Download as pdf or txt
Download as pdf or txt
You are on page 1of 39

Phys. Perspect.

16 (2014) 179–217
 2014 Springer Basel
1422-6944/14/020179-39
DOI 10.1007/s00016-014-0136-6 Physics in Perspective

Writing the Biography of Hans Bethe:


Contextual History and Paul Forman1
Silvan S. Schweber*

Some facets of the life of Hans Bethe after World War II are presented to illustrate how
Paul Forman’s works, and in particular his various theses—on mathematics and physics in
Wilhelmine and Weimar Germany, on physics in the immediate post-World War II period,
and on postmodernity—have influenced my biography of Bethe. Some aspects of the history
of post-World War II quantum field theory, of solid state/condensed matter physics, and of
the development of neoliberalism—the commitment to the belief that the market knows
best, to free trade, to enhanced privatization, and to a drastic reduction of the government’s
role in regulating the economy—are reviewed in order to make some observations regarding
certain ‘‘top-down’’ views in solid state physics in postmodernity, the economic and cultural
condition of many Western societies since the 1980s, the decade in which many historians
assume modernity to have ended.

Key words: Hans Bethe; Paul Forman; Weimar; Forman theses; postmodernity;
neoliberalism; effective field theory; nuclear theory; mathematical physics; physics
summer schools; Lamb shift; Mutually Assured Destruction (MAD); Nuclear
Utilization Target Selection (NUTS).

Introduction
In 1971, Paul Forman, then an assistant professor of history at the University of
Rochester, published a long article on the relation between Weimar culture and
acausality in quantum theory.2 In it, he argued that a sense of crisis had permeated
all aspects of life—including science—in the economically impoverished, socially
fickle, and politically riven Germany after its defeat in World War I. Widely read
ideological treatises such as Oswald Spengler’s Decline of the West made it
acceptable to reject former commitments to rationality, to progress, and to
modernist values. In this context, determinism and causality came under attack
and intuition and irrationalism gained standing. Forman claimed that to accom-
modate themselves to this hostile environment a number of prominent

* Silvan S. Schweber is the Koret Professor of the History of Ideas and Professor of Physics,
emeritus, at Brandeis University. Since the early 1980s, he has been an associate in the
department of the history of science at Harvard University. He is at work completing the
second volume of a biography of Hans Bethe.

179
180 S. S. Schweber Phys. Perspect.

mathematicians and physicists rejected or limited the validity of causality in


physics and incorporated acausality in the interpretation of quantum mechanics.
Forman’s thesis contradicted the then generally accepted belief (called ‘‘inter-
nalism’’) that the conceptual framework of science was autonomous and did not
depend on social conditions.3 If this Forman thesis has remained somewhat con-
troversial in the specificity of its foci (Weimar Germany, physics, quantum
mechanics), Forman’s more general thesis—that the cultural values prevalent in a
given place and period can influence the direction of research and the conceptu-
alization of the results of research within a scientific discipline—has become
widely accepted when writing the history of science.
After he became a curator at the National Museum of American History in
1972, Forman published a second thesis concerned the military funding of research
during the Cold War and how it affected the direction of the research carried out
by physicists in the United States.4 According to Forman, physicists eagerly
accepted money from the Department of Defense and from the Atomic Energy
Commission—the agency responsible for the development of atomic weaponry—
but their support altered the direction, and often the aims, of their research.
Forman’s claim was that the physicists accommodated themselves to this envi-
ronment by cultivating a false consciousness. As Heilbron put it succinctly: ‘‘They
worried about their purity and autonomy and managed to obtain perhaps 5 per-
cent of the military R&D budget for what they considered basic research; but they
neglected to take into account the other 95 per cent, which supported their col-
leagues and students in applied projects of direct military interest.’’5
A third Forman thesis deals with postmodernity and its manifestations, in
particular, what he terms ‘‘the Preposterous Primacy of Science Relative to
Technology Prior to Postmodernity,’’ interdisciplinarity, and neoliberalism.6 As
Philip Mirowski formulated it: The fundamental tents of neoliberalism are that the
market is an ideal information processor and that every successful economy is a
knowledge economy. The market, though a collective artifact, knows more than
any individual and can always provide solutions to problems, even to problems
caused by markets in the first place. Similarly, corporations can do no wrong.
Neoliberalism believes democracy to be the best state framework for an ideal
market, yet wants it to have as little power as possible. It conceptualizes politics as
if it were a market and it advances an economic theory of democracy by regarding
citizens as customers of state services. Thus, for neoliberalism education becomes
a consumer good, not the collective responsibility of molding students into able,
informed, responsible, caring social beings. And most importantly, neoliberalism’s
adherents believe that their vision of the good society will not come about ‘‘nat-
urally’’ but only by concerted political effort and organization.7
All three Forman theses are concerned with the impact of the prevalent cultural
and political values at a given time and place on the content and direction of
science, and all three have influenced me in the writing of my biography of Hans
Bethe (figure 1).
Vol. 16 (2014) Writing the Biography of Hans Bethe 181

Fig. 1. Paul Forman at the Smithsonian Institution, 2003. Courtesy of Paul Forman.

Nuclear Forces, the part of my biography of Bethe that covers his life until 1940,
was published in June 2012.8 By virtue of the radically changed and changing
context from 1940 on, the dramatic increase in the number and variety of activities
Bethe became involved in, the range of issues that have to be addressed, my
limited competence and my limited time, the second volume—qua contextual
history—will be different from the first in content, in connections, and in scope.
Nonetheless, I still want it to be an overview of the history of theoretical physics
from the mid 1940s till the late 1970s—in part because Bethe’s metaphysics, as
expressed in his contributions to the explanation of the Lamb shift, became the
dominant outlook of the theoretical physics community at the end of the twentieth
century. Thus, instead of dealing with the ‘‘pre-established harmony between
physics and mathematics,’’ as had been the case in the first volume, the second will
address the issue of quantum field theory becoming the language of theoretical
physics. Similarly, where the first volume made observations about underlying
local Jewish affinities, the second will be concerned with scientific universalism.
Before 1940, Bethe was interacting scientifically only with other physicists and
shunned politics, but after 1945 was a sought-after consultant to several industries
and was deeply engaged politically both at the local and national levels. Therefore,
the narration of the entanglement of all these activities—not to mention what was
happening in his personal life—will have to be addressed differently than in the
previous volume by virtue of scope and magnitude.9
To indicate the approach I have taken, I have chosen some episodes from
Bethe’s life that resonate with aspects of Paul Forman’s three theses. I believe that
investigating resonances between a given cultural, political, and social settings and
182 S. S. Schweber Phys. Perspect.

the modes of explanations of physical phenomena within that context is a fun-


damental aspect of the narration of the history of science. Given that physics is
primarily concerned with phenomena in systems that that can be isolated from
their surroundings and that the success of physics stems from the fact that it is able
to do so, it is important to extend investigations like Forman’s regarding Weimar
culture and its role in shaping the development of quantum mechanics to later
periods and other situations. Thus, except for some remarks concerning mathe-
matics and physics in Wilhelmine and Weimar Germany as they influenced
Bethe’s later activities, I shall be principally concerned in the present article with
physics in the immediate post-World War II period10 and with the third Forman
thesis. Hence, I have presented an account of Bethe’s metaphysics as revealed in
his Lamb shift calculation and an exploration of some of the enabling conditions—
as reflected in institutions and disciplines—for the dramatic advances in theoret-
ical physics after World War II. I conclude by making a claim concerning the
impact of neoliberalism on the approaches of some leading theorists to problems
in condensed matter physics in the post-1970 period.
The article is organized as follows: Section 1 briefly narrates some facets of the
young Bethe’s life not mentioned in Nuclear Forces and highlights the post-1940
part of his life. Section 2 is concerned with post-World War II physics. Subsec-
tion 2.1 deals with the practice of ‘‘elementary particle’’ theoretical physics and
adumbrates Bethe’s metaphysics by outlining his approach to the Lamb shift
problem. Section 2.2 is concerned with the mathematical physics discipline, its
formation and contributions. Section 2.3 deals with the self-image of physicists
after World War II and Bethe’s transmutation from being an ‘‘elementary particle’’
to a nuclear matter theorist; section 2.4 explains ‘‘effective field theories.’’ Sec-
tion 3 is concerned with Forman’s third thesis, and adumbrates the resonance
between neoliberalism and the ‘‘top-down’’ views of condensed matter physics
expounded by two leading condensed matter physicists. Section 4 highlights the
final decades of Bethe’s life.

Some Biographical Background


Bethe was born in 1906 in Strassburg, when it was part of Germany (figure 2). At
the time his father, Albrecht Bethe, was an Assistent to the Strassburg University
physiologist Richard Ewald. His mother was the daughter of a professor of
medicine at the university. Hans was the only offspring of the marriage; Bethe’s
father stimulated and nurtured his interest in science. Having spent some time in
the mid-1890s at the Naples Marine Biological Station under the influence of its
director, Anton Dohrn, a student of Ernst Haeckel, Albrecht Bethe became
deeply committed to an all-encompassing evolutionary view. Hans Bethe’s ana-
lysis and explanations of evolutionary, historical, physical processes—e.g., his
researches on energy generation in stars, on nucleosynthesis, on stellar evolution,
and those on the death throes of stars such as novae and supernovae11—were
Vol. 16 (2014) Writing the Biography of Hans Bethe 183

Fig. 2. Hans Bethe in the mid-1950s. Credit: Rose Bethe.

mathematical, quantified narrations of the processes involved12 and were focal


points of his dual career as a physicist and as an astrophysicist. I attribute the
stimulus for these interests and viewpoints to his father.13
In 1915, Bethe’s father accepted the directorship of the animal physiology
institute at the newly established Frankfurt University. Bethe attended Gymna-
sium in Frankfurt and in 1924 enrolled in Frankfurt University. Although Bethe
had outstanding young mathematicians as instructors in his mathematics courses at
Frankfurt University—in particular Carl Ludwig Siegel and Otto Szász—their
courses convinced him that he didn’t want to major in mathematics. The way
mathematics was being taught seemed to him to have no connection with the real
world, and no connection with the sciences.14
One of his physics teachers at Frankfurt University, recognizing Bethe’s
superior abilities, urged him to go to Munich to study with Arnold Sommerfeld.15
In the spring of 1926, just when Schrödinger’s seminal papers on wave mechanics
were being published in the Annalen der Physik, Sommerfeld accepted the twenty-
year-old Bethe to his seminar, then the outstanding school of theoretical physics in
Europe.
Göttingen (and its preeminent mathematician, David Hilbert) had been
responsible for the formal nature of the mathematics Bethe was taught in his
mathematics courses in Frankfurt. Another aspect of the Göttingen spirit deeply
influenced Bethe, namely Sommerfeld’s views concerning the special connection
between physics and mathematics that been instilled in him by Felix Klein, on the
one hand, and by Hilbert and by Hermann Minkowski, on the other. One facet of
Klein’s ‘‘pre-established harmony between physics and mathematics’’ considered
mathematics a tool that was created, sharpened, and expanded when used in
184 S. S. Schweber Phys. Perspect.

formulating physical theories and models. In a beautiful article on the construction


of knowledge, Timothy Lenoir pointed to Klein’s views that physical concepts and
theories are strongly coupled to the instruments and practices through which
phenomena are produced and stabilized.16 Klein believed that the distinction
between pure and applied science and between science and technology was
unfounded and unwarranted. Thus, the various institutes of applied science that
Klein founded at Göttingen were brought together under one roof. This outlook
molded Sommerfeld, who also embodied Hilbert’s and Minkowski’s versions of
the Kleinian ‘‘pre-established harmony between physics and mathematics.’’ For
Hilbert, pre-established harmony came to mean a pre-established unity of math-
ematics and the possibility of stipulating a set of axioms from which all of
mathematics would be derivable. For Minkowski, it came to mean the existence of
invariance principles that constrain all physical theories. In his famous 1905 paper
on the electrodynamics of moving bodies, Einstein had realized that two observers
in relative motion with respect to each other needed to agree on some parameters
in order to be able to transfer information between them. He made light, i.e. the
electromagnetic field, the carrier of information and made the velocity of light a
universal constant. The Lorentz transformation, which guaranteed that the speed
of light would be the same for all observers, was introduced as the transformation
connecting any two observers. The Lorentz group is also the invariance group of
Maxwell’s equations. It need not be the case that the transformation group used
for changing frames of reference be the same as the symmetry group of physical
laws. Einstein stipulated that the laws governing the dynamics of light and those
governing the dynamics of material particles manifest the same symmetry group
and adhere to the same transformation group for changing frames of reference.17
Minkowski fully recognized the power of Einstein’s insight and made it a general
principle, an expression of the ‘‘pre-established harmony between physics and
mathematics,’’ namely that symmetry, i.e. invariance, dictates the form of the
equations. In his general theory of relativity, Einstein made a further, and deeper,
contribution, that locality together with the symmetry of general covariance dic-
tate the form of the gravitational interactions. This became the basis of the great
advances in physics during the second half of the twentieth century.
Sommerfeld inculcated all his students with this dual metaphysics: on the one
hand, the value of mathematics in the physicist’s toolkit and on the other, seem-
ingly, its great content of physical reality.18 Given Bethe’s particular abilities and
inclinations, he saw mathematics as an all-important component in his toolkit as a
theoretical physicist—the Kleinian rather than the Einsteinian view of the pre-
established harmony between mathematics and physics—and was less concerned
with mathematics as a guide to the representation of physical reality.
Bethe’s off-scale ability to absorb and analyze huge amounts of complex data
and synthesize the materials into useful knowledge with the help of transparent,
readily understandable models was first demonstrated in the two encyclopedic
Handbuch der Physik articles he published in 1933.19 In them, he gave a masterly
Vol. 16 (2014) Writing the Biography of Hans Bethe 185

exposition of the application of quantum mechanics to atomic, molecular, and


solid state physics. Each was written in less than a year and reflected his uncom-
mon energy, his extraordinary powers of concentration, and his great ambition.
These articles set standards for subsequent contributions to these fields and have
remained classics to this day.
In 1933, after Hitler came to power, Bethe lost his position as an assistant
professor in Tübingen because his mother came from a Jewish family and had
converted to Protestantism as a young woman. Sommerfeld helped him obtain a
fellowship in England. In February 1935, he joined the physics department of
Cornell University, where he stayed until the end of his life.
During the 1930s the frontier of physics shifted to nuclear physics. Bethe
became an acknowledged leader in this field, co-authoring with Stanley Livingston
and Robert Bacher three lengthy articles in the Reviews of Modern Physics that
became known as the Bethe Bible of nuclear physics.20 His mastery of nuclear
physics made it possible for him to put forward in 1938 the explanation of energy
generation in stars for which he won the Nobel Prize in 1967.
Bethe’s experiences during World War II transformed his life. He was the
paradigmatic example why theoretical physicists proved to be so important in the
war effort and the early stages of the Cold War. It was his ability to translate his
technical mastery of the microscopic world—i.e. the world of nuclei, electrons,
atoms, and molecules—into an understanding of the macroscopic properties of
materials and into the design of macroscopic devices such as radar junctions and
atomic bombs that rendered his services so valuable at the Radiation Lab and at
Los Alamos.
In 2003, in a retrospective assessment of the Theoretical Division at Los Ala-
mos, the 97-year-old Bethe stated that the T-Division’s most important scientific
contributions during the war were the introduction and use of electronic com-
puters and the elucidation of the physics of implosions and explosions. The punch-
card-fed electromechanical IBM computers that Eldred Nelson, Stanley Frankel,
and Richard Feynman assembled made possible the extensive, complex compu-
tations required to solve the complicated physical and technical problems they
were dealing with, such as the use of uranium hydride instead of uranium metal or
the design of explosive lenses. John von Neumann, a consultant to the division,
observed the efficient work of the IBM machines and decided that ‘‘We must use
them more generally and they must work electronically.’’ Bethe concluded his
retrospective article with the statement that ‘‘the successful work of T-Division
during the war was the seed of modern computers.’’21
As the head of the theoretical division at Los Alamos, Bethe acquired mana-
gerial and entrepreneurial skills that proved valuable when he was to return to
Cornell. In discussions with Ezra Day, the president of Cornell at the time, he and
Robert Bacher negotiated the creation of the Laboratory of Nuclear Studies,
whose mission was ‘‘to investigate the particles of which atomic nuclei are com-
posed and to discover more about the nature of the forces which hold these
186 S. S. Schweber Phys. Perspect.

particles together.’’ It was to be strictly a disciplinary enclave dedicated to pur-


suing high energy physics for its own sake, with possible applications of nuclear
studies excluded.22 Thus, in contrast to other laboratories of nuclear studies
established at American universities after the war, such as the ones in Chicago,
Iowa, and MIT, the Cornell Newman Lab did not include biologists, chemists,
geologists, metallurgists, or nuclear engineers. This reflected Bethe’s ‘‘modernist’’
views of what a university was about and his determination that the researches
within the physics department at Cornell be ‘‘pure.’’23 Possible by-products of
physics research would be the province of Cornell’s applied science and engi-
neering departments.
By establishing this laboratory, Bethe became involved in the administration of
physics activities at Cornell. As at many of the other premier American univer-
sities, Cornell’s physics department after the war split into two semi-independent
subdivisions: the high energy, big science, Newman Laboratory with its synchro-
tron and Bacher its director, and the solid state component with the chair of the
physics department, Lloyd P. Smith, nominally its director. Given the tensions
between the two directors and between the fields they represented, an adminis-
trative committee had to be set up to oversee the operation of the department, in
particular, its teaching responsibilities and its appointments. The committee con-
sisted of the two directors and Bethe, who as the ‘‘holy ghost’’ made sure that the
channels of communication between the two directors remained open.24
What had been obvious to others at Los Alamos, namely, that his contributions
to making the bomb projects successful were exceptional and that no one besides
Oppenheimer had played as significant a role, only became apparent to Bethe
somewhat later.25 Thereafter, creating peaceful applications for nuclear power
became an obsession for Bethe. In the decade after the war, besides being an
extremely productive physicist, Bethe became a consultant to General Electric’s
Knolls Laboratory to help them design safe nuclear reactors, to the Nuclear
Development Corporation of America to develop shielding for nuclear reactors,
and to Detroit Edison to explore the feasibility of breeder reactors. Similarly, to
make the world safe from the use of nuclear weapons, Bethe devoted great efforts
to have the Lilienthal-Oppenheimer plan adopted as official United States policy.
The adoption of the Baruch plan to represent the United States position26 made it
clear to him that something like MADness (Mutual Assured Destruction) would
become the modus vivendi after the Soviet Union would develop its own atomic
weapons.
After the first Soviet bomb was detonated in late August 1949, 27 Bethe became
deeply involved in formulating a response to the new threat. Initially, he was very
much opposed to the crash program for building a hydrogen bomb that Truman
had ordered, but after Stanislaw Ulam and Edward Teller proposed a mechanism
that made fusion weapons plausible he became totally occupied with their feasi-
bility. He spent the academic year 1951/2 at Los Alamos helping design the
weapon because he believed that if the United States could build a fusion bomb, so
Vol. 16 (2014) Writing the Biography of Hans Bethe 187

could the Soviets—hence deterrence was the proper policy. Upon his return to
Cornell, he had an office in one of the Federal buildings on the campus, where he
and Frederic de Hoffmann devoted a good deal of their time to highly secret and
classified fusion work. He later also became a consultant to the AVCO (AViation
COrporation) research laboratory in Everett, Massachusetts, whose director was
Arthur Kantrowitz, a former colleague at Cornell, and there helped solve the
ablation problem connected with the reentry of ballistic missile payloads into the
atmosphere.28
In addition to all these involvements, Bethe was also an active member of the
American Physical Society and served as its president in 1954 during the critical
period that witnessed the revocation of Oppenheimer’s security clearance. This
action eliminated Oppenheimer as an advisor to the AEC and other governmental
agencies. The ruin of Oppenheimer29 sent a clear message to American physicists
about the limits of the roles they could play in shaping nuclear policy. Thereafter,
in the hope of being able to keep in check the more extreme elements making
recommendations or policy concerning nuclear weapons, Bethe became ever more
occupied with national security issues as a member of some of the highest echelon
governmental advisory committees by virtue of his expertise, his integrity, and the
respect he commanded. In the early 1960s as a member of the President’s Science
Advisory Committee (PSAC), Bethe played a crucial role in the negotiations with
the Soviet Union that led to the treaty of 1963 that banned tests of nuclear
weapons in the atmosphere, outer space, and underwater and underground tests
above a certain limit.30 As a member of PSAC, he also was instrumental in
bringing about the large governmental support given to high energy physics and in
making NASA a civilian agency.31
Richard Nixon’s election in 1968 marked a turning point in Bethe’s life. Bethe
was a Democrat and this political identification meant that his advice and expertise
were no longer sought by the government. This, however, freed him to start an
extremely productive career as an astrophysicist32 and make major contributions
to the elucidation of the structure of neutron stars, of supernovae and of binary
star systems, and to the unraveling of the solar neutrino problem.33
Remarkably, in addition to these fruitful but demanding scientific activities,
Bethe became deeply involved with energy policies after the energy crisis of
1973.34 He widely and forcefully argued for the reconsideration of nuclear power
in the aftermath of the oil embargo the Arab states had imposed following the
Yom Kippur war in 1973. And during that same period, as a consultant at AVCO,
he was investigating the possibility of separating the uranium isotopes using lasers
and designing high power ‘‘chirping’’ lasers for that purpose. Similarly, in 1983 and
thereafter he collaborated with Richard Garwin, Kurt Gottfried, Henry Kendall,
and other Union of Concerned Scientists (UCS) members to challenge the claims
President Reagan and Edward Teller were making for their Strategic Defense
Initiative (SDI) scenario.
188 S. S. Schweber Phys. Perspect.

Post-World War II Physics


Bethe’s peregrinations before coming to Cornell in 1935—doing physics in
Munich, Stuttgart, Cambridge, Rome, and Manchester—made him aware of what
Karin Knorr-Cetina succinctly stated about knowledge practices:
Epistemic cultures [are] those amalgams of arrangements and mechanisms—
bonded through affinity, necessity, and historical coincidence—which, in a given
field, make up how we know what we know. Epistemic cultures are cultures that
create and warrant knowledge.35
Before the war, theoretical physics was practiced differently in Munich, Stuttgart,
Cambridge, Rome, and Manchester: scientific knowledge was warranted differ-
ently in each of these places, though as results were reproduced, the community at
large eventually arrived at a consensus on the value of the work. The same could
be said more globally: for a host of reasons, physics was different and was prac-
ticed differently in France, Germany, Great Britain, Japan, the Soviet Union, and
the United States, if only because the sciences had become nationalized.36
Nonetheless, there was some general consensus about the accomplishments and
aims of theoretical physics. The 1930s brought a quantum-field-theoretical dem-
onstration that the electromagnetic interactions between charged particles could
be explained as due to photon exchanges,37 along with Fermi’s theory of b-decay,
and Yukawa’s suggestion that, in analogy to electromagnetic forces, the short-
range nuclear forces could be generated by the exchanges of a hitherto unobserved
massive particle.38 The ensuing conceptualization of physics became ever more
important. It consisted in
a) recognizing that the physical world—at the level of accuracy of possible
physical measurements and the corresponding theoretical representations—
could be considered hierarchically ordered into fairly well delineated realms39
and concerns: the cosmological—consisting of galaxies and their constituents,
their evolution and dynamics; the macroscopic—consisting of solids, liquids,
gases, their structure, properties, and processes; the molecular and atomic
realm; the nuclear; and the sub-nuclear.40
b) stipulating that the aim of physics was to identify, classify, and characterize
these various realms and their interrelations. The microscopic and submicro-
scopic levels became considered as more ‘‘fundamental’’ since it was believed
that one could reconstruct the higher levels in terms of the knowledge of the
properties and dynamics of the entities that populated the lower levels.
Furthermore, it was the task of the theories representing the lower levels to
account quantitatively for the empirically determined parameters that
described the ‘‘elementary’’ building blocks of the higher levels.
Thus, in the case of an atom, the mass, charge, and magnetic moment of the
electrons and of its nucleus enter as experimentally determined parameters in the
Vol. 16 (2014) Writing the Biography of Hans Bethe 189

non-relativistic Schrödinger equation. It is the task of nuclear physics to explain


and determine the value of the nuclear parameters. In his 1937 articles on nuclear
physics in the Reviews of Modern Physics, Bethe did just that in term of phe-
nomenological inter-nucleonic potentials determined from proton-proton and
neutron-proton scattering experiments. Subsequently, he and others attempted to
derive these potentials from meson theories41 despite the fact that there was little
confidence that quantum field theory was adequate to explain the nuclear domain
in terms of subnuclear constituents.42
After World War II—for a while at least—the practice of much of cutting-edge
experimental and theoretical physics became much the same all over the Western
world.43 The new instruments developed during the war to enhance the effec-
tiveness of radar transformed atomic and molecular physics. Similarly, the crystal
growing techniques developed during the war to obtain pure samples of germa-
nium and silicon and other metals transformed solid state physics. Transistors,
masers, and lasers were among the devices made possible later by these advances.
The article Forman published in 1995 in the Reviews of Modern Physics entitled
‘‘‘Swords into Plowshares’: Breaking New Ground with Radar Hardware and
Technique in Physical Research after World War II’’ provides the evidence to
justify my assertion with respect to experimental physics.44 In this remarkably
complete, well-organized survey, Forman reviewed the application in the years
immediately following World War II to fundamental experimental physics research
of the microwave instrumentalities developed during the war, now available as off-
the-shelf equipment: in molecular beam magnetic resonance spectroscopy,
molecular spectroscopy, radio astronomy, and the design of high energy accelera-
tors. Forman’s examination of the postwar applications was preceded by a survey of
the prewar physical research that depended on the availability of sources of elec-
tromagnetic radiation in the meter to submillimeter range.
The new experimental knowledge and practice in microwave technology (to
which Bethe had contributed significant theoretical underpinnings at Cornell from
1940 until 1942 and later at the Radiation Laboratory at MIT during his stay there
in 1942) became widely disseminated in the 27 volumes of the Radiation Labo-
ratory Series that Louis Ridenour edited, whose first volume on radar system
engineering appeared in 1947.45 These volumes allowed experimentalists to
build—in an almost algorithmic fashion—accurate sources of constant voltage and
constant current, fast amplifiers, pulse generators, high precision sources and
detectors of microwave radiation, and by extension, high power klystrons. After
the war, many of these devices could be acquired almost free of charge from the
depots the U.S. War Department had set up to make available the surplus
equipment to universities. The availability of these new pieces of equipment and
of the new statistical methods to detect signals in the background of noise created
new levels of precision and reliability in experiments, new standards in the
reproducibility of experiments, and thus a new standard for experimental practice,
a practice reinforced by the often parallel, competing experimentations.
190 S. S. Schweber Phys. Perspect.

Moreover, it should be pointed out—as Forman did—that although physicists


in the United States had developed many of the new instrumentalities, they were
not the first to apply them and reap the benefits from them. Their applications to
accelerators took place first in Great Britain;46 radio astronomy was initiated in
Australia.47
These instrumental advances had immediate repercussions in theoretical
physics. The best-known and one of the most consequential was the response to
Lamb and Retherford’s experiment on the fine structure of hydrogen and to Rabi,
Nafe, and Nelson’s accurate measurements of the hyperfine structure of hydrogen
announced at the Shelter Island Conference in June 1947.48 These experiments
stimulated crucial calculations49 by Bethe and by Julian Schwinger that initiated
the modern renormalization program and gave renewed faith to quantum field
theory. The ‘‘modern era’’ of quantum field theory was initiated by that confer-
ence. Steven Weinberg assessed its importance concisely: ‘‘It was not so much that
it forced us to change our physical theories, as it forced us to take them
seriously.’’50
At the Shelter Island Conference,51 Hendrik Kramers made a key presentation
of his formulation of the Lorentz theory of an extended charge in which structural
effects were encapsulated in the experimental mass of the particle. He indicated
how to reinterpret the formalism so as to obtain finite answers when self-inter-
actions are taken into account. Max Dresden in his biography of Kramers52
suggested that Kramers did not receive adequate credit for his contributions at
Shelter Island. Bethe’s Shelter Island notes indicate that he was right.53 To Bethe,
the pragmatist for whom numbers were always the criterion of good physics and
who had just been so deeply and successfully involved in the war effort calculating
numbers that translated into physical effects and measurable empirical data, the
challenge was to get the numbers out and account for the magnitude of the 2S–2P
level shift in hydrogen. Accounting for the empirical data would be explaining the
data. Perhaps one reason that Bethe did not acknowledge Kramers was that
Kramers’s approach was too model-dependent, too theoretical, and too far
removed from calculating numbers. For Bethe, the value of a novel idea was
gauged by whether it could help you calculate numbers that could be compared
with empirical data.
Bethe did have a much simpler and straightforward way than Kramers to
incorporate Kramers’s insight. He had noted that the quantum-electrodynamically
calculated self-energy of a free non-relativistic electron could be ascribed to an
electromagnetic mass of the electron, which, though divergent, should be added to
the mechanical mass of the electron. The only meaningful statements of the theory
involve the sum of the electromagnetic and mechanical masses, which is the
experimental mass of a free electron. In contrast to Kramers’ approach, Bethe’s
model-independent formulation of mass renormalization did not assume an
extended charge distribution for the electron. In contrast to Schwinger and
Weisskopf’s initial insight that a hole-theoretic calculation that computed the
Vol. 16 (2014) Writing the Biography of Hans Bethe 191

difference between the energies of two levels would be finite, Bethe formulated an
unambiguous prescription of mass renormalization in the non-relativistic case that
allowed computing the energy of each level.
After the conference was over, Bethe performed his famous non–relativistic
calculation on the train ride from New York to Schenectady. The paper in which
he proved that the level shift would be accounted for quantum electrodynamically
was completed three days after the conference ended and thereafter circulated to
the participants of the Shelter Island conference.

Post-World War II Theoretical Physics


After World War II, the practice of theoretical physics dealing with ‘‘elementary
particles’’ became much the same over much of the western world. The following
are some of the factors responsible for this homogenization:
1) The mimeograph machine and the ditto machine (with the characteristic
smells of their inks) had come into their own. They allowed the cheap and
rapid reproduction of lecture notes and preprints of papers. Thus, two
separate sets of notes of the lectures Schwinger delivered at Harvard on
nuclear physics during the academic year 1946-7 were written up and issued in
mimeographed form by John Blatt and by Morton Hamermesh. Both sets
were widely distributed and became the basis of numerous courses on
quantum mechanics all over the United States. The same was true of
Feynman’s 1949 Cornell lectures on advanced quantum mechanics and
Dyson’s 1951 Cornell course on quantum field theory.54 Furthermore, and
most importantly, the ditto machine allowed preprints of papers to be sent to
physics departments all over the world at nominal cost— a process that the
internet and the preprint depository arXiv now do much more efficiently and
democratically.55
2) The broad governmental support given after the war and during the Cold War to
the sciences (physics in particular) in the US, Europe, and the Soviet Union was
certainly one of the enabling conditions responsible for the important,
transformative, advances in physics from 1945 to 1955. In the US, governmental
funding made possible much larger physics departments in which a wide
spectrum of theoretical and experimental research activities were undertaken
and that supported greatly increased numbers of graduate students, research
associates, and postdoctoral fellows.56 A similar expansion of physics activities
took place in the Soviet universities and research institutes.
Within this widely expanding doctoral and postdoctoral framework of physics
education and overseen by a limited number of faculty members able to transmit
the recent advances, summer schools became important institutions that allowed
graduate students, postdoctoral fellows, and faculty members—regardless of their
institutional affiliation—to learn about recent developments in physics. Soon after
192 S. S. Schweber Phys. Perspect.

World War II ended, the Michigan Summer Symposium resumed the important
role it had played during the late 1920s and the 1930s when the most prominent
physicists lectured on the latest advances in theoretical and experimental physics
to graduate students from all over the US. In 1948, 1949, and 1950, Schwinger,
Feynman, and Dyson, respectively, lectured there on their researches in quantum
electrodynamics (QED). Mimeographed copies of their lectures became imme-
diately and widely available. The French summer school Les Houches, located
near Chamonix in the Alps, opened its doors in 1950. The Cargèse Summer
School, in Corsica, began its operation in 1951. In Italy, the International School of
Physics ‘‘Enrico Fermi’’ started holding annual summer schools in Varenna in
1953. In 1957, Brandeis University and the University of Colorado in Boulder
started their summer school in theoretical physics.57 The proceedings of all these
summer schools were published promptly and constituted comprehensive, valu-
able introductions to the advances in these fields. Their quick availability—in
mimeographed form until 1960—shaped the teaching of graduate courses in
theoretical physics all over the world.58
The French and Italian summer schools of the late 1950s and early 1960s played
another important role: they brought outstanding Soviet theorists to lecture and
thus informed their audiences of the important Soviet contributions to condensed
matter physics and introduced the Soviet physicists to their counterparts in the
West. Thus, the 1958 Les Houches Summer School devoted to the ‘‘Many Body
Problem’’ included as one of the lecturers the Soviet theorist Spartak Belyaev—an
important contributor to the introduction of the Feynman diagrammatic and the
Schwinger field-theoretic methods to solid state and nuclear physics.59
3) When in the fall of 1946 Robert Oppenheimer became the director of the
Institute for Advanced Study in Princeton (IAS) and a professor of physics
there, he used the institution to implement his universalist vision of science.60
In 1948, he began inviting a host of foreigners to the Institute. Until 1952,
these included: Faqir Auluck, Aage Bohr, Freeman Dyson, Léon van Hove,
Res Jost, Nicolaas van Kampen, Toichiro Kinoshita, Maurice Lévy, Cécile
Morette, Yoichiro Nambu, Abraham Pais, Giulio Racah, Abdus Salam, Sin-
Itiro Tomonaga, Hideki Yukawa, John Ward, and Chen Ning Yang.61 The
appointment of Cécile Morette in 1948 and that of Maurice Lévy in 1950 were
particularly consequential. The success of the postwar Michigan summer
school led Cécile Morette to have a similar school established in France. In
1950 she founded the Summer School of Theoretical Physics in Les
Houches62; Maurice Lévy organized the school in Cargèse in 1951. These
two summer schools were responsible for teaching the postwar generation of
French physicists modern quantum theory63 and quantum field theory.
Bohr’s institute in Copenhagen was another institution that hosted and brought
together theorists from various nations, including the Soviet Union. These efforts
Vol. 16 (2014) Writing the Biography of Hans Bethe 193

were greatly amplified by the European decision to built CERN and the basing of
Nordita, its theoretical division, at the Bohr Institute until CERN began operating
in Geneva in 1957.64

The Mathematical Physics Discipline


In his 2012 paper on the transition between modernity and postmodernity, Forman
noted that: ‘‘By the middle of the twentieth century disciplinary production and
curation of knowledge would become the most honored and emulated manifes-
tation of modernity’s high valuation of discipline—a valuation that had been rising
and spreading through Western societies for a full four centuries.’’65 The formation
of the modern, postwar discipline of mathematical physics had its beginning at the
IAS shortly after Oppenheimer became its director. In 1948, three young
theorists—Léon Van Hove, Res Jost, and Arthur Wightman—attended the
lectures on C* algebras that John von Neumann was delivering at the Institute.
Van Hove and Jost were fellows at the IAS; Wightman was at the time an
instructor at Princeton University. Together with Cyril Domb, Rudolf Haag,
Daniel Kastler, Arthur Jaffe, and others, they went on to establish mathematical
physics as a recognized, institutionalized, sub-discipline of physics and mathemat-
ics. The discipline of mathematical physics—with its demand of rigor and proof—
may indeed serve as a paradigmatic illustration of the role played by disciplinarity
in modernity, one of the central issues Forman addressed.
Roughly speaking, the difference between mathematical physicists and theo-
retical physicists is the following: theoretical physicists are interested in obtaining
explanations of physical phenomena, and in particular in accounting for exper-
imental data. They value convincing intuitive arguments in favor of the
interpretation of some physical fact. Mathematical physicists, on the other hand,
are not primarily interested in ‘‘getting the numbers out.’’ Rather, their aim is to
construct a mathematically rigorous, consistent representation of the phenomena
based on physical postulates and models. Their work is to meet the highest
standards of mathematical rigor. But a mathematically rigorous proof need not
mirror physical intuition, nor be generative of further physical questions.
Traditions of mathematical physics had been in existence for a long time with
its practitioners usually housed in mathematics departments. The one initiated by
Klein, Minkowski, and Hilbert at Göttingen66 had been seeded in the United
States, primarily at Princeton, through the immigration of Hermann Weyl, von
Neumann, and Wigner in the early 1930s. With Hitler’s coming to power and the
subsequent expulsion of Jewish scholars from their university posts, the Göttingen
tradition became transplanted to the United States, primarily in what became the
Courant Institute at New York University in New York. Another tradition,
principally concerned with the mathematics of classical mechanics67 (out of which
grew ergodic theory and rigorous probability theory) had been nurtured in France
and later blossomed in the Soviet Union.68 In Japan, Ryogo Kubo, Tosio Kato,
194 S. S. Schweber Phys. Perspect.

and others initiated a mathematical physics tradition concerned with statistical


mechanics and operator theory.
Here, I am concerned with the prehistory of mathematical physics as a
discipline to the extent of being able to characterize the focus and nature of the
researches carried out in various countries in the post–World War II period.69 In
the United States, the researches of Wigner and Valentine Bargmann on the
unitary representations of the Galileo and Lorentz groups greatly influenced the
works of Wightman and Jost. In Great Britain, Lars Onsager’s solution of the two-
dimensional Ising model stimulated the researches on phase transitions carried out
by Cyril Domb and his students. In the Soviet Union, the agenda was shaped by
the important advances made in probability theory, statistical mechanics, and
turbulence by Alexandr Khinchin, Andrey Kolmogorov, Nikolai Krilov, and
others during the 1930s.70 Nikolay Bogoliubov, Israel Gelfand, Mark Naimark,
and others invigorated the tradition during the 1940s and trained outstanding
students such as Yacov Sinai, Roland Drobushin, Vladimir Arnold, and Ludvig
Faddeev in the early 1950s. The foci of their researches—dynamical systems,
approach to equilibrium, chaos, turbulence—became important components of
mathematical physics.71
The young mathematical physicists who would mold the agenda in the West—
Wightman, Jost, Haag, van Hove—were initially differentiated from their Soviet
counterpart by their concern with relativistic quantum field theory, undoubtedly
stimulated by what had been presented at the Shelter Island and Pocono
conferences.72 All four had had close contacts with someone who had been
associated with Hilbert, Sommerfeld, or both, and thus had been touched by the
spirit of ‘‘the pre-established harmony of physics and mathematics.’’73 All four had
majored in mathematics as undergraduates and all four kept in close contact with
mathematicians, but held appointments in physics departments.74
What van Hove, Jost, Wightman, and Haag were able to accomplish in the
1950s was to transform what had been primarily individual activities into a
flourishing discipline. Rather than being concerned with problems close to applied
mathematics, applied science, engineering, and technology as was the case in the
Soviet Union, they and their students worked on problems in quantum mechanics
(rigorous definition of the operators appearing in its formulation, scattering
theory, existence of bound states, quantum logic, …), in quantum field theory
(axiomatic and latter constructive field theory, algebraic approaches,…), and in
statistical mechanics (thermodynamic limit, …); all subjects of great interest to
physicists.75 Streater and Wightman’s PCT, Spin and Statistics, and All That, Res
Jost’s The General Theory of Quantized Fields, and David Ruelle’s Statistical
Mechanics: Rigorous Results (1969) are the paradigmatic examples of this
approach.76
The growing influence of the discipline is made explicit by the fact that in 1960
the American Institute of Physics started publishing the Journal of Mathematical
Vol. 16 (2014) Writing the Biography of Hans Bethe 195

Physics devoted exclusively to mathematical physics. In 1965, Springer Verlag


began publishing Communications in Mathematical Physics, a journal committed
to even more rigorous expositions than the Journal of Mathematical Physics.
Similarly, by the mid-1960s several summer schools in theoretical physics devoted
part of their offerings either to axiomatic field theory or to rigorous results in
statistical mechanics.
Mathematical physicists made important contributions to the elucidation of
renormalization procedures in quantum field theory, the establishment of the
notion of an effective field theory, the clarification of spontaneous symmetry
breaking, and most importantly, to the explanation of the stability of matter.77
String theory, with its seminal contributions to pure mathematics78 and its
influential role in providing mathematical techniques for solving strong coupling
problems in quantum field theory when applied to condensed matter physics
problems,79 is indicative of the latest influence of the discipline.
At the 1971 Battelle Seattle Rencontre, Andrew Lenard, an important
contributor to the discipline, commented that ‘‘the maturing and growth of
modern mathematical physics is one of the striking intellectual developments of
the last two decades. Even more importantly, mathematical physics fosters a
cooperative and unifying spirit between practitioners of different areas of
expertise.’’80

The Self-Image of Physicists81


The contribution of physicists to the war effort was an important factor in bringing
about the homogenization of physics after the war. The war had engendered a
symbiotic relationship between physicists and the state. The Cold War intensified
this entanglement. The sciences, and physics in particular, became a means of
nurturing and displaying national greatness. Accelerators in the West came to
assume the role that cathedrals had played earlier, as did rockets in the Soviet
Union. Physicists readily accepted their new role in the partnership with the state,
and their hubris expanded accordingly.
For a while after the war, physicists saw themselves as being engaged in a
collective endeavor to discover if not the ultimate, then what seemed to be
immutable laws of physics and their consequences. All the subdisciplines of
physics—solid state physics, nuclear physics, cosmic ray physics—were considered
part of this epistemological enterprise.82 But the further chilling of the Cold War,
the Korean war, the subsequent lavish governmental support of physics, the
freeing of high energy physics from its dependence on cosmic rays for the
interpretation of high energy processes by virtue of the operation of new high
energy accelerators, the importance of solid state physics stemming from the
‘‘transistorization’’ of electronics, the introduction of efficient, cheap new modes of
computation, and the growing size of these subdisciplines all contributed to the
196 S. S. Schweber Phys. Perspect.

restructuring of physics into fairly well delineated, somewhat separated,


subdisciplines.
Bethe initially throve in the new world that had emerged after World War II
and made important contributions to quantum electrodynamics and nuclear
physics. His Lamb shift calculation was one of the high points of the early phase of
his postwar career, and the Matthew effect* magnified the importance of his
contribution as time went on, but the new environment also accentuated his
limitations.83 Julian Schwinger’s appointment as professor of physics at Harvard in
1946, rather than Bethe, was indicative that a new generation was taking over.84
Great rewards were to be gleaned from determining what were the founda-
tional theories describing the various subnuclear domains being unraveled by the
new synchrotrons and linear accelerators. As mentioned, Bethe had been
instrumental in establishing at Cornell the laboratory devoted to high energy
studies. With the appointment in early 1948 of Robert R. Wilson as its director, the
Newman Laboratory flourished and became an outstanding high energy labora-
tory. In 1950, its 300 MeV electron synchrotron began its operation, and Bethe
tried to mold himself into a high energy physicist. In spring and summer 1952, he
lectured on high energy physics at Los Alamos85 and during the academic year
taught graduate courses on the subject at Cornell. Bethe’s efforts with Dyson and
others from 1952 to 1954 to interpret the data that Fermi had obtained on the
scattering of pi mesons on nucleons met with limited success.86 Two younger
physicists, Geoffrey Chew and Francis Low, formulated a much more powerful
and generative approach.87 Once again it was brought home to him that he was
much better at applying foundational theories than at creating them even though
he had recognized very early that what were considered to be fundamental
theories—quantum electrodynamics, the various meson theories, the Fermi theory
of weak interactions, and Einstein’s general theory of relativity—were only
approximate theories, foundational88 for a particular class of phenomena delimited
by range on the energy scale or by the model or idealization underlying them.
The failure of Bethe and Dyson’s approach to pion-nucleon scattering as
compared to Chew and Low’s was the culmination of the ‘‘difficult’’ first half of the
1950s.89 During these years, Bethe had had to deal with McCarthyism; with the
moral and technical aspects of building hydrogen bombs;90 with the Oppenheimer
affair; and with the dramatically accelerating pace of research, particularly in
quantum field theory and in high energy physics. By the end of 1954, he recognized
that he no longer could keep up with the pace of developments in these fields,
given his other involvements and commitments. Important—but not totally

* The sociologist Robert K. Merton observed that if two scientists arrive at similar con-
clusions, the more eminent or famous scientist will often receive more credit. Merton coined
the term ‘‘Matthew Effect’’ to describe the phenomenon, from Jesus’ parable of the talents
as told in Matthew 25:29: ‘‘For unto every one that hath shall be given, and he shall have
abundance: but from him that hath not shall be taken even that which he hath.’’
Vol. 16 (2014) Writing the Biography of Hans Bethe 197

convincing—work in the nuclear many-body problem had then just been initiated
by Keith Brueckner.91 Given the richness of the available post-war experimental
data on nuclei and their structure, Bethe decided to concentrate his research
efforts in nuclear physics. From the mid ‘50s until the early ‘70s, the nuclear many-
body problem was the main focus of Bethe’s researches and that of his students.
In his 1937 Reviews of Modern Physics articles on nuclear structure,92 Bethe
had attempted to understand why some properties of nuclei showed discontinuities
when their protons or their neutrons numbered 2, 8, 50, and 82, just as was the case
with the binding energies of the noble gases in atomic physics.93 These proton and
neutron numbers were called ‘‘magic numbers.’’ In 1947, Maria Goeppert-Mayer,
Otto Haxel, Hans Jensen, and Hans Suess had given a phenomenological
explanation for the shell structure of nuclei, more specifically for the appearance
of these magic numbers, on the basis of an independent particle model.94 To
establish the shell model, one needed to understand why it is possible to consider
the nucleons moving in independent, single-particle orbitals and how such a
viewpoint could be reconciled with the compound nucleus model of Bohr. In 1954,
Keith Brueckner took the main step in this direction by applying techniques that
Kenneth Watson had developed in scattering theory. Brueckner and his
collaborators elaborated the Watson multiple-scattering formalism into a powerful
‘‘self-consistent’’ approach to handle many-body problems. They had difficulties,
however, giving adequate proofs for the validity of the results they had obtained.
As Bethe told Charles Weiner in his interview with him on May 8, 1972: ‘‘It
seemed to work all right. It seemed to be plausible, but it had certain very definite
flaws.’’95 Although Bethe was not part of the mathematical physics community,
Sommerfeld had impressed on him high standards of rigor and of mathematical
consistency. Removing the flaws in Brueckner’s theory became the first problem
he tackled. Bethe carefully went over all of Brueckner’s papers, and gave a more
transparent, lengthy exposition of Brueckner’s approach, but one that still did not
fully justify the method. He spent a sabbatical academic year 1955–6 in Cambridge
concentrating on the justification of Brueckner’s theory. He was given two
outstanding students to work with him: Jeffrey Goldstone and David Thouless. In
the first of his lectures on Brueckner theory, Goldstone asked some pertinent
questions that Bethe could not answer. Bethe put him to work on the theory and
the result was that, shortly thereafter, Goldstone and Bethe solved the problem on
how to handle the incorporation into the formalism of the effects of the hard-core
repulsion between nucleons at very short distances, which was part of the assumed
interaction potential.96 Thouless went back with Bethe to Cornell and obtained his
PhD there in 1958, writing an important thesis that described how to calculate two-
body correlations self-consistently.97
Goldstone was one the first theorists to introduce diagrams into many-body
theory with his 1957 paper.98 1957 was a banner year for many-body theory, with
the Bardeen, Cooper, Schrieffer (BCS) explanation of superconductivity in lead
198 S. S. Schweber Phys. Perspect.

the outstanding advance 99 1957 also witnessed the experimental confirmation that
parity was not conserved in the weak interaction, as had been intimated by Lee
and Yang.100 By virtue of the conference that Louis Michel organized in Lille on
Les problèmes mathématiques de la théorie quantique des champs, 1957 could also
be considered the year that the mathematical physics discipline was founded.101
October 1957 marked the launching of Sputnik, which had wide repercussions in
the United States. It was responsible for the creation of PSAC and for sizable
increases in the government’s budgets underwriting research and development in
the sciences and in technology, as well as the funding of science education.102
BCS became the point of departure for a reconceptualization of quantum field
theory. It indicated that in systems with an infinite number of degrees of freedom,
the state of lowest energy—the ground state—need not possess the (continuous)
symmetry exhibited by the Hamiltonian that determines the dynamics of the
system. This is what is meant by a ‘‘spontaneously broken symmetry.’’ The role
that symmetry plays in quantum field theory was enlarged and greatly extended by
BCS’s theory. This was the point of departure for establishing quantum field
theories as the appropriate formalism for the representation of all the foundational
theories as ‘‘effective field theories’’ describing the microscopic world down to
distances of the order of 10-17 cm.103

Effective Field Theories


An effective field theory is a description of phenomena in a certain energy regime
bounded by some energy (K) and in terms of wavelengths longer than some length
scale (h/K). An effective field theory assumes that the physics in the domain in
which it is valid can be given in terms of ‘‘elementary’’ entities out of which the
composite entities that populate that domain are built. These ‘‘elementary’’ entities
constitute the effective degrees of freedom appropriate to that scale. They depend
on a more ‘‘fundamental’’ theory only through a small set of parameters that enter
in the description of the dynamics of these entities.104
Relativistic quantum field theories implicitly make statements about arbitrarily
short space–time distances, and thus about arbitrarily high energies and momenta,
yet no conceivable experiment will be able to probe such distances. When
calculating the predictions of a given theory, the contributions stemming from
these high-energy, short-distance components are divergent, implying infinite
results. The perturbative renormalization program to circumvent these diver-
gences (developed by Weisskopf, Bethe, Schwinger, Feynman, and Dyson after
World War II) was given a new and deeper interpretation by Kenneth Wilson and
others105 in the early 1970s. Wilson was able to exhibit the effect of all possible
modifications of a given theory beyond a certain cut-off energy as a re-
parametrization of all possible interactions between the entities that are assumed
to populate the low energy domain of that theory. Furthermore, he showed that,
starting from any set of interactions at the cut-off scale, a low energy physics
Vol. 16 (2014) Writing the Biography of Hans Bethe 199

description at a given level of accuracy could be formulated that depended only on


a few relevant parameters.
The great accomplishment of Wilson, Weinberg, and others was to demonstrate
the universality of the low energy physics that resulted from the renormalization
process, thus justifying the use of effective field theories. There is an important
consequence of the justification of using effective field theories: as long as one
does not probe beyond the energy and length scales in which they are deemed
applicable, the description of the physics in their domain is not invalidated by
discoveries at lower levels. The issue becomes answering the question: ‘‘To what
extent can we reconstruct the world knowing the most ‘fundamental’ effective
theory now known?’’ That issue was addressed Philip Anderson in an important
article in 1971 entitled ‘‘More is Different.’’106 Knowledge of the foundational
theory does not imply that one can calculate or predict with the theory all the
possible stable or quasi-stable structures that can be created by composition.
There are limits imposed by computational complexity107 and limits in trying to
translate the mathematical language of the foundational theory into a vocabulary
adequate to describe geometrical conformations or informational transfers.

Epochal Break, Post-Modernity, and All That


It has been generally recognized that since the 1980s the sciences and technology
have undergone a profound transformation, reflecting, on the one hand, deeply
changed economic, political, and cultural contexts—and as a consequence the
nature and sources of the financial support of scientific and technological activi-
ties—and on the other, a changed view of what constitute foundational theories, as
well as vastly improved computational powers.108 Forman’s recent work is an
attempt to characterize and assess this transformation.109
The explanation for the multifaceted tranformation have for the most part been
concerned with the political, economic,110 and cultural factors, with less attention
paid to cognitive factors internal to the various scientific and engineering disci-
plines, except for molecular biology and biotechnology. In the physical sciences,
advances in quantum field theory in the mid-1970s justified a view of the physical
world that segmented it into different levels, each having a foundational theory—
its ‘‘effective’’ theory—that describes the dynamics of the entities that populate
that level. These ‘‘effective’’ (field) theories explicitly take into account the range
of energies in which the described systems can be probed and therefore the
accuracy with which they can be represented. The excluded high-energy, small-
distance effects are taken into account by appropriate local interactions that are
calibrated by experimentally determined parameters. As noted above, what is all-
important is that this highly accurate, robust, and stable description of the
microscopic world is not destabilized by the incorporation of new small-distance
effects provided by higher-energy—smaller-length-scale—findings.
200 S. S. Schweber Phys. Perspect.

What this implies is that a form of finalization has been given to the foundational
theory that describes atoms, molecules, and solids which then allows various highly
accurate models to be given for such entities as metals, insulators, superconductors,
superfluids, ferromagnets, liquid crystals, two-dimensional graphite systems, optical
lattices,…. People working in condensed matter physics, in photonics, in nanotech-
nology, on quantum computers,… are principally concerned with the creation of
novelty—of entities or effects that did not previously exist in the world—or concerned
with understanding the complexity and diversity that can emerge from composition and
are no longer concerned with establishing the foundational theory that governs the
interactions and determines the evolution of the structures that populate that domain.
Thus, except for a very small component of the practitioners in these fields, the agenda
is set by external factors, by the demands of specific novelty and complexity, by use-
fulness or efficiency, or by expectations—as is seen in nanotechnology, photonics, and
quantum computation. It has become difficult to differentiate these activities from
applied science, and in many cases from research and development in technology. In
the physical sciences, the robustness and the precision of the foundational theories at
the micro- and sub-microlevels are surely part of the reasons that the Bayh-Dole
legislation has had such consequential impact on the restructuring of universities since
the 1980s.111 Just as physics has been transformed, so has chemistry. Undoubtedly the
biological and medical sciences have been most deeply affected by the technical
advances in them: Crick and Watson, genetic codes, recombinant technologies, DNA-
sequencing, genome projects, bioinformatics,… And it is in the biological sciences that
the entrepreneurial aspects of the university are most visible.112
Forman has repeatedly emphasized that how the sciences evolved and
expressed themselves was deeply conditioned by their cultural and social context.
Although the evolution of science cannot be inferred by extrapolating current
scientific concepts, Forman believes that it can be predicted ‘‘to some extent by
considering the general social and cultural conditions under which scientific
knowledge is being produced at present and is likely to be produced in the
future.’’113 Given the robustness of foundational theories and of the institutional
frameworks in which they are created, Forman’s assertion seems to be valid at
present. It seems likely that neoliberalism will continue to help shape the policies
that govern the support of Western science and that corporate and private support
will deeply affect the agenda of the science created there.
Like all his other historical works, Forman’s recent articles on the deep structural
change we are witnessing are important, arresting pieces.114 These writings constitute a
new Forman thesis that makes a case for the primacy of science in modernity and of
technology in postmodernity, arguing that modernity entailed disciplinarity, postmo-
dernity antidisciplinarity. This new thesis is global in scope. Like his long articles on
acausality and Weimar Culture and those on the intellectual agenda of science in the
US during the Cold War, these essays are generative, influential, and controversial.
As Forman has himself emphasized, his approach is exclusively cultural. Thus
he demarcated postmodernity from modernity by ‘‘the abrupt reversal of culturally
Vol. 16 (2014) Writing the Biography of Hans Bethe 201

ascribed primacy in the science-technology relationship—namely, from the pri-


macy of science relative to technology prior to circa 1980, to the primacy of
technology relative to science since about that date.’’ At issue in this categorization
is not the ‘‘actual, factual relations between science and technology, but only the
putative, culturally presumed relations.’’115 Forman does not address what today in
postmodernity technology owes to science, and science to technology, nor what
was the case four or five decades ago or a century ago in modernity, except to
suggest that altered cultural presuppositions will likely continue to affect science in
practice. Forman emphasizes that the cultural valuation of science was held high in
the past by modernity’s ‘‘elevation of the public over the private and, more
importantly still, the belief that the means sanctify the ends, that adherence to
proper means is the best guarantee of a ‘truly good’ outcome.’’ In postmodernity,
on the contrary, the approach is top-down: ‘‘technology is the beneficiary, and
science the maleficiary, of our pragmatic-utilitarian subordination of means to
ends, and of the concomitants of that predominant cultural presupposition,
notably, disbelief in disinterestedness and condescension toward conceptual
structures.’’116
Even though Forman focuses primarily on the cultural dimension, he is,
nonetheless, open to some qualifications. Traditionally, technology is intrinsically
top-down in contrast to the traditional bottom-up approach of the sciences.
Yet already in the 1960s, the microchip technology had assumed a bottom-up
stance, and in recent years materials engineering and nanotechnology are
assuming a similar position and see themselves as scientific disciplines because of
their bottom-up approach. Bioengineering likewise, sees itself as a scientific dis-
cipline because of this. In fact, it is more accurate to characterize these fields as
interdisciplinary hybrids of science and engineering.117
Concerning economic factors, there seems to be a consensus that neoliberalism has
played a crucial role in effecting the profound transformation we are witnessing. I
want to sketch—painting with very broad brushes—a parallel between the Forman’s
Weimar and postmodernity theses. I would like to suggest that after 1975 in the
United States the Vietnam War and the student protests of the 1960s and early ‘70s
played the roles that World War I and the liberating Weimar culture did in Germany.
If acausality was at the center of Forman’s Weimar paper, I would propose that
information, uncertainty, and risk—the characteristics of the postmodern world—
and their omnipresence, conception, and management—by whom and by what
means—are central to the new Forman thesis. Similarly, Friedrich von Hayek and the
thought collective that formulated neoliberalism play the roles in postmodern times
that Spengler’s Decline of the West and Lebensphilosophie* did in the Weimar period.

* Lebensphilosophie (‘‘philosophy of life’’) emerged in Germany during the nineteenth


century as a reaction to the emphasis on science and rationalism in philosophy. It empha-
sized the meaning and purpose of life and had a anti-rational, Romantic component in its
outlook.
202 S. S. Schweber Phys. Perspect.

Neoliberalism entails a top-down view of the world as part of its metaphysics,


with ‘‘top-down’’ resonating with the primacy of ends over means. Robert
Laughlin’s recent book, A Different Universe: Reinventing Physics from the Bot-
tom Down illustrates that this ideology has infected physics as well. Laughlin is a
condensed matter theoretical physicist who, together with Horst Störmer and
Daniel Tsui, won the Nobel Prize for explaining the fractional Hall effects dis-
covered by Klaus von Klitzing. The message of Laughlin’s book is that
emergentism has triumphed over reductionism, with emergentism understood as
asserting that the foundational theory of a given level cannot be deduced from that
of a lower level, that the ‘‘reality’’ of a given level is more than the sum of the
constituent elements of a lower level. Its central dogma is that macroscopic objects
are the products of principle of organization and of collective behavior that cannot
be reduced to the dynamics of their ‘‘elementary’’ constituents. Although I do not
dispute that some form of emergentism has become part of the metaphysics of
physics, one of the issues involved becomes what is taken as the foundational
theory for a given level.118
Implementing von Hayek’s neoliberalism became part of Margaret Thatcher’s
and Ronald Reagan’s political agenda. Their election transformed the role of the
government in the UK and the US. Reagan’s election had further profound
consequences. I conclude addressing some of these by returning to Bethe.

The End
Bethe’s life ended on a somber note. Throughout his life, he had felt a heavy
responsibility for his contributions to the creation of nuclear weapons; he invested
huge efforts to constrain their developments and make nuclear energy have
peaceful applications. He found justification and consolation for his participation
in the development of fission and fusion bombs in the fact that the Soviet Union
and the United States did not come to blows in the face of many provocations after
the Cuban missile crisis of 1962. MADness had worked.
Ronald Reagan changed that. In March 1982, when asked during a news con-
ference whether a nuclear war was winnable, Reagan replied that there could not
be any winners in such a war: ‘‘everybody would be a loser.’’ Yet, acting at Rea-
gan’s request, Caspar W. Weinberger, his Secretary of Defense, and the Pentagon
were at that very moment drawing up plans for the possibility of waging a pro-
tracted nuclear war. On May 20, 1982, Reagan signed the National Security
Decision Directive (NSDD) 32 stating that ‘‘the United States will enhance its
strategic nuclear deterrent by developing a capability to sustain protracted nuclear
conflict,’’ and that ‘‘the modernization of our strategic nuclear forces … shall
receive first priority.’’ Further, on January 17, 1983, he signed NSDD-75, which,
while stressing deterrence, promoted a massive nuclear build-up so that the out-
come of a nuclear war with the USSR would be so devastating to it ‘‘that there
would be no incentive for Soviet leaders to initiate an attack.’’119 These directives
Vol. 16 (2014) Writing the Biography of Hans Bethe 203

led to regard the doctrine of Nuclear Utilization Target Selection (NUTS) as a


viable alternative to the dogma of Mutually Assured Destruction (MAD).120
NUTS gave rise to considerations of the possibility of making preemptive attacks
on the silos housing the Soviet nuclear weapons in order to destroy the weapons
before they could be used. To be successful, a preemptive strike requires the
possession of an overwhelmingly greater number of highly accurate nuclear
weapons than the enemy, and the US seemingly embarked on implementing this
viewpoint. The US commitment to NUTS can be inferred from the adoption at the
time of a number of first-strike weapons, e.g., the Trident II and Minuteman III
nuclear missiles, both highly accurate and able to destroy an enemy missile silo if
so targeted. An additional indication of the extent to which NUTS became policy
was the development of stealth bombers capable of carrying a large numbers of
cruise missiles, missiles that can likewise be made ‘‘stealthy’’ and thus able to evade
enemy radar detection. Moreover, cruise missiles, by virtue of the earlier suc-
cessful design of small fusion bombs, could be ‘‘nuclear tipped.’’121
Reagan’s massive arms build-up reached its climax in March 1983 when he
proposed his Strategic Defense Initiative (SDI), an anti-ballistic defense system
that would use ground and space-based X-ray lasers to destroy incoming Soviet
missiles armed with nuclear warheads. SDI focused on strategic defense and was to
complement the strategic offense doctrine of NUTS. Shortly after hearing Rea-
gan’s speech, Bethe went to Livermore to assess the capabilities of the X-ray laser
developed there, on which the SDI proposal was based. Though impressed by the
novelty of the X-ray laser, Bethe went away ‘‘highly skeptical [that] it would
contribute anything to the nation’s defense.’’122 He believed that a laser defense
shield was not feasible. To develop, build, and operate any defensive system was
very costly and easily rendered ineffective. Thus, the Soviets could easily send
thousands of cheap decoy missiles to overwhelm the defensive system during a
nuclear attack. Furthermore, Bethe believed that an anti-ballistic defense shield
would be viewed as a threat by the Soviets because it would limit Soviet offensive
capabilities but leave American offense capabilities intact. Bethe was convinced
that the only way to stop the threat of nuclear war was through diplomacy; he did
not believe that a technical solution to the Cold War existed. In these activities, he
was helped keeping abreast of developments by his colleague and close friend,
Kurt Gottfried—one of the organizers with Henry Kendall of the Union of
Concerned Scientists in May 1968—and by other colleagues at Cornell and else-
where, such as Richard Garwin, Henry Kendall, Franklin Long, and Judith Reppy.
Thus, in March 1984 Bethe helped write a 106-page report on SDI issued by the
Union of Concerned Scientists,123 whose conclusion was that ‘‘the X-ray laser
offers no prospect of being a useful component in a system for ballistic missile
defense.’’ Similarly, in 1988 he participated in the study conceived and initiated by
Paul Bracken, Richard Garwin, Kurt Gottfried, and Henry Kendall that examined
‘‘the role of crisis as a precursor to nuclear war and the extent to which the US and
the USSR could maintain control over such a chain of events.’’124 These are but
204 S. S. Schweber Phys. Perspect.

two among the many papers he wrote and the many discussions in which he
participated regarding nuclear weapons and their containment, along with the
prevention of nuclear war.125 Bethe’s anxieties concerning nuclear weapons cul-
minated in 1995 on the occasion of the fiftieth anniversary of the leveling of
Hiroshima with an appeal to scientists to take a Hippocratic oath not to work on
weapons of mass destruction:
As the Director of the Theoretical Division of Los Alamos, I participated at the
most senior level in the World War II Manhattan Project that produced the first
atomic weapons.
Now, at age 88, I am one of the few remaining such senior persons alive.
Looking back at the half century since that time, I feel the most intense relief
that these weapons have not been used since World War II, mixed with the
horror that tens of thousands of such weapons have been built since that time—
one hundred times more than any of us at Los Alamos could ever have
imagined.
Today we are rightly in an era of disarmament and dismantlement of nuclear
weapons. But in some countries nuclear weapons development still continues.
Whether and when the various Nations of the World can agree to stop this is
uncertain. But individual scientists can still influence this process by with-
holding their skills.
Accordingly, I call on all scientists in all countries to cease and desist from
work creating, developing, improving and manufacturing further nuclear
weapons—and, for that matter, other weapons of potential mass destruction
such as chemical and biological weapons.126
After George W. Bush became president in 2001, Bethe became additionally
concerned that independent scientific and technological advice was playing an
ever-smaller role in governmental policies. More specifically, he became very
disturbed by the actions of the Bush administration in disbanding many govern-
mental scientific advisory bodies and replacing a large fraction of the members of
the still-existing ones with people who were either drawn from the industrial
scientific community, whom he thought were less independent than scientists in
the academy or whom he believed were ideologically committed to the Bush
policies regardless of the scientific facts. With profound anguish, he observed the
paths taken by the Bush administration when addressing issues relating to nuclear
weaponry, test-ban treaties, the environment, and the dramatic increase in the
information it stamped secret. Secrecy prevented people from knowing. Only if
they had knowledge could they act rationally—and rationality was essential to
Bethe. He decried the Bush administration’s involvement in Iraq and the secrecy
involved in the justification for the military actions taken. And he lamented the
fact that the Bush administration was giving political and military considerations
priority over all other factors, including scientific realities, at a time when science
and technology were of paramount importance in making possible the US’s
Vol. 16 (2014) Writing the Biography of Hans Bethe 205

economic and social well-being. He came to regret the role he had played earlier in
making some of this possible. His despair stemmed from the fact that he had a
drastically different vision of the aims and responsibilities of the United States in
the world and of the role that science would play in its growth and evolution than
the one projected by the George W. Bush administration. Perhaps most painful
was that his faith in reason and rationality—which had given him hope, resilience,
and buoyancy all his life—had been deeply shaken and undermined.127

Acknowledgments
I am indebted to Jeffrey Goldstone, Kurt Gottfried, and Snait Gissis for very
valuable and useful discussions; to Paul Forman for extended talks regarding the
content of this article and for his critical reading of it and suggestions. The helpful
recommendations by Peter Pesic and Robert Crease, the editors of Physics in
Perspective, are likewise gratefully acknowledged.

References
1
This paper is based on an invited lecture delivered in November 2011 at the annual meeting of
the History of Science Society and in December 2011 on the occasion of Paul Forman’s retirement
at the Smithsonian Institution. It was a privilege to have been asked to deliver the lecture hon-
oring Forman. We have been close friends for the past thirty years. We both come out of physics
and have a special relationship to physics and physicists. I admire Paul Forman and his works
greatly, and he has influenced me deeply. His integrity, his comportment, and his writings made,
and continue to make, clear to me the responsibilities we have as historians. This paper is dedi-
cated to him as a token of my admiration, affection and respect.
When considering what Paul Forman has accomplished as a historian of science and as a curator,
and keeps on accomplishing as a historian, many commendations can be made. John Heilbron,
who has known Forman since his student’s days at Berkeley did so when commenting on Forman’s
oeuvre as a historian at a conference in Vancouver in 2005 honoring Forman: John L. Heilbron,
‘‘Cold War Culture, History of Science and Postmodernity: Engagement of an Intellectual in a
Hostile Academic Environment.’’ in Cathryn Carson, Alexei Kojevnikov, and Helmuth Trischler,
eds., Weimar Culture and Quantum Mechanics (Singapore: World Scientific, 2011), 2–20. The
editors’ introduction to the proceedings of the conference and Heilbron’s article therein detail the
magnitude of Forman’s accomplishments as a historian of science and the respect he is held in as
an outstanding scholar.
2
He had joined the department in the spring semester 1967 and completed his PhD dissertation
that summer in the history department of the University of California at Berkeley; Hunter Dupree
had been his thesis adviser. Paul Forman, ‘‘Weimar Culture, Causality, and Quantum Theory,
1918–1927: Adaptation by German Physicists and Mathematicians to a Hostile Intellectual
Environment,’’ Historical Studies in the Physical Sciences 3 (1971), 1–115; ‘‘The Reception of an
Acausal Quantum Mechanics in Germany and Britain,’’ in Seymor Mauskopf, ed., The Reception
of Unconventional Science, Seymor Mauskopf, ed., [AAAS] Selected Symposium 25 ([Boulder
Colo]: Westview Press, 1979), 11–50; ‘‘Kausalität, Anschaulichkeit, and Individualität; or how
Cultural Values Prescribed the Character and the Lessons Ascribed to Quantum Mechanics,’’ in
Nico Stehr and Volker Meja, eds., Society and Knowledge: Contemporary Perspectives in the
Sociology of Knowledge & Science, ([New Brunswick NJ]: Transaction Books, 1984), 333–48;
reprinted, 2nd revised edition (Transaction Books: New Brunswick NJ, 2005), 357–371.
206 S. S. Schweber Phys. Perspect.

3
For the argument that Darwin had been influenced by the intellectual and political views and
cultural values of powerful circles in England and Scotland, see Robert M. Young, ‘‘Malthus and
the Evolutionists: The Common Context of Biological and Social Theory,’’ Past and Present 43
(1969), 109–45; ‘‘The Historiographic and Ideological Contexts of the Nineteenth-Century Debate
on Man’s Place in Nature,’’ in M. Teich and R.M. Young, eds., Changing Perspectives in the
History of Science (London: Heinemann, 1973), 344–438; Darwin’s Metaphor: Nature’s Place in
Victorian Culture (Cambridge: Cambridge University Press, 1985). See also A. La Vergata,
‘‘Images of Darwin: A Historiographic Overview,’’ in D. Kohn, ed., The Darwinian Heritage
(Princeton, NJ: Princeton University Press, 1985), 901–929, and Ingemar Bohlin, ‘‘Robert M.
Young and Darwin Historiography,’’ Social Studies of Science 21 (1991), 597–648.
4
Forman, ‘‘Behind Quantum Electronics: National Security as Basis for Physical Research in the
United States, 1940–1960,’’ Historical Studies in the Physical Sciences, 18 (1987), 149–229.
5
Heilbron, ‘‘Cold War Culture’’ (ref. 1), 17.
6
Forman, ‘‘Recent Science: Late Modern and Post-Modern,’’ in Philip Mirowski and Esther-
Mirjam Sent, eds., Science Bought and Sold: Rethinking the Economics of Science (Chicago:
University of Chicago Press, 2002), 109–48; ‘‘The Primacy of Science in Modernity, of Technology
in Postmodernity and of Ideology in the History of Technology,’’ History & Technology 23 (2007),
1–152; ‘‘(Re)cognizing Postmodernity: Helps for Historians—of Science Especially,’’ Berichte zur
Wissenschaftsgeschichte 3 (2010), 157–75; and especially ‘‘On the Historical Forms of Knowledge
Productions and Curation: Modernity Entailed Disciplinarity, Postmodernity Entails Antidiscip-
linarity,’’ Osiris 27 (2012), 56–100.
7
This is abstracted from Mirowski’s ‘‘Postface’’ in Philip Mirowski and Dieter Plehwe, eds., The
Road from Mont Pèlerin: The Making of the Neoliberal Thought Collective (Cambridge, MA:
Harvard University Press, 2009). Mirowski also states that ‘‘Neoliberals see pronounced inequality
of economic resources and political rights not as an unfortunate by-product of capitalism, but as a
necessary functional characteristic of their ideal market system. Inequality is not only the natural
state of market economies, but it is actually one of its strongest motor forces for progress. Hence
the rich are not parasites, but (conveniently) a boon to humankind.’’ (438) See also Philip Mi-
rowski, Science-mart: Privatizing American Science. (Cambridge, MA: Harvard University Press,
2011).
8
Silvan S. Schweber, Nuclear Forces: The Making of the Physicist Hans Bethe (Cambridge, MA:
Harvard University Press, 2012).
9
I am aware that at most I can write succinct, contextually sensitive, narrations selected from
some of the important components of his life from 1940 to his death: Los Alamos; the Shelter
Island conference; his consulting for the General Electric Knolls Laboratory, Detroit Edison, and
later AVCO; his involvement with H-bombs; his sabbatical in Cambridge/England during the
academic year 1955–6; his shift from high energy to nuclear physics; nuclear matter; his serving on
PSAC; the Nobel prize in 1967; his involvement in the Cornell student rebellion 1968–71; his
becoming an astrophysicist; neutrinos and supernovae; star wars.
10
Forman, ‘‘‘Swords into Ploughshares’: Breaking new Ground with Radar Hardware and
Technique in Physical Research after World War II,’’ Reviews of Modern Physics 67 (1995),
397–455.
11
Hans A. Bethe, ‘‘My Life in Astrophysics,’’ Annual Review of Astronomy and Astrophysics 41
(2003), 1–14.
12
For an early example of this see Bethe, ‘‘Über die nichtstationäre Behandlung des Photo-
effekts,’’ Annalen der Physik 4 (1930), 443–449.
13
Bethe’s first exposure to the practice of science was in his father’s laboratory. There he became
aware of the amazing diversity of animal life and learned that individual behavior can never be
Vol. 16 (2014) Writing the Biography of Hans Bethe 207

considered by itself but must always be seen through interactions with the environment and all the
entities that make up that environment. He also first experienced science as a social activity in
which his father, his father’s Assistenten, Doktoranten, and laboratory assistants were in constant
interaction in a shared physical and intellectual environment. The view of science Bethe obtained
in his father’s laboratory was as a practice in which knowledge is created by experiments using
instruments that measure with limited accuracy, produce data that have to be analyzed statistically
and interpreted with mathematical models that idealize the context in which the interactions take
place. The aim of the knowledge produced in his father’s laboratory was how to understand the
complexity and diversity of the biological world. There was no attempt to find an ultimate theory
that would explain all biological phenomena. The mature Bethe was always skeptical of the
possibility of finding final theories in physics.
14
Szasz and Siegel (who both had studied in Göttingen) reflected Hilbert’s ‘‘modernist’’ views of
making mathematics an autonomous discipline as well as his idealistic views of mathematics. See
Jeremy Gray, ‘‘Modernism in Mathematics,’’ in Eleanor Robson and Jacqueline Stedall, eds., The
Oxford Handbook of the History of Mathematics, (Oxford: Oxford University Press, 2009),
663–683.
15
See Forman and Armin Hermann’s entry on Sommerfeld in the Dictionary of Scientific Biog-
raphy, Charles C. Gillispie, ed. (New York: Scribner’s, 1975), 12:525–532.
16
Timothy Lenoir, ‘‘Practical Reason and the Construction of Knowledge,’’ in Ernan McMullin,
ed., The Social Dimensions of Science (Note Dame, IN: University of Notre Dame 1992), 158–197.
17
See Hans-Jürgen Borchers, ‘‘Einstein’s Principle of Maximal Speed in Classical and Quantum
Physics,’’ in Rathindra Nath Sen and Alexander Gersten, eds., Mathematical Physics Towards the
21st Century (Beer-Sheva: Ben Gurion University of the Negev Press, 1994).
18
See in particular Karl von Meyenn, ‘‘Pauli’s Belief in Exact Symmetries,’’ in Manuel Garcia
Doncel, Armin Hermann, Louis Michel, and Abraham Pais, eds., Symmetries in Physics (1600–
1980) (Bellaterra [Barcelona]: Semineri d’Historia de les Ciènces, Universitat Autònoma de
Barcelona, 1987), 329–360. Von Meyenn quotes a letter of Pauli to Schrödinger written on January
27, 1955: ‘‘When I consider the matter where a theory is in need of improvement, I never start
from considerations about measurability but from such conclusions of the theory where the
mathematics is not correct.’’ Von Meyenn goes on: ‘‘Behind these words is [Pauli’s] deep con-
viction that the mathematical structure of physical theories possesses a greater content of reality
than the common intuition and direct experience.’’ (332)
19
Bethe, ‘‘Quantenmechanik der Ein und Zwei-Elektronenprobleme,’’ in Hans Geiger and Karl
Scheel, Handbuch der Physik, XXIV, Part I, Adolf Smekal, ed., Quantentheorie (Berlin: Julius
Springer Verlag, 1933), 273–560; Bethe and Sommerfeld, ‘‘Elektronentheorie der Metalle,’’ in
Geiger and Scheel, Handbuch der Physik XXIV, Part II (Berlin: Julius Springer Verlag, 1933),
333–622.
20
Bethe, Robert F. Bacher, and Milton S. Livingston, Basic Bethe: Seminal Articles on Nuclear
Physics (New York: American Institute of Physics and Tomash Publishers, 1986).
21
Hans A. Bethe, ‘‘Theoretical Division: The Beginning,’’ in Theory in Action: Highlights in the
Theoretical Division at Los Alamos 1943–2003. Volume I. Compiled by Francis H. Harlow and H.
Jody Shepard. LA–14000–H. History Report. Unclassified. (Los Alamos National Laboratory:
Theoretical Division, 2004), 1–5. For a more detailed account of T-Division’s involvement with
computers see Bethe, ‘‘Introduction,’’ in Sidney Fernbach and Abraham H. Taub, eds., Computers
and their Role in the Physical Sciences (New York: Gordon and Breach Science, 1970), 1–10. For
the continuation of that story see William Aspray, John von Neumann and the Origins of Modern
Computing (Cambridge, MA: MIT Press, 1990). It is also interesting to note that Bethe never
programmed a computer to solve the complex problems he was considering.
208 S. S. Schweber Phys. Perspect.

22
However, the Cold War context was such that Robert R. Wilson (who had severed all his ties
with making atomic weapons after he left Los Alamos, where he had been the head of the Nuclear
Physics Division), in 1950, when he was the director of the Newman Lab, designed a mobile
electron beam gun to destroy atomic bombs after he had learned of strong focusing. See Silvan S.
Schweber, ‘‘Defending against Nuclear Weapons: A 1950 Proposal,’’ Physics Today 60, no. 4
(2007), 36–41.
23
See Forman, ‘‘On the Historical Forms of Knowledge Production’’ (ref. 6).
24
See Paul Hartman, The Cornell Physics Department: Recollections and a History of Sorts
(Ithaca, NY: n. p., 1984).
25
Norris Bradbury, then director of Los Alamos, in 1950 asserted this when interviewed by the
FBI in connection with the renewal of Bethe’s Q clearance. I am indebted for this information to
Alex Wellerstein, who has studied Bethe’s FBI record.
26
See for example Silvan S. Schweber, In the Shadow of the Bomb: Bethe, Oppenheimer, and the
Moral Responsibility of the Scientist. (Princeton, NJ: Princeton University Press, 2000) for the
details of the plans.
27
Michael Gordin, Red Cloud at Dawn: Truman, Stalin, and the End of the Atomic Monopoly.
(New York: Farrar, Straus, 2009).
28
Bethe, A Theory for the Ablation of Glassy Materials, Issue 38 of Research report. Avco
Manufacturing Corporation, Avco Everett Research Laboratory, 1958–30 pages; H.A. Bethe and
M.D. Adams, ‘‘A Theory of the Ablation of Glassy Materials,’’ International Journal of Aero-
nautical and Space Sciences 26 (1959), 321–328.
29
See Priscilla J. McMillan, The Ruin of J. Robert Oppenheimer, and the Birth of the Modern
Arms Race (New York: Viking, 2005).
30
See Schweber, In the Shadow of the Bomb (ref. 26).
31
Zuoyue Wang, In Sputnik’s Shadow: The President’s Science Advisory Committee and Cold War
America (New Brunswick, NJ: Rutgers University Press, 2008).
32
Bethe, ‘‘My Life in Astrophysics,’’ in Gerald E. Brown and Chang-Hwan Lee, eds., Hans Bethe
and His Physics (Singapore: World Scientific 2006), 27–44.
33
The neutrino problem was concerned with the fact that there were far fewer neutrinos being
emitted by the sun than solar models had predicted. See the articles on neutrinos in Brown and
Lee, Hans Bethe and His Physics (ref. 32).
34
See Jeremy Bernstein, Hans Bethe, Prophet of Energy (New York: Basic Books, 1980); Boris
Ioffe, ‘‘Hans Bethe and the Global Energy Problems,’’ in Brown and Lee, Hans Bethe and His
Physics (ref. 32), 263–272.
35
Karin Knorr-Cetina, Epistemic Cultures: The Sciences Make Knowledge (Cambridge, MA:
Harvard University Press, 1999), 1.
36
See, e. g., David Edgerton, ‘‘Science in the United Kingdom: A Study in the Nationalization of
Science,’’ in John Krige and Dominique Pestre, eds., Science in the Twentieth Century (Amster-
dam: Harwood Academic Publishers, 1997), 759–776.
37
See, e. g., Enrico Fermi, ‘‘Quantum Theory of Radiation,’’ Reviews of Modern Physics 4 (1932),
87–132.
38
Laurie M. Brown and Helmut Rechenberg, The Origin of the Concept of Nuclear Forces
(Philadelphia: Institute of Physics Pub., 1996).
39
As a result of his researches and calculations in atomic and solid state physics, Bethe early on
had recognized implicitly what Dirac would state explicitly in the first edition of his The Principles
of Quantum Mechanics, namely that our representation of the physical world can be hierarchically
Vol. 16 (2014) Writing the Biography of Hans Bethe 209

ordered by virtue of Planck’s constant, h. Macroscopic systems, whose characteristic time T, mass
M, and length L, are such that ML2/T  h are described by classical mechanics; those for which
ML2/T & h are described by quantum mechanics. See the remarkable text on quantum
mechanics, Eyvind H. Wichmann, Quantum Physics (New York: McGraw-Hill, 1971).
40
These hierarchies are not independent: accurate measurements of atomic energy levels will
reveal nuclear and subnuclear properties. Similarly, the recent startling discovery of the presence
of cold dark matter—consisting of as yet undiscovered subnuclear entities—to make sense of new
cosmological observational data is proof of the linkage between the various levels. But it must also
be noted that these observations have not destabilized our amazingly accurate representations of
the atomic world. Needless to say, the linkage of the levels is made explicit as soon as one tries to
answer evolutionary questions.
41
And more recently in terms of the standard model.
42
See Brown and Rechenberg, The Origin of the Concept of Nuclear Forces (ref. 38).
43
See in this connection Michelangelo De Maria, Mario Grilli, Fabio Sebastiani, eds.,The
Restructuring of the Physical Sciences in Europe and the United States,1945–60: Proceedings of the
International Conference Held in Rome, Università ‘‘La Sapienza’’, 19–23 September 1988’’ (Sin-
gapore: World Scientific,1989); and therein, Forman, ‘‘Social Niche and Self-Image of the
American Physicist’’ pp. 96–104.
44
Forman, ‘‘‘Swords into Ploughshares’: Breaking New Ground with Radar Hardware and
Technique in Physical Research after World War II,’’ Reviews of Modern Physics 67 (1995),
397–455. See also Forman, ‘‘Into Quantum Electronics: the Maser as ‘Gadget’ of Cold-War
America,’’ in Paul Forman and José Sánchez-Ron, eds., National Military Establishments and the
Advancement of Science and Technology: Studies in Twentieth Century History (Dordrecht: Klu-
wer Academic,.1996), 261–326.
45
Louis N. Ridenour, Radar System Engineering (New York: McGraw-Hill, 1947).
46
See Forman, ‘‘‘Atom Smashers: Fifty Years’—Preview of an Exhibit on the History of High
Energy Accelerators,’’ IEEE Transactions on Nuclear Science NS–24 (1977): 1896–99.
47
J.S. Hey, The Evolution of Radio Astronomy (New York: Science History Publications, 1973).
48
The conference began on June 3 and ended on the 6th. See Schweber, QED and the Men who
Made It (Princeton, NJ: Princeton University Press, 1994); Willis E. Lamb, and Robert C.
Retherford, ‘‘Experiment to Determine the Fine Structure of the Hydrogen Atom’’ (Columbia
University Radiation Laboratory Report, 1946), 18–26; Lamb and Retherford, ‘‘Fine Structure of
the Hydrogen Atom by Microwave Method,’’ Physical Review 72 (1947), 241–243; John E. Nafe,
Edward B. Nelson, and Isidor I. Rabi, ‘‘Hyperfine Structure of Atomic Hydrogen and Deuterium,’’
Physical Review 71 (1947), 914–15.
49
I owe the notion of a ‘‘crucial calculation’’ to Howard Schnitzer. See Schweber, QED (ref. 48),
where the idea is applied to Bethe’s calculation of the Lamb shift in hydrogen and Schwinger’s
calculation of the anomalous magnetic moment of the electron using renormalization concepts.
Other examples come readily to mind: Einstein’s calculation of the advance of the perihelion of
Mercury using his formulation of general relativity; Pauli’s calculation of the spectrum of
hydrogen using Born, Jordan, and Heisenberg’s matrix mechanics; and many other instances in
modern particle and condensed matter physics.
50
Steven Weinberg, ‘‘The Search for Unity: Notes for a History of Quantum Field Theory,’’
Daedalus 106, no. 4 (1977), 17–35.
51
See Schweber, ‘‘Shelter Island Revisited,’’ History of Physics Newsletter 11, no. 33 (2011), 1, 8,
10–13.
210 S. S. Schweber Phys. Perspect.

52
Max Dresden, H.A. Kramers: Between Tradition and Revolution (New York: Springer-Verlag,
1987).
53
Bethe’s Shelter Island notes were found in 2011 in his mother’s trunk that had been stored in
the basement of Bethe’s house on White Park Road in Cayuga Heights, NY. The notes can now be
found in Bethe’s papers in the Rare Manuscript Division of the Cornell Library.
54
Since libraries only maintain copies of bound books, copies of lecture notes are only to be found
in the library of the institution where they were delivered or in the Nachlass of the lecturer.
Dyson’s lectures were recently reissued: Freeman J. Dyson, Advanced Quantum Mechanics.
Translated and transcribed by David Derbes. 2nd ed. (Hackensack, NJ: World Scientific, 2011). A
first edition had been issued in 2007.
55
Thus, the proceedings of the Pocono and Oldstone conferences, the follow-ups of the June 1947
Shelter Island conference, became available as dittoed notes within two months of when they were
held in the spring of 1948 and 1949. They disseminated the lectures that Schwinger, Feynman, and
Dyson had presented at them.
56
In the United States after the war, the GI Bill allowed large numbers of young men who had
served in the Armed Forces to be trained as physicists and transformed the demography of
physics: there were many more physicists and they were younger.
57
At that same time, NATO began supporting summer schools on various subjects at different
places in Europe.
58
For the diffusion of Feynman’s approach to quantum electrodynamics and of his diagrams, see
David Kaiser, Drawing Theories Apart: The Dispersion of Feynman Diagrams in Postwar Physics
(Chicago: University of Chicago Press, 2005). The role of lecture notes and summer school notes
supplement the views expressed therein regarding textbooks in graduate education. See also
Kaiser, ed., Pedagogy and the Practice of Science: Historical and Contemporary Perspectives
(Cambridge, MA: MIT Press, 2005).
59
The history of advances in theoretical physics during the twentieth century has often been
written so that developments in ‘‘fundamental’’ theories are seen as being the most important.
They thus have received a disproportionate emphasis. In addition, the Cold War—at least in the
West—has had as one of its consequences that the history of physics during the second half of the
century has often filtered out the Soviet and Russian contributions. The history of the solution of
the phase transition problem in the late 1960s makes it clear that: a) while quantum field theory
was in decline and viewed as being marginal among high energy physicists from the late 1950s to
the mid-1960s, the use of field-theoretic methods was a thriving enterprise in solid state and
condensed matter physics and the source of deep insights that would later be transferred to
relativistic quantum field theory; and b) besides Lev Landau, Soviet theoretical physicists had
made important, foundational contributions to condensed matter physics and to the unraveling of
the phase transition problem.
60
A vision now tainted for having tasted sin in building atomic weapons. That vision was spelled
out in his 1953 Reith lectures, in which he quoted Bishop Sprat’s 1667 history of the Royal Society.
I can readily transcribe Sprat’s statement so that it becomes Oppenheimer’s manifesto for the IAS,
and for physics at the IAS: ‘‘It is to be noted that [the members of the IAS] are to freely admit
Men of different religions, Countries, and Professions of Life. This they are obliged to do, or else
they would come far short of the Largeness of their own Declarations. For they openly profess, not
to lay the foundations of an American, British, German or Japanese mathematics or science; but a
mathematics and science of Mankind.’’
61
Incidentally, in recognition of all that Bethe had accomplished in applying quantum mechanics
and quantum electrodynamics to explain atomic and nuclear phenomena after the war, in the early
1950s Oppenheimer invited Bethe to join the Institute as a professor of physics. Bethe declined,
Vol. 16 (2014) Writing the Biography of Hans Bethe 211

feeling that his place was at Cornell. The IAS had become a finishing school for the brightest
young theorists and Bethe felt that with all his other commitments he could no longer chart new
directions in research for these young people to explore.
62
Les Houches has become a year-round school, whose lectures and publications continue to
influence the development of physics profoundly.
63
Quantum mechanics had acquired a new robustness during World War II. It had explained
quantitatively the properties of germanium used in radar receivers, the properties of matter at 50
million K, and could predict with fair accuracy the results of critical nuclear reactions. The
teaching of quantum mechanics thus gained new importance as a result of the wartime advances.
No one was able to highlight and demonstrate the new powers of quantum mechanics better than
the twenty-eight-year-old Julian Schwinger, who had become a professor of physics at Harvard in
the fall of 1946. His 1947 and 1948 courses on nuclear physics and quantum mechanics became
legendary. Attended not only by the graduate students at Harvard, but by a large fraction of the
physics community in the greater Boston area, two sets of notes of these lectures were written one
by John Blatt, the other by Morton Hamermesh, and both were widely disseminated and repro-
duced elsewhere. They became the basis of quantum mechanics courses all over the United States.
These notes are difficult to find since they were not printed, nor bound, and thus most libraries did
not preserve them.
64
See John Krige and Dominique Pestre in Peter Galison and Bruce Hevly, eds., Big Science: The
Growth of Large-Scale Research (Stanford: Stanford University Press, 1992); Armin Hermann
et al., eds., History of CERN (Amsterdam: North-Holland Physics Publications, 1987–1996).
65
Forman, ‘‘On the Historical Forms of Knowledge Production’’ (ref. 6), 63. The valuation of
disciplinarity in modernity and interdisciplinarity in postmodernity is a central concern of that
article. See also the discussion of the historical development of disciplinarity and interdiciplinarity
within American universities in Louis Menand, The Marketplace of Ideas (New York: W.W.
Norton, 2010).
66
See David E. Rowe, ‘‘Klein, Hilbert and the Göttingen Mathematical Tradition,’’ Osiris 5
(1989), 186–213; ‘‘Making Mathematics in an Oral Culture: Göttingen in the Era of Klein and
Hilbert,’’ Science in Context 17 (2004), 85–129. Besides all the researches on infinite dimensional
vector spaces carried out in Göttingen (stemming from the concerns of Hilbert and Minkowski
with Boltzmann’s gas theory, lattice vibrations, and black-body radiation), after Einstein became
involved with his theory of general relativity, Hilbert and Klein became deeply entangled in these
activities and generated important mathematical advances, e.g. the researches of Emmy Noether.
Some of the researches of Veblen and of Élie Cartan could similarly be characterized as math-
ematical physics.
67
In fact, until World War II the graduate courses in classical mechanics were often offered by
departments of mathematics.
68
Think of Poincaré, Borel, Kolmogorov, Sinai, …
69
It would be interesting to compare in detail the factors that operated in the various national
settings (US, France, Soviet Union, Germany, Switzerland,…) that made possible the emergence
the discipline of mathematical physics; to compare the different kinds of problems addressed in
the various settings and what these reflected; to compare the status of the discipline in the differing
settings; to see whether the practitioners became members of physics or mathematics departments.
In addition, one can ask what made it possible for Wightman, Jost, Haag, Kastler to create their
schools and for the students they trained to form a new discipline with all the accoutrements that
go with it, such as professional journals and prizes. Surely in the United States the restructuring of
the universities into research and teaching universities after World War II was an important factor.
There was a greater emphasis on research, with lavish government support as part of its pursuance
of the Cold War, and the accompanying overhead payments allowing universities to support
212 S. S. Schweber Phys. Perspect.

activities and functions not directly supported by the government, such as scholarship in the arts
and the humanities. Undoubtedly, during the Cold War era, national prestige and similar factors
were at play in the Soviet Union and elsewhere. See, e.g., Clark Kerr, The Great Transformation in
Higher Education, 1960–1980: The Uses of the University (Albany, NY: State University of New
York Press, 1991); Stuart W. Leslie, The Cold War and American Science: The Military-Industrial-
Academic Complex at MIT and Stanford (New York: Columbia University Press, 1993).
70
That mathematicians held important and influential positions in both the Soviet Academy of
Sciences as well as in the Soviet Atomic Energy establishment was another important determinant
of the focus of the Soviet research in this area.
71
For valuable insights into the evolution of Russian and Soviet mathematics, see the very
interesting volume A. A. Bolibruch, Yu. S. Osipov, and Ya. G. Sinai, eds., Mathematical Events of
the Twentieth Century (Berlin: Springer, 2000), in particular the articles by V. I. Arnold, L.
D. Faddeev, and V. S Vladimirov.
72
Bogoliubov seems to be the exception in the Soviet Union.
73
Wightman’s ties were with von Neumann and Wigner; von Neumann had a close working
association with Hilbert during his stay in Gottingen from 1927 until 1929 as a Privatdozent there.
See Steve J. Heims, John von Neumann and Norbert Wiener: From Mathematics to the Technol-
ogies of Life and Death (Cambridge, MA: MIT Press, 1980); William Aspray, John von Neumann
and the Origins of Modern Computing (Cambridge, MA: MIT Press, 1990). Wigner and von
Neumann were close friends since their teens, becoming close colleagues in Princeton in the early
1930s. Recall that the Institute for Advanced Study (IAS) was located on the campus of Princeton
University from 1933, the date of its opening, until 1939, when Fuld Hall was completed. The IAS
mathematicians had offices in Fine Hall, where the Princeton mathematics department was
located. Fine Hall is the building adjacent to the Palmer Laboratory, the home of the Physics
department, with open corridors to it. Jost was a student of Wentzel and Pauli and was Pauli’s
successor as professor of theoretical Physics at the ETH. Both Pauli and Wentzel were students of
Sommerfeld. Haag was a student of Fritz Bopp, who was a student of Heisenberg, who in turn was
a student of Sommerfeld. All three at some stage gave axiomatic formulations of what they
thought were the foundations of quantum field theory. Van Hove was closely associated with
Weyl, von Neumann, Wigner, Bargmann, and Wightman during his stay at the IAS from 1949 to
1954. Hendrik B. G. Casimir, ‘‘Léon Charles Prudent van Hove (10 February 1924–2 September
1991),’’ Proceedings of the American Philosophical Society, 136, no. 4 (1992), 602–606.
74
From the early 1930s on, Wigner was a member of both the physics and the mathematics
departments at Princeton.
75
See Arthur Wightman, ‘‘The Theory of Quantized Fields in the 50s,’’ in Laurie Brown, Max
Dresden, and Lillian Hoddeson, eds., Pions to Quarks: Particle Physics in the 50s (Cambridge:
Cambridge University Press 1989); ‘‘The Usefulness of a General Theory of Quantized Fields,’’ in
Yian Yu Cao, ed., Conceptual Foundations of Quantum Field Theory (Cambridge: Cambridge
University Press, 1999) for a history of the developments in axiomatic and constructive field
theory. See also Rudolf Haag, ‘‘Local Algebras: A Look Back at the Early Years and at Some
Achievements and Missed Opportunities,’’ The European Physical Journal H 35 (2010), 255–261.
76
Raymond F. Streater and Arthur S. Wightman, PCT, Spin and Statistics, and All That (New
York: W. A. Benjamin, 1964); Res Jost, The General Theory of Quantized Fields ([Providence,
R.I.], American Mathematical Society, 1965). David Ruelle’s book, Statistical Mechanics: Rigor-
ous Results (New York: A.W. Benjamin, 1969), which became known as ‘‘The Book,’’ gives a
thorough overview of the ways rigorous mathematical analyses had secured some of the foun-
dations of statistical mechanics and had established what kinds of systems could be describe.
77
Freeman J. Dyson and A. Lenard, ‘‘The Stability of Matter,’’ Journal of Mathematical Physics 8
(1967), 423–433. See also Freeman J. Dyson, ‘‘The Stability of Matter,’’ in M. Chrétien, E.P. Gross
Vol. 16 (2014) Writing the Biography of Hans Bethe 213

and S. Deser, eds., Statistical Physics, Phase Transitions, and Superfluidity (New York: Gordon and
Breach, 1968), 1:179–239. See also the informative and insightful article by Larry Spruch, ‘‘Ped-
agogic Notes on Thomas-Fermi Theory (and on Some Improvements): Atoms, Stars, and the
Stability of Bulk Matter,’’ Reviews of Modern Physics 63 (1991), 151–209 and the references
therein to the papers of Eliot Lieb and by Walter Thirring. Also Eliot Lieb, ‘‘Thomas-Fermi and
Related Theories of Atoms and Molecules,’’ in G. Velo and A. S. Wightman, Rigorous Atomic and
Molecular Physics (New York: Plenum Press 1981), 213–301; and Walter Thirring, ‘‘The Stability
of Matter,’’ in ibid., 309–326.
78
See, e. g., Pierre Deligne et al., eds., Quantum Fields and Strings: A Course for Mathematicians
(Providence, RI: American Mathematical Society, 1999), in particular Edward Witten’s lectures
on ‘‘The Dynamics of Quantum Field Theory’’ (2:1119–1158).
79
See for example Subir Sachdev, ‘‘What can Gauge-Gravity Duality Teach Us about Condensed
Matter Physics?’’ Annual Review of Condensed Matter Physics 3 (2012): 9–33; ‘‘Strange and
Stringy,’’ Scientific American 308, no. 12 (2013), 44–51; ‘‘The Quantum Phases of Matter,’’ (2011)
http://arxiv.org/abs/1203.4565, last accessed April 22, 2014.
80
Battelle Seattle Rencontres (1971). Statistical Mechanics and Mathematical Problems, ed.
A. Lenard (Berlin, New York: Springer-Verlag, 1973).
81
See Forman, ‘‘Social Niche and Self-Image of the American Physicist,’’ in Michelangelo De
Maria et al., eds., Proceedings of the International Conference ‘‘The Restructuring of the Physical
Sciences in Europe and the United States, 1945–60’’ (World Scientific: Singapore, 1989), 96–104.
82
Thus, until the 1960s the general examinations for the PhD at American universities included
questions on all parts of physics and students were expected to have a good grounding in all of
them.
83
Robert K. Merton, ‘‘The Matthew Effect in Science,’’ Science 159, no. 3810 (1968), 56–63; ‘‘The
Matthew Effect in Science, II: Cumulative advantage and the symbolism of intellectual property,’’
Isis 79 (1988), 606–623.
84
The wartime laboratories, the IAS, together with other postwar developments, such as the GI
bill, the ONR’s and the AEC’s support of research in physics at universities, and the contract
system associated with this support, were responsible for the creation in the US of a new gen-
eration of outstanding young theorists: Phillip Anderson, Geoffrey Chew, Leon Cooper, Sidney
Drell, Freeman Dyson, Murray Gell-Mann, Roy Glauber, Marvin Goldberger, Walther Kohn,
Norman Kroll, Francis Low, Quinn Luttinger, Marshall Rosenbluth, Arthur Wightman, …
85
Bethe, High Energy Phenomena. A course of lectures given at Los Alamos in the spring and
summer of 1952, concerning phenomena involving particles with energy in the range of hundreds
of Mev. Main emphasis is placed on the properties of [pi]-mesons and on a relativistic treatment of
the nucleon–nucleon interaction. (Los Alamos, 1953).
86
Freeman Dyson, Marc Ross, Edwin E. Salpeter, Silvan S. Schweber, M. K. Sudarshan, William
M. Vissher and Hans A. Bethe, ‘‘Meson-Nucleon Scattering in Tamm-Dancoff Approximation,’’
Physical Review 95 (1954), 1644–58.
87
Geoffrey F. Chew and Francis Low, ‘‘Effective-range Approach to the Low-energy p-wave
Pion-Nucleon Interaction,’’ Physical Review 101 (1956), 1570–9; ‘‘Theory of Photomeson Pro-
duction at Low Energies,’’ Physical Review 101 (1956), 1579–87. Perhaps one the reasons that the
success of Chew and Low’s approach may have affected Bethe so deeply is that he had formalized
the ‘‘effective-range’’ approach to low energy nucleon–nucleon scattering and was committed to
what later would be called an ‘‘effective field theory’’: Bethe, ‘‘Theory of the Effective Range in
Nuclear Scattering,’’ Physical Review 76 (1949), 38–50. See also Tran N. Truong, ‘‘Bethe-Schw-
inger Effective Range Theory and Lehmann and Weinberg Chiral Perturbation Theories,’’ paper
presented at ICFP 09, September 24–30, 2009, Hanoi, Vietnam.
214 S. S. Schweber Phys. Perspect.

88
One of the first people I heard make this distinction was Alexi Assmuth in a talk at Harvard in
the late 1980s on Philip Anderson and the SSC. Assmuth had characterized high energy theories
as ‘‘foundational.’’ The language of ‘‘foundations’’ had entered historical narratives through the
work of Clifford Geertz, particularly, his essays in The Interpretation of Cultures (New York: Basic
Books, 1973) and the history of science was undergoing its own cultural turn in the late 1980s.
89
I am calling them ‘‘difficult’’ to contrast them with Bethe’s appellation of the dismal 1930s as the
‘‘happy thirties.’’
90
See Schweber, In the Shadow of the Bomb (ref. 26) for Bethe’s involvement in overcoming the
difficulties that his colleague Philip Morrison was experiencing because of his liberal political
views.
91
For an overview of Brueckner’s work, see his lectures in the 1958 Les Houches summer school:
Keith Brueckner, ‘‘Theory of Nuclear Structure and of Many Body Systems’’ in The Many Body
Problem: Le Problème à n corps. Cours donnés à l’École d’Été de physique théorique (Les
Houches 1958), 47–242.
92
H.A. Bethe and R. Bacher, ‘‘Nuclear Physics. A: Stationary States of Nuclei,’’ Reviews of
Modern Physics 8 (1936), 82–229; Bethe, ‘‘Nuclear Physics. B: Nuclear Dynamics, Theoretical,’’
Reviews of Modern Physics 9 (1937), 69–244.
93
The binding energy of an additional proton beyond the number of protons in a closed shell was
less than that that of the protons in the closed shell; likewise for neutrons.
94
See Maria Goeppert Mayer, Elementary Theory of Nuclear Shell Structure (New York: Wiley,
1955).
95
Oral History Transcript of interview with Hans A. Bethe by Charles Weiner at Cornell Uni-
versity May 8, 1972. Niels Bohr Library, Center of the History of Physics; American Institute of
Physics, College Park, MD.
96
Goldstone, using a diagrammatic representation of the perturbative scheme, proved that the
contributions of all unlinked diagrams—the contributions of which had given Brueckner diffi-
culties in his non-diagrammatic, algebraic approach—added up to zero. Goldstone also proved
that each order of perturbation theory gives a contribution proportional to the total number of
particles, this to all orders of perturbation theory.
97
David James Thouless, ‘‘The Application of Perturbation Methods to the Theory of Nuclear
Matter.’’ PhD diss., Cornell University (1958).
98
Jeffrey Goldstone, ‘‘Derivation of the Brueckner Many-Body Theory,’’ Proceedings of the Royal
Society London A 239 (1957), 267–279.
99
John Bardeen, Leon Neil Cooper, and John Robert Schrieffer, ‘‘Theory of Superconductivity,’’
Physical Review 108 (1957), 1175–1204.
100
Chien Shiung Wu et al.,’’Experimental Test of Parity Conservation in Beta Decay,’’ Physical
Review 105 (1957), 1413–5; Tsung-Dao Lee and Chen Ning Yang, ‘‘Question of Parity Conser-
vation in Weak Interactions,’’ Physical Review 104 (1956), 254–8.
101
Les problèmes mathématiques de la théorie quantique des champs, (Lille: Colloques Interna-
tionaux du Centre Nationale de la Rechèrche Scientifique, 3–8 Juin 1957).
102
See Wang, In Sputnik’s Shadow (ref. 31).
103
My paper ‘‘Hacking the Quantum Revolution’’ substantiating this claim is being submitted to
Studies in History and Philosophy of Science. B: Studies in History and Philosophy of Modern
Physics.
104
Tom Banks, Modern Quantum Field Theory: A Concise Introduction (Cambridge: Cambridge
University Press, 2008), 137. Conversely, the success of the approach of describing physical
Vol. 16 (2014) Writing the Biography of Hans Bethe 215

phenomena in terms effective field theories is a reflection of the fact that appropriately isolated
physical phenomena in a certain energy regime, probed and analyzed by instruments able to
resolve effects only within a certain range of length scales, can be described most simply by a set of
effective degrees of freedom appropriate to that scale.
105
See G. Peter Lepage, ‘‘What is Renormalization?’’ ArXiv:hep-ph/0506330. Talk given at the
Theoretical Advanced Study Institute in Elementary Particle Physics (TASI). University of Col-
orado, Boulder, Colorado, in 1989; ‘‘How to Renormalize the Schrödinger Equation’’,
ArxXiv:nucl-th/9706029, 1997 Lectures given at the VIII Jorge Andre Swieca Summer School
(Brazil, 1997).
106
Philip W. Anderson, ‘‘More is Different,’’ Science 177 (1972), 393–6.
107
Ten or so component atoms are the current limit for accurate ab initio computations of
molecular structure.
108
There is a vast literature on the subject with which I am only superficially acquainted. Daniel T.
Rodgers, Age of Fracture (Cambridge, MA: Harvard University Press, 2011) is a most valuable
overview within the American context. In addition to Forman’s writings, I have found the fol-
lowing articles and books useful entries: Mirowski and Sent, Science Bought and Sold (ref. 6);
David Tyfield, The Economics of Science: a Critical Realist Overview (New York: Routledge,
2012); Alfred Nordmann, Hans Radder, and Gregor Schiemann, eds., Science Transformed?:
Debating Claims of an Epochal Break (Pittsburgh: University of Pittsburgh Press, 2011); Hans
Radder, ed., The Commodification of Academic Research: Science and the Modern University
(Pittsburgh: University of Pittsburgh Press, 2010); Philipp Mirowski, Science-Mart (ref. 7); Helga
Novotny et al, eds., The Public Nature of Science under Assault (Berlin: Springer 2005); Domi-
nique Pestre, ‘‘The Technosciences between Market, Social Worries, and the Political: How to
Imagine a Better Future,’’, in ibid., 29–52; Pestre, ‘‘The Historical Heritage of the 19th and 20th
Centuries: Techno-science, Markets, and Regulations in a Long-term Perspective,’’ History and
Technology 23, no. 4 (2007), 407–420. For an overview of current forms of government(ality) and
their interaction with science and technology, see Sheila Jasanoff, Designs on Nature: Science and
Democracy in Europe and the United States. (Princeton, NJ: Princeton University Press, 2005);
Pestre, ‘‘Challenges for the Democratic Management of Technoscience: Governance, Participa-
tion and the Political Today’’, Science as Culture 17(2) (2008), 101–19; Pestre, ‘‘Understanding the
Forms of Government in Today’s Liberal Societies: An Introduction,’’ Minerva 47, no. 3 (2009),
243–60.
109
See Forman, ‘‘From the Social to the Moral to the Spiritual: The Postmodern Exaltation of the
History of Science,’’ in Jürgen Renn and Kostas Gavroglu, eds., Positioning the History of Science
(Berlin: Springer Verlag, 2007), 49–55; Forman, ‘‘The Primacy of Science in Modernity’’ (ref. 6),
‘‘On the Historical Forms of Knowledge Production’’ (ref. 6).
110
See Chapter 2 of Rodgers’ Fracture (ref. 108) for a history of the changes of perspectives and
assumptions in the disciple of economics, including neoliberalism.
111
The Patent and Trademark Law Amendments Act—now known as the Bayh-Dole Act —was
enacted by the US Congress in December 1980. The legislation gave American universities, small
businesses, and non-profit organizations exclusive patenting rights of inventions and control and
property rights over intellectual materials that resulted from governmental funding. The legisla-
tion had been sponsored by Senators Birch Bayh of Indiana and Bob Dole of Kansas.
112
For an account of the transformation of American universities, in particular of the bio-medical
sciences, see Mirowski, Science-Mart (ref. 7).
113
‘‘What the Past Tells us about the Future of Science,’’ in José Manuel Sánchez Ron, ed., La
Ciencia y la Tecnologia ante el Tercer Milenio (Madrid: Sociedad Estatal España Nuevo Milenio,
2002), 27–37, on 27.
216 S. S. Schweber Phys. Perspect.

114
The scholarship that went into ‘‘The Primacy of Science in Modernity, of Technology in
Postmodernity, and of Ideology in the History of Technology’’ and his Osiris 2012 article ‘‘On the
Historical Forms of Knowledge Production and Curation: Modernity Entailed Disciplinarity,
Postmodernity Entails Antidisciplinarity,’’ is most impressive. The Primacy article is 72 pages
long; its 424 endnotes are set in small type and take up 56 pages and its bibliography, 24. The
Osiris paper is only 45 pages long. However, one third of most the pages are taken up by lengthy
footnotes listing an enormous number of articles and books referring to the subject matter being
discussed, and many of them are commented on, at times very critically.
115
Forman, ‘‘On the Historical Forms of Knowledge Production’’ (ref. 6), 58n7.
116
Forman, ‘‘The Primacy of Science in Modernity, of Technology in Postmodernity and of
Ideology in the History of Technology’’ (ref. 6).
117
The following are the introductory remarks in the web page of the MIT department of bio-
engineering: ‘‘The Department of Biological Engineering was founded in 1998 as a new MIT
academic unit, with the mission of defining and establishing a new discipline fusing molecular life
sciences with engineering. The goal of this biological engineering discipline is to advance fun-
damental understanding of how biological systems operate and to develop effective biology-based
technologies for applications across a wide spectrum of societal needs including breakthroughs in
diagnosis, treatment, and prevention of disease, in design of novel materials, devices, and pro-
cesses, and in enhancing environmental health. ’’ http://web.mit.edu/be/index.shtm, last accessed
April 22, 2014. The Stanford University web page reads:’’ The mission of [the] Department of
Bioengineering is to create a fusion of engineering and the life sciences that promotes scientific
discovery and the development of new biomedical technologies and therapies through research
and education. http://bioengineering.stanford.edu/, last accessed April 22, 2014.
118
In ‘‘Hacking the Quantum Revolution’’ (ref. 103), I take issue with Laughlin and Pines’s
formulation of the foundational theory that is the point of departure for Laughlin’s assertions in A
Different Universe. Robert B. Laughlin and David Pines, ‘‘The Theory of Everything,’’ Proceed-
ings of the National Academy of Sciences 97, no. 1 (2000), 28–31.
119
See Richard Holloran, ‘‘Protracted Nuclear War,’’ Air Force Magazine 91, no. 3 (2008). The
substance of NSDD 32 and NSDD 75 was divulged in the New York Times on May 30, 1983 in an
op-ed column by Holloran.
120
See Sonja Michelle Amadae, Rationalizing Capitalist Democracy: The Cold War Origins of
Rational Choice Liberalism (Chicago: The University of Chicago Press, 2003) and particularly
S.M. Amadae, ‘‘Cold War, Security Dilemma, and Prisoner’s Dilemma: Does Insecurity Ratio-
nalize Hegemony?’’ (unpublished manuscript).
121
The rhetoric of ‘‘nuclear tipped’’—rather than for example ‘‘having nuclear capabilities’’—was
a deliberate attempt to minimize the devastation and chaos that would be wreaked by a protracted
nuclear war.
122
William Broad, Teller’s War: The Top-Secret Story Behind the Star Wars Deception (New York:
Simon & Schuster, 1992).
123
Space-Based Missile Defense: A Report by the Union of Concerned Scientists (Cambridge,
MA:,Union of Concerned Scientists. March 1984).
124
Kurt Gottfried and Bruce Blair, Crisis Stability and Nuclear War (New York: Oxford Uni-
versity Press, 1988).
125
See Bethe’s list of publications in Hans A. Bethe, Selected Works with Commentary (Singa-
pore: World Scientific, 1997).
126
Hans A. Bethe, [Untitled Statement], Federation of American Scientists Public Interest Report
48, no. 5 (September-October 1995), 8.
Vol. 16 (2014) Writing the Biography of Hans Bethe 217

127
It should be noted that Forman’s view of the postmodern world is every bit as bleak as was
Bethe’s. See Forman, ‘‘The Primacy of Science in Modernity’’ (ref. 6), n420. There Forman states:
‘‘If postmodernity, such as it is, continues its advance—and I can see nothing short of a cata-
strophic alteration of the life conditions on this planet as capable of altering the ever wider spread
and deeper seating of this radically self-regarding individualism—then the consequent transfor-
mations of personality, culture and society will render the constructive endeavors of the past three
centuries increasingly irrelevant and unintelligible. Among those endeavors, science is especially
vulnerable. For if science is not regarded as separate and distinguishable from technology in some
culturally highly valued ways, and if the fact of scientific laws is not regarded as a greater miracle
than the fact that the machine works, then it is ‘curtains’ for the scientific enterprise.’’

Brandeis University
Waltham, USA
e-mail: schweber@brandeis.edu

You might also like