN: "The Spectroscopy of X-rays," pp. 137 and 141.
500 X-RAYS [CHAP. XII
( Obviously, ~ a: Vv.) The graph is seen to be nearly, but
not quite, a straight line. For comparison, there are also shown,
in Fig. 144, the corresponding graphs for the Li, Lu, and Lm limits
from Ag (Z = 47) to Th (Z = 90).
Not only is there regularity in the wave lengths of the absorp-
tion limits from one element to another, but there is also a
remarkable regularity in the magnitude of the mass absorption
coefficient, from wave length to wave length in a given element,
and from one element to another. Figure 145 shows, to a much
J4,--......,,,..~-,---=----~----==.........---=~~.------~---~~-
o'
+ IQt--~~-+-~~-+-~~-+~~-+-1--~~-+-~~~
c
a,
·u
i
~ 8
\.)
c::
0
+
~Gt---t-r---+~~-+-~~-+-~~.--~-+-~~~
lct
\I)
•'1)
~ 4~>-+~-+---~-+-~~-+-~~~~~~-+-~~~
0.010 0.015 0.020 0.025 0.030
3
j. Gn ang5froms)
Frn. 145.-Mass absorption coefficients of lead as a function of wave length,
showing the K discontinuity and the linear relation between µ/p and A. 3 •
larger scale, the mass absorption coefficients of lead 1 in the
region o to o' of Fig. 143, plotted as a function of the cube of
the wave length. The two parts of the graph on each side of the
K absorption limit aa' are seen to be straight lines with very
nearly equal intercepts of aboutµ/ p = 1 on the axis of ordinates.
We may, accordingly, write the empirical equation for the
variation ofµ/ p with X in the region oa as
!!.
p
= kX 3 +b (28a)
1 Data taken by the author, Phys. Rev., vol. 27, p. 1 (1925).
SEC. 11] LAWS OF ABSORPTION 501
and for the region a'o',
!!:.
p
= k'X 3 +b (28b)
where k and k' are the slopes of the respective lines and b is a
constant. To a first approximation, 1 the mass absorption
coefficient for a given absorber is a linear function of the cube of
the wave length, due regard being had to the spectral region in
which the equations (28) apply. The constant b is, in part at
least, determined by scattering, and it is frequently assumed that
b = (]'
p
where u / p is the mass scattering coefficient. Equations (28) are
frequently written in the form
!!:
p
= kX 3 + ~p (29)
A quantity of more fundamental significance than mass absorp-
tion coefficient µ/ p is the atomic absorption coefficient µa. This
is obtained by dividing the mass absorption coefficient, which is
expressed in grams as the unit of mass, by the number of atoms
per gram, No/A, where No is Avogadro's number and A is the
atomic weight, i.e.,
µ/p
µa= No/A
.Dividing equation (29) through by No/A, we have
µa = kaA 3 + <Ia (30)
k
where ka = No/ A and <Ia is the atomic scattering coefficient.
The term kaX 3 is sometimes abbreviated to Ta, so that
µa = Ta + <Ia
For reasons which will appear in Sec. 14, infra, Ta is called the
"fluorescent" absorption coefficient.
An equation of the type of equation (30) is applicable to the
atomic absorption coefficients of all substances on the short-
wave-length side of, and for a limited region beyond, their respec-
tive K absorption limits XK. Further, ka varies in a perfectly
1
There are slight systematic departures from the A3 law. See RICHT-
MYER, F. K.: Phys. Rev., vol. 18, p. 13 (1921); also, ALLEN, S. J. M.: Phys.
Rev., vol. 27, p. 266 (1926).
502 X-RAYS [CHAP. XII
systematic way with the atomic number of the absorber. It
has been shown 1 that ka is proportional to the fourth power 2
of the atomic number, i.e.,
ka = KZ 4
from C (Z = 6) to Pb (Z = 82), K being a proportionality con-
stant. Accordingly we may write as an empirical equation for
the atomic absorption coefficient µa of X-rays of wave length X
in an absorber of atomic number Z,
(31)
On the short-wave-length side of the K absorption limit, the value
of K is about 2.25 · 10- 2 , where X is expressed in centimeters.
For a considerable range beyond the K limit on the long-wave-
length side, the value of K is about 0.33 · 10- 2 •
This systematic variation of the absorption of X-rays with
atomic number Z and wave length X, as expressed empirically
by the comparatively simple equation (31), has prompted numer-
ous theoretical investigations in which attempts have been
made to derive a formula of the type of equation (31) from
fundamental concepts. A summary, even, of these theoretical
studies is quite beyond the scope of this book, but we shall
make brief reference to the subject later (p. 513), after we have
discussed the relation of the absorption of X-rays to the emission
of the characteristic X-ray spectra, to which latter subject we
shall now turn.
12. Characteristic X-ray Spectra.-Immediately following
the work of Bragg and of Moseley, discussed in Secs. 8 and 9,
above, the technique of X-ray spectrometry developed rapidly.
Many new characteristic lines were soon discovered, and, at the
present time, practically the entire X-ray spectrum of each ele-
ment has been mapped out. Following the notation of Barkla,
Bragg, and Moseley, these lines are classified into groups, respec-
tively; K, L, M, . . . The K group consists of four principal
lines known, in order of increasing wave length, as K'Y, K{3, Ka1,
and Ka 2 • In Table VI are given the wave lengths of these four
lines for several representative elements. For comparison, the
1 RicHTMYER, F. K., and WARBURTON, F. W.: Phys. Rev., vol. 22, p. 539
(1923).
2 This "fourth power" law was first proposed by BRAGG and PIERCE,
Phil. Mag., vol. 28, p. 626 (1914).
SEC. 12] CHARACTERISTIC SPECTRA 503
TABLE VL-WAVE LENGTHS IN .ANGSTROMS OF THE K LINES OF SOME
REPRESENTATIVE ELEMENTS
Element 'Y {3
S ................... 16 ...... 5. 0213 5.3609 5.3637 5.0123
Ca ................. 20 ...... 3.0834 3.3517 3.3549 3.0633
Fe ................. 26 1. 7406 1.7527 1.9323 1.9365 1.7377
Zn ................. 30 1.2810 1.2926 1.4321 1.4359 1.2963
Br ................. 35 0.9183 0.9308 1.0376 1.0416 0.9197
Mo ................. 42 0.6197 0.6314 0.7078 0. 7121 0.6184
Ag ................. 47 0.4861 0.4965 0.5582 0.5626 0.4850
I ................... 53 0.3748 0.3834 0.4325 0.4370 0.3737
w .................. 74 0.1790 0.1842 0.2088 0. 2135 0.1781
Pb ................. 82 0 .1412 0 .1461 0.1652 0 .1700 0.1385
U .................. 92 0.1084 0.1119 0.1264 0 .1309 0.1075
wave lengths AK of the K absorption limits are shown in the last
column. It is to be noted that, in each case, the wave length of
the K limit is slightly less than the wave length of the gamma line.
That these four lines of the K group, or series, are intimately
connected with each other and with the K absorption limit is
readily shown by considering the voltage which must be applied
to the X-ray tube to generate the lines. Referring to the curves
shown in Fig. 141 for the energy distribution in the continuous
spectrum of tungsten, it is see11 that these curves show no such
characteristic "humps" as were observed by Bragg and are shown
in Fig. 139. As the voltage applied to the tube is raised, however,
the minimum wave length Ao moves toward shorter wave lengths,
according to the law given in equation (26) :hvm = Ve. Now, it is
found that when the voltage applied to the X-ray tube has been
raised to such a value that Ao = XK, all four lines of the K series
are simultaneousiy produced, and with further increase in applied
voltage, these four lines all increase in intensity at exactly the same
rate. In the case of tungsten, for example, for which Ax = 0.1781
Angstrom, by use of equation (27) (or of the equation X =
0
12,345/V Angstroms, on p. 387), we find that Ao = AK when the
applied voltage is 69,300 volts. When higher voltages are
applied to a tungsten-target X-ray tube, all the four lines of the
K series of tungsten appear. The lines are absent for lower
voltages, ~ven though the applied voltage be such that Ao is
greater than the wave length of the Ka lines. From the values of
AK, given in Table VI, one can calculate at once the voltage
necessary to generate the K series of lines of the several elements.
504 X-RAYS [CHAP. XII
A further, and still more intimate, relation between the K
series of lines and the absorption limits is shown by considering
the frequencies 1 of the lines and the limits. In columns two and
TABLE VII.-RELA'l,ION OF THE FREQUENCIES OF THE Ka LINES OF
TUNGSTEN TO THE FREQUENCIES OF THE K AND L ABSORPTION LIMITS
Absorption limit Emission lines
I
v
}...,
.Angstroms
-
R ..,....--.....,k: }... -Rv
~,~
~
K .......... 0.17806 5,116.8
Lr ......... 1.024 889.9 4,226.9 Ka3(?) ..... m1ssmg
Lrr · · · · · · · · · 1.0726 849.6 4,267.2 Ka2 ....... 0.21352 4,267.8
Lm ........ 1.2136 750.9 4,365.9 Ka1 ....... 0. 2088514, 366. 3
three of Table VII are shown the wave lengths and the values of
v / R for, respectively, the K, Lr, Ln, and Lm absorption limits of
tungsten. 2 In column four is shown the differences in v / R
values between the K limit and each of the three L limits. The
second part of the table shows the directly observed values of A
and of v/ R for the Ka1 and the Ka2 lines of tungsten. It is seen
that the observed value of v/R for the Ka1 line is almost exactly
equal to the dijference between the value of v/ R of the K limit and
that of the Lrn limit. Similarly, the v / R value of the Ka 2 line is
equal to the difference between the value of v / R of the K limit
and that of the Lu limit. (There is, however, no known emission
line which corresponds to the difference between v / R of the K
limit and that of L1 • We shall make mention of this later.)
Putting this relation in symbols, we have (the notntion is obvious)
(~Ll - (it -(;tIJI (32a)
(;L2 - (;t (;t (32b)
In part to avoid writing the large numbers required to give frequencies,
1
the frequency vis usually divided by the Rydberg constant R, giving v/R.
This quantity is proportional to frequency, since R is a constant. It is to
be pointed out that the symbol, namely R, which, in accordance with
current practice, we have here used for the Rydberg constant, differs from
that used in preceding chapters.
2 Data from SmGBAHN, Zoe. cit.
SEC. 12] CHARACTERISTIC SPECTRA 505
Or, expressed in frequencies P,
PKal = PK - PLrn (32'a)
PKa2 = PK - PLn (32'b)
Equations (32') are exactly analogous to the equations which
we discussed in Chap. IX, for the lines of a spectral series. In
these equations for a spectral series, the frequency of a line, P,
is given by the difference between two terms, or quantities, which
themselves, as was pointed out, must have the dimensions of
frequency. Similarly, the quantities PK, PLn and PLrn of equations
(32') may be thought of as X-ray terms, the differences between
which give the frequencies of X-ray emission lines. In the
series of lines in the visible and the ultra-violet part of the spec-
trum, the existence of these terms was, in general, inferred from
the relations between the lines. In the X-ray case, the existence
of the terms is made directly evident from the absorption limits,
such as are shown in Fig. 143.
This illustration, showing that the frequencies of the Ka
lines are related to the K and L absorption limits, leads to a
law of far-reaching importance in X-ray spectra: The frequencies
of X-ray lines are given by the differences between the frequencies
of absorption limits. By means of this principle, it is possible,
step by step, from the observed data on X-ray lines and limits,
to build up the whole scheme of X-ray snectra and absorption
limits. For example, the frequency of the K{3 line should be
given by the difference between the frequency of the K absorp-
tion limit and that of some other limit, the frequency of which
we can determine from the known frequency of the K{3 line and
that of the K limit, thus
PKfJ = PK - PM
where PM is the unknown frequency. The value (Table VII) of
P/R for K{3 is 4,942.9. Subtracting this from the value of P/R for
the K limit, namely, 5,116.8, gives for PM the P/R value 173.9,
0
corresponding to a wave length of approximately 5.2 Angstroms.
We should, therefore, expect to findo an absorption limit in tungs-
ten in the neighborhood of 5.2 Angstroms. It is difficult to
make direct measurements of absorption coefficients at that
wave length, but the existence of limits in this region is readily
demonstrated by indirect means: the differences between lines
in the L series and the L limits give M limits in the same way
506 X-RAYS [CHAP. XII
that the differences between the Ka lines and the K limit give L
limits.
The L series contains some 25 or 30 lines designated, in an
exceedingly chaotic notation, by a, (3, or 'Y, with various sub-
scripts 1, 2, s . . . The wave lengths and v/R values of some
of these lines are given in Table VIII. (The reason for the
grouping into the three columns L 1 , Ln, and Lm will appear
presently.)
TABLE VIIL-WAVE LENGTHS AND v/R VALUES OF SOME OF THE LINES
IN THE L SERIES OF TUNGSTEN
Lr II Lu II Lur
Line X I v/R II Line I X I v/ R 11 Line I X I v/R
{34 • ••••••••. 1,299 701.7 1/ • •••••••• 1,418 642.8 l . ........ 1,675 544.0
{33 •••••••••. 1,260 723.2 f31 ..•••... 1,279 712.4 a2 .••••••. 1,484 617.8
'Y 2 •••••••••• 1,066 855.0 'Y 5• ••••••• 1,129 807.0 a1 •••••••. 1,473 618.5
'}'3 •••••••••• 1,060 860.0 'Yl •••••••• 1,096 831.8 {32 . ....... 1,242 733.8
'Y 4 ..•••.•••. 1,026 887.8 'Y 6 .••.•••• 1,072 850 {35 . ....... 1,212 751.6
Limit .......... 889.9 Limit ......... 850 Limit ......... 751
Recalling, from above, that by combining PK with vKfJ, an
"M" absorption limit was predicted in the neighborhood of
v/R = 173.9, we observe that by subtracting from the L1 limit
v/R = 889.9 the v/R values of the {34 and f3s lines in the table,
we get v/R values of 188.2 and 166.7, respectively. These are
two more terms-M terms, or levels-in the spectrum of tung-
sten. The term 173.9 predicted from the K{3 line is, therefore,
really two terms. 1 Similarly, we get two more terms, namely,
207 and 138, by combining the Lu limit with the lines r, and
f31. The combination of the Lm limit with l and a2 gives the
terms 207 and 133, the former of which is the same as the com-
1For, turning the procedure around, by subtracting these two limits
from the K limit of tungsten v / R = 5, 116.8, we should predict two lines in the
K series, the v / R values of which should be 4, 928.6 and 4, 950.1. The value of
v/R given above for the K{3 line, namely, 4,942.9, falls between these two,
which suggests that the K{3 line, originally thought to be a single line, is
really a close doublet. Such actually proves to be ''the case, as recent
measurements of the K{3 line of some of the lighter elements show. For
example, Allison and Armstrong (Phys. Rev., vol. 26, p. 701 (1925)) find that
the molybdenum KB line consists of two components, the respective wave
lengths of which are 0.631571 and 0.631009 Angstroms.
SEO. 12] CHARACTERISTIC SPECTRA 507
bination Ln - r,. We therefore have five terms of limits-they
are called M limits-which we arrange in descending order of
v/ R, as follows:
v/R x,
ANGSTROMS
M 1 . . . . . . . . 207 5.30
Mu ....... 187 5.85
Mm ...... 167 6.55
M1v· ...... 138 7.95
Mv ....... 133 8.25
These are the five M absorption limits in the X-ray spectrum of
tungsten. 1 In a sjmilar way, it is found that there are seven
N limits. A very large number of the observed lines in X-ray
spectra can be accounted for by suitable combinations of these K, L,
M . . . limits. The K series of lines results from combinations
between the K limits and the limits L, M, N . . . of lower
frequency. The L series of lines results from combinations
between one or the other of the L limits and limits M, N . . .
of lower frequency; and so on. 2
Some lines, however, which might be predicted on the basis of
this principle of the combination of limits are not found. We
have seen one example of this in Table VII: there is no ljne
corresponding to the combination K - Lr. Similarly, there is no
line corresponding to Lr - Mv; ortoLu - Mu. Of the 15mathe-
matically possible combinations between the three L limits and
the five M limits, only seven result in actual lines. The remainder
are "ruled out" by a selection principle, similar to that operating
in optical spectra, to which we shall make reference in the next
section.
1 It is noted that if differences be taken between the L limits and the other
£-series lines shown in Table VIII, much smaller quantities than those
obtained for the M limits result. For example, Lr - 'Ya gives 29.9; Ln -
'Y1 gives 18; etc. These are the N absorption limits.
The value 173.9 obtained by the combination K - K{3 is a mean value
between the Mu and Mm limits and is not a separate M ]imit.
2 Lines accounted for in this way are frequently called "diagram lines,"
since they fit into the energy-level diagram discussed in Sec. 13. Associ-
ated with many of these diagram lines are found fainter lines which
appear to be satellites of the diagram lines, and which do not fit into the
energy-level diagram shown in Fig. 146. For a discm,sion of these satel-
lites (also called "non-diagram lines'' or "spark lines"), see Sommerfeld:
"Atomic Structure and Spectral Lines;" also Robinson, Phil. Mag., vol. 4,
p. 763 (1927).
508 X-RAYS [CHAP. XII
13. X-ray Energy-level Diagrams.-(a) With the empirical
data of X-ray spectra, presented in the preceding section, now
before us we are in position to combine Moseley's conclusion
(Sec. 9) regarding the origin of the Ka line 1 with the concept of
the arrangement of electrons in atoms, discussed at the end of
Chap. XI, to explain the origin of the X-ray spectral lines.
Moseley concluded, from an application of Bohr's theory,
that the Ka line arises from the electron jumps from orbits
of quantum number n = 2 to orbits of quantum number n = 1.
We saw, in Chap. XI, that the electrons in the normal atom
occupy various orbits characterized by total quantum numbers
1, 2, 3 . . . There are, in the heavier atoms at least, 2
electrons in orbits of quantum number n = 1; there are 8 elec-
trons in orbits of quantum number n = 2; for n = 3 there are 18
electrons. The 8 L electrons (see Appendix II) are divided
into three subgroups; the 18 M electrons, into five subgroups, the
subgroups being characterized by certain azimuthal quantum
numbers k and inner quantum numbers j. Each of these
s:ibgroups corresponds to a definite energy level within the atom.
Likewise, we have seen that, because of the part which they
play in the origin of X-ray lines, the absorption limits
K, L 1 , Lu . . . must represent energy levels within the atom.
Bohr's theory suggests an intimate connection between these
observed absorption limits or energy levels and the several
groups of electrons: The two most strongly bound electrons
are in the K level and are called K electrons. The next
higher energy level, the 211 level, contains 2 electrons and
obviously corresponds to the Lr absorption limit. Then
comes the 221 level, with 2 more electrons, corresponding to the
Lu limit. The 222 level, with 4 electrons, completes the "L
shell" and corresponds to the Lnr absorption limit. Continuing
in this way, the various Mand N absorption limits are identified
with the several nki groups of electrons, as shown in Appendix
II. The absorption limits are thus characterized by the various
quantum numbers nk'J of the electron groups with which they
are respectively associated. For example, the Lu limit has
quantum number 221; the Mrv limit has quantum number 332;
and so on.
1The resolution in Moseley's apparatus was not sufficient to separate the
components a 1 and a 2 of the Ka line, which was later shown to be a doublet.
SEC. 13] ENERGY-LEVEL DIAGRAMS 509
We can now construct an energy-level diagram, exactly
analogous to the corresponding diagrams for optical spectra
(Figs. 102 and 103), to show the origin of the lines in X-ray spec-
tra. Figure 146 shows such a diagram for the X-ray spec-
trum of uranium (Z = 92), the energy levels K, L, M, N and O
being plotted to a logarithmic scale, as in Figs. 102 and 103.
n kj
5 3 3 Ov
5 3 2 Ou ..__
5 2 2 O.m
5 2 I On
4 I I OI
4- 44 ~T~
·.w/
4 4- 3 NJi
4 3 3 Wv~
4 3 2 No
4 2 2 Nor
4 2 r N1I.,, -2
4 I I 'NJ
3 3 3 lMv-
3 '2. 2 ~~
3 3 2
7II/l.
3 2 I W.Ir \
3 I r Mi~ JJif;' M 9 MIJl .MJY Mv
i14h~:J t6 JI~')
~ lzU ~ raj ? n1nil1f -<· -- - -M Sen'es -- -- ~ -
2 '2. 2 lkr4-
3
2 2 I ~1l~' ~,I,, .,
'2 I I lLz '-lT z;;; Lm
-<- - - - - L Ser/es .. - - -;>
J_p__~~
KSeries
I I I K
Frn. 146.-The X-ray energy-level diagram for uranium showing the complete
- 4
scheme of X-ray spectra.
The three columns of figures at the left of the diagram give the
nki designation of the various levels. The vertical lines indicate
the lines of the K, L, and M series.
A selection rule similar to that mentioned in Chap. X is also in
operation here, namely: Only those interlevel transfers are
possible for which
llk = + 1
.6.j = + 1 or O
510 X-RAYS [CHAP. XII
where ~k and ~j represent the changes in the quantum numbers
k and j from the initial level to the final level. It is seen that this
principle 1 "rules out" such transfers as Lr ~ K, Mr ~ L 1 , etc.
Transfers from both levels Lu and Lm to K are possible. These
two give rise to the "doublet" Ka.1 and Ka.2. Similarly, the
K(3 line, which results from transfers from Mu and Mm to K,
should be a doublet, a close doublet, since the levels Mu and
Mm are close together when compared to the energy of the K
level. Likewise, the transfers from Nu and Nm to K yield a
doublet, as do also the transfers from On and Om. These
last four lines are so close together that they cannot be resolved,
and, together, they constitute the K'Y line as observed.
(b) We can now elaborate somewhat the picture, suggested
at the end of Sec. 9, of the process of the production of X-ray
lines. Considerjng the lines of the K-series, which are produced
by electrons falling from outside levels into the K level, it is
obvious that before such transfers as give rise to the K lines can
occur, a vacant place must first be made in the K shell, since this
shell can contain 2, and only 2, electrons. To accomplish the
removal of one of the K electrons requires that the atom shall be
given sufficient energy, by some process, to remove that electron
to infinity, since, in the normal atom, there are no vacant places
(orbits) between the K shell and the outside of the atom. Once there
is a vacant place in the K shell, the place may be filled by elec-
trons dropping from any one of the outside orbits, subject to the
limitations of the selection principle, into the K orbit. The
totality of such transfers in a large number of atoms gives rise
to the entire K series. The relative intensities of the lines are
governed by the relative probabilities of transfer from the
various outside orbits to the K orbit. These probabilities are
determined, in part, by the relative numbers of electrons in the
several orbits. For example, since there are twice as many
electrons in the Lm level (4 electrons) as in the Ln level (2
electrons), it may be expected that the chance that an electron
will drop from the Lm level to a vacancy in the K shell is about
1 The quantum numbers k andj as here used for the selection rule in X-ray
spectra are, so far as the rule is concerned, similar to the corresponding
numbers in optical spectra. Sommerfeld, however (" Atomic Structure and
Spectral Lines"), in applying the doublet-separation law (equation (83),
Chap X) in explaining X-ray doublet separations, such as Ka.1 - Ka2,
identifies j with the azimuthal quantum number. For discussion of this
point, see ANDRADE, "The Structure of the Atom," Chap. XIII.
SEC. 14] PHOTOELECTRIC PHENOMENA 511
twice as great as the chance of a similar transfer from the Ln
level. The Ka. 1 line should be twice as intense, therefore, as
the Ka. 2 line. This is found experimentally to be the case.
Only if the energy Ve of the electrons which strike the target
is greater than hvK, where V is the potential difference applied to
the tube and vK is the frequency of the K absorption limit, can
the removal of an electron from the K level take place. This is
simply another way of stating the experimental result, mentioned
in Sec. 10, that the K series of lines is produced only when the
voltage applied to the X-ray tube is such that the minimum wave
length of the continuous spectrum becomes less than the K
absorption limit of the material of the target.
(c) As one proceeds from uranium (for which Fig. 146 was
constructed) toward elements of lower atomic number, the
electron groups of higher quantum numbers progressively dis-
appear (see Appendix II), and therewith, also, the corresponding
energy levels as well as the X-ray lines originating from these
levels. In copper, for example, there are no O levels and only
one N level. One should not expect to find, therefore, in the
X-ray spectrum of copper such lines (see Fig. 146 for their origin)
as L{35, L{31, £,y6, £,y3, etc.
14. Fluorescence and the Photoelectric Effect. (a) The
Absorption of X-rays.-It was early observed that X-rays cause
the photoelectric expulsion of electrons from materials onto
which they are incident. According to Einstein's photoelectric
equation
}~mv 2 = hv - wo
a beam of incident radiation of frequency v will free photoelec-
trons with kinetic energy }~mv 2 whenever hv > w 0 where w0 is
the work required to free the electron from the surface radiated.
We have seen in the preceding section that the work required
to remove a K electron from an atom is hvx. If, therefore, a
beam of X-rays of frequency v falls onto a substance the fre-
quency of the K absorption limit of which is vx, it is to be
expected that, by virtue of the photoelectric process, K electrons
may be photoelectrically expelled from (some of) the atoms of the
substance if
If, however,
512 X-RAYS [CHAP. XII
no photoelectric em1ss10n from the K shell can result. This
concept at once offers an explanation of the general features
of the absorption curves, such as Fig.143. Let Fig. 147 represent
diagrammatically the absorption curve for a susbtance which has
a K absorption discontinuity at XK, frequency vK. If radiation
of wave length X1 and frequency Vt falls on a slab of the absorber,
where v 1 > vK, photoelectrons will be expelled from the K shell
with kinetic energies given by
Jimv 2 = hv - hvK (33)
Photoelectrons will come not only from the K level but also
from the L, M . . . levels as well, the residual kinetic energies
FIG. 147.
of the electrons coming from these latter levels being corre-
spondingly greater than that of the K photoelectrons. The
energy required to cause the removal of these electrons comes
from the incident X-ray beam. For every electron photoelec-
trically expelled, whether it be a K, an L, or an M photoelectron,
a quantum of energy hv is removed from the incident beam-i.e.,
due to the photoelectric process, the beam suffers absorption as
it passes through the material.
If the incident beam of X-rays has a wave length X2, however,
and frequency v2 such that v2 < vK, the incident quantum of
energy is not sufficient to remove electrons from the K shell
SEC. 14] PHOTOELECTRIC PHENOMENA 513
(since hv < hvK), and only L, M . . . photoelectrons will
result. For any frequency v of incident radiation such that
v > vK, there will be K, L, M . . . photoelectrons; while if
v < vK, there will be only L, M . . . photoelectrons. Hence,
the sudden break or discontinuity in the absorpiion curves at XK.
Excepting for scattering, practically the entire absorption of
X-rays is accounted for in this way. We may, therefore, divide
the absorption coefficient µa into two parts, viz.,
µa =Ta+ (J'a
where Ta is t~e photo@lectric (or fluorescent) absorption coeffi-
cient and (J'a is the scattering coefficient.
According to the curves in Fig. 143, the absorption increases
rapidly as one approaches an absorption limit from the short-
wave-length side. This means that the nearer the frequency v
of the incident radiation is to VK the more photoelectrons are
expelled; or the greater is the probability that an incident quantum
will collide with an atom in such a way as to expel an electron from
the atom. It is the variation of this probability with wave length
that gives rise to the shape of the absorption curves as expressed
by the empirical equation (31):
µa = KZ 4X3 + (J'a (31)
Partly on the basis of the general laws of probability, partly from
considerations based on both quantum theory and classical
physics, several investigators have derived theoretical formulre
for the atomic absorption coefficient. 1
Of these formulre, that of de Broglie agrees best with experi-
ment in so far as it predicts the absolute magnitude of absorption
coefficients on the short-wave-length side of the K absorption
limit. The formula derived by de Broglie is
(34)
where Ta represents that part of the total atomic absorption
coefficient µa, due to the photoelectric effect above mentioned;
nK, nL . . . are the numbers of electrons in the K, L . . .
shells (i.e., 2, 8, 18 . . . ) ; and EK, EL . . . stand for the
1 COMPTON, A. H.: Phys. Rev., vol. 14, p. 249 (1919); dE BROGLIE, L.:
J our. phys., vol. 3, p. 33 (1922); KRAMERs, H. A.: Phil. Mag., vol. 46, p. 836
(1923); OPPENHEIMER, J. R.: Zeit. fiir Physik, vol. 41, p. 268 (1927).
For a discussion of these formulre, see RICHTMYER, F. K.: Phys. Rev., vol. 27,
p. 1 (1925); and vol. 30, p. 755 (1927).
514 X-RAYS [CHAP. XII
energy hvx, hvL • • • necessary to remove an electron from
the K, L . . . limit. The other quantities have meanings
previously assigned to, them. In using the formula, all the
terms within the bracket are to be used to compute absorption
coefficients Ta on the short-wave-length side of the K limit. For
the region A > AK, the first term nKEi is to be dropped; for A
> AL, the first two terms are to be dropped; and so on.
While, however, the empirical formulm, such as equation
(31), and theoretical formulre, such as equation (43) give a
close approximation to the experimental facts, these formulre are,
in fact, approximations only. The correct law of the absorption
of X-rays is yet to be discovered.
(b) Fluorescence.-ln the visible spectrum, the characteristie
radiation emitted by a substance when radiated with light of
wave length shorter than that of the characteristic radiation is
called "fluorescent radiation." We have already mentioned
the fluorescence of X-rays in connection with the work of Barkla
(Sec. 4). The explanation of this phenomenon follows, at once,
from the above discussion of the photoelectric effect. When
X-rays of frequency v > vK, where vx is the frequency of the K
2.bsorption limit, fall upon a substance, electrons may be photo-
electrically expelled from any of the orbits K, L, M . . . The
filling of these vacant places by electrons falling from outside
orbits results in the emission of the entire characteristic spectrum
of the elements. If, however, vL < v < vK, electrons are not
expelled from the K orbits, and, accordingly, only the L, M . . .
series of lines appear in the fluorescent spectrum-in accord
with the observations of Barkla.
(c) Direct Observation of the Photoelectric Effect of X-rays.-
These photoelectric phenomena together with the existence
within the atom of the various energy levels K, L 1 , Lu . . .
occupied by electrons receive direct confirmation from observa-
tions of the velocities of photoelectrons expelled from a given
element by X-rays of known frequency.
Let a beam of X-rays of frequency v fall upon a substance the
frequencies of the absorption limits of which are vK, vL 1 , vLu
. . . , where v > vK. Photoelectrons may be expelled from
any level. Because of the different amounts of energy hvK,
hvL , hvL • • . needed to remove electrons from the several
I II
limits, the velocities of these photoelectrons as they leave the
atom will not all be the same but will depend on the respective
SEC. 14] PHOTOELECTRIC PHENOMENA 515
levels from which they have originated. Thus, the velocity
vK of the photoelectrons coming from the K level 1s given by
Einstein's equation 1
p,~mvl = hv - hvK (35a)
The L 1 photoelectrons have velocity VL ,
1
likewise given by
(35b)
Since vK > vL 1 , it follows that the L 1 photoelectrons, after leav-
ing the atom, have a greater velocity than the K photoelectrons.
Among the experiments directly confirming this conclusion
are those of Robinson, 2 whose apparatus is shown diagram-
matically in Fig. 148. A beam of X-rays of frequency 3 v enters
\~
'' ,
,' t...'
I' r
1t-f011ochromaf/c I B
X-Rcrgs w p .U.1 p
Frn. 148.-Robinson's magnetic spectrograph for studying the photoelectric
action of X-rays.
through a thin window W a highly evacuated brass box BB and
falls upon a target T of the material under investigation. Photo-
1 Neglecting the change of mass with velocity. In an experiment such as
Robinson's, the velocity of the electrons becomes such that variation of mass
with velocity must be taken into account.
2 Proc. Roy. Soc., (London) vol.104, p. 455 (1923); Phil. Mag., vol. 50, p.
241 (1925). For previous articles on the same subject, see: ROBINSON and
RAWLINSON: Phil. Mag., vol. 28, p. 277 (1914); Hu, K. F.: Phys. Rev.,
vol. 11, p. 505 (1918); DE BROGLIE: J our. phys., vol. 2, p. 265 (1921);
ELLIS, Proc. Roy. Soc., (London) vol. 99, p. 261 (1921).
3 It is impossible by present experimental means to get a strictly mono-
chromatic beam of X-rays of sufficient intensity for such experiments as this.
Accordingly, use is made of the fact that, with suitable exciting voltage, the
Ka lines from an X-ray target are much more intense than the accompanying
radiation of the other wave lengths and, therefore, serve as a "monochro-
matic" beam, particularly if the beam is first passed through a filter of
suitable thickness, the wave length of whose K limit is just shorter than the
wave length of the Ka lines. (See HULL, A. W.: Phys. Rev., vol. 10, p. 661
(1917).)
516 X-RAYS [CHAP. XII
electrons are expelled from the surface of T in all directions and
with various velocities, as indicated by equations (35). The
whole apparatus is placed jn a known magnetic field H, at right
angles to the plane of the paper. H can be varied at will. The
photoelectrons will describe circles of radii r given by, neglecting
the correction for change of mass with velocity,
mv 2
Hev = - - (36)
r
where v is the velocity of the photoelectrons moving at right
angles to the field. Some of these electrons will pass through
the narrow slit S and, continuing their circular path, will strike
the photographic plate PP. If the electrons leaving T have
velocities v1, v2 • • • they will move in circles of radii r1,
r 2 • • • and will strike the photographic plate at points L1,
L 2 • • • As shown in the figure, the arrangement is such as
to "focus" electrons leaving the different parts of the target with
the same velocity onto the plate at such positions as L1 and L2,
the diameter of the circles being the distance between S and L1
or L2.
Robinson found on the plate a number of "lines" correspond-
ing to groups of electrons of different velocities v1, v2, etc., the
magnitudes of which could be readily determined from the posi-
tion of the lines on the plate and the magnetic field H, by use of
equation (36). Knowing v, the kinetic energy of each group of
electrons could then be computed, giving the term ~imv 2 in
equation (35):
~imv 2 = hv - hvo
where hvo stands for the quantities hvK, hvL1 , hvLu . . . The
term hv is known from the frequency of the incident radiation.
The several values of hvo are, thus, at once obtainable, and from
them the absorption limits vx, vL1, VLn
This experiment gives as nearly as one could wish a direct
method of proving (1) that electrons exist at various energy
levels within the atom and (2) that the numerical evaluation of
these levels by spectroscopic methods is correct, since, as Table
IX shows, Robinson's results are in complete agreement with the
data of X-ray spectroscopy. It is most remarkable that Robinson
finds one K level, three L levels, five M levels, and, in the case of
Bi (Z = 83), six of the seven N levels, the levels Nv1 and Nvu
being too close together to be resolved in his apparatus.
SEC. 15] THE COMPTON EFFECT 517
TABLE IX.-X-RAY ABSORPTION LIMITS DETERMINED FROM THE MAGNETIC
SPECTRUM OF PHOTOELECTRONS (AFTER ROBINSON); COMPARISON
WITH SPECTROSCOPIC METHOD
Thorium Tungsten Silver Manganese
(Z = 90) (Z = 74) (Z = 47) (Z = 25)
Level
Robin-I Spectros-\ Robin-I Spectros-1 Robin-I Spectros-1 Robin-I Spectros-
son copy son copy son copy son copy
K .......... ..... ..... ..... . .... . .... . .... 480.1 484
Li .......... ..... ..... ..... . .... 280.0 280.0 55.3 57.1
Lu ......... ..... ..... ..... . .... 260.1 260.4 48.8
Lrn ........ ..... . .... . .... . .... 246.2 247.9 }47.9 48.0
Mi ......... 379.5 381.5 208.5 206.8 53.4
} 47.1
Mu ........ 355.1 354.7 189.4 188.2 43.5
Mnr ....... 297.0 298.0 167.5 166.7 39.0 41.3
Miv ....... 258.3 256.5 139.0 137.0 27.9
A{v ........ 244.5 244.9 133.0 132.4 } 23.8 27.5
NJ .. ......
Nu ........
Nn1 .......
92.8
86.0
69.0
97.9
90.0
71.5
l
)
31.4
42.9
35.0
30.0
N1v ........ 56.1 17.6
17.2
Nv ......... l48.4 47.7 17.1
Nvi ........ 28.2
24.4
Nvn ....... 27.4
(Data is given in terms of J/ 0/ R where P O is the frequency of the limit, and R is the
Rydberg constant.) The columns headed "Robinson" give Robinson's observations.
Those headed ''Spectroscopy" give the values of the levels as determined from X-ray
spectroscopy.
15. The Scattering of X-rays; the Compton Effect.-Mention
was made, in Sec. 3, that Barkla, from his measurements of the
scattering coefficients, computed the number of electrons per
atom by use of Sir J. J. Thomson's formula for scattering which
was based on the ether-pulse (classical) theory of X-rays. This
was the first unambiguous determination of the number of
electrons per atom. In spite of this success, there is no phase
of modern physics which better illustrates the inadequacy of
classical theory than the scattering of X-rays.
(a) Total Scattering Coefficients.-According to equation (12),
of Sec. 3(b), the mass scattering coefficient for X-rays should be
<r 81r e 4 n
p= 3 m 2c 4 p (12)
where p is the density of the scattering material and n is the
number of (scattering) electrons per cubic centimeter. Accord-
ing to this equation: (1), O"/p should be independent of both
scattering material and wave length; and (2), the numerical
value of O" / p should be very nearly 0.2, as may be verified by put-
518 X-RAYS [CHAP. XII
ting into equation (12) the numerical values of the constants
e, m, and c and the values of n and p for some scattering material,
such as carbon or aluminum.
For the lighter scattering materials and for X-rays of moderate
wave length, the scattering coefficients approximate the value
required by classical theory, namely, 0.2. But for hard X-rays,
0
say "X. = 0.1 Angstrom, and for some such scatterer as carbon,
the total absorption coefficient, scattering plus fluorescent absorp-
tion, 1 is only 0.15; while for gamma rays, the wave length of
which is of the order of 0.02 Angstrom, the total absorption
coefficient is, approximately, 0.06. For this latter wave length,
the scattering coefficient must, therefore, be less than one-third
of the value required by classical theory. Thus, not only do scatter-
ing coefficients vary with wave length, contrary to classical
theory, but also the values of the coefficients for hard X-rays
and gamma rays are very much smaller than required by equation
(12).
(b) The Angular Distribution of Scattered X-ray Energy.-The
ether-pulse theory of X-rays led to equation (19)
1¢ = I e24 4 R\ (1
me
+ cos2 <J,) (19)
for the intensity I~ of X-rays scattered per electron at an angle
¢ from the direction of the incident beam, the intensity of which
is I. R is the distance between the scatterer and the place where
the scattered energy is measured. According to this equation,
the scattering should be symmetrical forward and backward
about a plane passing through the scatterer and perpendicular
to the direction of the incident rays. For ¢ = 0, the value of
the bracket in equation (19) is 2; for ¢ = 1r/2, the value is
1. That is, the energy scattered in the forward direction
should be twice that scattered at 90 degrees from the incident
beam. The full line in Fig. 149 gives the predicted distribution
for various angles from the direction of incidence, according to
equation (19).
The measurements of Barkla and Ayres 2 for the scattering
from carbon at various angles, using moderately soft X-rays,
are shown by the small circles. The observations agree very
1 For tables and literature references, see COMPTON, A. H., "X-rays and
Electrons."
2 Phil. Mag., vol. 21, p. 270 (1911).
SEC. 15] THE COMPTON EFFECT 519
well with the predictions, except in the forward direction for
angles less than 45 degrees from the incident direction. In this
forward direction, the scattering is greater than predicted. For
short X-rays, this dissymmetry increases, the backward scatter-
ing becoming relatively less with increasing frequency. For
gamma rays, the dissymmetry becomes very great, as is shown by
the dots which represent the scattering by iron of the gamma
rays from radium C as measured by Compton. 1 In this latter
case, the scattering is almost entirely in the forward direction.
As in the case of the magnitude of the scattering coefficient, the
120 90 60
120 60
Frn. 149.-The angular distribution of scattered X-rays from carbon as a
sea tterer. The full line is the prediction of classical theory; the circles are
the observations of Barkla and Ayres for moderately soft X-rays; the dots are
the observations of Compton for gamma rays.
discrepancy between the classical theory increases with increas-
ing frequency of the scattered radiation.
(c) The Wave Length of the Scattered Radiation; the Compton
Ejfect.-If X-rays consist of wave trains of definite frequency,
then a primary beam of X-rays of frequency v incident onto a
scattering material, such as carbon, aluminum, or the like, should
set the electrons of the scatterer into forced vibrations with
the frequency v; and these electrons should then radiate a second-
ary, or scattered, beam of X-rays of the same identical frequency
as the primary beam. It was early found that, in general,
secondary radiation is less penetrating than the primary-a
fact which was later explained by the presence in the secondary
beam of the characteristic fluorescent radiation, which, as we
1 Phil. Mag., vol. 41, p. 749 (1921). These measurements are not plotted
to the same scale in Fig. 149 as are those of Barkla and Ayres.
520 X-RAYS [CHAP. XII
have seen above, is always of longer wave length, and, there-
fore, less penetrating, than the exciting primary beam. In a
scattering material of very low atomic number, such as carbon,
however, the fluorescent radiation is of such a long wave length
as to be absorbed by even a thin film of air. Yet even with
carbon as a scatterer, the secondary beam was found to be some-
what less penetrating, and presumably, therefore, of somewhat
longer mean wave length, 1 than the primary-a fact quite mex-
plicable on classical theory.
J\--=-0. 7078A
:n
U)
t-
<J)
c
w
- -0.02J6A- -
a
Wavelength
Frn. 150.-The spectrum of scattered X-rays, showing the unmodified line P,
and the modified or "shifted" line S (Compton).
In a spectrometric investigation of scattered radiation, A. H.
Compton 2 showed that when monochromatic radiation falls
upon a scatterer, the scattered beam, at least from the lighter
elements-lithium, carbon, aluminum-as scatterers, is com-
posed of two lines, one corresponding in wave length to that of
the primary beam and the other being of definitely longer wave
length. Figure 150 shows the spectrum of scattered radiation
from a carbon scatterer when irradiated by the Ka line of molyb-
denum, the scattered radiation being observed at an angle of
1 Because of the known variation of the coefficient of absorption with
wave length.
2 Phys. Rev., vol. 21, p. 715; vol. 22, p. 409 (1923).
SEC. 151 THE COMPTON EFFECT 521
90 degrees from the incident radiation. The vertical line aa
gives the wave length of the primary radiation. There are seen
to be two maxima, or lines, one of which P corresponds exactly
to the primary radiation; the other S is "shifted" toward
longer wave lengths from the position of the primary, the differ-
ence in wave length between
a
the shifted and the unshifted com-
ponents being 0.0236 Angstrom. Classical theory could not
account for the presence of this shifted component. Its explana-
tion by Compton is one of the most striking applications of the
quantum theory.
Compton boldly applied the extreme quantum picture of
radiant energy, according to which a beam of radiant energy
consists of a stream of localized "pellets," i.e., quanta, of energy
of magnitude hv. These quanta, since they are in motion (with
the velocity of light) and possess energy, must also possess
momentum. When,~ therefore, they collide with material par-
ticles, such as atoms or electrons, it might be expected that the
collisions would be governed by the two laws of mechanics:
(1) the conservation of energy and (2) the conservation of
momentum.
We must first compute the momentum of a quantum which
possesses energy hv and is moving with velocity c. Proceeding
as in the computation, on the basis of the kinetic theory of gases,
of the pressure produced by the molecules of a gas, we first com-
pute the momentum which a stream of quanta, incident onto
and absorbed by a surface, communicates to that surface in 1
second, since such transfer of momentum per square centimeter
per second is equal to the pressure produced by the incident beam
of radiation. Let I be the intensity of the incident beam of
radiation, assumed monochromatic and of frequency v. (I is
the energy contained in a volume of the radiation 1 sq. cm. in
cross-section and c cm. long.) The total number N of quanta,
each of energy hv, contained in this volume, is, then,
N = !_
hv
Let the momentum of each quantum be p. Since these N quanta
are incident onto 1 sq. cm. of the surface in 1 second, the total
momentum P transferred from the radiation to the surface per
square centimeter per second is
I
P = hvp
522 X-RAYS [CHAP. XII
But by the ordinary laws of mechanics, Pis numerically equal to
the pressure which the radiation exerts on the surface, and we saw,
in Chap. VII, Sec. 6, that, irrespective of any theory as to the
nature of radiation, the pressure exerted by radiation falling
normally onto a surface is numerically equal to the energy
density VI of the radiation. Therefore, P = VI and
I
p·-=VI
hv
But i/; is obviously given by
I
if; = -
c
I I
·. · p hv = c
from which it follows, at once, that
hv
p= (37)
c
That is, the momentun p of a quantum is obtained by dividing
the energy hv of the quantum by the velocity of light c.
,,~~0
~~,
' ' ',
~.e Ve, ,
~ .,., '
~e
_ _ _E..:_:_h_2"_o_ _ _~ - ~ ________' '~
0
i. a /'sfi_,A /
Po= ·"(o ~' /
~ ~('~ /
/~.,,,. /
'?,>~~~/
Fm. 151.-Vector diagram showing the conservation of momentum when a
quantum is scattered by an electron.
Having now the energy of the quantum hv and its momentum
hv/c, we can apply to the collision between a quantum and an
electron the laws of the conservation of energy and of momentum,
in exactly the same way as we should solve the problem of the
elastic collision of two billard balls.
Let an incident quantum, energy Ea = hvo and momentum
Po = hva/c, collide with an electron a (Fig. 151), mass m, and
initially at rest. After the collision, the electron will move_ in a
SEC. 15] THE COMPTON EFFECT 523
direction making an angle of ¢, say, with the initial direction of
motion of the quantum, while the quantum itself will move in a
direction making an angle () with its initial direction. Qualita-
tively, one sees, at once, that the energy E 0 of the quantum after
collision and, therefore, also the frequency, must be less than that
of the quantum before collision, since some of the energy of the
incident quantum must have been given to the electron if it has
been set in motion.
We can now write the equations for the conservation of energy
and momentum. Let
Ea = hv 0 = the energy of the quantum before collision
E 0 = hv 0 = the energy of the quantum after collision, the direc-
tion of motion making an angle () with the initial
incident direction
1
Em = mc 2 ( - 1 2
- 1) = the kinetic energy 1 of the electron
v 1 - fj
p 0 = hvo = the momentum of the quantum before collision
c
p0 = hvo = the momentum of the quantum after collision
c
Pm = V 1mv- f32
== the momentum of the electron 1
fj = v/ c, where v is the velocity of the electron
From the law of the conservation of energy, we write,
Eo = Eo + Em
i.e.,
hvo = hvo + mc 2
(v 1 - 1)
1 - fj2
(38a)
1 The theory of relativity leads to the following expression for the kinetic
energy Ek of a mass m moving with velocity v:
Ek= mc 2 (y1-1- - 1)
1- (32
where {3 = v / c is the ratio of the velocity of the body to the velocity of light.
Our ordinary expression for the kinetic energy, i.e., }2mv 2, is the first term in
the expansion of this equation and is sufficiently accurate for {3 < < 1.
Since v = {3c, we may write the equation for momentum p:
m{3c
p=
v'1 - {3 2
where V m gives the variation of mass with velocity.
1 - (32
!J24 X-RAYS [CHAP. XII
From the law of the conservation of momentum, we can write for
the x-components of momentum (i.e., parallel to Po),
-hvo = -hvo cos (} + m(3e cos '¥,1,. (38b)
e e y 1 _ (j2
And for the y-components,
hv 0 • m(je .
0=-smO+_; Slll<p (38e)
e -v 1 - (j2
In these three independent equations, we have, given vo the
frequency of the incident quantum, the four unknown quantities
v 0 , O, ¢, and {3. By simultaneous solution, we can obtain three
of them in terms of a fourth, say e, which is the direction with
respect to the incident beam in which the quantum is" scattered''
as a result of the particular collision.
To solve these equations, set
hvo h hv 0 h
me 2 = µo = me).) me 2 = µo = me'Ao
where Xo and 'A. 0 are the wave lengths (whatever that may mean in
terms of this "picture"!) of the incident and the scattered
quantum. By slight rearrangement, equations (38) then take
the following respective forms:
µo - µo = y' 1 1- (j2
- 1 (39a)
µo - µ9 COS (J = _I (3 COS <f:, (39b)
-v 1 - (j2
. (} (3 .
µosm =-_ 1 2
sm¢ (39e)
-v 1 - (j
Squaring equation (39a) and rearranging gives equation (40);
and squaring equations (39b )and (39e) and adding gives equation
(41), viz.,
2 1
2
µ0+ 2
µ9 - 2µ0µ9 +
y' 1 - (32 - l _ (3 2 + 1 (40)
µ 02 + µo2 - 2µoµo cos (} = (32
_ (j 2 (41)
1
Subtracting equation (41) from equation (40) gives, by aid of
equation (39a),
µo - µo = µoµo(l - cos (}) (42)
1 1
. ·. - - - = (1 - cos e)
µo µo
h
. ·. Ao - Ao = -(1 - cos(}) (43)
me
SEC. 15] THE COMPTON EFFECT 525
by replacing the equivalent of µ. Putting m the numerical
values of h, m, and c gives
. Ao - Ao = dA 0 = 0.0243(1 - cos 0) (43')
0
where A is expressed in Angstroms.
This important equation states that when an incident radiation
of wave length Ao is scattered at an angle O, the wave length Ao of
the scattered radiation should be greater than that of the incident
radiation by the quantity 0.0243(1 - cos 0), which, for a given
angle O, is constant whatever the incident waue length. When the
scattering angle is 90 degrees, the "shift" dA90 in wave length
between the primary and the scattered beam should be
0
dA90 = 0.0243 Angstrom
exactly in accord with Compton's measurements, quoted on
page 521. Further, the shift should be entirely independent of
the material of the sea tterer and for various angles O of sea ttering
should be proportional to (1 - cos 0).
-(d) These predictions have been abundantly confirmed by
the investigations of Compton and his collaborators,1 Ross, 2
J. A. Becker/ Allison and Duane,4 and others. For details,
the student is referred to the original articles.
By use of equations (42) and (38a), it is readily found that the
energy Em of the recoiling electron is
E = µ(l - cos 0)
hvo (44)
m 1 + µ(l - COS 0)
and from the construction of the Fig. 151, it is evident that
the electron must move in the forward direction, i.e., </> < 1r/2.
Following Compton's prediction of their existence, these ''recoil''
electrons, as they are called, were discovered by C. T. R. Wilson 5
and W. Bothe. 6 They have been studied quantitatively by
Compton and Simon 7 and others. Bless, 8 by the method of the
1 Phys. Rev., vol. 21, pp. 207, 483, 715 (1923); Proc. Nat. Acad. Sci., vol.
10, p. 271 (1924).
2 Proc. Nat. Acad. Sci., vol. 9, p. 246 (1923); Phys. Rev., vol. 22, p. 524
(1923).
3 Proc. Nat. Acad. Seri., vol. 10, p. 342 (1924).
4 Proc. Nat. Acad. Sci., vol. 11, p. 25 (1925); Woo, Y. H.: Phys. Rev.,
vol. 27, p. 242 (1926).
5 Proc. Roy. Soc., (London) vol. 104, p. 1 (1923).
6 Zeit. fur Physik, vol. 20, p. 237 (1923).
7 Phys. Rev., vol. 25, p. 306 (1925).
8 Dissertation, Cornell University (1927). Phys. Rev., vol. 29, p. 918
(1927).
526 X-RAYS [CHAP. XI
magnetic "spectrograph," similar to Robinson's method for
photoelectrons, has measured the value of Em for the recoil
electrons and has found that the results are in agreement with
Compton's equation (44).
The presence of the so-called "unmodified" line (P, in Fig.
150) may be accounted for as follows:In setting up equations (38),
it was assumed that the mass with which the quantum collides
is that of a "free" electron or, at least, that the kinetic energy
Em given to the electron by the collision is sufficient to remove
the electron from an atom in which the electron may be "bound."
If, however, the energy required to remove the electron is greater
than Em, the electron will not be removed from the atom, and we
may regard the collision as taking place between the electron
and the atom as a whole, the mass of which, M, is far greater than
that of the electron. If M be substituted for m in equation
(43), it is readily seen that the computed change of wave length
is so small as to be quite beyond the possibility of detection.
A certain proportion of collisions between quanta and the
electrons of a scattering substance will involve electrons so loosely
bound that they are ejected from the atom; a change of wave
length in the scattered radiation then results. In other collisions,
the energy given to the electron will be insufficient to remove the
electron from the atom, and no change of wave length results.
The scattered radiation should, therefore, contain both the
shifted and the unshifted lines, exactly as Compton found.
It can be shown that the relative intensities of these components
should depend on the wave length of the incident radiation and
on the atomic number of the scatterer. These predictions are
also confirmed, qualitatively, by experiment.
(e) This radical picture of the quantum structure of radiation,
which is so successful in explaining the Compton effect, finds
curious confirmation when one computes, by the application of
the same concepts, the radius of the electron from the measured
value of the magnitude of the X-ray scattering coefficients.
We may "explain" the phenomenon of the scattering of X-rays
by assuming that some of the quanta of an incident beam in
passing through matter are deflected from their paths by colli-
sions with electrons, the remaining quanta proceeding through
the scattering material undeviated. If the electron is a sphere
of radius a, we may think of it as obstructing the free passage
of quanta over an area 1ra 2 and that a quantum is deviated, or
SEC. 15] THE COMPTON EFFECT 527
scattered, whenever it hits this area. Let an incident beam of
radiation be of such intensity that Q quanta per second strike
a square centimeter of a thin slab of absorbing material of
density p and thickness dx, which contains n electrons per cubic
centimeter. Then the fraction dQ/Q of the Q quanta which are
scattered is evidently the same as the ratio of the total projected
area dA of all the electrons in the square centimeter to 1 sq. cm.,
i.e.,
Proceeding exactly as in the derivation of the mass absorption
coefficient in equation (2), we find that the mass scattering
coefficient er/ p should be
p p
Now, the measured value of the mass scattering coefficient for
carbon is approximately 0.15. And n/ p, the number of electrons
per gram (p = density) of carbon, is
~o Z = 6.06 ~ 1023 X 6 = 3.03 X 1023
\_where No= Avogadro's number, A =
atomic weight, Z =
atomic number. Putting these values in the above equation
and solving for a, the radius of the electron, gives
a = 4 X 10- 13 cm. (about)
From equation (22) of Chap. V we have another method of
computing the radius of an electron on the assumptions that the
electron is a sphere of radius a and mass m with its charge e
uniformly distributed over the surface of the sphere-admit-
tedly a very highly artificial concept. Then
2 e2
m = 3 ac 2
Knowing m ( = 9 X 10- 28 grams) and e ( = 4.77 X 10- 10 e.s.u.),
we can compute a. We find
a = 2 X 10- 13 cm. (about)
Although these two values of a, computed from absolutely
dissimilar data, are not equal, it is quite remarkable that they are
528 X-RAYS [CHAP XII
at least of the same order of magnitude. The quantity a cannot,
in all probability, have the precise physical meaning which either
of these two computations assign to it. Yet it is a quantity which
has the dimensions of a length and is, in some way, associated
with the electron. How-we must leave to future research.
Further, we may raise the question: Precisely under what
circumstances does a collision between a quantum and an atom
(or an electron) result in the "scattering" of the quantum, as
contrasted with the circumstances under which a collision results
in the photoelectric expulsion of the electron? In the former
case, only a small part (or none at all) of the energy of the
quantum is lost; in the latter case, the entire energy of the
quantum disappears. It is probable that an unambiguous
answer to this question can be made only after we have a much
more fundamental knowledge than we now possess of the nature
of radiation and of the atomic mechanism.
16. The Refraction of X-rays. (a) Measured by Use of
Bragg's Law.-The discovery of the refraction of X-rays and the
subsequent measurement of indices of refraction are excellent
mustrations of the great advances which have been made in
the technique of X-ray measurements since the discovery of the
action of the crystal grating. One of the first observations
recorded by Roentgen was that no measurable refraction of
X-rays was found. Subsequent, and more precise, measurements
seemed to confirm this result. For example, Barkla,1 in a very
careful investigation, failed to detect any refraction oand con-
eluded that the refractive index for wave length 0.5 Angstrom
differed from unity by less than 0.000006.
The first positive evidence that X-rays are measurably
refracted came from the work of Stenstrom, 2 who showed from
accurate measurements of wave length that Bragg's law of the
reflection of X-rays from crystals
nA = 2d sin (}
does not yield identical values when the wave length of a given
line is computed from different orders of reflection. Hjalmer3
1 Phil. Mag., vol. 31, p. 257 (1916).
2 Dissertation, LUND (1919).
3 Quoted by S1EGBAHN, "The Spectroscopy of X-rays" (English transla-
tion), p. 22.
SEC. 16] REFRACTION 529
found, for example, that the wave length of the Fe Ka 1 line as
measured in the first order by reflection from a gypsum crystal
(d = 15.155 Angstroms) was 1.9341 Angstroms;
o
while measure-
ments in the sixth order gave 1.9306 Angstroms, nearly 0.2 per
cent less.
This apparent failure of the Bragg formula was shown to be
due to the refraction of the beam of X-rays as it entered the
crystal, the deviation of the beam being such as to indicate that
the index of refraction is less than unity. In Fig. 152 is shown the
path of a ray as it enters the surface SS of a crystal at a glancing
@' 8, RerlecHn Plan(? p
p
of Afom.s
Frn. 152.-Refraction of a beam of X-rays entering the surface of a crystal,
showing the reason for the apparent failure of Bragg's law.
angle (} and is incident onto the plane of atoms PP at an angle O'.
Bragg's law in the form
n'J./ = 2d sin o'
where 'J....' is the wave length in the crystal, gives the true law of
reflection at the crystal plane. Measurements give O, however,
and it is 'J...., the wave length in air, that is desired. It can be
shown 1 that the correct wave length 'J.... can be obtained by the
use of a slight correction factor to Bragg's equation, viz.,
. (} [ 1 -
n'J.... = 2d sm 1sin-
2
µ]
0
(45)
where µ is the index of refraction of the X-rays in the crystal.
This equation contains the unknown quantities 'J.... and µ. By
making measurements in two different orders n1 and n2, it is
possible to solve simultaneously for 'J.... and µ. In this way,
Hjalmer, 2 for example, obtained the following data for the
quantity o = 1 - µ for gypsum:
1
See COMPTON, A. H.: "X-rays and Electrons," p. 212.
2
Ann. Physik, vol. 79, p. 550 (1926).
530 X-RAYS [CHAP. XII
TABLE X.-lNDEX OF REFRACTION, µ = 1- o, OF GYPSUM (HJALMER.
loc. cit.)
X, o, 2
.Angstroms times 10 6 X
times 10 6
1.537 9.0 3.85
1.932 15.4 4.11
2.498 23.1 3.70
3.025 27.5 3.00
3.378 37.9 3.32
3.926 60.4 3.92
4. 718 74.9 3.37
5.166 82.6 3.10
Classical theory gives for the index of refraction µ of a S1ub-
stance for waves of frequency 1 v
(46)
where ni is the number of electrons per unit volume which have
frequency Vi and the summation is to be extended to all the
different groups of electrons in the refracting material. If, as
in X-rays, (1) the index of refraction differs from unity by only
a small quantity o, as shown by Hjalmer's results, and (2) the
frequency v of the incident radiation is very much higher than
any of the frequencies vi of the refracting substance,2 we may,
neglecting v! in comparison with v2, write equation (46) in the
form
ne 2
µ = 1 - 2 (47)
21rmv
where n is the total number of electrons per unit volume. Since
µ = 1 - o, we have
ne 2
o= 21rmv2 ( 48)
or, changing v into wave length X,
"\
o - ne 2 = a constant (49)
I\
2
21rmc 2
1 See, also, equation (94), Chap. IV, which is the same formula expressed
in slightly different symbols.
2
This is true for X-rays of moderately short wave length and refracting
materials of not too high atomic number.
SEC. 16] REFRACTION 531
That is, the quantity o/'A. 2 should be a constant as long as the
frequency of the refracted X-rays is much greater than the fre-
quency of any groups of electrons within the refracting substance.
The measurements of Larson 1 have shown that for mica the
qu::\ntity o/'A. 2 is substantially constant over a wave-length range
1.5 < 'A < 8.3 Angstroms. This constancy, however, should
not hold in the neighborhood of a critical frequency vi. Hjal-
mer's measurements for the index of refraction of gypsum,
CaS0 4 , given in Table X, cover a wave-length range which
includes the K limit of Ca and the K limit of S. If the values
of o/'A.2, given in the table, be plotted against wave length, a
graph of the type shown diagrammatically in Fig. 153 results,
unique points or discontinuities apparently occurring at the wave
Ca Klimif S Kl/mH .A
Fw. 153.-Variation of index of refraction through the K absorption limits of
the a toms in a crystal.
lengths 'A.1 and 'A2, which correspond approximately to the wave
length of the calcium K absorption limit ('A = 3.06 Angstroms)
0
and the sulphur K absorption limit ('A = 5.01 Angstroms).
On the basis of the classical theory, such discontinuities as are
observed in Fig. 153 indicate resonance phenomena, it being
assumed that there are present in the refracting medium resonat-
ing systems the natural frequencies of which correspond to the
discontinuities in the curve. These critical frequencies shown in
Fig. 153, however, correspond to the K absorption limits of Ca
and S and not to the frequencies which these substances radiate,
which frequencies are the K series of lines. It might be
expected from classical theory that these latter frequencies should
be identical with the resonating frequencies. But such is not
the case. Again, a formula derived on classical theory and, in
part, successful comes into conflict with experiment.
1 Zeit. far Physik, vo1. 35, p. 401 (1926).
532 X-RAYS [CHAP. XII
(b) Direct Measurement by a Prism.-Although the index of
refraction for X-rays differs only very slightly from unity, its
direct measurement by a prism has been accomplished by several
observers. Among the most interesting measurements are those
of Davis and Slack, 1 whose method is instructive. A beam of
X-rays from a target is reflected from a crystal K1 (Fig. 154) at
such an angle that a monochromatic beam, say the characteristic
is--~--.-~~-,-~~-r-~--,r--~--,
_L_
.,.
/I
,,
/ IP
-.J
\
\
\
\
\
OO 4 8' 12' 16~ 20"
Angular Pos·1tion of Crystal Ki (in seconds of arc),
A~ Without Prism in PC4th of Beam
B~ With Prism in Path of Beam
Frn. 154.-The "double" X- Frn. 155.-The shift in a spectral line
ray spectrometer of Davis and produced by refraction through a prism
Slack for measuring index of re- (Davis and Slack) .
fraction of X-rays by the prism
method.
Ka1 line of the target, passes through a slit S and falls onto a
second crystal K2. When K 2 is exactly parallel to K1, this beam
is reflected from K2 and enters the ionization chamber I. If
K2 be turned in either direction from exact parallelism, the ioniza-
tion current decreases rapidly to zero, the apparent "shape" of
the line, or beam, being shown as A (Fig. 155), in which the
abscissre are seconds of arc from an arbitrary zero. The point a
can be located very exactly.
1 Phys. Rev., vol. 27, p. 18 (1926).
SEC. 16] REFRACTION 533
If, now, a prism P be placed in the path of the beam, as shown,
the beam is deflected-in the opposite direction to that in which
a beam of light is deflected, since µ < 1. The angle through
which the beam is deflected in its passage through the prism
can be determined by observing the new position in which the
crystal K 2 must be placed to reflect the beam. The shift in the
position of the line A, which is the Ka1 line of Mo (A = 0.7078
Angstrom), when a prism of aluminum with a refracting angle
of 166 degrees is placed in the beam, is shown as B (Fig. 155).
The angular displacement L\. between A and B is 5.85 seconds.
From the ordinary law of refraction by a prism, it follows that the
index of refraction, when, as in this case, L\. is very small, is
µ = 1 + sm. 2L\. cot
()
2
. L\. ()
.·. o = 1 - µ = - sm
2 cot
2
where (} is the angle of the prism. The results obtained by
Davis and Slack for a prism of aluminum for the two wave lengths
0
A = 0.7078 and A = 1.537 Angstroms are as follows:
X0 = 0.7078 X
0
= 1.537
ANGSTROM ANGSTROMS
Angle of prism. . . . . . . . . . . . . . 166 degrees 116 degrees
Mean angle of deviation. . . . . 5. 62 seconds 5. 53 seconds
o= 1 - µ,observed ........ 1.68 X 10- 6 8.4 x 10- 6
o,computed ................ 1.77 X 10- 6 8.36 X 10- 6
The computed values of o are obtained by use of equation (49)
ne 2 A2
0 =-- (49')
21rmc2
where n, the number of electrons per unit volume of aluminum,
is computed from the known values of Avogadro's number N 0 , the
atomic weight A, atomic number Zand the density of aluminum
p, VIZ:
(50)
The computed values of o are seen to be in excellent agreement
with the observed, thus, apparently, confirming the classical
theory of refraction on which equation (49) is based.
534 X-RAYS [CHAP. XII
Subsequently, Davis and Slack 1 reported measurements on the
index of refraction of several substances for the Mo Ka1 line
0
(A = 0. 7078 Angstrom) as follows:
REFRACTING
ELEMENT o= 1 - µ,
Ag ......... . 5.85 x 10-6
Cu ......... . 5.95
S .......... . 1.39
Al ......... . 1.68
C .......... . 1.23
(c) Measurement by Total Refiection.-Since the index of refrac-
tion of a material for X-rays is less than unity, a beam of X-rays
Si
I
Frn. 156.-Doan's apparatus for measuring the index of refraction of X-rays
by the method of total reflection.
incident onto a polished surface at a sufficiently small glancing
angle should be totally reflected. One of the first measurements
of X-ray refraction was by A. H. Compton,2 using this method.
The measurements of Doan 3 are jnstructive.
Doan's apparatus, following Compton's method, is shown
schematically in Fig. 156. A beam of X-rays from a target T,
after passing through a slit 81, falls onto a crystal K, by means of
which one of the characteristic lines in the spectrum of T may be
reflected at a very small glancing angle onto the mirror M.
With a sufficiently small angle of incidence, this beam is totally
reflected from M and falls onto a photographic plate PP at R.
M is very slowly turned during an exposure; and when the
critical angle is reached, reflection ceases. This critical angle
1 Phys. Rev., vol. 27, p. 296 (1926).
2
Nat. Research Council, Bull. 20, p. 50 (1922).
3
Phil. Mag., vol. 4, p. 100 (1927).
SEC. 16) REFRACTION 535
can be determined by noting the distance on the plate between
the point 0, which indicates the position of the direct beam, and
the extreme edge of the record of the reflected beam. The
metals under study were sputtered onto a highly polished optical
glass surface.
The following are some of the results obtained by Doan:
TABLE XL-INDEX OF REFRACTION OF X-RAYS BY THE METHOD OF TOTAL
REFLECTION (Do AN, lac. cit.)
Wave length,
0 Substance Critical angle
o= 1 - µ,
Angstroms times 106
0.7078 Ni 10' 15" 4.42
0.7078 Ag 11' 42" 5.76
1.389 Ni 16' 9" 10.98
1.389 Cu 19' 36" 16.2
1.537 Ni 24' 40" 25.5
1.537 Cu 20' 24" 17.7
1.537 Ag 26' 42'' 30
1.537 Au 31' 24" 41.6
A comparison of Doan's value of o for Ag for A = 0.7078 Ang-
strom with that obtained by Davis and Slack shows excellent
agreement between the two methods.
Much more experimental work remains to be done before suffi-
cient data on indices of refraction are available to serve as an
adequate guide to theories of absorption, scattering, and refrac-
tion of X-rays, all three phenomena being more or less inter-
related. It is very significant, however, and is suggestive
of future possibilities, that experimental data on the refraction
of X-rays yield, by aid of equations (49') and (50), a value for the
number of electrons per atom which is in exact agreement with
that obtained by other methods, as is evidenced by the fact that
the observed values of o agree with those computed by these
equations.
CHAPTER XIII
THE NUCLEUS
In Chap. X, we referred to Rutherford's experiments on the
scattering of a particles and to the hypothesis of the nuclear type
of atom which these experiments suggested. This was the first
evidence of the existence of nuclei within atoms. The subse-
quent development of the Bohr atom model and its use in explain-
ing the origin of spectral lines seemed to confirm Rutherford's
hypothesis and gave some information concerning the mass
of, and the charge associated with, the nucleus: Practically the
entire mass of the atom is contained in the nucleus, and its
charge is equal to +ze, where Z is the atomic number of the
atom and e is the charge on the electron. The data of spectros-
copy and chemistry, however, yield no further information
regarding the nucleus. Our present knowledge of the nucleus
comes, in large part, from two other fields: (1) radioactivity
and (2) the precision measurements of the masses of atoms by
the so-called "mass spectrograph" developed by Aston in
England and by Dempster in America. In this chapter, we shall
give a very brief survey of these two fields, leaving it to the reader
to make such further study of the literature, preferably from
original sources, as the importance of the subject warrants.
p ART I. THE MASSES OF ATOMS
1. Positive Rays.-Measurements of the masses of atoms
are made by observing the deflection produced on positively
charged ions of the substance under study by the combined
action of an electric and a magnetic field, the principles of the
methods used being somewhat similar to the method by which
J. J. Thomson first measured the value of e/m for electrons.
These positive ions in motion are frequently called positiverays.
Two sources of positive rays are customarily employed: (1) the
canal rays (see Chap. IX, Sec.15 (b)) originating near the cathode
of a tube containing gas at a low pressure through which an
electrical discharge is passing; and (2) the positive ions emitted
536
SEC. 1] POSITIVE RAYS 537
by salts when heated 1 under certain conditions. The first
source is used in studying those substances which can be con-
veniently introduced into the discharge tube in gaseous or vapor
form. Th~ second is used when the substance is available only
in the solid state.
For the early work on the deflection of positive rays by the
combined action of electrostatic and magnetic fields, the student
is referred to the original papers of J. J. Thomson. 2 Thomson's
method is shown diagrammatically in Fig. 157. B is a large
discharge tube, the cathode K of which is perforated with a very
small hole through which pass the positively charged particles
originating in the region immediately in front of K. These
positively charged particles emerge from the opposite end of
K as a narrow bundle of canal rays, the velocity of which depends
B
J
K J
FIG. 157.~Thomson's positive-ray spectrograph.
on their charge and mass and on the potential V applied to the
tube. JJ is a waterjacket for cooling the cathode. M1 and M 2
are the poles of an electromagnet, of which A1 and A2 are the
soft iron pole pieces electrically insulated from M 1 and M 2 by
insulating strips I. A1 and A2 can thus be maintained at any
desired potential difference, so that in the space between A 1 and
A2 we may have a magnetic field and, parallel thereto, an electric
field. A positively charged particle moving from left to right
in this space will experience a deflection in the plane of the paper,
due to the electrostatic field, and at right angles to the plane of
the paper, due to the magnetic field. After leaving the space
between A 1 and A 2, the deflected particle moves in the field-
free, evacuated space inside the "camera" C and falls on the
1See RICHARDSON: "The Emission of Electricity from Hot Bodies."
2Phil. Mag., vol. 13, p. 561 (1907); vol. 20, p. 752 (1910); vol. 21, p. 225
(1911); vol. 24, p. 209 (1912); THOMSON, J. J.: "Positive Rays of Electricity,"
2d ed. (1921).
.538 THE NUCLEUS: MASSES OF ATOMS [CHAP. XIII, PART I
photographic plate PP. Thomson found that the plate, after
development, showed a series of parabolas, the constants of which
were dependent on the ratio E / M, where Eis the charge carried
by the particle (E must be, of course, an integral multiple of
the charge e) and M is the mass of the particle.
Let a particle, as it enters the space A1A2, be 'moving with
velocity v, initially parallel to and coincident with the x-axis
of a system of rectangular coordinates with origin in the space
A1A2. Let the mass of the particle be M and its charge E.
Also, let Y and {3 be, respectively, the electrostatic and the
magnetic fields, and S the length of the path through the fields,
it being assumed that the fields are constant over the length S
and that they terminate sharply. As a result of the passage
through the field, the particle will experience a deflection y'
(from the x-axis), due to the electric field, given by
y' = i2 YE
M v
(S) 2
(la)
and a similar deflection z' in the z direction due to the magnetic
field, assuming that the deflection is small compared to the radius
of the circular path, given by
z' = ! {3E 8 2 (lb)
2 Mv
After leaving the fields, the particle moves in a straight line.
If the distance of the photographic plate PP from the space
A1A2 is large compared to the length of the path S, the point
where the particle strikes the plate will have coordinates y and
z which are, respectively, proportional to y' and z'. It follows
readily that a relation between y and z is given by (eliminating
v from equations (1))
E fj2
22 = C - - y (2)
MY
where C is a constant depending on the dimensions of the appa-
ratus. This is the equation of a parabola. Accordingly,
particles having various velocities v as they enter the field space
A1A2, but having the same ratio E/M, should make a parabolic
trace on the plate, as Thomson found. For the same value of
the fields /j and Y, particles with a different ratio of E / M should
produce on the plate another parabola. The fact that these
SEC. 2] ISOTOPES 539
traces found by Thomson were reasonably sharp indicated that
atoms of a given kind all have the same mass.
By a systematic study of the relative positions of the various
parabolic traces appearing on a series of plates, Thomson was
able to determine the origin of the traces. Traces due to H+,
Ht, o+, O~, co+, etc., were identified. Knowing the masses of
these "standards," the masses of atoms producing other traces
could be determined. Thomson found that when neon, atomic
weight 20.2, was introduced into the discharge tube, instead of a
single trace, there were two traces corresponding to atomic weights
20.0 and 22.0 and suggesting that there are two kinds of neon
atoms, one having atomic weight 20.0 and the other having
atomic weight 22.0. A direct method was thus available for
measuring the masses of atoms. This method was substantially
modified and rendered more precise by Aston, whose "mass
spectrograph" has very greatly extended our knowledge of atoms.
2. Isotopes. (a) Aston's "Mass 8pectrograph."-Astons'
improvements 1 in the methods of positive-ray analysis consist
p
Po.s/-1/ve
Fia. 158.-Schematic arrangement of Aston's "mass spectrograph."
(1) in securing greater _dispersion and (2) in bringing all ions
with a given ratio E / M to a focus (instead of spreading them out
into a parabola), thereby securing greater sensitivity. The
principle employed by Aston is represented in Fig. 158. The
positive rays from a discharge tube (not shown) pass through a
narrow slit 8 1 in the cathode, thence through a second slit 82, from
which they emerge into the space between the metal plates P1
and P 2 between which can be maintained an electric field of
any desired intensity. This field causes a deflection of the par-
ticles toward P 2 , the amount of the deflection being greater the
less the velocity of the particles. Since the positive-ray stream
1 Aston's papers are found in the Phil. Mag., vols. 38 to 49 (1919 to 1925);
see also Nature, vol. 116, p. 208 (1925) and vol. 117, p. 893 (1926). AsToN's
"Isotopes," 2d. ed. (1924), also gives an account of the subject.
540 THE NUCLEUS: MASSES OF ATOMS [CHAP. XIII, PART I
contains a wide range of velocities, the stream will be broadened
as it passes through the field, the more swiftly moving ions
passing through the wide slit 83 on the side a and the more slowly
moving ones passing on the side b. After leaving S3, this diverg-
ing stream of ions enters a magnetic field at right' angles to the
plane of the paper, maintained between the circular pole pieces
M of an electromagnet. This magnetic field causes deflections,
as shown, the more slowly moving ions being deflected more than
the faster ones, the result being that when the ions emerge from
the magnetic field they are moving in converging directions in
such a way that they are brought to a "focus" at some point L.
It turns out that, with suitable design of the apparatus, the locus
of point L for ions having various ratios of E / M is nearly a
straight line passing through point 0. Consequently, a photo-
graphic plate placed in position PP will record a number of
"lines," each line corresponding to a particular ratio E/M. In
part, from the known dimensions of the apparatus and, in part,
by introducing into the discharge tube certain known substances
-H2, 02, CO, C02, etc.-as calibrating standards, the masses of
other ions can be determined.
The lines produced by such ions as O+, ot, CO+, Nt, cot,
which are made up of atoms the atomic weights of which are, as
measured, exact integers (e.g., 16.00 for oxygen; 12.00forcarbon;
etc.) are, in each case, single lines, indicating that the atoms of
each of these substances all have the same mass. When, how-
ever, neon, the atomic weight of which, namely, 20.2, departs
measurably from an integer, is introduced into the tube, no line
corresponding to the mass 20.2 is found, but, instead, there are
two lines, the one, stronger, corresponding to a mass of 20.0, and
the other, weaker, to a mass of 22.0. Unmistakably, there are, as
Thomson showed, two kinds of neon atoms which differ from each
other only in having different masses, which masses are integers
when the mass of the oxygen atom is taken as 16.00.
Similarly, when chlorine is introduced into the tube, no line is
observed which corresponds to the chemically determined weight
of chlorine, namely, 35.46, but, instead, there are two lines corre-
sponding to the integer values 35.0 and 37.0, the former being the
more intense line. There are, thus, two kinds of chlorine atoms
differing from each other in atomic weight but being identical as
regards all their other chemical and physical properties. Hence,
they are called isotopes. Chlorine is, thus, a mixture of two kinds
SEC. 2] ISOTOPES 541
of atoms, of respective masses 35.0 and 37.0 in such proportion
that the average mass per atom of a large number of atoms-
which is the mass given by chemical determinations-is 35.46.
(b) Dempster's Method.-Dempster's method of measuring the
masses of atoms differs from Aston's in that (1) the photographic
plate is replaced by an electrical method of measurement and (2)
the disposition of the electric and magnetic fields is materially
different. Dempster's apparatus 1 is shown schematically in
Fig. 159. An electrically heated metal cylinder A, which serves
as anode, has on its front surface f a salt of the element under
study. This heated salt is bombarded by electrons from the hot
wire ww, which is heated by the battery B1, the wire being main-
a
w
4~-"\;?- w
+H-
B~
c
FIG. 159.-Dempster's "mass spectrograph."
tained at a potential difference V 2 of - 30 to - 60 volts with
respect to A by means of the battery B2. Dempster found that,
when so bombarded, the anode emitted positively charged ions.
These ions, after passing through the slit 81, are accelerated
toward the slit 82 by means of the potential difference V 3 of some
800 to 1,000 volts. On emerging from the slit 82, the ions enter
the space abcdef in which is maintained a uniform magnetic
field, perpendicular to the plane of the paper, by means of which
the ions are caused to move in a circular path toward the slit 8 3 ,
the radius r of the circle being determined by the velocity v of the
ions as they enter 82, their mass M, and charge E, as well as the
1
For Dempster's articles, see Phys. Rev., vol. 11, p. 316 (1918); vol. 18,
p. 415 (1921); and vol. 20, p. 631 (1922).
542 THE NUCLEUS: MASSES OF ATOMS [CHAP. XIII, PART I
intensity of the magnetic field H, according to the well-known
equation
Mv 2
HEv=-
r
When the ions move in circles defined by the three slits 8 2 , 8 3 , and
84, they pass through 84 and fall upon the metal plate p, which
thus acquires a positive charge at a rate which can be deter-
mined by the electrometer E. (Dempster used a sensitive elec-
troscope.) For a given ratio of E / M and field H, the value of r
+c
Q)
L.
c..
::)
(..)
Atomic Mass
16:10 16.00 15.90 15.80
mo rn o 1740 nso
Potential Difference(V3) 1n volts
Frn. 160.-The oxygen (O+) line with Dempster's apparatus.
and, therefore, of the current I registered by the electrometer
depends on the potential difference V 3 • A curve plotted between
I and Vs shows sharp maxima, each of which corresponds to a
definite value of E / M. In part, from the known constants of
the apparatus and, in part, by use of known ions, Dempster was
able to identify these maxima with definite ions. A sample of
the data for the ion O+ is shown in Table I. Figure 160 shows
graphically the relation between V3 and I. Column three of
Table I gives the computed value of M (in terms of O = 16).
Of course, the finite width of the graph is due to the necessity of
employing finite slit widths, since all the ions entering the slit 84
SEC. 2] ISOTOPES 543
TABLE I.--DEMPsr.rER's DATA FOR THE loN o+
Va M
I
(volts) (computed)
1,758 7.3 15.76
1,752 7.3 15.81
1,746 15.6 15.87
1,740 38.6 15.92
1,734 51 15.98
1,728 58 16.03
1,722 37 16.09
1, 716 21 16.14
1, 710 10 16.20
0
5.9 6.0 6.1 6.8 6.94- 7.0 7.2
Atomic Weights
Frn. 161.-The isotopes of lithium, as measured by Dempster.
in this experiment were o+ ions. The maximum of the curve
corresponds to the weight of the oxygen atom.
When employing a salt containing lithium, Dempster found
two maxima in the neighborhood of the atomic weight of lithium,
6.94, as shown in Fig. 161, in which the abscissre are the values
of M, the computed mass of the ion, as determined semiempiri-
cally from the constants of the apparatus. The one maximum A
corresponds within the limits of error of measurement to atomic
weight 7.00; the maximum B corresponds to 6.00. There is no
indication of atoms of lithium with weight 6.94, the chemically
determined atomic weight of lithium, which is indicated in Fig.
161 by the, dotted line. Dempster, by this method, has analyzed
544 THE NUCLEUS: MASSES OF ATOMS [CHAP. XIII, PART I
a number of elements-magnesium, lithium, calcium, and zinc-
and has found them to be made up of isotopes, the masses of
which are, within the limits of error of measurement, integers in
terms of the mass of the oxygen atom as 16.
(c) Isotopes of the Elements.-With few exceptions, practically
all of the more important elements have now been studied, and
many of them have been found to consist of mixtures of atoms,
i.e., isotopes, which have very nearly integer atomic masses.
The isotopes for the various elements are shown in Appendix
I (b), in connection with the periodic table of the elements. Some
of the elements, for example mercury, are observed to be a mix-
ture of many different kinds of atoms, all having practically
integer masses. We thus return to Prout's hypothesis and, now
on the basis of unambiguous experimental evidence, assert that
the masses of all atoms are integer multiples of a unit mass which is
one-sixteenth of the mass of the oxygen atom. The presumption
is inescapable that this unit mass must be a definite physical
entity and must be the primordial substance out of which all
atoms are built. The two kinds of lithium atoms, which have
masses of 7.00 and 6.00, must contain, respectively, 7 and 6 of
these units; the two kinds of chlorine atoms contain, respectively,
35 and 37 of these units; and so on.
What is this unit? We may guess that it is the nucleus of the
hydrogen atom, the mass of which is very nearly, but not quite,
unity, namely, 1.0078, and which has a charge +e. This
particle is called the "proton." If this surmise be correct, the
helium nucleus should contain 4 protons and should have a
mass of 4.0312 and a charge of +4e. The helium nucleus,
however, i.e., the a particle, actually has a mass of 4.00 and a
charge of +2e. We can explain these facts by assuming that
(1) the a particle is made up of 4 protons (total charge +4e)
and 2 electrons (total charge -2e), giving to the a particle a net
charge of +2e; and (2) when these 4 protons and 2 electrons
combine to form an a particle, there is a loss of mass of 0.0312
gram, which is transferred into radiant energy, perhaps by a kind
of superchemical reaction. According to the theory of relativity,
there is a definite equivalence between radiation and matter,
such that a definite mass of matter mis equivalent to a quantity
of radiation E, according to the relation
E = mc 2
SEC. 2] ISOTOPES 545
where c is the velocity of light. For every 4 grams of helium
formed in this way from hydrogen, there should be a loss of mass
of 0.0312 gram and a consequent emission of radiant energy E
equal to
E = 0.0312 X (3 X 10 10 ) 2 ergs
-an amount of energy sufficient to raise the temperature of some
200 tons of water from O to 100°C. ! Clearly, this is a cosmical,
and not a terrestrial, process.
Once the a particle is formed, it appears to be an exceedingly
stable structure, since there is no evidence of any kind indicating
the breaking up of these particles. They can serve as units,
perhaps in connection with protons, in building up the nuclei
of the heavier elements. The oxygen nucleus, for example,
should on this hypothesis consist of 4 a particles, while the
nitrogen nucleus, which has atomic mass 14, cannot be made up
entirely of a particles but might contain 3 a particles, 2 protons,
and 1 additional electron to give it a mass of 14 and a net charge
of +7e.
This surmise is, in part at least, confirmed by the remark-
able experiments of Rutherford, in which the atoms of the lighter
elements are bombarded by the swiftly moving a particles
ejected from radioactive materials. In these experiments
Rutherford has obtained unmistakable evidence of the disruption
of the atomic nuclei, and he finds that protons can be "knocked
out" of nitrogen atoms but not out of oxygen atoms. For an
account of these experiments, the reader is referred to the original
articles. 1
(d) The Separation of the I sotopes.-There have been numerous
attempts to separate the isotopes of the elements in appreciable
quantities, but these attempts have met with only partial success.
The usual methods of separating elements from each other are
based upon differences in chemical and physical properties. As we
have seen in previous chapters, the properties of a given atom are
determined by the number and arrangement of the extranuclear
electrons. The isotopes of a given element, however, are identical I
both as to the number and arrangement of these electrons.
1 RUTHERFORD: Phil. Mag., vol. 37, p. 537 (1919) and vol. 41, p. 307
(1921); Nature, vol. 109, p. 614 (1922); RoTHERFORD and CHADWICK:
Phil. Mag., vol. 42, p. 809 (1921); vol. 44, p. 417 (1922); vol. 48, p. 509
(1924). Many other articles dealing with this important subject will be
found.
546 THE NUCLEUS: MASSES OF ATOMS [CHAP. XIII, PART I
The two kinds of lithium atoms have 3 extranuclear electrons
each. They have, therefore, identical chemical properties;
identical 1 spectra; and are quite indistinguishable from each
other except for those properties which depend only on mass.
We may picture the Li(6) nucleus as containing 1 a particle,
2 protons, and 1 electron; and the Li(7) nucleus as containing
1 a particle, 3 protons, and 2 electrons. In either case, the
atomic number, i.e., the net positive charge on the nucleus, is 3.
Incidentally, since the average mass of the lithium atom is 6.94,
we conclude that Li(7) is much more abundant than Li(6), which
probably means that the nuclear structure of Li(7) is more stable
than that of Li(6).
The methods of separating isotopes, therefore, can be based
only on such differences in behavior as result from differences in
mass. The mass spectrograph does actually separate isotopes of
the elements. By replacing the photographic plate PP in Aston's
apparatus (Fig. 158) by suitably disposed slits, we should be
able to collect one kind of neon atoms, say, in one compartment
and the other kind in another. This method has actually been
tried but the number of atoms obtainable by this method is so
exceedingly minute that it is almost impossible to get a quantity
of any isotope sufficient for any test, physical or chemical.
A method based on differences in rates of evaporation as a
result of different atomic masses has been somewhat more
successful. In mercury, for example, which has several isotopes,
the lighter constituents should evaporate from the surface of
(liquid) mercury at a slightly greater rate than the heavier
constituents. Bronsted and Hevesy 2 allowed mercury vapor
evaporating from the liquid maintained at a temperature of some
50°C. to condense onto a surface cooled by liquid air. After
several "fractionations" they prepared from nearly 3 litres
of mercury two portions of about 0.2 cc. each which differed
from each other in density by a few hundredths of 1 per cent-
1 There is, however, distinct evidence showing that different isotopes of a
given element emit slightly different band spectra; also minute differences
in the frequencies of line spectra have been observed due to difference of
mass of the nuclei in somewhat the same way as the greater mass of the
helium nucleus results in a larger value of the Rydberg constant for that
element than for hydrogen. See "Molecular Spectra in Gases," Nat.
Research Council Bull. 57, Chap. V.
2 Nature, vol. 106, p. 144 (1920); vol. 107, p. 619 (1921). Phil. Mag.,
vol. 43, p. 31 (1922).
SEC. 3] THE PACKING EFFECT 547
not a complete separation, but sufficient to serve as a confirma-
tory check on the conclusions from the mass spectrograph.
3. The Packing Effect.-The data obtained by the use of the
mass spectrograph indicates, beyond question, that the nuclei of
atoms are built up of certain units, which units are probably
protons and electrons. We saw however, that in order to explain
the fact that the mass of the helium atom is not quite four times
the mass of the hydrogen atom,it was necessarytoassumethatthe
'' packing'' of 4 protons and 2 electrons together to form the
helium nucleus results in a loss of mass. This loss of mass in
the formation of nuclei from protons and electrons is called the
packing effect. There is every reason to believe that the packing
effect should play a part in the formation of all nuclei, although
the circumstance that the masses of all atoms heavier than
helium are whole numbers, according to the above mentioned
measurements of Aston and of Dempster, indicates that in the
grouping together of alpha particles to make nuclei the packing
effect must be small. For example, the mass of the oxygen atom
is almost exactly four times the mass of the helium atom. When
4 a particles unite to form an oxygen nucleus, the loss of mass is
proportionately very much less than in the formation of a helium
nucleus from protons. In the case of the Li(7) atom, which,
presumably, contains 1 alpha particle and 3 protons-the mass of
the electrons is negligible-the mass should be slightly more than
7.00, namely, 7.023, if the 3 protons are not subjected to the
packing effect. The difference between 7.00 and 7.023 is almost
within the limits of experimental error. The study of the packing
effect in atoms of higher atomic number than helium demands
more precise data on atomic masses.
Aston has attacked this problem and has evolved an improved
mass spectrograph, 1 which has a resolving power 2 of 1 part in 600
and a precision in the determination of mass in terms of the mass
of the oxygen atom as 16.000, of 1 part in 10,000. This remark-
able achievement has resulted in the discovery of several new
isotopes, and in the determination of masses with a precision
heretofore unknown. Some of these new measurements by Aston
are given in Table II. Column three of this table shows the
1 AsToN, F. W.: The Bakerian Lecture, Proc. Roy. Soc. (London), vol.
115, p. 487 (August, 1927).
2
That is, two atoms whose masses differ by only 1 part in 600 are shown
as separate '' lines ''
548 THE NUCLEUS: MASSES OF ATOMS [CHAP. XIII, PART I
"mass number" of the several atoms, i.e., the number which
gives the number of protons in the nucleus. We thus have
associated with each atom three fundamental quantities-atomic
number, mass number, and mass. To these may be added a
fourth-the chemically determined atomic weight, or, more cor-
rectly, the combining weight. For chlorine, for example, these
four quantities are as follows:
Atomic number. . . . . . . . . . . . . 17
Mass numbers. . . . . . . . . . . . . . 35 and 3 7
Masses . . . . . . . . . . . . . . . . . . . . . 34. 983 and 36 . 980
Combining weight. . . . . . . . . . . 35. 458
It is observed from Table II that, except in the case of oxygen,
which is taken as a standard, and also of fluorine, the masses of
atoms are not whole numbers but differ from whole numbers by
very small amounts. These precise measurements give for each
TABLE IL-THE MASSES OF ATOMS BY ASTON'S PRECISION MASS
8PECTOGRAPH
Packing
Atomic Mass Atomic
Element fraction,
number number masses
times 104
H ................... 1 1 77.8 1.00778
He .................. 2 4 5.4 4.00216
Li(6) ................ 3 6 20 6.012
Li(7) ................ 3 7 17 7.012
Bo(lO) ............... 5 10 13.5 10.0135
Bo(ll) ............... 5 11 10.0 11.0110
C ................... 6 12 3.0 12.0036
N ................... 7 14 5.7 14.008
0 ................... 8 16 (standard) 16.000
F ................... 9 19 0.0 19.000
Ne(20) ............... 10 20 0.2 20.0004
Ne(22) ............... 10 22 2. 2(?) 22.0048
P ................... 15 31 -5.6 30.9825
Cl(35) ............... 17 35 -4.8 34.983
Cl(37) ............... 17 37 -5.0 36.980
Ar(36) ............... 18 36 -6.6 35.976
Ar(40) ............... 18 40 -7.2 39. 971
As .................. 33 75 -8.8 74.934
Br(79) ............... 35 79 -9.0 78.929
Br(81) ............... 35 81 -8.6 80.926
Xe(134) .............. 54 134 -5.3 133.929
Hg(200) ............. 80 200 +o.8 200.016
SEC. 4] BECQUEREL'S DISCOVERY 549
of the atoms measured the packing fraction-as Aston calls it-
which is "the gain or loss of mass per proton when the nuclear
packing is changed from oxygen to that of the atom in question."
The packing fractions for the several elements are given in
column four of Table IL The packing fraction of carbon is
+0.0003, and its mass is 12.0036. Aston finds that starting with
hydrogen, the packing fraction decreases rapidly with succeeding
elements, reaches a minimum of -0.0009 in the neighborhood of
elements of atomic number 30 to 40, and then increases again,
becoming +0.00008 for mercury. High packingfractions (taking
0 = 16.000 as standard) indicate looseness of packing or com-
parative instability of the nucleus. This increase in the packing
fraction toward the end of the periodic table of the elements may
perhaps foreshadow the nuclear instability which is the cause of
radioactivity. At any rate, these precise determinations of the
masses of atoms offer, perhaps, a new line of attack on the prob-
lem of nuclear structure.
p ART IL RADIOACTIVITY
The mass spectrograph yields precise data concerning the
masses of the nuclei of atoms but gives no direct clue to the
structure of the nucleus. From the circumstantial evjdence
that the masses of all the atoms are very nearly whole numbers,
we infer that atomic nuclei are built up of fundamental units of
some kind, which units we may assume are the so-called ''pro-
tons." For direct evidence, meager though it is at present, we
turn to the field of radioactivity.
4. Becquerel's Discovery.-The discovery of the phenomenon
of radioactivity, although quite accidental, -resulted directly
from the discovery of X-rays. Roentgen had showed that X-rays
are emitted by those parts of the discharge tube which are bom-
barded by the cathode rays. This bombardment was also
accompanied by the emission of the well-known greenish or
bluish fluorescence. The question arose: Is fluorescence always
accompanied by the emission of X-rays?
Several investigators had apparently found that fluorescent
bodies activated by sunlight gave out a type of radiation which,
like X-rays, was able to pass through black paper and to affect a
photographic plate. In February, 1896, a few months after the
discovery of X-rays, Henri Becquerel1 was trying an experiment
1 BECQUEREL'S papers appear in Compt. rend., vol. 122 (1896).
550 THE NUCLEUS: RADIOACTIVITY [CHAP. XIII, PART II
of this kind using as the fluorescing substance the double sulphate
of uranium and potassium. After preparing the experiment and
while waiting several days on account of inclement weather for
sunshine, Becquerel discovered that the specimen even in the
dark emitted a radiation which penetrated the black paper, thin
sheets of metal, and other substances opaque to ordinary light
and that exposure of the fluorescing substance to sunlight had no
effect on the phenomenon.
Becquerel soon found that this radiation was emitted by
uranium irrespective of its state of chemical combination and
that there was no connection whatever between this phenomenon
and phosphorescence. Further, the phenomenon was found to
be quite independent of the temperature of the uranium com-
pound. It was later discovered that these rays from uranium
possess the power of discharging electroscopes by rendering the
air through which they pass conducting.
This property of" radioactivity," as it was called, was found to
be possessed by several substances. Among these substances
were thorium and two new elements-polonium and radium-
discovered by M. and Mme. Curie, the latter substance being
more than a million times more active than uranium.
5. The Types of Radioactive Radiations.-(a) Like X-rays,
the rays from radioactive materials affect a photographic plate,
cause fluorescence, and ionize gases
r through which they pass. Unlike
X-rays, the radioactive rays are of
three types, known, respectively, as
the a-rays, the /3-rays, and the ,'-
rays. There are various ways of
showing the existence of these vari-
ous kinds of rays. Schematically, if
(3 a small quantity of radium prepara-
tion be placed at the bottom of a
small hole drilled into a heavy metal
p p
block B, the emerging rays can be,
Frn. 162.-Schematic represen- ideally, divided into the three groups
ta tion of the three types of rays
from radioactive materials. by use of a magnetic field of suit-
able strength at right angles to the
plane of the paper and directed away from the reader, as is
shown in Fig. 162. One group is bent into a circular path
to the right and will cause an impression on a phot?graphic
SEC. 5] TYPES OF RADIATIONS 551
plate PP. These are the {3-rays. From the direction of their
deflection, it follows that they must be negatively charged
particles. By studying quantitatively their deflection in mag-
netTo and electric fields, it was shown that these particles are
electrons which are ejected from radioactive materials with,
in some cases, very high velocities, there being, in general, differ-
ent groups with different velocities from any one material. A
second type of radiation is deflected by strong fields toward the
left. This type consists of positively charged particles called
"a particles," which were shown to possess a ratio e/m much
smaller than that for the /3-rays; in fact, a ratio of the order of
magnitude of that for atoms. These a particles were found to
have a mass 4 (taking the mass of the oxygen atom as 16) and to
carry a charge +2e. This identified them with the nuclei of
helium atoms. The third type of radiation, the ')'-rays, proceeds
undeviated by either electric or magnetic fields, has a very high
penetrating power, and is now known to consist of electromag-
netic radiations of very short wave length lying, in general, in
the spectral region beyond the shortest X-rays.
These three types of rays are further differentiated from each
other by their penetrating power. The a-rays are absorbed by
a few cemtimeters of air at ordinary pressure. They are reduced
in intensity one-half by passing through 0.005 mm. of aluminum.
Their initial velocities are of the order of 2 X 10 9 cm.-sec- 1 • The
{3-rays are, roughly, one hundred times more penetrating, since it
requires something like 0.5 mm. of aluminum to reduce their
intensity to half. The initial velocities of the {3-rays, in some
instances, exceed 99 per cent of the velocity of light. The ')'-rays
are able to penetrate many centimeters of even so dense a metal as
lead.
(b) The technique which has been developed for studying,
classifying, and interpreting the phenomena of radioactivity
constitutes almost a separate science by itself. For details, the
reader is referred to special treatises 1 and to the original articles.
As an illustration, we shall describe briefly the methods used by
Rutherford and his collaborators in determining the properties of,
and in identifying, the a particles.
1RUTHERFORD: "Radioactive Substances and Their Radiations" (1913);
SoDDY: "The Chemistry of the Radio Elements" (1914); HEVESY and
PANETH: "Radioactivity," translated by R. W. Lawson (1926); KovARIK
and McKEEHAN: "Radioactivity," Nat. Research Council, Bull. No. 51
(1925).
552 THE NUCLEUS: RADIOACTIVITY [CHAP. XIII, PART II
First, by studying the deflections produced by electric and
magnetic fields Rutherford determined the ratio E / M where
E is the charge carried by the a particle and M is its mass. His
apparatus for this purpose is shown, schematically, in Fig. 163.
A fine wire W coated with an active deposit of radium C, which
emits a particles, was placed in a groove in a block B within a
highly evacuated vessel which was in a very strong magnetic
field perpendicular to the plane of the figure. The a particles
emitted by W move with velocity v in circular paths and some
of them pass through the slit S and fall upon the photographic
s
A B
FIG. 163.-Rutherford's apparatus FIG. 164.-Rutherford's apparatus
for measuring the magnetic deflection for measuring the electrostatic deflec-
of a rays. tion of a rays.
plate PP at L1 with the magnetic field directed away from the
reader, and at L 2 with the field reversed. The two traces L1 and
L 2 were sharp, so that the radius r of the circular path could be
quite accurately determined. This gave, since the field H was
known,
M
Hr= Ev
To determine v it was necessary to measure the deflection
produced by an electrostatic field. This proved to be difficult
because of the very small deflections. The wire W and block B
(Fig. 164) were arranged as in the previous experiment. Imme-
diately over the wire were two metal plates A and B some 5 cm.
long and 0.2 mm. apart, between which could be maintained an
electric field of the order of 20,000 volts per centimeter. The a
SEC. 5] TYPES OF RADIATIONS 553
particles, in passing through this field, are deflected at right angles
to their path and by reversing the field two "lines" L1 and L 2
were produced on the photographic plate PP. From the distance
between these lines and the dimensions of the apparatus, it was
possible to determine the deflection produced by the electro-
static field and thus to get a numerical value of the quantity
M 2
E v • (The method used in getting this expression is some-
what similar to that used in getting equation (la).) By com-
bining these two experiments both E / M To E!ecfrom~fer
and v could be determined. Rutherford
found that, for the a particles from
radium C, v = 2.06 X 109 cm. per second
and E/M = 5,070 e.m.u. per gram.
The value of E / M for the hydrogen atom
is (see Chap. VI, Sec. 4) about double this,
namely, 9,650. As in the case of the elec-
tron, to determine M it became neces- c
sary to measure E. A known quantity w
of radium C was deposited on a plate P
(Fig. 165) placed in a highly evacuated
vessel. The a particles, after passing
through a window W of known area and
covered with very thin aluminum, fell p
upon the metal plate C, giving thereto the
charge which they carry. This charge ac- Frn. 165.-Rutherford's
quired by the plate in a known time was apparatus for measuring
measured by an electrometer. To obtain the charge carried by the
a-r a y s fr om a known
the charge per particle it was necessary quantity of radioactive
material.
to know the number of particles.
Rutherford and Geiger measured ("counted") by means of
the apparatus shown in Fig. 166, the number of a particles
emitted per second by a known quantity of radium C. A known
quantity of radium C was deposited on a disk D which was
placed inside the highly evacuated vessel A at a known distance
from a small circular opening S of known area which was covered
by a sheet of mica thin enough to allow the passage of the
a particles into a brass chamber C. This chamber was
evacuated to a pressure of several millimeters of mercury and
had at its center an insulated wire WW which, by means of a
battery B, was maintained at a potential, with respect to the
554 THE NUCLEUS: RADIOACTIVITY [CHAP. XIII, PART II
walls of the cylinder, just less than the critical discharge potential.
When an a particle entered C through the aperture S the ioniza-
tion caused by the passage of the particle lowered the critical
potential by an amount sufficient to allow the passage through
the cylinder of a momentary current, which could be detected
by a ''kick" in the electrometer E. In this way, Rutherford
and Geiger were able to detect the passage of single a particles.
The quantity of radium on D, the distance DS, and the area of
S were so adjusted that several particles per second entered the
chamber C. Thus, knowing the rate at which the a particles
passed through S, the aperture which S subtended at D, and the
quantity of radium on D, it was possible to determine the total
number of a particles emitted per second per gram of radium.
The total number of particles emitted per second by the
radium preparation in the experiment illustrated in Fig. 165
Ji' a
JD .A w w
Frn. 166.-Apparatus of Rutherford and Geiger for "counting" a particles.
could thus be computed and thence the charge per particle.
It was found that each a particle carried a charge of about
9.3 X 10- 10 e.s.u.-approximately twice the value 4.77 X 10- 10
of the charge per electron. From this it follows that the a
particle has (1) a charge +2e and (2) from the previously
determined value of E / M, a mass approximately four times
the mass of the hydrogen atom. In other words, the a particle
is a helium atom with a double charge.
This conclusion regarding the nature of the a particles was
confirmed, both qualitatively and quantitatively, by the observa-
tion that helium is produced as a result of radioactive disintegra-
tion. Rutherford showed, from the measurements on counting
a particles, that 1 gram of radium in radioactive equilibrium with
its surroundings should emit 1.3 X 10 10 a particles per second.
If the a particle is a (charged) helium atom 1 gram of radium
should, in 1 year, emit about 0.160 cc. of helium. A direct
SEC. 6] ORIGIN OF RAYS 555
laboratory study of the rate of emission of helium from radio-
active materials, as well as a study of the evolution of helium
from radioactive minerals, gave results in satisfactory quantita-
tive agreement with the quantity predicted from the rate of a-
particle emission.
The evidence is quite complete, therefore, that the a particles
are helium atoms with a charge +2e. Since the neutral helium
atom has 2 (extranuclear) electrons, the conclusion is obvious,
therefore, that the a particles are helium nuclei.
6. Origin of the Rays.-From the fact that the a-rays have
atomic mass 4, one concludes, at once, that they must come from
the nuclei of atoms, since in our present concept of atomic struc-
ture there is no place in the extranuclear region of an atom for
particles of mass so large as that of the a particle. And it
follows at once, that a radioactive atom which has lost an a
particle has thereby decreased in atomic weight by 4 units-
it must, therefore, be a different type of atom. The a particles
emitted by a given radioactive element all have the same velocity,
from which we conclude that each a particle is expelled as a
result of a definite atomic process-a disintegration. That each
disintegrating atom gives out 1 a particle was shown by actually
counting the number of a particles given out by a known amount
of a radioactive substance in a given time and comparing the
number so obtained with the known rate of disintegration of
the material.
The origin of the {3-rays cannot be so definitely fixed, partly
because we know that the atom possesses electrons in its extra-
nuclear structure, partly because the {3-ray emission from a
given radioactive element consists, in general, of {3 particles
having a wide range of velocity. The most swiftly moving have
a velocity which approaches, within a small fraction, the velocity
of light. Others have much smaller velocities. The magnetic
spectrum of the {3-rays from some of the radioactive elements
shows sharp lines, indicating groups of {3 particles with definite
velocities. And it is probable that some of them come from the
nucleus; others, by a kind of secondary photoelectric process,
from the extranuclear structure.
The "!-rays have been studied, in part, by reflection from
a crystal grating, as in the case of X-rays and, in part, by a
secondary photoelectric method in which the velocities of the
photoelectrons liberated by "!-rays was measured by the mag-
556 THE NUCLEUS: RADIOACTIVITY [CHAP. XIII, PART II
netic spectrograph, and from these velocities the frequencies of
the exciting radiation could be computed. In this way, it is
found that the '}'-rays cover a wave-length range extending
0 0
from about A = 0.005 Angstrom to beyond A = 1.0 Angstrom.
The rays of shorter wave length in this region must certainly
come from the nucleus, for we can fix definitely the shortest rays
which the extranuclear structure of any atom can radiate. This
is the K'Y line of uranium which has a wave length of 0.1085
0
Angstrom. Any line radiation of shorter wave length must
come from the nucleus.
Primarily, then, the phenomena of radioactivity have their
origin in nuclear processes, although, as secondary effects, the
extranuclear structure is involved.
7. The Radioactive Disintegration Series.-A systematic
study of the radioactive process and of disintegration products
has led to the generalization that there are several series of the
radioactive elements, the elements of a given series being so
arranged that each is a disintegration product of a preceeding
element. Thus, the uranium atom when it disintegrates emits
an a particle, mass 4, and charge +2e. Uranium has an atomic
mass of 238 and a (net) nuclear charge of +92e. After emitting
the a particle the remaining nucleus must have mass 234 and a
charge +90e. This is an element of atomic number 90. It is
called "uranium X1" (UX1). Although possessing atomic
number 90, UX1 cannot be thorium, for which also Z = 90,
since thorium has atomic mass 232.
The UX1 atom, in disintegrating, emits a /3 particle, charge
- e, from its nucleus. The effect of this is to increase the net
nuclear positive charge to +91 and therefore, also, the atomic
number to 91 but to leave the mass unchanged. This new
element is called UX2. It, in turn, emits a {3 particle. The
resulting product, uranium II, (U II) has atomic number 92, and
an atomic mass of 234. It is, therefore, an isotope of uranium.
U II emits an a particle, producing ionium, mass 230, atomic
number 90. Ionium is, therefore, an isotope of UX1. Ionium,
in turn, emits an a particle and becomes radium, atomic mass
226, atomic number 88. A direct determination of the atomic
weight of radium by chemical methods yields 226. Radium,
then, is a disintegration product from uranium. Continuing
in this way radium G is finally reached which has an atomic
mass of 206 and an atomic number of 82. Radium G is there-
SEC. 7] DISINTEGRATION SERIES 557
fore an isotope of lead. Radium G is not radioactive and is,
therefore, ·the end of the series.
Another series of disintegrations starts with thorium, atomic
number 90 and atomic mass 232, and ends with an element of
atomic number 82 and atomic weight 208. This is also an isotope
of lead. These two series are shown by the conventional diagram
of Fig. 167. The elements in the same vertical rows are isotopes;
240
238
236
t;f
234
232
¥~
,rt(
230 /" Io
228 ~ a,
1'ht' ~~h,
226
~£/_/ '
11,/ "" /
~m'/
~ 220 r5/
)Efll
+
4 218 11ial ,,,"
LI;~(/
215
2\4
212.
210
HJ ~1
\JV7 "Ci C'
11
~;.,i•p)
')vy?:~~
~
20& HP/
206
204
~ P'b .Bi
a1 82. 83 84 85 86 81 88 89 90 91 92 ~3 I
Atomic Number
Frn. 167. -The radioactive disintegration series of uranium and thorium.
while the elements in the same horizontal rows are isobars, i.e.,
elements which have the same atomic mass but different chemical
properties, since they have different atomic numbers and there-
fore different extranuclear electron structures.
The end products of these two series are isotopes of lead with
atomic masses 206 and 208, respectively. Ordinary lead has
atomic weight 207.2. The atomic weight of lead, however,
depends somewhat on the source. Uranium-bearing minerals
usually contain lead-a fact in itself confirming the series of
radioactive transformations from uranium to lead. The same is
true of certain thorium-bearing minerals. Further, the atomic
558 THE NUCLEUS: RADIOACTIVITY [CHAP. XIII, PART II
weight of lead coming from uranium-bearing minerals is only a
little over 206, while lead associated with Norway thorite has an
atomic weight of 207.9, which is nearly the value 208 predicted
for the end products of the thorium series.
These radioactive transformation series, therefore, from
uranium and thorium to lead, are the result of a spontaneous
disintegration of the atoms, more correctly of the nuclei, this dis-
integration being accompanied by the emission of a particles,
{3 particles, and ')'-rays. We thus have direct evidence that the
atomic nuclei of the radioactive elements, at least, and, pre-
sumably, also of the other elements, contain a particles and
electrons, perhaps also protons, all held together in equilibrium
by some kind of mechanism or forces of which we have no
knowledge.
We can make guesses-they are guesses only-as to the number
of a particles, protons (call them "'YJ particles"), and electrons
(" {3 particles") which the radioactive nuclei may have. For
example, uranium has atomic number 92 and atomic mass 238.
From the atomic mass, we may conclude that the uranium
nucleus contains 59 a particles and 2 'YJ particles (238 = 4 · 59 +
2). If so, the total positive charge on the nucleus must be (2 · 59
+ 2) or 120. But from the atomic number of uranium 92, we
know that the net positive charge is +92. The uranium nucleus
must, therefore, contain 28 electrons, and we may write sym-
bolically the composition of the uranium nucleus as a5 9{3 28 'Y] 2 •
Or, we might guess instead that the uranium nucleus contains
58 a particles and 6 "IJ particles. The composition would then
be a5sf33o'YJ6· In the absence of experimental evidence, one guess
is as good as another.
8. Gamma-ray Spectra.-Direct measurements of the wave
lengths of some of the ,,-rays have been made by use of the
crystal grating. The wave length is so small in comparison,
however, with the grating constants of available crystals that
only the longer wave lengths can be measured in this way. Con-
sequently, an indirect method has been developed based on the
magnetic spectrum either of the secondary {3-rays (photoelec-
trons) produced by ')'-rays or, later, of the primary {3-rays
themselves.
The method employed by Ellis 1 has been followed, with modifi-
cation, by subsequent investigators. · The principle of the
1 Proc. Roy. Soc. (London), vol. 99, p. 261 (1921).
SEC. 8] GAMMA-RAY SPECTRA 559
method is identical with that used by Robinson (p. 515) in his
investigation of the photoelectrons produced by X-rays. In
Robinson's work, X-rays of known frequency caused the emission
from various energy levels within the atom, of photoelectrons.
By measuring by the so-called "magnetic spectrograph," the
energies with which the various groups of electrons leave
the atom, Robinson was able to compute the energy values of the
levels from which they (the photoelectrons) originate. The pro-
ceedure is reversed in the experiments of Ellis. Knowing the
energy levels from which the photoelectrons emitted by ')'-rays
come, and the energies of these electrons as they leave the atom,
the energy (hv) of the rays can be determined. In Ellis' experi-
ments the source of the secondary rays was placed at W (Fig. 168)
near a lead block B. W was a small
glass tube 1 cm. long and 0.7 mm. in
diameter, the inside walls of which were
---\.:.::P===L:::!::::=..P
coated with a deposit of radium C which
served as a source of the ,,-rays under
study. Around W was wrapped a foil B
of the metal M-Pb, Pt, W, etc.- w
which was to serve as the source of Frn. I68.-The {J-ray spectro-
the secondary {3-rays. The ')'-rays pass- graph of Ellis.
ing out through the walls of the tube
excite secondary (3-rays, or photoelectrons, in the metal foil. The
whole apparatus is placed in a magnetic field (and, of course, in
a highly evacuated enclosure). The photoelectrons pass, in cir-
cular paths, through the wide slit S and are "focused" onto the
photographic plate PP at L. On developing the plate, a number
of "lines" are found corresponding to various electron energies
which can be determined by knowing the magnetic field and the
diameter L W of the circular paths.
The following data (Table III) taken from Ellis' paper will
serve to illustrate the method of analyzing the results: For any
given metal M there was observed a number of lines on the plate
of which three were more intense. These same three appeared on
the plate for each element but were shifted in position in such a
direction as to indicate that the corresponding photoelectrons had
less energy the greater the atomic number of M. This is shown
in the upper part of Table III, in which the energies of the photo-
electrons are expressed in equivalent volts. 1 Ellis concluded
1This is the potential difference in volts through which the electron would
have to fall to give it the observed energy.
560 THE NUCLEUS: RADIOACTIVITY [CHAP. XIII, PART II
from a study of the data that these electrons were expelled from
the K levels of the respective atoms by three different ,y-ray
frequencies. Applying Einstein's photoelectric equation
hv =Ek+ WK
where Ek is the observed kinetic energy of the photoelectron and
WK is the energy required to remove the electron from the K level
of the atom M, one obtains at once the energy hv of the three
incident ,y-rays Data from each metal M is found in this way
to give the same group of three ')'-rays, as the lower half of Table
III shows. These then are three lines in the ')'-ray spectrum of
radium C, the respective wave lengths of which are 0.0518,
0
0.0425, and 0.0355 Angstrom. In a similar way, energy values
of many other lines in ')'-ray spectra have been measured.
TABLE IIL-DETERMINATION OF THE WAVE LENGTH OF ')'-RAYS FROM THE
ENERGY OF SECONDARY {3 RAYS (ELLIS)
Tung- Plati- Ura-
Lead
Emit tor~ sten num n1um
82
74 78 92
Secondary {3-ray energies 1.66 1.58 1.49 1.22
m equivalent v o 1 ts 2.20 2.12 2.03 1. 74
times 10-5 2.76 2.69 2.60 2.31
Ex of emittor-~ 0.693 0.782 0.891 1.178
Mean x,
0
energy Angstroms
Energy (hv) of -y-rays in 2.35 2.36 2.38 2.40 2.37 0.0518
equivalent volts times 2.89 2.91 2.92 2.92 2.90 0.0425
10-s 3.46 3.46 3.49 3.48 3.47 0.0355
Investigations of the primary /3-rays emitted by radioactive
bodies showed that the energies of these rays are in accord with
the hypothesis that some of them are, in reality, photoelectrons
coming from one or the other of the levels in the extranuclear
structure of the radioactive atom itself as a result of the photo-
electric action of the ')'-ray as it passes from the nucleus where
it originates out through the extranuclear structure. Conse-
quently, in later work, the secondary metal radiator M is dis-
pensed with and the source W (Fig. 168) is a fine wire coated
with the radioactive deposit under study. 1
1 ELLIS and SKINNER: Proc. Roy. Soc. (London), vol. 105, p. 60 (1924).
SEC. 9] NUCLEAR ENERGY LEVELS 561
9. Nuclear Energy Levels.-The measurements by Ellis and
Skinner 1 of the ')'-ray spectrum of radium Band C, and by Black 2
of the ')'-ray spectrum of thorium products have yielded suffi-
cient data so that it has been possible to postulate a system of
energy levels in the nucleus, transitions between which give rise
to ')'-ray lines in exactly the same way that X-ray lines originate
in transitions between levels in the extranuclear structure of
the atom. The wave lengths and energy values (in equivalent
volts) of some of the lines in the ')'-ray spectrum of radium B,
together with the "name" of the line, are shown in the first
three columns of Table IV.
Following a procedure commonly applied to other spectral
regions, Ellis and Skinner found it possible to postulate a series
of seven energy levels A, B, C . . . transitions among which
give rise to the fourteen lines given in Table IV, assuming the
applicability of the ordinary quantum principles of the origin
of spectral lines. The interlevel transitions giving rise to the
several lines are given in column four, and an energy-level dia-
gram is shown in Fig. 169. Tables V and VI give similar data
TABLE IV.-SoME LINES IN THE ')'-RAY SPECTRUM OF RADIUM Band THE
ENERGY LEVELS IN THE RADIUM B NucLEus (ELLIS AND SKINNER)
Wave Energy, I Proposed energy levels
Line length, volts times Origin in the radium B
Angstroms 10-s j nucleus, volts times IQ-5
C2* 1.37 0.090 B~C A= 0
C1 0.230 0.537 A~B B = 0.537
Cs* 0.196 0.629 A~c C = 0.625
C3* 0.115 1.073 E~F D = 2.572
C4* 0.098 1.250 F~G E = 2.942
E1 0.0625 1.947 c~D F = 4.048
E2 0.0616 2.035 B~D G = 5.31
E3 0.0513 2.404 B~E
E4 0.0480 2.572 A~B
Es 0.0451 2.733 D~G
E6 0.0419 2.942 A~E
E9 0.0351 3.511 B~F
EI6 0.0263 4.684 c~a
E11 0.0257 4.800 B~G
* Measured by Rutherford and Andrade by crystal reflection.
1 Proc. Roy. Soc. (London), vol. 105, pp. 165, 185 (1924).
s Proc. Roy. Soc. (London), vol. 106, p. 632 (1924); vol. 109, p. 166 (1925).
562 THE NUCLEUS: RADIOACTIVITY [CHAP. XIII, PART II
obtained by Black for two members of the thorium series. It
may, or may not, be significant that no lines correspond;~g to
certain of the (mathematically possible) transitions were found.
While much remains yet to be done in securing and in inter-
preting data on the ')'-ray spectra of the radioactive elements,
when one considers the difficulty of the measurements it is a
Ao
1 Cs C,e B
c ,
__
16 IJ)
Z-2 )(
D v>
-+-
E ·c
3 :::>
~
C')
5 '-
Q)
c;:
F 41.lJ
c
G
5
Frn. 169.-Energy-level diagram for the radium B nucleus, showing the origin
of the -y rays (Ellis anil Skinner).
scientific achievement of the first magnitude to have isolated
and measured with acceptable precision the various ')'-ray lines. 1
TABLE V.-THE -y-RAY SPECTRUM OF MESOTHORIUM (BLACK, Zoe. cit.)
Wave Energy, Proposed
Linc number 0
length, volts times Origin levels, volts
Angstroms 10-5 times 10-5
1 0.213 0.578 A~B I A= 0
2 0.0955 1.291 B~C B = 0.58
3 0.0671 1.838 A~C C = 1.84
5 0.0365 3.38 c~D D = 5.22
6 0.0267 4.62 B~D E = 9. 72
7 0.0135 9.14 B~E
8 0.0127 9.69 A~E
1 The bulletin by Kovarik and McKeehan (Zoe. cit., p. 551) gives an
excellent account of measurements of radioactive phenomena and their
interpretation up to 1925. Extensive tables are given showing data on a-,
{1-, and ')'-rays.
SEC. 10] STRUCTURE OF NUCLEI 563
TABLE Vl.-THE ')'-RAY SPECTRUM OF THORIUM D (BLACK, lac. cit.)
Wave Energy,
Line
length, volts times
number 0
10-a
Angstroms
1 0.279 0.41
2 0.0791 1.45
3 0.0543 2.11
4 0.0491 2.33
5 0.0452 2.53
6 0.0442 2.59
7 0 .0510 2.79
8 0.0393 2.91
9 0.0221 5.17
10 0.0171 6.58
10. The Structure of the Nuclei of the Radioactive Elements.-
A mechanism capable of emitting a-rays, /3-rays, and a whole
spectrum of monochromatic '}'-rays must possess a well-ordered,
though probably very complex, structure. What type of struc-
ture does the existing data on radioactivity suggest? Perhaps
it would be better not to attempt to form a "picture" of nuclear
structure and processes. Nevertheless, the whole history of
physics shows that" pictures," models, and analogies have played
an indispensable part in the growth and development of all
branches of physical theory. Most of us can think better in
terms of models and geometrical relations.
Rutherford has attacked this problem 1 and has proposed a
type of structure for the nucleus which in a rough way is anal-
ogous to the structure proposed by Bohr for the extranuclear
part of the atom. The problem of nuclear structure is somewhat
more complex than that of the extranuclear structure, since the
latter contains only electrons while the former contains a par-
ticles, electrons, and, perhaps, protons.
Rutherford advances the suggestion that the nuclear structure
of the radioactive elements comprises a central core, the radius of
which is less than 1 X 10- 12 cm., around which out to a distance
of the order of 1.5 X 10- 12 cm. from the center are found elec-
trons. From 1.5 X 10- 12 cm. to 6 X 10- 12 cm. is a "shell"
occupied by electrically neutral satellites in orbital rotation about
the core. These satellites are helium-like structures consisting
1 Phil. Mag., vol. 4, p. 580 (September, 1927).
564 THE NUCLEUS: RADIOACTIVITY [CHAP. XIII, PART II
of an a particle to which are attached 2 electrons. Under the
action of the strong electric field of the nucleus, the satellites
become "polarized" in somewhat the same way as a neutral
conductor becomes polarized by and is therefore attracted toward
a charged body, and it is this force of attraction which holds
the satellite in its orbit. A given configuration of the entire
system corresponds to one of the energy levels, or states, sug-
gested by Ellis and Skinner.
As a result of some kind of nuclear catastrophy, the equilibrium
is disturbed and the positive part of the satellite, i.e., the a par-
ticle, under the influence of repulsive forces, is projected from
the structure with high velocity and the 2 electrons are pulled
in toward the core around which they then circulate with a
velocity "close to that of light." Occasionally one of these
electrons is hurled from the system (how?) and becomes a high-
speed {3 particle. As a result of this disturbance to the equilib-
rium of the system, the structure settles "down" to a new
state, thereby emitting a ')'-ray, the quantum energy of which
is equal to the difference in the energy of the two states. This
')'-ray in passing out through the extranuclear structure acts
photoelectrically on the electrons, causing the expulsion of
the lower-speed {3 particles such as those measured by Ellis and
Skinner in the {3-ray emission from radioactive substances.
This picture is not so significant as is the fact that sufficient
data has been accumulated from a study of radioactive phenom-
ena to make it possible to postulate a structure for the nuclei of
atoms and to visualize thereby some of the processes which take
place in the disintegration of atoms. What next? At the
beginning of the century, there was a vague impression that the
atom consists of electrical charges. We saw that Thomson's
picture gave way, before experimental evidence, to the Ruther-
ford-Bohr type of atom in which the nucleus was thought of as
a point charge, just as the atom itself in the early chemical
theories played the part of a point mass. Now, the nucleus
itself succumbs to analysis, protons and electrons playing the
part of point charges in the proposed structure. Will these, in
their turn, yield to analysis and be assigned a structure? Or have
we reached the end of the series?
Physics of today seems quite secure, in spite of the chasm
between classical theory and quantum theory, but perhaps some
day some Copernicus will appear who will completely overturn
SEC. 10} STRUCTURE OF NUCLEI 565
our present exceedingly complex structure of physical theories
and concepts and show us a beautiful simplicity in the laws of
nature. If physics continues to grow at the present geometri-
cally increasing rate, the physicist of a half-century hence will
welcome such a revolution with open arms.
APPENDIX I(a)
ATOMIC NUMBERS AND ATOMIC 1 WEIGHTS
Number Symbol Atomic Isotopes 2
Name
weight 1
First period
1 H Hydrogen 1.0077 1.0077
2 He Helium 4.00 4.0
Second period
3 Li Lithium 6.939 7, 6
4 { Ql3
Be Beryllium
Glucinium
l 9.02 9
f
5 B Boron 10.82 11, 10
6 c Carbon 12.000 12
7 N Nitrogen 14.008 14
8 0 Oxygen 16.000 16
9 F Fluorine 19.00 19
10 Ne Neon 20.2 20, 22
Third period
11 Na Sodium 22.997 23
12 Mg Magnesium 24.32 24, 25, 26
13 Al Aluminum 26.96 27
14 Si Silicon 28.06 28, 29, 30
15 p Phosphorus 31.024 31
16 s Sulphur 32.065 32, 34, 33
17 Cl Chlorine 35.458 35, 37
18 A Argon 39.91 40, 36
Fourth period
19 K Potassium 39.095 39, 41
20 Ca Calcium 40.07 40, 44
21 Sc Scandium 45.10 45
22 Ti Titanium 47.9 48
23 v Vanadium 50.96 51
24 Cr Chromium 52.01 52
25 Mn Manganese 54.93 55
26 Fe Iron 55.84 56, 54
27 Co Cobalt 58.97 59
566
APPENDIX 567
ATOMIC NUMBERS AND ATOMIC WEIGHTS (Continued)
Atomic Isotopes 2
Number Symbol Name
weight 1
Fourth period (continued)
28 Ni Nickel 58.69 58, 60
29 Cu Copper 63.57 63, 65
30 Zn Zinc 65.38 64, 66, 68, 70
31 Ga Gallium 69.72 69, 71
32 Ge Germanium 72.38 74, 72, 70
33 As Arsenic 74.96 75
34 Se Selenium 79.2 80, 78, 76, 82, 77, 74
35 Br Bromine 79.916 79, 81
36 Kr Krypton 82.9 84, 86, 82, 83, 80, 78,
81
Fifth period
37 Rb Rubidium 85.44 85, 87
38 Sr Strontium 87.62 88, 86
39 y Yttrium 89.0 89
40 Zr Zirconium 91 90, 94, 92
41 { Nb
Cb
Niobium
Col um bi um } 93.1
42 Mo Molybdenum 96.0
43 Ma Masurium 97.5-98.8*
44 Ru Ruthenium 101.7
45 Rh Rhodium 102.91
46 Pa Palladium 106.7
47 Ag Silver 107.880 107, 109
48 Cd Cadmium 112 .41 110, 111, 112, 113,
114, 116
49 In Indium 114.8 115
50 Sn Tin 118. 70 120, 118, 116, 124,
119, 117, 122, 121,
112, 114, 115
51 Sb Antimony 121.77 121, 123
52 Te Tellurium 127.5 128, 130, 126
53 I Iodine 126.932 127
54 Xe Xenon 130.2 129, 132, 131, 134,
136, 128, 130, 126,
124
Sixth period
55 Cs Cesium. 132.81 133
56 Ba Barium 137.37 138, (136), (137)
568 APPENDIX
ATOMIC NUMBERS AND ATOMIC WEIGHTS (Continued)
Number Symbol Name Atomic Isotopes 2
weight 1
Sixth period (continued)
57 La Lanthanum 138.91 139
58 Ce Cerium 140.25 140, 142
59 Pr Praseodimium 140.92 141
60 Nd Neodymium 144.27 142, 144, 146, 145
61 Il Illini um 146.0*
62 {Sm
Sa
Samarium
\
J
150.43
63 Eu Europium 152.0
64 Gd Gadolinium 157.26
65 Tb Terbium 159.2
66 { Ds
Dy
Dysprosium } 162.52
67 Ho Holmium 163.4
68 Er Erbium 167.7
69 {Tm Thulium l 169.4
Tu
J
70 Yb Ytterbium 173.6
71 {Lu
Cp
Lutecium
Cassiopeim
l
)
( 175.0
72 Hf Hafnium 178.6
73 Ta Tantalum 181.5
74 w Tungsten 184.0
75 Re Rhenium 187 .4*
76 Os Osmium 190.8
77 Ir Iridium 193.1
78 Pt Platinum 195.23
79 Au Gold 197.2
80 Hg Mercury 200.61 202, 200, 199, 198,
201, 204 196
81 Tl Thallium 204.4
82 Pb Lead 207.20 206, 207, 208, (209),
(203,) (204), (205),
83 Bi Bismuth 209.00 209
84 Po Polonium 210(?)
85
86 rn
Em
Nt
?
Radon
?
Ra-emanation
Ni ton
)
212*
222
Seventh period
87 ? ? 223*
88 Ra Radium 225.95
89 Ac Actimum 229*
APPENDIX 569
ATOMIC NuMBERS .A.ND ATOMIC WEIGHTS (Continued)
Atomic Isotopes 2
Number! Symbol Name weight 1
Seventh period (continued)
90 Th Thorium 232.15
91 { Pa
UX2
Protoactini um
Uranium X2 } ?
234*
92 u Uranium 238.17
1 From International Critical Tables, except those marked*·
2 That is, mass numbers of constituent atoms, where known.
3 Alternative names and symbols are given in smaller type.
* Washburn, E. W., J our. Am. Chem. Soc., vol. 48, p. 2351 (1926). (Estimated by a
graphical method.)
APPENDIX I(b)
THE PERIODIC TABLE, GIVING ATOMIC NUMBERS AND ISOTOPES
Periods I I II III IV v VI VII VIII II
I {1 1.6"677
1
2
He
4
2 (a) 80, 78, 76, 82, 77, 74
(b) 84, 86, 82, 83, 80, 78, 79, 81
(c) 110, 111, 112, 113, 114, 116
1 4 25 36 41 58 69 710 8 (d) 120, 118, 116, 124, 119, 117, 122, 121,
31
II { L" Be B c N 0 F Ne 112, 114, 115
7, 6 9 11. 10 12 14 16 19 20, 22 (e) 129, 132, 131, 134, 136, 128, 130, 126,
1 12 213 314 415 5 16 617 7 18 8 124
III { 11 Mg Al Si p s CI Ar (f) 202, 200, 199, 198, 201, 204
Na
23 24, 25, 26, 28 27 28, 29, 30 31 32, 34, 33 35, 37 40, 36
I19 20
1 Ca
2 21
Sc
3 22
Ti
4 23
v
5 24
Cr
6 25
Mn
7
Fe
26 8127 Co 9128 Ni 101
IV K 39, 41 40, 44 45 48 51 52 55 56, 54 59 58, 60
01 11 so 14 33 15 34 16 35 17 36
r
29 12 31 13 32
--1
0 l 63, 65
Cu
64, 66, 68,
Zn
70 69, 71
Ga
74,
Ge
72, 70 75
As
(a)
Se
79, 81
Br Kr 1811
(b)
Rb
38
Sr
2 39
Yt
3 40
Zr
4 41
Cb
5 42
Mo
6 43
Ma
7 44
Ru 8145 Rh 9146 Pd 101
v 85, 87
1 88, 86 89 90, 94, 92
47 11 48 12 49 13 50 14 51 15 52 16 53 17 54 18
107, 109
Ag
(c)
Cd
115
In
(d)
Sn Sb Te
121, 123 128,130,126 127
I
t) 11
r 55 1 56 2 57-71 72 18 73 19 74
w
20 75 21 76
22177 23178 241
VI j Cs133, (?)
Ba
138, 136 * Hf Ta Re Os Ir Pt
3211 * 57
79 25 80 26 81 27 82 28 83 29 84 30 85 31 86 Rare Earths
Bi Nt
l Au
(f)
Hg Tl
3 90
Pb
209
5 92
Po
6
? La 61 II
58 Ce
59 Pr
62 Sa
63 Eu
65
66
67
Tb
Dy
Ho
69 Tu
70 Yb
71 Lu
vnlf
87 1 88 2 89 4 91
? Ra Ac Th Pa u 60 Nd 64 Gd 68 Er
Explanation of the Table.-The Roman numerals in the top row designate thP "columns" of the periodic table. ThP Roman numerals at the left
give the several ''periods" according to Bohr's arrangement (Appendix I(c)), and the italic numerals in the upper right-hand corners of the several
spaces give the ordinal numbers of the elements in these periods. Atomic numbers are indicated by boldface type in the upper left-hand corners. Iso-
topes are given by the numbers at the bottoms of the spaces.
APPEN DIX J(c)
BOHR'S PERIODIC TABLE OF THE ELEMENTS
Period Period
VI VII
55Cs-- 87-
56Ba--8 8Ra
--89Ac
90Th
59Pr 91Pa
Period Period 60Nd 92U
IV V 61Il
'i9K -37Rb 62Sa
20Ca --38Sr 63Eu
21Sc -39Y 64Gd
Period 22Ti --40Zr 65Tb
II 23V -41Nb 66Dy
3Li --llNa 67Ho
Period / / 4Be--12M g 25Mn- 43Ma 68Er
I / 5B --13Al 26Fe - 44Ru
lH// 6C --14Si 27Co - 45Rh
"-, 80 --168 29Cu - 47Ag
. "9F --17Cl 30Zn --48Cd
32Ge --50Sn
33As --51Sb
34Se --52Te
35Br -531
'36Kr - - 54Xe 79Au
80Hg
81Tl
82Pb
83Bi
84Po
8.5-
86Nt
571
APPENDIX II
ARRANGEMENT OF ELECTRONS IN ORBITS AccoRDING TO FooTE 1
:f. K Lr Lu Lnr Mr Mn Mnr Mry My Nr Nu Nnr Nry Ny Nyr Nvrr Or On Our OIV Ov Pr Pn Pnr Pry Q[
-- - -
n 1 2 3 4 5 6 7
Period Element - - --
k 1 1 2 2 1 2 2 3 3 1 2 2 3 3 4 4 1 2 2 3 3 4 1 2 2 3 1
- -
J 1 1 1 2 1 1 2 2 3 1 1 2 2 3 3 4 1 1 2 2 3 3 1 1 2 2 T
I
I lH 1
2 He 2
II I 3 Li 2 1
4 Be 2 2
5B 2 2 1
6C 2 2 2
vl 7N 2 2 2 1
-l
~ I 80
9F
2
2
2
2
2
2
2
3
lONe 2 2 2 4
m-i- 11
12
Na
Mg
2
2
2
2
2
2
4
4
1
2
13 Al 2 2 2 4 2 1
14 Si 2 2 2 4 2 2
15 P 2 2 2 4 2 2 1
16 S 2 2 2 4 2 2 2
17 Cl 2 2 2 i 2 2 3
18 A 2 2 2 i 2 2 4
CTI- 19
20
K
Ca
2
2
2
2
2
2
4
4
2
2
2
2
4
4
1
2
-
21 Sc 2 2 2 4 2 2 4 1 2
22 Ti 2 2 2 -i 2 2 4 2 2
23 V 2 2 2 4 2 2 4 3 2
24 Cr 2 2 2 4 2 2 4 4 1 1
25 Mn 2 2 2 4 2 2 4 4 1 2
26 Fe 2 2 2 4 2 2 4 4 2 2
27 Co 2 2 2 4 2 2 4 4 3 2
28 Ni 2 2 2 4 2 2 4 4 4 2
29 Cu 2 2 2 4 2 2 4 4 6 1
30 Zn 2 2 2 4 2 2 4 4 6 2
31 Ga
G 2 2 2 4 2 2 4 4 6 2 1
32 Ge 2 2 2 4 2 2 4 4 6 2 2
33 As 2 2 2 4 2 2 4 4 6 2 2 1
34 Se 2 2 2 4 2 2 4 4 6 2 2 2
35 Br 2 2 2 4 2 2 4 4 6 2 2 3
36 Kr 2 2 2 4 2 2 4 4 6 2 2 4
v 37
38
Rb
Sr
2 2
2
2
2
4
4
2
2
2
2
4
4
4
4
6
6
2
2
2
2
4
4
1
2
2
39 Y 2 2 2 4 2 2 4 4 6 2 2 4 1 2
40 Zr 2 2 2 4 2 2 4 4 6 2 2 4 2 2
41 Cb 2 2 2 4 2 2 4 4 6 2 2 4 4 1
42 Mo 2 2 2 4 2 2 4 4 6 2 2 4 4 1 1
43 Ma 2 2 2 4 2 2 4 4 6 2 2 4 4 2 1
44 Ru 2 2 2 4 2 2 4 4 6 2 2 4 4 3 1
45 Rh 2 2 2 4 2 2 4 4 6 2 2 4 4 4 1
46 Pd 2 2 2 4 2 2 4 4 6 2 2 4 4 6
47 Ag 2 2 2 4 2 2 4 4 6 2 2 4 4 6 1
48 Cd 2 2 2 4 2 2 4 4 6 2 2 4 4 6 2
49 In 2 2 2 4 2 2 4 4 6 2 2 4 4 6 2 1
50 Sn 2 2 2 4 2 2 4 4 6 2 2 4 4 6 2 2
51 Sb 2 2 2 4 2 2 4 4 6 2 2 4 4 6 2 2 1
52 Te 2 2 2 4 2 2 4 4 6 2 2 4 4 6 2 2 2
53 I 2 2 2 4 2 2 4 4 6 2 2 4 4 6 2 2 3
Vl 54 Xe 2 2 2 4 2 2 4 4 6 2 2 4 4 6 2 2 4
'l
c.,.;i 2 4 1
VI 55 Cs 2 2 2 4 2 2 4 4 6 2 2 4 4 6 2
56 Ba 2 2 2 4 2 2 4 4 6 2 2 4 4 6 2 2 4 2
57 La 2 2 2 4 2 2 4 4 6 2 2 4 4 6 2 2 4 1 2
58 Ce 2 2 2 4 2 2 4 4 6 2 2 4 4 6 1 2 2 4 1 2
59 Pr 2 2 2 4 2 2 4 4 6 2 2 4 4 6 2 2 2 4 1 2
60 Nd 2 2 2 4 2 2 4 4 6 2 2 4 4 6 3 2 2 4 1 2
61 Il 2 2 2 4 2 2 4 4 6 2 2 4 4 6 4 2 2 4 1 2
62 Sm 2 2 2 4 2 2 4 4 6 2 2 4 4 6 5 2 2 4 1 2
63 Eu 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 2 2 4 1 2
64 Gd 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 1 2 2 4 1 2
65 Tb 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 2 2 2 4 1 2
66 Dy 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 3 2 2 4 1 2
67 Ho 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 4 2 2 4 1 2
68 Er 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 5 2 2 4 1 2
69 Tm 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 6 2 2 4 1 2
70 Yb 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 7 2 2 4 1 2
71 Lu 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 1 2
72 Hf 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 2 2
73 Ta 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 3 2
74 W 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 2
1 Trans. Am. Inst. Min. Metal. Eng., vol. LXXIII, p. 628, February (1926).
C)l
--1
~
ARRANGEMENTS OF ELECTRONS IN ORBITS AccoRDING To FooTE (Continued)
* I KI Lr Lu Lui I M1 Mu Mm Mrv Mv I Nr Nu Nur Nrv Nv Nv1 Nvu 01 Ou Orn Orv Ov I Pr Prr Pru P,v IQI
I
Period Element
n j 1 I 2
I 3
I 4
I 5
I 6 \7
k I i I 1 2 2
I 1 2 2 3 3
I 1 2 2 3 3 4 4
I 1 2 2 3 3 4 I 1 2 2 3 Ii
I 1 I 1
j 1 2
I 1 1 2 2 3 I
1 1 2 2 3 3 4 I 1 1 2 2 3 3
I 1 1 2 2 I 1
VI 75 Rei 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 1 2
76 Os1 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 2 2
. 77 Ir 1 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 3 2 ~
78 PtI 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 4 2 ~
79 Au 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 6 1 ""tj
80 Hg 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 6 2 ttj
81 Tl 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 6 2
82 Pb 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 6 2
1
2
~
83 Bi 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 6 2 2 1 t1
..........
84 Po 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 6 2 2
85 - 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 6 2 2
2
3
~
86 Rn 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 6 2 2 4
VII 87- 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 6 2 2 4 1
88 Ra 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 6 2 2 4 2
89 Ac 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 6 2 2 4 1 2
90 Th 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 6 2 2 4 2 2
91 Pa 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 6 2 2 4 3 2
92 U 2 2 2 4 2 2 4 4 6 2 2 4 4 6 6 8 2 2 4 4 6 2 2 4 4 2
1 Alternative arrangements are suggested by Foote for these elements.
*K, L, M . . . = X-ray levels.
n = total quantum number.
k = azimuthal quantum number.
j =
inner quantum number.
APPENDIX III
SOME IMPORTANT CONSTANTS 1
c velocity of light . . . . . . . . . . . . . . 2. 9986 X 10 10 cm. sec. - 1
e electronic charge ............ . 4. 77 4 X 10-10 e.s. u.
e' electronic charge ............ . 1 . 592 X 10- 20 e.m. u.
m electronic mass ............. . 8 . 999 X 10-23 grams
e
5. 305 X 10 17 e.s. u. gm. -1
m
e'
1. 769 X 10 7 e.m.u. gm.-1
m
h Planck's constant ........... . 6. 554 X 10-27 erg-second
F the faraday ................ . 9. 6500 X 104 coulombs
R gas constant ................ . 1. 9869 calories deg. - 1-mole-1
R gas constant ................ . 8. 315 X 10 7 erg-deg. -1-mole-1
No Avogadro's number ......... . 6. 061 X 10 23 mole- 1
ko Boltzmann's constant ........ . 1 . 372 X 10-16 erg-deg. -1
mn mass of the hydrogen atom ... . 1 . 663 X 10- 24 grams
u Stefan's constant ............ . 5. 709 X 10-5 erg-cm. - 2-sec. - 1deg- 4
W Wien's displacement constant .. 0. 2885 centimeter degrees
N 00
Rydberg's wave number ..... . 1.0930 X 105 cm.-1
Grating space of calcite ...... . 3.0288 X 10-s cm.
Wave length of red cadmium
line ....................... . 6438. 4696 X 10-s cm.
1 From the International Critical Tables Vol. I. p. 18.
575
INDEX
A a rays, 550
charge carried by, 553
Abbreviated notation, Spectral electrostatic deflection of, 552
series, 305 magnetic deflection of, 552
Aberration of light, 48 Alpha particle, deflection of by
Abnormal Zeeman effect, 322* Rutherfords atom, 340
"Absorbing power," 186 by Thomson's atom, 338
Absorption bands, infrared, 120 excess scattering of, 339
coefficient of, 1861 454 in passing through matter, scat-
mass, 455 tering of, 337
limits, K and L, 498 Ampere, 64, 65
Mosely diagram for, 499 Ampere's rule, 64
(X-ray), determined from pho- Anaxagaras, 7
toelectric effect, 517 Angular distribution of scattered
of energy by atoms, 383 X-ray energy, 518
of X-rays, 454, 511 of scattered X-rays, 463
empirical formula for, 502 momentum, quantum rules for,
laws of, 497 365
theoretical formula for, 513 Angstrom, 290, 292, 296
variation with atomic number, Apparent failure of Bragg's law, 529
502 Arab~ 14
"Absorptive power," 186 Arago, 58, 59, 65, 78
Absorptivity, 182, 186 Arcetri, 24
and em1ss1ve power, relation Archimedes, 12, 13
between, 184 Argon, arrangement of electrons in
Academy of Sciences, Royal, 32 417
Accelerated charge, energy radiated Aristarchus, 11, 12, 13, 31
by, 128 hypothesis of, 5
rate of radiation from, 132 Aristotle, 8, 9
special cases, radiation by, 133 assertion of, 3, 7
Achromatic lenses, 37 doctrines of, 19, 20
Adams, 275, 277 estimate of size of earth, 10
Adiabatic process, 195 prestige of, 12
Alberti, 155 Arrangement of electrons in argon,
Al Hazen, 14, 31 417
Alkali elements, effective quantum in atoms, 412-446
numbers of, 439 in neons, 416
Allen, H. S., 273 in orbits, 572
Allen, S. J. M., 501 in Thomson atom, 336
Allison, 525 Assignment of quantum numbers to
a particles, "counting," 554 electron orbits, 444
* References of major importance indicated by bold-face type.
577
578 INTRODUCTION TO MODERN PHYSICS
Aston, 536, 539, 547 Balmer, 295
Aston's "Mass Spectograph," 539 Balmer's formula for hydrogen spec-
Atom, number of electrons per, 331 trum, 295
stationary states of the, 383 series, 317, 350, 364, 388, 397
model building, problem of, 333 fine structure of lines of, 378
for heavier elements, static, 425 origin of on Bohr's theory, 356
Atomic members, table of, 566 Band spectra, 320, 407
frequencies, Einstein's equation spectrum of, nitrogen, 321
for the, 264 Bands, "electronic," 320
heat, classical value of, 253 "rotation," 320
Einstein's equation for, 257 vibration, 320
test of, 258 Barberini, 22
of gases, classical theory, 278 Barkla, 331, 344, 463, 466, 468, 487,
of solids, Einstein's theory, 256 518, 528
heats at very high tern per a tures, Barkla's determination of electrons
274 per atom, 331
at very low temperatures, 272, Bartholinus, 57
274 Basilewski, 452
Compton's theory of, 275 Becker, J. A., 525
Debye's theory of, 268 Becquerel's discovery of radioactiv-
N ernst-Lindemann, formula for, ity, 549
267 Bennet, 190
of solids with temperature, Benoist, 290, 453, 454
variation of, 250 Bergmann, 304
of various elements, 249 Bernoulli, Daniel, 46
hypothesis, of Democritus, 8 Bertin-Sans, 455
number, 344 {3-ray spectograph, 559
number and properties of inert {3-rays, 550
gases, 413 Birge, 246
properiies, spectral series and, 314 Bjerrum, 284
scattering coefficient, 501 Black, 561
structure, early views on, 330 "Black Body," 183
weights, table of, 566 distribution of energy in spectrum
Atoms, absorption of energy by, 383 of, 204
and electrons, collisions between, emissive power of a, 187
385 radiation, 177, 183
excitation of, 383 energy distribution in, 227
masses of, 536 spectral distribution of, 203
Atwood, George, 46 Black, James, 47
Avogadro's constant, 120 Bless, 525
number, 171, 219 Bleunard, 453
Ayres, 518 Bodies, falling, Aristotles' views on,
Azimuthal quantum number, 366, 11
433 Bohr, 289, 313, 346, 433, 446, 486
orbits, energy corresponding to,
B 352
Babinet, 59 radiation from, 353
Babylonians, 6 Bohr's extension, nuclear atom
Bacon, Roger, 17 model, 350
INDEX 579
Bohr's formula for the hydrogen Cario, 392
spectrum, 355 Carlisle, 63
for frequency of spectral lines, Carnot, 54
354 engine, 194
postulates, 351, 354 Carnot's cycle, 54, 179, 194, 203
theory, Balmer's series, origin of efficiency of, 196
on, 366 for a perfect gas, 196
elliptical orbits in, 364 Cathode rays, J. J. Thomson's
Boltzmann, 194 experiments on, 150
constant, best numerical value of, measurement of velocity of, 151
Stefan-, 202 properties of, 149
the Stefan-, 200 Cauchy, 59
law, experimental verification of Cavalieri, 33
the Stefan-, 201 Cavendish, Henry, 18, 50, 51
the Stefan-, 194, 200, 203, 243, experiment, 50
244 Centimeter, electrons per cubic, 120
Boltzmann's constant, 219, 232, 252 Chaldeans, 5
"ether engine," 197 Chamber, ionization, 453
Born and Karman, theory of, 274 Characteristic frequencies, 260
Bothe, W., 525 compressibilities, and, 263
''Bound'' electrons, 166 from Debye's formula, 273
Boyle, Honorable Robert, 46 radiation, 178
"Boyle's Law," 46 temperatures, 259
Brackett series, 364, 397 for several elements, (Einstein),
spectrum of hydrogen, 368 259
Bradley, 48 secondary radiation, X-rays, 467
Bragg, W. H., 483 X-ray spectra, 502
Bragg, W. L., 480, 502 X-rays, discovery of, 456
Bragg's discovery of monochro- Charge carried by a-rays, 553
matic characteristic radiations, Chemical elements, periods of the,
483 427
law, 476 Chemical properties of the lighter
apparent failure of, 529 elements, 418
Brewster, Sir David, 35, 45 Clairau t, 46
Bronsted, 546 Classical physics, 1, 4
Brougham, Henry, 57 theory of specific heats of solids,
Brown, Robert, 220 252
Brownian movements, 220 thermodynamics, application to
Bruno, 18 the radiation problem, 205
Buisson, 327, 3 79 value of atomic heat, 253
Bury, 417, 429 Coblentz, 247
spectroradiometer, 203
c Coblentz, W. W., 202, 212
Cabeo, 33 Coefficient of absorption, 186, 454
Cadmium line, wave length of, 292 Coefficients for X-rays, scattering,
Calculus, 46 517
Caloric theory of heat, 53, 54 Collision of the second kind, 392
Canal rays, 536 Collisions between a toms and
Carbon, static atom model for, 421 electrons, 385
580 INTRODUCTION TO MODERN PHYSICS
Color vision, Maxwell's work on, 74 Cruikshank, 63
Combination lines, 311, 312, 363 Crystal grating, elementary theory
"Commentaries on the Motion of of, 473
Mars,'' 30 Crystal diffraction grating, 472
Compressibilities and characteristic Crystals, grating space of some, 482
frequencies, 263 Cubical atom, 417
Compton, A. H., 275, 235, 513, 518, Current density, 79, 81
534 Currents, displacement, 86
Compton, K. T., 159, 161
Compton, 519, 520 D
effect, 362, 517
theory of, 523 D lines, 298
wave length shift in, 525 D' Alambert, 46
Compton's theory of atomic heats, da Vinci, Leonardo, 17
275 Dalibard, 50
Conditions for equilibrium in Thom- Dal ton, 8, 330
son's atom, 334 Danz, 274, 277
Conservation of energy, 68 Dark lines in solar spectrum, 59
law of, 53 Davis, 394, 532
Constant, best numerical value of, Davy, Sir Humphrey, 54, 60
Stefan-Boltzmann, 202 de Broglie, 513
Boltzmann's, 219, 232, 252 de la Rive, 65
for H and He, Rydberg, 361 Debye, 259
molecular gas, 219 Debye's equation, test of, 271
"Stefan-Boltzmann," 200 for the specific heat, 271
"Stefan's," 200 formula, characteristic frequen-
variation in, Rydberg, 362 cies, from, 273
Continuous X-ray spectrum, 489 for atomic heats, 277
intensity of, 495 theory of atomic heats, 268
limiting wave lengths in, 490 Decomposition of water, 63
maximum frequency in, 492 Degree of freedom, mean energy per,
Convergence frequency, 306 232
frequencies for lithium, 306 Degrees of freedom, 214
Convergent wave number, 302 associated with ether vibrations,
Coolidge, Dr. W. D., 451, 493 221
Coolidge tube, 451 per unit-length of a vibrating
X-ray tube, 451 string, 222
Copernicus, 12, 13, 14, 17, 27, 48 per unit volume for ether vibra-
Copernican system, 11, 14, 28, 48 tions, 226
Corpuscles, 333 of an enclosure, 226
Corpuscular theory of light, 173 rotation, suppression of, 286
photoelectric effect, and the, 173 sta tis ti cal distribution of energy
'' Cosmographicum, Mysterium,'' among, 229
29 suppression of, 285
Coulomb, Charles A., 50, 51 vibration, suppression of, 287
Creation and destruction of matter, 7 Dember, 167
Crescent phases of Venus, 21 Democritus, 5, 7, 8
Crookes, 190 Dempster, 536
dark space, 148 Dempster's "mass Rpectograph.," 541
INDEX 581
Density of radiation, 182 Distribution of electrons in orbits,
Descartes, 33, 36 433
Deslandres, 321 of velocities, Maxwell's law of, 229
Destructive interference of second- schemes for electrons in atoms, 445
ary wave trains, 475 Ditscheiner, 296
Determination of grating space for Doan, 534
NaCl, 481 Dolland, 37, 47
Dewar, 293, 297 Doppler effect on spectral lines, 327
di Novara, 15 Doppler's principle, 206, 207, 326
Diagrams, energy-level, 396 "Double" X-ray spectrometer, 532
Dielectric constant, 111 Doublet separations, table, 309
'' Dialogues of the Two New Sci- series, relations between, 308
ences," 24 Doublets in the spectrum of sodium,
Diatonic gas, molecular heat of, 280 298
molecule, quantum conditions Drude, 190, 261, 332
for rotation of, 408 Du Fay, 49, 190
Differential equations of electro- Duane, 111, 525
magnetic wave, 94 Dulong and Petit, 248
Differential equations of wave empirical law of, 248
motion, 98 law of, 251, 271
Difficulties with theories concerning Dynamic atom model, 412, 433
photoelectric energy, 170 Dynamo, Faraday disk, 66
Diffraction bands, 56
grating, crystal, 4 72 E
first use of, 291 e' / m for electrons, from cath ode
of X-rays, early observations, 457 rays, 152
Diffuse series, 299 for photoelectrons, Lenard's
Direct measurement by a prism, method, 155
refraction for X-rays, 532 Lenard's values, 157
observation of the photoelectric ratio of, 142
effect of X-rays, 514 value from Zeeman effect, 148
Discovery of electromagnetic waves, Earth, Aristotle's estimate of size
110 of, 10
of Uranus, 48 Early search for series relations in
Roentgen's, 447 spectra, 292
Disintegration series, radioactive, theories, X-rays, 457
556 Eclipses, 5, 7
Dispersion, 136 of Jupiter's moons, 48
formula, Sellmeier's, 118 Effect, Compton, 517
Lore11tz theory of, 155 of variation of mass of electron on
ofligh~ 3~ 11~ 142,332 elliptical orbits, 374
"Displacement constant, Wien's," Zeeman, 143
212 Effective quantum numbers, 434
currents, 86 electron orbits determined
law, energy-temperature, 209 from, 438
wave-length-temperature, 206 in lithium, 436
Wien's, 205, 212, 242 method of determining values
laws, experimental confirmation, of, 435
211 of alkali elements, 439
582 INTRODUCTION TO MODERN PHYSICS
Ehrenfest, 284 Electromagnetic theory of radiation
Eighteenth century, electricity pres sure, 190
during the, 49 unit of electric current, the, 81
heat during, 46 of quantity of electricity, the,
light during the, 47 81
mechanics during, 46 wave, the, 101, 129
Einstein, 165, 220, 264, 283 differential equations, of, 94
characteristic temperatures, for flow of energy in, 103
several elements, 259 intensity of, 108
Einstein's equation for atomic heat, plane-polarized, 103
257 velocity of, 108, 109
for the atomic frequencies, 264 waves, discovery of, 110
test of, atomic heat, 258 index of refraction for, 115
photoelectric equation, 165, 174 measurements of velocity of,
theory, atomic heat of solids, 256 111
of specific heats, 284 Electromagnets, Henry's work on,
Elastic collisions, 388 71
Electric and magnetic fields, sym- Electron, discovery by Sir J. J.
metrical relations between, 93 Thomson, 148
charge, equivalent mass of, 128 orbits, assignment of quantum
convection, 123 numbers to, 444
current, discovery of, 63 determined from effective quan-
electrostatic unit of, 79 tum numbers, 438
the electromagnetic unit of, 81 orientation of, 443
motor, Faraday's discovery of, 64 radius of the, 527
wave, 101 Electronic band spectra, 411
Electrical constitution of matter, 143 bands, 320
nature of matter, 109 charge, e, 141
units, electromagnetic system of, direct measurement of, by H.
80 A. Wilson, 153
electrostatic system of, the, 78 Electrons, 118, 153
Electricity during the eighteenth deflection of in passing through
century, 49 matter, 332
theory of, one-fluid, 50 "free," 154
two-fluid theory of, 50, 52 from ca th ode rays, e' / m for, 152
Electrolysis, discovery of, 63 in atoms, arrangement of, 412-
the laws of, 68, 140 446
Electromagnet, invention of, 70 distribution schemes for, 445
Electromagnetic equations, 92, 102 in orbits, distribution, 433
field, Maxwell's equations of, 83, per atom, Barklas determination
91 of, 331
induction, 70 from sea ttering of X-rays, 463
Faraday's discovery of, 65 number of, 331
Faraday's laws of, 67 per cubic centimeter, 120
system of electrical units, 80 Electronic charge, accepted value
theory, 75, 76 of, 154
fundamental tenets, 121 charge, Millikan's measurement
Maxwell's, 67 of, 154
of light, the, 78-122, 108 mass, accepted value of, 154
INDEX 583
Electroscope, invention of, 50 Energy, distribution in black-body
Electrostatic deflection of X-rays, radiation, 227
552 Planck's equation for, 237
field strength, 80 equipartition of, 214
system of electrical uni ts, the, 78 -level diagram for hydrogen atom,
unit of electric current, 79 399
of quantity of electricity, the, sodium atom, 401
79 for the radium B, 562
Electrostatics, Cavendish's work on, diagrams, 396
51 X-ray, 608
Elster and Geitel, 138, 139, 157, 158, levels, nuclear, 661
167 per degree of freedom, 218
Elemental Theory of crystal grating, radiated by accelerated charge,
473 128
Elements, (after Bohr) periodic radiation of, according to electro-
table of the, 571 magnetic theory, 110
Elliptical orbits, effect of variation released in formation of helium,
of mass of electron on, 374 545
energy corresponding to, 369, -temperature displacement law.
376 209
in Bohr's theory, 364 theorem of equipartition of, 216
penetration of, 437 total, of, ether pulse, 131
Ellis, 558, 560, 561, 564 Enhanced lines in helium, 319
"Emission," 186 spectra, 318
Emissive power, monochromatic, Equation, Einstein's photoelectric,
179, 186 165
of a black body, 187 Equipartition of energy, 214
relation between absorptivity theorem of, 216
and, 184 Equivalence of matter and energy,
total, 179, 186 544
''Emissivity,'' 186 Equivalent mass of electric charge,
monochromatic, 180 128
total, 180 of heat, mechanical, 55
Empedocles, 7 "Ether," 109
Empirical formula for absorption engine, Boltzmann's, 197
of X-rays, 502 pulse, intensity of, 131
Empirical laws of absorption of X- origin of, 129
rays, 497 theory of X-rays, 458, 458
law of Dulong and Petit, 248 thickness of, 129
total energy of, 131
Energy, conservation of, 68
''string" theory of light, Thom-
contained in the magnetic field
son's, 175
surrounding moving charge,
Euclid, 12
127 Euler, 46, 190
corresponding to elliptical orbits, Excitation by collision, 391
369, 376 by radiation, 389
density of radiation, 108, 182, Excitation of atoms, 383
202, 222 of sodium vapor, 386
of solar radiation, 192 potentials, 385
584 INTRODUCTION TO MODERN PHYSICS
Excited state, 384 Formulre for spectral series, 305
Experiment of Friedrich, Knipping, radiation, 205
and Laue, 478 Fowler, 306, 320
Experimental confirmation, dis- Foucalt, 59
placement laws, 211 Fraction, packing, 548
Experiments and theories, some Franck, 392
early X-ray, 460 Franklin, Benjamin, 50
"Extinction coefficient," 261 Franklin's kite experiment, 50
Fraunhofer, 59, 291
F '' Free '' electrons, 154, 166, 168
Fabry, 290, 327, 379 Freedom associated with ether
Failure of classical theory, scatter- vibrations, degrees of, 221
ing of X-rays, 521 degrees of, 214
Falling bodies, Aristotle's views on, energy per degree of, 218
11 per unit length of a vibra ting
Galileo's experiment on, 20 string, degrees of, 222
Faraday, Michael, 2, 4, 60 per unit volume for ether vibra-
Faraday, 70, 72, 93, 123, 140, 141, tions, degrees of, 226
143, 330 of an enclosure, degrees of, 226
disk dynamo, 66 Frequency of spectral lines, Bohr's
effect, the, 69 formula for, 354
Farraday's discovery of, electric of the light, effect of, on velocity of
motor, 64 photoelectrons, 162
laws of electromagnetic induction, Fresnel, 58, 65, 78, 190
67 Friedrich, 4 73
Field strength, (electrostatic), 80 Friedrich, Knipping and Laue,
magnetic, 80 experiment of, 478
Fine structure components of Ha, "Fundamental series," 304
wave-number differences be- tenets, electromagnetic theory,
tween, 381 121
of Ha, 382
of spectral lines, 372 G
Fizeau, 25, 59
Flow of energy in electromagnetic Galileo, Galilei, 18, 57
wave, 103 lenses, 36
Fluorescence, 178 Galileo's experiment, 3
Fluorescent X-radiation, Barkla's on, falling bodies, 20
discovery of, 4 72 Galvani, 60
X-rays, 470, 514 Gamma-ray spectra, 668
Fluorine, static a tom model for, 420 '}'-ray spectrum of mesothorium, 662
Fluxions, 45 of radium B, 661
Foote, 385, 393 of thorium D, 663
Foote, P. D., 446 'Y rays, 550
Formula, Rayleigh-Jeans, the, 227, wave length of, 660
236, 241 "Gamma's rays," 178
Rayleigh's, 205 Gas, molecular heat of, diatonic, 280
test of, Rayleigh-Jeans, 227 monatonic, 279
Thiessens, 206 Gases, classical theory, atomic heat
Formula, W ein's, 206, 228 of, 278
INDEX 585
Gases, molecular heat of, variation Heat, some early measurements of,
with temperature, 281 specific, 249
Gassendi, 33 Heats, atomic, of various elements,
Gehrcke, 379 249
Geiger, 339, 342, 344, 345 at very high temperatures, 274
Gellibrand, 33 at very low temperatures, 272,
General equations of wave motion, 274
97
Heath, 5, 6
Geocentric hypothesis, 6
Heisenberg, 446
Gilbert, 32
Helium, Bohn's formula for the
Goucher, 394
Graph, Moseley, 487 spectrum of, 358
Grating space, determination of for energy released in formation of,
NaCl, 481 545
of some crystals, 482 enhanced lines in, 319
Greek theories of vision, 8 Rydberg constant for, 319
Greeks, the, 5 Helmholtz, 55, 69
Gregory, James, 37 Henry, 111
Gravitation, inverse-square law of, Henry, C., 452
43 Henry, Joseph, 70
Kepler's speculations on, 31 Herschel, William, 48
law of, 12 Hertz, 110, 121, 136
Newton's work on, 42 experiments of, 1
principle of universal, 45 Hicks, 294, 303, 320, 324
Gray, Stephen, 49 Hj aimer, 530
Griffith, 158 Hooke, Robert, 3$, 46
Grti.neisen, 255 Hoor, 166
Guericke, 33 Houstoun, R. A., 328
Houstoun, W. V., 379
H Huggins, 296
h, best value of, 246 Hull, 192
numerical value of from radiation Hull, A. W., 515
data, 245 Hund, 446
value from photoelectric effect Hurmuzescu, 453
Hand He, Rydberg constant for, 361 Huyghens, 3~ 4~ 175
Ha, fine structure of, 382 wave theory of light, 4
Hafnium, 446 Hydrogen atom, energy-level dia-
Haga, 473 gram for, 399
* Hall, 37, 47 stationery states of, 397
Hallwachs, 138 -like terms, 318
Halm, 315 lines, wave lengths of, 296
Hartley, 293 molecular heat of, variation with
Hartley's law, 293, 297 temperature, 282
Harvey, 546 spectrum, Balmer's formula for,
Heat during the eighteenth century, 295
46 Bohr's formula for the, 355
materialistic theory of, 4 7 Hypothesis, atomic, of democritus, 8
mechanical equivalent of, 55 geocentric, 6
mechanical theory of, 54 Prout's, 330
586 INTRODUCTION TO MODERN PHYSICS
I Joule, 53, 55, 69
Imbert, 455 Jupiter, moons of, 21
Jupiter's moons, eclipses of, 48
Index of refraction, 112
for electromagnetic waves, 115 K
X-rays, 457
classical formula for, 530 K and L absorption limits, 498
Induced electromotive forces, 93 K characteristic secondary radiation,
Inelastic collisions, 388 485
Inert gases, atomic numbers and K fluorescent radiation, 472
properties of, 413 Klines, wave lengths of, some, 503
melting points and boiling Karman, theory of Born and, 27 4
points of, 414 Kaufmann, 372
Infra-red absorption bands, 120 Kayser, 297
Inner quantum numbers, 402, 405, Kelvin, 55, 71
442 scale, 202
selection rules for, 406 Kemble, 281, 284
Inquisition, 24 Kepler, 16, 18, 26, 48, 190
Intensity of continuous X-ray spec- planetary motion, 35
trum, 495 Kepler's laws, 30, 44
of electromagnetic wave, 108 second law, 28
of ether pulse, 131 speculations on gravitation, 31
of illumination, photoelectric cur- third law, 30
rent, effect of, 167 work on optics, 31
of radiation, 180 Keplerian telescope, 31
Interference, 175 Kerr, 69
fringes, 56 Kinetic theory of matter, 77
principle of, 56 Kircher, 33
Invention of electroscope, 50 Kirchoff, 185
of, telescope, 20 Kirchoff's law of radiation, 185
Inverse square law of electrostatics, Kirkpatrick, 496
51 Kite experiment, Franklin's, 50
of gravitation, 43 Knipping, 473
of radiation, 131 Kramers, 497
Ionization, 383 Kramers, H. A., 513
chamber, 453 Kulenkampff, 497
measurements, X-rays, 462 Kunz, 158, 167
potential, 389 Kurlbaum, 205
potentials, 394
of mercury vapor, 395 L
Ionized atoms, terminology for, 413 Lablesse, 453
Isothermal process, 195 Ladenburg, 165, 167
Isotopes, 639 Lagrange, 46
of the elements, 644 Langmuir, 413, 417
separation of the, 646 Larson, 531
·Ives, 167 La Place's Law, 82
Lau, 379
J Law of the conservation of energy, 53
Jeans, 221 of gravitation, 12
Joffe, 165 of Dulong and Petit, 251, 271
INDEX 587
Laws, Kepler's, 30, 44 Lines of force, 67
of electrolysis, 68, 140 Lipperhy, 20
of scattering of a-particles, 342 Lithium, convergence frequencies
Lay, 445 for, 306
Lebedew, 192 effective quantum numbers m,
Leibnitz, 46 436
Lenard, 155, 157, 163, 393 static atom model for, 418
Lenard, P., 139 Liveing, 293, 297
Lenses, achromatic, 37 Loeb, 220
Lewis, 262 Lorentz, H. A., 143, 372
Lewis, G. N., 413 Lorenz, L., 142
Lewis, W. M. C., 251 Lorentz theory of dispersion, 155
Leyden jar, 50 Loria, 392
Light, aberration of, 48 Lummer, 289
classical theory of, conflict with and Pringsheim, 201, 202, 204,
photoelectric effect, 173 214
corpuscular theory of, 38, 173 Lyman, 357
and photo-electric effect, 173 series, 319, 364, 388, 397
dispersion of, 37, 59, 116, 142, 332 spectrum of hydrogen, 367
during the eighteenth century,
47 M
indices of refraction of, for various
substances, 112, 113 MacLean, 111
Newton's speculations concerning Magnetic deflection of a-rays, 552
nature of, 38 field, due to wire carrying current,
polarization of, 57 82
quanta, 165 effect of, on vibrating charge,
refraction of, 111 146
the electromagnetic theory of, 78- produced by moving charge,
122, 108 123
Thomson's '' ether string'' theory strength, the, 80
of, 175 pole, the unit, 80
velocity of, 25, 109 wave, 101
Fizeau's measurement, 59 Magnus, A., 250, 254, 255, 274, 277
Romer's measurement, 48 Malus, 57
wave theory of, 56 law of, 58
waves, transverse, Young's sug- Marangoni, 453
gestion of, 58 Marci, 33
Lighter elements, chemical proper- '' Mars, Commentaries on the
ties of the, 418 Motion of," 30
Limiting wave lengths in continuous motion of, 28
X-ray spectrum, 490 Marsden, 342, 344, 345
Lincean Society, 32 Mascart, 293, 296
Lindemann, 251, 255, 265 Mass absorption coefficient, 455
Lindemann's melting point formula, coefficients, variations of, with
265, 270 wave length, 500
Line spectra, series relations in, Mass number, 548
289-329 of electron, ratio to mass of
Linear oscillator, Planck's, 346 hydrogen a tom, 359
588 INTRODUCTION TO MODERN PHYSICS
Mass sea ttering coefficient, 501 Merritt, 155
spectograph, Aston's, 639 Mersenne, 33
Dempster's, 541 Mesothorium, ')'-ray spectrum of,
Masses of a toms, 636 662
by precision mass spectograph, Method of determining values of
effective quantum numbers, 435
648
Michelson, 290, 327, 328, 379
Mastlin, 27
Micron, 290
Matter and energy, equivalence of,
Miletus, Thales of, 6
544 Milky Way, 7
ere a tion and destruction of, 7 Millikan, 163, 165, 167
electrical constitution of, 143 Millimicron, 290
electrical nature of, 109 Modern physics, 1
kinetic theory of, 77 Mohler, 385, 393
Maximum frequency in continuous Molecule, moment of inertia of,
X-ray spectrum, 492 monatonic, 279
Maxwell, 51, 78, 86, 111, 123, 190, 229 Molecular gas constant, 219
Maxwell, James Clerk, 72 heat of gases with temperature,
Maxwell's distribution law, 286 variation of, 281
electromagnetic theory, 67 of hydrogen, variation with tem-
of light, 1, 4 perature, 282
equations of electromagnetic field, theory, Maxwell's work on, 75
83,91 spectra, 407
law of distribution of velocities, Momentum of a quantum, 521
229 quantum rules for, angular, 365
work on color vision, 74 radial, 365
on molecular theory, 75 Monatonic gas, molecular heat of,
Mayer, 69 279
Mayer, R. J., 54 molecule, moment of inertia of,
McLennan, 445 279
Mean energy per degree of freedom, Monochromatic characteristic radia-
232 tions, Bragg's discovery of, 483
Measurement by total reflection, emissive power, 179
refraction of X-rays, 634 emissivity, 180
Mechanical equivalent of heat, 54, Moons of Jupiter, 21
55 Mosley, 345, 413, 415, 485, 502, 508
Mechanics during the eighteenth diagram, absorption limits, 499
century, 46 graph, 487
Meggers, 385, 393 Moseley's law, 486
Meissner, 379 values for wave lengths of X-ray
Melting point and characteristic lines, 486
frequencies, 265 Motion of Mars, 28
formula, Lindemann's, 265, 270 Movements, Brownian, 220
Melting points and boiling points of Moving charge, energy contained in
inert gases, table, 414 magnetic field surrounding,
Mendenhall, 296 127
Mercury vapor, ionization poten- force acting on in a magnetic
tials, 395 field, 126
resonance potentials of, 396 magnetic field produced by, 123
INDEX 589
Moving electron, relativity formula Oersted, 64
for mass of, 373 Optics, Kepler's work on, 31
Multiple scattering, 339 Opticks, Newton's, 39
Multiplicities in spectral lines, 404 Oppenheimer, J. R., 513
Mysterium cosmographicum, 29 Orbits, privileged, 354
Orientation of electron orbits, 443
N Origin of the photoelectrons, 166
of quantum theory, 177-247
National Academy of Sciences, 72 of radioactive rays, 555
Nature of light, Newton's specula- of spectral lines, 330-411
tion's concerning, 38 Oxygen, static atom model for, 421
Neon, arrangement of electrons in,
416 p
Nernst, 251, 255, 266
N ernst-Lindemann formula for Packing effect, 64 7
atomic heats, 267 fraction, 548
Newton, 2, 4 7, 173, 175, 190, 291 Padua, 15, 20
Newton, Isaac, 25, 34 Pascal, 33
Newton's contempories, 46 Paschen, 379, 380
"Op ticks," 39 series, 397
"Principia," 44 spectrum of hydrogen, 367
"queries," 40 Pauli, 446
rings, 56 Penetration of elliptical orbits, 437
work on gravitation, 42 Periodic table of the elements, 570
Nichols, 192, 261 after Bohr, 571
Nicholson, 63 Periods of the chemical elements,
Nitrogen, band spectrum of, 321 427
static atom model for, 421 Perfect absorber, 187
Normal Zeeman effect, 322 gas, Carnot's cycle for a, 196
Novak, 453 radiator, 187
Nuclear a tom, 330-411 Permeability, 111
difficulties with, 345 Perot, 290
model, Bohr's extension, 360 Perrin, 149
Rutherford's, 340 Phase integral, the, 346, 366
energy levels, 661 Philolaus, 6, 12, 15
type of atom, Rutherford's, 279 theory of, 11
Nuclei of the radioactive elements, Phosphorescence, 178
structure of, 663 Photoelectric cell, 158
Nucleus, the, 636-666 current, 138
Number, Avogadro's, 219 effect of intensity of illumina-
Numerical value of h, from radiation tion, 167
data, 245 effect of potential of cathode
Numerical value of Planck's con- on, 156
stant, from radiation data, 245 effect, 136-176
and the corpuscular theory of
0 light, 173
conflict with classical theory of
Observatory of U raniborg, 26 light, 173
Octet, an, 417 discovery by Hertz, 136
590 INTRODUCTION TO MODERN PHYSICS
Photoelectric effect, influence of Planets, revolution of, ancient data,
temperature on, 166 6
of X-rays, 511 Plato, 7, 9
direct observation of the, 514 Plutarch, 15
time lag in, 172 Pohl, 473
energy, source of, 168 Poisson, 52
equation, Einstein's, 165, 174 Polarization by reflection, discovery
mechanism, difficulties with theo- of, 57
ries concerning, 170 of X-rays, 465
phenomena, resonance theory of, Polarized light, rotation of plane of
170 polarization of, 69
Photoelectricity, 76 Positive-ray spectograph, Thom-
'' Photoelectrons,'' 156 son's, 537
e' /rn for, Lenard's values, 157 rays, 636
effect of frequency of the light Postulates, Bohr's, 351, 354
on velocity of, 162 fundamental, concerning static
energy given to, 163 atom, 417
Lenard's method, e' /rn for, 155 Potentials, excitation, 385
maximum velocity of, 160, 161 ionization, 394
number of atoms involved m Poynting's theorem, 106
production of, 171 Precision mass spectograph, masses
origin of the, 165 of atoms by, 548
velocity distribution curves for, Pressure, effect on spectral lines,
168 328
Photographs, X-ray, 452 of radiation, 196
Physics, classical, 1, 4 radiation, experimental determi-
modern, 1 nation of, 193
Picard, 44 Prevost, 178
Pickering series, 319 "Principia," Newton's, 44
Pisa, 19, 20 Principle, Doppler's, 206, 207
leaning tower of, 3 of interference, 56
Planck, 165, 229, 289 of relativity, 373
"Planck's constant," 164, 240 "of Universal Gravitation," by
best value of, 246 Newton, 45
determination of, from X-ray Principal series, 299
data, 494 Pringsheim, 289
numerical value of, from radi- Privileged orbits, 354
ation data, 245 radii of, 352
equation for energy distribution Production of X-rays, 468
in black-body radiation, 237 Properties of "cathode rays," 149
formula, 234
Prout, 330
linear oscillator, 346
Prou t's hypothesis, 330
quantum hypothesis, 313
radiation law, 229 Ptolemy, 14
verification of, 240 of Alexandria, 13, 16
theory of radiation, radical nature on refraction, 13
of, 238 Plirbach, 15
Plane-polarized electromagnetic Pythagoras, 6, 417
wave, 103 Pythagorean system, 15
INDEX 591
Q Radiation resonance, 389
similarity to perfect gas, 193
Quantum, 164, 313 thermal, 177
conditions for rotation of diatonic Radii of privileged orbits, 352
gas molecule, 408 Radioactive disintegration series,
momentum of a, 521 556
rules for angular momentum, 365 radiations, types of, 550
for radial momentum, 365 rays, origin of, 555
theory, 122 Radioactivity, 549
birth of the, 237 Becquerel's discovery of, 549
of specific heats, 248-288 Radium, 550
origin of, 177-24 7 B, energy-level diagram for the,
{luantity of electricity, elertro.. 562
magnetic unit of, 81 X-ray spectrum of, 561
electrostatic unit of, 79 Radius of the electron, 527
Quartz, index of refraction glycerine, Ramsauer, 163
dielectric constant of, 114 Rare-earth period, 430
'' Queries,'' Newton's, 40 Ratio of e/m, 142
to mass of hydrogen atom, mass
R of electron, 359
Rayleigh, 214
Radial momentum, quantum rules Rayleigh-Jeans formula, the, 227,
for, 365 228, 236, 241
quantum number, 366 test of, 227
"Radiating power," 186 radiation law, 221
Radiation, black-body, 177, 183 Lord, 221, 390
by accelerated charges, special Rayleigh's formula, 205
cases, 133 Rays from certain solids, residual,
characteristic, 178 262
density it,, 191 Reflecting telescope, 37
density of, 182, 191 Reflection, polarization by, dis-
energy density of, 108, 182, 202, covery of, 57
222 Reflectivity, 182, 260
excitation by, 389 Refracting telescopes, 37, 47
formulm, 205 Refraction for X-rays, direct meas-
from vibrating charge, 133 urement by a prism, 532
intensity of, 180 index of, 112
inverse-square law of, 131 of light, 111
Radiation law, Planck's, 229 of X-rays, 528
Rayleigh-Jeans, 221 early observations, 457
verification of, Planck's, 240 measurement by total reflec-
pressure of, 190 tion, 534
electromagnetic theory of, 190 Ptolemy on, 13
experimental determination of, Regiomontanus, 15
193 Regnault, 249
radical nature of, Planck's theory Reinforcement of secondary wav~
of, 238 trains, 475
rate of, from accelerated charge, Relation between absorptivity and
132 emissive power, 184
592 INTRODUCTION TO MODERN PHYSICS
Relation between doublet series, 308 Runge, 297
to absorption limits, X-ray emis- Rutherlord, 33~ 486, 563
sion lines, 505 Rutherford-Bohr atom model, SUC;,-
Relations between spectral series, cesses of, 366
304 Rutherford's nuclear atom, 340
between triplet series, 310 type of a tom, 279
Relativity formula for mass of mov- Russell, 405
ing electron, 373 Russell-Saunders notation for
change of mass, 373 spectral terms, 405, 442
principle of, 373 Rydberg, 297, 312, 415
Representative lines, Zeeman pat- constant, for H and He, 361
terns of some, 324 for helium, 319
Residual rays from certain solids, 262 numerical values of, 303
Resistance, change in, of selenium, variation in, 362
136 Rydberg's formula for spectral lines,
Resonance of sodium vapor, 389 302
potentials of mercury vapor, 396 for spectral series, 297
radiation, 389 Rydberg-Schuster law, 308, 363
Reststrahlen, 260
Revolution of planets, ancient data, s
6
Richardson, 159, 161 Saunders, 111, 405
Richtmyer, F. K., 501, 502, 513 Satellites, 311
Righi, 456 Scales, thermometric, 4 7
Ritchie, 184 Scattered radiation, wave length of,
Ritter, 63 619
Ritz, 312 X-ray energy, angular distribu-
Robinson, 168, 515, 559 tion of, 618
magnetic spectrograph, 515 X-rays, angular distribution of,
Roentgen, 2, 44 7 463
discovery, 44 7 Scattering coefficient, atomic, 501
Romer, 48 for X-rays, 331
Ross, 525 mass, 501
"Rotation" bands, 320 X-rays, 462, 617
of the sun, 21 multiple, 339
spectrum, 409 of alpha particles, excess, 339
"-vibration" bands, 320 of alpha particles in passing
spectrum, 410 through matter, 337
Rowland, 292 of a-particles, laws of, 342
Rowland, Prof. H. A., 123 of X-rays, 517
Royal Academy of Sciences, 32 electrons per a tom, from, 463
Royal Institution, 53, 60 failure of classical theory, 521
Royal Society, 32 Thomson's theory, 460
Rubens, 205, 260, 261, 262 single, 341
Rubinowicz, 382 X-rays, early observations, 455
Rudolph the Second, 27 Scheiner, 33
Rudolphine tables, 27 Schematic arrangement for study-
Rumford, Count, 2, 4, 53 ing secondary radiation, X-
Rumford's experiment, 53 rays, 470
INDEX 593
Schiaperelli, 6 Sodium, sharp and diffuse series of,
Schroedinger, 272 300
Schuster, 293, 458 vapor, excitation of, 386
law, Rydberg-, 363 resonance of, 389
Schwers, 277 Solar constant, 182
Screening effect, 436 radiation, energy density of, 192
• Second law, Kepler's, 28 spectrum, dark lines, 59
Secondary radiation, X-rays, charac- Sommerfeld, 361, 375, 446, 473
teristic, 467 Sound, early measurements of
schematic arrangement for velocity of, 33
studying, 470 Source of photoelectric energy, 168
variation of quality of, 471 Spark spectra, 318
wave trains, destructive inter- Specific heat, Debye's equation
ference of, 475 for, 271
reinforcement of, 475 first measurements of, 4 7
Seemann, 482 some early measurements of,
"slit," 483 249
Selection principle, 380, 382 heats, Einstein's theory of, 284
rules for inner quantum numbers, of solids, classical theory of,
406 262
Selenium, change in resistance in, quantum theory of, 248-288
136 Spectograph, {3-ray, 559
Self-induction, Henry's discovery of, Spectra, band, 320, 407
71 electronic band, 411
Separation of the Isotopes, 646 gamma-ray, 658
Series relations in line spectra, 289- Spectral distribution of black-body
329 radiation, 203
in spectra, early search for, 292 lines, Bohr's formula for fre-
Sellmeier, 118 quency, 354
Sharp, 362 breadth of, 327
and the diffuse series of sodium, Doppler effect on, 327
the, 300 effect of external physical condi...
series, 299 tions on, 321
Siegbahn, 483 fine structure of, 372
Simon, 525 multiplicities in, 404
Single scattering, 341 origin of, 330-411
Skinner, 560, 561, 564 pressure, effect on, 328
Hlack, 532 Rydberg's formula for, 302
Smith, 445 temperature, effect on, 326
Hmith, Main, 446 series, abbreviated notation, 305,
Smithsonian Institution, 71 311
Snell, 33 and atomic properties, 314
Society, Lin re an, 32 formulm for, 305
Royal, 32 notation, (Fowler's), 310
Socrates, 6 relations between, 304
Sodium atom, energy-level diagram, Rydberg's formula for, 297
401 terms, 308
the stationary states, 400 significance of, 312
static atom model for, 420 types of, 304
594 INTRODUCTION TO MODERN PHYSICS
Spectral terms, Russell-Saunders Straube!, 456, 468
notation for, 405 Structure of nuclei of the radio-
(X-ray) energy-distribution curve, active elements, 663
489 Successes of Rutherford-Bohr atom
Spectrometer, X-ray, 480 model, 366
Spectroradiometer, the, (Coblentz), Sule, 453
203 Sun, rotation of, 21
Spectrum, Newton's discovery of, 36 Sunspots, 21, 23
of helium, Bohn's formula for, 368 Suppression of degrees of freedom,
of hydrogen, Brackett Series, 368 286
Lyman Series, 367 rotation, 286
Paschen Series, 367 vibration, 287
of sodium, doublets in the, 298 System, Copernican, 11, 14, 28, 48
rotation, 409 Pythagorean, 15
vibration, 410 Systems of units, (electric), ratio of,
Standards of wave lengths, 292 81
Stark effect, 326
T
Static atom, 413
fundamental postulates con- Tables, Rudolphine, 27
cerning, 417 Telescope, invention of, 20
model, 412, 416 Keplerian, 31
for carbon, 421 reflecting, 37
for fluorine, 420 refracting, 37, 4 7
for heavier elem en ts, 425 Temperature, effect on spectral
for lithium, 418 lines, 326
for nitrogen, 421 measurement, 202
for oxygen, 421 Terminology for ionized atoms, 413
for sodium, 420 Terms, "hydrogen like," 318
Stationary states of the atom, 383 significance of spectral-series, 312
of the hydrogen atom, 397 spectral series, 308
sodium atom, 400 Test of Debye's Equation, 271
Sta tis ti cal distribution of energy Thales of Miletus, 6
among degrees of freedom, 229 Theorem, Poynting's, 106
Stebbins, 158 Theoretical formula for absorption
Stefan, 194 of X-rays, 513
''Stefan-Boltzmann constant,'' 200 Theory of electricity, one-fluid, 50
best numerical value of, 202 two-fluid, 50, 52
law, 194, 200, 203, 243, 244 heat, caloric, 53, 54
experimental verification of the, materialistic, 47
201 mechanical, 54
"Stefan's constant," 200 of light, classical, conflict with
law, 203 photoelectric effect, 173
Stern, 283 corpuscular, 38
Stewart, 155 Thermal Radiation, 177
Stoletow, 138, 167 Thermometric scale, 4 7
Stokes, 458 Celsius, 47
Stokes' law, 469 Fahrenheit, 4 7
Stoner, 446 Reaumur, 47
Stoney, 293 Thiessen's formula, 206
INDEX 595
"Third law," Kepler's, 30 Variation of atomic heats of solids
Thompson, Benjamin, 53 with temperature, 260
Thompson, Elihu, 452 of mass absorption coefficients,
Thompson, S. P., 450 with wave length, 500
Thomson atom, 331 with atomic number, absorption
arrangements of electrons in, of X-rays, 502
336 Velocity distribution curves for
conditions for equilibrium in, photoelectrons, 168
334 maximum, of photoelectrons, 160,
Thomson, Sir J. J., 2, 140, 150, 175, 161
331, 338, 372, 458, 517, 536 of cathode rays, measurement of,
discovery of the electron by, 151
148 of electromagnetic wave, 108, 109
Thomson's positive-ray spectograph, of light, 25, 109
537 Romer's measurement, 48
Thorium D, ')'-ray spectrum of, of sound, early measurements of,
663 33
Torricelli, 25, 33 Venus, crescent phases of, 21
Total emissive power, 179, 181 Vibrating charge, effect of magnetic
emissivity, 180 field on, 146
quantum number, 366, 433 radiation from, 133
Trains, wave, 135 Villari, 452
Transition series, 429 Vision, Greek theories of, 8
Transmissivi ty, 182 Viviani, 25
Transverse light waves, Young's Volta, 61
suggestion of, 58
Triplet series relations between, 310 w
Trowbridge, 111 Wave length of light, first deter-
Tungsten, emissivity of, 187 minations of, 56
"Two New Sciences, Dialogues of," of scattered radiation, 619
24 shift in Compton effect, 525
Tycho, observations of, 16, 18, 26, 48 -temperature displacement law,
Tyndall, 194 206
effect, 469 units and methods of measure-
_Types of radioactive radiations, 660 ment, 290
of ')'-rays, 660
u lengths of hydrogen lines, table,
Unit magnetic pole, the, 80 296
Units and methods of measurement, of some K lines, 603
wave length, 290 of X-ray lines, Moseley's valneR
Uraniborg, observatory of, 26 for, 486
Uranium, X-ray energy-level dia- standards of, 292
gram for, 509 X-ray, early estimates of, 473
Uranus, discovery of, 48 first measurements of, by
Urban VIII, 22 Bragg, 484
motion, differential equations of,
v 98
Vallombrosa, 19 general equations of, 97
Van Vleck, 281, 284 number, 291
596 INTRODUCTION TO MODERN PHYSICS
Wave number convergent, 302 X-ray spectrum, 484
differences between fine-struc- continuous, 178
ture components of Ha, 381 tube, Coolidge, 451
trains, 135 early form, 451
Way, milky, 7 wave lengths, early estimates of,
Wagner, 497 472
Walter, 473 X-rays, 447-636
Warburton, F. W., 502 absorption of, 464, 611
Water, decomposition of, 63 classical formula for index of
index of refraction of, 118 refraction, 530
Weber, 250 diffraction of, early observations,
White, Andrew D., 23 467
Wien, 214 discovery of characteristic, 456
Wien, W., 473 early observations, scattering, 466
'' Wien' s displacement constant,'' early theories, 467
212 "ether pulse" theory of, 468
law,'' 205, 212, 242 fluorescent, 470, 511
formula, 206, 228 index of refraction, 467
law, 241, 289 ionization measurements, 462
Wiechert, 458 polarization of, 466
Wilson, H. A., 152 production of, 468
Wilson, C. T. R., 525 photoelectric effect of, 614
Winchester, 167 refraction of, 628
Wind, 473 early observations, 467
Winkelmann, 456, 468 scattering, coefficient for, 331, 462
Wood, R. W., 389 scattering of, 517
W oestyn, 249 Thomson's theory, 460
W ollaston, 64 some early experiments and
theories, 460
x sources of, 460
X-radiation, Barkla's discovery of, X-unit, 290
fluorescent, 4 72
y
X-ray data, determination of
Planck's constant from, 494 Young, 74, 78, 175, 291
determined from photoelectric Young, Thomas, 4, 54, 56
effect, absorption limits, 617
early estimates of wave lengths, z
473 Zeleny, 167
emission lines, relation to absorp- Zeeman, 143, 148
tion limits, 505 effect, 121, 143, 152, 154, 322
energy-distribution curve, spect- abnormal, 322
ral, 489 elementary explanation of, 144
energy-level diagrams, 608 frequency change in, 145
diagram for uranium, 509 normal, 322
first measurements of wave shift, 322
lengths, by Bragg, 484 pattern, 322
photography, 452 patterns of some representative
spectrometer, 480 lines, 324
''double,'' 532 Zollner, 190