0 ratings 0% found this document useful (0 votes) 405 views 792 pages Anthony Duncan-The Conceptual Framework of Quantum Field Theory-Oxford University Press, USA (2012)
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content,
claim it here .
Available Formats
Download as PDF or read online on Scribd
Go to previous items Go to next items
Save Anthony Duncan-The Conceptual Framework of Quantum... For Later The Conceptual Framework
of Quantum Field Theory
Anthony Duncan
OXFORD
UNIVERSITY PRESSOXFORD
UNIVERSITY PRESS
Great Clarendon Street, Oxford, OX2 GDP,
United Kingdom
Oxford University Press is a department of the University of Oxford.
If furthers the University’s objective of excellence in research, scholarship,
and education by publishing wo-ldwide. Oxford is a registered trade mark of
Oxford University Press in the UK and in certain other countries
© Amhony Duncan 2012
‘The moral rights of the author have been asserted
First Edicion published in 2012
Impression: 1
All rights reserved. No part of this publication may be reproduced, stored in
a retrieval system, or transmitted, in any form or by any means, without the
prior permission in writing of Oxford University Press, or as expressly permitted
by law, by licence or under terms agreed with the appropriate reprographics
rights organization. Enquiries concerning reproduction outside the scope of the
above should be sent to the Rights Department, Oxford University Press, at the
address above
‘You must not circulate this work in any other form
‘and you must impose this same condition on any acquirer
British Library Cataloguing in Publication Data
Data available
Library of Congress Cataloging in Publication Data
Data available
BN 978-0-19-957326-4
Printed and bound by
CPI Group (UK) Ltd, Croydon, CRO 4YY.
Links to third party websites are provided by Oxford in good faith and
for information only. Oxford disclaims any responsibility for the materials
contained in any third party website referenced in this work.Preface
In the roughly six decades since modern quantum field theory came of age with the
introduction in the late 1940s of covariant field theory, supplemented by renormaliza-
tion ideas, there has been a steady stream of expository texts aimed at introducing
each new generation of physicists to the concepts and techniques of this central area of
modern theoretical physics. Each decade has produced one or more “classics”, attuned
to the background, needs, and interests of students wishing to acquire a proficiency
in the subject adequate for the beginning researcher at the time. In the 1950s the
seminal text of Jauch and Rohrlich, Theory of Photons and Electrons, provided the
first systematic textbook treatment of the Feynman diagram technique for quantum
electrodynamics, while more or less simultaneously the first field-theoretic attacks on
the strong interactions were presented in the two-volume Mesons and Fields of Bethe,
de Hoffmann, and Schweber. The 1960s saw the appearance of the massive treatise
by Schweber, Introduction to Relativistic Quantum Field Theory, which addressed
in much greater detail formal aspects of the theory, including the LSZ asymptotic
formalism and the Wightman axiomatic approach. The dominant text of the late 1960s
was undoubtedly the two-volume text of Bjorken and Drell, Relativistic Quantum
Mechanics and Relativistic Quantum Fields, which combined a thorough introduction
to Feynman graph technology (in volume 1) with a more formal introduction to
Lagrangian field theory (in volume 2). In the 1970s the emergence of non-abelian
gauge theories as the overwhelmingly favored candidates for a successful field-theoretic
description of weak and strong interactions coincided with the emergence of functional
(path-integral) methods as the appropriate technical tool for quantization of gauge
theories. In due course, these methods received full treatment with the appearance
in 1980 of Itzhykson and Zuber’s encyclopedic Quantum Field Theory. In a similar
way, the surge to prominence of supersymmetric field theories throughout the 1980s
necessitated a full account of supersymmetry, which is the sole subject of the third
volume of Weinberg’s comprehensive three-volume The Quantum Theory of Fields,
the first edition of which appeared in 1995.
With such a selection of classic expository treatises (not to mention many other
fine texts not listed above—with apologies to authors of same!) one may well doubt
the need for yet another introductory treatment of quantum field theory. Nevertheless,
in the course of teaching the subject to graduate students (typically, second year) over
the last 25 years, I have been struck by the number of occasions on which important
conceptual issues are raised by questions in the classroom which require a careful
explanation not to be found in any of the readily available textbooks on quantum
field theory. To give just a small sample of the sort of questions one encounters in the
classroom setting: “Of the plethora of quantum fields introduced to describe Nature at
subatomic scales, why do so few (basically, only electromagnetism and gravity) have
classical macroscopic correlates?”; “If there are many possible quantum fields availableiv Preface
to ‘represent’ a given particle, can, or in what sense does, quantum field theory
prescribe a unique all-time dynamics?”; “If the interaction picture does not exist,
as implied by Haag’s theorem, why (or in what sense) are the formulas derived in
this picture for the S-matrix still valid?”; “Are there non-perturbative phenomena
amenable to treatment using perturbative (i.e., graph-theoretical) methods?”; and so
on. None of these questions require an answer if one’s attitude in learning quantum
field theory amounts to a purely pragmatic desire to “start with a Lagrangian and
compute a process to two loops”. However, if the aim is to arrive at a truly deep
and satisfying comprehension of the most powerful, beautiful, and effective theoretical
edifice ever constructed in the physical sciences, the pedagogical approach taken by
the instructor has to be quite a bit different from that adopted in the “classics”
enumerated above.
In the present work, an attempt is made to provide an introduction to quantum
field theory emphasizing conceptual issues frequently neglected in more “utilitarian”
treatments of the subject. The book is divided into four parts, entitled respec-
tively, “Origins”, “Dynamics”, “Symmetries”, and “Scales”. Although the emphasis is
conceptual-the aim is to build the theory up systematically from some clearly stated
foundational concepts—and therefore to a large extent anti-historical, I have included
two historical chapters in the “Origins” section which trace the evolution of the modern
theory from the earliest “penumbra” of quantum-field-theoretical phenomena detected
by Planck and Einstein in the early years of the twentieth century to the emergence,
in the late 1940s, of the recognizable structure of modern quantum field theory, in the
form of quantum electrodynamics. The reader anxious to proceed with the business
of logically developing the framework of modern field theory is at liberty to skim, or
even entirely omit, this historical introduction.
‘The three remaining sections of the book follow a step-by-step reconstruction of
this framework beginning with just a few basic assumptions: relativistic invariance,
the basic principles of quantum mechanics, and the prohibition of physical action at a
distance embodied in the clustering principle. The way in which these physical ingre-
dients combine to engender some of the most dramatic results of relativistic quantum.
field theory is outlined qualitatively in Chapter 3, which also contains a summary of
the topics treated in later chapters. Subsequent chapters in the “Dynamics” section
of the book lay out the basic structure of quantum field theory arising from the
sequential insertion of quantum-mechanical, relativistic, and locality constraints. The
rather extended treatment of free fields allows us to discuss important conceptual
issues (e.g., the classical limit of field theory) in greater depth than usually found in
the standard texts. Some applications of perturbation theory to some simple theories
and processes are discussed in Chapter 7, after the construction of covariant fields for
general spin has been explained. A deeper discussion of interacting field theories is
initiated in Chapters 9, 10, and 11, where we treat first general features shared by all
interacting theories (Chapter 9) and then aspects amenable to formal perturbation
expansions (Chapter 10). The “Dynamics” section concludes with a discussion of
“non-perturbative” aspects of field theory—a rather imprecise methodological term
encompassing a wide variety of very different physical processes. In Chapter 11 we
attempt to clarify the extent to which certain features of field theory are “intrinsically”Preface v
non-perturbative, requiring methods complementary to the graphical expansions made
famous by Feynman.
In the “Symmetries” section we explore the many important ways in which
symmetry principles influence both our understanding and our use of quantum
field theory. Of course, at the heart of relativistic quantum field theory lies an
inescapable symmetry of critical importance: Lorentz-invariance, which, together
with translational symmetry in space and time, makes up the larger symmetry of the
Poincaré group. The centrality of this symmetry explains the dominance of Lagrangian
methods in field theory, even though from a physical standpoint the Hamiltonian
would appear (as is typically the case in non-relativistic quantum theory) to hold
pride of place. The role played by Lorentz-invariance in restricting the dynamics of a
field theory is the main topic of Chapter 12, which also includes an introduction to the
extension of the Poincaré algebra to the graded superalgebra of supersymmetric field
theory. Discrete spacetime symmetries, and the famous twin theorems of axiomatic
field theory—the Spin-Statistics and TCP theorems—are the subject of Chapter
13. The discussion of global symmetries, exact and approximate, in Chapter 14
leads naturally into the very important topics of spontaneous symmetry-breaking
and the Goldstone theorem. The “Symmetries” section of the book closes with a
treatment of local gauge symmetries in Chapter 15, which imply remarkable new
features not present in theories where the only symmetries are global (i.e., involve a
finite-dimensional algebra of spacetime-independent transformations).
With the final section of the book, entitled “Scales”, we come to perhaps the
most characteristic conceptual feature of quantum field theory: the scale separation
property exhibited by theories defined by an effective local Lagrangian. Given that
essentially all of the information obtained from scattering experiments at accelerators
concerns asymptotic transitions (i.e., the infinite time evolution of an appropriately
prepared quantum state, terminated by a detection measurement) it is critically
important for theoretical progress that the probabilities of such transitions not depend
in a sensitive way on interaction details at much smaller distances than those presently
accessible in accelerator experiments (roughly, the inverse of the center-of-mass energy
of the collision process). The insensitivity of field theory amplitudes to our inescapable
ignorance of the nature of the interactions at very short distances (or equivalently,
high momentum) is therefore of central importance if we are to infer reliably an
underlying microdynamics from the limited phenomenology available at any given
time. Remarkably, in this respect quantum field theories are far kinder to us than their
classical (particle or field) counterparts, where non-linearities almost always introduce
chaotic behavior which effectively precludes the possibility of accurate predictions
of state evolution over long time periods. The technical foundations needed for
examining these issues are taken up in Chapter 16, which contains an account of
regularization, power-counting, effective Lagrangians, and the renormalization group.
Applications to the proof of perturbative renormalizability, and a discussion of the
“triviality” phenomenon (the absence of a non-trivial continuum limit), follow in
Chapter 17. Chapters 18 and 19 then explore important features of the behavior
of quantum field theories at short distance (e.g., the operator product expansion
and factorization) and long distance (in particular, the complications in defining the
correct physical state space in unbroken abelian and non-abelian gauge field theories).vi Preface
To the beginning student, quantum field theory all too often takes on the appear-
ance of a multi-headed Hydra, with many intertwined parts, the understanding of any
‘one of which seems to require a prior understanding of the rest of the frightening
anatomy of the whole beast. The motivation for the present work was the author's
desire to provide an introduction to modern quantum field theory in which this rich
and complex structure is seen to arise naturally from a few basic conceptual inputs, in
contrast to the more typical approach in which Lagrangian field theory is presented as
a theoretical fait accompli and then subsequently shown to have the desired physical
features.
Much (perhaps most) of the attitude towards quantum field theory expressed in this
book is the result of innumerable conversations, over four decades, with colleagues and
students. For laying the foundations of my knowledge of field theory I wish especially
to thank my predoctoral and post-doctoral mentors (Steven Weinberg and Al Mueller,
respectively). In the case of the present work I am extremely grateful to Estia Eichten,
Michel Janssen, Adam Leibovich, Max Niedermaier, Sergio Pernice, and Ralph Roskies
for reading extensive parts of the manuscript, and for many useful comments and
suggestions. Any remaining solecisms of style or content are, of course, entirely the
responsibility of the author.Contents
Origins I: From the arrow of time to the first quantum field
1.1 Quantum prehistory: crises in classical physics
1.2 Early work on cavity radiation
1.3. Planck’s route to the quantization of energy
1.4 First inklings of field quantization: Einstein and energy fluctuations
1.5 The first true quantum field: Jordan and energy fluctuations
2 Origins II: Gestation and birth of interacting field theory:
from Dirac to Shelter Island
2.1 Introducing interactions: Dirac and the beginnings of quantum
electrodynamics
2.2 Completing the formalism for free fields: Jordan, Klein, Wigner,
Pauli, and Heisenberg
2.3. Problems with interacting fields: infinite seas, divergent integrals,
and renormalization
3 Dynamics I: The physical ingredients of quantum field theory:
dynamics, symmetries, scales
4 Dynamics 11: Quantum mechanical preliminaries
4.1 The canonical (operator) framework
4.2. The functional (path-integral) framework
4.3 Scattering theory
4.4 Problems
5 Dynamics III: Relativistic quantum mechanics
5.1 The Lorentz and Poincaré groups
5.2 Relativistic multi-particle states (without spin)
5.3 Relativistic multi-particle states (general spin)
5.4 How not to construct a relativistic quantum theory
5.5. A simple condition for Lorentz-invariant scattering
5.6 Problems
6 Dynamics IV: Aspects of locality: clustering, microcausality,
and analyticity
6.1 Clustering and the smoothness of scattering amplitudes
6.2 Hamiltonians leading to clustering theories
6.3 Constructing clustering Hamiltonians: second quantization
6.4 Constructing a relativistic, clustering theory
6.5 Local fields, non-localizable particles!
Rowe
30
31
40
46
57
69
70
86
96
106
108
108
111
114
121
125,
130
132
133,
138
144
159viii
7
8
9
10
1
Contents
6.6 From microcausality to analyticity
6.7. Problems
Dynamics V: Construction of local covariant fields
7.1 Constructing local, Lorentz-invariant Hamiltonians
7.2. Finite-dimensional representations of the homogeneous
Lorentz group
7.3 Local covariant fields for massive particles of any spin:
the Spin-Statisties theorem
7.4 Local covariant fields for spin-} (spinor fields)
7.5 Local covariant fields for spin-I (vector fields)
7.6 Some simple theories and processes
7.7 Problems
Dynamics VI: The classical limit of quantum fields
8.1 Complementarity issues for quantum fields
8.2 When is a quantum field “classical”?
8.3. Coherent states of a quantum field
8.4 Signs, stability, symmetry-breaking
8.5 Problems
Dynamics VII: Interacting fields: general aspects
9.1 Field theory in Heisenberg representation: heuristics
9.2. Field theory in Heisenberg representation: axiomatics
9.3 Asymptotic formalism I: the Haag-Ruelle scattering theory
9.4 Asymptotic formalism II: the Lehmann-Symanzik-Zimmermann
(LSZ) theory
9.5 Spectral properties of field theory
9.6 General aspects of the particle-ficld connection
9.7 Problems
Dynamics VIII: Interacting fields: perturbative aspects
10.1 Perturbation theory in interaction picture and Wick’s
theorem
10.2. Feynman graphs and Feynman rules
10.3 Path-integral formulation of field theory
10.4 Graphical concepts: N-particle irreducibility
10.5 How to stop worrying about Haag’s theorem
10.6 Problems
Dynamics IX: Interacting fields: non-perturbative aspects
11.1 On the (non-)convergence of perturbation theory
11.2 “Perturbatively non-perturbative” processes: threshhold bound
states
11.3. “Essentially non-perturbative” processes: nou-Borel-summability
in field theory
11.4 Problems
164
169
171
171
173
sues
184
198
202
215
219
219
223
228
234
238
240
241
253,
268
281
289
297
304
307
309
314
325,
341
359
371
374,
376
386
400
41112
13
14
15
16
Contents
Symmetries I: Continuous spacetime symmetry: why we need
Lagrangians in field theory
12.1. The problem with derivatively coupled theories: seagulls, Schwinger
terms, and T* products
12.2 Canonical formalism in quantum field theory
12.3 General condition for Lorentz-invariant field theory
12.4 Noether’s theorem, the stress-energy tensor, and all that. stuff
12.5 Applications of Noether’s theorem
12.6 Beyond Poincaré: supersymmetry and superfields
12.7 Problems
Symmetries II: Discrete spacetime symmetries
13.1 Parity properties of a general local covariant field
13.2 Charge-conjugation properties of a general local covariant field
13.3 Time-reversal properties of a general local covariant field
13.4 The TCP and Spin-Statistics theorems
13.5 Problems
Symmetries III: Global symmetries in field theory
14.1 Exact global symmetries are rare!
14.2 Spontaneous breaking of global symmetries:
14.3. Spontaneous breaking of global symmetrie:
14.4 Problems
the Goldstone theorem
dynamical aspects
Symmetries IV: Local symmetries in field theory
15.1 Gauge symmetry: an example in particle mechanics
15.2 Constrained Hamiltonian systems
15.3 Abelian gauge theory as a constrained Hamiltonian system
15.4 Non-abelian gauge theory: construction and functional integral
formulation
15.5 Explicit quantum-breaking of global symmetries: anomalies
15.6 Spontaneous symmetry-breaking in theories with a local
gauge symmetry
15.7. Problems
Scales I: Scale sensitivity of field theory amplitudes and
effective field theories
16.1 Scale separation as a precondition for theoretical science
16.2 General structure of local effective Lagrangians
16.3 Scaling properties of effective Lagrangians: relevant, marginal,
and irrelevant operators
16.4 The renormalization group
16.5 Regularization methods in field theory
16.6 Effective field theories: a compendium
16.7 Problems
ix
414
414
421
426
431
443
464
469
470
ATA
ATT
478
485
487
489
492
495
507
509
509
512
519
529
544
552
565
569
570
571
574
581
588
595
608x Contents
17
18
19
Scales II: Perturbatively renormalizable field theories
17.1 Weinberg’s power-counting theorem and the divergence structure
of Feynman integrals
17.2. Counterterms, subtractions, and perturbative renormalizability
17.3. Renormalization and symmetry
17.4 Renormalization group approach to renormalizability
17.5 Problems
Scales III: Short-distance structure of quantum field theory
18.1 Local composite operators in field theory
18.2 Factorizable structure of field theory amplitudes: the operator
product expansion
18.3 Renormalization group equations for renormalized amplitudes
18.4 Problems
Scales IV: Long-distance structure of quantum field theory
19.1 The infrared catastrophe in unbroken abelian gauge theory
19.2. The Bloch—Nordsieck resolution
19.3 Unbroken non-abelian gauge theory: confinement
19.4 How confinement works: three-dimensional gauge theory
19.5 Problems
Appendix A The functional calculus
Appendix B Rates and cross-sections
Appendix C Majorana spinor algebra
References
Index
610
613
645
652
660
662
664
679
698
708
aaa
7131
Origins I: From the arrow of time
to the first quantum field
1.1 Quantum prehistory: crises in classical physics
The first indications of serious inadequacies in the framework of classical physics—
deficiencies which eventually could only be resolved by the introduction of quantum-
theoretical concepts and methods—can already be found in Maxwell’s discussion of
the anomalously low specific heat of gases in the mid-1870s. Nevertheless, the birth of
quantum theory as such, and in particular the clear identification of a new fundamental
constant of Nature characteristic of quantum phenomena, is usually located in the
year 1900 with Planck’s invention (Planck, 19000) (“derivation” would perhaps be too
charitable a term) of a novel formula for the distribution of energy over frequencies
in thermal radiation (namely, electromagnetic radiation in the interior of a sealed
enclosure which has been allowed to come to thermal equilibrium with the walls of the
enclosure, themselves maintained at a fixed temperature T). ‘The problem of thermal
cavity (or “blackbody” radiation would in modern terms be regarded as one of thermal
quantum field theory, so we have the strange situation that the first historical impetus
to the discovery of quantum principles actually lay in a problem of quantum field
theory, which is generally supposed to be a much later invention arising from a fusion of
quantum, relativistic, and locality principles. In fact, Planck’s papers of 1900 make no
reference to quantization of the electromagnetic field-energy (a concept which Planck
would continue to resist strenuously until the mid-1920s): the energy quantization
principle for Planck is strictly a statement about the distribution of energy among
the idealized material oscillators constituting the walls of the enclosure, and general
principles of thermodynamics are then brought to bear to fix the electromagnetic
energy distribution which must necessarily obtain once equilibrium between these
oscillators and the interior radiation has been achieved. Only five years later, in his
remarkable paper entitled “A heuristic point of view concerning the creation and
conversion of light” (Einstein, 19052), Einstein was to extend boldly and explicitly
the idea of energy quantization to the electromagnetic field itself, introducing the idea
of “light quanta” (in modern language, photons). From a conceptual (if not technical)
point of view, this paper therefore marks the true birth of quantum field theory.
In order to understand the origins of quantum theory in the problem of black-
body radiation, which to physicists of Planck’s generation must have appeared
quintessentially classical (merging as it did well-established principles of Maxwellian
electromagnetic theory and thermodynamics), we need to push our time horizon back
a quarter century or so and survey the overall situation in classical physics at the start2 Origins |: From the arrow of time to the first quantum field
of the final quarter-century of the 1800s. The three great: edifices of classical physics
Newtonian mechanics (amplified and deepened, of course, by the contributions of
Laplace, Lagrange, Hamilton, and many others), electromagnetic theory, only recently
completed by Maxwell (in Philosophical Transactions, vol. 155, 1865), and, somewhat
later, put in a form recognizable to the modern student of the subject by Hertz,
and thermodynamics, which reached essential conceptual completeness at the hands
of Clausius, also around 1865—stood as precise descriptions of natural phenomena,
each apparently unassailable in its natural domain of applicability. In a sense, further
progress keeping strictly within the limits of each of these disciplines had become
difficult or impossible. However, precisely at this time, natural phenomena requiring
the simultaneous application of more than one of these formal structures began to
demand the attention of physicists. It is possible to trace the origins of the core
disciplines of twentieth-century physics—quantum theory, statistical mechanics, and
relativity—to developments at the interfaces of the three basic classical frameworks.
We can summarize these developments very briefly as follows:
1. Developments at the interface of thermodynamics and electromag-
netic theory.
The discovery and classification of solar spectral lines by Fraunhofer, and the
development of spectroscopy as an analytical tool by Bunsen and Kirchhoff in
the 1850s, led naturally to an investigation of the radiation emitted by hot
bodies, and thus to the study of the relation between thermodynamics and
electromagnetic phenomena. It was immediately recognized by Kirchhoff that
the intensity and frequency of the radiation emitted by a perfectly absorbing
body had a fundamental significance. Various arguments—some of a rigorous
thermodynamic nature, others of an heuristic character—led, by 1896, to a
widely accepted form for this “blackbody distribution” (the Wien Law). The
experimental failure of this “law” was the final stimulus which led Planck to
(reluctantly) advance the quantum hypothesis in his seminal papers (Planck,
1900) in Annalen der Physik, 1900.
»
. Developments at the interface of mechanics and thermodynamics.
Attempts to reconcile classical mechanics, regarded as the underlying dynamical
description of all phenomena, with the formal principles of thermodynamics led
to the development of kinetic theory by Clausius, Maxwell, and Boltzmann. The
more general framework developed by the last of these has since come to be
called statistical mechanics. Boltzmann was the first to understand clearly the
role of statistical and probabilistic considerations in reconciling mechanics with
heat theory (as thermodynamics was called at the time). However, his views were
highly controversial at the time, especially with regard to his claim of a statistical
origin for the phenomenon of irrever 'y in thermal physics. Although he would
prove to be absolutely right on this point, Boltzmann’s methods proved incapable
of explaining the observed specific heats of gases: this difficulty, early (before
1875) recognized by Maxwell (Maxwell, 1875) as a serious anomaly in classical
theory, would only finally be removed by the application of quantum ideas in the
1920s, fully fifty years after the problem was first recognized.Early work on cavity radiation 3
3. Developments at the interface of electromagnetic theory and
mechanics.
‘The historical and conceptual preeminence of classical mechanics implied a spe-
cial status which led naturally to an attempt to interpret all natural phenomena
in mechanical terms. In particular, the attempt to weld clectromagnetic theory
with mechanics (begun by Maxwell himself) by main force led by the second half
of the nineteenth century to the development of a profusion of aether theories of
increasing complexity and artificiality. The failure to produce any direct evidence
for the existence of an aether in optical experiments (by measuring relative motion
of the aether and the optical apparatus of choice) grew from a mere annoyance
into an outright crisis with the null result of the experiments of Michelson (1881)
and Michelson and Morley (1887). The entire class of complicated and messy
dynamical aether theories concocted to surmount this impasse were demolished
with one stroke by Einstein (Einstein, 19054) in 1905, by accepting the kine-
matical structure natural to electromagnetic phenomena as generally valid in the
mechanical sphere also.
Tn our outline of the conceptual origins of quantum mechanics and quantum field
theory, the first item above holds pride of place: firstly, because by common consent
quantum theory begins with Planck’s introduction of a new universal constant of
Nature in his blackbody distribution formula of 1900, and secondly, because, as we shall
see later, the first explicitly quantum field-theoretic calculation, Jordan’s derivation of
the mean-square energy fluctuations in a subvolume of a cavity containing (a one-
dimensional version of) electromagnetic radiation in the final section of the Drei-
Ménner-Arbeit (1925) of Born, Heisenberg, and Jordan (Born et al., 1926), came
directly out of an attempt to reproduce a remarkable result of Einstein dating from
1909 (Binstein, 1909b,a) in which the apparent paradox of simultaneous wave and
particle behavior of light was first exposed with full clarity. The essence of quantum
field theory is to provide a unified dynamical framework in which these apparently dis-
parate behaviors can coerist in a conceptually consistent fashion. The electromagnetic
radiation contained in a cavity at thermal equilibrium therefore plays a central role
in the conceptual origins both of quantum theory generally and quantum field theory
in particular. For this reason we shall retell in this chapter the history of thermal
radiation in some detail, paying particular attention to those aspects important for
understanding the conceptual origins of quantum field theory. The story will lead
us continuously from the early arguments surrounding the role of the Second Law
of Thermodynamics (and the “arrow of time” it implies) in blackbody radiation, to
the appearance of the first truly quantum-field-theoretical analysis of electromagnetic
radiation. We begin in the next section by describing some important milestones on
the way to the understanding of blackbody radiation as it stood in 1900 when Planck
took the first steps along the road to modern quantum theory.
1.2 Early work on cavity radiation
The fact that heated bodies glow with a color and intensity varying with their temper-
ature must surely have been apparent in prehistoric times (at least since the discovery4 Origins |: From the arrow of time to the first quantum field
of fire!). The precise nature of thermal radiation became the subject of intense study
in the nineteenth century, and as we shall see, led directly to the discovery of the
quantum principle. Wedgewood, the porcelain manufacturer, observed in the 1790s
that heated bodies all became red at the same temperature. The Scottish physicist
Balfour Stewart noted (in 1858) that a block of rock salt at 100° C strongly absorbs
the radiation emitted by a similar block at the same temperature, and suggested the
rule of equality of radiating and absorbing power of bodies for rays of any given type
(ie., wavelength). About a year later (independently of Stewart) Kirchhoff put all
of this phenomenology into a comprehensible framework by the use of very general
thermodynamic arguments (Kirchhoff, 1859, 1860).
The arguments given by Kirchhoff established the universal character of blackbody,
or “cavity”, radiation—the radiation filling the interior of a hollow material cavity, the
walls of which are maintained at a fixed temperature 7’ It is important to understand
at the outset the role of the cavity in these arguments. After heating the cavity to
the desired temperature 7, a fixed amount of radiant energy fills the interior. To
examine the nature of this radiation we are at liberty to drill a very small hole in
the walls, as the small amount of radiant energy emerging will not sensibly disturb
the established equilibrium. Note that from the point of view of the external world,
the punctured cavity is essentially “black”, in the sense that any radiation entering
through the pinhole will have to scatter around in the interior of the cavity for a
very long time before having an opportunity to escape. Such a cavity is therefore
(effectively) a perfect absorber, or “black body”. The essence of the problem is that
radiation in the cavity is forced to interact with the walls (or contents, if any) of the
cavity until thermal equilibrium is reached.
Kirchhoff showed that the energy (per unit volume, and per wavelength interval) of
the cavity radiation was uniform and isotropic throughout the interior. The arguments
he gave were subsequently simplified by Pringsheim. The latter observed that by
inserting a reflecting plane surface at some point inside the cavity, the equality of
radiation in opposite directions follows, as otherwise the unequal radiation pressure
exerted on the two faces would allow the spontaneous conversion of heat to work,
thereby violating the Second Law of Thermodynamics. The existence of a pressure
exerted by radiation reflected from a surface is crucial to these arguments: the precise
relation (to be discussed further below) between this pressure and the energy density of
the radiation had been derived earlier by Maxwell (Maxwell, 1873). Similar arguments
employing two mirrors can be used to prove the isotropy and homogeneity of the
radiation in the cavity. The existence of filters selectively absorbing and transmitting
radiation of different wavelengths means that these statements hold separately within
each interval of wavelength. Thus, we can define a function ¢(A, 7) such that (A, T)AA
is the radiant energy per unit volume between wavelengths ,\ + AA anywhere in the
cavity.
"The use of the term “blackbody”, though historically predominant, frequently confuses the beginning
student, who wonders quite naturally how a truly black body can radiate! In the forthcoming discussion we
prefer the use of the term “cavity (or thermal) radiation”. The German terminology, “Normal Spektrum",
is not particularly illuminating either.Early work on cavity radiation 5
Next, one easily sees that (A, 7) is the same (universal!) function for any cavity,
irrespective of size, shape, or material constitution. This can be established by
connecting two cavities at the same temperature by a thin tube allowing radiation
to pass in either direction. Unequal radiation densities would result in unequal fluxes
down the tube, and hence in a spontaneous flow of heat between two systems at
the same temperature. Again, by placing filters which pass only a limited range of
wavelengths in the tube, this equality is found to hold in each wavelength interval
a
The universal function ¢(A, 7) appears in the description of radiation emitted from
any hot surface in the following way. At thermal equilibrium the radiation impinging
on the surface must be balanced by the total radiation leaving, as otherwise the surface
would grow progressively cooler or hotter. Suppose the surface (say, the interior wall
of our cavity discussed above) absorbs a fraction Ay of the incident radiation flux ¢,
(at wavelength A). It is clear that the radiation emitted (as opposed to reflected), or
the “emissive power” of the surface By must just equal Aye), or equivalently, the ratio
E)/Ay of emissive power to absorption coefficient is a universal function (essentially
our old friend @) of wavelength and temperature. This result, of course, lies at the
core of Stewart’s observations mentioned above. It also explains why thermos flasks
are silvered (making A, small) in order to get them to radiate less.
The total energy density of cavity radiation (at all wavelengths) is evidently just
the integral of Kirchhoff’s function ¢:
aT) = footer (a)
In 1879 Josef Stefan proposed on the basis of some preliminary experiments the
form
(7) = aT* (1.2)
with a a universal constant. In 1884 Boltzmann derived this formula thermodynam-
ically, essentially by the following argument. Consider a spherical cavity of radius r.
Maxwell had shown that radiation of energy density p exerts a radiation pressure
4p. Thus the internal energy is U = $rp and the heat absorbed in an infinitesimal
reversible expansion is given by the First Law of Thermodynamics as?
dQ = dU +4W = ag ar®p) + dar? xen (1.3)
with the corresponding entropy change
4 dy 4 dp dT
a8 = 4/7 = 409 @ 4 Cant = Sui OE 4 Cart a
pig epg arta ep OA)
2 We remind the reader that U and S are state functions, unlike work W and heat Q: only the changes
in the latter are meaningful, whence the difference in notation in the associated differentials W, dQ.6 Origins |: From the arrow of time to the first quantum field
Equating 498 to 238 we find
dp _ 4,
ar" T° (18)
which immediately implies the T* law stated above, and now known as the Stefan—
Boltzmann Law.
Serious attempts to measure experimentally the intensity and spectral composition
of blackbody radiation began with Langley, the American astronomer, who invented
the bolometer, a device for measuring the intensity of radiation in the infrared, using
the principle of temperature-dependent resistance of a thin filament placed in the path
of the rays after they were refracted through a rock-salt prism. These measurements
(1886) extended up to about wavelengths of 5. He found “a real though slight
progression of the point of maximum heat towards the shorter wave-lengths as the
temperature rises” and the asymmetric form of the maximum of (A,7), steeper on
the shorter wavelength side. By 1895 the greatly improved measurements of Paschen
established the rule that the wavelength Am of maximum intensity was inversely
proportional to the temperature T. Paschen’s measurements led him to propose, in
1896, the form
60,7) = BY exp(-4) (1.6)
with the constant C’ somewhere in the range of 5-6.
In 1893 Wien derived, using purely thermodynamic arguments, an important
constraint on the Kirchhoff function $(A,T). Consider once again the spherical cavity
used above in the derivation of the Stefan—Boltzmann Law. Imagine that the sphere
undergoes slow, adiabatic compression, where the radius contracts steadily at speed v
(v << ¢). At every reflection from this inwardly contracting sphere light of wavelength
2 suffers a Doppler shift to wavelength \(1—2v/c). During a contraction by Ar,
occurring in time Ar/v, the light undergoes cAr/2ur reflections across a diameter of
the sphere (it can be shown that light not incident perpendicular to the walls suffers a
smaller Doppler shift cach time, but is reflected correspondingly more frequently: the
net result is the same). The result is a total blue shift to wavelength
20, gar Ar
a-)#a~a- ya
so the wavelength is shifted by 44 = —4* in this adiabatic compression. As there is
no heat transfer dS = 0 and (see Eq. (1.4) above)
Tor ago
aoa aa ala aia,
T
dra
Te
37
7
1
sora a.
r
The essence of the thermodynamic argument given by Wien lies in the observation
that the adiabatic process described here gives at every stage cavity radiation inEarly work on cavity radiation 7
equilibrium at the new temperature T' inversely proportional to r. If the result were
otherwise, producing more or less radiation in some wavelength band than appropriate
for blackbody radiation at the new temperature, introduction of filters absorbing
in this range immediately allows the generation of temperature differences and a
consequent violation of the Second Law.
In the time interval At of the adiabatic compression, the radiant energy
originally in the wavelength band (A, + 4), namely
grr", r)AX (17)
is shifted down to the interval (A(1 — 244), (A + AA)(1 — 24¢)), and increased by the
amount of adiabatic work done against the radiation pressure, which is
(4er)(0At)(40,7)A) (18)
Thus
4 a(n — vAt)SO(A(L - , r—vasara — “At
3 7
= gros) - (Grr? wane, r)AX (1.9)
Expanding to first order in At, one finds the equation
or
a. a5
gy + rg, )O%A) =0
so that,
$= fA/r) = FAT)
Wien phrased this result slightly differently. Since
1
o> orslOP)
it follows that if iT, = A2T», then
+
sen) = 70027)
?
which Wien called the “T®” law, but is now commonly called the Wien Displacement
Law. If we know the radiation function (1,71) at temperature 7), the law allows us
to “displace” it into the appropriate curve for any other temperature Tp, as
(A, Ta) = Fol ap, qT)8 Origins |: From the arrow of time to the first cuantum field
It follows immediately from the Wien Displacement Law that the total energy density
or) = [ ax0.7)
= { ans fear)
oe
ie f de,
:
x
(x) =aT* (1.10)
satisfies the Stefan-Boltzmann Law (provided, of course, that the integral converges:
a condition by no means to be taken for granted, as we shall see). Another imme-
diate corollary is the result later verified experimentally by Paschen (but suggested
previously by several workers in this field, notably H. F. Weber) that the maximum in
wavelength displaces inversely with the temperature. Finally, the Wien Displacement
Law immediately fixes the value of the constant C in the form proposed by Paschen
(see (1.6)) to be 5 exactly. Independently of Paschen’s work, Wien in 1896 arrived
at the form (1.6) on the basis of an ad hoc assumption concerning the emission of
radiation by molecules distributed according to a Maxwellian velocity distribution.
This form, which is now a complete specification of the Kirchhoff function ¢, was
called the Wien Distribution Law, and was to play, with its “corrected” version, the
Planck Distribution Law, a critical role in the evolution of attempts to understand
quantization of the electromagnetic field.
1.3. Planck’s route to the quantization of energy
Max Planck, the father of quantum theory, was born in Kiel, Germany, in 1858, and
attended the Gymnasium (high school) and University in Munich before going to
Berlin for his doctoral degree, where he had classes from Helmholtz and Kirchhoff.
His doctoral thesis concerned the application of thermodynamics (2 la Clausius) to
problems of “Evaporation, Melting, and Sublimation”. Planck was fascinated by the
extraordinary scope, power and (apparent) infallibility of the energy conservation
principle—or First Law of Thermodynamics—and accorded to the Second Law of
Thermodynamics (with its concomitant “arrow of time”) an equal degree of validity. He
was therefore convinced that the Second Law could not rest on the purely mechanical
foundations of Boltzmannian gas theory, in which the reversibility objections of
Loschmidt and Zermelo (the latter a Planck assistant in Berlin) would necessarily lead
to spontaneous processes (albeit rare) where the entropy decreased.* The consideration
of the paradox of Maxwell’s demon also led Planck to the rather peculiar conclusion
that the Second Law could never be valid in @ system comprised of discrete particles
Instead, Planck began to investigate (in 1897) the possibility that irreversible
thermal phenomena could somehow be traced back to irreversible processes in a
continuous medium—in particular, the electromagnetic field. The archetypal process
considered by Planck was the apparently irreversible conversion of plane radiation
5 For a beautiful retelling of this remarkable period in the development of statistical heat theory, see the
biography by Martin Klein of Paul Ehrenfest (Klein, 1970).Planck's route to the quantization of energy 9
incident on a charged oscillator into outgoing spherical waves. This subject was
explored with great thoroughness in a series of five papers in the Berliner Berichte
(1897-99) entitled “ber irreversible Strahlungsvorginge” (“On irreversible radiation
processes”) (Planck, 1900a). The subject of absorption and re-emission of electromag-
netic radiation from an oscillator led Planck naturally into the subject of thermal
cavity radiation. Here the oscillators constitute the material of the walls of the cavity,
absorbing and re-emitting the radiation in the interior. The universal character of the
thermal radiation discussed above allowed Planck the freedom of making a very simple
model of the constituent particles of the walls (essentially charged simple harmonic
oscillators), as the spectral distribution of the cavity radiation would have to be
independent of the specific material constitution of the cavity once equilibrium is
reached.
The irreversibility that Planck relies upon in his radiation studies can be seen
clearly in the damped oscillator equation that he derived as a prelude to his studies
of the coupled field-oscillator problem:
2. 2 gs
nit + ke — a = eF cos(2nvt) (1.11)
The third (“radiation damping”) term has three time-derivatives and evidently
changes sign under time-reversal. It arises because the damping force times the velocity
must give the power lost to radiation, which is proportional to the acceleration of the
charged particle squared. The average of the third term above times the velocity
42 over a cycle of the periodic system is easily seen to be the same as the average
power radiated, by a single integration by parts. Planck was particularly impressed
by the fact that the irreversibility in this system arises without any recourse to non-
conservative processes, in which ordered energy is lost (as in friction or air resistance)
to disordered heat. Instead, the energy appears to flow irreversibly from an ordered
source (an incoming plane wave incident on the oscillator) to an equally ordered form:
outgoing spherical radiation.
In 1898 Boltzmann succeeded in convincing Planck (in a paper entitled “On
the supposedly(!) irreversible radiation processes” (Boltzmann, 1898)) that the hope
of deriving irreversible phenomena. from electromagnetic theory without additional
statistical assumptions was bound to fail, as Maxwell’s equations are just as invariant
under time-reversal as those of classical mechanics. In fact (Boltzmann claimed), in
the course of a careful derivation of radiation damping one is forced to apply boundary
conditions to the fields which amount to a field analog of the assumption of molecular
disorder implicit in the Boltzmann approach to gas theory.* Planck admitted this
promptly and abandoned the attempt at a “microscopic” explanation of irreversibility
based on electrodynamics.
In the fifth of his papers on irreversible radiation processes (Planck, 1899), Planck
derived a crucial formula relating the distribution function for cavity radiation to the
average energy of his fictional oscillators (at equilibrium). Before stating this formula,
4See (Klein, 1970), (Kubn, 1978) for masterful expositions of the remarkable developments at the
interface of mechanics and heat theory summarized all too briefly above.10 Origins I: From the arrow of time to the first quantum field
aslight change in notation will be convenient. Let p(v,T)dv be the energy/unit volume
of cavity radiation in the frequency interval (v,1 + dv), where vA = ¢, |dv| = dA, so
that
pv, T)dv = ant) gaa = (A, T)dd
2 ce
p.t) = ~9(a,7) = So(2,7) (1.12)
In terms of p(v,T), the T® law takes the form
eT) = 4(Z) (1.13)
and the Wien Distribution Law is
p(v,T) = av’ exp(—Bv/T) (1.14)
where the new constants a, @ are related to those appearing in the Paschen result
(1.6) by a= 4,8 = 4 (recall that the constant C was fixed previously by purely
thermodynamic reasoning to be 5). The equation derived by Planck, obtained by
equating at equilibrium the energy absorbed and emitted by the oscillator, stated
simply
E(vo,T) (Yo, T) (1.15)
~ Be?
It relates the average energy of an oscillator of natural frequency 1,
the blackbody distribution function. That such a relation must exist is physically
clear. Planck showed that if the left-hand side exceeded the right, energy would flow
from the oscillators to the electromagnetic field, while if the intensity of radiation at
vy became large enough that the right-hand side exceeded the left the oscillators
would tend to absorb energy from the field. The importance of this equation (a
full derivation of which we must unfortunately forego, in the interests of brevity) in
Planck’s intellectual journey can scarcely be overemphasized: it allowed him to restrict
the application of energy quantization to the material oscillators alone (left-hand side
of (1.15)), while relying on the equilibrium condition to transfer the resultant average
distribution of energy by main force, as it were, to the continuous electromagnetic
radiation (right-hand side) in the interior of the cavity. Planck would continue to
insist on the continous, purely classical character of electromagnetic radiation for the
next 25 years.
In his final paper on irreversible radiation processes (see (Planck, 1900a)), Planck
gave a “derivation” of the Wien Law based on purely thermodynamic arguments
together with the crucial formula (1.15) above. This was done by making a plausible
assumption for the form of the entropy J of the oscillator as a function of energy E,
using for the inverse temperature T~! = 35, and solving for E as a function of T.
Planck showed that his assumption for S(E) implied that the entropy of the whole
system (oscillators plus radiation) would necessarily increase in time, in agreement
with the Second Law of Thermodynamics. He was also under the (as it later turnedPlanck's route to the quantization of energy 11
out, erroneous) impression that this was the only possible choice for $(Z) consistent
with the Second Law. Consequently, at this point Planck was quite convinced that
he had finally managed a complete derivation of the blackbody spectrum from pure
thermodynamics (even if he had now to agree with Boltzmann that the Second Law
had a statistical rather than absolute significance, even in radiation phenomena)
On the afternoon of Sunday, 7 October 1900, Planck was visited at home by an
experimental colleague from the Physikalische-Technische Reichsanstalt (the Physical-
‘Technical Imperial Institute, or PTR), H. Rubens. He learnt from Rubens that recent
experiments at the PTR had established incontrovertible deviations from the Wien
Distribution Law on the infrared (low-frequency) side. In particular, the intensity was
roughly proportional to temperature in this regime, instead of the saturation at high
temperatures implied by the Wien Law (1.14). Planck realized that a more general
form for the oscillator entropy S(E) would in turn allow the derivation of a modified
distribution law
. 1
= BI
OT) =o I (1.16)
which clearly reproduces the Wien Law at higher frequencies, but behaves like
|
~ ST LA?
2 (117)
in the infrared (small v), showing the desired linear behavior with T. This interpolating
formula, which Planck appears to have constructed in the few hours following the visit
of Rubens, was checked within the next week and a half and found to match exactly
the experimental data.
Planck was perfectly aware that his interpolating formula was nothing more than
an enlightened guess at this stage, and he began right away to search for a proper
understanding of the formula (1.16). His strategy was preciscly the inverse of the one
he had followed heretofore. He used (1.15) to obtain the average oscillator energy,
assuming the validity of the Planck distribution (1.16):
hv
BON sp Buf
(1.18)
where h = 2%. He then reconstructed the corresponding expression for oscillator
entropy as a function of energy, using the thermodynamic relation TdS = dE valid
for a reversible transformation involving transfer of heat but no external work. Here,
oscillators of a fixed natural frequency v (called vp above) are considered. Solving
(1.18) for 1/T’ as a function of E:
(1.19)12 Origins |: From the arrow of time to the first quantum field
and integrating, one obtains,
— i
= es [oe + hv) —In(Z))dE
h E E E
= gilt pin + iw!
apart from an irrelevant, integration constant. The problem now shifted to finding a
“fundamental” explanation for this last expression.
By this point in late 1900, Planck had been converted to Boltzmann’s statistical
approach, and he now adopted the teckniques used by the latter for gas theory in
an attempt to establish (1.20) by microstatistical reasoning. Thus, the entropy was
to be determined by taking the logarithm of the number of available microscopic
states consistent with the stated macroscopic parameters, S = kIn(W) (here k is
Boltzmann's constant). Like Boltzmann, Planck introduced a finite-energy unit € to
facilitate the counting. The total energy Ey shared by N oscillators was a (large!)
integer P number of these units, Ey = Pe. Planck then “counted” W by simply
computing the number of ways in which P units of energy could be distributed among
the N oscillators. The combinatorial formula needed for this can be derived rapidly
using a characteristically elegant trick due to Ehrenfest. Write out a string of P energy
units €, with dividers to indicate how many units belong to the first, second, etc.,
oscillator:
(1.20)
eelecelelee...lee
There are P of the « symbols and N — 1 dividers. First assume that all these symbols
are distinguishable. There are then (P-+N —1)! ways of ordering them. As the
dividers and energy units are (separately) indistinguishable, we have overcounted by a
factor (N — 1)!P!. Thus the desired result (using Stirling’s approximation to evaluate
the factorials of large numbers) is
(P+N-1)!
$= kw Sy yr
) ~ k(N + P)In(N + P) — Pin(P) — Nin(N))
The average entropy of each oscillator S =
oscillator is E = Ey = Fe. Consequently
4Sw while the average energy of a single
S=HQ+ =)t
(1.21)
This is exactly the relation (1.20), provided we identify = hv, 4 = k. In other words,
the derivation of the new distribution formula forced Planck to keep the energy units ¢
finite, even at the end of the calculation. Setting ¢ to zero here, as Boltzmann had done
at the end of his gas theory calculations, would be equivalent to setting the constant
h to zero, which would lead to an incorrect distribution law (the Rayleigh-Jeans Law,Planck's route to the quantization of energy 13
to be discussed further below). Apparently, the oscillators in the walls of a cavity
were only allowed to have energies in integer multiples of the basic energy “quantum”
¢ = hv! The arguments outlined above were presented in Planck’s paper in Annalen der
Physik 4 (1901),p. 553, “Uber das Gesetz der Energicverteilung im Normalspectrum”
(“On the law of energy distribution for the normal [i.e., blackbody] spectrum”). The
famous Planck's constant h appears here for the first time, In modern notation, the
blackbody distribution thus takes the form
8x hy
“S exp(hv/kT) —1 (1.22)
pv, T) =
From the experimental fits, Planck determined h = 6.55 x 107?” erg/sec, and k (Boltz-
mann’s constant)= 1.346 x 10-1 ergs/degree. The latter value allowed Planck to
obtain the first decently accurate value for Avogadro's number N = R/k (where R
is the gas constant).
It is a strange historical irony that Planck’s modification of the Wien’s Law,
motivated by the pressure of the Kurlbaum-Rubens experimental results, was actually
a move towards a “more classical” result: as Einstein was to emphasize in his epochal
1905 paper (Einstein, 19052), in which the revolutionary idea of field quantization
was introduced, the Wien Law is in a sense an extreme manifestation of the quantal
properties of light. The deviations observed from this law in the infrared by Kurlbaum
and Rubens are harbingers of the reappearance of the classical wave-like aspects of
electromagnetic phenomena. To understand this we must realize that despite Planck’s
heroic efforts to obtain a rigorous and unique classical result for the distribution func-
tion of cavity radiation throughout the 1890s, leading up to the quantum-theoretically
correct Planck distribution (1.22), the first derivation of the blackbody distribution
based on a consistent and full application of classical principles is actually due to Lord
Rayleigh. In a short (two-page) paper published in 1900 (Rayleigh, 1900) Rayleigh
derived the correct classical form of the distribution function from the classical
equipartition theorem applied directly to the electromagnetic modes in the cavity.
Consider a cubical LxLxL box containing electromagnetic radiation in the form of
standing waves. A typical standing wave mode takes the form
mymE, . meTy . MsTz
) sin( re
sin( 2H) sin( ST)
where the associated frequency is v = 3571 and 71 is the vector with (positive) integer
Cartesian components (n1,n2,n3)- The number of such modes in the shell ((7i), || +
djit|) (octant of positive components only!~—an error of Rayleigh’s later corrected by
Jeans, see below) is evidently
1 Bs
gitinPdin| = 4a V7dv
and each of these modes receives a total of 2kT at equilibrium by the equipartition
principle (namely, 447 each into electric and magnetic field energy, and each of two
polarization modes). Thus the energy per unit volume in the field in the frequency
interval (¥,v + dv) is14 Origins |: From the arrow of time to the first quantum field
pv, T)dv
: RT (4 a Pdi = 8 ekT dy (1.23
Ae ee eae )
‘a result which has since become known as the Rayleigh-Jeans Law. (The error
mentioned above of an overall factor of eight made by Rayleigh in his original paper was
subsequently corrected by Jeans. As Pais points out in his biography of Einstein (Pais,
1982), the correction was made also in Einstein's 1905 paper on the light quantum, so
the result should perhaps more properly be called the Rayleigh-Jeans-Einstein Law.)
Rayleigh was perfectly aware that this result could not be correct: the total energy
contained in the cavity radiation, when integrated over all frequencies, would then be
infinite! Instead, he assumed that it was correct only for the “graver modes” (i.e., lower
frequencies) and that: the distribution wes modified for some as yet unknown reason.
at higher frequencies (Rayleigh simply inserted an exponential suppression factor at
high frequencies, and the resultant formula was in fact his final result). In any event
the simple linear dependence on temperature in the Rayleigh-Jeans Law flies in the
face of experience: a bar of steel at room temperature (300 K, say) does not emit
radiation at one-tenth the blinding intensity of a bar at 3000 K ! The infinite amount
of energy present in the classical radiatior. field under equipartition would later (1911)
be referred to by Ehrenfest (Ehrenfest, 1911) as the “ultraviolet catastrophe”
Of course, if Planck had finished his Boltzmannian calculation of the average
oscillator energy by taking the energy units € to zero, as Boltzmann had done
previously in his discussion of gas theory, 1e would have arrived precisely at Rayleigh’s
result (though he does not seem to have been aware of Rayleigh’s work during the
critical period leading up to the 1901 paper), as the Rayleigh-Jeans Law is simply the
h — 0 limit of the Planck distribution. That he did not do so is probably due to a
combination of reasons:
1. He does not seem to have regarded equipartition as a fundamental guiding
principle to the same extent as other physicists of a more “mechanist” bent.
Planck attacked the problem from the point of view of the behavior of the
oscillators at thermal equilibrium, rather than by directly considering the modes
of the electromagnetic field itself, which would have led much more quickly to
the (wrong!) classical result.
The result obtained by setting the energy units to zero would not have agreed
with the Wien Law, with which Planck had started and which he knew to be
empirically correct at higher frequencies.
n
bad
1.4 First inklings of field quantization: Einstein and energy
fluctuations
Although Planck succeeded in obtaining an absolutely correct expression for the
equilibrium thermal frequency distribution of electromagnetic radiation in a cavity,
there is absolutely no indication that he supposed any sort of energy quantization to
hold for the electromagnetic field itself. Instead, the (at this point frankly magical)
effect of the energy quantization imposed on the material oscillators receiving from
and transferring energy to the radiation in the interior was forcibly transferred to
the electromagnetic field via the equilibrium formula (1.15). The field itself, PlanckFirst inklings of field quantization: Einstein and energy fluctuations 15
was to insist for almost another full quarter century, was a continuous, fully classical
entity regulated by Maxwell’s equations. The situation was to change dramatically
with Einstein’s remarkable 1905 paper, “On a heuristic point of view concerning
the creation and conversion of light” (Einstein, 1905a). Although this paper is now
commonly referred to as the “photoelectric paper”, Einstein spends much more time
in it on an analysis of the volume-dependence of blackbody radiation (pp. 92-102)
than on the brief discussion (pp. 104-105) of the photoelectric effect.
‘After pointing out that a strictly classical analysis must necessarily lead to the
Rayleigh-Jeans result (1.23), with its inescapable concomitant ultraviolet catastrophe,
Einstein goes on to analyse cavity radiation in the high-frequency domain, drawing
some extraordinarily non-classical conclusions from the quintessentially “classical” (at
least from an historical point of view) Wien Law. Einstein’s approach in this paper is
radically different from Planck’s. He focusses first and foremost on the thermodynamic
and statistical properties of the electromagnetic radiation in the interior of the cavity.
Taking a cavity of volume Vo and considering only the electromagnetic radiation in
the frequency interval (v,v + dv), the energy E of such radiation in the high-frequency
domain where Wien's Law (1.14) holds is given by
8rhv?
o
EB Voe" BF dv (1.24)
Solving this equation for 4 and repeating the integration procedure of (1.20) to obtain
‘an expression for the entropy S of the electromagnetic radiation in this frequency
interval, one finds (the 0 subscript indicates that the radiation in the entire cavity of
volume Vo is being considered—we shall shortly consider radiation in a subcavity)
KE E
So =~, (ln (Geaenagae) ~ 13 (1.25)
‘The same amount of radiation confined to a smaller volume V would lead to an entropy
$ with exactly the same form as (1.25) but with Vo replaced with V. Accordingly, the
difference in entropy for the two situations is
S-—S =—b(— = kin( eM (1.26)
Vo
‘The fundamental Boltzmannian association of entropy with the probability W of the
associated microstates of the system, S = kln W, then leads to the conclusion that
_
Wa
Elke (1.27)
i.e., that the probability of an energy fluctuation leading to a concentration of all the
electromagnetic radiation in the frequency interval (v,v + dv) in the subvolume V
of the full cavity Vo takes exactly the form which we would expect if that radiation
consisted of = “mutually independent energy quanta” (cach of energy hv) moving
freely throughout the cavity, in complete analogy to the behavior of molecules in a
gas. This is as far from the classical picture of electromagnetic radiation as extended
waves subject to mutual (destructive and constructive) interference as it is possible to16 Origins J: From the arrow of time to the first quantum field
get. The result (1.26)—extraordinarily simple, but profoundly baffling, from a classical
point of view—clearly had a deep impact on Einstein’s thinking. He was to hold firmly
to the concept of energy (and later momentum) quantization of the electromagnetic
field over the next 20 years—a period of time in which the majority of physicists
were firmly on Planck’s side and resistant to any notion of quantization of the sacred
classical Maxwellian fields.
The centrality of blackbody radiation to Einstein’s thinking about the nature of
the electromagnetic field is clear once one reflects on the number of occasions on
which he would return to the subject: to take the most prominent cases, in 1909 in
two papers (Einstein, 19096,a) (one entitled “On the present status of the radiation
problem”, the other “On the development of our conceptions on the nature and
constitution of radiation”) in which energy fluctuations were once more used as a
diagnostic for exposing the underlying properties of radiation, and in 1917, in the
famous “A-B coefficients” paper (Einstein, 1916, 1917), of critical importance in
the later development of dispersion theory by Kramers, and thereafter in the 1925
development of matrix mechanics at the hands of Heisenberg, Born, and Jordan.5 Here
we briefly review Einstein's results of 1909, which proved to be a critical inspiration
for Jordan’s introduction in 1925, in the last section of the “hree-Man” paper of
Born, Heisenberg, and Jordan (Born et al., 1926), of the first true quantum field.
In returning to the problem of energy fluctuations in cavity radiation, Einstein
decided to relax the simplifying assumption of high-frequency (or low-density) radia-
tion described by the Wien Law, and to enquire into the implications of the full Planck
distribution (1.16), valid at all densities and frequencies, for the fluctuation properties
of thermal radiation. In this case, instead of considering the highly non-Gaussian
process whereby a fluctuation would concentrate 100% of the radiation energy in
a given interval (v,v+dv) in a subvolume V (giving the result (1.27), later to be
called “Einstein’s first fluctuation theorem” by Jordan) Einstein decided to calculate
the mean-square energy fluctuation of the energy in this interval in the subvolume
V. The formula for such mean-square fluctuations is a standard result of statistical
mechani
2 2 dE) :
(Any) = eet (1.28)
where T is the temperature and (£2) the mean energy, which in this case is clearly just
Vplv,T)dv. We can distinguish three interesting choices for the energy distribution
p(v,T) and corresponding mean-square energy fluctuation. We shall distinguish the
results obtained for the mean-square energy fluctuation in each case by a subscript
indicating the assumed form for the universal Kirchhoff distribution function p(v,T):
“RJ” for the completely classical Rayleigh—Jeans form, for the Wien Law, and
“P” for the final result of Planck. In the case of the Rayleigh-Jeans Law valid at low
frequencies,
5 For a thorough study of the role played by dispersion theory in the birth of modern quantum mechanics,
see the two-part paper by M. Janssen and the present author (Duncan and Janssen, 20074,6)