The Tiger and The Shark Empirical Roots of Wave-Particle Dualism
The Tiger and The Shark Empirical Roots of Wave-Particle Dualism
The Tiger and The Shark Empirical Roots of Wave-Particle Dualism
Bruce R. Wheaton
with a foreword by Thomas S. Kuhn
CAMBRIDGE
UNIVERSITY PRESS
CAMBRIDGE UNIVERSITY PRESS
Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, Sao Paulo
Published in the United States of America by Cambridge University Press, New York
www.cambridge.org
Information on this title: www.cambridge.org/9780521250986
Cambridge University Press 1983
A catalogue record for this publication is available from the British Library
1 Introduction i
Part I The introduction of temporal discontinuity, 1896-1905
2 The electromagnetic impulse hypothesis of x-rays 15
3 The analogy between y-rays and x-rays 49
Part II Ionization and the recognition of paradox, 1906-1910
4 Secondary rays: British attempts to retain mechanism 71
5 The appeal in Germany to the quantum theory 104
Part III Seeking an electrodynamic solution, 1907-1912
6 Localized energy in spreading impulses 135
7 Problems with visible light 168
Part IV Interference of x-rays and the corroboration of paradox,
1912-1922
8 Origins of x-ray spectroscopy 199
9 Quantum transformation experiments 233
Part V The conceptual origins of wave-particle dualism,
1921-1925
10 Synthesis of matter and light 263
Epilogue: The tiger and the shark 302
Bibliography 309
Index 347
FOREWORD
THOMAS S. KUHN
PREFACE
1
Wheaton, Photoelectric effect (1971). Abbreviations used in the footnotes and
the bibliography are identified in the notes on sources.
2
For example, see Eisberg, Fundamentals (1961), 76-81; Jammer, Conceptual
development (1966), 35.
xvi Preface
After an invaluable exposure to the modern descendant of the
European culture within which these issues developed, I was again
fortunate to have the opportunity to work on the more general
historical problem of wave-particle dualism with Thomas Kuhn
as guide. When I extended my original concerns to include high-
frequency radiations, I found that the experimental evidence
amassed by 1911 concerning x-rays played a far more significant
role in preparing physicists to accept dualistic theory than did
evidence regarding ordinary light. My doctoral dissertation
formed the second stage in the development of this book.3 In it I
clarify the extent to which the then standard impulse interpreta-
tion of x-rays led to an implicit tension of a kind destined to bring
reformulation of classical concepts regarding the distribution of
energy in all forms of radiation in the 1920s.
But the story does not end there, and in this book I bring the
discussion to its historical conclusion. The results of attempts to
force consistency on electromechanical interpretations of radia-
tion rebounded to affect also the theory of matter. While reconsid-
ering Einstein's lightquantum hypothesis in 1921, Louis de Broglie
tried to find a theory of light that would combine the macroscopi-
cally incompatible representations of wave and particle. He based
this work on the solid foundation provided by his elder brother's
experimental corroboration of Einstein's photoelectric law for
x-rays. The result in 1923 was a hypothetical synthesis of matter
and light: Each was to be considered to possess both particle and
wave attributes that make their presence felt to a greater or lesser
extent according to experimental conditions. The successful elabo-
ration of this remarkable hypothesis signaled the end of strictly
deterministic representations of both matter and of light.
The reader may wonder why the name of Albert Einstein does
not figure more heavily in this account of evolving understanding
that neither a wave nor a particle characterization of radiation is
alone sufficient. After all, Einstein introduced the lightquantum
hypothesis in 1905 and early on recognized that a synthetic theory
was needed. The reasons are straightforward. Einstein's revolu-
tionary hypothesis had almost no followers and very little influ-
ence before 1921, and the growth of its acceptance by others marks
the limit of our direct concerns. His lightquantum was not initially
intended to apply either to x-rays or to y-rays, but rather only to
3
Wheaton, Nature ofx- and gamma rays (1978).
Preface xvii
visible and ultraviolet light. Before 1911 it was not clear to most
physicists that x-rays and visible light are common species of
radiation. Consequently the experimental evidence that had accu-
mulated by that time that x-rays and y-rays do transfer energy
in individual units did not justify Einstein's lightquantum in the
eyes of most physicists. The difficulties of any corpuscular theory
in explaining interference properties of radiation were too formi-
dable. And for almost a decade following Niels Bohr's quantum
theory of the atom in 1913, most mathematical physicists showed
little interest in the paradoxes that complicated a consistent elec-
tromechanical interpretation of free radiation.
Thus, our story developed quite independently of the imagina-
tive and prescient statistical treatment of light of which Einstein
was the chief architect. Einstein's own developing realization of
the need for duality in radiation theory has been discussed histori-
cally by Martin Klein.4 Here we analyze other physicists' efforts to
wrestle with closely related issues, discuss their rejection of and
eventual accommodation to dualistic ideas in physical theory, and
trace the way in which this development came to alter ideas not
simply about radiation and matter but about humans' ability to
construct consistent models of physical phenomena based on their
experiences in the macroscopic world of human senses.
The present narrative evolved out of an extensive search in the
physics literature of the period, all significant results of which
appear in the notes. Documents in several manuscript collections,
some in archives and others privately held, were invaluable for
reconstructing contemporary insights and opinions, for support-
ing evidence, and for opening new areas for research. In particular,
I benefited from examination of the Archive for History of Quan-
tum Physics,5 the Bohr archive, the Cherwell papers, the Einstein
archive, Paul Langevin's papers, the Lorentz papers, holdings of
the Nobel archives, the archives of the Paris Academy of
Sciences, the Rutherford papers, the Schwarzschild papers, and
the archives of the Solvay Institute. Full identifications and loca-
tions of these and other collections are given in the notes on
sources. I am grateful to these and other repositories for the kind
access granted me, and to the holders of literary rights for permis-
sion to quote documents. I was greatly assisted in my research by
4
Klein, Natural philosopher, 3 (1964), 3-49.
5
Described in Kuhn, Heilbron, Forman, and Allen, Sources (1967).
xviii Preface
concurrent work on the Inventory of sources for history of twen-
tieth-century physics, being led to several important and formerly
untapped collections.6 The immense quantity of data the ISHTCP
has compiled and their as yet incomplete description encourage
further mining of their riches.
An undertaking of this magnitude is possible only with the help
of colleagues and of prior work in the field.71 am pleased here to
acknowledge debts of both sorts. I am grateful for the comments of
several anonymous reviewers of the typescript before publication.
I have already mentioned the importance of Martin Klein's work
on the early work of Einstein; other studies by him are mentioned
in the Notes. The quantum interpretation of the Compton effect in
1922 was an event of great significance; Roger Stuewer's detailed
review of the evolution of Compton's approach to it made it
unnecessary to recount that episode in depth here.8 Although
Stuewer's and my interpretations differ on the importance of the
electromagnetic impulse hypothesis of x-rays and about the renais-
sance of interest in the lightquantum, his study is a significant
addition to the meager literature on the development of experi-
mental physics in this century.9
I have been fortunate, professionally and personally, in my own
introduction to history of science to have had as guides both
Thomas Kuhn and John Heilbron. I have benefited greatly from
discussions with both on many aspects of physical theory. Profes-
sor Kuhn's reassessment of the conceptual origins of quantum
theory has influenced this study in many ways.10 Only one who has
had the benefit of his insightful criticism and sympathetic ear can
understand how much of a debt I owe him. His conceptualization
of scientific advance encourages historical analysis within which
the humanistic aspects of science are no longer submerged in
mythic objectivity. We are all in his debt for that insight.
6
This inventory identifies the location, author, recipient, and approximate date
of half a million letters from or to physicists active between 1896 and 1952.
It will soon be published in microfiche/book form. Contact Springer Verlag,
New York for information on ordering.
7
For background on the material basis of national style in physics, see Forman,
Heilbron, and Weart, HSPS, 5 (1975), 1-185.
8
Stuewer, Compton effect (1975).
9
For the neglect of experimental physics, see Heilbron and Wheaton, Literature
(1981).
10
Kuhn, Black-body theory (1978).
Preface xix
I would be pleased if this study of radiation theory were taken as
complementary to John Heilbron's analysis of concurrent devel-
opments in atomic theory.11 Matter and light are the two perceiv-
able manifestations of nature. Their interaction provided the evi-
dence that showed that electromechanical representations of
nature would fail. But although the classical-quantum compro-
mise inherent in the Bohr theory of the atom acted to delay
recognition of failure for the electromechanical atom until the
early 1920s, the paradoxical behavior of radiation was evident to
some as early as 1908. The problems that led to the rejection of
deterministic physical theory presented themselves early, and
most clearly, in attempts to bring consistency to the theory of
radiation.
BRUCE R. WHEATON
11
Heilbron, Atomic structure (1964); contribution in Twentieth century physics
(1977X40-108.
NOTES ON SOURCES
PUBLISHED SOURCES
All works cited in short form in the footnotes will be found in the
bibliography under the author's name, but the footnote references
are intended to be sufficient to lay hands on the work itself. In
general, and following Library of Congress practice, names of
journals are listed under the name of the issuing organization.
Because many are cited frequently I have used the following
abbreviations:
AHES Archive for History of Exact Sciences
AJP American Journal of Physics
AJS American Journal of Science
Amsterdam Koninklijke Akademie van Wetenschappen te Am-
sterdam. Wis- en Natuurkundige Afdeeling
AP Annalen der Physik
ASPN Geneve Bibliotheque universelle. Archives des sciences phy-
sique et naturelle, Geneve
AusAAS Australasian Association for the Advancement of
Science
BAAS British Association for the Advancement of Science
Berlin Mb, Sb Konigliche preussische Akademie der Wissenschaf-
ten, Berlin. Physikalisch-mathematische Klasse,
Monatsberichte, Sitzungsberichte
BJHS British Journal for the History of Science
Bologna Reale Accademia della Scienze delFIstituto di Bologna
CR Academie des Sciences, Paris. Comptes rendus hebdo-
madaires des seances
DSB Dictionary of Scientific Biography, 14 vols., C. C. Gil-
lispie, ed. (New York, 1970-6)
XX11 Notes on sources
GDNA Gesellschaft Deutscher Naturforscher und Arzte
Gottingen Konigliche Gesellschaft der Wissenschaften zu Got-
tingen
Halle Naturforschende Gesellschaft zu Halle
Heidelberg Sb Heidelberger Akademie der Wissenschaften, Sitz-
ungsberichte (listed by Abhandlung number)
HSPS Historical Studies in the Physical Sciences
JFI Franklin Institute, Journal
JHI Journal of the History of Ideas
JP Journal de physique et le radium
JRE Jahrbuch der Radioaktivitdt und Elektronik
Konigsberg Physikalisch- okonomisch Gesellschaft zu Konigsberg
Lincei Reale Accademia dei Lincei, Rome
Manchester MP Manchester Literary and Philosophical Society,
Memoires and Proceedings
Munchen Sb Konigliche Bayerische Akademie der Wissenschaften
zu Munchen, Sitzungsberichte
NAS National Academy of Sciences, Washington, D.C.
NNGC Nederlandsch natuur- en geneeskundig congres
NRC National Research Council, Washington, D.C.
PCPS Cambridge Philosophical Society, Proceedings
PM Philosophical Magazine
PPSL Physical Society of London, Proceedings
PR Physical Review
PRI Royal Institution, Proceedings
PRS RSL, Proceedings
PTRS RSL, Philosophical Transactions
PZ Physikalische Zeitschrift
RGS Revue general des sciences pures et appliquees
RDS Sci trans Royal Dublin Society, Scientific Transactions
RSL Royal Society of London
SFP PVRC Societe Frangaise de Physique, Proces-verbaux et re-
sume des communications
SHPS Studies in History and Philosophy of Science
TCPS Cambridge Philosophical Society, Transactions
TPRRSSA Royal Society of South Australia, Transactions and
Proceedings and Report
VDNA GDNA, Verhandlungen
VDpG Deutsche physikalische Gesellschaft, Verhandlungen
(usually bound with Berichte)
VI Victoria Institute
WAS Washington Academy of Science
Wien Sb Kaiserlich-Konigliche Akademie der Wissenschaften,
Notes on sources xxiii
Wien. Mathematisch-naturwissenschaftlichen Klasse,
Sitzungsberichte
Wiirzburg Sb Physikalisch-medizinische Gesellschaft zu Wiirzburg,
Sitzungsberichte
ZP Zeitschrift fur Physik
Zurich Physikalische Gesellschaft zu Zurich
UNPUBLISHED SOURCES
A work of this kind depends on both published and unpublished
records of contemporary opinion and results. We are fortunate
that the subject at hand developed after the typewriter was in-
vented and before the telephone was widely used. Carbon copies of
typed letters double the chance that they will be retained, and the
lack of telephone use encourages creation of a written record.
In my research I used many sources of unpublished documents,
and I am indebted to several people for permission to quote from
them: Victor de Pange for permission to quote Maurice de Broglie,
the American Friends of the Hebrew University to quote Albert
Einstein, Mme. Luce Langevin to quote Paul Langevin, Dr. Bent
Nagel to quote transactions of the Nobel Prize committee, and
Ruth Braunizer for permission to quote Erwin Schrodinger. I am
grateful to Professor Louis de Broglie, who graciously set aside
several hours for a discussion with me about this research. Of
course I, not he, am answerable for the interpretation given here.
I am no less grateful to many other libraries and archives in
Europe and America for making their collections of correspon-
dence available to me. Consulting them was essential for back-
ground information, for leads to new sources, and to ensure that
the general picture I painted was true to life. Many of the letters
cited here are quoted in part or whole in other published works.
References appear in the microfiche index An inventory of pub-
lished letters to and from physicists, 1900-1950, listed in the
bibliography under Wheaton and Heilbron (1982). Many letter
collections related to modern physics are now described in detail in
the Inventory of sources for history of twentieth-century physics (a
database soon to be published in microfiche/book form by Sprin-
ger Verlag) including all that are noted below, save the Nobel
archives.
xxiv Notes on sources
AHQP. Archive for history of quantum physics. Berkeley, Copenhagen,
Minneapolis, New York, Philadelphia, and Rome. Microfilm
references follow the form "JC, y" where x is reel number, y is
section number. For a description, see Kuhn, Heilbron, Forman,
and Allen, Sources (1967). Sources marked with a *are available
on microfilm at several of these locations
Bohr (Niels) archive.* Niels Bohr Institutet, Copenhagen, Denmark
L. de Broglie dossier. Archives of the Academie des Sciences, Paris
Cherwell (F. A. Lindemann) papers. Nuffield College Library, Oxford,
England
Ehrenfest (Paul) archive.* Museum Boerhaave, Leiden, Netherlands
Einstein (Albert) papers. Institute for Advanced Study, Princeton, New
Jersey
Langevin (Paul) papers. Formerly privately held in Paris, now at "Fonds
des ressources historique Langevin," Ecole superieur de physique
et chemie de la ville de Paris, France.
Lorentz (Hendrik) papers.* Algemeen Rijksarchief, the Hague, Nether-
lands
Nobel archives. Nobelstifting, K. Vetenskapsakademiens Nobelkommit-
teer, Stockholm, Sweden
Planck (Max) papers.* Staatsbibliothek Preussischer Kulturbesitz, Berlin
Richardson (Owen) papers.* Humanities Research Center, University of
Texas at Austin
Rutherford (Ernest) papers. Cambridge University Library, Cambridge,
England
Schrodinger (Erwin) papers.* Privately held in Alpbach, Austria
Schwarzschild (Karl) papers. Niedersachsische Staats- und Universi-
tatsbibliothek, Gottingen
Solvay (Institute) archives. Archives de 1'Universite Libre, Brussels, Bel-
gium
Sommerfeld (Arnold) papers. Sondersammlung, Deutsches Museum,
Munich
1
Introduction
17
Jaumann, Wien Sb, 106:11a (1897), 533-50.
18
Lodge, Electrician, 36 (1896), 438-40.
19
[W. Thomson], PRS, 59 (1896), 270-3.
20
[W. Thomson], Baltimore lectures (1904), 141.
21
Reported in Nature, 53 (1896), 528. Stokes, Manchester MP, 41 (1897), no. 15,
p. 2. Gifford, Nature, 54 (1896), 172.
22
J. J. Thomson, Nature, 53 (1896), 391-2.
20 The introduction of temporal discontinuity, 1896-1905
tion, diffraction, or reflection.23 The principal motive of the longi-
tudinal wave thus removed, Ockham's razor, which slices away
multiple hypotheses where one will suffice, more than compen-
sated for the lack of definite evidence about polarization. By the
fall of 1896, transverse waves had become "the surviving hypoth-
esis."24
Major support for the transverse wave came from Henri Bec-
querel's discovery of radioactivity in 1896, for Becquerel had
claimed to find polarization in his uranium rays.25 At the end of
1896, most physicists agreed with Thomson when he described the
Becquerel rays as resembling "the Rontgen rays more closely than
any kind of light known hitherto."26 The Becquerel rays seemed to
be intermediate between light and x-rays. This view "greatly
strengthened the position . . . that the x-rays were of the nature
of ultra-ultraviolet light."27 Because the Becquerel rays reportedly
showed reflection and refraction, the possibility of these properties
was more easily projected onto x-rays. Lodge concluded, late in
1896, that "it has become almost certain that the x-rays are simply
an extraordinary extension of the spectrum," that is, periodic
transverse waves of high frequency.28
propagation
ment is everywhere zero (i.e., to the left of, the right of, or fully
within the impulse) or its effect arises from the integral across the
full impulse. But according to Stokes' assumption, this integral
vanishes; viewed from the side, the net contribution at P will then
be zero. (2) If r < PQ, no part of the pulse can affect P. (3) The only
contribution at P comes from the lenticular shaded region for
PS > r > PQ. The total transverse displacement at P is just the
sum of the contributions from this region, attenuated in the ratio
f Of + r). Stokes showed that this region was quite small. For a
pulse of width icr6 in., and for reasonable values of r and f of 4 in.,
the radius of the active zone is only 0.002 in. The complete de-
crease in pulse intensity to the side of the shadow occurs in only
1/250 in., below the threshold of contemporary photographic
resolution.
J. J. THOMPSON ON PULSES
Joseph John Thomson, Cavendish Professor of Physics at Cam-
bridge University, carried out the detailed calculation of Stokes'
hypothesis. Thomson was an expert on Maxwell's theory. Further-
more, he was actively collecting proofs that cathode rays are
The electromagnetic impulse hypothesis of x-rays 25
subatomic particles, having recently proclaimed them to be such.
The impulse hypothesis of x-rays, a corollary to the particulate
interpretation of cathode rays, was bound to attract his interest.
Thomson had considered the longitudinal wave hypothesis in an
early study.42 But he quickly recognized the merits of Stokes' idea
and, in December 1897, put it to a test.43
An electron is the source of isotropically diverging electric lines
of force. If the electron moves at a velocity that is small compared
to that of light, this set of lines remains symmetric as it moves
along with and centered on the electron. The accompanying mag-
netic field is arranged in circles lying in successive planes perpen-
dicular to the direction of electron motion. If the electron stops
abruptly, it takes some small time for that fact to be broadcast to
the surrounding field. The propagation of the dislocation outward
along the field lines constitutes the electromagnetic impulse. Fig-
ure 2.3 illustrates the growth of the spherical impulse at succes-
sive equal intervals of time after the electron stops.44 The center
marked "x" indicates the position the electron would have occu-
pied had it not stopped. The field outside the impulse remains
centered on "x," whereas the field inside diverges from the station-
ary electron.
Using Figure 2.4, one may easily calculate the ratio of tangential
to radial electric force. At time t after the impact, it will be equal to
the ratio
NP' vt sin 6
{2A)
NN7 J {2A)
AE ?A
(2.2)
Anrd c
where c = r/t is the velocity of wave transmission along the field
42
J. J. Thomson, PCPS, 9 (1896), 49-61.
43
J. J. Thomson, PM, 45 (1898), 172-83. Thomson mentioned in his Princeton
lectures in October 1896 that electromagnetic impulses should arise from
cathode ray impacts. Discharge of electricity (1898), 192-3.
44
Heaviside, PM, 27 (1889), 324-39; PM, 44 (1897), 503-12.
t=3
AH = v sin 6 (2.3)
ro
The point is that the transverse effect of the impulse drops only as
i/r, whereas the radial force drops as i/r2. Thus, the electrody-
namic effect of the impulse decreases less rapidly than does the
normal radial electric force, and will predominate at great dis-
tances from the electron.
In 1903, Thomson calculated the energy residing in one impulse.
In a given direction 6 from the direction of electron motion, the
energy density is
sin
J A 2 = ___ (2.4)
4TI ^" 4^ rS
AH is the magnitude of the magnetic force, and ft is the magnetic
permeability of the medium (in a vacuum, // = 1). Integrating this
expression over the volume of the expanding spherical shell gives
the total energy carried by the impulse:
2JU
F = (2.5)
^total 3S
45
In 1897 Thomson published only hisfindingson the magnetic force; those on
the electric force followed in 1903.
28 The introduction of temporal discontinuity, 1896-1905
time
-5/2c +5/2c
46
J. J. Thomson, Conduction of electricity (1903), 540.
sin 2 0 ,
And1 J j r2
Integrating over a hemisphere:
"* 2nr sin3 6
And1 Jof
*J 0
r1
f sisin3 Ode
rn
Jo
Since
-7T/2 2
s i n 3 M ? =
=
3' 3 J
47
Stokes, Papers, 5 (1905), 265-6.
The electromagnetic impulse hypothesis of x-rays 29
By 1899, Kelvin too had accepted the impulse hypothesis. His
original concern had been that an extremely high-frequency longi-
tudinal wave was necessary to explain the low absorption of x-rays.
He analyzed the analogous case of oscillations in a mechanical
system to chart the relative amplitudes of the transverse and
longitudinal waves when the wavelength was about equal to the
molecular spacing in matter.48 He concluded that the amplitude of
the condensational wave might drop to negligible values for fre-
quencies above io12 per second.49 No explicit denial of the longitu-
dinal wave can be found in Kelvin's writings, but it is difficult not
to interpret his several attempts in 1899 to calculate the theoretical
properties of single, detached impulses as tacit approval of the
impulse hypothesis of x-rays.50
57
Wiechert, Konigsberg Schriften, 37 (1896), 1-48; AP, 59 (1896), 283-323,
especially 321-2.
58
"Irregular light blows." There is no question that Wiechert used jdhe in the
temporal sense, and he specified a sense for Stoss that made it clear that he
meant individual impulses.
59
Haga and Wind, Amsterdam Verslag, 7(1899), 387-8, 500-7.
60
Wind, Amsterdam Verslag, 7 (1899), 12-19. Among other results called into
question were those of Fomm, Munchen Sb, 26 (1896), 283-6.
Figure 2.6. Haga and Wind's photograph of x-ray diffraction by a slit.
Left: Exposure 5, collimation 25 /an, distance from slit to photoplate
1 cm, exposure 29 hr. No discernible diffraction. Right: Exposure 2,
collimation 14 /zm, distance from slit to photoplate 75 cm, exposure
100 hr. Broadening of the image appears at approximately 8 to 9 jum.
The dark region at the lower right is the image of a reference hole drilled
in one of the 0.5-mm-thick platinum plates that form the slit. It is
elongated in the right-hand exposure because of angular extension of
the x-ray source. [Haga and Wind, Amsterdam Verslag, 7(1899), facing
502.]
The electromagnetic impulse hypothesis of x-rays 33
possible sources for the broadening and concluded that it must
have arisen from diffraction of x-rays.
f
J c
f(t)e~iwt dt (2.6)
66
Sommerfeld used the standard symbol for wavelength A to characterize the
impulse width, although he stated quite explicitly that no periodicity is to be
expected in the impulses. I will use the neutral symbol S throughout this study
to indicate spatial impulse width.
67
Sommerfeld's initial statement of the failure of classical diffraction theory
begins with two statements of the problem, one physikalisch ausgedruckt and
the other mathematisch ausgedruckt.
68
Sommerfeld, PZ, 1 (1899), 105-11; PZ, 2 (1900), 55-60.
36 The introduction of temporal discontinuity, 1896-1905
screen
impulse
in Figure 2. io. The smaller the pulse width, the less the discernible
diffraction.
To relate the calculation to the results from Groningen, Som-
merfeld had to introduce approximations. He needed to convert
electromagnetic amplitude to a measure of photographic blacken-
ing. Then again, Wind and Haga's data came from slit diffraction;
38 The introduction of temporal discontinuity, 1896-1905
the introduction of the second edge, a problem Sommerfeld had
not solved completely in his 1896 study, introduced error. As an
approximation, he took the total broadening of the image as due
half to each side. The result was a calculated x-ray pulse width of
33 A, compared to the Dutch estimate of a i-A periodic wave-
length.
In the second phase of his study, Sommerfeld the physicist
retracted what Sommerfeld the mathematician had done. After
some discussion with Wind on sources of error in the photograph,
he revised his calculation.69 This time he reworked the problem
explicitly by taking both edges of the slit into account. "Refraction
from a slit," he said, "often turns out very different from refraction
at an edge."70 He now found that the "dark core" of the wedge-slit
image, his interpretive sketch of which is shown in Figure 2.11,
would cease at a slit width of \f$Sr/io. Wind's best estimate was
that this occurred at 9 //m; Sommerfeld derived a pulse with a
spatial width d= 1.3 A and a temporal width of 4 x io~19 sec. He
thought this a "much more trustworthy" estimate. To his friend,
the astrophysicist Karl Schwarzschild, an expert on photochemical
response, Sommerfeld wrote, "with such a small [pulse width] it is
easy to understand why x-rays are neither absorbed nor refracted.
Molecules simply cannot follow such short impacts."71
Sommerfeld went on to propose that a continuum of radiation
waveforms exists, one that extends from the purely periodic sine
waves of monochromatic light to the temporally localized aperi-
odic impulses of x-rays. There would generally be a mixture of
pulse widths in an x-ray beam. He had proven to his satisfaction
that x-ray impulses are aperiodic; the lack of characteristic diffrac-
tion fringes was evidence for their intrinsic temporal discontin-
uity.72 Thus, Sommerfeld had a physical if not a formal reason to
reject the consideration of an impulse as if broken into its Fourier
spectrum, as others were suggesting. Wind, for example, was
surprised to have his photograph successfully analyzed in terms of
aperiodic pulses; recall that he had first claimed the diffraction
69
Details of Wind's criticism exist in his several letters to Sommerfeld between
April and August 1900. Sommerfeld papers; AHQP 34, 14.
70
Sommerfeld, PZ, 2 (1900), 55.
71
Sommerfeld to Karl Schwarzschild, 24 September 1900. Schwarzschild papers.
72
Discussion following Wind's presentation to the seventy-second Naturfor-
scherversammlung, PZ, 2(1900), 297-8.
The electromagnetic impulse hypothesis of x-rays 39
9 junn
Oscillations at A
only in plane normal
to beam 1.
Oscillations at B
only perpendicular
to beams 1 and 2
As seen at B
-S
Fl
3
P. and [M] Curie, CR, 727(1898), 175-8.
4
Rutherford, PM, 47(1899), 109-63.
5
The "diffusely reflected" x-rays were largely secondary electrons, but this was
not known at the time. See Chapter 4.
6
P. and [M.] Curie, CR, 729(1898), 714-16.
The analogy between y-rays and x-rays 51
TRANSFORMING WAVES INTO PARTICLES
It came, therefore, as a surprise to many when, in 1899, Becquerel
rays were first deflected by a magnet. They bent in precisely the
manner expected of negatively charged particles.7 Soon afterward,
it was demonstrated that the ray trajectories also bend in an
electrostatic potential.8 Part of the credit for this recognition must
go to the new and more intense sources of radioactivity, polonium
and radium, provided by the Curies in 1898. Credit also must go to
the widespread interest, particularly in Germany, in techniques for
the study of cathode rays by magnetic deflection. In any event, the
nature of the Becquerel rays was now in dispute. As one reviewer
put it, "not content with attributing to Becquerel rays all the
properties of x-rays, physicists and chemists have found that these
rays also possess properties hitherto attributed to cathode rays
alone."9
Pierre Curie showed that Rutherford's division of Becquerel
rays into a- and ^-components was also expressed in their mag-
netic deflection. Only the /^-component seemed to bend. The
a-rays traveled on, apparently unaffected.10 A photoplate, exposed
to uranium rays that had passed through a magnetic field, thus
showed two exposed regions. Rutherford's division of Becquerel
rays into components offered the first hope of systematic study of
uniform properties. No longer were the seemingly variable proper-
ties of the rays approachable only through the heroic chemical
processes of the Curies, Friedrich Giesel, and others. Around 1899,
an effluorescence of research interest occurred in this field, which
was increasingly referred to as radioactivity.11 Figure 3.1 shows the
number of papers on related topics for the early years cited in an
authoritative review by Max Ikle.12
7
Three independent demonstrations were given. The earliest was by Friedrich
Giesel at Braunschweig, AP, 69 (1899), 834-7, following strong hints by Julius
Elster and Hans Geitel. See also S. Meyer and von Schweidler, PZ, 1 (1899),
90-1, 113-14; CR, 729(1899), 996-1001.
8
Dorn, Halle Abhandlungen, 22 (1901), 39-43, 47-50.
9
Stewart, PR, 11 (1900), 155.
10
P. Curie, CR, 130(1900), 73-6.
11
Becquerel in Guillaume and Poincare, Congres international, 3 (1900), 47 - 78.
Crookes, PRS, 66 (1900), 409-23.
12
Ikle, JRE, 1 (1904), 413-42 is more complete than Wilson, Nature, 70 (1904),
241-2.
52 The introduction of temporal discontinuity, 1896-1905
250-
200- 1
150- I
100-
/*
50 H
f f \i 1 f < 1 1
1896 97 98 99 1900 01 02 03
30
Trenn, Isis, 67(1976), 61-75.
31
M. Curie, Theses (1903), 48, 80, 83, 138.
32
Becquerel, CR, 7^6(1903), 1517-22.
33
Rutherford, Nature, 68 (1903), 163.
34
Nature, 68 (1903X 610.
35
R. J. Strutt, PRS, 72 (1903), 208-10.
The analogy between y-rays and x-rays 57
Kelative Ionisations.
Relative 1 onidation.
Gas. Relative
density. Rout gen
a rays. /3 rays. 7 rajs- rays.
Figure 3.3. R. J. Strutfs data on relative ionization of gases by a-, /?-, y-,
and x-rays. [PRS, 72 (1903), 209.]
H 07 11* 42
Air 10 10 10
HOS 12 6 9
SO., 22 8 23
Chloroform 43 32 46
Carbon Tetrachloride 53 45 4*9
Methyl Iodide 50 72 135
and therefore interacts not with entire atoms but with their consti-
tuent electrons. Equal volumes of gases contain equal numbers of
molecules, so that the gas density is proportional to the molecular
weight. According to Thomson's influential ideas about atomic
structure, the number of electrons in an atom is proportional to its
atomic weight; thus, it is also proportional to the density of the gas.
A particle of subatomic dimension that passes through the gas will
lose energy in proportion to the number of electrons it encounters.
Therefore, particle absorption should be proportional to the den-
sity of the gas.
McClelland was uneasy, however, about whether spreading
waves could be absorbed by gases in the same proportion. "It
would seem difficult to explain such a law," he said.44 He did not
have in mind the difference one might imagine to exist between the
absorption of periodic waves and impulses by bound electrons.45
Rather, McClelland was echoing reservations that his former
teacher, J. J. Thomson, had recently expressed about the isotropic
44
McClelland, PM, 8 (1904), 77.
45
For a periodic wave, not all electrons have natural frequencies sufficiently
close to resonance to be appreciably excited. But a sharp impulse will excite all
electrons. An organ tone transfers energy only to the closely tuned strings of a
harp; a pistol shot excites them all.
The analogy between y-rays and x-rays 61
character of x-ray pulses; McClelland tried to find justification for
Thomson's nascent ideas about the structure of the aether. If a
spherical impulse or periodic wave spreads over areas that are large
with respect to the size of the atoms, it will no longer, McClelland
asserted, interact with constituent electrons. Rather, it must be
absorbed by entire atoms, or groupings of atoms. Since the num-
ber of molecules per unit volume of a gas is constant for the same
pressure and is independent of the gas density, McClelland
doubted that spherical impulses could be absorbed in proportion
to the gas density. Nor was he alone in his concern. William Henry
Bragg in Australia raised the same issue more forcefully at the
same time. But Thomson's and Bragg's important views will be
considered in the next chapter.
Eve's discovery that hard x-rays follow the same law of absorp-
tion in gases as do particles permanently tied the fortunes of the
y-ray to those of the x-ray from 1904 on. The results appeared only
in a footnote in Rutherford's book Radioactivity, published that
year.46 They brought an end to the cautious balancing of three
alternative models for the y-rays thatfilledthe text itself. But in the
second edition of the book, published the following year, Ruther-
ford's discussion of y-rays is extensively simplified and reorganized
around the newly justified impulse hypothesis. The idea of impulse
y-rays was widely circulated in Rutherford's book, which was
quickly recognized as a working digest of essential research in
radioactivity and was published in several editions over the next
decade.
48
Paschen, AP, 74(1904), 171.
49
Ibid., 389-405.
50
Becquerel in Guillaume and Poincare, Congres international, 3 (1900), 63-9.
Starke, VDpG, 5(1903), 14-22.
51
Paschen, PZ, 5 (1904), 502-4.
52
McClelland, RDS Sci trans, 8 (1904), 99-108. Eve, Nature, 70 (1904), 454.
The analogy between y-rays and x-rays 63
Figure 3.5. Diagram of Paschen's experiment that made him think that
y-rays deposit negative charge. [AP, 14 (1904), 390.]
12
Wiechert, AP, 69(1899), 739-66.
13
Lenard, AP, 8 (1902), 149-98. Despite claims to the contrary, even by histo-
rians, Lenard did not here demonstrate that photoresponse varies with light
frequency. He only claimed that it varies with the type of light used-arc light,
spark light, or the type of electrode metal. The word frequency appears no-
where in his study; wavelength appears only once, and in a different context.
Compare Jammer, Conceptual development (1966), 35; Kleinert and Schon-
beck, Gesnerus, 35 (1978), 318; Hendry, Annals of science, 37 (1980), 64.
British attempts to retain mechanism 75
hypothesis formed the basis of almost all physicists' understanding
of the photoelectric effect.14
The triggering hypothesis owed some of its appeal to its compati-
bility with contemporary hypotheses of atomic disintegration. It
seemed to promise a means of investigating the mysterious decay
of radioactive elements. Thomson adopted triggering as an expla-
nation of the photoelectric effect in 1905.15 Between 1905 and 1907,
his student, J. A. McClelland, devoted a series of studies to the
corroboration of the triggering hypothesis for the emission of
^-particles by y-rays. McClelland's chief weapon in this study was
the observed monotonic increase he found in the total ionizing
power of the secondary rays, including what he thought were
secondary y-rays, with the atomic weight of the scattering material.
He took this as evidence that the primary beam, a mixture of /?-
and y-rays from radium, could induce atomic disintegration in
matter.16 If the electrons were free before interacting with the rays,
McClelland explained, "it is difficult to see how such remarkable
relations should exist between the atomic weight and the intensity
of the secondary radiation."17
It was no great step to expand the capabilities of the triggering
hypothesis to include the excitation of electrons by x-rays. Wil-
helm Wien in Wiirzburg concluded in 1905 that the triggering
mode offered the only realistic explanation.18 Thomson publicly
announced his adoption of the triggering hypothesis for x-ray
ionization in 1906.19 His student, P. D. Innes, found direct sup-
porting evidence the following year; like the velocity of photoelec-
trons, the velocity of the secondary electrons from x-rays appears
to be fully independent of the intensity of the incident radiation.20
For a brief period in 1906, the influence of the triggering hypoth-
esis was at its height. It was widely thought to explain the observed
velocities of electrons ejected from matter by light, by x-rays, and
14
Wheaton, HSPS, 9(1978), 299-323.
15
J. J. Thomson, PM, 10 (1905), 584-90.
16
McClelland, RDS Sci trans, 9 (1905), 1-8, 9-26, 37-50. Righi also found
evidence of a correlation with atomic weight. Lincei Atti, 14 (1905), 556-9.
17
McClelland, RDS Sci trans, 9 (1905), 36.
18
Wien, AP, 18 (1905), 991-1107. See Chapter 5.
19
J. J. Thomson, PCPS, 13 (1906), 322-4; Corpuscular theory (1907), 320-1.
Some of the evidence that convinced Thomson came from Bumstead, PM, 11
(1906), 292-317.
20
Innes, PRS, 79A (1907), 442-62.
76 Ionization and the recognition ofparadox, 1906-1910
by y-rays. Indeed, for a while it seemed as if the study of these
velocities would reveal the distribution of electron energy within
the atom and shed light on the mysterious process of radioactive
disintegration. These were two of the most pressing fields of physi-
cal research in the first decade of the century.
Potential
34
Ibid., 77. The passage actually reads "unable to act as a whole molecule," but
from the next sentence it is clear that this is a misprint.
35
These ten papers and one letter may be found under "Bragg, W. H." in the
indexes of PM, 8, 10, 11, 13, and 14 (1904-7).
84 Ionization and the recognition ofparadox, 1906-1910
In most of these experiments, Bragg adopted Thomson's means
to measure the ionizing strength of the a-beam: A standard poten-
tial difference is applied between electrodes enclosed in the irra-
diated gas. The potential is then increased until the current stimu-
lated by the beam achieves its fully saturated value. In 1906 Bragg
discerned that some source other than the incident a-particle
beam was ionizing the gas. He quickly dismissed the possibility
that the electrode potential alone could be the cause. "If it were,"
he explained, "there would be no saturation current"; potential
gradients substantially greater than the few hundred volts per
centimeter at which saturation occurs give ions sufficient energy to
ionize atoms themselves.36 An avalanche effect occurs, and the
current increases very rapidly. Rather, Bragg suspected that the
new ionizing emanation might be the x-rays expected to arise
when the charged a-particles strike gas molecules.
But herein lay a significant problem. The ionization caused by
a- or /^-particles should saturate. Each particle can interact with
only a limited number of molecules. But the x-rays produced by
collisions should radiate in all directions (in the orthodox view)
and the impulses should affect all molecules in their vicinity
equally. Since large numbers of atoms would therefore be affected,
the ionization current should not saturate. Bragg, like Thomson,
knew that the current does saturate at microscopically small levels.
It appeared to Bragg that far too few electrons were being released.
Thus, Bragg confronted the paradox of quantity, only hinted at in
his 1904 claim for "waves so small." It was reaffirmed in a particu-
larly clear operational form for Bragg, and he could not ignore it
without upsetting his ongoing a-particle studies.
Bragg suggested a possible explanation for the ionization due to
x-rays produced by a-rays. Perhaps, he thought, the secondary
effect of the x-ray is fully expended within a microscopically small
distance from its source. Within that short distance, the impulse
might retain sufficient energy to ionize an atom. But this explana-
tion clearly could not be complete. The current induced by ordi-
nary x-rays from a discharge tube also saturates. This occurs even
after the rays have traveled distances equal to billions of atomic
diameters. Even if one allowed the impulse to be severely restricted
in its spatial width, after traversing these long distances it would
36
W. H. Bragg, PM, 11 (1906), 623.
British attempts to retain mechanism 85
have diffracted to encompass regions much larger than atoms. The
ionization data indicated that soft x-rays might do this, but hard
x-rays and y-rays acted like particles: They did not appear to
diffuse their ionizing power in space at all. By 1906 it was clear to
Bragg that something was seriously amiss in the widely accepted
hypothesis of x-ray impulses, and he reaffirmed the extension to
x-rays of the corpuscular status he had assigned to y-rays in 1904.
HOMOGENEOUS X-RAYS
Barkla had shown in 1903 that x-rays scattered off light elements
reradiate with a quality roughly equivalent to that of the primary
rays. In this way, he sought corroboration of Thomson's simple
picture of the interaction of pulses and matter. In the course of his
polarization studies in 1904 and 1905, he extended his survey to
some of the heavier elements - calcium (Ca), iron (Fe), copper
(Cu), zinc (Zn), platinum (Pt), and lead (Pb) - and found that the
secondary x-rays are consistently less penetrating than the primary
ones.55 He invoked Thomson's "thunder and lightning" clause in
52
W. H. Bragg, TPRRSSA, 31 (1907), 96.
53
"When I first put forward the neutral pair theory I was ignorant of the work of
Einstein . . . I did not think of carry ing over the idea to the theory of light; on
the contrary, I had hopes of proving that no connection existed between the
two kinds of radiation." W. H. Bragg, Studies in radioactivity (1912), 192-3.
54
W. H. Bragg, PM, 14 (1907), 429-49.
55
Barkla, PTRS, 204A (1905), 467-79.
British attempts to retain mechanism 91
i"
In
g 70 11 \ Sb
\
1 so V
A
i
s
0 y 7 V
;
1
j. 0 10 20 30 40 50 60 70 80 90 100 110 ieO 130 140 150 160 170 180 190 2(fl 210
ATOMIC WEIGHT
ATOMIC WEIGHT
12m
1HK)
AirJ
1000
J
I*
>n
ir
tot.
>
T 1
c
p
Ntt
Atomic Weip-ht.
59
Kleeman, PM, 74(1907X618-44.
60
Ibid., 643.
61
Barkla and Sadler, PM, 14 (1907), 408-22. Barkla, Nature, 75 (1907), 368.
94 Ionization and the recognition ofparadox, 1906-1910
particular transparency for the material of their origin. Secondary
rays from copper, for example, passed with lower than normal loss
through copper. Barkla called this a "property which appears to be
unknown among x-ray beams hitherto experimented upon."62
The announcement of this new sort of x-ray came in September
1907. But the full significance of Barkla's discovery only emerged
slowly over the next year as his investigation proceeded. Barkla
was encouraged to develop and extend his interpretation of the
result by the controversy that quickly developed over Bragg's
paper on neutral pairs. It appeared in England just a month after
Barkla's paper on homogeneous x-rays.
chamber at the top, the ionization was slightly greater than when
the two top plates were interchanged. This was so regardless of the
order of plates on the bottom. Kleeman's finding that a metal
emits secondary rays in rough proportion to its atomic weight thus
implied that the net ionizing radiation scattered in the "through"
direction was greater for aluminum than for lead.
When the order of plates was changed on the bottom with no
change on the top, 44 percent larger currents occurred with lead
closest to the chamber. Thus, the number of rays scattered back,
against the direction of the primary beam, was greater for lead than
for aluminum. Since all four layers were too thin to absorb appre-
ciable y-rays, either aluminum or lead or both emitted more
secondaries in the forward compared to the backward direction.
This, they claimed, was "fatal" to any transverse impulse or wave
theory of y-rays.
Madsen soon claimed that what he took to be secondary y-rays
also are asymmetrically emitted from matter.70 Bragg completed
the quatrain with an investigation of the weaker effect in secondary
x-rays. There, deep within Barkla's own x-ray research domain,
70
Madsen, TPRRSSA, 32 (1908), 163-92; PM, 17 (1909), 423-48. For the
conditions under which Madsen carried on his researches, see his correspon-
dence with Bragg in Home, Historical records ofAustralian science, 5:2 (Nov.
1981), 1-29.
British attempts to retain mechanism 99
71
Bragg found the asymmetry he sought. But Bragg, always confi-
dent and somewhat parochial, had quite neglected the momentum
carried by an electromagnetic wave. The error was first mentioned
by Charlton Cooksey at Yale when he presented evidence of an
equivalent asymmetry in the distribution of secondary electrons
stimulated by x-rays.72 For the moment, nothing was made of the
oversight.
Bragg's commitment to the neutral-pair hypothesis led to a
second important discovery in 1908. In May he listed six properties
of the y-rays that, he claimed, were best interpreted in terms of the
neutral-pair view. One, on the transformation of y-rays, had pre-
viously been buried in raw experimental data: "the penetration
and therefore the speed of the /? radiation . . . produced [by
y-rays] increases with the penetration of the y radiation to which it
is due."73 All neutral pairs of equivalent penetrating power should
release electrons of equal speed. The velocity of the secondary
betas should be independent of the number of neutral pairs and
should depend only on their penetrability.
This result was difficult for any impulse hypothesis to explain. If
one adopted the triggering hypothesis, the velocity of secondary
electrons should depend on the type of scattering material. Bragg's
conviction led directly to a test, and he was vindicated.74 As with
the photoelectric effect, measured velocities of the secondary fi-
particles were independent not only of the intensity of the incident
y-radiation but also of the nature of the scattering material.
Bragg was convinced that the neutral-pair theory explained "all
the known properties of the y-rays much more simply and com-
pletely" than did the impulse hypothesis.75 He conveniently sum-
marized his position in a series of largely rhetorical questions
related to the possible origin of secondary betas: (1) both the
particle and the energy arise from the atom (in essence, the trigger-
71
W. H. Bragg and Glassen, TPRRSSA, 32 (1908), 301-10; PM, 17 (1909),
855-64.
72
Cooksey, Nature, 77 (1908), 509-10. Thomson had pointed out in 1907 that
x-ray impulses would carry momentum. In the "bright speck" hypothesis of
light, to which we shall return in Chapter 6, he claimed that x-rays would "have
all the properties of material particles." PCPS, 14 (1907), 424.
73
W. H. Bragg, Nature, 78 (1908), 271.
74
W. H. Bragg and Madsen, TPRRSSA, 32 (1908), 35-54; PM, 16 (1908),
918-39.
75
W. H. Bragg, Nature, 78 (1908), 271.
100 Ionization and the recognition ofparadox, 1906-1910
ing hypothesis); (2) the particle comes from the atom but its energy
is transformed from the incident radiation (the impulse or wave
hypothesis) and (3) both the electron and the energy come from the
primary y (the neutral-pair hypothesis). Of (1) he inquired, how
could the penetrability of the y-ray affect the energy of the ^-parti-
cle at all? Also, why does the energy not vary with the type of atom?
Finally, how could one explain the asymmetry in secondary emis-
sion, as if the y-rays found the "guns pointed in the direction in
which they are travelling themselves?"76 Explanation (2) hardly
fared better. To use one of Bragg's favorite expressions of later
years, it was as if a plank dropped from a height of 100 meters into
the sea sent out a circular impulse that, after spreading over
thousands of kilometers, concentrated its effect on another plank,
giving it sufficient energy to impel it 100 meters into the air! The
y-ray or x-ray somehow delivered its energy bundle to a single
electron. But replace the bundle by a neutral pair, he exclaimed,
"and the whole affair seems simple enough."77 A cathode-ray
electron, incident on the anticathode, picks up a neutralizing
positive charge and becomes an x-ray or y-ray. On collision it
breaks into its component parts, and the electron continues on
with close to its original velocity.
76
W. H. Bragg and Madsen, PM, 16 (1908), 934.
77
Ibid., 936.
British attempts to retain mechanism 101
Rontgen pulses," and is therefore inhomogeneous. These scattered
x-rays are polarizable; this is a consequence of the unidirectional
cathode beam that produced the primary x-rays. The other type of
x-ray is, according to Barkla, "a homogeneous radiation charac-
teristic of the element emitting it, and produced by the motion of
electrons uncontrolled by the electric force in the primary
pulses."78
The emphasis on the characteristic nature of the secondary
x-rays was new and was offered largely as a challenge to Bragg.
Barkla had found that these homogeneous rays characteristic of
heavy elements are emitted isotropically, and they show no incli-
nation to polarize. Bragg was hard-pressed to find any explanation
using neutral pairs for the characteristic nature of the secondary
rays. Neutral pairs are, by definition, independent of the atoms
that scatter them. The hypothesis denies detailed interaction be-
tween a pair and an atom; it consequently cannot explain how the
pairs scatter so differently depending on the weight of the atom.
The discovery of homogeneous x-rays forced careful study of
their properties; it was quickly recognized that the results of all
prior experiments might have been affected by their selective but
strong influence.79 They also provided a potential source of x-rays
of definite penetrating power needed for detailed analysis of x-ray
absorption. Anticipating that fine distinctions in penetrating
power might be discerned, Barkla discovered that the secondary
rays from some elements - silver (Ag), tin (Sn), antimony (Sb),
and iodine (I) - each possess two distinct components. He called
the less penetrating component A, the more penetrating one B,
and realized that the periodicity he had charted the year before
(Figure 4.2) was due to the overlap of the two species of secondary
x-ray (cf. Figure 4.8). Within two years he had renamed the species
L and K; he thought it "highly probable that series of radiations
both more absorbable and more penetrating exist," and he wanted
to leave room on both ends for additions.80
Unlike Bragg, Barkla could at least hint at a plausible mecha-
nism for the production of the characteristic secondary x-rays.
When a pulse disturbs an atom, electrons within it should oscillate
78
Barkla and Sadler, PM, 16 (1908), 576.
79
Barkla and Sadler, PM, 17 (1909), 739-60. Barkla and Nicol, Nature, 84
(1910), 139. Barkla, PCPS, 15 (1909), 257-69.
80
Barkla, PM, 22 (1911), 4o6n.
102 Ionization and the recognition ofparadox, 1906-1910
^ (5.1)
3
McCormmach, HSPS, 2 (1970), ix-xx, 41-87. Miller, Relativity (1981).
4
Einstein, AP, 77(1905), 132-48.
5
See Klein, Natural philosopher, 2 (1963), 59-86; 3 (1964), 3-49.
6
Klein, Science, 757(1967), 509-16.
7
Wien, AP, 52(1894), 132-65; AP, 58 (1896), 662-9.
The appeal in Germany to the quantum theory 107
Here E is the total energy contained in the sample and b is a
constant.
According to Ludwig Boltzmann, the entropy change at con-
stant temperature in a gas of particles can be expressed as a
function of the gas constant R and Avogadro's number N. It is
10
Wheaton, Photoelectric effect (1971); HSPS, 9 (1978), 299-323.
11
Einstein,^, 20(1906), 199-206.
The appeal in Germany to the quantum theory 109
stein's paper of 1905 therefore marks the origin of the quantum
theory as we understand it today.12
Second, Einstein's paper contains no reference to x-rays or to
y-rays. Einstein did note that the properties of light he wished to
emphasize were most pronounced at high frequency, where
Wien's law was valid. But he did not suggest, nor is he likely to
have believed, that similar considerations should be applied to
x-rays. X-rays and periodic light were then thought to be essen-
tially different forms of radiation. Viewed retrospectively, Ein-
stein's lightquantum hypothesis solved the dual paradox just then
being recognized for any classical theory of x-rays. But it also
introduced serious difficulties for radiation theory that most physi-
cists justifiably thought insurmountable.
Einstein's hypothesis of lightquanta was not taken seriously by
mathematically adept physicists for just over fifteen years. The
reasons are clear. It seemed to be an unnecessary rejection of the
highly verified classical theory of radiation. Most physicists re-
jected the idea as merely the resuscitation of the discredited emis-
sion theory of light. In the domain of visible or ultraviolet light, not
even a hint of the x-ray paradoxes then existed. How lightquanta
could possibly explain interference phenomena was always the
central objection. Einstein offered very little explanation; reconcil-
iation of the lightquantum with interference was a problem with
which he wrestled for the rest of his life. In 1921 he characterized
the conflict to his friend Paul Ehrenfest as something fully capable
of driving him to the madhouse.13
Although Einstein's theory of relativity was soon accepted by
leading physicists, his lightquantum hypothesis was not. Einstein's
own study of the problems of radiation had almost no influence on
the work done by others before the early 1920s. For this reason, we
shall not discuss it in further detail here. Some very useful histori-
cal discussions have been provided, but it is a story yet deserving
scrutiny.14
12
Kuhn, Black-body theory (1978), argues that Einstein and Ehrenfest first
quantized the energy of oscillators in 1905-6. For another view, see Klein,
AHES, 1 (1962), 459-79-
13
Einstein to Ehrenfest, 15 March [1921]. AHQP 1, 77.
14
Klein, Natural philosopher, 2 (1963), 59-86. McCormmach, HSPS, 2 (1970),
41-87.
110 Ionization and the recognition ofparadox, 1906-1910
WILHELM WIEN AND THE ENERGY OF X-RAYS
The first person to perceive the connection of Einstein's unortho-
dox lightquantum hypothesis with the nature of x-rays was Wil-
helm Wien, since 1900 ordentlicher professor of physics at the
University of Wlirzburg. Wien was a proponent of the electromag-
netic view of nature. He was used to thinking that the properties of
matter are artifacts of the continuous electromagnetic field, not
consequences of actual material particles. He believed that there
was great value in the idea that the electron mediates between the
events in the electromagnetic field that we call radiation and those
that we ascribe to atoms. Therefore, his concepts were closely
allied to the electron theory of H. A. Lorentz, in which interaction
between matter and radiation occurs only by means of the elec-
tron.15
But Wien deviated from Lorentz in two important respects.
First, Wien thought that the electron itself has no material exis-
tence; its apparent mechanical properties are due exclusively to
electrodynamic effects. Second, he believed that the electron is
compressible, and it was this idea that first directed his attention to
x-rays.16 He gave the electron a dipole moment that could change
with time. In 1904 he treated x-ray impulses as if they arose from
impacts of this deformable electron.17 He thought the impulse
hypothesis was the "most probable" explanation of x-rays but later
that year admitted that "theoretical conceptions of x-rays can in
no way be considered certain."18 He hypothesized that some
3 x io~13 ergs of energy should be radiated away in each electron
impact, and to test this prediction he turned to experiments on the
heat produced in material bodies by x irradiation.
The results appeared in a Festschrift dedicated to Adolph
Wiillner in 1905.19 Wien found by experiment that no more than
one part in a thousand of the heating capacity of the cathode rays
actually appears as x-ray energy. Joseph Larmor had shown in
1897 that the rate at which energy radiates from an accelerated
particle with charge e is:20
15
Hirosige, HSPS, 1 (1969), 151-209.
16
Wien, Archives Neerlandaises (2), 5 (1900), 96-107.
17
Wien, PZ, 5 (1904), 128-30.
18
Wien, JRE, 1 (1904), 215-20.
19
Wien, Festschrift Adolph Wiillner (1905), 1-14.
20
Larmor, PM, 44 (1897), 503-12.
The appeal in Germany to the quantum theory 111
(5 5)
T" '
where a is the mean acceleration. Wien used a more recent form
for the total energy derived by Max Abraham to take account of
what we would now classify as relativistic effects from a rigid
spherical electron accelerating over a time interval r:21
* 71 (5-6)
24
Wien, Festschrift Adolph Wullner (1905), 7.
25
Wien,ylP, 75(1905), 1003-4.
26
Ibid., 1005.
The appeal in Germany to the quantum theory 113
revolves in the inverse-square potential surrounding a positive
charge; the orbit is of atomic radius icr 7 cm. An electron moving
5 x io7 cm/sec, slower even than electrons released by light in the
photoelectric effect, should, according to equation 5.6, radiate its
5 x io~13 erg at the rate of io"6 erg/sec. The total energy of the
electron is gone in io~7 sec!27 As early as 1905, Wien had concluded
that grave difficulties confronted the triggering hypothesis, not
simply for x-rays but for visible light as well. But he had no better
alternative to offer at the time.
Actually, the problem is not as severe as Wien implied. In the
article in which Larmor derived expression 5.5, he went on to
suggest that radiative losses might drop almost to zero for atomic
systems with more than one electron.28 This would be true as long
as the vector sum of all electron accelerations remains zero, a
situation most easily achieved when two electrons describe the
same circular orbit at opposite ends of a diameter. J. J. Thomson
developed this idea in 1903 for radiation from multielectron orbits,
finding that the energy radiated per electron drops by a factor of
roughly 1,000 for each additional electron in the ring when the
particles move at a velocity of 0.0 ic.29 Each of six electrons in a ring
radiates only 10"17 of the energy predicted by expressions 5.5 or
5.6. But of this work Wien was apparently unaware.
Wien was convinced very early that there is very little difference
between x-rays and periodic light. His use of the term Rontgen-
welle to describe x-ray pulses, as quoted above, indicates the close
connection in his mind of the two forms of radiation. At the same
time as he assumed the similarity of x-rays and light, most British
physicists were emphasizing their differences. Very likely, Wien's
conflation of periodic wave and discontinuous pulse was a product
of his experience in conjuring the localized properties of atoms out
of the continuous electromagnetic field. It is, of course, possible to
describe an impulse as the sum of many periodic waves, each of
which extends to infinity in space. In any event, the connection
Wien saw between x-rays and light, combined with his concern to
find an explanation for x-ray ionization, increased his interest in
any attempt to explain the photoelectric effect on some basis other
than the triggering hypothesis.
27
The radiated power is given on p. 1007 as io6 ergs/sec, but the sign of the
exponent is clearly a misprint.
28
Larmor, PM, 44 (1897), 503-12.
29
J. J. Thomson, PM, 6 (1903), 681.
114 Ionization and the recognition ofparadox, 1906-1910
QUANTUM IMPULSES
In the spring of 1907, Wien found what he had been looking for:
Einstein's unorthodox treatment of the photoelectric effect using
the lightquantum. In all likelihood, Wien discovered Einstein's
paper by himself. He was fully conversant with the literature on
cavity radiation, particularly Planck's work, and Einstein's paper
was closely related. Furthermore, Wien was aware of Einstein's
new relativity theory and recognized early the profound nature of
Einstein's thought. A direct influence may also have contributed to
Wien's awareness of Einstein's quantum studies. Wien's student,
Jakob Laub, completed a dissertation on secondary electrons early
in 1907.30 At about the same time, he prepared a seminar talk at
Wiirzburg on the theory of relativity.31 Wien was sufficiently
impressed to suggest that the younger man travel to Bern to meet
and learn from Einstein.32 This Laub did. After a short visit, he
returned in the spring of 1907 to Wiirzburg. The following year, he
went back to Bern to collaborate with Einstein.33 It is not at all
unlikely that Laub's close early contact with Einstein widened
Wien's appreciation for Einstein's work, even if Laub's enthusi-
asm for Einstein's lightquantum hypothesis provided nothing
new, and was not entirely acceptable, to Wien.
Wien'sfirstuse of the quantum hypothesis, the Planck theory, as
he called it, arose out of his interest in the problems of cavity
radiation. It was an attempt to extend Planck's quantum treat-
ment of heat radiation to the discrete optical emission spectrum.34
Wien sought the conditions under which emission lines might, like
cavity radiation, be expressed solely as a function of intensity and
frequency for a given temperature. But Wien soon recognized
another application of the quantum transformation rule, one that
he saw clearly only after reading about Einstein's treatment of the
photoelectric effect. It might hold the key to the difficulties yet
unresolved for x-ray ionization. "A simple generalization of the
Planck radiation theory to x-rays," he suggested in November
1907, "would not only explain, but demand, the production of the
30
Laub, AP, 23 (1907), 285-300.
31
Laub, AP, 23 (1907), 738-44; AP, 25 (1908), 175-84.
32
Seelig, Albert Einstein (1954), 85-6; Albert Einstein, Leben und Werk (i960),
i2off. For more on the Laub-Einstein connection, but only with regard to
relativity theory, see Pyenson, HSPS, 7 (1976), 83-123.
33
Einstein and Laub, AP, 26 (1908), 532-40; AP, 26 (1908), 541-50.
34
Wien, AP, 23 (1907), 415-38.
The appeal in Germany to the quantum theory 115
high velocities of secondary electrons. [It would] as well provide
yet another way to calculate the wavelength of the active x-rays."35
Note Wien's use of the word "wavelength" to characterize x-ray
impulses. The lack of distinction in his mind between impulses
and waves here yielded a benefit. The quantum transformation
relation requires that radiation be expressed in terms of a fre-
quency. An impulse has no frequency. But Wien could now
interpret the inverse pulse duration r as analogous to a frequency.
Thus, the circle was complete; by 1907 Wien characterized the
energy in x-rays not in terms of pulse amplitude but in terms of
pulse width alone. He set the kinetic energy of the electron mv2/i
equal to h/r and then derived S from the relation S = ex = ihc/mv2.
The result was S = 6.75 x io~9 cm, considerably closer than his
1905 estimate to Sommerfeld's derived value for the x-ray pulse
width. In support of his use of the quantum relation, Wien cited
the experiments by P. D. Innes showing that the maximum elec-
tron velocity stimulated by x-rays, like that stimulated by periodic
ultraviolet light, does not depend on the intensity of the x-rays but
only on their quality.36 But neither, in the first approximation, did
it seem to depend on the material from which electrons are ejected,
so Wien explained just where the analogy to light breaks down.37
Wien employed the quantum relation to achieve a specific
result; he was not particularly concerned to find a modellmdssig
interpretation for why the relation should be employed. The
quantum transformation relation offered quantitative success
without recourse to the suspect triggering hypothesis. Wien's ex-
tension to x-rays of Einstein's relation for visible light was made
possible by his close identification of periodic light with impulse
x-rays, an identification that followed in turn from Wien's accept-
ance of the electromagnetic view of nature. In December 1907, a
month after the new calculation of S, Wien declared in an impor-
tant review of radiation theory, "there seems to me to be no basis
for any essential distinction between x-rays and light."38
The quantum transformation relation was not synonymous
35
Wien, Gottingen Nachrichten (1907), 599. Laub soon repeated Wien's experi-
ment. AP, 26 (1908), 712-26.
36
Innes, PRS, 79A (1907), 442-62.
37
Wien neglected the energy required to release an electron from the metal after
it has left the atom. He felt that this was negligible relative to x-ray energies but
important for visible light.
38
Wien, Wurzburg 6(1907), 107.
116 Ionization and the recognition ofparadox, 1906 -1910
with Einstein's lightquantum hypothesis. Wien accepted the
former but not the latter. One could not assume that the energy of
x-rays exists in "indivisible quanta [in the way that] matter consists
of atoms."39 His analysis had been directed solely to the produc-
tion of x-rays by electron impacts; he did not treat the conceptually
more demanding case of x-ray absorption. Only in the latter case
need one explain how an isotropic pulse can concentrate its total
energy on an individual electron.
46
Stark, PZ, 8 (1907), 91411; 9 (1908), 767-73.
47
Einstein to Stark, 7 December 1908. Quoted in A. Hermann, Sudhoffs Ar-
chiv, 50(1966), 272.
48
Stark, PZ, 9(1908), 767-73.
49
Einstein to Stark, 2 December 1908. A. Hermann, Sudhoffs Archiv, 50
(1966), 276; Fruhgeschichte (1969), 88.
The appeal in Germany to the quantum theory 119
quantum hypothesis to calculate the high-frequency absorption
limit for the stimulation of molecular spectra.50 He soon went on
with a student to analyze the relation between photoelectric sensi-
tivity andfluorescencein some fifty-seven organic molecules.51
In the course of these studies, Stark returned repeatedly to the
close connection he perceived between ionization, the photoelec-
tric effect, andfluorescence.Even before recourse to the quantum
in 1907, he had expressed his confidence in the interconnection.52
With the success of the quantum principle demonstrated, his
confidence increased. "Absorption of light," he said, "has as its
consequence the ionization of the absorbing atom or molecule,
that is, the excitation of photoelectric cathode rays, accompanied
by afluorescenceconforming to Stokes' law."53 Stokes' law, which
stated that the frequency of incident light must equal or exceed
that of emitted fluorescent light, lacked a classical explanation. In
the spring of 1908, Stark expressed his appreciation of Einstein's
success in "illuminating another dark point offluorescencephe-
nomena" by giving the straightforward quantum theoretical basis
for Stokes' law. This prompted his request of Einstein in 1908 for
an elaboration of the subject, suitable for chemists' understanding,
to appear in the Jahrbuch.54
A review article by Charles Barkla, which spelled out the de-
tailed properties of characteristic x-rays for the first time in the
German language, appeared early in 1908 in Stark's Jahrbuch.55
Stark, as editor, had known of its content for some time. In that
paper, Barkla revealed that the characteristic secondary x-rays
possess precisely the trait that Stark had emphasized in his long
study of optical absorption: "properties of the chemical elements
that are periodic functions of atomic weight."56 Barkla emphasized
that "without exception" the stimulation of a homogeneous radia-
tion requires a primary beam of greater penetrability than that
produced. Further, the intensity of the primary beam could be
50
Stark, PZ, 9(1908), 85-94, 356-8.
51
Stark and Steubing, PZ, 9 (1908), 481-95; 661-9.
52
Stark, PZ, 5 (1907), 8 in.
53
Stark, PZ, 9(1908), 92.
54
Stark to Einstein, 11 February 1908. Hermann, Sudhoffs Archiv, 50 (1966),
272-3.
55
Barkla, JRE, 5 (1908), 264-324.
56
Stark, JRE, 5 (1908), 124-53.
120 Ionization and the recognition ofparadox, 1906 -1910
altered almost at will without affecting the hardness of the sec-
ondary.57
Stark had already converted in his own mind x-ray "hardness"
or "penetrability" to frequency. He recognized immediately that
Barkla had found the x-ray analog to optical fluorescence. The
hardness threshold for the stimulation of characteristic x-rays was
a simple case of Stokes' law offluorescence.Stark was very early in
this realization; it was to be three years before Barkla would speak
of fluorescent x-rays as "an extension of Stokes' Law." Stark was
quick to ensure that his quantum view of characteristic x-rays
appeared in English in Nature.58
Stark was unusual among German physicists in his early recog-
nition of the problems surrounding the paradox of quantity. Be-
fore adopting the quantum relation, he puzzled over the "small
fraction of the total number of molecules" that are placed in a
"chemically different state" after absorbing light energy.59 Follow-
ing contemporary work by Lenard on fluorescence, Stark came to
the conclusion in 1908 that the effect of light is "localized in
definite atomic groups."60 Then Barkla raised the analogous issue
for x-rays. The primary impulse "passes over every single elec-
tron" in all atoms of the scattering substance, but "only a small
fraction of the total number of electron [radiation] sources pro-
duce the homogeneous [secondary] radiation."61 Not only did
Barkla provide the evidence for Stark's extension of fluorescence
to x-rays, he pointed out the single most difficult observation to be
explained in a theory of spreading pulses, whether treated with the
quantum or not.
57
Barkla, JRE, 5 (1908), 229, 306, 308, 309, 317, 318, 319.
58
Stark, Nature, 77(1908), 320.
59
Stark, PZ, 5(1907), 8211.
60
Stark, JRE, 5(1908), 149.
61
Barkla, JRE, 5 (1908), 320, 323.
The appeal in Germany to the quantum theory 121
worked so well to explain observed frequency limits and secondary
electron velocities that little doubt remained in his mind about the
validity of the relation. Although Stark adopted the quantum
relation, he tended more than most German physicists to interpret
the elementary processes in terms of conceptual mechanisms.
Unlike most of his German colleagues, Stark took very seriously
the observation that x-ray ionization affects only a very small
fraction of atoms. Sometime before 1909, this forced him to adopt
the lightquantum as a physical reality.
Stark may have adopted the lightquantum by the time of his talk
to the Naturforscherversammlung held in Cologne in 1908. There
was a marked shift in his terminology. What had formerly been the
Planck'sche Elementargesetz, or Planck's law, became the Licht-
quantenhypothese. His talk at the meeting was entitled "New
observations on canal rays with respect to the lightquantum hy-
pothesis," and he considered the significant new case of spectral
lines radiated after ion collisions.62 He had undertaken the experi-
ments, he claimed, "as a proof of the lightquantum hypothesis,"
and expressed the hope that so suspect a motive might not dis-
credit his results. "This hypothesis is so unusual and so strongly
opposes the customary concept of emission and absorption of light
that the general reservations against it are perfectly understand-
able," he said.63 All he actually used in the study was the quantum
transformation relation, which by that time had been adopted by
Planck, Wien, Einstein, and others. That Stark so clearly recog-
nized the controversy into which he headed suggests strongly that
he had already accepted the spatially localized lightquantum.
If Stark had previously been ignorant of Einstein's work, he was
no longer. He cited all three relevant papers, doubtless assisted by
the reprints Einstein had sent him. He spoke freely about the
transfer of radiant energy in units of lightquanta, although such a
view clearly did not require accepting the lightquantum per se.64 In
an Anhang added to the printed version of his talk, he reviewed the
successes of the hypothesis, significantly crediting it solely to
Einstein. His playing down of Planck's contribution may have
62
Stark, PZ, 9(1908), 767-73.
63
Ibid., 768.
64
As Einstein is reputed to have put it: "Although beer is sold only in pint bottles,
it does not follow that it exists only in indivisible pint portions." Frank,
Einstein (1953), 71.
122 Ionization and the recognition ofparadox, 1906 -1910
been due to his acceptance of Einstein's 1906 claim that Planck's
work was a consequence of the more basic lightquantum assump-
tion. There was at that time general misunderstanding ofjust what
Planck had done in 1900.65 Einstein, who had hoped to attend the
meeting in Cologne at which Stark spoke, decided instead to spend
his short vacation elsewhere. But he was naturally exceedingly
interested in Stark's remarks because he saw how close Stark had
come to accepting the "revolutionary" lightquantum.66
I suggest that Stark had already done so. The terminology of his
Cologne talk strongly suggests it; but radiation from canal rays -
light produced by mechanical impact - did not require spatially
localized rays in the same way as did absorption of light. Stark had
only to express his results in a manner that would allow him to
refer back to this work as consistent with lightquanta. Stark had
other reasons to pause before taking so bold a step. He knew how
controversial the topic was. And he was at that moment embark-
ing on a bit of political maneuvering to obtain a regular appoint-
ment at Aachen. For this he was counting on, and getting, strong
support from the former resident physicist, Arnold Sommerfeld.67
Sommerfeld was the foremost German proponent of the im-
pulse hypothesis for x-rays, precisely the object of Stark's quantum
assault. Immediately after he received the post, with Sommerfeld's
endorsement, Stark offered Einstein a position as his assistant.
Einstein wisely declined.68 The next month, Einstein published a
masterful review of the state of radiation theory and began to
formulate a statistical treatment of radiation based on large num-
bers of lightquanta.69 Stark wrote that the new view interested him
greatly.70 Four months later, Stark proclaimed publicly that he had
adopted the spatially localized lightquantum for x-rays.71
His paper was entitled "On x-rays and the atomic constitution of
radiation." The critical transition to the lightquantum was de-
65
Klein, AHES, 1 (1962), 459-79. Kuhn, Black-body theory (1978).
66
Einstein to Stark, 2 December 1908. A. Hermann, Sudhoffs Archiv, 50
(1966), 276.
67
Stark to Sommerfeld, 12 October 1908. A. Hermann, Centaurus, 12 (1968),
41-2.
68
Einstein to Stark, 6 April 1909. A. Hermann, Sudhoffs Archiv, 50 (1966), 278.
69
Einstein, PZ, 10 (1909), 185-93. See also the discussion between Einstein and
Ritz, ibid., 224-5, 323~4-
70
Stark to Einstein, 8 April 1909. A. Hermann, Sudhoffs Archiv, 50 (1966), 278.
71
Stark, PZ, 10(1909), 579-86.
The appeal in Germany to the quantum theory 123
signed to win converts directly from the impulse theory. First, he
calculated the mean time between electron impacts in the x-ray
tube; typically operating at 30,000 V and passing 3 X io~4 amps,
the tube produced 1014 collisions every second. On the other hand,
the accepted pulse width, io~8 cm, implied a mean pulse duration
of some io~19 sec. It was extremely unlikely, Stark claimed, that
two or more pulses would combine their effect on the same atom.
The impulses were separated by an interval io5 times longer than
the duration of any pulse. Note that Stark implicitly assumed that
ionization begins instantaneously when the radiation arrives; he
had explicitly claimed this before, but did not do so now.72
Stark sketched a vague mechanism that he thought might ex-
plain interference for visible lightquanta. Optical lightquanta are
radiated from "centers" in atoms that are generally far closer
together than the wavelength. Thus, a "space-time superposi-
tion" of lightquanta arises that has a "complex of specifically
ordered properties." It was not clear what Stark meant by this, and
it was likely that Stark himself did not know. In some unexplained
manner, the superposition of lightquanta would apparently direct
quanta toward regions of constructive interference and away from
regions of destructive interference. For wavelengths that are short
compared to the separation of the centers, on the order of io"8 cm
or less, the complex does not arise and the radiation acts like more
or less directed units.73 Stark concluded that the "oscillating en-
ergy" of each impulse must therefore be contained within a vol-
ume of space equal to (c/v)3. He no longer bothered to qualify his
use of the then standard frequency symbol to characterize x-rays,
while at the same time his argument depended on the rays' spatial
and temporal limitations.74 This article formally opened Stark's
attempt "as quickly as possible to bring the conflict between the
lightquantum hypothesis and its older opposing concepts from the
sphere of speculation and theoretical discussion to the firm ground
of experiment."75
72
Stark had recognized the significance of an immediate response in optical
fluorescence. He thought that his chemical interpretation explained "the fact
that the onset and cessation of fluorescence occurs almost at the instant the
illumination start and stops." Stark and Steubing, PZ, 9 (1908), 491.
73
Stark,/>Z, 9(1909), 582.
74
Ibid., 583. The standard symbol for frequency was n, not v.
75
Ibid.
124 Ionization and the recognition ofparadox, 1906-1910
>30*
76
Stark, PZ, 70(1909), 902-13.
77
Cooksey, Nature, 77(1908), 509-10. W. H. Bragg and Glasson, TPRRSSA, 32
(1908), 301-10.
The appeal in Germany to the quantum theory 125
Electron beam
direction
79
Bassler, AP, 28 (1909), 808-84. Sommerfeld, Scientia, 51 (1932), 48. Other
important evidence was provided by Kaye, PCPS, 15 (1909), 269-72.
80
Sommerfeld, PZ, 70(1909), 969.
81
Ibid.
The appeal in Germany to the quantum theory 127
within the atom. He thereafter called these fluorescent rays. The
partially polarized, aperiodic transverse impulses created by elec-
tron impacts he called the Bremsanteil or braking part of the rays.
This division was not original; the British and Stark had made it
before.82 But although Sommerfeld allowed the homogeneous
fluorescent rays a wavelength, he categorically denied it to the
Bremsanteil "For this, as before," he said, "one would more
correctly speak of pulse width rather than wavelength."83 Although
Sommerfeld here first admitted that the Planck relation might be
needed to treat the fluorescent x-rays, he was certain that it had
"nothing to do with the Bremsanteil" The division between the
two types of x-rays seemed so strict to Wien that he quickly wrote
to be assured that Sommerfeld did not conceive of either type as
"nonelectromagnetic" in origin.84
Then Sommerfeld turned to the defense of the electromagnetic
theory of pulses. "Recent speculations on the lightquantum," he
said, "have tarnished our trust in the validity of electromagnetic
theory."85 To set the picture straight, he showed that the directed
field amplitudes AE and AH at a great distance from an electron,
stopped instantaneously from a velocity v, are both proportional to
(v/c) sin 6
(5.8)
1 (v/c) cos 6
Here 6 is the angle between the direction of the acceleration and
the direction of interest. For the more realistic case in which the
deceleration occurs over a finite time interval, he argued that the
field amplitudes become proportional to
Sin
(5.9)
[1 -(v/c) cos 6]3
Energy radiates instantaneously at a rate given by the product of
these vectors; the total energy radiated comes from integration
over the duration of the impulse.86 Sommerfeld found that the
total energy flux varies with 6 as
82
Stark had suggested that there was a distinction between impulse and fluores-
cent x-rays, PZ, 10 (1909), 579-86. He was one of the first physicists to point
up the close parallels between x-rays and light, their identity being a viewpoint
that, he said, was "gaining ground." See also Stark, PZ, 10 (1909), 614-23.
83
Sommerfeld, PZ, 70(1909), 970.
84
Wien to Sommerfeld, 27 December 1909. Sommerfeld papers; AHQP 34, 13.
85
Sommerfeld, PZ, 10 (1909), 976.
86
Poynting, PTRS, 175 (1884), 343-61.
128 Ionization and the recognition ofparadox, 1906-1910
70
sin2 6 / 1
S(6) = - 1 (5.10)
cos0 \[l - (v/c) cos 6]4
The maximum of this expression occurs at an angle with cosine
equal to v/c. The closer the initial velocity is to c, the greater is the
fraction of energy radiated toward the front. The resulting inten-
sity distribution is shown in Figure 5.4. Roman numerals denote
the initial electron velocities c/10, c/5, and c/3 in increasing order.
The maximum in the intensity at about 6 = 6o for case III agreed
well with Stark's measurement for an initial electron velocity
near c/3.
Sommerfeld commented that the asymmetry he had derived
classically could be visualized in a simple diagram. In Figure 5.5,
point / marks the beginning of the electron's deceleration. Point x
marks the position that the electron would have occupied had it
not slowed at all. The expanding impulse front is centered on /.
The trailing edge of the impulse emanates from point/ where the
deceleration ends. The resulting pulse width varies with the azi-
muthal angle 0, as shown exaggeratedly in the figure. The field
vectors are maximum where the density of the radial force lines is
greatest. This will occur in the plane perpendicular to the instanta-
The appeal in Germany to the quantum theory 129
neous velocity vector of the electron. But it is not until after the
pulse has passed a point in space that the field, so to speak, knows
that any acceleration has occurred. The maximum energyfluxwill
always be directed through the region of the impulse that intersects
an imaginary plane perpendicular to the acceleration vector and
centered on the moving point x Consequently, the maximum
x-ray intensity will always be in the forward direction, and will
occur at smaller angles 6 the greater the initial electron velocity.
But the net energy flux is compounded of both the quantity of
pulses and the quality of each. Sommerfeld turned to Wien's
analysis of 1905 to point out something "that appears to be impor-
tant in connection with recent observations."87 Figure 5.5 directly
answers Stark's claim that the penetrating power of the x-rays
decreases for increasing azimuth. The impulse width increases
with angle 6 up to 1800. Thomson had shown that the energy
carried by a pulse varies as i/d; but he had not noticed that S will
87
Sommerfeld, PZ, 10 (1909), 975.
130 Ionization and the recognition ofparadox, 1906 -1910
vary with the azimuthal angle around the pulse. In pointing this
out explicitly, Wien and Sommerfeld were providing a significant
reformulation of the definition of radiant intensity.88
If an electron oscillates periodically about a fixed point, it
radiates a periodic wave that varies in amplitude with the azimuth,
but does not vary in wavelength or in frequency. An impulse acts
differently. Both its amplitude and its pulse width vary with the
azimuth. The electromagnetic impulse hypothesis offered a con-
ceptual route, unavailable with periodic waves, whereby the prop-
erties associated classically with wave amplitude could be trans-
ferred to pulse duration or frequency. When it became clear that a
quantum theory of x-rays was needed, Sommerfeld's and Wien's
analyses justified assigning a significant role in defining the energy
content of impulse radiation to the impulse's duration rather than
to its amplitude.
Stark had placed himself in so indefensible a position that an
essential point he wished to convey was obscured. In describing the
radiation from a single electron, Stark had said that its energy
density far from the electron "has equal magnitude in all direc-
tions."89 Sommerfeld showed without question that Stark's claim
was simply wrong. Stark had confused the equality of energy in all
directions with the constancy of energy in a given direction. When
Sommerfeld pointed out his error, Stark retreated to a position
somewhat closer to his intent. What he now sought to explain with
the lightquantum was the fact that "the electromagnetic radiation
from a single electron does not spread to a volume that grows as the
square of the separation from the emission center."90 He now said
that what he had called the Atherwellenhypothese was not compat-
ible with Maxwell's electromagnetic equations.91 He seems to have
meant, in accordance with the successes of the quantum hypoth-
esis, that a true aether wave must be able to transfer energy in this
non-Maxwellian way.
Stark's position was no clearer to others. Einstein commented
privately that Stark had "once again produced pure dung."92
88
Wheaton, HSPS, 11 (1981), 367-90.
89
Stark, PZ, 10 (1909), 903.
90
Ibid.
91
Stark to Sommerfeld, 6 December 1909. Hermann, Centaurus, 12 (1968), 46.
92
"Stark hat wieder einmal gediegenen Mist produziert, Sommerfeld wohl die
Beweiskraft jenes Phanomens uberschatzt." Einstein to Laub, 16 March 1910.
From a transcription by Laub in the Einstein papers.
The appeal in Germany to the quantum theory 131
Sommerfeld was unable to fathom what Stark meant and con-
cluded that his caricature of classical electromagnetic theory had
only been a straw man set up especially for the kill. "Your aether-
wave hypothesis," he wrote, "is electromagnetically impossible
and has neither followers nor interest; it makes no sense [for you]
to oppose it by experiment."93 In the ensuing discussion, Stark
deleted all reference to his unfortunate phrase "equal magnitude
in all directions." Instead, he emphasized only that energy will
remain constant in a given direction.94 Sommerfeld knew perfectly
well that Stark had changed his stance, and called the later exposi-
tion an "artfully dissected transcription" of the original version.95
Stark'sfinalpaper on the x-ray asymmetry reported only empirical
findings that were no longer in doubt. No interpretation of the
results, which would have brought him back into the dispute, was
provided.96 In any given direction from the source of an x-ray
impulse, the electromagnetic amplitude decreases as the pulse
spreads over larger volumes of space; the pulse width likewise
changes. But the frequency of a periodic disturbance remains
constant in a given direction and is equal in all directions. Already
used to interpreting pulse duration as an intrinsic frequency for
x-rays, Stark saw its directed constancy as the key to the preserva-
tion of the energy units required by the quantum relation. Som-
merfeld's impulse did not fulfill that condition. When Stark tried
to point this out after the fact, it appeared only that he was
switching to a new argument to save face. That may have been the
case, but Stark had also raised an important question. Not only did
the electromagnetic theory have to show that energy is localized, it
also had to demonstrate how that energy stays localized as the
"wave" propagates.
In the first volume of a long exposition on the Principles of
atomic dynamics, published in the midst of the controversy with
Sommerfeld, Stark stated his general concern. He opposed what he
called the uncritical application of traditional continuum me-
chanics to the new field of elementary processes. "All objections
raised to date against the lightquantum hypothesis are based on
93
Sommerfeld to Stark, io December 1910. A. Hermann, Centaurus, 72 (1968),
48-9.
94
Stark, PZ, 11 (1910X30.
95
Sommerfeld, PZ, 11 (1910), 101.
96
Stark, PZ, 11 (1910), 107-12.
132 Ionization and the recognition ofparadox, 1906 -1910
incorrect or unprovable assumptions," he said.97 He began to do
experiments on visible and ultraviolet light.98 Even stronger claims
for the lightquantum appeared in the second volume of Stark's
Atomic dynamics, entitled Elementary radiation, which appeared
in 1911. There Stark reformulated the lightquantum in terms more
suited to the temperament of German physicists. It should be
thought of as a "deformation in the energetic medium" that
remains localized.99 Stark could afford to come somewhat closer to
Sommerfeld's position because by 1911 it was clear that the energy
of x-rays is distributed far less isotropically than had been thought.
From this point on, the issue was no longer whether x-rays diverge
spherically from a source, but whether they diverge at all. And it
was far from clear that any electrodynamic theory, with or without
the new definition of frequency and intensity, could achieve the
degree of localization that ionization experiments implied the
energy of x-rays possessed.
97
Stark, Electrischen Quanten (1910), vii.
98
Stark, PZ, 11 (1910), 179-87.
99
Stark, Elementare Strahlung (1911), 283-4.
PART III
Seeking an electrodynamic solution
1907-1912
Localized energy in spreading impulses
2
E. Ladenburg, VDpG, 9(1907), 504-14.
Localized energy in spreading impulses 137
the large number of atoms provided the apparent continuum
effect? This proposal was disagreeable to anyone familiar with
chemistry and conflicted with spectroscopic evidence for atomic
homogeneity. Could it possibly be that within a single atom so
large a number of vibrational or rotational frequencies are repre-
sented? To account for this, extremely large numbers of electrons
would have to reside in each atom.
Just the year before, Thomson had challenged the prevailing
notion that atoms contain large numbers of electrons.3 Based on
results of scattering experiments, Rutherford and Bragg had sup-
posed that most of the atomic weight of any atom consists of
electrons; a single atom of hydrogen would possess a swarm of
1,000 electrons. But Thomson, combining evidence from ^-parti-
cle absorption, x-ray scattering, and optical dispersion, revised the
estimate downward by a factor of io3. This left about as many
electrons per atom as the atomic weight in units of hydrogen
atoms.4 It was difficult to see how so few electrons could execute
mechanical oscillations of enough different frequencies to explain
Ladenburg's findings. The paradox of quality, eclipsed by the
triggering hypothesis since 1902, began to exert its influence on
Thomson.
The year before, Thomson had suggested to P. D. Innes that
some further tests be made on x-rays by analogy to the photoeffect.
Innes had found that the velocities of secondary electrons due to
x-rays are independent of the intensity of the x-rays and depend
only on their quality.5 The similarity to photoeffect behavior had
convinced Innes that the triggering mechanism should be ex-
tended to x-rays. And Thomson, like Wien before him, temporar-
ily adopted the triggering hypothesis for x-ray ionization. Now
Thomson realized that, whatever the theoretical difficulties for a
triggering explanation of the photoeffect, the same considerations
would apply to ionization by x-rays. Within a month of Laden-
burg's announcement, and just as Bragg's paper on neutral-pair
3
J. J. Thomson, PM, 11 (1906), 769-81.
4
Heilbron, AHES, 4 (1968), 269-74.
5
Innes used a variable spark gap to measure the x-ray tube potential and,
consequently, the x-ray quality. He found that changing the gap from 5.5 to
16 cm produced a 7 percent increase in the maximum velocity of released
electrons, PRS, 79A (1907), 442-62.
138 Seeking an electrodynamic solution, 1907-1912
x-rays appeared in England, Thomson returned to the problem of
radiation.6
The problem was especially acute for x-rays. Whereas the
maximum velocity of electrons released by ultraviolet light is on
the order of io8 cm/sec, Innes had corroborated Dorn's figure of
several times io9 cm/sec for the velocity of x-ray electrons. The
energy differs by a factor of about io3; Thomson believed it to be
io4. Velocities of electrons released by y-rays are even greater. Even
if the atom encompassed the profusion of electron motions re-
quired by the triggering explanation of the photoelectric effect, an
entirely new set of motions, with energy io3 times greater, would
be required by the x-ray data. The y-rays needed an even more
energetic set. It was clear to Thomson that, at least for x-rays and
y-rays, the energy could not come from the atom but must be
transformed from the radiation itself. The double paradox of
quantity and quality convinced him in 1907 that the energy of the
new radiations comes in units, the "bright specks on a dark
background" that he had first discussed in 1903.7
Thomson had already calculated the size of the x-ray unit. In
1906 he had used the absurdity of the result to support the trigger-
ing hypothesis.8 The transverse electricfieldamplitude AE acts on a
charge e to induce a velocity measured to be at least io9 cm/sec.
The net change in electron velocity must be (AE e/m)dt where dt is
the time the pulse takes to pass the electron. The known value of
e/m implies that AEdt is approximately 60 g cm/emu/sec. But the
energy density in the expanding impulse at the position of the
ionized electron is
AE2S/4nc2 = AE2 dt/Anc (6.1)
Substituting 60 for AEdt, and putting an upper limit of io~5 cm for
<5, Thomson found that almost 1 cal/cm2 has to be provided by the
x-ray impulse.9 The results of his earliest analysis showed that even
in the direction of maximum radiant intensity, 6 = 900, a large
6
J. J. Thomson, PCPS, 14 (1907), 417-24. This answers McCormmach's query
on why Thomson should have revived his 1903 speculations on light at this
time. McCormmach, BJHS, 3 (1967), 372, 374.
7
J. J. Thomson, Electricity and matter (1904).
8
J. J. Thomson, PCPS, 13 (1906), 321. It was, he said, "an altogether inadmissi-
ble amount" of energy. The following year, he interpreted the result to indicate
the extreme coarseness of the distribution of energy "units" in x-rays.
9
J. J. Thomson, Conduction of electricity, 2nd ed. (1906), 320-1.
Localized energy in spreading impulses 139
20
enough energy density occurs only within io~ cm of the origin of
the x-ray pulse.10 This is inside the decelerating electron itself!
In 1907 Thomson turned the argument around. If the energy
was bound up in units, he said, the spatial density of those units
needed only to be correspondingly low. For the faint but visible
light intensity of io"4 erg/cm2/sec, there would be only 1 unit/liter
of space - "exceedingly coarse," Thomson said.11 "As these units
possess momentum as well as energy they will have all the proper-
ties of material particles," he continued. "Thus we can readily
understand why many of the properties of the y-rays resemble
those of uncharged particles moving with high velocities."12
The implicit reference to Bragg's neutral-pair hypothesis was
clear enough. Thomson tested Bragg's hypotheses: Were the slow
electrons emitted by different causes of ionization actually of
uniform velocity? It appeared that they were. Thomson found that
the electrons released by cathode rays had about the same velocity
as did those elicited by canal rays. The velocity was independent of
the type of irradiated gas and of its pressure.13 Early in 1908,
Thomson concluded that this common effect could be explained if
"a doublet consisting of a corpuscle and a positively charged
particle" was removed from the atom.14 He soon agreed, at the
height of the Bragg-Barkla controversy, that Bragg's hypothesis
had much to offer, not just for x-rays but for y-rays as well.15
But Thomson did not believe that x-rays are particles. He sought
a solution within quasi-classical electromechanics. The coarse-
grained aether, he thought, would rescue the advantages of a wave
model of radiation and still explain the dual paradoxes. The energy
transmitted in x-ray and y-ray impulses is, in his interpretation,
concentrated in specific directions, so that few atoms are hit, and
those receive close to the full complement of radiant energy. He
expected to be able to explain the similar behavior of periodic light
waves in much the same way.
10
J. J. Thomson, PM, 46 (1898), 528-45.
11
J. J. Thomson, PCPS, 14 (1907), 423-
12
Ibid., 424.
13
Thomson, PCPS, 14 (1908), 541-5. Fuchtbauer, PZ, 7 (1906), 748-50.
14
Ibid., 545.
15
J. J. Thomson, PCPS, 14 (1908), 540. Just what the effect of this was on Barkla,
who saw himself as the defender of Thomson's views, is not clear. But the
message was limited to the small readership of the PCPS. Not until 1910 did
Thomson raise these issues in the widely distributed PM.
140 Seeking an electrodynamic solution, 1907-1912
THOMSON'S RESTRUCTURED AETHER
In 19 io, after two years of silence on the nature of light, Thomson
announced that prior concepts about the microscopic structure of
the aether must be in error.16 According to his "coarse-grained
aether" hypothesis, radiant energy does not propagate spherically
from its source because the structure of electric field lines sur-
rounding each electron is microscopically discontinuous. Now he
went even further, suggesting that each electron is the seat of only a
single "tube" of electric force; its energy is radiated only into one
narrow cone. Macroscopic charged bodies appear to exert electric
force isotropically only because of the large number of electrons
they contain. Each of these electrons broadcasts its influence in
only one specified direction in space.
The impulse produced by a rapid displacement of the electron is
represented as a transverse kink in the electric field, as in Figure
6.1. The kink travels outward along the cone of solid angle Cl
defined by the tube. The transverse electric force is approximately
time
27
Laby, PCPS, 15 (1909), 106-13. See also Laby and Burbidge, Nature, 87
(1911), 144; Campbell, PCPS, 15 (1909), 136.
28
Campbell, PCPS, 15 (1909), 117-36. By "atomic" theories of radiation,
Campbell had in mind "at least two:" Thomson's unit hypothesis and
"Planck's theory (as interpreted by Stark)," PM, 19 (1910), 190.
29
Campbell, PCPS, 15 (1909), 310-28, 513-25.
30
Others, including Meyer and Regener, typically used a "Bronson" resistance -
a gas cell in which low conductivity is induced by a source of ionizing
radioactive material. Campbell understandably criticized Meyer for introduc-
ing an additional source of fluctuation and replaced the Bronson resistance by
a liquid measure of xylol and ethanol, calibrated at 9 x io10 1.
146 Seeking an electrodynamic solution, 1907-1912
tions of interest.31 A few months later he reported failure, for the
same reason, in a test on x-rays.32
Although Campbell reached no experimental conclusion, he
considerably advanced the analysis of the problem. The mean
square fluctuation in the number of electrons registered from a
single photocell is
~ (~ + ~tf)dt, (6.7)
where TV is the mean number of "light disturbances" (spherical
impulses or localized units) that appear in a unit of time.33 The
period of illumination is dt, and co is the number of electrons
released by each disturbance; today we would call co the quantum
efficiency. But co can vary from one ionization event to another,
and t] measures the instantaneous deviation of co from its mean
value characteristic of the cathode atoms, the light frequency, and
light intensity. Campbell fixed attention on the change in rf when
the illuminated area of the photocathode becomes larger and the
photocurrent consequently increases. The larger area serves to
integrate irregularities in electron receptiveness to the light over a
larger sample of atoms. According to the classical wave theory of
light, waves spread to encompass all electrons in the illuminated
surface, so that fluctuations in co should approach zero when larger
areas are illuminated. For the rival theory, this is manifestly not
the case. A directed light uni^affects electrons only along a specific
direction from the source; rf should remain roughly constant as
the cathode area increases.
If one increases the intensity of the light while holding the
illuminated area of the photocathode fixed, the current also
changes. Under all circumstances the current is proportional to
Nco, the total number of electrons released per unit time. For
31
The Nernst lamp was designed to compete with carbon-filament light bulbs. It
needed no surrounding vacuum and produced a higher ratio of illumination to
electrical power. Nernst sold the patent rights to his lamp to the Allgemeine
Elektrizitats Gesellschaft in the mid-1890s for a million marks, a good-sized
fortune at the time. But the zirconium rare-earth compound used as the
filament requires preheating to bring its internal resistance down to allow
steady operation, and the lamp was never produced commercially. See Men-
delssohn, Walther Nernst (1973), 44-6.
32
Campbell, PZ, 11 (1910), 826-33. _
33
Campbell, PCPS, 15 (1909), 314, shows rf rather than rf, but this is clearly a
misprint.
Localized energy in spreading impulses 147
spreading waves, the number of impulses TV is independent of the
distance from light to cell. Thus, the decrease in current with light
intensity must be due to a corresponding decrease in the potency
of each event, co. Neglecting the variation in co, the mean square
fluctuation in ionization current is then proportional to co2. Since
the light intensity must be proportional to co, x2 should be propor-
tional to the square of the light intensity. But it was Campbell's
stated "hope" to find evidence against the "spherical wave theory"
of light.34 For the unit hypothesis, co remains constant; each
ionizing event, if it occurs at all, delivers the full unit of energy.
Thus, it must be N that changes with the intensity of the light. A
smaller solid angle of light containing fewer energy bundles will hit
the photocathode as the source is moved away. In this case,
fluctuations should depend directly on N, that is, in proportion to
the first power of the light intensity.
But Campbell's aims were obstructed again. Wheji he included
the anticipated_variation in co by putting back the rf term in the
expression for x2, both cases were reduced to simple proportion-
ality of fluctuation with light intensity. Having considerably ex-
tended the analysis, he was left with no more than a plausibility test
of his method. He derived an order-of-magnitude estimate for the
elusive quantity co, and compared his estimate to what he called
Planck's theory, but which was actually derived from Einstein.35
Einstein predicted that co = i under all circumstances. Abram
Ioffe's reworking of Ladenburg's experimental photoeffect data for
the energy of electrons released from zinc suggested that co = 3.36
Campbell s independent data on fluctuations in the current from
zinc-amalgam photocathodes indicated that co = 3.1, a factor 100
times lower than that expected from any spherical wave theory of
light.
ANISOTROPIC y-RAYS
ei2
that is, directly proportional to the ratio of the currents themselves.
But for the anisotropic case, TV varies while co remains constant.
Thus, the ratio of fluctuations is
Nxcoe \I~N~2 (6.9)
ei2
150 Seeking an electrodynamic solution, 1907-1912
that is, proportional to the square root of the ratio of currents.
Meyer's data weakly indicated that the anisotropy was the more
likely result, the mean ratio of fluctuations in each of two cases
being at least 50 percent closer to the square root of the ratio of
currents than to the ratio itself.
Meyer did not deny that the impulse hypothesis might explain
his result. He did wonder aloud whether Sommerfeld's success in
finding asymmetry in x-rays to answer his Aachen colleague,
Stark, could also explain the degree of anisotropy he had now
found in y-rays. Early in 1911, Sommerfeld demonstrated that it
could.
44
Walter, AP, 66 (1898), 74-81.
45
He published extensively in the Fortschritte auf dem Gebiete der Rontgen-
strahlen, a journal serving those with medical interest in x-rays. For a discus-
sion of Walter's role in setting early radiological standards, see Serwer, Radia-
tion protection (1976).
46
Walter, KDM4 (1901), part 2, p. 43',PZ, 3(1902), 137-40. Haga and Wind, AP,
10 (1903), 305-12. Walter, Fortschritte auf dem Gebiete der Rontgenstrahlen,
9(1906), 223-5.
47
Franck, VDpG, 70(1908), 117-36. Marx, VDpG, 70(1908), 137-56, 157-201.
Franck and Pohl, VDpG, 10 (1908), 489-94. Marx, AP, 28 (1909), 37-56,
153-74; 33 (1910), 1305-91. Franck and Pohl, AP, 34 (1911), 936-40. Marx
stopped the exchange with the question, "Are my experiments on the velocity
of x-rays by interference of electrical waves explainable?" AP, 35 (1911),
397-400.
48
Walter and Pohl, AP, 25 (1908), 715-24; 29 (1909), 331-54. Pohl, Physik der
Rontgenstrahlen (1912), 144.
152 Seeking an electrodynamic solution, 1907-1912
experiments had been encouraged by the new theoretical objec-
tions raised by Stark in 1908 against an aether model of x-rays.
The chief motive behind Sommerfeld's analysis of y-rays was his
growing interest in the quantum transformation relation {not the
lightquantum). Late in 1910, H. A. Lorentz had questioned
whether the quantum hypothesis could explain absorption of
radiation at all without some additional hypothesis to limit the
spatial extent of radiant energy.49 In 1909 Sommerfeld had denied
that the quantum relation had anything to offer to the analysis of
impulse radiation; in 1911 he explicitly reversed that statement.50
The y-ray study was intended to answer Lorentz and was Sommer-
feld's first use of the quantum transformation relation. It was
intended to show how the quantum relation, formulated in terms
of frequency, could be extended to nonperiodic events such as the
ejection of an electron by a disintegrating atom.
The same asymmetry that had answered Stark's objections
regarding x-rays also served to define the localization of y-ray
impulses. The two cases are remarkably similar even though in one
case the electron is stopping and in the other it is starting. The
electron accelerates from rest at point / toward the right in Figure
6.4 and attains its final velocity at point / The resulting field
exhibits a variation in pulse width around the azimuth similar to
that of the decelerating electron in Figure 5.5. One qualitative
difference is that the quiescent and dynamic fields are inter-
changed; here the former, which Sommerfeld called the electro-
magnetic atmosphere, is the outerfield.51Another difference is that
once the electron reaches relativistic velocities, the dynamic field
within the inner sphere deviates from the schematic symmetry
shown in Figure 6.4. The field of a relativistic electron becomes
restricted to the region between two cones facing each other, each
with its vertex on the electron and symmetrically oriented about
the axis of acceleration.52 As the velocity approaches c the angle
49
Lorentz, PZ, 11 (1910), 1234-57. These matters will be discussed at length in
Chapter 7.
50
Sommerfeld, Miinchen Sb, 41 (1911), 43n.
51
Ibid., 4.
52
Heaviside, PM, 27 (1889), 324-39. J. J. Thomson had concluded in 1898 that
the electrodynamic effect of stopping an electron abruptly from a speed close
to that of light is twofold. The roughly planar impulse of the surrounding
dynamic field would continue as though the electron had not stopped. At the
same time, a spherical impulse would arise, centered on the arrested electron
and of essentially infinitesimal thickness. Thomson, PM, 45 (1898), 172-83.
Localized energy in spreading impulses 153
per unit time at a distance r from the point /, and at an angle 6 from
the acceleration axis as seen from point /, is given by
e2a2 sin2 6 (6.10)
\6n2c3r2 (1 - (v/c) cos Of
Here a is the magnitude of the acceleration and v is the final
velocity of the electron. The factor sin2 6 arises from the transverse
nature of the impulse; its amplitude falls to zero in the forward and
backward directions and presents a maximum at 900 to the axis.
This is true regardless of the time duration of the acceleration. The
modulating factor (1 (v/c) cos 9f produces the asymmetry in
the radiated energy. It arises from the classical interpretation of
wave intensity and explains the localization of y-ray energy in the
forward direction.
Localized energy in spreading impulses 155
EXPANDING-RING y-RAYS
The energy is radiated into a hollow cone that is symmetric about
the axis of electron acceleration. Integrating equation 6. io over the
acceleration time produces the total energy radiated into the cone
defined by the angle 9:
sin20 1
(6.11)
cos 6 [(I (v/c) cos 6)4
For values of v close to c, and for small values of 6,
62
=s 16 eyeg = 16/06 (6.12)
-v/c + 62/2)
Figure 6.6 is a plot of S(9) for two cases v = o.gc and v = 0.99c. In
the former case 5max lies at about 150, and in the latter about 50,
from the axis of acceleration. The energy never quite rises above
zero in the forward direction but is the more tightly constrained
156 Seeking an electrodynamic solution, 1907-1912
around the axis the closer the final velocity is to c.54 Corrections for
nonlinearity in the acceleration and for the relativistic increase in
electron mass cause only a small broadening of the cone.
The net result is a ring-shaped concentration of energy. The
impulse is still a temporally discontinuous event, so that the energy
always lies at the intersection of the hollow cone with the expand-
ing spherical shell. Thus it forms a ring, as in Figure 6.7, that
expands in proportion to distance as it travels in the same direction
as its parent ^-electron. "One can see from this," Sommerfeld
continued hopefully, "that for very great velocities the energy
emission is almost corpuscular."55
"It is quite remarkable," Sommerfeld said, "that on the basis of
the Maxwell theory, that is, pure undulatory theory, one is led
under certain conditions to consequences that approximate the
Newtonian emission theory, and that appear to have more similar-
ity to Newtonian light projectiles than to Huygens' spherical
waves."56 In fact, Smax grows in proportion to (1 v/c)~3; thus,
energy radiated into the cone of maximum effect resulting from
acceleration to 0.99c is io3 times greater, as well as being more
localized, than that due to the slower 0.9c electron.
Thus far, the calculation was entirely classical. But to compare
results to experiment, Sommerfeld needed data, as yet unavail-
able, on the precise nature of the acceleration. He wanted to obtain
an expression for Efi/Ey, the ratio of total ^-particle energy to that
which appears in the resulting y-pulse. Sommerfeld expressed the
54
Sommerfeld calculated the extremum values of 6 where S dropped to 0. iSmax.
For v/c = 0.9, these limits are 12.5 to 470, for v/c = 0.99, they are 4 to 160.
55
Sommerfeld, Munchen Sb, 41 (1911), 13. A misprint on p. 10,1.12 puts /? for y.
56
Ibid., 4.
Localized energy in spreading impulses 157
unknown acceleration in terms of the distance / over which it
occurs and obtained the result
Element
Jo
f
Thus, Sommerfeld reformulated the quantum condition in terms
(6.15)
Bragg had held fast to the neutral-pair hypothesis. It was partly the
completeness of Barkla's analysis for the softer x-rays that gradu-
ally pushed the domain of their discussion closer to that of the hard
x-rays, where Bragg held the advantage. When Rutherford and
Geiger showed in 1908 that the oparticle carries twice the charge
of the electron,77 Bragg cast aside the a-particle as the positive
constituent of a pair in favor of an undefined "positive electron."
In 1909 Bragg responded directly to Thomson's proposal that each
electron radiates into a single narrow cone. By then, overwhelming
evidence had accumulated for the forward scattering of secondary
products of x-rays and y-rays. Bragg claimed that even Thomson's
"energy bundles of very small volume" must interact with individ-
74
Campbell, PZ, 13(1912), 81-3.
75
Ibid, 83.
76
Stark, PZ, 13 (1912), 161-2. E. Meyer, PZ, 13 (1912), 253-4. Laby and
Burbidge, PRS, 86A (1912), 333-48. Burbidge, PRS, 89A (1913), 45-57.
77
Rutherford and Geiger, PRS, 81A (1908), 162-73. Bragg and Madsen, PM, 16
(1908X938.
164 Seeking an electrodynamic solution, 1907-1912
ual atoms to explain this behavior. "It seems hard," he sighed in
this, his last paper before returning from Australia to England, "to
understand the distinction between such bundles and entities
generally classed as material."78
But in 19io, following his removal from Adelaide to Leeds, a
variety of influences caused Bragg to recognize that his neutral-
pair hypothesis could not be the full answer to the problems of
radiation. First, he came into closer contact with the continental
literature. The debate in Germany between Stark and Sommerfeld
mirrored many of the issues still unresolved between himself and
Barkla. A review article on his position, prepared for both British
and German readers in 1910, illustrates the more reflective and less
dogmatic expression of his hypothesis, which he tailored for a
European audience. When a neutral pair is disrupted, the constitu-
ent electron proceeds on course; "the form of the entity may
change, y into /?, X into cathode ray, and so on," he said, "but there
is so little change in anything but form that practically we may
assume a continuity of [particle] existence."79
So too did the narrowing gap between x-rays and visible light
influence Bragg's position. In 1910 two independent tests, one by
Bragg's compatriot R. D. Kleeman, showed that forward scatter-
ing was also found in the photoelectric effect using ultraviolet
light.80 Bragg had never presumed to extend his hypothesis to
periodic light waves;findingthat they too demonstrated one of the
chief arguing points for neutral pairs caused him to step back for
reflection. Perhaps there was some truth in the growing conviction
of many physicists by 1911, British as well as German, that
Barkla's characteristic x-rays bridged the gap between periodic
light waves and impulse x-rays.
Bragg was also impressed by how far toward a particlelike
concentration of y-ray energy Sommerfeld had managed to come.
But Bragg felt that Sommerfeld had not gone far enough. Soon
after reading the y-pulse study, he wrote to Sommerfeld to com-
plain that the rings still dissipated their energy too rapidly.81 How
78
W. H. Bragg and Glasson, PM, 17 (1909), 855-64. Bragg gave a final talk on the
scattering of x-rays at the January 1909 meeting of the Australasian Associa-
tion for the Advancement of Science but no text remains, Report, 72 (1910),
113.
79
W. H. Bragg, PM, 20 (1910), 396.
80
Kleeman, Nature, 83 (1910), 339; PRS, 84A (1910), 92-9. Stuhlmann, Nature,
55(1910), 3 i i ; / W , 20(1910), 331-9. See Chapter 9.
81
W. H. Bragg to Sommerfeld, 17 May 1911. Steuwer, BJHS, 5 (1971), 271-2.
Localized energy in spreading impulses 165
could the ring-pulse pass on virtually its entire energy to an
electron after traversing macroscopic distances? But while he ar-
gued with Sommerfeld the dangers of continued attempts to make
the pulse hypothesis conform with the ionization data, Bragg
simultaneously hedged his bet. His basic point, he concluded, was
not the physical nature of x-rays but the claim that in ionization,
particlelike accumulations of energy are transferred to single elec-
trons. One to one. He could avoid the quaking ground of radiation
theory by separating the x-rays per se from the ionization event.
With his new colleague at Leeds, H. L. Porter, Bragg proposed at
the Royal Society that x-rays and y-rays do not ionize atoms
directly. One need only assume that a neutral pair disintegrates
before ionization can occur. It is the resulting high-speed electron
that strikes and ionizes the atom, they claimed.82
Just two weeks before, C. T. R. Wilson had reported to the Royal
Society on his most recent experiments to make visible the tracks
of ionizing particles by means of vapor condensation.83 Here was
the clear evidence that Bragg had long sought to demonstrate
unambiguously that x-rays ionize gases in precisely the same
manner as do particulate /?-rays. Wilson's photograph, reproduced
in Figure 6.9, shows only multiple short tracks; there is no sign of
the diffuse cloud that ionization by spreading impulses would
seem to require. Bragg made certain that the message got to
Sommerfeld. If the energy of the rays were not concentrated in
unchanging particlelike units, he claimed, exceedingly long pe-
riods would be necessary for sufficient energy to build up on a
single electron from a succession of impulses.84
More perplexed than ever by the ambiguous action of x-rays and
y-rays, Bragg returned to the dual nature he had first ascribed to
x-rays in 1907.85 He began to call for a theory of x-rays, and to a
lesser extent of y-rays, that would embrace both particle and wave
characteristics. The existence of two sorts of x-rays, the polarizable
but inhomogeneous rays from the x-ray tube and the unpolarized
characteristic rays from irradiated matter, reflected for him a
82
W. H. Bragg and Porter, PRS, 85A (1911), 349-65. W. H. Bragg, PM, 22
(1911), 222-3; 23 (1912), 647-50.
83
C. T. R. Wilson, PRS, 85A (1911), 285-8.
84
W. H. Bragg to Sommerfeld, 7 July 1911. Stuewer, BJHS, 5 (1971), 273.
85
W. H. Bragg, TPRRSSA, 31 (1907), 94-8. He had suggested that along with the
neutral pairs, impulses coexist in an x-ray beam. This was to answer Marx's
velocity measurements.
166 Seeking an electrodynamic solution, 1907-1912
hv = eAE cos cot [(C Bco) sin cot + Ccot cos cot] dt (7. lg)
Jo
The first two terms drop out when integrated over a whole number of oscilla-
tions, and because cos2 a = (1 + cos 2a)/2,
r . /I +cos2a)A ,
hv= eAEcot( \dt
To reconcile this result with Lorentz' formula (7.1), we need to express the
field strength AE in terms of Lorentz' /, which is an energy flux, / = AE2l^nc;
then substitute the classical radius of the electron Si = ie2l^m [M. Abraham,
AP, 10 (1902), 151]; and recognize that r = n/v = nk/c:
4hc2
r
Jo
(T- U)dt = h/2n
/
s
#*=
S
>
i y /
i/
/
J
// 1 / +-^
-*/y-
\
'
\
\
\
/ air
2
\\
\
/
/
r x
Figure 7.1. Graph used by Debye and Sommerfeld to derive Stokes' law
for the photoelectric effect. [AP, 41 (1913), 890.]
1909 10 11 12 13 14 15 17 18
light
1909 10 11 13 14 15 16 17 18
2
Friedrich, Knipping, and Laue, Miinchen Sb, 42 (1912), 303-22. Reprinted
with additional notes in AP, 41 (1913), 971-88.
3
Forman, AHES, 6 (1969), 38-71.
4
Sommerfeld, AP, 38 (1912), 473-506.
Origins ofx-ray spectroscopy 201
Fig. S
+ + '
CO " +
+ & oo
13
Laue, AP, 41 (1913), 989-1002.
14
Lorentz, Amsterdam Verslag, 21 (1913), 911-23.
Origins of x-ray spectroscopy 205
singly-directed
doubly-directed
primary beam
al
T~
OpTical pa^
A=
SB
ab - at sin
* 4 * cos
%*& \ - zosZB c ^ sin* a
Tkvs
A-
Degrees.
Figure 8.6. Bragg's evidence for multiple orders of x-ray spectra. [Bragg
and Bragg, PRS, 88A (1913), 413.]
the form and spacing of the peaks did not change appreciably when
different crystals were used to diffract the x-rays. The same pattern,
shown in Figure 8.6, was found at just slightly different angles for
the x-rays reflected from rock salt, iron pyrite, zincblende, potas-
sium ferrocyanide, potassium bichromate, calcite, quartz, and
sodium ammonium tartrate. At first, Bragg tried to discern a
harmonic progression in the angle of the peaks. The distribution of
the peaks was remarkably like that produced by visible light on one
side of an optical grating. In each case three peaks, A, B, and C,
were accompanied by three more at proportionately greater
angles. The distribution looked like multiple orders generated by
optical emission spectra. The new spectroscopy had been suitably
rewarded with x-ray emission lines.
Bragg measured the absorption coefficient for each of the three
line sets. He found that each set (e.g., C,, C2, C3) has a distinct
penetrability independent of the crystal. He had found "three sets
of homogeneous rays," and he concluded, "since the reflection
angle of each set of rays is so sharply defined, the waves must occur
in trains of great length. A succession of irregularly spaced pulses
could not give the observed effect."30 Once they had estimated the
spacing of atom planes in the crystals, the Braggs gave the first
unambiguous estimate of^the wavelength of a homogeneous x-ray.
30
W. H. and W. L. Bragg, PRS, 88A (1913), 436.
212 X-ray interference, 1912-1922
The value, 1.78 X io~8 cm, was considerably longer than Laue's
hypothetical wavelengths, but in line with Sommerfeld's pulse
width estimates.
By 1913, x-ray emission lines were not unexpected; Barkla had
spoken in 1911 offluorescentx-rays as a "line spectrum."31 But the
sharpness of the angular definition was unexpected. Even so, the
i/ioth degree angular resolution of the Bragg apparatus prevented
the full realization of homogeneity. The conflict between x-rays of
definite wavelength and the paradoxes of x-ray absorption was
demonstrated clearly enough to Bragg. "The problem remains,"
he said, "to discover how two hypotheses so different in appear-
ance can be so closely linked together."32 Bragg's call for a dualistic
theory of light grew louder.33 But no resolution appeared possible.
The Braggs saw in the new spectroscopy the means to a detailed
understanding of crystal structure. From mid-1913 on, their efforts
were directed primarily to this end. They did not continue their
investigation of the nature of x-rays, and for that reason their
otherwise commendable work ceases to be of immediate concern
to us.
CONTINENTAL REACTION TO
CRYSTAL INTERFERENCE
43
Moseley and Darwin, PM, 26 (1913), 228.
44
Moseley to his sister, [2 February 1913]. Heilbron, Moseley (1974), 200- 1
45
Bohr, PM, 26 (1913), 1-25.
46
Moseley and Darwin, PM, 26 (1913), 228.
216 X-ray interference, 1912-1922
share those unwavelike characteristics. Interest in x-ray research
grew by a factor of three from 1911 to 1913.47
Whatever the reservations Sommerfeld held before the discov-
ery of x-ray interference, they vanished afterward. He repeatedly
spoke on x-rays before general audiences in the following years,
using the Laue result as the central theme of his talk. In one of the
first of these talks, delivered in April 1913 to the German Society
for the Advancement of Instruction in Mathematics and Natural
Science, he explained that "impulses and waves are both deter-
mined by Maxwell's equations." He saw no conflict in the exis-
tence, side by side, of both impulses and periodic waves in x-rays.
It was nothing like the conflict "between the wave theory and
corpuscular theory of light."48 He also described Lorentz' concerns
about the form of the Bremsstrahlung pulses, a ratherfinepoint for
his audience. He was intensely interested in that analysis, and had
already begun an extension of Lorentz' work that he published two
years later. Although a full explanation of the origin of the periodic
fluorescent x-rays was still lacking, "one thing can be discerned,"
he said; "the theory is on the right track."49
Stark's response to crystal diffraction of x-rays was more
guarded. He had been carrying out experiments on the mechanical
effects he thought should be produced in crystals by impacts of a-
and canal-ray particles.50 He thought he had found evidence that
the particles are preferentially deflected along the "avenues" of the
crystal. In a paper published shortly after Laue's announcement,
Stark claimed to have come already to the conclusion that the
same behavior would be shown by the x-ray lightquanta. "It is
remarkable," he added gratuitously, "that I had come to these
views on the basis of the light-cell hypothesis even before I knew of
the experiments . . . taken as evidence for 'interference' of x-
rays."51
47
The Beibldtter zu den Annalen der Physik lists approximately twenty-five
papers on x-rays in each year from 1907 to 1910. The year 1911 has only twelve.
Thereafter, the numbers rise rapidly: 33, 51, 64, 61, and 63 in 1916.
48
Sommerfeld, Naturwissenschaften, 1 (1913), 708. Sommerfeld spoke fre-
quently on the new experiments on x-ray interference in the following years.
Munchener medizinische Wochenschrift, 62 (1915), 1424-30; 63 (1916),
458-60.
49
Sommerfeld, Naturwissenschaften, 1 (1913), 706.
50
Stark and Wendt, AP, 38 (1912), 941-57.
51
Stark, PZ, 75(1912X975.
Origins of x-ray spectroscopy 217
Stark's inflated claims did not lead him to attempt a reconcilia-
tion of lightquanta with interference once it became clear that
more was involved than channeling of particles. By mid-1913 he
accepted the Braggs' explanation of crystal diffraction.52 Subse-
quently, he renounced the lightquantum. Interference, he said,
demands that lightquanta extend along the direction of propaga-
tion, and thus are not localized. If, on the other hand, lightquanta
are localized from front to back, their spectral lines should be
conspicuously broadened in an electricfield.53Stark would win the
Nobel Prize in physics in 1919 for his experimental researches on
line broadening in electric fields. Yet in 1912 he could find no
spectral evidence that lightquanta are localized. Soon thereafter,
Stark, the first supporter of the lightquantum hypothesis after
Einstein, altogether dropped the nature of radiation as a topic of
research.
Laue answered Stark's early claims, and in the process showed
that the Bragg reflection law, nX = id sin #, is a special case of his
own method. It applies, he said, when the interfering rays are "so
spectrally inhomogeneous as to be incapable of interfering after a
difference in [optical path] of only a few wavelengths."54 He still
thought that the spots were due to a few homogeneous component
x-rays. The Braggs, he felt, were pursuing only a special case of his
more general analysis.
Laue concluded with the frank admission that the effect on the
interference maxima of the thermal agitation of lattice points was
"less than any expectation," but he was unable to explain why.
The lattice point vibration was treated in detail in a pair of papers
by Debye. Sommerfeld's former assistant had just finished the
joint paper on the photoelectric effect, discussed in Chapter 7.
Debye gave an influential talk on the quantum theory of specific
heats and heat conduction at the 1913 Wolfskehl lecture series in
Gottingen.55 The theory, introduced by Einstein in 1907, became
the first subject after cavity radiation to excite widespread interest
in the quantum theory.56 Debye focused on Planck's remarkable
prediction of a "zero-point" energy of atoms at absolute zero
52
Stark, PZ, 74(1913), 319-21.
53
Stark, VDpG, 76(1914), 304-6.
54
Laue, PZ, 74(1913X422.
55
Debye, Kinetische Theorie der Materie (1914), 17-60.
56
Klein, Science, 148 (1965), 173-80. Kuhn, Black-body theory (1978).
218 X-ray interference, 1912-1922
temperature. He turned to the thermal influence on crystal diffrac-
tion as a possible source of corroborating experimental data.
In July 1913 Debye concluded that thermal motion would have
little or no observable effect on the definition of the Laue spots.57
He calculated the expected decrease in intensity of the interference
spots located at increasing angles from the beam direction. This
relation depends on the effective wavelength of the x-rays and
allowed Debye to derive values for x-ray wavelengths indepen-
dently of the crystal parameters required by both Laue and
Bragg.58 The end result was a relation giving the decrease in spot
intensity for increasing order of the spectrum.59 The more com-
pressible the crystal, that is, the weaker the interatomic binding
forces, and the higher the temperature, the faster the intensity
decreases. In this way, Debye provided the means to reconcile
Laue's and the Braggs' interpretations of interference. If the higher
orders of spectra decrease quickly enough in intensity, the contin-
uous x-ray spectrum may be largely free of the superposed high
orders. Such a spectrum would therefore be a more accurate
reflection of the Fourier spectrum of the impulses that produce it
than would a spectrum including the higher orders.
In the fall of 1913, Laue renounced his appeal to monochro-
matic primary x-rays. At the Naturforscherversammlung in
Vienna, Walther Friedrich reported that "it is, for the first time,
experimentally demonstrated [that] the absorption coefficient for
x-rays is a measure of the wavelength."60 He specifically denied the
assertion that only a few wavelengths in the primary beam are
responsible for the interference spots. In the discussion that fol-
lowed, Laue protested that Friedrich had credited him with a
viewpoint that he had "never held," that the primary x-ray beam is
itself monochromatic. If he had not been entirely clear before, he
claimed, it was due to his "strong reservations" about the "only
possible explanation that he had seen" at the time.61 So, Laue
brought his own interpretation into line with that of the Braggs
while still retaining its generality. Indeed, he now thought that the
57
Debye, VDpG, 15 (1913), 678-89. For the state of this issue before Debye's
treatment, see Forman, AHES, 6 (1969), 38-71
58
Debye, VDpG, 75(1913), 738-52.
59
Ibid., 857-75.
60
Friedrich, PZ, 74(1913), 1081.
61
Ibid, 1085.
Origins of x-ray spectroscopy 219
primary beam consists of a "wide continuous spectrum." It was
particularly fortunate, he said, that the Braggs, Moseley, and
Darwin had found "isolated monochromatic x-rays" so that cor-
roboration of the basic laws of interference could be found.62
By 1915 Sommerfeld could speak "with complete confidence of
'x-ray light' distinguished from normal light by precise determina-
tion of its wavelength."63 As early as 1900, he had suggested the
division between periodic and aperiodic waves, two ends of a
continuum stretching from monochromatic light to the x-rays.
This idea was now widely applied to distinguish the two types of
x-rays. The characteristic x-ray lines are like monochromatic light;
Sommerfeld compared them to "distinctly colored spectral
lines."64 Like white light, the Bremsstrahlung impulses may be
thought of either as irregular pulses or as a superposition of Fourier
component waves. No longer as concerned as he had been in 1900
with keeping the two concepts separate, he concluded, "it is more a
question of ease and purpose which way of speaking we employ."65
Sommerfeld developed this idea of interchangeability between
impulse and continuous spectrum representations in a careful
study of the spectrum of Bremsstrahlung pulses in 1915. He ex-
tended Lorentz' argument against singly directed pulses using
Fourier spectrum analysis to do so, and showed inter alia that a
doubly directed square pulse has a spectrum that peaks at a
frequency approximately equal to nc/d, rather than at zero fre-
quency like the singly directed impulse.66 Moreover, it was clear
from the analysis that inherent in the very nature of the problem
was an unavoidable tradeoff between well-defined frequency and
well-defined pulse width. The more temporally restricted is the
impulse, the more component frequencies have to be added to
produce it, and the notion of a single frequency for the impulse
becomes less well defined. Conversely, the very sharp x-ray emis-
sion lines, which define virtually pure frequencies, have to be
thought of as periodic disturbances extending over macroscopic
distances in space. Just where in space the effect of the x-ray is
located then becomes unclear.
62
Laue, PZ, 74(1913), 1077.
63
Sommerfeld, Naturwissenschaften, 4 (1916), 1.
64
Ibid., 5.
65
Sommerfeld, Naturwissenschaften, 1 (1913), 708.
66
Sommerfeld, AP, 46(1915), 721-48.
220 X-ray interference, 1912-1922
It was not so much that physicists prior to this period were
unable to visualize x-ray impulses as a superposition of periodic
waves, but that there had been little advantage in doing so. The
temporal discontinuity of impulses had attracted attention to
them originally. Most physicists in thefirstdecade of the twentieth
century thought of x-rays as short temporal events rather than as
the net effect of many extended processes in the aether. After 1912
that view became dated, and the new physics of x-rays exploited
more and more the Fourier tradeoff between temporal discontin-
uity and frequency. Evidence for the interference of x-rays did not
convince physicists that x-rays are periodic waves; rather, it con-
vinced them that, whatever the true nature of radiation, x-rays act
in all respects just like ordinary light.
CHARACTERISTIC y-RAYS
Rutherford's hypothesis that y-rays are the electromagnetic im-
pulses produced by the expulsion of ^-electrons from the atom
remained purely qualitative until Sommerfeld's 1911 study. In this
view, one would expect to find a correlation between ^-particle
velocity and y-ray penetrating power. Although the latter property
could be measured, not until 1910 were methods developed to
determine with accuracy the velocities of released ^-electrons.
Consequently, there was little more than fragmentary evidence
before 1912 that y-rays possess properties characteristic of their
parent element. McClelland had found in 1904 that the y-rays from
radium are not homogeneous in penetrating power; their absorp-
tion does not follow the exponential law.67 A student of Ruther-
ford's in Montreal, Tadeusz Godlewski, noted the following year
that y-rays from different sources exhibit differing penetrability.68
Richard Kleeman, Bragg's former co-author, found at least two
distinct absorption coefficients combined in the y-rays from ra-
dium.69 In his experiments in 1904 Paschen had found two peaks in
the velocity distribution of the ^-particles emitted by radium,70
and Kleeman suggested in 1907 that these are intimately related to
67
McClelland, PM, 8 (1904), 67-77.
68
Godlewski, PM, 70(1905), 375-9.
69
Kleeman, PRS, 79A (1907), 220-33.
70
Paschen, AP, 14 (1904), 389-405; PZ, 5 (1904), 5O2"4-
Origins of x-ray spectroscopy 221
71
the two strong y-ray components he had found. Otto von Baeyer,
Otto Hahn, and Lise Meitner found more compelling evidence of
/?-particle velocity peaks in 19 io.72
There was little evidence, however, that the perceived groupings
of y-rays are characteristic of their radioactive source, even if the
/^-particle velocity peaks seem to be. At first, Rutherford was
reluctant to believe that the patterns in y-ray penetrating power
could be, as he later put it, "ascribed to a characteristic radiation
set up in the radiator by the primary y-rays."73 When D. C. H.
Florance measured the intensity distribution of the y-rays scattered
from lead in 1910, Rutherford interpreted the results as proof that
y-rays cannot be stimulated by external influences on matter.74 But
the following year, J. A. Gray demonstrated that y-rays - or what
appeared to be y-rays - could, like x-rays, be stimulated in matter
by incident /^-particles.75 This set to rest several vague doubts:
Some radioactive substances are known to emit a significantly
lower intensity of y-rays than would be expected were the y-rays a
product of /^-particle emission. Actinium is a case in point. Ruth-
erford thought that this might be due to "some peculiarity in the
mode of emission of thefirays," and his doubts about characteris-
tic y-rays began to fade.76
First, Rutherford set Moseley to work to verify that only one
^-particle is emitted by any single atom. Moseley showed that the
elements called radium B and radium C emit only one ^-electron
when an atom decays. In the meantime, Gray found what at the
time seemed convincing evidence that y-rays can stimulate sec-
ondary y-rays in matter,77 although it is clear in retrospect that
these were high-frequency characteristic x-rays. Initially, Ruther-
ford doubted Gray's claim that these secondary "y-rays" are char-
71
Eve, PM, 11 (1906), 586-95. Kleeman, PM, 14 (1907), 618-44; 75 (1908),
638-63.
72
von Baeyer and Hahn, PZ, 11 (1910), 488-93. von Baeyer, Hahn, and
Meitner, PZ, 72(1911), 273-9.
73
Rutherford, Radioactive substances (1913), 282. Although it had very little
influence on his research, Rutherford came very close to adopting Thomson's
unit hypothesis for x-rays in 1913, p. 84.
74
Florance, PM, 20 (1910), 921-38.
75
Gray, PRS, 85A (1911), 131-9. Important preparatory work was done by
Starke, Le Radium, 5 (1908), 35-41; and by Davisson, PR, 28 (1909), 469-70.
76
Rutherford, Radioactive substances (1913), 271.
77
Gray,/W, 26(1913), 611-23; T^S, 56^ (1912), 513-29;^7^(1912),489-501.
222 X-ray interference, 1912-1922
acteristic of the element from which they arise. But in 1912 he
began to take seriously evidence that the groups of y-rays of
homogeneous penetrating power correspond to the peaks in /?-
electron velocity. At about the same time that the news of crystal
diffraction of x-rays came from Munich, Rutherford concluded
that "the transformation of energy from the /? ray form to the y ray
form or vice versa takes place in definite units which are character-
istic for a given ring of electrons."78
Rutherford now demanded the evidence for characteristic y-rays
that he had denied before. He set about to find it himself, and he
soon felt he had succeeded. Frederick Soddy, with his wife and
Alexander Russell, had shown in a series of experiments between
1909 and 1911 that, aside from endpoint behavior, y-rays are
absorbed exponentially along most of their path through lead.79
This is the classical symptom of homogeneity. With a student,
Rutherford measured absorption coefficients for y-rays spontane-
ously emitted from several elements, finding that radium B gives
three "distinct types" and radium C a single type.80 The coeffi-
cients differed markedly; in units of inverse centimeters of alumi-
num, they are shown in the penultimate column of Figure 8.7, a
table from a slightly later publication.81 In a burst of overconfi-
dence, Rutherford declared that the y-rays of coefficient 45 from
radium D were, with "little doubt," the "characteristic radiation of
the L type to be expected from an element of atomic weight 2 io."82
By one of those misapprehensions that so enrich and color the
history of experimental science, Rutherford was here quite misled.
The electromagnetic spectrum of a radioactive element consists of
more than just the y-rays characteristic of it. The exiting y-rays
and other energetic emissions stimulate the surrounding electronic
structure of the atom, releasing, in addition, the characteristic
x-ray spectrum of the element. It was these high-frequency charac-
teristic x-rays that Rutherford studied, calling them "y-rays" all the
while. We commit a historical anachronism to explain this point
here, and certainly no discredit is to reflect on Rutherford. Indeed,
78
Rutherford, PM, 24 (1912), 453-62.
79
F. and W. Soddy and Russell, PM, 19 (1910), 725-57. Russell, PRS, 86A
(1911X240-53.
80
Rutherford and H. Richardson, PM, 25 (1913), 722-34.
81
Rutherford and H. Richardson, PM, 26 (1913), 937-44.
82
Ibid., 331. Richardson soon used the y-ray absorption coefficients to fix the
atomic weight of uranium at 230. PM, 27 (1914), 252-6.
Origins ofx-ray spectroscopy 223
y Rai/$ of the Thorium and Actinium Products.
INTERFERENCE OF y-RAYS
After the discovery that x-rays can interfere, many physicists
thought that y-rays ought to show interference effects too. Moseley
started his x-ray investigations with this idea in mind. Walther
Friedrich gave two talks to the Wednesday Munich Colloquium
early in 1913 on interference of x-rays and then turned directly to a
review of analogous y-ray research.84 It was Rutherford who actu-
ally put x-ray diffraction apparatus to work and showed that
"y-rays" can show interference effects. His results, obtained with
Edward Andrade, were announced in the May 1914 issue of the
Philosophical magazine*5
Exceedingly long exposure times were required to capture the
penetrating "y-rays" on photographic plates. To make the diffrac-
tion angles as large as possible, the rays from radiums B and C were
reflected at glancing angles from rocksalt crystals. This decreased
the effective grating spacing. The y-ray source was placed in a
strong magnetic field to deflect all ^-particles and secondary elec-
trons. As a final precaution, Rutherford used Maurice de Broglie's
technique of rotating the analyzing crystal to overcome its inher-
ent irregularities; he worried about "contorted and undulating
surfaces." A very slow scan by the crystal accentuates the lines of
interest by smearing out ghost lines produced by periodic irregu-
larities in the atom planes.86
83
Rutherford and Andrade, PM, 27 (1914), 855.
84
The Register of the Miinchener physikalisches Mittwochscolloquium, 1 July
1912, 19 February 1913, 7 May 1913. AHQP, 20.
85
Rutherford and Andrade, PM, 27 (1914), 854-68. A brief first announcement
was made in Nature, 92 (1913), 267.
86
M. de Broglie, CR, 157 (1913), 924-6. See Heilbron, Isis, 58 (1967), 482-4.
Origins of x-ray spectroscopy 225
KADIUM B. PLATINUM.
low v high v
92
Bohr to Kossel, 19 July 1921. Bohr archive; AHQP BSC 4, 2.
93
His inspiration occurred between September and October 1914. VDpG, 16
(1914), 898-909, 953-63. Heilbron, his, 58 (1969), 451-85.
228 X-ray interference, 1912-1922
incident electron beams for his dissertation under Lenard at Hei-
delberg.94
With Kossel's suggestion that characteristic x-ray lines originate
in electron transitions to the innermost rings of the Bohr atom,
x-rays gained theoretical credentials equivalent to those of visible
light. Both were, in the context of the Bohr theory, the product of
downward electron transitions: visible light to high-lying electron
rings, x-rays to low-lying ones. Sommerfeld's subsequent elabora-
tion of the Kossel scheme further established the validity of the
hypothesis. Theoretical and empirical evidence had converged,
justifying further the equivalence of x-rays and light.
In the first months of 1915, evidence mounted that Kossel was
right. Ionization of the atom must precede the emission of charac-
teristic x-rays. Transitions to refill the lower rings produce the
characteristic lines. As Heilbron has indicated, this mechanism
would not likely have suggested itself to Bohr, who had already
rejected ionization as a prerequisite for optical emission. In Kos-
sel's hypothesis, the absorption edges directly measure the energy
levels in the atom. Differences in absorption edges allow calcula-
tion of the fine splitting of the main quantum levels, and this soon
became the chief empirical means of refining the Bohr theory.95 It
also enabled Sommerfeld and others to make predictions about
the number of electrons that shield the nucleus at each level of the
atom. The main significance of the new understanding of x-rays
was that it provided a means to probe more deeply into the
structure of the atom, not to stimulate interest in the associated
x-ray photoelectric effect. It was not until 1921 that the latter
problem was seriously examined, as we shall see in a later chapter.
In the meantime, the success of the Bohr theory attracted away the
talents ofjust those physicists who were most capable of producing
a theoretical solution to the problems of free radiation.
3
Innes, PRS, 79A (1907), 442-62. Cooksey, AJS, 24 (1907), 285-304. Laub, AP,
26 (1908), 712-26.
4
Seitz, PZ, 77(1910), 707.
5
Stuhlman, Nature, 83 (1910), 311; PM, 20 (1910), 331-9; 22 (1911), 854-64.
Rubens and Ladenburg found in 1907-before the Bragg-Barkla debate clari-
fied its importance-that a difference of 1:100 exists between the photocurrent
stimulated in the backward and forward directions from a gold foil i o 5 cm
thick. VDpG, 9(1907), 749-52.
6
Kleeman, Nature, 83 (1910), 339; PRS, 84A (1910), 92-9.
7
Stuhlman, PM, 20 (1910), 339.
8
Cooksey, PM, 24 (1912), 37-45.
236 X-ray interference, 1912-1922
light. And the number of experimental studies on the photoefFect
peaked in 1912, as can be seen in Figure 7.2.
As discussed briefly in Chapter 7, the empirical understanding of
the precise form of the photoelectric relation was lacking even as
late as 1913. First, it was difficult to ensure that photocathode
materials used in tests were free of impurities. The mose sensitive
cathods are the alkali metals with loosely bound outer electrons.
These substances are subject to rapid oxidation, and emission
from the oxidized products clouded the results of many experi-
ments. Second, one had to be careful to separate the effects of the
normal photoefFect from the selective photoefFect discovered in
1910.
Two other influences were responsible for the strong experimen-
tal interest in the photoefFect around 1912. Of greatest significance
was Lenard's renunciation of the triggering hypothesis the year
before, an event that removed the basic presupposition that un-
derlay most previous serious studies of the phenomenon. Second
was the discovery of x-ray crystal diffraction, a result that fully
identified x-rays with ordinary light. It consequently raised for
visible light all the paradoxes of x-rays and y-rays.
One of the most significant experiments on the photoefFect in
this peak year 1912 was done by Owen Richardson and Karl
Compton at Princeton. Richardson's interest in the photoefFect
had been partly stimulated by his close ties to J. J. Thomson. He
had been a student in Trinity College, had taken a first in the
natural sciences tripos, and had done research under Thomson.9 In
1906 he had gone to Princeton University as a faculty member, and
it was there that he began research on the photoefFect with his
former student, Karl Compton. Dissatisfied with previous failures
to remove impurities, they directed single wavelengths of ultravio-
let light onto photocathodes that had been carefully scraped clean
and then kept in a vacuum.10 Their results appeared to substantiate
Einstein's relation - that light frequency is proportional to the
square of electron velocity - a result that Richardson later admit-
ted he had not expected to find.11 Criticism by Robert Pohl and
9
Swenson, DSB, 11 (1975), 419-23.
10
O. Richardson and K. Compton, PM, 24 (1912), 575-94. Hughes, PTRS,
212A (1913), 209-26.
11
O. Richardson, PRS, 94A (1918), 269.
Quantum transformation experiments 237
Peter Pringsheim in 1913 brought the explicit form of the relation-
ship into question again, but it was thereafter clear that no reso-
nance peaks were to be found.12 Richardson explicitly pointed out
the similarities between x-rays and light, reflected on the paradoxes
that stood in the way of a consistent theory of x-rays, and rehearsed
the data for asymmetrical photoemission. The last, he remarked,
might lead one "to the conclusion that [like x-rays, light]
. . . consists of a shower of material particles."13
To avoid this unpalatable conclusion, Richardson offered a
treatment of the photoeffect "from a rather wider standpoint"
than either lightquanta or waves, "without making any definite
hypothesis about the structure of the radiation."14 Expressed in its
fullest form in 1914, Richardson's derivation of the photoelectric
relation rested on no assumptions about the nature of light or
about the structure of the atom. It was a purely thermodynamic
argument that bypassed completely the issues that led to the
double paradox of quality and quantity. Photoemission of elec-
trons was considered to be analogous to the evaporation of a gas.15
Richardson assumed the Planck law for the energy distribution in
the radiation, and also assumed that the number of electrons
emitted by the light is proportional to its intensity. The result
predicted a sharply defined wavelength threshold for emission,
analogous to a latent heat of evaporation from the metal of the
cathode. Moreover, he showed that the maximum energy of an
electron has to be proportional to hv C, where C, like the
photoelectric work function, is a constant characteristic of the
metal.
But Richardson's interest in the photoelectric effect was largely
motivated by concerns other than the nature of light. An under-
standing of the photoeffect was important to his study of the
12
Pohl and Pringsheim, VDpG, 15 (1913), 637-44.
13
O. Richardson, PM, 25 (1913), 145.
14
Ibid.
15
O. Richardson, PM, 27 (1914), 476-88. Richardson's interest in the photoef-
fect grew out of his research on the "glow discharge," the emission of electrons
from hot metals. The earliest statement of the method was in PM, 23 (1912),
594-627, in particular 615ff. See also PM, 24 (1912), 570-4; a note appended
in proof to the last paper appeared separately in Science, 36 (1912), 57-8; PR,
34 (1912), 146-9. Richardson's theory deserves closer scrutiny than it has
received. For a beginning, see Stuewer, Minnesota studies in philosophy of
science, 5, (1970), 246-63.
238 X-ray interference, 1912-1922
electron discharge from a heated cathode.16 The "glow discharge"
soon became his central research subject, and in 1916 he sought an
explanation of it in terms of a photoeffect stimulated by the light
emitted from the hot cathode.17 Although Richardson's photoelec-
tric research was restricted to purely phenomenological tests that
contributed precise data on a series of metals not analyzed before,
he continued to watch the developing evidence about the photoef-
fect carefully.18 Having formulated his own non-lightquantum
hypothesis, he was confident that a resolution would be found for
the problems of the absorption of radiation that would avoid the
"restricted and doubtful hypothesis used by Einstein."19
1.2 Conta< :t
Potem ial
Correc tion
.8
.4 y*
intensit
Thresh Did Frejuency = 57.0 x10 1 3 Hz Di ect Ob servatic
ofTh resholc Freque ncy
.4
1.2
1
55x10 13 60 65 70 75 80 85 90 95 100 55 59 63 ;10 1 3
Frequency (Hz)
Figure 9.1. Millikan's data for lithium corroborating Einstein's photoelectric law. [Courtesy of the Exploratorium, San
Francisco. From Millikan, PR, 7(1916), 377.]
Quantum transformation experiments 241
cathode and detector, shifting it upward to the dashed line. The
intercept of the corrected relation accurately marked the threshold
frequency for emission, measured separately and shown in the
inset on the right side of Figure 9.I.27 Closer corroboration of
Einstein's linear law, equation 5.1, could not be desired. To my
knowledge, Millikan's results were never questioned, and were
quickly recognized by leading European physicists to be definitive.
But acceptance of the Einstein law of the photoelectric effect did
not carry with it acceptance of the hypothetical lightquantum.
British French
German Italian
1900 01 02 03 04 05 06 07 08 09
Despite the equivalent number of British and German studies, the experimen-
tal work that was cited in this period was almost entirely British.
242 X-ray interference, 1912-1922
power and the atomic weight of the anticathode.29 The early
experiments were done without the benefit of Barkla's sharply
defined homogeneous x-rays. In 1910 Charles Sadler, who had
collaborated with Barkla in the discovery of characteristic second-
ary x-rays, found that the beam of secondary electrons emitted
from metals irradiated by a single fluorescent x-ray "line" is itself
always absorbed exponentially.30 This implied that the released
electrons were all of similar velocity. Sadler found that the velocity
was independent of both the intensity of the incident x-ray and the
nature of the tertiary radiating substance. He discovered, however,
that the electron velocity increases linearly with the atomic weight
of the secondary radiating substance that produced the homogen-
eous x-rays. Beatty found at roughly the same time that the
absorption of secondary electrons varies linearly with the absorp-
tion of the various homogeneous x-rays he used to stimulate
release of the electrons.31 But the most important relationship that
tended to confirm the close connection between the characteristic
secondary x-rays and the characteristic optical emission spectra of
chemical elements was discovered by Richard Whiddington in
1911.
Whiddington studied the secondary x-rays from chromium and
other elements at the Cavendish Laboratory in Cambridge. The
rays obediently appeared, with their characteristic coefficient of
absorption of 136 cm"1 in aluminum, only when the potential of
the discharge tube exceeded a certain threshold. When Whidding-
ton substituted various materials as anticathodes using Kaye's
techniques, he found that there was a definite order in the pene-
trating power of the emitted secondary x-rays. Thisfindingwas not
unexpected. Barkla had shown enough to suggest that such rela-
tionships should hold. But Whiddington thought that he had
found a possible violation of Stokes' law for the fluorescent nature
of secondary x-rays. "There is reason to believe," he hinted, "that
this secondary radiation can be extracted by a primary radiation
somewhat less penetrating than itself."32 Thus, it was extremely
important for him to refine his apparatus so that the electron
energy required to stimulate a given x-ray line could be measured
accurately.
29
Kaye, PTRS, 209A (1908), 123-51.
30
Sadler, PM, 19 (1910), 337-56.
31
Beatty, PCPS, 15 (1910), 416-22; PM, 20 (1910), 320-30.
32
Whiddington, PCPS, 15 (1910), 575; PRS, 85A (1911), 99-118.
Quantum transformation experiments 243
T O ELECTROSCOPE
Velocity, v
Element (108 cm/sec) A A/v
Al 20.6 27 1.3
Cr 50.9 52.5 1.03
Fe 58.3 56 0.96
Mn 61.7 58.6 0.95
Cu 62.6 63.2 1.01
Zn 63.6 65.1 1.02
Sc 73.8 78.9 1.07
mark the straight line through the slit from the radioactive wire
used as the source. The curved paths followed by /^-electrons in the
uniform magnetic field depend on the velocity of the electrons.
Since different-velocity electrons would pass through the slit on
orbits of different radii of curvature, the bands in the photographic
record indicated that there were distinct peaks in the velocity
distribution.
Once the technique was known, modifications were instituted
that led to the recognition of a remarkable focusing effect. In 1911
von Baeyer, Hahn, and Lise Meitner turned the photoplate so that
it lay in the plane of /^-particle motion and the entire course of the
electrons was traced, in much the same manner as Villard had
done in the experiments in which he discovered the y-rays.40 Jean
Danysz, working in the laboratory of Marie Curie, placed the
plate so as to take advantage of a full 1800 deflection of the elec-
trons.41 This arrangement was soon recognized to have a distinct
40
von Baeyer, Hahn, and Meitner, PZ, 12 (1911), 273-9, 378-9; 13 (1912),
264-6.
41
Danysz, CR, 753 (1911), 339-41, 1066-8; Le Radium, 9(1912), 1-5.
Quantum transformation experiments 247
JUL
F0 = - f (9.2)
e ^
where Vo is the critical potential and Xo the corresponding x-ray
wavelength. Equation 9.2 soon became known in America as the
Duane-Hunt law, and was recognized for many years to be the
most accurate of all means to measure h.55
The Coolidge tube was also important to the research group at
Manchester that carried on the x-ray studies begun by Henry
Moseley. Charles Darwin, Moseley's collaborator, had hoped that
monoenergetic electron beams might be isolated using magnetic
deflection. One could then use them to produce x-rays and see if
the wavelengths, measured by absorption and diffraction, corre-
spond to the quantum law. His early experiments were prevented
by the variation in x-ray output due to the unsteady vacuum in the
experimental tube. The work was soon dropped altogether when
he, like Moseley, left Manchester for the war.
In 1915 Rutherford resumed the work for a short while using a
54
Duane and Hunt, PR, 6 (1915), first 166-71. Webster argued that finding a
threshold would cast doubt on any impulse representation of Bremsstrahlung
because the Fourier spectrum of an impulse has no high-frequency limit. This
is true in principle but, as Figure 2.7 illustrates, the amplitude of the high-
frequency components in the pulse drop rapidly toward zero. For any practical
experiment there is an effective threshold. Webster, PR, 6 (1915), 56.
55
Birge, PR Supplement, 1 (1929), 57.
252 X-ray interference, 1912-1922
tottage in thousands.
new Coolidge x-ray tube. With the few students still in the labora-
tory, he tried to "see how far [the quantum] relation holds for the
excitation of x-rays by electrons."56 Up to 115,000 V the tube was
supplied with electricity from a motor-driven Wimshurst genera-
tor; from there up to 175,000 V they used an induction coil capable
of producing a 20-in. spark. Because the tube produced x-rays of
unprecedented hardness, the effective wavelengths were very small
and the angles required by crystal diffraction were correspondingly
minute. To estimate the wavelengths, they were forced to use the
imprecise relation of absorption coefficient to wavelength.
The results of the survey are shown as Figure 9.6. The calculated
frequency of what Rutherford termed the end radiation is essen-
tially the highest-frequency x-ray produced by the indicated volt-
age. They found systematic deviation from the quantum relation
at high potentials; rather than increasing in proportion to the
applied potential, the end frequency reached a maximum value at
about 142 kV. Rather than following the Planck relation hy = E,
the data conformed to hv = E kE2, where k is a constant.
Rutherford realized that these data were only approximate. By this
56
Rutherford, J. Barnes, and Robinson, PM, 30(1915), 352.
Quantum transformation experiments 253
time, he was convinced that the Bohr-Kossel model of x-ray
production could explain the systematic deviation as energy lost to
the electron as it penetrates to the more tightly bound orbits within
the atom. But in 1915, knowledge of x-ray absorption edge energies
was insufficient to allow evaluation of further tests, and Ruther-
ford, pressed by wartime work, did not pursue any.
Rutherford's primary reason for undertaking this study had
little to do with x-rays; he still hoped to find evidence in support of
his hypothesis that characteristic y-rays originate in the electronic
structure of the atom. He did find some x-ray frequencies at high
potentials that were shorter than those estimated from absorption
coefficients of what he still thought were selected y-ray lines. But
the saturation in frequency made it clear that a definite limit
prevented the production of x-rays as penetrating as true hard
y-rays such as those from radium C. He wrote only one more paper
on x-rays, and then after the war moved back into experiments
entirely on radioactivity. The most influential voice in Britain on
x-rays continued to be Charles Barkla's.
Kedge
low v high v
CORPUSCULAR SPECTRA
Maurice de Broglie was the eldest son in a family traditionally
prominent in French politics. In 1906 he assumed the hereditary
title "Due de Broglie," and at about the same time he outfitted a
private laboratory in his Parisian residence.5 Although educated in
physics at the College de France, he was a scientific amateur, a
species common in the nineteenth century but increasingly rare in
the twentieth. In 1908, when he defended his theses following work
with Paul Langevin, his research concerned ionization of gases and
Brownian motion.6 He soon took an interest, then uncommon for
a Frenchman, in the new quantum theory. He was secretary to the
First Solvay Congress in 1911 and, with Langevin, edited its pro-
ceedings.7 With the discovery of x-ray interference in 1912, he
reorganized his research plans. The following year he made a
notable discovery: The slow rotation of a crystal greatly reduces
the spurious x-ray interference maxima caused by periodic imper-
fections in the crystal lattice.8 The rotation smooths out what
Moseley had called the "horrid diffraction fringes" caused by
natural ruling errors in the crystal, which greatly confused early
crystallographic analysis.9 The new technique was an important
step forward and was, for example, indispensable to Rutherford in
his 1914 measurement of what he thought were y-ray wavelengths.
De Broglie did some of the most significant experimental re-
4
M. de Broglie, Scientia, 27 (1920), 110-11.
5
Weill-Brunschvicg and Heilbron, DSB, 2 (1973), 487-9. La Varende, Les
Broglie (1950), 265-320. The laboratory was first established in de Broglie's
apartments in rue Chateaubriand; later, another was set up in rue Lord Byron a
block away in the 8th arrondissement.
6
M. de Broglie, Theses (1908).
7
Solvay I.
8
M. de Broglie, CR, 157(1913), 924-6. See also Heilbron, Ms, 5^(1967), 482-4.
9
Moseley to his sister, 13 August [1913]. Heilbron, Moseley (1974), 207-8.
Synthesis of matter and light 265
search on x-ray absorption edges. In 1914, on the suggestion of
Bragg and Siegbahn, he successfully interpreted anomalous ab-
sorption bands recorded on a photoplate for x-rays from silver as if
they were due to preferential absorption of the x-rays by the silver
halide emulsion.10 There soon developed a lively race among
Siegbahn, Wagner, and de Broglie to measure the precise frequen-
cies of the absorption edges. Wagner found that there were two L
edges.11 In confirming this result in 1916, de Broglie found a
third.12 At war's end he dedicated his research to x-ray absorption;
the frequencies of the edges gave the most direct measurement of
the energy levels of the Bohr atom. The discovery of three L edges,
five M edges, and so on directly shaped ideas about the electronic
structure of the atom as interpreted by Bohr and Sommerfeld. De
Broglie had entered a front-rank research field and possessed the
skill and facilities to make significant contributions.
By January 1921, de Broglie realized the potential value of /?-ray
spectroscopy to his research. He had noted in 1920 that x-rays
stimulate secondary ^-electrons,13 but now recognized that this
held the key to a study of energy transformation in the absorption
of x-rays. His inspiration likely had much to do with Alexandre
Dauvillier, who joined de Broglie's research group late in 1920.
Dauvillier's research had already moved from studies of x-ray
absorption to an examination of the kinetic energies of the second-
ary electrons released by x-rays.14 He began to investigate the
chemical effects of electrons on photoemulsions and carefully
studied the work by Barkla in which the electron range was used to
fix electron energy. Early in 1921, Dauvillier completed the most
thorough review in the French language on the photoelectric
effect, with the explicit intention to apply that knowledge to the
case of x-rays. "The fastest electron emitted by a body exposed to
[x-rays] possesses precisely the same velocity as the fastest electron
in the cathode beam that produces the x-rays," he claimed.15
Dauvillier was also perplexed that the energy granted an electron
by x-rays is independent of the distance over which the rays have
10
M. de Broglie, CR, 158 (1914), 1493-5-
11
Wagner, AP, 46 (1915), 868-92.
12
M. de Broglie, CR, 163 (1916), 352-5.
13
M. de Broglie, SFP PVRC, 1920 (1921), 33-
14
Dauvillier, Theses (1920).
15
Ledoux-Lebard and Dauvillier, Physique des rayons x (1921), 208.
266 Conceptual origins of wave-particle dualism, 1921 -1925
traveled.16 "The electromagnetic and corpuscular theories seem to
be very difficult to reconcile," he said. "The problem of the nature
of radiation remains unresolved."17
In January 1921 a South African physicist, encouraged by Owen
Richardson, published a study that further directed the Parisian
physicists' attention to these problems.18 Lewis Simons, like Rich-
ardson, was convinced that Barkla's antiquantum claims were
simply wrong: A single homogeneous x-ray beam gives rise to an
entire spectrum of electron velocities, and different amounts of
energy are lost to each electron depending on its original position
in the atom. Simons worked hard, but unsuccessfully, to extract
convincing evidence for this claim from ionization and electron-
range data. About ^-ray spectroscopy, the suggestive early work by
Robinson and Rawlinson, and the incomplete experiments by Hu,
Simons had apparently not heard.
Maurice de Broglie had the great advantage of knowing all of
these prior experiments. But he was skeptical of the claim that
x-ray energy is always absorbed by electrons in integral multiples
hv. When hefirsttook up the detailed evaluation of electron energy
in February 1921, he thought he had found evidence that electrons
leave the atom with quanta of energy hv and therefore must have
absorbed more than that quantized amount from the x-rays.19 His
studies of x-ray absorption edges and his understanding of the
Bohr atom made it clear that some energy must be spent to remove
each electron from the atom. This seemed a significant issue to
clarify for presentation at the Third Solvay Congress, to which he
had been invited and which would meet in two months' time.
Besides, a recent disappointment in his attempt to enter the Paris
Academy of Sciences made a certain broadening of his experimen-
tal inquiry attractive.20
And so, Maurice de Broglie, intrigued by the paradox that
16
Ibid., 399.
17
Ibid., 407-8.
18
Simons, PM, 41 (1921), 120-40. De Broglie knew of Simons' paper, see Solvay
III, p. 85n.
19
M. de Broglie, SFP PVRC, 1921 (1922), 10.
20
On 22 November 1920, Maurice de Broglie was nominated to replace Adolphe
Carnot in the Academie des Sciences. On the 29th, after four votes, Jules Louis
Breton was elected, de Broglie having received only one of about sixty votes
cast in each round. CR, 171 (1921), 1031, 1045. One of his biographers reports
that Maurice de Broglie was later elected on the basis of his studies on the
photoeffect. Sudre, Revue des deux mondes (July-Aug. i960), 577-82.
Synthesis of matter and light 267
23
M. de Broglie to Lindemann, 12 April 1921. Cherwell papers.
24
M. de Broglie, CR, 172 (1921), 275.
25
Ibid., 806-7.
Synthesis of matter and light 269
Element
Ba W Pt Pb U
Figure 10.3. Debye's figure relating x-ray energy and direction to scat-
tered electron energy and direction. [PZ, 24 (1923), 164.]
(10.2)
/2
On the other hand, the most directly comprehensible consequence
of the theory of special relativity is that a moving clock appears to a
stationary observer to be running more slowly than it does to an
observer moving along with it. According to relativity theory, the
observed frequency of the particle oscillator is
(10.3)
h Vl-(tyfc) 2
where n is an integer. This may be rearranged to the form
T
=nh (10.9)
VI ~{vplcf
Equation 10.9 is precisely the action-integral expression for al-
lowed electron orbits in the Bohr atom, a mathematical condition
that had lacked, until now, any physical interpretation.107
Many years later, de Broglie described himself as always having
had "a very 'realist' conception of the nature of the physical world
and little given to purely abstract considerations."108 He had found
in 1923 what had eluded Bohr, a realistic conceptual explanation
for the discrete orbits in the Bohr atom, and it obviously impressed
him mightily. "It is almost necessary to suppose," de Broglie said
at the time, "that the electron orbit is only stable if the fictional
wave reencounters the electron in phase."109 And the following
year he repeated, "This beautiful result . . . is the best justifica-
tion that we can give for our way of addressing the problem of
quanta."110
115
La Varende, Les Broglie (1950), 333. See also the comments of another
member of de BroghVs thesis committee, Charles Mauguin, in Louis de
Broglie (1953), 434-
116
Langevin to L. de Broglie, 27 July 1924. "Autographes," L. de Broglie dossier.
117
Langevin to L. de Broglie, 13 January [1925]. Ibid. (The original is quoted in
German.)
118
Lande, PZ, 24 (1923), 441-4.
119
Forman, his, 59 (1968), 156-74; HSPS, 2 (1970), 153-261.
120
Cassidy, HSPS, 10 (1979), 187-224.
121
Serwer, HSPS, 8(1911), 189-256.
122
Kronig in Theoretical physics (i960), 5-39.
298 Conceptual origins of wave-particle dualism, 1921 -1925
123
theory. Bohr found the Compton effect a particularly intransi-
gent obstacle. It seemed to prove that individual interactions of
x-rays with electrons conserve energy and momentum. Even
Bohr's daring statistical interpretation of energy conservation now
seemed insufficient to circumvent the lightquantum. The story
need not be repeated here how the American theorist, John Slater,
approached Bohr at this time with the idea that lightquanta are real
and that their emission and absorption by atoms are dictated by a
"virtual radiation field."124 The resulting development of the
Bohr-Kramers-Slater theory of atomic "virtual oscillators" reaf-
firmed in a transitional form Bohr's statistical interpretation of
radiant interaction.125 Although experimental tests of major pre-
dictions of the BKS theory by Bothe and Geiger failed to justify the
idea, by the time the experimental results were known, the issues
had become superfluous.126
The basic ideas inherent in the statistical approach of the BKS
theory were used by Bohr's assistant, Hendrik Kramers, later
joined by Werner Heisenberg, to treat the thorny problem of the
dispersion of light by the Bohr atom.127 This successful work
strongly influenced the young Heisenberg in 1925 when he found a
mathematical formalism that solved most of the pressing problems
that the Bohr theory faced.128 Heisenberg's matrix mechanics,
although directed to the problems of atomic theory, was stimu-
lated in essential ways by the renewed appreciation of Einstein's
lightquantum hypothesis.
In the meantime, a separate line of development based on de
Broglie's hypothesis of phase waves led to another means of ana-
lyzing the difficulties of atomic theory. This followed the route
from de Broglie through Einstein to Erwin Schrodinger. The
resurgence of interest since 1921 in Einstein's lightquantum had
led him to try to restate its inherent flaws. He had known since
1909 that lightquanta alone were impotent to resolve the difficul-
ties that faced radiation theory - hence his plea to Ehrenfest that
123
Bohr, AP, 71 (1923), 228-88.
124
Klein, HSPS, 2 (1970), 1-39. Slater in Wave mechanics (1973), 19-25.
125
Bohr, Kramers, and Slater, PM, 47(1924), 785-802.
126
Bothe and Geiger, ZP, 26 (1924), 44; ZP, 32 (1925), 639-63. See also
A. Compton and Simon, PR, 25 (1925), 306-13; 26 (1925), 289-99.
127
Kramers, Nature, 773 (1924), 673-4. Heisenberg and Kramers, ZP, 31 (1925),
681-708.
128
Heisenberg, ZP, 33 (1925), 879-93.
Synthesis of matter and light 299
129
the problem was sufficient to drive him to the madhouse. In-
stead he developed a concept of a Gespensterfeld, or "ghost field,"
that might itself carry little or no energy, yet guide the trajectories
of lightquanta in accord with laws of wave optics.130 Little came of
the concept, but it bore sufficient similarity to de Broglie's ideas so
that Einstein was quick to recognize what was significant in the
heterodox suggestion from France.
When Einstein encountered de Broglie's thesis, he was just
developing a new statistical approach to the kinetic theory of an
ideal gas using a method suggested by the Indian physicist S. N.
Bose. Bose had derived the black-body law using a statistical model
of interacting particles based on the lightquantum, entirely free of
assumptions from classical electromagnetic theory.131 In 1925 Ein-
stein pointed out that de Broglie's hypothesis allowed one to treat
equivalently the fluctuations in black-body radiation and the sta-
tistical fluctuations of a gas according to Bose's statistics. He then
went on to propose a possible experimental test for diffraction in
atomic beams.132 The particular statistical application of de Brog-
lie's hypothesis that Einstein made in 1925 "stuck the matter right
under [Schrodinger's] nose."133
Among the physicists now interested in the lightquantum,
Erwin Schrodinger was in a unique position to appreciate the
merits of de Broglie's phase wave hypothesis. In 1922 he had called
attention to a "remarkable property of the quantum orbits of an
individual electron."134 A mathematical restriction that acts on an
electron in a periodic orbit suggested that with each revolution the
electron "adjusts" itself to its new point in four-dimensional
space-time. Thus, Schrodinger had already recognized a periodic,
spatially extended effect of each electron in space. It has been
suggested that Louis de Broglie's reputation among most German-
129
Einstein to Ehrenfest, 15 March [1921]. AHQP 1, 77.
130
Einstein, Berlin Sb (1923), 359-64, was as close as Einstein came to publishing
his ideas. See Lorentz, Problems (1927), 156-7, who lectured on the idea at
Caltech in 1922; Lorentz to Einstein, 13 November 1921, Einstein archive.
131
Bose, ZP, 26 (1924), 178-81.
132
Einstein, Berlin Sb (i924)> 261-7; (1925), 3-14.
133
Schrodinger to Einstein, 23 April 1926. Einstein archive. Quoted in Przibram,
Briefe zur Wellenmechanik (1963), 24. Schrodinger's first use of de Broglie's
hypothesis was in PZ, 27(1926), 95-101. For the influence of Einstein's work
in kinetic theory on Schrodinger's derivation of wave mechanics, see Klein,
Natural philosopher, 3 (1964), 3-49.
134
Schrodinger, ZP, 72(1922), 13-23.
300 Conceptual origins of wave-particle dualism, 1921 -1925
speaking physicists, notably those such as Bohr and Sommerfeld
who were working on atomic spectroscopy, was not high.135 But
Schrodinger, like Einstein and de Broglie, was committed to rela-
tivistic and statistical formulation of theory; he was not directly
concerned with problems of atomic structure.
Late in 1925, Schrodinger began a four-paper series in which he
extended the formal analogy between Hamiltonian mechanics of
particles and wave optics.136 Using de Broglie's hypothesis, Schro-
dinger reinterpreted quantum transitions as changes in the vibra-
tional mode of the superposition of "phase waves" that character-
ize the entire atomic system. He was relieved to have found an
alternative interpretation of quantum theory that avoided what to
him was the uncomfortable notion of mechanically indescribable
electron jumps from one orbit to another. The transition of an
electron could now be interpreted as a change in the vibrational
mode of the entire atom. The quantum wave mechanics that
resulted from Schrodinger's analysis was a direct descendant of
physicists' accommodation to the lightquantum hypothesis.
In the spring of 1926, Schrodinger showed that Heisenberg's
matrix mechanics, which Schrodinger said had "frightened" and
"repulsed" him by its purely formal character, was mathemati-
cally equivalent to his own wave mechanics.137 Schrodinger wrote
to de Broglie in March 1926 to describe his work. "It is possible to
calculate all the Born-Heisenberg matrix elements by simple
differentiations and quadratures if one has the complete solution
of the wave equation," he said.138 He was especially gratified to
135
Raman and Forman, HSPS, 1 (1969); 291-314. Hanle, Isis, 68 (1977), 606-9.
136
Schrodinger, AP, 79 (1926), 361-76. See also Wessels, SHPS, 10 (1979),
311-40.
137
Schrodinger, AP, 79 (1926), 734-56. Eckart, PR, 28 (1926), 711-26.
138
Schrodinger to L. de Broglie, 26 March 1926. (L. de Broglie dossier.) Extracts
from the original letter written in Schrodinger's French, follow [reprinted
with permission of Ruth Braunizer and l'Academie des Sciences de Paris].
"Pour des phenomenes petites ou fines il faut aussibien dans la mechanique
que dans Foptique remplacer le traitement geometrique par un traitement
ondulatoire-on pourrait aussi dire par un traitement physique,-en tenant
compte de ce que la longueur d'onde estfinieet non infiniment petite comme
le suppose l'optique geometrique (et de meme l'ancienne mechanique)." "II
est possible de calculer toutes les elements de matrices de Born-Heisenberg
par des simples differentiations et quadratures, quand on possede la solution
complete de l'equation d'onde mentionnee plus haut, c'est a dire un systeme
complet de 'Eigenfunktionen'."
Synthesis of matter and light 301
report that the wave mechanical formulation based on de Broglie's
hypothesis replaced the purely "geometrical formalism" by a wave
treatment, "one could also say by a physical treatment."
Two formerly independent research fields - the spectroscopic
analysis of atoms and the statistical analysis of radiation - were
brought into contact in 1925 by the long-delayed discussion of the
physical reality of lightquanta. By 1926 each had developed into
equivalent formulations of the new quantum mechanics. That
there were two independently formulated versions of the new
quantum mechanics was in large part an effect of the virtual
separation of the two parent fields - atomic theory and radiation
theory - for the preceding decade. And this, in a larger sense,
constituted the fullest corroboration of Louis de Broglie's remark-
ably fruitful but decidedly unorthodox proposal that matter and
light are fundamentally one and the same thing.
Epilogue
ELECTRON DIFFRACTION
The wave properties of light and the localized properties of matter
are familiar characteristics. The accumulated experimental evi-
3
Sommerfeld, Scientia, 51 (1932), 41-50.
304 Epilogue
All works mentioned in the notes and text are listed here by author
or editor. A few liberties have been taken in the interests of
consistency. Lords Cherwell, Kelvin, and Rayleigh appear here
under their family names - Frederick Lindemann, William
Thomson, John W. Strutt and Robert J. Strutt - even though
many of their works are published under their baronial title.
Similarly, English transliteration is used for Russian physicists
who published papers in a German or French transliteration; in
particular, Chwolson appears here as Khvol'son, Joffe as IofFe.
Names with prepositions (de Broglie, Des Coudres, van der Waals)
are listed under the final part.