You are on page 1of 297

SOL I D STAT E I NSU R R E CT ION

SOLID
STATE
I NSUR R ECTION
HOW T H E SCI E NCE OF SU B STA NCE
M A DE A M E R ICA N PH YSICS

M AT T E R
JOSE PH D . M A RT I N

U N I V E RSI T Y OF PI T TSBU RGH PR E S S


Published by the University of Pittsburgh Press, Pittsburgh, Pa., 15260
Copyright © 2018, University of Pittsburgh Press
All rights reserved
Manufactured in the United States of America
Printed on acid-free paper
10 9 8 7 6 5 4 3 2 1

Cataloging-in-Publication data is available from the Library of Congress

ISBN 13: 978-0-8229-4538-3


ISBN 10: 0-8229-4538-X

Cover art: Dave Barfield/National MagLab


Cover design: Joel W. Coggins
For my mother, who taught me to think before asking why
CONTENTS

Acknowledgments ix
List of Abbreviations xiii

INTRODUCTION
What Is Solid State Physics and Why Does It Matter? 3

1. The Pure Science Ideal and Its Malcontents 18


2. How Physics Became “What Physicists Do” 32
3. Balkanizing Physics 54
4. The Publication Problem 76
5. Big Solid State Physics at the National Magnet Laboratory 102
6. Solid State and Materials Science 119
7. Responses to the Reductionist Worldview 135
8. Becoming Condensed Matter Physics 152
9. Mobilizing against Megascience 170

CONCLUSIONS 199

Notes 213
Bibliography 243
Index 267
ACKNOWLEDGMENTS

The coziest of coffee shops and pubs sometimes dedicate a few square feet of
wall to the books that patrons have penned at their tables, and a writer might
requite with a tip of the hat to a gemütlich retreat that helped overcome para-
lyzing bouts of writer’s block. I wish I could recognize one clean, well-lighted
place, but this project intersected with a peripatetic phase of my life; it took
shape in Minneapolis and Saint Paul, Minnesota; Philadelphia, Pennsylva-
nia; Waterville, Maine; East Lansing, Michigan; and Leeds, Kenilworth, and
Cambridge, England—and while I was to-ing and fro-ing among them. With
that in mind, I thank instead the (dearly departed) Federal Aviation Adminis-
tration ban on the use of electronic devices during taxi, takeoff, and landing.
More than once, that forced respite sent my mind meandering toward some
of this book’s central arguments, which found their earliest form in frantic
scrawl on air sickness bags.
Those moments bore fruit only because I was traveling between institu-
tions populated with the best sort of people. The Program for History of
Science, Technology, and Medicine at the University of Minnesota made
even conceiving of this project possible. Michel Janssen and Sally Gregory
Kohlstedt steered it deftly from its jejune beginning to something approach-
ing maturity, with healthy assists from Alan Love, Bob Seidel, Ken Waters,
and Bill Wimsatt. The Physics Interest Group and the working papers sem-
inar of the Minnesota Center for Philosophy of Science put several of these
chapters through their paces and I benefited from conversations with Will
Bausman, Victor Boantza, Nathan Crowe, Lois Hendrickson, Maggie Hofius,
Adrian Fisher, Amy Fisher, Xuan Geng, Cameron Lazaroff-Puck, Barbara
Louis, Charles Midwinter, Aimee Slaughter, Jacob Steere-Williams, and many
others.
I have spent two invaluable years in residence at the Consortium for His-
tory of Science, Technology, and Medicine (once when it was still the Phil-
adelphia Area Center for History of Science). I recommend it to everyone I
meet. Babak Ashrafi is a scholarly force multiplier; on top of being one of the
clearest-thinking critics I have encountered, he has built a community ideal
x SOLID STATE INSURRECTION

for enriching projects like this one. My fellow Philly fellows Sarah Basham,
Rosanna Dent, Lawrence Kessler, Kurt MacMillan, Alicia Puglionesi, and
Michelle Smiley consistently challenged me to think in new ways, and this
book is the better for it. While in Philadelphia, I had the privilege to haunt
the halls of the Chemical Heritage Foundation (CHF), home to some of the
sharpest readers in the East. I am grateful for invaluable feedback from the
CHF writing group, where Carin Berkowitz, Ben Gross, Roger Eardley-Prior,
James Voelkel, and numerous other participants prompted me to hone vari-
ous chapters.
My colleagues at Colby College—Jim Fleming, Paul Josephson, and Lenny
Reich—and Michigan State University—Rich Bellon, James Bergman, Marisa
Brandt, Megan Halpern, Rebecca Kaplan, Dan Menchik, Richard Parks,
Isaac Record, and Catherine Westfall—provided me with strong, supportive
communities as this book was taking shape. Catherine in particular has been
my champion since the very early stages of this project. She showed me how
the history of solid state physics could find an audience. Most recently, the
Department of History and Philosophy of Science at the University of Cam-
bridge has offered an ideal environment in which to see this project through
its final stages. I have also been the beneficiary of fruitful comments from
and conversations with Joan Bromberg, Bob Crease, Clayton Gearhart, Greg
Good, Lillian Hoddeson, Catherine Jackson, Jeremiah James, Christian Joas,
Leo Kadanoff, Bill Leslie, Kathy Olesko, Peter Pesic, Greg Radick, Michael
Riordan, Ann Robinson, Richard Staley, James Sumner, Andy Warwick, Ben
Wilson, and Andy Zangwill. My errors are in spite of them.
Research for this book was made possible by the generosity of the Univer-
sity of Minnesota Graduate School; the Friends of the Center for History of
Physics, American Institute of Physics; the American Philosophical Society;
the Chemical Heritage Foundation; the Consortium for History of Science,
Technology, and Medicine; the Minnesota Center for the Philosophy of Sci-
ence; and the University of Chicago Special Collections Research Center.
These organizations funded the research that permitted me to contribute
further to the towering debt the historical profession owes to the fabulous
archivists and librarians who have assumed the unenviable task of bringing
the mountains of paper the Cold War generated to heel.
Abby Collier and the University of Pittsburgh Press have been a delight
to work with throughout. I am sorely in hock to Abby for her patience, per-
ceptiveness, and unfailingly good advice, and to the press’s reviewers and
ACKNOWLEDGMENTS xi

copyeditor for their careful reading of the manuscript and thoughtful criti-
cisms that much profited the final version.
Finally, my deepest thanks to Margaret Charleroy, who makes it all worth-
while. Those flights that had me scribbling frantically onto air sickness bags,
or in the vanishing margins of in-flight magazine ads for “America’s Best
Doctors,” were mostly because the early phases of our careers kept us sepa-
rated by many miles, and, at times, continents. With patience and acuity, she
read large portions of the writing that resulted. This book is her fault.
LIST OF ABBREVIATIONS

ACS American Chemical Society


AEC Atomic Energy Commission
AIME American Institute of Mechanical Engineers
AIP American Institute of Physics
AIPH American Institute of Physics. Office of the Director, Records
of Elmer Hutchisson, 1948–1966. Niels Bohr Library and
Archives, College Park, MD
AIPK American Institute of Physics. Office of the Director, Records
of H. William Koch, 1932–1988. Niels Bohr Library and
Archives, College Park, MD
AIPM American Institute of Physics. Governing Board Meeting Min-
utes, 1931–1990. Niels Bohr Library and Archives, College
Park, MD
AIPS American Institute of Physics. Office of the Secretary Records,
1931–2000. Niels Bohr Library and Archives, College Park,
MD
APS American Physical Society
APSM American Physical Society Meeting Minutes and Membership
Lists, 1902–2003. Niels Bohr Library and Archives, College
Park, MD
APSR American Physical Society. Records. Niels Bohr Library and
Archives, College Park, MD
ARPA Advanced Research Projects Agency
ASM American Society for Metals
AvHP Arthur von Hippel Papers. MIT Archives and Special Collec-
tions, Cambridge, MA
xiv SOLID STATE INSURRECTION

BAR University of Chicago, Office of the President, Beadle Admin-


istration Records. University of Chicago Special Collections
Research Center, Chicago
BES Basic Energy Sciences (United States Department of Energy)
BKS Bohr-Kramers-Slater theory
COSMAT Committee on the Survey of Materials Science and Engineering
CRS American Physical Society, Division of Solid State Physics.
Correspondence of Roman Smoluchowski, 1943–1947. Niels
Bohr Library and Archives, College Park, MD
CSSP Cyril Stanley Smith Papers. MIT Archives and Special Collec-
tions, Cambridge, MA
DFD Division of Fluid Dynamics (American Physical Society)
DFT Density functional theory
DOD Department of Defense
DOE Department of Energy
DSSP Division of Solid State Physics (American Physical Society)
EPWP Eugene P. Wigner Papers. Princeton University Archives,
Princeton, NJ
FBP Francis Bitter Papers, 1925–1967. MIT Archives and Special
Collections, Cambridge, MA
FBPS Felix Bloch Papers. Stanford University Archives, Stanford, CA
FSP Frederick Seitz Papers. University of Illinois Archives, Urbana
GPHP Gaylord P. Harnwell Papers. University of Pennsylvania Ar-
chives, Philadelphia
HAR University of Chicago, Office of the President, Hutchins Ad-
ministration Records. University of Chicago Special Collec-
tions Research Center, Chicago
HBP Harvey Brooks Papers. Harvard University Archives, Cam-
bridge, MA
IDL Interdisciplinary laboratory (Advanced Research Project
Agency)
IRE Institute of Radio Engineers
ISM Institute for the Study of Metals (University of Chicago)
LIST OF ABBREVIATIONS xv

JCP Journal of Chemical Physics


JCSP John C. Slater Papers. American Philosophical Society, Phila-
delphia
JFI James Franck Institute (University of Chicago)
JHVVP John H. Van Vleck Papers. Niels Bohr Library and Archives,
College Park, MD
KBP Kenneth Bainbridge Papers. Harvard University Archives,
Cambridge, MA
KKDP Karl K. Darrow Papers. Niels Bohr Library and Archives, Col-
lege Park, MD
LAR University of Chicago, Office of the President, Levi Adminis-
tration Records. University of Chicago Special Collections
Research Center, Chicago
LIR Laboratory for Insulation Research (Massachusetts Institute of
Technology)
LKP Leo Kadanoff Papers. University of Chicago Special Collec-
tions Research Center, Chicago
MAB Materials Advisory Board (National Research Council)
MIT Massachusetts Institute of Technology
NAL National Accelerator Laboratory
NAS National Academy of Sciences
NML National Magnet Laboratory
NMLR Francis Bitter National Magnet Laboratory. Records. MIT
Archives and Special Collections, Cambridge, MA
NRC National Research Council
NSF National Science Foundation
PTDR American Institute of Physics, Physics Today Division Re-
cords, 1948–1971. Niels Bohr Library and Archives, College
Park, MD
RGPP Robert G. Parr Papers. Chemical Heritage Foundation, Phila-
delphia
RLE Research Laboratory of Electronics (Massachusetts Institute
of Technology)
xvi SOLID STATE INSURRECTION

RRL Radio Research Laboratory (Harvard University)


RRLR Radio Research Laboratory. Records. Harvard University
Archives, Cambridge, MA
SDI Strategic Defense Initiative
SLAC Stanford Linear Accelerator
SSC Superconducting Super Collider
WSP William Shockley Papers. Stanford University Archives, Stan-
ford, CA
SOL I D STAT E I NSU R R E CT ION
introduction

WHAT IS SOLID STATE PHYSICS


AND WHY DOES IT MATTER?

Solid state physics sounds kind of funny.


—GREGORY H. WANNIER, 1943

The Superconducting Super Collider (SSC), the largest scientific instrument


ever proposed, was also one of the most controversial. The enormous parti-
cle accelerator’s beam pipe would have encircled hundreds of square miles
of Ellis County, Texas. It was designed to produce evidence for the last few
elements of the standard model of particle physics, and many hoped it might
generate unexpected discoveries that would lead beyond. Advocates billed
the SSC as the logical apotheosis of physical research. Opponents raised their
eyebrows at the facility’s astronomical price tag, which stood at $11.8 billion
by the time Congress yanked its funding in 1993. Skeptics also objected to
the reductionist rhetoric used to justify the project—which suggested that
knowledge of the very small was the only knowledge that could be truly fun-
damental—and grew exasperated when SSC boosters ascribed technological
developments and medical advances to high energy physics that they thought
more justly credited to other areas of science.
To the chagrin of the SSC’s supporters, many such skeptics were fellow
physicists. The most prominent among them was Philip W. Anderson, a No-
bel Prize–winning theorist. Anderson had risen to prominence in the new
field known as solid state physics after he joined the Bell Telephone Labora-
tories in 1949, the ink on his Harvard University PhD still damp. In a House
of Representatives committee hearing in July 1991, Anderson, by then at
4 SOLID STATE INSURRECTION

Princeton University, testified: “Particle physics is a narrow, inbred field, and


it is easy for the particle physicists to create an external appearance of una-
nimity of goals.”1 This was not a smear against the intellectual viability of the
SSC—Anderson conceded that the science it would enable would be unim-
peachably sound. Rather, it was a reaction against the tendency of some par-
ticle physicists to equate their subdisciplinary priorities with those of physics
writ large. It was a challenge to the position high energy physics had enjoyed
as the most prestigious branch of American science for much of the Cold War.
The opposition Anderson and his like-minded colleagues mounted
against the SSC throughout the late 1980s and early 1990s, which played
out in congressional committees, scientific publications, and popular media,
laid bare deep divisions that had remained largely hidden to nonphysicists up
to that point. Physicists simply did not openly oppose funding for a project
championed by colleagues in a neighboring specialty, especially an under-
taking so high profile as the Super Collider. That reality had preserved the
illusion that physicists were unanimous in their goals for decades. Anderson
and his allies, by exposing rifts within the physics community, shattered that
illusion. They introduced policymakers and the American public to solid
state and condensed matter physics.2 These fields, although they had rep-
resented a healthy plurality of physicists since at least the early 1960s, had
nevertheless remained comparatively obscure. So, therefore, had their inter-
ests. Increased visibility of solid state and condensed matter physics in policy
circles heightened awareness of their distinct perspective on the identity and
purpose of physics, which differed substantially from the one politically savvy
nuclear and high energy physicists had been selling in the halls of power, with
considerable success, since the end of the Second World War.
The standoff between the SSC’s advocates and its critics was just the most
recent and most public encounter in a long, intricate, and often troubled re-
lationship between those physicists who investigated complex physical sys-
tems and those who probed the minutest constituents of matter and energy.
Anderson’s testimony cut to the heart of the controversy behind the SSC:
the high energy physics community, which wielded its intellectual prestige to
sway patrons and policymakers alike, was wont to assume that its parochial
interests represented the common mission of all of physics. But physics in
the second half of the twentieth century was far from monolithic, and, from
Anderson’s perspective, could not be adequately served with monolithic
laboratories.
This book tells the story of how solid state physicists, by developing an
INTRODUCTION: WHAT IS SOLID STATE PHYSICS AND WHY DOES IT MATTER? 5

identity and a set of intellectual priorities that suited their professional goals,
redefined the boundaries and mission of American physics during the Cold
War. The research program to which the SSC belonged was rooted in a pure
science ideal dating to the late 1800s, which had motivated the founding of
the American Physical Society (APS) in 1899. But, almost from its inception,
the APS was beset by demands that it do more to represent those physicists
who plied their trade in industry. Solid state physics grew from a tension at
the heart of American physics between the pure science ideal and the needs
of industrial and applied physicists who constituted an increasing proportion
of its membership as the twentieth century wore on. Once established within
the APS in the late 1940s, solid state grew rapidly into the largest subfield
of American physics, developing a set of interests, outlooks, and goals that
at times aligned with and at other times clashed with the ideals dominant
in other areas of physics. Those interests, outlooks, and goals helped define
the scope of American physics and shape the identity of American physicists
through the Cold War.

WHAT IS SOLID STATE PHYSICS?


This deceptively simple question has some deceptively simple answers: solid
state physics is the study of the physical properties of solid matter; it is a sub-
field of physics, the most populous in the United States for much of the later
twentieth century; it is the branch of condensed matter physics that studies
solids with regular crystal lattice structures. Those answers are true within
their respective domains, but they gloss over a bevy of bedeviled details. Re-
search into the properties of solids has a long history, but it was not until the
mid-twentieth century that physical research on solids became the focus for a
new discipline. Yes, the physicists who founded solid state physics and built
it into the largest segment of the American physics community were primarily
concerned with understanding the behavior of regular solids, but that casts
only the palest illumination on those factors that make the field worthy of
historical attention. Solid state physics is notable for what it is not as much as
for what it is. When it formed in the 1940s, solid state physics defied deeply
rooted ideological presumptions—most centrally the pure science ideal—that
the American physics community held dear. As a result, it helped redefine the
scope of physics itself in a way that would shape its role in Cold War America.
Solid matter—rigid though it is—was ill-adapted for building the bound-
aries of a discipline when solid state physics emerged.3 The physical con-
cepts, theoretical methods, and experimental techniques used to investigate
6 SOLID STATE INSURRECTION

solid matter were often just as readily turned to not-so-solid matter—super-


conductivity, observed in some solids at low temperatures, is closely related
to superfluidity, another low-temperature phenomenon. A semantically strict
definition of solid state physics would include the former, but not the lat-
ter (a nettlesome inconsistency that would contribute to the rise of “con-
densed matter physics” as a preferred term in the 1970s and 1980s). Further-
more, the vast expanse of questions physicists could ask about solids, and
the equally diverse range of techniques they could use to investigate those
questions, made for a diffuse field that lacked a set of central motivating ques-
tions or techniques to provide conceptual cohesion. As the editors of Out of
the Crystal Maze: Chapters from the History of Solid-State Physics noted in
1992: “The field is huge and varied and lacks the unifying features beloved
of historians—neither a single hypothesis or set of basic equations, such as
quantum mechanics and relativity theory established for their fields, nor a
single spectacular and fundamental discovery, as uranium fission did for nu-
clear technology or the structure of DNA for molecular biology.”4
The argument that the solid state of matter is itself a discrete physical
phenomenon carries some prima facie plausibility, but it did not appear that
way from the standpoint of physical theory in the 1940s. Although solidity
was an easily identifiable trait of some material aggregates, the properties
of solids could not be reliably characterized by a consistent theoretical ap-
proach. Whereas Maxwellian electrodynamics served as a single framework
with which electromagnetic phenomena could be addressed, and physicists
could reach for the laws of thermodynamics anytime they wanted to discuss
heat, solids were a medium in which electromagnetism, heat, and most other
physical phenomena might persist. It would be plausible to suggest that
quantum mechanics provides a basis from which it is possible to understand,
or even derive, most if not all the properties of solids. However, such an enter-
prise was unfeasible in the mid-1940s. Investigating solids instead required
employing a number of theoretical approaches, both quantum and classical.
Solids invited a similarly colorful array of experimental techniques. Physicists
explored their properties at the extremes of low temperature and high pres-
sure. They zapped them with neutrons, electrons, and various frequencies of
electromagnetic radiation. They chemically doped them and blasted them
with ultrasonic waves. They poked and prodded them with other solids.
Solid state physics was a big tent, both theoretically and experimentally, and
so the impetus for its formation cannot be found by searching for a consistent
set of techniques or practices.
INTRODUCTION: WHAT IS SOLID STATE PHYSICS AND WHY DOES IT MATTER? 7

Because it could not claim an origin in any one research tradition or re-
gime of practice, solid state was, by the traditional standards of discipline
formation, an unusual category. Before the Second World War, physics was
understood to be divided into phenomenological categories like thermo-
dynamics, acoustics, optics, mechanics, electromagnetism, and quantum
mechanics.5 After the Second World War, a field appeared that claimed as
its domain thermodynamics, acoustics, optics, mechanics, electromagnetism,
and quantum mechanics in solids (and sometimes in other phases of matter
too). Isidor Isaac Rabi’s exclamation upon learning of the discovery of the
muon—“Who ordered that?”—is perhaps a more fruitful starting point for
gaining purchase on the slippery history of solid state physics.6 Whose inter-
ests did a field with such an unorthodox constitution serve? What changes
in the physics community allowed it to form? How did that formation come
about? Given the field’s rapid growth into the most populous segment of
post–Second World War American physics, what consequences propagated
as a result of its heterodoxy and the changes that permitted it? In short, why
did the field come to exist at all and how did it influence physics as a whole?
Addressing those questions reveals that solid state physics was much more
than a provincial subfield, subsidiary to the primary narratives of American
physics. It was integral to negotiating the identity of physics and essential for
maintaining its prestige in Cold War America.
Telling this story requires trading in some well-worn categories, of which
historians tend to be rightfully suspicious. Categories like pure science, or ba-
sic and applied research, are problematic. A great deal of work has shown that
so-called pure science was adulterated with worldly interests, and that the
artificial and not altogether coherent distinction between basic and applied
research fails to hold in practice. But historians also recognize the power
these categories possessed as regulative ideals that guided the way scientists
organized their professional lives. Mario Daniels and John Krige have shown
how “basic” and “applied” research functioned as political tools for Cold
War scientists, permitting them some control over the circulation of knowl-
edge in a context governed by military secrecy regimes.7 I approach these
categories from a similar perspective and show how pure science, basic and
applied research, fundamental research, and other value-laden designations
were tools for disciplinary as well as national politics, and therefore reveal
the ideals and convictions that gave meaning to physicists’ active efforts to
systematize their professional lives.
8 SOLID STATE INSURRECTION

THE PROMINENCE OF PHYSICS IN COLD WAR AMERICA

Taking solid state and condensed matter physics as a central object of his-
torical inquiry requires approaching old questions from a new perspective.8
A great deal of historical work addresses the question of why the Super-
conducting Super Collider failed, for example, but it might be more appro-
priate to ask why it ever had a chance to succeed in the first place.9 The US
government had spent over a billion dollars on a scientific project before,
but the Manhattan Project was principally an engineering endeavor, single-
mindedly focused on a military objective during a time of war.10 How did it
even become conceivable that a single facility dedicated to uncovering ab-
stract knowledge might consume similar resources in peacetime? It would be
tempting to answer this question by pointing to the considerable prestige and
influence physics garnered from the Manhattan Project. High energy physics,
which emerged from nuclear physics, had earned the latitude to pursue ab-
stract research. Nuclear physics, after all, was exceedingly abstract, even into
the 1930s, and it had resulted in the most fearsome weapon the world had
ever seen by 1945.11
This familiar story reflects aspects of the exalted heights physics attained
in Cold War American society, but it neglects what most physicists were
actually doing. For all its visibility, high energy physics, which cast itself as
the intellectual heir to nuclear physics, constituted only around 10 percent
of the American physics community at the time of the SSC’s cancellation.
Most physicists were not probing atomic viscera at cathedralesque accel-
erator facilities; they were investigating the properties of the type of matter
that surrounds us and finding new things to do with it. Historians require a
fuller accounting of those activities before claiming a perspective capable of
explaining the place of physics in Cold War American society. It is easy to see
how the historical trajectory of fields like solid state physics depended on its
relationship with nuclear and high energy physics. Less obvious is the fact
that this dependence was reciprocal, and that solid state—a diverse, messy
field with a complicated and shifting set of conceptual dependencies—in
some respects better represents physics as a whole than do its more revered
siblings.
After the Second World War, solid state physics, plasma physics, poly-
mer physics, and other specialties devoted to complex matter grew rapidly.
Physicists working in these fields quickly came to dominate the American
physics community, at least numerically. Nevertheless, the smaller proportion
INTRODUCTION: WHAT IS SOLID STATE PHYSICS AND WHY DOES IT MATTER? 9

of physicists who studied the elementary components of matter and the most
distant celestial objects capitalized most fully on the postwar prominence of
physics. They were the most recognizable to the public, wielded the greatest
influence in government, commanded the bulk of the considerable intellec-
tual prestige physics enjoyed in the postwar era, and nurtured intellectual
ideals that reinforced those advantages. The contrarian spirit apparent in An-
derson’s testimony against the SSC emerged over decades as a response to
this attitude, becoming central to the identity of American solid state physics.
In addition to exposing long-standing disagreements about the mission
and purpose of physics, the demise of the SSC symbolized the end of the era
in which physics reigned as the undisputed sovereign of American science.
As the SSC faltered, the Human Genome Project gathered momentum on
promises that it would revolutionize biology and medicine, and surpassed
physics in both public approbation and policy influence.12 The exalted posi-
tion physics had held during the Cold War is nonetheless a remarkable his-
torical phenomenon. Even toward the end of the Second World War, Ameri-
can physicists worried that their field was little known beyond a small group
of professionals. The exceptions to this rule were iconic figures like Albert
Einstein, whose fame was bound up in the legendary unfathomability of his
theories.13 After the war, leaders in the physics community gained national
celebrity and became familiar faces in Washington, DC, as they assumed
powerful advisory roles, shaped national policy, and shepherded in an era of
generous government funding for science.14 The question of how physicists
first attained this position is somewhat different from the further question of
how they then maintained it for half a century.
An appeal to the Manhattan Project, and other wartime contributions,
does provide a powerful answer to the first of these questions. The $2 bil-
lion the United States government invested in the Manhattan Project went in
part toward developing a physical infrastructure that provided the template
for the national laboratory system.15 The psychological immediacy of nuclear
weapons helped figures such as J. Robert Oppenheimer and Freeman Dyson
position themselves as public intellectuals.16 The urgency of the nuclear arms
race created opportunities for physicists to become deeply engaged with
weapons policy, which in turn gave them clout on a wide array of public pol-
icy issues.17 The success of wartime nuclear research, which quickly turned
abstruse knowledge about the submicroscopic world into a weapon that ir-
revocably reconfigured geopolitics, goes a long way toward explaining the
exalted position of physics in early Cold War American politics and society.
10 SOLID STATE INSURRECTION

This explanation is less than sufficient, however, to account for the con-
tinued prominence of physics through the early 1990s, which included the
growth of high energy physics, a field that claimed little economic, techno-
logical, or military relevance but nonetheless commanded billions of tax-
payer dollars to build and operate research facilities of unprecedented scale.
“Megascience,” as Lillian Hoddeson, Catherine Westfall, and Adrienne Kolb
have christened it, became the standard mode of research for the most visi-
ble physics research after the Second World War.18 From the vantage point
offered by a quarter century’s distance from the SSC’s demise, however,
megascience seems like a Cold War fever dream. For how long is it reasonable
to assume that the memory of the Manhattan Project sufficed to convince
policymakers that high energy physicists should continue to enjoy a blank
check from the Atomic Energy Commission (AEC), and later, the Depart-
ment of Energy, especially when they routinely denied that their work came
with practical offshoots?
The remarkable history of nuclear physics in the 1930s and 1940s no
doubt contributed to the rapid growth of high energy physics soon after the
Second World War. As Audra Wolfe explains in her history of Cold War sci-
ence and technology: “High-energy physics thrived within the institutional
culture of the Cold War because the AEC—the agency that bankrolled it—
believed in the inherent relevance of nuclear science to the national interest.
What nuclear physics wanted, nuclear physicists got.”19 This explanation
captures the psychology of the 1950s and early 1960s, but it becomes less
adequate later in the Cold War. Although they claimed the same ancestry,
nuclear physicists and high energy physicists had formed distinct commu-
nities by the late 1960s. The former was deeply intertwined with the inter-
ests of the national security state, whereas the latter was uncompromising in
its commitment to pursuing knowledge with no evident applications.20 The
more high energy physics established its bona fides as a field unsullied by
practical concerns the less it should have been able to trade on the promise
of relevance to national defense, even though it represented an investment in
national prestige. What explains the continued—and indeed ostentatious—
success high energy physics enjoyed with federal patrons that ended only
with the SSC’s demise in 1993?
Missing from previous accounts is the contribution of solid state and re-
lated research to the image and identity of physics. As Anderson observed
when he lamented the unanimous front high energy physicists presented,
those viewing physics from the outside were often not equipped to distin-
INTRODUCTION: WHAT IS SOLID STATE PHYSICS AND WHY DOES IT MATTER? 11

guish between the various subfields and research communities of which it was
composed. To many policymakers, physics was physics. It generated arcane
knowledge about the natural world and it produced fantastic gadgets. Those
two functions were connected in some way; therefore, the field was deserv-
ing of support. Policymakers generally accepted the judgment of the most es-
teemed representatives of the field as to how that support should be allocated.
Sarah Bridger’s Scientists at War recounts the recollections of New Mexico
senator Clinton Anderson, who admitted weighing scientific evidence based
on his instinctual trust of the individual expert delivering it, rather than on an
attempt to understand the scientific content of the evidence.21 Habits such as
these ensured that the politically best-placed physicists enjoyed considerable
sway over the image of the field, which shaped federal funding priorities.
High energy physicists’ success maintaining high levels of federal sup-
port, however, depended on provinces of physics with less political clout con-
tinuing to churn out research with near-term technological and economic rel-
evance. The military made rapid and expedient use of semiconductor-based
electronic components and improved materials. The burgeoning American
consumer culture eagerly embraced the technological products of physical
research such as transistors, integrated circuits, and improved bakeware and
stereo equipment. American industry found uses for lasers, superconducting
magnets, nuclear magnetic resonance techniques, and bespoke alloys. These
originated in solid state physics and allied fields, but as long as high energy
physicists succeeded in presenting their work as archetypical and policy-
makers remained incurious about the field’s internal diversity, the benefits of
such advances accrued to its more prestigious branch. High energy physics,
in short, maintained its success in part because the accomplishments of solid
state physics continually renewed in the minds of federal patrons the associ-
ation between physics as a whole and the technical, economic, and military
benefits of a few of its endeavors. A thorough appreciation of the growth of
solid state physics through the Cold War is therefore a prerequisite for under-
standing physics as a whole in one of the most auspicious eras in its history.

THE SCOPE OF THE BOOK


In 1899, the year the American Physical Society was established, its found-
ing president Henry Rowland wrote: “Where, then, is that person who ig-
norantly sneers at the study of matter as a material and gross study? Where,
again, is that man with gifts so God-like and mind so elevated that he can
attack and solve its problem?”22 He referred to late nineteenth-century strug-
12 SOLID STATE INSURRECTION

gles to understand the structure and behavior of atoms and molecules. The
sentiments he described nonetheless colored physical investigations of solids
and other complex matter throughout the twentieth century. Solid state phys-
ics often drew sneers from those who fancied that their own studies attained a
greater degree of elegance and looked down their noses at “Schmutzphysik,”
or “squalid state physics.” These pejoratives, the stuff of water-cooler ban-
ter rather than published invective, are attributed to Murray Gell-Mann and
Wolfgang Pauli, respectively. In addition to serving particle physicists in their
efforts to exalt their own studies, they provided a rallying point for solid
state physicists, who found motivation in opposing such condescension.23
Far from being the grimy and inelegant enterprise high energy physicists
derided, they insisted, solid state physics posed gnarly conceptual and prac-
tical problems that inspired noteworthy leaps of theoretical imagination and
experimental virtuosity.
The great irony of the derision directed at solid state physics is that the
things that offended other physicists’ sensibilities—its focus on complex,
real-world systems, its connections to industry—were the very same things
that helped renew the warrant for blue-skies research so valued by those hurl-
ing the insults. This book offers a history of the American solid state physics
community with the goal of illuminating how attention to it and similar fields
can reveal dependencies of this type and thereby enrich, and perhaps even
reform, our understanding of twentieth-century physics. It presents a story
about the organizational structures of American physics and the ideas that
shaped it, following the professional societies, journals, laboratories, and po-
litical interventions, as well as the discourses and disagreements that influ-
enced what forms they took. These structures both reflected and reinforced
what it meant to be a physicist in the eras in which they were built, and they
changed in response to shifting ideas of professional identity and disciplinary
purpose. Changing them was often a way to enact a vision of the field, of
where it should go, what it should be, and whom it should serve. Through
each of the changes traced here, solid state took another step toward reshap-
ing American physics in its own image.24
Appreciating how solid state physics changed the collective identity of
American physics requires understanding what came before. That is the goal
of the first two chapters, which describe the dominant ideals of American
physics that were established in the first half of the twentieth century. Chapter
1 charts the rise of the “pure science” ideal, which Henry Rowland mixed
into the mortar of the American Physical Society. Rowland saw the society
INTRODUCTION: WHAT IS SOLID STATE PHYSICS AND WHY DOES IT MATTER? 13

as a refuge for unfettered, curiosity-driven scientific inquiry that would in-


sulate physicists from questions of technological applicability or economic
relevance. The centrality of this ideal for the power brokers of the American
physics community ensured that industrial physics, as it grew throughout the
1920s and 1930s, was relegated to the periphery. The increasing relevance
of industry to the physics community, however, led industrial physics to seek
professional satisfaction. A slew of new societies and publication outlets filled
needs that the APS and the Physical Review, its flagship journal, deigned to
address.
Industrial physicists were not content to suffer their near exclusion from
the key institutions of American physics in silence, however. Chapter 2 fol-
lows the machinations they undertook as mid-century approached, while the
physics community at large set about consolidating the resources and influ-
ence it had won with its wartime labors. Improving the position of industrial
physicists required crafting a new understanding of what physics was and
how it should be organized. Whereas traditionalists viewed physics, and its
boundaries and subdivisions, as founded in the structure of the natural world,
advocates of industrial representation instead viewed disciplinary boundaries
as affairs of convention that could be restructured at will to meet contempo-
rary needs. The rise of this latter attitude paved the way for the emergence of
solid state physics, a category that made little sense according to the tradi-
tional way of looking at physics in terms of discrete classes of phenomena and
the practices used to investigate and explain them.
The pure science ideal remained a potent force in American phys-
ics through the remainder of the twentieth century, and solid state physics
emerged from the industrial insurrection against it. Chapters 3 and 4 chart
the discipline as it established its first institutions and grew into the largest
constituency of American physics. In chapter 3, I introduce the “group of
six,” an alliance of physicists determined to create institutional space for in-
dustrial and applied researchers within the APS. Led by General Electric’s
Roman Smoluchowski, the group of six organized to form what would even-
tually become the Division of Solid State Physics (DSSP), the first institu-
tional expression of the field. They would not succeed without stirring up
considerable controversy, however. The push to found a new APS division
that would be friendlier to industrial researchers led some to worry that such
efforts would compromise the society’s purpose, and therefore the unity of
American physics. Those tensions persisted in spite of attempts to resolve
them within the DSSP, and the push and pull between a desire for unity and a
14 SOLID STATE INSURRECTION

need for more specialized professional representation would define the field’s
early years.
The physics discipline’s rapid growth through the 1950s presented
pressing challenges, and these are the subject of chapter 4. Solid state physics
outstripped even the rapid inflation of the ranks of all physicists. The large
pool of applied and industrial physicists who were underserved by the APS
flocked to the new solid state division and helped establish the field’s legit-
imacy. The journal infrastructure, which struggled to accommodate expan-
sion across physics as a whole, felt the greatest pressure from solid state’s
rapid growth. Discussing the publication problem offered a means to negoti-
ate lingering disquiet about the identity of solid state physics. Some favored
establishing new publications and building stronger alliances with chemis-
try and engineering, whereas others fought hard to keep the field ensconced
in physics. The latter view would win out and solid state’s commitment to
securing its place within American physics ensured that the discipline as a
whole would come to embrace constituencies that challenged the strong pure
science ideology that defined its early decades and engage more fully with the
military, economic, and industrial needs of the Cold War.
The resolution of this issue and the beginnings of a stable professional
identity for solid state physics came just in time for conditions that would
test it. Chapters 5 and 6 both explore the influence on solid state physics of
the mid-1960s funding crunch. The US government, especially the military,
had funded all manner of scientific research in the immediate post–Second
World War years with a generosity that bordered on the haphazard. In the
mid-1960s, funding for science began to tighten. Conditions that had favored
indiscriminate growth gave way to an era of red-in-tooth-and-claw com-
petition that sowed bitterness between disciplines competing for the same
dwindling funds. The tensions between those who sought to explore solid
state’s technical potential and those who wanted to position it as a source of
fundamental physical knowledge had not resolved, even as the field’s institu-
tional situation had stabilized. These two chapters consider how this tension
led different research groups to find different niches in the shifting funding
ecology. Chapter 5 examines the possibility presented by following the lead
of high energy physics and pursuing large facilities for basic research, such as
the National Magnet Laboratory at the Massachusetts Institute for Technol-
ogy. A somewhat different opportunity, discussed in chapter 6, came in the
form of materials science, which remained a generous font of federal funding
and provided an outlet for solid state’s applied ambitions.
INTRODUCTION: WHAT IS SOLID STATE PHYSICS AND WHY DOES IT MATTER? 15

Chapter 7 introduces a sharp reaction against the technological legacy


of solid state physics: the philosophical defense of emergence developed by
Philip W. Anderson. Anderson, responding to the subordinate professional
position solid state physicists occupied in the physics community, penned
“More Is Different,” a 1972 Science article challenging the reductionist pic-
ture of the physical world that had become gospel within particle physics.
The reductionist position maintained that the most fundamental knowledge,
and therefore the most important, was to be found among the smallest con-
stituents of matter and energy. Anderson’s argument that fundamental knowl-
edge could be had at all levels of physical complexity became a rallying cry
for the solid state community. The battle for intellectual recognition would
lead some to distance themselves from solid state’s industrial roots, heighten-
ing internal tension between the requirements of funding solid state research
and a quest for intellectual esteem. Acknowledging the intellectual value of
the concepts that were necessary to appreciate the behavior of complex mat-
ter, solid state physicists argued, would necessitate rewarding their field with
both greater esteem and financial support that was not linked to technological
deliverables.
This strategy led some to abandon the name solid state physics in favor
of a new designation, condensed matter physics. Quantum mechanical treat-
ments of complex matter had developed considerably by the 1970s. They
could by then be more successfully applied to fluids—such as liquid helium—
amorphous solids, and other systems that did not submit to simplification so
readily as regular solids than they could in the 1950s, when solid state phys-
ics formed. The growing importance of these research areas made the field’s
nominal restriction to solids increasingly uncomfortable, providing all the
more reason to favor a name change. Chapter 8 traces the transition to con-
densed matter physics. The new name aimed to delineate a field that could
claim the accomplishments of solid state and could make a more serious case
that it belonged within the intellectual core of physics.
Public debates over the merits of the Superconducting Super Collider,
the focus of chapter 9, prompted both solid state and particle physicists to
defend their intellectual and professional ideals in a high-stakes context.
Particle physicists relied heavily on the reductionist rhetoric that had served
them so well during the Cold War. Conscious, though, that the context had
changed, many of them also embellished this justification with sometimes im-
modest claims about the spin-off benefits of large-scale accelerator research.
Solid state physicists rallied in opposition to what they considered an ex-
16 SOLID STATE INSURRECTION

travagance. They made an aggressive case that basic research funding could
be better spent in their own backyard. Opposition to the SSC rested on the
claims that solid state was just as fundamental as particle physics, that funding
exploratory solid state research with no strings attached would produce more
socially and technologically valuable results as a matter of course, and that
the concentration of federal physics funding in large facilities damaged other
areas of research. This view complemented the vision of physics that had
been incubated in American solid state and condensed matter physics, and
that aimed to synthesize the physics community’s long-standing pure science
ideal with a commitment to its technological and economic relevance. The
SSC’s demise, because it marked the limits of the big science program that
had dominated physics spending for decades, represented a public victory
for an alternative to the hard-line pure science outlook that had been main-
tained in part by the technical contributions of solid state physics throughout
the Cold War.
The original Star Wars trilogy tells the story of a ragtag band of misfits,
many of whom are adept at manipulating a force pervading everyday matter,
who ally to mount an insurrection against the established order and help de-
stroy a giant, partially built beam machine. The history of American solid
state physics, as chronicled in these chapters, followed much the same plot.
The field was cobbled together from a diverse assortment of research tradi-
tions, the only common element of which was a focus on the forces govern-
ing the matter that surrounds us—and how to manipulate it. Its formation
represented a rejection of the traditional power structure of the American
physics community, which exalted pure science and held applications in
lower esteem. And it came to public prominence when many of its influential
practitioners mobilized to help bring down the SSC. (The Super Collider,
admittedly, was not designed for the express purpose of destroying planets,
but some on the fringes have suggested that similar machines might have just
that effect.)25
Many solid state physicists adopted a rebel mindset, marginalized as they
were by the low status accorded applied physics and their more powerful
colleagues’ derision of their intellectual efforts. Their professional machi-
nations were calibrated to challenge this status quo. It is in this sense that
the establishment and growth of solid state physics constituted a form of re-
bellion. Much like political uprisings, the solid state insurrection responded
to specific grievances. It reflected the interests of industrial physicists, who
railed against the predominant ideals of American physics and its traditional
INTRODUCTION: WHAT IS SOLID STATE PHYSICS AND WHY DOES IT MATTER? 17

modes of professional organization. Subsequent efforts solid state physicists


mounted to rearrange the institutions of American physics sought freedom
of disciplinary association, or more equitable distribution of resources. It
would be a gross exaggeration to say that solid state physicists threw off the
hegemony of the pure physics ideal, but this need not weaken the metaphor.
Insurrections, after all, do not always lead to overthrow. They can also involve
a new integration, one that brings into the center groups whose interests were
previously on the peripheries. The conclusion of this book reflects on how
we can understand the history of American solid state physics in just that way.
1

THE PURE SCIENCE IDEAL


AND ITS MALCONTENTS

We form an aristocracy, not of wealth, not of pedigree, but of intellect


and of ideals, holding him in the highest respect who adds the most
to our knowledge or who strives after it as the highest good.
—HENRY ROWLAND, 1899

When Alexis de Tocqueville visited the American continent in the 1830s,


he was struck that “hardly anyone in the United States devotes himself to
the essentially abstract and theoretical portion of human knowledge.”1 De-
mocracy, combined with the nominal egalitarianism of American culture, he
contended, incentivized near-term practical gain over the aimless surmise
that Europe’s aristocratic traditions and hereditary wealth afforded. Abstract
knowledge was a luxury of the traditional elite, and so languished in a society
that disdained elitist traditions. When the American physics community co-
alesced at the turn of the twentieth century, it would do so in explicit opposi-
tion to this tendency, crafting a vision of “pure” science designed to keep the
pragmatism of American culture at bay and to encourage a scientific culture
that could stand alongside the established physics traditions of Europe.
Physics in nineteenth-century Europe was the province of a social and
intellectual elite. In Britain, it was dominated by veterans of the notoriously
demanding Cambridge mathematical tripos, open in practice only to well-
heeled members of the Anglican establishment.2 In the Germanic states,
physics had secured a stable place in secondary education by the mid-1800s,
but its internationally recognized researchers were drawn from the upper
echelons of society and its institutions, such as the Physikalisch-Technische
THE PURE SCIENCE IDEAL AND ITS MALCONTENTS 19

Reichsanstalt, were by and for the ruling classes.3 But neither of these au-
gust traditions claimed to be “pure” in the way the term was understood by
American physicists at the turn of the century. As abstract as the mathemat-
ics that powered British thermodynamics and electromagnetism was, it was
unapologetically linked to British industry, in particular the telegraph cables
and steam engines that sustained its global empire. And following unification
in 1871, industrial progress became a similarly potent concern for German
physicists.4 When American physicists envisioned their field as a pure intel-
lectual endeavor, above and apart from the practical demands of society, they
understood themselves to be emulating their European counterparts. To the
extent that they adopted Europe’s sensibilities, however, they did so with the
fervor of the converted. They crafted a new ideal, all the more staunch be-
cause it took the practical bent of American culture as its foil.
The pure science ideal remained a cherished element of American physi-
cists’ identity throughout the first half of the twentieth century, but this does
not mean that all or even most of the physics practiced in the United States
was remote from technological and economic concerns.5 The American
Physical Society, founded in 1899, grew quickly beyond its thirty-six charter
members, and an appreciable portion of its growth in the first few decades of
the twentieth century was in the industrial sector. Many American physicists
individually saw no reason to shy away from industry, but the institutions
of American physics worked to marginalize industrial physics and maintain
pure science as a regulative ideal. This background is crucial to appreciate the
rise of solid state physics, which represented an industrial incursion into the
pure science citadel that had been erected, and for the most part successfully
defended, through the first half of the twentieth century. If solid state physics
emerged as an institutional salve for a conflict of ideals within American phys-
ics, then addressing the ideals that suffused the early institutions of American
physics is necessary to understand the conditions that made it possible.6

HENRY ROWLAND AND THE PURE SCIENCE IDEAL


At the end of the nineteenth century, Henry Augustus Rowland, one of the few
Americans to enjoy an international reputation in physics, observed much the
same state of affairs as Tocqueville. The practically minded Thomas Edison
was the public face of American science, and Rowland lamented that “much
of the intellect of the country is still wasted in pursuit of so-called practical
science which ministers to our physical needs but little thought and money
20 SOLID STATE INSURRECTION

is given to the grander portion of the subject which appeals to our intellect
alone.” Rowland and thirty-five others founded the American Physical Soci-
ety (APS) to minister to the intellect.7
Rowland had spent 1875–76 studying in Europe, during which time his
views on the proper conduct of science crystallized.8 He worked in Herman
von Helmholtz’s Berlin laboratory, where he grew to admire the German re-
search university and the continental tradition of theoretically oriented sci-
ence pursued by a cultural and intellectual elite. These experiences served
him well in the appointment he assumed on his return to the United States,
at the newly formed Johns Hopkins University, which was founded with a
research mandate. In Rowland’s eyes, universities were only one element of a
strong scientific community, which also required robust professional institu-
tions to act as a bulwark against the economic enticements to technical work
that suffused American culture. He understood “pure science”—by which he
meant unfettered pursuit of truth about the natural world, free from the de-
mands of immediate useful application and the allure of pecuniary reward—
not only as a prerequisite for scientific truth, but also as a moral imperative.
“Let us hold our heads high with a pure conscience while we seek the truth,”
he implored his fellow physicists at the second meeting of the APS.9
The categories of “pure” and “applied” science as Rowland understood
them were at once clearly delineated and inextricably linked. Rowland might
have scorned the Edisonian pursuit of profit, but not so much that he sought
to deny the place of scientific knowledge as a wellspring of novel know-how.
Rather, he insisted that the pursuit of knowledge could only function properly
and yield those benefits when it was insulated from the diversionary influence
of mercantilism.10 And he thought that strong professional institutions could
provide such protections.
Rowland died in 1901, but his elitism was woven into the fabric of the
American Physical Society, which began as a distinctly ivory tower institu-
tion.11 In 1902, the society’s third year, only 4 of its 144 members reported
job titles or affiliations that reflected industrial employment.12 Industrial
membership grew gradually in the following decades. As of July 1920, ap-
proximately 60 percent of its membership was affiliated with academic insti-
tutions, compared with about 24 percent in industry and 9 percent in gov-
ernment jobs.13 Despite such growth in the industrial sector, the officers of
the society all claimed university affiliations in 1920, as did the vast majority
of the seventeen-member council, which included just two employees of gov-
ernment laboratories and one representative from industry.14
THE PURE SCIENCE IDEAL AND ITS MALCONTENTS 21

Growth in the society’s industrial membership, and the contrasting con-


tinuity of its academic leadership, indicate the extent to which the APS was
an island, and sought to remain one. Historians have consistently and force-
fully challenged the sharp distinction between pure and applied science that
Rowland sought to maintain (and the post–Second World War distinction
between basic and applied research that was its heir). As David F. Noble
observed in 1977, and many others have reinforced since, the late nine-
teenth and early twentieth centuries witnessed a closer connection between
science and technology. The resulting emphasis on science-based, profit-
oriented industrial production led Paul Lucier to dub it the Tinseled Age. A
basis in physics and chemistry became a mark of professional identity for the
emerging engineering disciplines and industrial development became firmly
linked with scientific research, especially during and in the wake of the First
World War.15
Nor did this science–industry alliance emerge in opposition to an exist-
ing, entrenched pure science ideal. Graeme Gooday observes that the visions
of pure science championed by Rowland in the United States and by Thomas
Henry Huxley in Britain were new. As much as Rowland understood his vi-
sion to be rooted in the great intellectual traditions of Europe, the idea that
taxpayers and philanthropists had an obligation to bankroll intellectual pur-
suits unmoored from the practical questions of the age was novel.16 A selec-
tive reading of history of the physical sciences led Huxley to conclude that
“practical advantages . . . never have been, and never will be, sufficiently at-
tractive to men inspired by the inborn genius of the interpreter of Nature, to
give them courage to undergo the toils and make the sacrifices which that
calling requires from its votaries.”17 Thermodynamic and electromagnetic
theory, the hugely successful and intellectually revered physical accomplish-
ments of Huxley’s own age, in fact owed both a material and intellectual debt
to the expanding technological infrastructure of the industrial revolution.18
The APS, which thumbed its nose at industrial capitalism, proceeded
with the firm conviction that influence did and should flow from abstract
knowledge to practical use, but not the other way around. That position did
not get easier to maintain as the twentieth century wore on. The First World
War in particular strengthened the connection between abstract research and
technical implementation that Rowland and his cohort had hoped to resist.
The demands of the First World War led many to lament the artificiality of
the pure/applied division. John J. Carty used his presidential address to the
American Institute of Electrical Engineers in 1916 to observe:
22 SOLID STATE INSURRECTION

Arising out of this agitation comes a growing appreciation of the importance


of industrial scientific research, not only as an aid to military defense but as
an essential part of every industry in time of peace. . . . I consider that it is the
high duty of our institute and of every member composing it, and that a similar
duty rests upon all other engineering and scientific bodies in America, to im-
press upon the manufacturers of the United States the wonderful possibilities
of economies in their processes and improvements in their products which are
opened up by the discoveries in science.19

The University of Chicago botanist John Merle Coulter contended, “The


public has begun to recognize the fact that pure and applied science are not
mutually exclusive fields of activity, but complementary, and therefore public
support for pure science has been growing, and as a consequence the prac-
tical achievements of pure science in connection with the war, it bids fair to
enter upon its own public estimation and support.”20 To those outside of
physics, and to some within, the relationship between science and industry
appeared reciprocal rather than hierarchical.
That perception was bolstered by a genuine strengthening of the connec-
tion between physics and industry. Industrial research laboratories blossomed
during the first decades of the twentieth century. The General Electric Re-
search Laboratory, founded in 1900, offered a proof of concept that inspired
other industrial concerns to invest in research. The AT&T Bell Telephone
Laboratories, established in 1925, fostered a research culture that encour-
aged its scientists to remain open to the unexpected in the hopes of maintain-
ing its advantage in electronic communications technology. Bell became the
first industrial laboratory to produce a Nobel Prize winner in physics when
Clinton Davisson won in 1937 for a 1927 experiment that demonstrated elec-
tron diffraction, confirming a theoretical prediction of Louis de Broglie.21
The wave behavior of electrons illustrates the blend of pure and applied
research fostered at Bell. It was at once a fundamental physical discovery and
something a telephone company, which was in the business of transmitting
electrons, would very much like to know. The promise of the possibility, at
least, to conduct curiosity-driven research, along with the higher salaries
the private sector could offer, lured many physicists into industry—and the
prominence that Bell in particular achieved, in part on account of Davisson’s
prize, positioned it to become the most influential solid state physics hub
after the Second World War. Through the 1920s and 1930s, industrial lab-
oratories not only employed an appreciable proportion of American physi-
THE PURE SCIENCE IDEAL AND ITS MALCONTENTS 23

cists but also generated an appreciable proportion of the papers published


in American physics journals.22 The cultural differences between industry
and academia nevertheless remained sharp. Industrial research, even when
directed toward scientific insight, required its own style, one that understood
insight and applications to be of a piece, and that did not rank them within a
strict value hierarchy.23
Although American industry was becoming much enamored of physicists,
American physics—or at least its flagship society—had little affection for in-
dustry. The prevailing attitude through mid-century is reflected in a piece of
doggerel that made the rounds at MIT’s Radiation Laboratory in 1944, cele-
brating Isidor Isaac Rabi’s Nobel Prize by praising his self-abrogative disdain
for the comparative riches available to physicists who went corporate:
Now all you bright young fellows with your eyes upon the stars,
You graduate assistants who subsist on peanut bars
If industry should woo you with two hundred bucks a week
Refuse the job and say, without your tongue in your cheek,

It ain’t the money


It’s the principle of the thing
It ain’t the money
There’s things that money can’t buy
It ain’t the money
That makes the nucleus go round
It’s the philosophical ethical principle, we keep telling ourselves, of the thing.24

The song conveyed the sense of moral superiority that came with resisting
the higher salaries industry was able to offer, along with the consensus that
the most interesting intellectual work remained the province of university
research.
Within a context that favored closer contacts between science and com-
merce, the pure science ideal that the APS staunchly maintained prompted
institutional growth elsewhere in American physics. Three new societies
formed between 1916 and 1929 representing narrower specialties, each
with a prominent focus on instrumentation and/or applications. The Optical
Society of America grew out of Eastman-Kodak’s research laboratories and
was intended to serve the needs of a growing group of industrial researchers
studying interactions between light and matter.25 The Acoustical Society of
America, whose first meeting was hosted at the Bell Telephone Laborato-
24 SOLID STATE INSURRECTION

ries in 1929, also had industrial roots and was directed toward engineering
interests.26 That same year saw the formation of the Society of Rheology,
dedicated to the newly named science that studied the deformation of mat-
ter. Its founders embraced their field’s potential applications, noting in their
announcement of the new society: “Heraclitus was probably correct in say-
ing that ‘everything flows’ and the major problems of great industries dealing
with nitrocellulose paint, varnish, artificial textiles, metals, rubber, etc., have
to do with elastic deformation and flow.”27
None of these groups would have found a warm welcome in the APS.
Wallace Waterfall, founding member of the Acoustical Society, recalled that
“if anybody had come along then with the idea of setting up divisions of the
Physical Society and having the Acoustical Society become one of those divi-
sions, why, that wouldn’t have gone over at all.”28 The Physical Society’s con-
scious decision to spurn applications created a need for new professional out-
lets that served the growing community of applied and industrial physicists.

ACCOMMODATING APPLIED PHYSICS


Members of application-oriented professional organizations who were
trained in physics continued to think of themselves as physicists, even if the
APS was not their professional home. If it hoped to remain a cohesive com-
munity, American physics would need a larger corral. On May 3, 1931, rep-
resentatives from the three newly formed societies and the APS assembled at
the Cosmos Club in Washington, DC. The occasion was the first meeting of
the American Institute of Physics (AIP).29 The new organization was princi-
pally a publishing operation coordinating the collective print output of these
organizations—and, some months later, of the American Association of Phys-
ics Teachers, itself newly established. The AIP, an organization of organiza-
tions, must have seemed unwieldy to many in a community that until recently
had been so small. It was a recognition, however, of “the fact that there was
then no single society that drew together all those whose primary scientific
interest was in the field of Physics,” as the committee charged with planning
the AIP’s post–Second World War activities would recall in 1945.30
The AIP was a concession to the fact that American physics was becom-
ing larger and more diffuse. Abraham Pais, in his reflections on the Physical
Review—the undisputed journal of record of American physics for the better
part of the past century—recalled that in the 1930s dedicated physicists could
still read “the green monster,” as it was known, cover to cover and maintain a
panoramic view of the field.31 Even in 1931, however, that level of dedication
THE PURE SCIENCE IDEAL AND ITS MALCONTENTS 25

Articles in Physical Review, 1920 –1960


1400

1200

1000
Articles Published

800

600

400

200

1957
1924
1921

1931

1940

1942
1943
1944
1945
1946
1947
1948
1949

1951
1934

1954
1920

1922
1923

1925
1926

1928
1929
1930

1932
1933

1935
1936

1938
1939

1950

1952
1953

1955
1956

1958
1959
1960
1937
1927

1941

Year

Figure 1.1. Number of articles published in Physical Review between 1920 and 1960.

was rare and exercising it beyond the capabilities of all but the most dogged
readers.
Pais’s recollections must also be considered within the broader context
of American physics publishing. Before 1929, the Physical Review and the
Journal of the Optical Society of America were the only dedicated outlets for
scholarly work in physics in the United States. That changed with the ap-
pearance of several new journals in the late 1920s and early 1930s. Reviews
of Modern Physics sought to make contemporary research more digestible by
summarizing lively areas in short review articles.32 Review of Scientific In-
struments launched in 1930, by which time the publication expansion was
already jarring to some. In the editorial that began the inaugural issue, Floyd
K. Richtmyer acknowledged: “The number of scientific and technical peri-
odicals to which any worker in either pure or applied science must refer has
increased so rapidly in recent years as to raise in some minds the question of
the desirability of taking steps to discourage the starting of new journals.”33
The background to his remarks included publications such as Journal of the
Acoustical Society of America and Journal of Rheology, both founded in 1929.
Nor was Review of Scientific Instruments the last American physics journal
to appear around 1930—Physics, the Journal of Chemical Physics, and the
American Physics Teacher would follow between 1931 and 1933.
The number of articles the Physical Review published reached a local
26 SOLID STATE INSURRECTION

maximum in 1931, when the new journals stemmed what had been a steady
rise through the 1920s (figure 1.1). It would not reach the same level again
until after the Second World War.34 The panoramic view of physics one
would achieve by diligently taming the green monster twice monthly thereby
changed in the 1930s. The landscape shifted as whole areas of physics mi-
grated to new outlets and the Physical Review focused more intently on keep-
ing abreast of new and exciting developments in nuclear physics and quan-
tum mechanics.
Considering the experience of an archetypical Physical Review–reading
APS member over the period from 1925 to 1935 exposes a clear shift. A typ-
ical issue of the journal in the mid-1920s included some theoretical work,
including papers on quantum phenomena, but it also published a great many
articles that would have found a home in more specialized journals just a few
years later. By the mid-1930s, the theory quotient was higher and the journal
was dominated by nuclear and quantum papers. The growth of new outlets
in intervening years did not threaten the Physical Review’s status as the com-
munity’s journal of record, but it did mean that its profile was more sharply
defined than it had been. It was no longer a general interest journal, at least
not so far as the growing constituency of industrial and applied physicists was
concerned. Anyone operating on the assumption that the APS represented
American physicists and that the Physical Review published what was im-
portant to know about current physics would have perceived a sharpening
of Henry Rowland’s pure science mission, rather than a dilution, even as the
importance of applied physics grew within the rest of the community.
The journal’s reputation also changed markedly over this span. John Van
Vleck later recalled: “The Physical Review was only so-so, especially in the-
ory, and in 1922 I was greatly pleased that my doctor’s thesis was accepted for
publication by the Philosophical Magazine in England. . . . By 1930 or so, the
relative standings of The Physical Review and Philosophical Magazine were
interchanged.”35 John Torrence Tate became editor in 1926, the same year
full-blooded quantum mechanics emerged in Europe. Van Vleck and others
credited this turnaround to Tate’s eager embrace of nuclear and quantum
physics, both of which advanced rapidly in the 1920s and 1930s. Van Vleck
coauthored a biographical memoir for the National Academy of Sciences
with Tate’s doctoral student, mass spectroscopist Alfred Nier, in which they
praised Tate for showing “rare judgment and common sense in not delay-
ing by much refereeing noteworthy papers dealing with various applications
of quantum mechanics; this was important, for America was somewhat at a
THE PURE SCIENCE IDEAL AND ITS MALCONTENTS 27

disadvantage compared to the centers of Europe, where the [quantum] rev-


olution had germinated.”36 Pursuing a more narrowly theoretical focus and
tackling the exciting changes initiated in Europe raised the Physical Review
to a journal of international standing, but at the expense of its relevance to a
broad swath of its domestic readership.
Tate was wise to the fact that Physical Review had narrowed its scope on
his watch. With this in mind, he launched a new journal, Physics, in 1931. It
was designed to serve the growing cadre of applied physicists, whose publi-
cations until that point had been “scattered through a number of engineer-
ing, chemical and industrial journals.” Physics, which would be renamed the
Journal of Applied Physics within a few years, hoped to attract contributions
from applied researchers who felt alienated by the narrowing focus of Physi-
cal Review and to reaffirm their identity as physicists. Tate commented on the
publication’s mission:
The Physical Review has hitherto been the only outlet provided by the Amer-
ican Physical Society for the publication of original research. As you may
have noted the Review has more than doubled in volume during the past six
years and has become more and more the exponent of the purely introspective
side of physics. This is but the natural result of the fundamental and radical
changes in the logical framework of the science which have attracted the atten-
tion of an ever increasing group of active physicists. But fascinating as are the
developments in atomic physics and the quantum mechanics, they do not by
any means represent the whole of the science, nor are they more interesting
or valuable than the original work of the greater number of professional phys-
icists who are applying physical methods and principles to the problems of
other sciences and the industries.37

Physics was both an overture to applied physicists and an act of demarcation


that placed them outside its center. Applied research would not be left to dif-
fuse into chemical and engineering journals and the industrial gray literature,
but neither would it be granted a berth on the flagship vessel of the American
Physical Society.

THE MAP OF PHYSICS, 1939


In 1939, Bernard H. (Bern) Porter drew a “Map of Physics” (figure 1.2).
Porter was a graduate student in physics at Brown University. The following
year he would be enlisted into the Manhattan Project, on which he worked at
Princeton University and Oak Ridge until he quit three days after the bomb-
28 SOLID STATE INSURRECTION

Figure 1.2. Bern Porter’s Map of Physics, 1939. The caption reads: “Being a Map of
Physics, containing a brief historical outline of the subject as will be of interest to phys-
icists, students, and laymen at large. Also giving a description of the land of physics
as seen by the daring souls who venture there. And more particularly the location of
villages (named after pioneer physicists) as found by the many rivers. Also the date of
founding of each village. As well as the date of its extinction. And finally a collection of
various and sundry symbols frequently met with on the trip.” Reproduced with per-
mission of Mark Melnicove, literary executor for Bern Porter, mmelnicove@gmail.com.
From Bern Porter Collection, Colby College, Special Collections, Miller Library, Water-
ville, Maine

ing of Hiroshima, traumatized and disillusioned with physics. He would in-


stead follow his passion for art, which he used as an outlet for his lifelong
struggle with the responsibility he felt he shared for the development of nu-
clear weapons. But in 1939 he remained enamored of physics. His map cel-
ebrated the pioneering instinct of the “daring souls” who venture into the
frontiers of physical knowledge. It also encoded an idea about what physics
was that was central to the American pure science ideal.
Porter’s map is both an artifact and a literal illustration of the habits of
mind that relegated applied and industrial research to the fringes of the
American physics community. It represents the various provinces of physics
—mechanics, sound, electricity, magnetism, light, heat, and astronomy—as
geographical regions linked to one another by energy, depicted as a river fed
THE PURE SCIENCE IDEAL AND ITS MALCONTENTS 29

Figure 1.2a. From Bern Porter’s Map of Physics, 1939.

Figure 1.2b. From Bern Porter’s Map of Physics, 1939.

by mechanical and electromagnetic tributaries (and a reservoir of radioactiv-


ity), feeding into an ocean labeled “Research: The Future of Physics” (figure
1.2b). Physics, thus represented, is conceptually unified and historically con-
tinuous, defined by phenomena that existed in the world (figure 1.2a). Phys-
icists, who give their names to the villages dotting the landscape, are the ones
who expose those phenomena and deduce the rules governing them. Tech-
nology is a foreign land in this rendition. Physics is out there in the world;
physicists are those daring souls called to discover it on the downstream
journey toward the frontiers of research. Conceiving of physics in this way
encouraged the organizational assumptions that precluded industrial physics
from gaining purchase in the American Physical Society.
This outlook was further ill-equipped to accommodate a field like solid
state physics. It is difficult to imagine how Porter might have represented
solid state had he updated the map a decade later. Solid state physicists hailed
from almost all the distinct geographical regions represented on the map. The
30 SOLID STATE INSURRECTION

difficulty of locating their work within this visual schema illustrates the fact
that solid state physicists did not explore a discrete region of physics in the
traditional sense. Solid state physics was not a self-contained assembly of top-
ics and methods that could be conveniently represented as a river, island,
continent, or other natural outcropping of the disciplinary landscape.
Porter channeled an ethos of classification that was characteristic of the
previous century’s science. Nineteenth-century natural philosophers often
understood taxonomy as an essential piece of their mission, operating from
the conviction that proper classification could reveal the order inherent in
nature. The same ethos extended to classifying the sciences themselves.38
William Whewell wrote in 1840: “A sound classification must be the result,
not of any assumed principles imperatively implied to the subject, but of an
examination of the objects to be classified;—of an analysis of them into the
principles in which they agree and differ. The Classification of Sciences must
result from the consideration of their nature and contents.”39 The assumption
that the sciences themselves, like the objects of their study, possessed intrinsic
features that allowed them to be distinguished naturally and unambiguously
continued into the twentieth century and shaped attitudes among American
physicists. That assumption encouraged resistance to categories like indus-
trial physics, and indeed solid state physics, which did not slot neatly into a
perceived natural order.
Respect for that perceived order guided the institutions of American phys-
ics in the first half of the twentieth century, even while industrial physicists
became an appreciable proportion of the community. By 1933, the institu-
tional landscape had attained a local equilibrium. The AIP administered the
eight American physics journals that remained after the Journal of Rheology
ceased publication in 1932. The ranks of the APS, which remained the prin-
ciple society for American physics, swelled in response to the frisson gen-
erated by nuclear and quantum physics, generous foundation support that
allowed physicists with degrees from elite universities to supplement their
studies in Europe, and the influx of émigrés fleeing the clouds gathering over
Central Europe. Much of this growth reinforced the society’s commitment to
pure science.40 The increase of the community’s size, and the role of indus-
trial and applied physics within it, nonetheless represented a continuing chal-
lenge to the traditional ideals of the discipline. It was less the potency of the
pure science ideal itself than it was its entrenchment in key institutions that
made it difficult to dislodge. The power brokers of American physics, who
THE PURE SCIENCE IDEAL AND ITS MALCONTENTS 31

understood its goal to be the extraction of raw knowledge from nature, at-
tempted to keep the field pure with the battlements of institutional structure.
Those battlements would hold until the pressures of the Second World
War precipitated further reorganization. Solid state physics, when it emerged
in the late 1940s, constituted the first successful insurrection against pure
science fundamentalism in the APS. The establishment of a discipline that
catered to the needs of industrial physicists and was organized in a way that
paid little heed to natural categories represented the most substantive changes
to the foundational identity and ideological commitments of the American
physics community since its founding. The following two chapters explore
how this challenge to traditional community ideals unfolded.
2

HOW PHYSICS BECAME


“WHAT PHYSICISTS DO”

When Solomon said that “a good name is rather to be chosen than great
riches,” he knew what he was talking about.
—OLIVER E. BUCKLEY, 1944

“Physics” was the name Oliver E. Buckley, president of Bell Laboratories,


had in mind when he opened his address to the National Research Coun-
cil’s Conference of Physicists with the above line. Buckley worried that this
term evoked nothing concrete to the average American. The nuclear bombs
detonated over two Japanese cities would fling physicists into the forefront
of American public consciousness in August 1945, but the Manhattan Proj-
ect remained shrouded in secrecy when Buckley spoke in May 1944. Well-
publicized contributions to the war effort would likely have made radar famil-
iar to a substantial segment of the American public, but it was rarely linked
to physics in the popular press.1 Albert W. Hull, the outgoing president of
the American Physical Society, had remarked earlier that year, “It is a rare
occurrence that a census taker has ever heard of a physicist, and the task of
explaining is such that one is often tempted to register as a chemist.”2 Faced
with this deficit, Buckley insisted that professional identity was the most pri-
mal challenge American physicists faced and encouraged the conference at-
tendees to consider how it shaped activities from undergraduate teaching to
government advising. He asked his colleagues to confront difficult questions
about who they were and what they did. What was physics? Who could claim
to be a physicist? Who got to decide?
Daniel J. Kevles’s often-quoted response to these questions is “physics is
HOW PHYSICS BECAME “WHAT PHYSICISTS DO” 33

what physicists do.”3 Though it might seem tautological, this slogan makes
the serious point that historians, pace Whewell, should avoid the impulse to
seek some essential, context-independent core of the fields they study. Dis-
ciplines are historically contingent social entities that might be assembled in
different ways at different times by different people acting on different mo-
tives. The contingency of human-built categories is gospel to today’s histori-
ans, who are likely to accept this point unflinchingly. The same, though, can-
not be said for the people who constructed those categories in the first place.
Many mid-twentieth-century physicists, notably those who controlled
the American Physical Society (APS), would have bristled at the suggestion
that their field consisted in anything other than a set of preexisting empirical
regularities, which they took as their task to discover and formalize. Under-
standing “physics” to refer to something existing in the world, they might
have suggested a different slogan: physics is what physicists pursue. To these
traditionalists, who were also the fiercest defenders of the pure science ideal,
“physics” remained a fixed set of natural phenomena, whose structure de-
termined who was a physicist and who was not. They suggested that anyone
whose principal interest was not the discovery and elaboration of general
physical principles belonged more properly in engineering, chemistry, metal-
lurgy, or another field.
As the community of American physicists grew, however, populating in-
dustrial laboratories and seeking concessions to the demands of technical and
applied work, a larger portion of the community began to insist that physi-
cists, not nature, held authority over the scope and organization of physics.
Stanford University’s William W. Hansen, replying in 1943 to his colleague
David L. Webster’s suggestion that physics was defined by the pursuit of nat-
ural physical laws, wrote: “I wouldn’t want to attempt a precise definition,
but it would seem that your criterion sets the sights terribly high. How many
physicists do you know who have discovered a law of nature? You have, I
know, and so has Compton and perhaps one or two others I don’t know or
think of. But really, it seems to me, this privilege is given only to a very few of
us. Nevertheless the work of the rest is of value.”4 The rest tended to agree.
Perspectives more sympathetic to the Kevles dictum began to exert their
influence in the mid-1940s. Some outside the traditional core of academic
physics, and even some within, came to understand their discipline as a com-
munity with the latitude to define and redefine itself as it proved convenient.
The circumstances of the mid-1940s presented opportunities that made re-
evaluating the traditional definition of the physicist convenient indeed.
34 SOLID STATE INSURRECTION

Buckley’s 1944 lecture addressed the anxieties that came to the fore as
these competing agendas for defining physics were brought into tension by
the centrifugal effects of growth and diversification. Advocates of both per-
spectives developed a number of strategies for answering the questions Buck-
ley posed, and these competed for traction as American physicists assayed
the possibilities that the post–Second World War environment would offer
them. By the time the American physics community found its footing in the
postwar era, the understanding that physics was “what physicists do”—and
so could change if physicists started to do different sorts of things—had mo-
tivated changes in the discipline’s institutional structure, with wide-ranging
consequences for how it would develop through the Cold War.
With the foregoing in mind, I propose my own, equally glib alternative to
the Kevles dictum: physics is what physicists decide it is. American physicists
in the 1940s did not merely realize that physics could be understood as the
sum of their activities, or a relevant subset of them; they seized upon their
agency to organize their discipline so as to proactively delineate the types
of activities that fit within it. That agency scarcely needed to be exercised
in the era when the community was small and divisions within it slight. As
the population boomed and the institutional character of physicists’ employ-
ment changed entering the mid-century, however, that agency became criti-
cal.5 Physics became what physicists did during the Cold War by means of
concerted efforts on the part of a small group of individuals working within
the institutional constraints of the American Institute of Physics (AIP) and
the APS. Their vision was opposed by a group of influential traditionalists
who maintained that simply earning a degree in physics did not make one a
physicist—or at least was not sufficient for one to remain a physicist—and
actively sought to keep some of the things that physicists did from working
their way into the definition of the field.

PHYSICS AS NATION: THE UNITED STATES OF PHYSICS


In the early 1940s, “the physicists’ war” raged. It had earned that name even
before the transformative influence of nuclear weapons research became
clear.6 Physicists recognized that the Second World War would bring them
new relevance to American society; in 1943 and 1944, as the war’s end came
into focus, a scramble to shape postwar physics began. Opinion pieces dis-
cussing existing challenges and future development of physics peppered
physics journals. The relationship between industry and academia garnered
significant attention. The arrangement reached in the 1930s, which allowed
HOW PHYSICS BECAME “WHAT PHYSICISTS DO” 35

the APS to rededicate itself to pure research while the raft of societies serving
applied interests was held in loose confederation by the AIP, was unsatisfac-
tory to many. Wartime nationalistic rhetoric was in the air. Physicists breathed
it in, exhaling it again in their discussions of their discipline’s future. Those
who favored a more ecumenical approach to defining physics, which the AIP
embodied, rallied around the notion of a “United States of Physics,” com-
posed of many provinces, each with its own local character, held together by
a political commitment to the unity of the discipline. In the face of a growing
population of industrial researchers, the question of what position industrial
physics would occupy in this union loomed large.
The stakes of debate over industry’s place in postwar physics were, in
the mid-1940s at least, more ideological than practical. Only after the war
would the wider physics community gain a clear understanding of how the
abundance of federal dollars would reshape their research. Furthermore,
physicists weathered the ravages of the great depression better than most,
and thus funding was not so prominent a concern as it would become just
a few years later.7 Discussions in the final years of the Second World War,
carried out almost exclusively by those physicists not sequestered conducting
secret research for the Manhattan Project, turned instead on the question of
dignity. For adherents to Rowland’s traditional pure science ideal, physics
maintained its disciplinary dignity through its status as a calling, rather than
a profession. But the growing mass of applied and industrial researchers for
whom physics was a profession first—those who saw physics as an intellectu-
ally rewarding path to a comfortable middle-class existence, and who would
drive what David Kaiser has called “the postwar suburbanization of Amer-
ican physics”—saw their professional dignity impugned by calls that their
interests be kept at a remove from the real business of physics.8
Although they differed in their preferred solutions, both these groups
shared the concern that industry and academia had drifted too far apart. The
issue appeared in many screeds on the future of physics.9 In 1943, Thomas
H. Osgood of Michigan State College of Agriculture and Applied Science (as
Michigan State University was then known) put the problem thus: “Both in
the past and now, technical physicists have known too little about the work,
both in research and teaching, in which their academic colleagues are en-
gaged; and an even more lamentable ignorance of the practical problems of
the age which are being solved by physicists in industry has been displayed
by those who train students in the rudiments of physics in our educational
institutions.”10 Osgood voiced frustrations that many industrial researchers
36 SOLID STATE INSURRECTION

felt as they tried to maintain their identity as physicists within a field that
considered applied work intellectually subordinate.
Many similar expressions of frustration appeared in the Review of Scien-
tific Instruments column “Contributed Points of View,” in which physicists
could sound off about professional issues—a function that Physics Today
would assume in 1948. In the February 1945 issue, Morris Muskat of Gulf
Research Laboratory gave a detailed account of the professional challenges
that industrial physicists faced and described the ways in which existing
institutional structures failed to serve them. The industrial physicist “must
work intimately with the chemist, the electronics engineer, the acoustical en-
gineer, the color expert, the hydraulics engineer, or whoever has given him
the problem and will make use of its solution when achieved,” Muskat wrote.
Industrial physicists generally worked on problems of someone else’s design
and for companies that were unlikely to pay expenses for conference travel.
Working in this context, Muskat suggested, alienated industrial researchers
from their academic colleagues who dominated the APS, membership in
which became, for the industrial physicist, “a traditional ‘hangover’ from his
youthful professional pride of his school days,” serving no useful professional
function.11
Muskat urged the APS to become more responsive to its industrial mem-
bers, especially by supporting the formation of special interest divisions. This
suggestion was an alternative to a proposal popular among many applied
physicists that would have seen the AIP replace the APS as the official orga-
nization of American physicists. The principle advocate of this solution was
Gaylord P. Harnwell, the head of the physics department at the University of
Pennsylvania, where he would later serve as president. As editor of Review
of Scientific Instruments, Harnwell enjoyed a front-row seat to the wrangling
over the future of physics that played out in the journal, and he used his edi-
torial pulpit to nudge it in the directions he found most productive.
In an editorial published in August 1943, Harnwell set down the chal-
lenges as he saw them and articulated his favored solution. The problems
facing physics included the large number students, both undergraduate and
graduate, whose training had been derailed by war work and who would
emerge from their wartime assignments without adequate foundational train-
ing.12 “There will be more physicists after the war, but the great majority of
them will have the technical or craftsman’s attitude toward the science rather
than the professional or academic point of view,” Harnwell predicted. This
posed a challenge, especially alongside the increased social and political
HOW PHYSICS BECAME “WHAT PHYSICISTS DO” 37

cachet physics was beginning to enjoy. Harnwell’s principle concern was


that the technical accomplishments of physics, which were relatively easy
to communicate to nonspecialist audiences, would come to dominate pub-
lic awareness of the field and undermine support for abstract research. This
concern, along with worries about establishing effective training programs
and finding stable sources of research support, led Harnwell to conclude that
professional organization should be at the forefront of physicists’ discussions
about how to conduct their affairs after demobilizing.13
A few months later, Harnwell offered his affirmative response to these
troubling questions. He called for “a more perfect union” in his editorial
opening the February 1944 issue of Review of Scientific Instruments, and pro-
posed “a new American Physical Society to replace the existing associated
Societies and to include all members thereof for the purpose of broadly and
liberally advancing the science and technology of physics in all its branches.”
Echoing others who lamented institutional rifts between the search for general
principles and their applications, Harnwell insisted: “We must see to it also
that no schism develops between the academic and the industrial physicist.
In some respects the former tends to be the specialist and the latter the gen-
eral practitioner or clinician, but without the unification of both their efforts
in a common society, neither can be fully effective.”14 Harnwell fretted that
both academic and industrial physics might suffer if either could not draw on
the other to achieve its goals, a threat he found severe enough to propose a
dramatic overhaul of the professional structure of physics.
Harnwell championed a federalist view of American physics. He advo-
cated a central organization more inclusive than the APS but stronger than
the AIP, which he considered too loose a confederacy to effectively serve all
physicists. He wrote favorably of a potential industrial division within the
proposed new society, observing that a similar entity in the American Chem-
ical Society had grown rapidly, serving as a nursery for a number of spin-off
divisions. This vision of physics would give wide latitude for new groups to
form and define novel areas of interest, while holding them together under the
aegis of a strong central organization.
Harnwell struck a nerve; the invitation closing his editorial to comment
further on how American physics might promote unity garnered many re-
plies, some brief, others expansive, over the next several months. Wallace
Waterfall and Elmer Hutchisson responded favorably to the call for unity,
using their own editorial soapbox in Journal of Applied Physics, and in so
doing made Harnwell’s implicit federalism explicit.15 Their May 1944 edi-
38 SOLID STATE INSURRECTION

Figure 2.1. Map of the “United States of Physics.” The initialisms expand as follows:
ASME (American Society of Mechanical Engineers); ASH & VE (American Society of
Heating and Ventilation Engineers); AIEE (American Institute of Electrical Engineers);
IRE (Institute of Radio Engineers). Reproduced from Wallace Waterfall and Elmer
Hutchisson, “Organization of Physics in America,” Journal of Applied Physics 15, no. 5
(1944), 407–9, with the permission of the American Institute of Physics

torial revolved around a diagram they titled the “United States of Physics”
(figure 2.1). “Obviously a strong central organization in physics is needed,”
they concluded on the basis of the apparent fragmentation in the diagram. “A
single society with many subject matter divisions or a ‘union’ of many ‘states’
might accomplish the desired unity in physics provided the proper balance
between ‘federal’ power and ‘states’ rights’ is maintained.” Meetings and pub-
lications, they argued, could be the province of the “states,” whereas issues
of broad interest to physicists and the “‘colonization’ of virgin territory”—
represented by the dashed empty space on the right of the diagram—could be
left to the central organization.16
Not all those covered by this unification were enthusiastic about their
inclusion. Alfred N. Goldsmith, longtime RCA research engineer, wrote to
the editor on behalf of the Institute of Radio Engineers (IRE), which he had
cofounded in 1912, and which is represented in Waterfall and Hutchisson’s
map as dealing with the applications of electronics.17 Goldsmith complained
that Waterfall and Hutchisson had inappropriately characterized the “related
societies” as narrow, far-flung provinces of American physics: “It is not un-
HOW PHYSICS BECAME “WHAT PHYSICISTS DO” 39

usual for a society or institute to define or delimit its own activities and to
determine, in its own best judgment, the field of its activities. However, it is
perhaps unique to find one learned society gratuitously defining the scope
of other scientific organizations.” In defiance of the narrow characterization
of the IRE’s role in the “United States of Physics,” Goldsmith observed: “Of
necessity, problems of mechanical construction, optical theory and design . . .
and acoustical theory . . . have been exhaustively treated in the Proceedings of
the I. R. E.”18
Goldsmith’s ire exposed weaknesses in the federalist approach that ulti-
mately doomed the idea of a strong, unifying umbrella society to replace the
AIP. A plan that ceded overall control of fundamental research to the Ameri-
can Physical Society ran afoul of other societies’ interests. Organizations like
the IRE aimed not just to support applied research, but to foster a give and
take between fundamental and applied research that was rare within the con-
fines of the APS. The unification scheme also drew the outer boundaries of
physics too sharply for some and failed to account for the roles played by
smaller societies that expanded into other fields. In short, the strong federal-
ism of Harnwell, Waterfall, and Hutchisson elided the real benefits fragmen-
tation had provided those specialties that were poorly served by the APS and
underestimated the extent to which subjugation under an umbrella society
would be unwelcome among these constituencies.
Goldsmith’s entry into the conversation is notable for another reason: as
an electrical engineer, he represented an explicit fear of the helmsmen of the
inchoate field of solid state physics. Electrical engineering and physics, once
of a piece, had drifted apart early in the twentieth century.19 Some physicists
worried that the study of solids would follow the same path, losing the poten-
tial to benefit from heightened postwar prestige and isolating itself from prob-
lems of purely theoretical interest. General Electric’s Roman Smoluchowski,
in his 1943 manifesto initiating the effort to found the APS Division of Solid
State Physics, anticipated the increased demand for physics training after the
war, especially in new subfields concerned with complex matter, and insisted,
“We would like them to remain branches of physics rather than to become
new . . . types of ‘engineering.’”20
Similar commentary on the looming institutional changes in American
physics dotted the letters section of the Journal of Applied Physics and Re-
view of Scientific Instruments over the months following the publication of
Waterfall and Hutchisson’s editorial. Applied physicists, for the most part,
favored a federalist approach in which a large umbrella society with loose
40 SOLID STATE INSURRECTION

membership criteria would confer unity on physics without organizing that


unity around a central spire of pure research. Three physicists from the Case
School of Applied Science (which would merge with Western Reserve Uni-
versity in 1967) wrote to offer a “hearty second” to Harnwell’s proposal, and
recorded feeling “confident that these groups [societies such as the Optical
Society, Acoustical Society, and Association of Physics Teachers] would ben-
efit immensely by giving up a little of their autonomy in order to unify the pro-
fession of physics in America.”21 In the same letters column, Thomas Osgood
added his support and echoed a popular dissatisfaction among applied phys-
icists with the status quo, writing: “To say that the activities of this society do
not now meet the needs of the majority of American physicists is to cast no
slur upon the American Physical Society, but rather to emphasize that physics
has so broadened its scope since the end of the last century that the academic
and theoretical subjects which are the special concern of this society may
legitimately make no immediate appeal to many active physicists.”22 Applied
researchers preferred the prospect of participating in a unified field of physics
to the possibility of increasing their autonomy by forming their own societies
and journals. This preference fueled their frustration with the APS, which
by this time was under the sway of, and therefore catered most strongly to,
theorists—and the experimentalists who tested theoretical claims—most of
whom held academic appointments.
The tenuous consensus within the applied physics community was that
political unity should be maintained and that it should be built on a broad,
inclusive view of physics—although disagreements remained on the issue of
federal authority versus states’ rights. From the federalist perspective, the
American Institute of Physics, rather than the American Physical Society, sat
at the appropriate scale for a flagship society. Applied physicists saw the APS
as a haven for pure researchers—a label that representatives of the APS did
little to disavow. They advocated expanding the powers of the AIP, which
included groups such as the Optical Society and the Acoustical Society and
which promoted more freely the mixing of theory and experiment, basic and
applied research, so that academic interests would not dominate the central
governing body of American physics.23 Unity, in this context, meant preserv-
ing and promoting mechanisms that encouraged generous application of the
title “physicist” so as to increase dialogue both within and across physical
specialties.
The federalist approach to postwar physics represented the growth of the
conviction that physics did need to be understood, at least in part, as “what
HOW PHYSICS BECAME “WHAT PHYSICISTS DO” 41

physicists do,” the consequence being that those physicists doing things not
traditionally recognized by the APS deserved representation nevertheless.
The orthodoxy within the American Physical Society was somewhat differ-
ent. APS officials, council members, and fellows were unenthused by the idea
that their society be unseated in favor of the AIP, or a new entity made in its
image, as the principal organization of American physics. Despite the dissat-
isfaction of applied physicists, the APS still wielded the power necessary to
definitively influence how the field would be structured. The modularity of
the APS’s divisional scheme, which had been instituted in 1931 and allowed
organized interest groups to form a division of the society around a well-
defined topic of physics, gave it an advantage in negotiations about the orga-
nization of American physics. Frederick Seitz’s address at the APS meeting of
January 1945 titled “Whither American Physics?” articulated a position that
expressed the stance of those advocating for the American Physical Society.24
Seitz, like many of his contemporaries, saw physics as naturally split into
two parts: “In the first place it contains a body of knowledge which has in-
trinsic value as a form of culture. This component is commonly called ‘funda-
mental’ or ‘pure’ physics. . . . In the second place, physics serves as a source of
fundamental knowledge for a majority of the important fields of engineering.”
Unlike a considerable number of his colleagues, Seitz exhibited little con-
cern about the growing rift between the two branches, as exemplified by the
academic/industrial split, emphasizing that “the terms ‘fundamental physics’
and ‘applied physics’ are in no sense synonymous with ‘academic physics’
and industrial physics.’” In keeping with the traditional view that emphasized
continuity with Henry Rowland’s vision for pure science, he defended the
role of the APS as an institutional organ for basic research. “The principal
aim of the Society,” Seitz claimed, had been “to publish a journal and arrange
meetings in which fundamental physics was emphasized.” If the society ex-
panded its scope to include an emphasis on applied research, he continued,
then no other organization would remain to protect the interests of explor-
atory research.25
On these grounds, Seitz argued that divisions should proceed within the
APS in such a way, (a) that they did not encourage compartmentalization,
and (b) that they did not lead to too much emphasis being placed on applied
physics. On point (b), Seitz opined: “The danger from this source is partic-
ularly great at present because the vast majority of physicists is concerned
with problems of applied physics. This includes many men who were hither-
to concerned only with pure physics. A large number of these men desire
42 SOLID STATE INSURRECTION

quite naturally to continue this type of work after the war and may, as a result,
feel that the society to which they belong should be adjusted to suit their new
interests.” Although he recognized the same trend toward applied work as
those who favored restructuring physics around the AIP, Seitz preferred to
retain the privileged place of basic research. He suggested that, rather than
expanding the APS or elevating the AIP, applied physicists could find a pro-
fessional outlet in existing AIP member societies and engineering societies,
or start a new association for applied physics to address their needs.26
Seitz argued that fundamental research was the driving force behind all of
physics. Physics did not need to be institutionally unified, according to this
stance, because applied work necessarily relied on advances in foundational
basic research: “The great importance of fundamental physics as a spring for
the well of technology assures us that the development of this field has social
value even if we adopt the most hard-headed attitude towards society.” Basic
researchers, in other words, need not be worried about the safety of their
social importance, because it would be guaranteed by the dependence of ap-
plied research on basic research. Applied physicists and basic physicists had
“no basic quarrel on the issue of whether or not fundamental physics should
be pursued, even though they may feel that their objectives lie apart.”27 Phys-
ics, for Seitz, was conceptually unified—he did not perceive organizational
discomfit as a threat to unity—and the first objective of the American Physical
Society should be to serve and advocate for the physicists who investigated
basic concepts.
Powerful allies such as the well-connected APS secretary Karl K. Darrow
and George W. Stewart, who was president of the APS in 1942, joined in
Seitz’s defense of the APS. Unlike Seitz, these figures coincided with Harn-
well and company in perceiving the increasing rift between the academic and
industrial communities as a threat to the unity of American physics. Their
principal concern, however, was with conceptual purity rather than politi-
cal unity. They sought to prevent industrial interests from diluting the APS’s
avowed commitment to pure science, and organized to prevent the forma-
tion of an APS division devoted to industrial or applied physics. The APS
council, impelled by such concerns, succeeded in blocking several proposals
for an industrial division. Industrial physicists were becoming progressively
more chagrinned by the fact that the society was not geared to their needs.
The council fielded letters to this effect through the early 1940s. One in-
dustrial researcher, W. W. Lozier of the National Carbon Company, griped
that the programs for APS meetings were not published early enough for em-
HOW PHYSICS BECAME “WHAT PHYSICISTS DO” 43

ployees of industrial laboratories engaged in deadline-driven practical work


to request leave and suggested that the society generate notices earlier. On
this count, the council determined, without noting the irony, that changing
existing procedures would be “impractical.”28 Another, A. V. Ritchie, com-
plained that APS publications did not list specific job titles when printing
names of industrial researchers, and thus conveyed no indication of the indi-
vidual’s rank with an organization.29 By 1943, the council was aware enough
of the growing desire among physicists employed in private enterprise to par-
ticipate in the affairs of the society to address, however tepidly, “the topic
loosely described as ‘encouragement of industrial physicists.’”30 Encouraging
industrial physicists, however, did not extend to granting them a division of
their own.
In the September 1943 issue of Journal of Applied Physics, Darrow deliv-
ered a shot across the bow to proponents of an industrial division, reporting
the APS council’s intention to veto any efforts to reshape the Physical Society
in a way that would further isolate academia and industry:
Letters have been received conveying the idea of a “Division of Industrial
Physics.” The Council is disinclined to favor this idea, which incidentally
appears to be precluded by the language of Article IX. One of the things most
greatly to be desired is a unification of the physicists called “academic” and
the physicists called “industrial.” It is important to avert the danger of a lack
or loss of interest by either group in the problems and the enterprises of the
other. Now, the establishment of a “Division of Industrial Physics” would
work in exactly the opposite direction. That way lies the peril of forming two
contrasting groups, whose aims and ambitions ought to be alike but would
infallibly deviate more and more as the years go on.

Darrow did not unilaterally oppose APS divisions and had backed the APS
council’s approval of the Division of Electron and Ion Optics. He encour-
aged proposals for additional divisions, but cautioned that they “be limited
in scope to a particular field or fields of physics.”31 Darrow’s stance amounted
to an official defense of the values on which the society had been founded.
The APS council, through Darrow’s voice, stood by the assumption that
physics, if it were to be subdivided at all, should only be carved along clear
conceptual lines. Article IX, to which Darrow referred, had been added to
the APS constitution in 1931 in tandem with the formation of the American
Institute of Physics. It permitted APS divisions to form around “specified
subject or subjects in physics.”32 Darrow, by suggesting that this language
44 SOLID STATE INSURRECTION

precluded a division dedicated to industrial physics, upheld the assumption


that a “subject” within physics had conceptual roots. Borders, as established
by divisions, would reflect natural categories of knowledge rather than pre-
vailing professional or institutional categories. This attitude, intentionally or
not, favored an academic approach to physics.33 Academic physicists were
more inclined to devote themselves to single, clearly defined research areas
predicated on their interests and the resources available at their institutions,
and to build their professional identities within those research areas. For
them, conceptually defined APS divisions made good sense; such divisions
would likely include the bulk of other physicists with whom they would want
to interact.
Some generational forces were at play in the APS council’s resistance to
an industrial division. The members of the APS council in 1943 included, in
addition to Darrow, president Albert W. Hull, vice-president Arthur J. Demp-
ster, treasurer George Pegram, along with elected members Joseph C. Boyce,
William F. G. Swann, Ferdinand Brickwedde, Alpheus W. Smith, and Henry
DeWolf Smyth. Only Brickwedde and Boyce, both with 1903 birthdays, had
been born in the twentieth century and six of the nine were over fifty. Their
collective scientific work was largely in the early twentieth-century American
tradition of academic-style, instrument-focused research. Having matured
amid the prewar orthodoxy, they would have been naturally suspicious of any
proposed category that deviated from society tradition and so offered a strict
interpretation of the requirement that new divisions be subject-based.34
An arrangement that permitted only narrowly conceived topical divisions
was challenging for industrial physicists. Rather than focusing in depth on a
particular research area, an industrial research physicist would often shift fo-
cus to meet the needs of individual projects. Physicists whose work in acous-
tics was contingent on a project with a fixed timeline, after which they might
turn to thermodynamics or mechanics, had little incentive to join the Acous-
tical Society of America. The same logic would apply to divisions. Topical
divisions, even if they marked out areas of primarily applied interest, would
not have been a strong draw for industrial physicists who were often expected
to adapt their skills to a wide range of topics as exigencies demanded.
Darrow lamented the growing rift between academia and industry, but his
proposed solution, which admitted only subject-based divisions, narrowly
understood, bolstered the traditional academic view in which physics was
unified, and formed its identity, on the basis of the character of the knowledge
it pursued. His concern that industrial physicists were becoming isolated
HOW PHYSICS BECAME “WHAT PHYSICISTS DO” 45

amounted to the suspicion that they might drift too far from the mainstream
of basic research and stop doing work that qualified as physics. Insisting that
divisions maintain topical foci would force industrial physicists to define
themselves conceptually in order to affirm their identity as physicists.

BELLING THE CAT AT THE CONFERENCE OF PHYSICISTS


The fable of the bell and the cat tells of a group of mice plotting to control
a marauding cat. The mice debate strategies back and forth before settling
on a proposal that calls for hanging a bell around the cat’s neck, so that they
might hear it approaching. But when it comes to deciding who will take up
the task of attaching the bell to the cat’s collar, volunteers are scarce and ex-
cuses abundant. Karl K. Darrow, secretary of the American Physical Soci-
ety, invoked this fable to describe the state of American physics in the early
1940s. The community was in broad agreement that change was afoot and
that American physicists should do something to manage that change and
prepare for post–Second World War conditions, but few were willing to com-
promise their own interests for the good of the collective.
The full spectrum of perspectives described above was on display at the
National Research Council (NRC) Conference of Physicists, a “‘town meet-
ing’ of physics,” in the words of American Institute of Physics director Henry
A. Barton.35 The idea of holding such a meeting originated within the AIP
Board of Governors, motivated by an uptick in efforts to form new APS divi-
sions. The AIP board took up the matter of a group petitioning the APS for
a division devoted to applied spectroscopy during its March 1944 meeting,
considering whether the APS was the appropriate venue for such an organi-
zation and discussing what the AIP’s role should be in mediating such efforts
in the future. The discussion moved from the particulars of the proposal to
the general wisdom of supporting interest groups within the AIP member so-
cieties, which led to the even more abstract question of how deliberate orga-
nization of the physics community could be directed to the advantage of the
field. Barton suggested that the National Research Council might have funds
available to support a meeting discussing such questions.36
The hastily assembled conference convened at the American Philosoph-
ical Society in Philadelphia two months later, May 19–21, 1944, “for the
purpose of discussing the present and postwar problems facing physics.”37
The problems singled out for attention included education from high school
through the graduate level, the value and needs of industrial physicists, pro-
fessional standards and the possibility of instituting discipline-wide accred-
46 SOLID STATE INSURRECTION

itation mechanisms, how to manage the relationship between physics and


government during peacetime, and promotion of the value of physics to the
American public. The conference was a premeditated effort to plot the pro-
fessional contours of postwar physics, motivated by an optimistic slant on the
ability physicists had to control their own destiny in the postwar world. The
participants clearly appreciated, before the Second World War had ended,
that the central importance of physics in the war effort presented an oppor-
tunity to increase the field’s peacetime profile as well. Since physicists were
conspiring to structure their national profile, nationalist metaphors again
emerged as the natural way to express their goals.
Richard M. Sutton, secretary of the conference and a Haverford College
physicist widely respected for his teaching, deployed the nation metaphor in
his introductory remarks:
As in an adjacent building here in Independence Square, wise men met to
frame the Constitution of these United States, so we are in a sense met to
frame the Constitution of Physics, perhaps I should say “The United States of
Physics,” on such lines that future generations may acknowledge that we did
our work well. Our purpose is to take some of the preliminary steps with the
expectation that what we discuss here will, by the democratic workings of our
various organizations, take root in the soil of our societies and contribute to
the strength of our whole organism. . . . There is no question here of academic
physicists vs. industrial physicists, or of experimentalists vs. theoretical phys-
icists, and there has been a conscious effort made to have all sides of physics
represented.38

In singling out those specific dichotomies, Sutton reinforced their impor-


tance, which, despite the conference’s egalitarian mission, was on display
throughout. Sutton was nonetheless correct to note the broad represen-
tation at the gathering. Government physicists were in attendance, joining
academic physicists from private and public research universities as well as
small colleges. Representatives from such well-established industrial research
institutions as AT&T’s Bell Laboratories participated alongside physicists
from smaller industrial enterprises that were just beginning to dedicate sub-
stantial resources to research. Fragmentation was a growing fact of life, but
the NRC meeting can be understood as a genuine attempt to bridge the gaps
that divided American physicists. Its diversity affords a cross-section of views
surrounding the future of physics, especially industrial physics, in the 1940s.
The question driving the meeting was an existential one: what does it
HOW PHYSICS BECAME “WHAT PHYSICISTS DO” 47

mean to be a physicist? Oliver E. Buckley of Bell Telephone Laboratories


articulated common concerns that physics would be debased if it strayed too
far from its defining ideals, which he understood in the conventional manner
of the time as the orientation toward the formulation of general principles.
Buckley identified the question of how “physicist” should be defined as key
to postwar professional development and supported developing accredita-
tion mechanisms for both institutions and individuals. He championed in-
dustry as a viable career path for physicists, but also cautioned that although
the growth of physics in industry “is fine for the physicists and should lead
to benefits for them,” it could only do so “if it does not at the same time so
overpopularize the profession as to lead to its degradation.”39
Protecting prestige ranked alongside revamping education as a postwar
goal for the majority of physicists, both academic and industrial. Buckley
continued: “Many who are not physicists will see nourishment for themselves
in adopting the title. Such a result would obviously be unfortunate, for it is
only by restricting the use of the name physicist to those who ennoble it that
these benefits can be made enduring.”40 Protecting physics as a prestigious
appellation meant actively demarcating its responsibilities from those of other
scientists and technicians, most notably chemists and engineers. Demarca-
tion was especially important in industry, where physicists were usually not
corralled within departmental structures that respected disciplinary integrity
and needed to define how their responsibilities and their status differed from
their collaborators within large research groups.
Seconding Buckley’s advocacy of accreditation, Cornell’s Roswell C.
Gibbs judged it “a mistake in making appointments to designate a physicist
as an engineer or as a chemist even though it may be the easy way to secure
approval of a new appointment by an Executive Board or Officer or to elicit
favorable consideration from a Draft Board for an ‘essential man’ in the war ef-
fort.” Being accredited as a physicist, and properly referred to as such would,
in Gibbs’s view produce “the effect upon the individual in maintaining his
morale, in giving him a proper feeling of prestige, in developing a sense of
belonging to a group with interests and points of view in common with [his]
own, in promoting loyalty to the profession he has chosen, and in encourag-
ing him to associate himself with suitable scientific organizations from which
he will derive stimulation and other benefits and to the development of which
he can direct his own efforts and support.”41
Gibbs’s anxiety was related to Buckley’s, but it was distinct in one mean-
ingful respect: Buckley aimed to keep interlopers out of physics; Gibbs strove
48 SOLID STATE INSURRECTION

to keep credentialed physicists in. Whereas Buckley expressed concern that


those who were not physicists would usurp the name, Gibbs sought to pre-
empt the temptation some individuals might feel to adopt another name for
the sake of expedience. Buckley worried that the status and prestige of phys-
ics would be diluted if the name were not controlled; Gibbs predicted that
prestigious accomplishments might bolster the status of competing fields if
physicists were not properly ensconced in a robust disciplinary structure.
The concerns were not mutually exclusive, and both were relevant when con-
sidering the growth of industrial physics. The parallel impulse to police the
boundary of physics from within and without also reflects a larger ambiva-
lence about industry, which appeared in mid-1940s discussions about the
future of physics as both a potential area for useful growth and as a threat to
physicists’ traditional values.
Harvard’s Edwin Kemble, one of the pioneers of American quantum the-
ory, described both the opportunities and the dangers industry represented
by setting out the challenges graduate education faced. He observed what
was by then a widely accepted fact: “Greater emphasis on the industrial ap-
plications of physics will be necessary and desirable after the war is ended.
Industry will need more physicists, and physicists will need industrial jobs.”
Turning to the challenge of preparing graduate students for industrial posts
while remaining true to traditional ideals of physics, Kemble continued: “We
shall need to give more electronics, chemistry, metallurgy, and shop work. At
the same time I, for one, should be very sorry to see anything like the con-
version of graduate training into a glorified engineering course.”42 Kemble
voiced a broader ambivalence with respect to the proper place of industrial
physics. It was a growth area in the 1940s, offering abundant employment for
physics PhDs, and its technological focus helped cement the social relevance
of physics. Yet industrial work often did little to bolster the traditional ideals
of physics. Kemble’s pejorative reference to engineering expressed the feeling
many academic physicists shared that knowledge how remained less noble
than knowledge why.
Kemble’s stance illustrates an imbalance in the relationship between aca-
demia and industry. Although the industrial sector could boast faster growth
and a larger population, all physicists, save the rare autodidact, were trained
in an academic context. The path to protecting industrial interests therefore
cut through academic territory. Building industrial concerns into the future of
American physics meant convincing graduate advisers that students should
be exposed to industrial problems, skills, and job opportunities. R. Bowling
HOW PHYSICS BECAME “WHAT PHYSICISTS DO” 49

Barnes of the American Cyanamid Company noted that the place of the phys-
icist in industry could only be realized more fully through attention to “the
type of training that our future physicists are given.” Barnes foresaw “golden
opportunities ahead for industrial physicists,” but cautioned that “to take ad-
vantage of these opportunities . . . and to make the best use of his knowledge
of physics it will be imperative that he be able to speak and understand the
language of his fellow scientists.”43 To this end, he supported broad training
for physics students in the rudiments of general fields such as biology, chem-
istry, and geology, and also exposure to specific technological growth areas
like rubber and petroleum.
Barnes’s remarks stirred up considerable controversy. Several discus-
sants, including Mervin Kelly of Bell Laboratories and other industrialists,
argued that strong training in foundational skills and concepts outweighed
broad exposure to other fields, even among applied physicists. The reaction
against Barnes’s advocacy of what Kemble might have called “glorified engi-
neering” indicates that although industrial physicists were keen to see their
interests reflected in graduate training, they were hesitant to do so at the risk
of ghettoizing themselves. Protecting industrial physicists, for most, did not
mean tracking doctoral students into basic or applied subprograms; rather
it meant, as Ralph A. Sawyer suggested, expanding the scope of necessary
foundational training in physics to include fields such as geometrical optics
and hydrodynamics.44
Effectively managing the impending growth of industrial physics was not
just a matter graduate training, but also required organizational encourage-
ment. G. P. Harnwell, having further developed his views on how postwar
physics should be organized, observed in his address that “the needs that
should be supplied by and for physicists have simply outgrown the exist-
ing organizational framework.” In fact, according to Harnwell: “There can
scarcely be said to be at present any organization of physicists. The Ameri-
can Physical Society is not sufficient[ly] broadly based and representative; it
is properly an exclusive rather than an inclusive society.” Harnwell repeated
what had become a mantra, stating that “physics is a unified discipline deal-
ing with matter and energy in all their forms and interactions,” continuing:
“But this corpus of concern is so broad that the internal structure of the
society must be braced with the beams and cross-ties of special interests.”
Harnwell advocated a “horizontal and vertical divisional structure,” in which
physicists could be represented both by institutional context and topical
interest, which he felt would allow society to better organize meetings and
50 SOLID STATE INSURRECTION

distribute publications such that they would reach the greatest number of
interested readers. Similar discontent with the narrow goals of the APS was
widespread and Harnwell’s idea of an inclusive umbrella organization was
one prominent solution under consideration.45
The notion that physics should be reimagined under a new, broader
framework was not universally beloved, however. Although Harnwell’s talk
won support from industrial physicists who felt that an expansively conceived
society would better fit their needs, it also met pointed criticism. Harvey
Fletcher, then of Bell Labs, opened the discussion with the complaint, “Dr.
Harnwell’s thesis would seem to indicate that we have done wrong in form-
ing the Chemical Society, the Astronomical Society, the Physical Society, and
others, as we have grown from the original Philosophical Society into these
branches.”46 Fletcher would later serve as president of the APS, but he spoke
at the NRC with his feet firmly planted in the Acoustical Society, and sug-
gested from this standpoint that fragmentation was a natural and unavoidable
by-product of growth.
Mervin Kelly, also of Bell, suggested that Harnwell was indulging utopian
fantasies and that it was more pragmatic to work with the APS as it existed
rather than attempting to craft a new, suitably complete, large-scale edifice.
Karl Darrow, the APS secretary, seconded this view, suggesting that the
scheme for divisions that APS was just beginning to implement be given a
chance to work before the physics community considered subjecting itself
to sweeping changes.47 Together, these voices favored allowing the organic
processes of institution building to work before considering unilateral, top-
down action.
Following the airing of a range of proposals favoring large structural
changes, which were suffused with the same optimism about the capacity of
physicists to shape their fate that motivated the congress, the meeting ended
on a conservative note. Karl Darrow, commenting on the dubious likelihood
that many of the proposed actions could be implemented in an orderly fash-
ion, remarked: “I am reminded of the old story of the mice who decided to
bell the cat. It seems that in this case the American Institute of Physics has
been invited to bell the cat.”48 Darrow’s quietist conclusion about restructur-
ing was perhaps somewhat disingenuous, in that it conveniently aligned with
his conviction that the APS was adequately equipped to handle the pressures
of postwar demographic changes and that its more limited understanding of
physics should be protected. But despite the failure of the likes of Darrow to
come around to the view that the organizations of American physics required
HOW PHYSICS BECAME “WHAT PHYSICISTS DO” 51

substantial restructuring, the NRC conference generated momentum for


those seeking to implement such reforms, and the following few years would
indeed visit significant change upon both the AIP and the APS.

REORGANIZING PHYSICS
Following on the heels of the NRC conference, the AIP formed a Policy
Committee on the Reorganization of Physics to distill the range of opinions
hashed out at the meeting into a set of recommendations. The committee,
chaired by John Tate, included many who were sympathetic to the federalist
view, including Harnwell, Gibbs, and Buckley.49 The premise from which the
committee began was that changes in the nature and composition of Ameri-
can physics since the formation of the AIP necessitated reevaluating its core
responsibilities. The foremost change they identified was the fact that “inter-
est in the science of physics is now much more widespread than formerly. It is
no longer so much concentrated in academic circles and extends into a host
of industries and into the border ground of other sciences. The number of
academic, institutional, and industrial workers who identify themselves with
physics has approximately doubled in the past decade, and the post-war era
promises a much greater expansion.”50 The committee acknowledged that it
drew the boundaries of physics broadly enough that it would likely embrace
many people otherwise classified as chemists, engineers, and metallurgists,
but nevertheless insisted that recognizing the interest such researchers main-
tained in physics was necessary for the health of the community.
The proposal for reorganization became the template for a new AIP con-
stitution, adopted in February 1946. It reflected the federalist sympathies of
the committee and fell in line with an understanding that physics needed to
be organized to reflect what physicists did. The AIP would continue its pub-
lishing responsibilities, and would add a new category of individual mem-
bership, which would be extended automatically to individual members of
AIP member societies, but which could also be acquired by joining the AIP
directly as an associate member. The AIP dropped this membership category
shortly thereafter, but its inclusion in the initial version of the new constitu-
tion speaks to the evolved role the organization envisioned for itself. Also
rolled into the AIP’s new mission was the publication of a true general inter-
est journal, which would provide a venue for any physicist to publish work
that was interesting and accessible to a wide array of colleagues, and which
would in some measure counteract the topical specialization of physics. That
proposal would lead to the establishment of Physics Today in 1948.
52 SOLID STATE INSURRECTION

The new AIP constitution provided the apparatus necessary to permit


something akin to Harnwell’s vision to come to pass. The AIP, however,
never grew into a society to which physicists felt any allegiance. The APS,
which had the most to lose if the AIP succeeded in expanding its mission,
finally made concessions to the groups of industrial and applied researchers
who had been agitating for greater representation. By mid-1945, two divi-
sions had managed to navigate through the APS council and receive official
recognition, the Division of Electron and Ion Optics and the Division of High
Polymer Physics. But as more proposals reached the council, requesting
representation for industrial physics, physics and aviation, chemical phys-
ics, spectroscopy, and other specialties, the council balked. The very name
“division” embodied everything figures like Darrow feared in the specter of
specialization. At its May 5 meeting, the APS council placed a moratorium on
the formation of new divisions, pending the resolution of outstanding reorga-
nization efforts, including those at the AIP.
The vast changes in the status and purpose of physics that marked the
early post–Second World War period are often cast as brought about by dis-
tinctive Cold War pressures and incentives. But the institutional machina-
tions that laid the groundwork for those changes were under way before the
Second World War came to a close. The most important prerequisites that
led to the reconfiguration of American physics were negotiated through the
early 1940s and had imposed themselves on the infrastructure of American
physics before the most salient features of the Cold War context—the fed-
eral funding environment, virulent anticommunism, and the cloud of nuclear
fear, in particular—came into focus.51 These changes could only come about
because some physicists came to think of their field as the sum of what physi-
cists did. As William A. Wildhack, a National Bureau of Standards physicist,
wrote to support the idea of a general, AIP-like organization to replace the
APS: “The policies of the new AIP should be framed on the realization that
its aims can be achieved only by widespread and active membership support,
and that this support can be retained only by an organization which is as
much concerned with service to scientists as with service to science.”52 The
locus of professional power did not shift fully to the AIP, as some hoped, but
the new prevalence of this way of thinking did spur institutional change.
Late in 1946, the APS council again turned its attention to the question
of divisions, this time with knowledge of the AIP’s plans. On September 19,
the council approved an official policy regarding divisions, composed of five
points:
HOW PHYSICS BECAME “WHAT PHYSICISTS DO” 53

1. The Society shall continue the policy of establishing and supporting Divi-
sions as provided for in the constitution.
2. All members of Divisions shall be Members or Fellows of the Society.
3. In the Division now enrolling Associate Members, these shall be permitted
to continue in their present status until the reorganization of the American
Institute of Physics makes provision for individual members.
4. Any member of the Society may enroll in a Division on payment to the
Society of an initiation fee of two dollars.
5. Divisional expenses considered normal by the Council shall be met by the
Society, and no divisional dues may be collected or assessed by a Division
unless authorized by the Council.53

The policy, especially points (2) and (3), was designed to prevent divisions
from bleeding outside the bounds of the APS and thereby to reinforce the
society’s control over the scope and character of the discipline. Divisions al-
lowed groups with interests that were not traditionally represented by the
APS to build space for themselves within it, as long as they submitted to the
authority of the larger society. By allowing an expanded program of division
formation, albeit grudgingly, the APS shored up what authority it retained to
shape the definition of physics, forestalling somewhat Harnwell’s federalist
vision. That is not to say that efforts to expand the scope of physics failed.
The end of the moratorium on division formation opened the door to a new
division that would rapidly become the society’s largest and most influential,
while embodying many of the principles Harnwell defended: the Division of
Solid State Physics.
3

BALKANIZING PHYSICS

When we sit on the lawn of the Bureau of Standards, we do not want


to feel “qu’un sang impur abreuve nos sillons.”
—JOHN H. VAN VLECK, 1944

Like the transistor or the microwave oven, solid state physics was itself an
industrial innovation. Physical investigation of the properties of solid matter
could boast a long tradition by the 1940s, but solid state physics only became
a distinct professional entity in the United States upon the founding of the
American Physical Society’s Division of Solid State Physics (DSSP) in 1947.
The DSSP emerged from the institutional machinations explored in the pre-
vious chapter, in which industrial physicists struggled for a greater role in the
community while traditionalists defended a more restrictive and conceptually
purer vision of physics. The DSSP resulted from an effort spearheaded by in-
dustrial physicists, which aimed to negotiate between these competing views
of how the post–Second World War physics community should be unified.
The qualitative and quantitative growth of American physics disrupted
traditional modes of institutional governance and notions of professional
identity. Topical divisions of the American Physical Society (APS) emerged
as the preferred salve for destabilizing expansion. Divisions gave smaller in-
terest groups an institutional outlet while also keeping them under the aegis of
the APS, which defined and enforced professional norms more strictly than a
looser alliance could. The DSSP, as well as responding to an uptick in interest
in the physics of solids, can be understood as an attempt to enact an inclusive,
outward-looking identity for physics, contrasting traditional notions of what
BALKANIZING PHYSICS 55

a physicist was, within the institutional constraints and opportunities of the


mid-1940s.
Spencer Weart describes the process through which solid state physics
formed, noting: “When we speak of the emergence of solid-state physics . . .
we do not mean the creation of something de novo. We mean a grand rear-
rangement of an entire array of specialties, old and new, into a novel constella-
tion.” Weart identifies the idiosyncrasies that characterized the development
of solid state physics, a field that broke the rules of discipline formation fol-
lowed by its sibling subfields, such as nuclear physics. Rather than growing
from a seed, Weart observes, specialties sometimes “rearrange themselves
into fields at a single time when conditions reach some critical point.”1
The lesson Weart draws from the formation of solid state physics is that
“a few thoughtful people were able to sway the balance in favor of a com-
munity that was intellectually and socially open, yet internally coherent—a
solid-state physics community.”2 Examining the physics community in the
1930s, he finds “no clear tendency to unify around the general subject of
solids,” but that, even so, “there was indeed a unifying force. . . . This force
was the growing intellectual unity of the subject. Solid-state physics could
become an intellectual community only after its cognitive parts had drawn
together in the minds of some physicists.” He suggests that a “‘unified theory’
had been created through the coming of quantum mechanics,” and identifies
the “intellectual unity offered by the concept of a solid,” as a force working in
favor of solid state’s cohesion.3
This chapter revisits the story of solid state physics’ founding with the
goal of suggesting that the unity the solid state synthesis supported, although
powerful and effective at managing the rising tide of specialization, was some-
what less than conceptually coherent. The quantum mechanical approach
to crystalline solids that had emerged by the 1940s was indeed unified and
coherent, but it was not the whole of solid state physics. The distinctive in-
stitutional rearrangement that produced the larger solid state community, in-
cluding at the outset a number of research traditions beside the one focused
on the quantum mechanics of regular solids, reflected ongoing debates about
the identity of American physics. In that sense, solid state was a political alli-
ance rather than a conceptually unified community. It was internally hetero-
geneous and quickly became fractious. Although the possibilities opened up
by the application of quantum techniques to solids were a critical antecedent
for the rise of solid state physics, another equally important precursor was
an ideological shift that made it possible for American physicists to conceive
56 SOLID STATE INSURRECTION

of a subfield defined by professional and institutional objectives. Solid state


physics, although its founders strove for unity, was not defined by the same
type of unity beloved of other areas of physics.
Some of the research programs that made up solid state physics were
indeed conceptually coherent, but focusing on the unity within these areas,
which often were part of traditions dating to before the quantum revolution,
neglects the role of political unity in providing what little cohesion the field
as whole possessed. Analyzing solid state as a fractious constellation, bound
into a loose alliance in the service of common professional goals, clarifies
the reasons for its formation, the characteristics of its evolution through the
Cold War decades, and ultimately, as discussed in later chapters, its role in
the eventual formation of new categories, such as condensed matter physics
and materials science. The story of the DSSP’s origins, therefore, tells not
just how one of the most important provinces of Cold War physics estab-
lished itself, but how a new sense of identity for American physicists, one that
embraced applications, interdisciplinary exchange, and industrial relevance,
became possible.

MAKING DIVISIONS
The founding of the Division of Solid State Physics was a circuitous affair,
lasting almost four turbulent years straddling the end of the Second World
War. The idea for an APS division that would cater to an inclusive cross sec-
tion of industrial and academic researchers interested in some aspect or an-
other of the physics of solids began to germinate in November 1943 in Evans-
ton, Illinois. Northwestern University hosted an APS meeting that included
a symposium on the physics of rubber. One purpose of the symposium was
to discuss a petition that had been circulated earlier in the year, garnering
thirty-one signatures, in support of a division representing the physics of
high-polymeric materials. The industrial lineage of this division is clear from
its originally prosed title: Division of Textile Physics. The more highbrow
reference to high polymers was adopted to mollify the APS council and its
prejudice for subject-based divisions. On the power of the Evanston petition
and the verbal support expressed at the meeting, the APS council authorized
the Division of High-Polymer Physics, the society’s second.4
Roman Smoluchowski (figure 3.1), a General Electric (GE) research
physicist, was in attendance. Smoluchowski had come to the United States
from Poland, fleeing German occupation, a few years earlier. In Warsaw he
had led the department of the physics of metals, and the rubber symposium
BALKANIZING PHYSICS 57

Figure 3.1. Roman Smoluchowski in the General Electric Research Laboratory, ca.
1944. Credit: AIP Emilio Segrè Visual Archives, courtesy of Roman Smoluchowski

emboldened him to launch a similar effort on behalf of his preferred class of


materials. Before the year was out, Smoluchowski had recruited five other
physicists with an interest in metals research to advocate for an APS divi-
sion of metals physics. This “group of six,” so named by Karl K. Darrow, was
strong on industrial researchers. It included Smoluchowski’s GE colleague
Saul Dushman, along with Sidney Siegel (Westinghouse) and William Shock-
ley (Bell Labs). It included one representative each from the academic and
industrial sectors. Frederick Seitz of the Carnegie Institute of Technology,
who represented the academy, had previous industrial experience, having
worked at General Electric from 1937 to 1939. Thomas A. Read of Frankford
Arsenal, representing the public sector, had been a Westinghouse Research
Fellow from 1939 to 1941. The group of six signed their names to a letter
58 SOLID STATE INSURRECTION

Smoluchowski wrote entitled “The Present War Is a Physicist’s War,” which


asked “metal physicists” to weigh in on two questions: “First, do you think
that some sort of a cooperation among metal-physicists is necessary and ad-
visable or not; and second, what form this cooperation should take: a small
committee, a section of The Physical Society, or something else maybe?”5
The group of six distributed the letter to fifty-three of their colleagues
nationwide who had a research interest in metals, which, in Smoluchowski’s
judgment, promised rapid postwar growth “not only from the point of view
of fundamental science, but also from the point of view of practical problems
in industry.”6 He chose “metals” both because industrial researchers worked
primarily with metallic substances and because it implied a willingness to
collaborate with metallurgists, much as he did in his day-to-day work at Gen-
eral Electric.7 Collaboration outside of physics was central to Smoluchow-
ski’s vision and he resisted other names on the grounds that “cooperation
from purely metallurgical quarters may be more active if we are ‘all out for
metals.’”8
Metal physics proved more fractious than polymer physics and the letter
elicited mixed responses. The most pointed opposition came from John H.
Van Vleck, by then ensconced at Harvard University and heading the theory
group of the Radio Research Laboratory, which pursued wartime research
into radar countermeasures. Van Vleck, in his reply, wrote, “I must confess
that I am a rather violent opponent of what I like to call ‘Balkanization’ of the
American Physical Society, be it either geographical or by subject matter.”
Van Vleck praised “the spirit of a Quaker meeting” that characterized prewar
meetings, at which “one can go to hear the papers, in case one is interested, or
sit on the lawn and talk to one’s friends, in case one is not,” and where “any-
body who wants to can give a ten-minute paper on any subject having even
the remotest connection with physics.”9 An essay Darrow wrote for Physics
Today in 1951 also references the tradition of fleeing to the outdoors during
APS meetings. By way of giving physicists advice on how to deliver a paper,
Darrow suggested that familiar dry monotone of most APS talks was in part
responsible for chasing attendees to the lawn, noting “the popularity of the
saying that when a meeting of the American Physical Society is going on, the
members are in the corridors or on the lawn instead of listening to the speak-
ers” (figure 3.2). Physicists could learn something from stage actors: “People
with tickets to South Pacific are not standing around in Forty-fourth street
when the curtain is up.”10
BALKANIZING PHYSICS 59

Figure 3.2. Karl Darrow (left) chats with Henry Barton, American Institute of Physics
director from 1931 to 1957, outdoors during an American Physical Society meeting
in Washington, DC, 1960. Credit: Emilio Segrè Visual Archives, American Institute of
Physics, Physics Today Collection

Van Vleck further criticized plans for a new division on the grounds that
it would pollute APS meetings with nonphysicists: “The idea that various
groups whose main interest is not physics must be coddled, in order to make
them members of the American Physical Society, has never appealed to me,
as just mere numbers is not everything. The American Chemical Society
is, to my mind, a prime example of this point. It seems to me that the con-
tinual tendency to section the Physical Society and establish a lot of sub-
organizations will tend to put it on what I may term a ritualistic and/or
new-deal-bureaucratic basis.”11 Van Vleck favored an informal environment,
unencumbered by substructures. He worried further that divisions, if they
maintained their own membership, would harbor interlopers who had spe-
cial interest in the subject the division represented—metals, for instance—but
had no inclination to consider questions that were characteristic of physics
as he understood it, as the search for generalizable laws. Opposition of this
character was consistent with the view of the APS as a pure research organi-
60 SOLID STATE INSURRECTION

zation. Van Vleck, through his foundational work on the quantum theory of
magnetism, had contributed much to the physics of metals but was interested
in them only insofar as they provided test systems for foundational princi-
ples.12 By “groups whose main interest is not physics,” Van Vleck meant not
just chemists, metallurgists, and engineers, but applied researchers whose in-
terest in applications cast doubt on their commitment to pure physics.
Van Vleck’s opposition to a division represented a minority of the metals
physicists polled, but that minority exerted outsized influence. Among the
founders of the American theoretical physics community, Van Vleck held an
exalted place. He was one of the few Americans to participate actively in the
quantum revolution of the 1920s, and this inspired veneration sufficient to
lend his opinions considerable weight.13 He was a good friend of APS secre-
tary Karl Darrow, who was responsible for overseeing the formation of new
divisions.14 No decisive evidence shows Van Vleck exerting direct influence
on the division-formation process through Darrow, but their frequent corre-
spondence gave Van Vleck repeated opportunities to reiterate his distaste for
the carving up of the APS into interest groups, and his expression of such
opposition predated Smoluchowski’s proposal. Upon receiving a survey
regarding the Division of Electron and Ion Optics, for example, Van Vleck
wrote to Darrow:
In reply to your questionaire [sic], I do not wish to be enrolled in the electron
microscope division of the Physical Society. Apparently we are now to have
vertical as well as horizontal Balkanization of the American Physical Society.
I enclose herewith a copy of the Constitution that I ran across in my files.
Apparently it is the original form written by our wise forefathers at Indepen-
dence Hall, before any new deal amendments. It seems to carry no provisions
for either type of Balkanization, or entering into the real estate business. Are
you sure that all these items are legal?15

Article IX, which added a provision for topical divisions in 1931, would
not have appeared in an original version of the APS constitution. Given
Van Vleck’s standing, his persistence, and their close relationship, the way
Darrow dealt with proposed divisions likely owed something to Van Vleck’s
objections.
Smoluchowski took pains commensurate with Van Vleck’s stature to as-
suage his concerns. He aimed to skirt both of Van Vleck’s objections while
still acknowledging their validity and demonstrating deference: “Everybody,
I think, will agree that ‘Balkanization’ of the American Physical Society would
BALKANIZING PHYSICS 61

be very undesirable,” he wrote in response to Van Vleck’s skepticism about


the group of six’s proposal: “What some of us consider worthwhile is (if I
may use your political analogy) an attempt to prevent ‘an invasion and par-
tition by powerful neighbors’ (i.e. chemists and metallurgists) at the same
time avoiding the dangers of specialization. I do hope that if such a group is
created it will not interfere with the democratic spirit of the meetings of the
Society which all of us enjoy so much.”16 Van Vleck, unappeased, extended
the political analogy, describing Smoluchowski’s stance as “a little like the
idea of Germany before 1914 that if it did not expand, it would be gobbled up
by Russia and France. As far as I can see, the Physical Society is a going con-
cern, and its membership list is not shrinking.” His hesitancy about opening
the door of the APS too widely rested on a concern “that our meetings will
have too many hangers-on whose main interest is not that of pure physics or
science. When we sit on the lawn of the Bureau of Standards, we do not want
to feel ‘qu’un sang impur abreuve nos sillons.’”17
Van Vleck emphasized what Smoluchowski had missed in his first reply:
his concern was not that chemists and metallurgists would infiltrate the APS,
but that expanding the scope of the APS beyond its traditional focus on pure
physics would rob it of its intellectual integrity and social intimacy. Van Vleck,
a physicist before he was a scientist or a theorist or a metals researcher, cast
his lot with those who felt similar allegiance to physics as a cohesive intellec-
tual enterprise. Smoluchowski, resigned to Van Vleck’s opposition, replied
only that he did not anticipate that the type of organization he had in mind,
directed mostly toward organizing special symposia at APS meetings, would
undermine the Quaker spirit of those meetings.18
Van Vleck and other opponents of divisions made more successful en-
treaties to Frederick Seitz, who was more favorably disposed to view the APS
as a haven for pure researchers. In early May 1944, while the group of six
was in the process of organizing a petition to the APS council for a metals
division, Seitz wrote to Darrow suggesting that the question of new division
be postponed until after the war. He explained his suggestion by referencing
conversations he had conducted with opponents of the division: “These are
men whose opinions I respect very much, and I was impressed with the fact
that they feel as if divisionalization will be a catastrophe.”19 Darrow was puz-
zled by Seitz’s effort to undermine the group of six’s petition, replying, “I am
not sure whether this is a definite withdrawal of the request by your group for
a Division of Metal Physics, but since the Council has no formal petition for
such a division before it, it has not yet taken action about such a division nor
62 SOLID STATE INSURRECTION

is it likely to.”20 Darrow expressed his conviction that further subdivisions


of the APS were inevitable, and that the process was essential as the society
grew and required additional structures to handle practical matters such as
organizing meetings and symposia.
Seitz, an astute politician with a finger ever in the wind, appeared relieved
that Darrow had drawn him back from the skeptical brink, replying:
You have put the entire situation in a light that makes me feel much clearer
about the objectives at hand. As you undoubtedly gathered from my previous
letter, I have been subject to considerable bombardment from many hands
concerning the part I had played in connection with the proposal to start a
division of “Metal Physics.” I now realize that divisionalization is not an evil
in itself, and that the success or failure of the American Physical Society in
the future depends entirely upon the way in which the members exercise the
rights that they have under the existing organization.21

The proposal for a new division would proceed, despite the opposition.

FROM METALS PHYSICS TO SOLID STATE PHYSICS


Darrow recognized the growth of divisions as inevitable, but in deference to
Van Vleck and other traditionalists held fast to the interpretation of the APS
constitution that required divisions to form around subject areas—the source
of his resistance to proposals for a division of industrial physics. He sought to
bridge the ideological divide by recommending that the group of six pursue
a solid state physics division, rather than a metals physics division, which
mitigated Van Vleck’s objections somewhat while still giving Smoluchowski’s
energetic group the institutional space they craved.22
Seitz was swayed by Darrow’s suggestion that the proposal for a metals di-
vision be broadened to encompass solids generally. He offered a mea culpa to
Smoluchowski for wavering in his conviction while being “bombarded very
heavily with views against divisionalization” before asking: “Do you have a
deep-seated objection to using the name ‘Division on the Physics of Solids?’
This would undoubtedly open the doors to a much wider group, such as
those interested in pigments and ceramics, and would make the aspects of
divisionalization a little less grim.”23
The discussion over whether to organize around metals or solids cut to
the core of the disagreements over the proposed division’s mission. Léon
Brillouin, an expert on the quantum theory of crystals who had fled France
for the United States at the beginning of the Second World War, expressed the
BALKANIZING PHYSICS 63

heart of the opposition to metals as a subject category in his reply to the group
of six’s letter: “The distinction between metals and other solids has no sci-
entific basis and is only a matter of engineering—The physicist cannot draw
a border between the two cases, and most of our recent knowledge of metals
was the result of experiments and theories worked out on the non metallic
crystals. So let us speak of Physics of solids in general.”24 From the perspective
of basic physics, the solid state was the superior category because it did a
better—if imperfect—job of preserving topical coherence within the division.
In contrast, the choice of metals as an organizing principle betrayed too starkly
the industrial origins of the original proposal. Smoluchowski, responding to
Sidney Siegel’s suggestion that the entire solid state be considered, reported
hearing “the opinion that most industrial research is done in the domain of
metals,” although he admitted that he could not be sure himself.25
It took considerable persuasion before Smoluchowski would accept sol-
ids over metals as the focus of the new division. Despite his deference to Van
Vleck’s concerns and the pains he took to emphasize that the division was
motivated by the best interest of physics and physics alone, one of his driving
motivations was collaboration with chemists and metallurgists, which he took
to be essential to the health of physics. He envisioned an organization that
would keep the study of metals tied closely to physics, but also saw it as a tool
to organize joint symposia with societies such as the American Society for
Metals (ASM) and the American Institute of Mechanical Engineers (AIME).
Smoluchowski did have good reason to emphasize the collaborative po-
tential with these societies. Metallurgy before the Second World War had as-
pirations to improve its standing among the sciences. When the University of
Pennsylvania was working to expand its research profile in the late 1930s, for
instance, metallurgy was one area it identified as offering a substantial return
on investment. Gaylord P. Harnwell, recently appointed chair of the physics
department, attempted to poach metallurgist Gerhard Derge from Carnegie
Tech, noting that “a number of alumni and interested friends of the Univer-
sity have proposed that the interest in Metallurgy be expanded,” and em-
phasizing “a complete unanimity of opinion among us that the graduate and
research aspect of the development should be conceived of as a pure scientific
program in the Physics and Chemistry of solids rather than endeavoring to tie
it too closely to classical empirical metallurgy.”26
Further, metallurgy programs and industries focused around research
into the properties of metals were hiring physicists. In the National Research
Council’s 1946 survey of US industrial laboratories, only one, Bell Tele-
64 SOLID STATE INSURRECTION

phone Laboratories, listed “solid state” as a research area.27 In the same vol-
ume, references to metals, alloys, steel, metallurgy, and similar terms appear
on almost every page and in the entries for the great majority of laboratories
that list physicists among their research staff. Smoluchowski’s sense that met-
als were central to the work of industrial physicists was not misplaced. But
similar data indicate that he need not have been so concerned about the exact
name of the field. Only a handful of laboratories in the 1950 edition of the
same report list solid state physics among their research areas or solid state
physicists among their staff, but by 1956 the term was in common circulation
and in 1960 it was ubiquitous.28
Smoluchowski eventually relented. He wrote Seitz at the end of May
1944: “I quite agree that from the point of view of the A.P.S. as a whole ‘sol-
ids’ are to be preferred, but I think that from the point of view of cooperation
with other societies, ‘Metals’ are more appropriate. This cooperation and this
bridging of the gap between metal physicists and metallurgists is to my mind
one of our main objectives.”29 Seitz replied on June 14:
Regarding the title of the division, I honestly do not believe that we should
worry about what the metallurgists would consider a good title. From the
experience I had in Pennsylvania with organizing cooperative meetings, I
believe I can state honestly that the cooperation with the metallurgists will
have to originate on our side. If there are joint meetings with the ASM or the
[AIME], it will be because someone like you or I has been aggressive enough
to approach them. The exact name of the division will not play a role in this
negotiation. As a result, I still hardly favor the use of the word “solid” instead
of the word “metal.” In addition, I think we should remember that the persons
interested in pigments, glass, and the like, who will be interested in the divi-
sion, are very large in number, and are no less organized than the metallurgists.
I believe it would be a mistake if we expressed interest in one of these groups,
to the exclusion of the other.30

Seitz had been an assistant professor of physics at the University of Pennsyl-


vania in the late 1930s during Harnwell’s efforts to expand the institution’s
profile in metallurgy. The work Seitz was pursuing on solids had been a crit-
ical element of Harnwell’s sales pitch to potential hires, so Seitz would have
had ample opportunity to develop a considered position on the relationship
between physics and metallurgy. Seitz’s influence paved the way for Smo-
luchowski to finally abandon his commitment to metals. When Smoluchowski
met with Darrow two days later on June 16, the result was an agreement to
BALKANIZING PHYSICS 65

circulate a petition for the formation of a representational organ for physics


of the solid state and to arrange a symposium to discuss the state of the field
and plan future organizational efforts.31
The shift from metals to the solid state did not dampen Smoluchowski’s
enthusiasm for ensuring that the growth of new fields of inquiry, especially
those with industrial relevance, remained linked to the APS. The name
change quelled some opposition from those who insisted on topical divisions
or none at all, but the proposed DSSP, as demonstrated by subsequent events,
remained an effort to integrate industrial researchers with the academic
community by seeking a closer association between self-identified pure and
applied researchers.

FOUNDING THE DIVISION OF SOLID STATE PHYSICS


The question of what sort of organization should be established had remained
in the background up to this point. It became central to the discussion after
Smoluchowski and Darrow agreed to move ahead with the proposal for a
solid state division. The group of six’s original letter had asked whether a full-
fledged division or a smaller standing committee would be more favorable.
In June 1944, the APS council appointed the group of six as a committee
to organize a symposium at which the question would be put to a group of
interested physicists.32 To this end, the group of six published the results
of their survey in the Journal of Applied Physics with a one-page statement
entitled “Physics of the Solid State.” This published statement gave a softer
pitch than “The Present War Is a Physicist’s War,” but advanced essentially
the same argument and raised essentially the same questions.33
Rather than advocating for new organizations, the statement plainly pre-
sented the range of options available and advertised the symposium, to be
held at the January 1945 meeting of the APS in New York, at which they
would be hashed out. Alongside the possibility of a new division, the article
outlined the option of forming a standing committee within the APS, which,
although it would require modification of the APS constitution to allow for
such entities, would not maintain membership rolls and would thus make
less of an impact on the identity and allegiance of individual physicists. The
statement recognized among the community of solids researchers “a desire
to avoid, as much as practical, too formal an organization,” and noted that
avoiding divisionalization would skirt “the dangers of overspecialization and
[stress] the unity of interest and of purpose of all physicists.”34 Such dainty
rhetoric sought to defang the likes of Van Vleck, while recognizing the fact
66 SOLID STATE INSURRECTION

that solid state was such a large category and that institutionally reifying it
would group together physicists who otherwise had little in common.
The group of six spent the fall and winter of 1944 organizing a symposium
for the January APS meeting. They took pains to balance institutional affili-
ations of the participants, their stated positions on divisions, as well as theo-
retical, experimental, and applied research. Van Vleck, for example, agreed to
speak on the theory of ferromagnetism on the condition that his participation
would not be presented as an endorsement of Balkanization.35 The Journal
of Applied Physics, in which the statement announcing the symposium was
published, was both a preferred outlet for industrial researchers and one of
the few publications where a good balance of industrial and academic con-
tributions was in evidence.36 The venue of publication and makeup of the
symposium indicated that this new field would aim to bridge the divide be-
tween academic and industrial communities while also including an applied
constituency that had previously had little say in APS affairs.
From an organizational standpoint, the symposium resulted in a short-
term stalemate. Featuring voices both for and against a division, the sympo-
sium, like the circular letter before it, produced an overall preference for a
committee rather than a division. Smoluchowski wrote to Van Vleck follow-
ing the symposium: “My own feelings are quite optimistic now: I hope we
will be able to have a ‘solid’ committee acting according to our original plans,
avoiding at the same time the dangers of ‘pressure groups’ and other draw-
backs which you have mentioned.”37 This sentiment expressed the broad
consensus that the needs of the inchoate solid state community could be
met by either appointing or electing a committee to organize meetings and
symposia. The committee would have no permanent membership and collect
no dues.
Van Vleck used the attention the symposium generated to consolidate
opposition to divisions. He proposed a prophylactic measure against fur-
ther division-making—a standing Committee on Programs within the APS
that would have representation from a range of subject areas and hold the
responsibility for organizing meetings so as to benefit each. A petition, ad-
dressed to Darrow and signed by Van Vleck and a number of his like-minded
colleagues, argued: “It is preferable to have a committee of the Society on the
whole subject of symposia rather than a separate committee for each special
field, for two reasons: In the first place, the danger of catering to pressure
groups would be avoided, and in the second place, there would be furnished
a better safeguard that the number of symposia would be neither excessively
BALKANIZING PHYSICS 67

large nor too small.”38 Smoluchowski and the group of six endeavored to
work with Van Vleck on this proposal. Upon receiving a copy of Van Vleck’s
letter, Smoluchowski replied: “I see no reason why each group of physicists
representing a definite broad field should not elect one or two men to your
General Symposium Committee which would serve as outlined in your letter.
However, I do think these men should be elected, not appointed.”39 Smo-
luchowski was concerned that younger physicists felt alienated from the soci-
ety’s systems of governance and would benefit from an organizational struc-
ture that encouraged them to participate more fully in society business. In
the absence of a committee or a division dedicated to solids, Smoluchowski
insisted on elections to determine the constitution of any general committee
in order to ensure that the system was not oligarchical. Van Vleck, in keeping
with his preference for informality, found such democratic machinery dis-
tasteful. He considered contested elections a source of ill will and uncon-
tested elections pointless formalities.40
Although Smoluchowski and Van Vleck had come closer to agreeing on
a sequence of practical measures, their core disagreement about how the so-
ciety should operate persisted. Van Vleck preferred a small society, concep-
tually unified and focused closely on pure research, which would organically
produce the individuals willing and able to conduct the committee work nec-
essary to arrange meetings. Smoluchowski, on the other hand, saw a need to
introduce institutional apparatus in the society in order to achieve his aims.
Seitz pointed out another consequence of Van Vleck’s plan, writing to
Smoluchowski, “I believe that a large number of individuals whose primary
interest is in applied physics will still be disgruntled.”41 Seitz foresaw steep
postwar growth in industrial employment of physicists and suggested that, if
steps could not be made to address their needs within the Physical Society, a
separate society for applied physics be established. Nonetheless, he agreed to
support Smoluchowski’s proposal, which would combine basic and applied
work within the solid state division.42 Seitz’s conviction that applied physi-
cists would be dissatisfied by a general program committee is indicative of the
widespread acknowledgment that industrial and applied physicists generally
favored more, rather than less structure within the APS.
These discussions were ultimately academic. Efforts to organize a com-
mittee, either topical or general, were thwarted by a constitutional technical-
ity; provisions for standing committees did not exist in the APS constitution.
Since the purpose of APS divisions, as already in practice by the Division of
High-Polymer Physics and the Division of Electron and Ion Optics, closely
68 SOLID STATE INSURRECTION

mirrored the proposed functions of the solid state committee and of the
general program committee, Darrow and the APS council deemed that an
amendment to the constitution in order to allow standing committees was un-
justified. The group of six shelved their efforts until the end of the war, when
physicists were less encumbered by war work and travel restrictions and the
political maneuvering required to marshal enough support for the solid state
division would be simpler.
Their effort was resurrected late in 1946, after the APS council lifted its
moratorium on new divisions. Amid the controversy surrounding the wis-
dom of divisions, the APS council had appointed a committee, headed by
Edward Uhler Condon, to study the issue and craft a more permanent pol-
icy.43 At the same time, “Authorization was . . . obtained to tell the ‘Group
of Six,’ who want to organize a Division of the Solid State, to go ahead with
their plan.”44 This by no means represented an end to the debate over the
role of divisions. At the next meeting, Condon’s committee conferred with
the council, with the result “that both the Committee and the council found
themselves confronted with irreconcilable viewpoints, and the matter has to
go over to the January meeting.”45
Substantial progress would have to wait until May of the following year,
when the council again, in Darrow’s sardonic paraphrase, “turned to its favor-
ite pastime of discussing the question of Divisions.”46 Delegates from the Di-
vision of Electron and Ion Optics and the Division of High-Polymer Physics
complained about the council’s heavy-handed approach to crafting policies
curtailing their activities. Divisional representatives had been excluded from
the committee that drafted the society’s policy regarding divisions, and as a
result they found their bylaws, and the activities permitted by them, abruptly
constrained. The society clamped down on the practice of divisions granting
associate memberships to individuals who were not APS members and pro-
hibited divisions from collecting their own dues. Darrow reported that “the
Divisions felt themselves completely at sea owing to the authority possessed
by the Council to make such changes without consultation with Divisional
officers.”47 The reaction prompted the council to allow representatives from
the divisions to take part in any future discussions pertinent to divisional ac-
tivities, but refused to forestall implementing the policy it had adopted.
Darrow reported the resolution: “The President shall appoint a commit-
tee of five members of the Council to study the relations between the Divi-
sions and the Council necessary to implement the policy adopted in September
1946. (The [emphasized] phrase is an essential part of this motion: sugges-
BALKANIZING PHYSICS 69

tions that it be deleted were made, but did not eventuate in any amendment
of the motion).”48 The council, in other words, reasserted central authority
over the divisions. The 1946 policy was designed to ensure that divisional
activities would be governed by the APS, rather than promoting narrower
interest groups. The council was particularly determined to restrict divisional
membership to APS members. It ruled: “The Division of High-Polymer
Physics may keep its present Associates indefinitely, but neither it nor any
other Division may elect Associates henceforward,” and further required that
“the By-Laws of the two elder Divisions shall be examined and brought into
conformity” with the rules the council had set out.49
The council’s active policing of any activities that might threaten the social
and conceptual cohesion of the society or usurp authority over membership
or fees failed to deter the DSSP’s boosters. The day after getting the go-ahead
from the society, Smoluchowski—who had moved from GE to the Carnegie
Institute of Technology in the fall of 1946—wrote to the council: “In view of
the recent favorable decision of the Committee on Divisions (under the chair-
manship of Dr. Fletcher) and the return of more normal conditions the ‘group
of six’ requests the Council to consider again its petition and to approve a
Division of Solid State in accord with the recommendations of the Commit-
tee on Divisions.”50 Just over a month later, at the November 30 council meet-
ing, the formation of a “Division of Solid-State Physics” was authorized.51 Its
status was officially recognized at the June 1947 APS meeting in Montreal.52
An announcement to the APS membership defined the DSSP’s scope
“to comprise all theory and experimental research pertaining to the physics
of the solid state, such as metals, insulators, phosphors, all crystalline sub-
stances, etc.”53 Metals had been substantially downgraded in the division’s
mission, but although the expansion of scope made for an entity with slightly
less fuzzy conceptual boundaries, it also created opportunities for it to en-
compass all styles of research, old and new, pure and applied. Adopting the
solid state as a demarcating line did not draw sharp conceptual distinctions
as much as it avoided them entirely. By compromising between the society’s
demand for topical continuity and the desire to represent industrial and ap-
plied physicists equally, the DSSP became a big tent.
The DSSP, and the loose community of physicists it represented, can be
seen as a disciplinary experiment. The proposal for a metals division exposed
several growing rifts within the physics community. As industry was becom-
ing a considerable force in American physics, the rift between academia and
industry grew into a matter of deeper concern. Two incompatible concep-
70 SOLID STATE INSURRECTION

tions of unity emerged in the face of questions about how physicists should
position themselves following the war. The first followed directly from the tra-
ditional view that physics should be a field dedicated to pure science. Unity
in this sense, of a field defined by the object of its study, derived its meaning
from the basic concepts that were supposed to correspond to natural catego-
ries. Unity in the second sense was political. It embodied the idea that a com-
munity could come together for a common purpose from far-flung provinces
as long as it was guided by a strong central organization. This type of unity,
in contrast to conceptual unity, required continual institutional maintenance.
American solid state physics, in the process of navigating these rifts, or-
ganized physicists not on the basis of shared techniques or conceptual tools,
but by professional entente. It was a category imposed on a group of physi-
cists who had a shared interest, but this interest was not in an encapsulated
realm of physical inquiry; it was in bridging an institutional gap and creat-
ing organizational representation for groups that were otherwise marginal-
ized. The question of how unity should be understood lay at the core of this
development.

THE POLITICAL UNITY OF SOLID STATE PHYSICS


Both those proposing the DSSP and those opposing it, those pushing for
greater influence for applied physics and those seeking to protect the primacy
of pure research, pledged allegiance to the unity of physics. They disagreed
about what it meant for physics to be unified. Solid state physics embodied
a political view of unity. Unlike the view of physics as unified by its core con-
cepts, which grew from the same nineteenth-century sensibilities that favored
classification schemes based on natural phenomena, the political view of unity
allowed physicists a great deal of latitude to define both the outer boundaries
of their subject and its internal structure. The institutional possibility of a
field as large and unorthodox as solid state physics required some physicists
to start behaving as though physics really was just what physicists do. Solid
state physics, as the first field organized on such a basis, was a classification
unlike any other before it. Its success, in the long run, would legitimize classi-
fying physics on the basis of transient professional needs.
The unusually broad character of solid state physics was evident as early
as the January 1945 symposium where the proposal for a division (or a com-
mittee) was first submitted to communal scrutiny. The slate was strong with
theoretically sophisticated talks. Gregory H. Wannier of the University of
Iowa outlined new applications of statistical methods to cooperative phenom-
BALKANIZING PHYSICS 71

ena.54 Van Vleck gave a review of theoretical approaches to ferromagnetism,


beginning with phenomenological treatments of the early twentieth century
and continuing through descriptions of competing quantum mechanical ap-
proaches. These approaches were based on exchange interactions, which,
although part of the standard discourse among quantum theorists since the
late 1920s, were still sufficiently obscure within the general American physics
community that Van Vleck needed to caution his audience that “[exchange]
forces cannot be described in simple intuitive language.”55
The symposium also demonstrated a commitment to the applications of
solid state physics, and included contributions from Richard M. Bozorth and
Howell J. Williams of Bell Labs, who described their work as “understanding
of the behavior of magnetic materials in apparatus developed as a part of the
war effort.”56 Watertown Arsenal’s Clarence Zener framed his treatment of
the fracture stress of steel using unabashedly applied rhetoric, beginning by
noting that the “sinews of warfare, namely guns, projectiles, and armor, are
made of steel,”57 and Otto Breck of the Shell Development Company argued
that the increasing importance of catalytic processes in chemical technology
merited further attention to the topic from physicists.58
The broad cross section of work on display was representative of the
similarly broad range of approaches and questions the DSSP ultimately uni-
fied. Van Vleck’s concern with how the exchange interaction might provide
a more robust causal account of ferromagnetism had little to do conceptually
with Zener’s interest in the phenomenology of steel. The fact that they both
addressed some property of solid matter was a superficial commonality at
best. Van Vleck’s discussion of ferromagnetism outlined how the ascent of
quantum mechanics permitted a new causal understanding of the phenome-
non. Such a presentation assumed his idea of the natural ordering of physics:
the investigation of general laws allowed application to, and explication of,
specific systems. Van Vleck, representing cutting edge theoretical work, was
welcomed into the solid state union, where he found himself among the likes
of Zener and Breck, with their overtly applied thrust, and Percy Bridgman,
who, under the auspices of solid state, continued his long-standing research
program on the physics of materials under high pressures, which aimed at
qualitative description of the bulk properties of materials and avoided delving
into quantitative generalization.59
Grouping these disparate enterprises within the same province of physics
guaranteed that the new field would not respect existing, conceptually de-
fined barriers between basic and applied physics. Solid state defined borders
72 SOLID STATE INSURRECTION

around a new area of physics based on convenience. The solid state was an
expedient category because it was broad enough to encompass such a wide
range of topics. Its breadth assured that it would not discriminate against
industrial or applied physicists, who could often not state their focus area
narrowly, allowing the DSSP to span academic and industrial territories that
were otherwise isolated from each other.
The strong applied component and expansive scope of solid state physics
that came with its institutionalized form are further evident in its pedagogy.
The first textbook to describe physical approaches to solid matter compre-
hensively, Frederick Seitz’s Modern Theory of Solids, appeared in 1940. It
focused on the transition from classical to quantum approaches, with par-
ticular emphasis on the approximation methods that made regular crystal-
line solids susceptible to quantum mechanical description.60 Charles Kittel’s
Introduction to Solid State Physics became the standard text after its second
edition in 1955.61 The second edition expanded the textbook by about two
hundred pages over the original 1953 printing. Much of the additional mate-
rial dealt with practicalities that would be relevant to engineers and industrial
physicists. Compared with Seitz’s formalism-heavy style, Kittel’s approach to
theory was straightforward, in most cases relegating full quantum mechanical
treatments to the appendixes. Kittel’s textbook also dedicated more space to
applications, addressing in detail, for example, the properties of alloys and
the behavior of transistors, illustrating concepts with descriptions of exper-
imental techniques and appeals to easily observable laboratory phenomena.
Having become the standard text, Introduction to Solid State Physics rep-
resented a field with a strong applied inflection. As John J. Hopfield remarked
in his recollections of his training in solid state at Cornell in the 1950s: “The
weakness of the book was that it left you (as a theorist) with no idea of where
to start to develop a deeper understanding of any of the topics covered.”62 It
was the textbook Mildred Dresselhaus adopted when hired to teach a theory
of solids course at MIT that would be more accessible to engineers than the
highly abstract style that dominated John Slater’s physics department. Her
theory of solids course addressed the shortcomings Hopfield identified by
supplementing Kittel’s book with 302 pages of handwritten, photocopied
notes providing a methodical presentation of crystal structure and lattice dy-
namics leading into a detailed presentation of the electronic states of solids,
which the course dwelled on because “for most of the practical applications
of solids to our technological development, it is probably the electronic prop-
erties that are of the greatest interest.”63
BALKANIZING PHYSICS 73

Seitz and Kittel wrote for different audiences. Seitz assumed a stronger
background of his readers, targeting graduate students and practicing phys-
icists. Kittel’s text was designed to be accessible to undergraduates. The
differences nonetheless ran deeper. By the 1950s, solid state had not only
been established as a much broader enterprise than Seitz’s treatment would
suggest, but its first major coup, in the form of the transistor invented at Bell
Laboratories, had come from industrial quarters.64 To be marketable in the
1950s, a text on solid state physics had to take into account the range of the
field’s applications, not just its conceptual structure, and remain accessible to
chemists and engineers. As Kittel noted in his preface: “Solid state physics
is a very wide field.”65 Discontent with the name “solid state physics,” which
persisted long after the name was validated by the APS and emblazoned on
textbook covers, was a symptom of a deeper dissatisfaction with a category
possessed of little inherent cohesion.
The field nevertheless managed to hang together, if loosely, on the
strength of common professional objectives. The political unity manifested
by the formation of a solid state community in the United States differed in
three substantial ways from conventional conceptual unity. First, it was insti-
tutionally imposed. Solid state physics could be said to be unified because it
was guaranteed cohesion through institutional representation. Because solid
state was so diverse, it required an organizational infrastructure if its various
sectors were to avoid being annexed by other areas of physics, other sciences,
or branches of engineering. As MIT electrical engineer Arthur von Hippel
observed in 1942: “The fence between the two fields [physics and electri-
cal engineering] is falling into disrepair. The electrical engineer has to learn
and to apply atomic physics in order to understand and improve his new
tools, and the physicist is beginning to talk about ‘high Q’s’ and ‘characteris-
tic impedances’—and seems to like it.”66 The physicists who could speak the
language of electrical engineering—or mechanical engineering, chemistry, or
metallurgy—tended to be those who would be classified as solid state physi-
cists. The weak conceptual boundaries that kept these fields apart meant that,
if physicists interested in certain types of solid state problems were to be kept
within physics, they would need institutional support and encouragement.
Second, solid state was a malleable union. Its form was not supposed to
be objectively fixed by any facts about the physical world. Such flexibility car-
ried strategic potential. It allowed solid state physics to define itself in such a
way that it might grow, adapt, and compete for funding and prestige. It could
change its scope without endangering its standing. A solid state physicist
74 SOLID STATE INSURRECTION

could explore a new area of industrial interest, for instance, without trans-
gressing the topical boundary of the field. Such flexibility proved critical as
industry became a more prominent element of American physics in the post–
Second World War era.
Third, political unification allowed solid state to embrace the applied
consequences of scientific research. Conceptual unity assumed that engi-
neering applications of scientific knowledge lay permanently outside any uni-
fied field of physics. Solid state took a more flexible approach. By so actively
seeking to provide applied and industrial physicists representation, it linked
the traditional basic research arm of the physics community with a growing
industrial sector in which technological needs demanded to be filled more
forcefully than explanatory lacunae. These differences ensured that solid
state remained a viable subfield, but also branded it an outsider. The field’s
path over the subsequent decades reflects both the flexibility it enjoyed and
the difficulties it confronted as a result.
The debate over unity was, at core, a debate over what shape physics
would take in the postwar community. It played out as a turf war within the
APS. The growing constituency of industrial and applied physicists was
out of step with the society’s traditional focus on basic research. Industrial
growth was outstripping academic growth, and a strong industrial presence
in the APS threatened to alter the society’s character by suggesting a broad-
ening of its mission into areas some thought should not qualify as physics.
Applied physicists were no less emphatic about their identity, however,
pointing to their training and to the centrality of physical principles to their
work, denying that manipulating and applying these principles made them
less worthy of inclusion in the field than those who set out to discover them.
They promoted institutional mechanisms that would allow them to operate
within the APS while still maintaining a measure of autonomy. Building on
G. P. Harnwell’s big tent ideal, in which the term “physicist” would be gen-
erously bestowed, advocates for industrial and applied physicists sought to
reform American physics by promoting a broad topical scope and erasing
topical and institutional value distinctions among its membership.
John Van Vleck championed the traditionalist point of view in response
to this challenge, staunchly maintaining that researchers with strictly applied
interests fell outside of the APS mission and that including them would dilute
the atmosphere of free exchange that characterized prewar American physics.
His idea of unity was a purely conceptual one: physics was unified by a set
of first principles that constituted the targets of physical investigation. The
BALKANIZING PHYSICS 75

search for and manipulation of those principles held physics together largely
undifferentiated. Van Vleck correspondingly opposed attempts to build
bridges over which researchers who were not interested in those questions
might swarm.
The DSSP, taking as its central mission the problem of bridging the gap
between industry and academia, emerged amid the tension between these
opposing views of physics. To a limited extent, it succeeded in reconciling
them. Karl Darrow and the APS council perceived a danger in allowing in-
dustrial physicists to lose the ability to identify with the physics community,
but also insisted that industrial physicists’ primary allegiance within the APS
be to a topic area and not to industrial applications per se. Because solid state
physics emerged within the Physical Society’s divisional structure, it could
not be oriented overtly toward industrial interests and instead broadened its
scope to the point where it served in a fashion similar to the big tent phys-
ics community Harnwell envisioned, but on a more limited scale. The grand
compromise that resulted in the DSSP aimed to fulfill the APS mission, not
by closing off the division to those doing applied work, as Van Vleck would
have liked, but by bringing applied physicists into contact with their basic
counterparts within the confines of the society.
In order to make this compromise work, the DSSP sought political unity.
So as to maintain a wide-ranging field, serving both academic and industrial
physicists working on both basic and applied problems, the framers of Amer-
ican solid state physics distanced themselves from the ideal of conceptual
unity that had characterized the mission of the prewar APS. Solid state phys-
ics, to the extent it was a distinct unit, was distinct not by virtue of a well-
framed research program or a common experimental approach, but by virtue
of a community consensus, imposed and maintained by institutional decree.
4

THE PUBLICATION PROBLEM

The Physical Review continues to grow and to have financial problems.


We have heard statements to the effect that probably no single individual
is interested in more than one-tenth of the contents of the Review.
—ALAN T. WATERMAN, 1955

Solid state physics, newly demarcated, grew quickly under the charge of its
nascent American Physical Society (APS) division. From the group of six and
the small confederation of fifty-odd metals physicists who had guided its for-
mation, the Division of Solid State Physics (DSSP) grew to almost five times
that size by the time the membership was submitted to its first unofficial cen-
sus in 1948.1 By 1961 the division enrolled around eight hundred physicists,
who constituted approximately 5 percent of the American Physical Society’s
total membership, at a time when few joined divisions.2 The division’s ability
to swell its ranks marked its viability shortly after its formation, but member-
ship was only one dimension of the division’s growth. It also developed an
increasing measure of autonomy, raising questions about its relationship with
the APS. And its members contributed to the flood of papers that strained
the capacities of existing journals. These factors combined to force a reck-
oning in the 1950s about the mission of solid state physics. Did it aspire to
maximize its collaborative potential with neighboring fields, or to prove that
it belonged among the pantheon of pure physics?
In 1950, APS secretary Karl Darrow, at the APS council’s behest, sug-
gested to the chairmen of the three divisions then established that they con-
solidate their contributed papers and symposia at the March meeting of the
Physical Society, a practice that went gradually into effect over the following
THE PUBLICATION PROBLEM 77

few years.3 Van Vleck remained concerned for the unity of physics, pushing
unsuccessfully for the March meeting to be discontinued and divisional meet-
ings moved to June in order to stem the flow of professional congresses that
he felt exacerbated topical, temporal, and geographical rifts. Although Van
Vleck’s opinion carried considerable weight, he was unable to muster wide-
spread support for his position; lacking a clear consensus, the council elected
to leave matters as they stood.4
Divisional hegemony over the March meeting grew. By the beginning of
the 1960s, the DSSP had undertaken so much upon its own authority that
Frederick Seitz, who was then serving on the APS council, was moved to
chastise Elias Burstein, secretary-treasurer of the DSSP: “[The] Division of
Solid State Physics has been getting out of hand, and is using loopholes to
take independent action that seem improper to me and probably will to the
Committee.”5 Seitz was not alone in his assessment that the DSSP was over-
reaching. Karl Darrow submitted a letter to the APS council complaining,
“The Division of Solid State Physics gives at times the impression of acting as
though the March meeting were its own private affair to locate as it chooses.”6
The APS power structure mobilized to bring the obstreperous division to
heel.
The scrap developed because the DSSP had, without approval from the
APS council, made arrangements for its membership to attend the 1961
March meeting in Monterey, California. After the council, ignorant of the
DSSP’s plans, accepted an invitation from Buffalo, New York, they were
forced to backpedal on promises made to Buffalo hotels and conference cen-
ters. Those charged with issuing the red-faced mea culpas were understand-
ably miffed. Seitz warned Burstein that the practice of planning meetings,
especially in conjunction with other societies, without the blessing of the APS
would “cause endless confusion and undermine the prestige of the APS,” and
continued: “Should there be a substantial feeling at present among a group
of solid state workers that the APS is too confining, the group has the choice
of starting its own organization outside the Society. It cannot, however, have
complete autonomy and still enjoy the prestige and privileges of the APS.”7
Burstein had, in the past, expressed the view that it would be “more desir-
able for the Division to have an APS meeting to itself, except for occasional
planned joint meetings with other Divisions of APS,” but showed no indica-
tion of wanting to split from the Physical Society entirely.8
The Monterey episode indicates two facets of solid state physics’ devel-
opment through the 1950s. First, it was rapid and robust. The division’s
78 SOLID STATE INSURRECTION

expansion in size and influence allowed it to mediate interactions between


solid state physicists and associated groups in the American Chemical Soci-
ety and the International Union of Pure and Applied Physics when a decade
earlier it was scarcely more than a modestly conceived mechanism for arrang-
ing symposia within APS meetings. Second, however, it was somewhat out
of step with the larger organization and sometimes chafed at the strictures of
a centralized society. The tensions between the DSSP and the APS, which
were sparked by the DSSP’s increasing autonomy, realized some of Van
Vleck’s fears about Balkanization. Despite Seitz’s suggestion that the DSSP
form a separate society if it could not operate within the confines of the APS,
such a possibility does not appear to have received serious consideration.
Nonetheless, the friction generated by the division’s independent functioning
demonstrates the compartmentalizing influence divisions had over a subfield
that was only loosely organized just ten years earlier.
As solid state established itself within American science, these two facets
predominated in determining its goals. Solid state was viable, but it was also
perched precariously on the boundaries of the physics community. As Seitz’s
emphasis on prestige indicates, the DSSP gained more from its association
with the APS at this stage of its development than the society at large gained
from the division’s activities, especially when those activities proceeded with-
out the knowledge or approval of the council. As it managed its growth and
negotiated its place within American physics, solid state physics honed its
identity as a physical subspecialty and negotiated its place within the phys-
ics community. This chapter explores how growth, both within solid state
physics and in the larger community, challenged the fledgling discipline,
and demonstrates that the strategies the field’s early leaders adopted to meet
those challenges cemented solid state within physics, while preserving the
close connections with neighboring disciplines its early leaders had taken
pains to cultivate.
The most pressing way in which boundary concerns manifested them-
selves involved what became known among physicists as “the publication
problem.” As the population of the American physics community in general
—and the solid state community in particular—swelled, the established pub-
lication outlets strained under the pressure of increased submissions and
ballooning publication costs. APS and American Institute of Physics (AIP)
journals, the Physical Review especially, developed damaging publication
delays, threatening priority in fast-moving fields and prompting physicists
to clamor for a solution. These strains pressured the emerging discipline of
THE PUBLICATION PROBLEM 79

solid state physics to reevaluate its identity by confronting anew the question
of what audience it sought to reach. Some solid state physicists saw this as an
opportunity to escape the confines of the Physical Review and build a closer
association with chemical and metallurgical publications outlets. Others per-
ceived the opportunity to start a new journal dedicated to solid state phys-
ics, asserting the field’s independence.9 Still others fought for a resolution
within the established order, which would maintain solid state’s newly won
place alongside the other subdisciplines of physics with which it had, until
then, shared space in general physics journals, including the highly regarded
Physical Review.
The third option, which reaffirmed solid state’s identity as a field of phys-
ics, eventually carried the day, but not before a gut-check moment for the
new field. Wrestling with questions about what solid state was positioned to
accomplish, and with whom it should be communicating, helped to define a
clearer sense of the field’s mission and overcome a portion of the anomie that
characterized its early adolescence. By providing solid state with a clearer
sense of place within the field of physics, the publication problem also fanned
the first pangs of animosity between solid state and high energy physics,
which would become a central theme of the subsequent decades. As the tec-
tonics of the American physics community shifted in the postwar years, the
resultant tremors spurred solid state physicists to take a clear stand on where
their discipline would be situated and with whom it would cast its lot. That
decision had long-ranging consequences for the terms on which solid state
interacted with neighboring fields, both inside and outside physics.

BACKGROUND TO THE PUBLICATION PROBLEM:


THE JOURNAL EXPLOSION
When the Physical Review became a journal of international note under the
editorship of John Torrence Tate in the 1930s, its publication rates shot up.
It published 172 articles in 1925, the year before Tate took the helm. In 1931,
it achieved a pre–Second World War peak of 380. New journals launched
around this time siphoned away some articles from the Physical Review,
tempering its growth, but for the rest of the decade its publication rate held
steady at around 300–350 papers per year (see figure 1.1).10
The war precipitated a publication lull. The self-imposed embargo on
publication of nuclear research, the diversion of physicists to the war effort,
and the breakdown of international scientific communication conspired to
decimate the journal’s output, which fell as low as 78 articles in 1945. But
80 SOLID STATE INSURRECTION

it rebounded with a vengeance, clearing its prewar high-water mark in 1948


and more than doubling it by 1953. In 1956, Physical Review published
1,212 articles, more than three times 1931’s volume, despite the fact that
competing journals witnessed similar increases. The increase in publication
reflected an increase in the number of physicists. The rate of PhD produc-
tion, which had increased through the 1930s, only accelerated following the
war. It also reflected an increase in the abundance of research funding. Not
only were physicists more abundant, they were also flusher than they ever had
been.11 The population of credentialed physicists was ballooning, and they
had ready access to the resources required to translate their labors into pa-
pers. The Physical Review was unprepared for the deluge that the confluence
of these factors caused.
As the journal and the community it represented grew in size, both diver-
sified in character, a change driven by the wave of specialization Van Vleck
was doing his level best to stem. Specialization was accompanied by demo-
graphic shifts, which became an avid point of discussion in Physics Today, a
monthly magazine founded in 1948. The magazine, conceived as a forum for
articles of general interest from all branches of physics, was itself a response to
a diversifying discipline. It had first been proposed at the National Research
Council’s 1944 Conference of Physicists in Philadelphia, which had estab-
lished the need for stronger communication mechanisms linking physicists
and empowered the AIP to pursue it. Physics Today appeared, according to
AIP director Henry Barton, as an “Institute journal suitable for circulation to
all physicists . . . a readable report and discussion of what concerns physics
and physicists—today.”12 Advocacy for the magazine through its early years
centered on its capacity to address the challenge of a fragmenting community:
“I am sure you are as keenly interested as the rest of us,” Gaylord P. Harn-
well wrote to Frederick Seitz in 1950, “in making a go of Physics Today in
order that the Institute may take on some unity and character through having
a journal reaching all its members.”13 A general character was Physics Today’s
most distinctive attribute. It presumed that some issues were of interest to all
physicists, and provided a forum in which to discuss them.
In a note opening the inaugural issue, editor David A. Katcher wrote, “As
fields of research become more and more specialized, the knowledge shared
by research workers in their technical journals is becoming a secret under-
stood only within the specialized field,” and described the new journal as a
prophylactic against insularity:
THE PUBLICATION PROBLEM 81

Physics Today is for the physicist, to inform him in comfortable, everyday


language, of what goes on and why and who goes where. But it is also for the
chemist, the biologist, the engineer, to tell them of the science towards which
they are driven by so many of their investigations; it is for the student, the
teacher, the lawyer, the doctor, and all who are curious about physics; it is for
administrative officials who deal with research; it is for editors and writers
whose profession puts them midway between what is done and how it should
be reported; it is for you, whatever reason brought you to this page.14

Katcher’s quixotic hopes for the magazine’s reach tells us more about the
problems that precipitated its founding than its immediate impact. It was
no longer possible, in the days of rapidly growing research output in more
and more subspecialties, for a physicist to stay current on the whole range
of issues that might hold potential interest. As solid state had shown the year
before, the hoary fundamentalists of the old APS no longer had a monopoly
on what could or could not be called physics. Keeping the chaotic range of
new subfields in some kind of rational order required efforts to open common
lines of communication. Physics Today responded to what Charles Weiner
has called “the spirit of the forties.”15 Physicists, as they became ensconced in
their specialties to a degree they had not been before, required an outlet that
reaffirmed their shared identity and protected their mutual claim to postwar
public approval.
Physics Today debuted with a cover featuring J. Robert Oppenheimer’s
iconic porkpie hat resting on the 184-inch cyclotron at Lawrence Berkeley
Laboratory (figure 4.1), affirming the status of nuclear physics as the cover
story of the late 1940s. Nevertheless, early issues of Physics Today were scru-
pulously attentive to breadth. The abbreviated eight-issue run comprising the
magazine’s first calendar year included features on cyclotrons, neutrinos, and
liquid helium, but also explored connections between physics and cancer,
electrical phenomena in the atmosphere, the origin of the earth, and oceanog-
raphy.16 Promoting unity meant adopting a catholic editorial philosophy that
welcomed perspectives from what previously would have been considered
fringe provinces of physics. Physics Today actively courted any scientists, re-
gardless of institutional affiliation or professional status, who self-identified
as physicists or thought that physics research might be useful for their own
work.
The magazine’s second year brought indications that these efforts had
82 SOLID STATE INSURRECTION

Figure 4.1. Cover of the first issue of Physics Today. Reproduced from Physics Today 1,
no. 1 (1948), with permission of the American Institute of Physics. Cover photo ©
University of California, Lawrence Berkeley National Laboratory

been successful at promoting outreach, if not at instilling unity. Physics To-


day would not become a mouthpiece for the emerging nuclear/particle phys-
ics constituency that was in the process of establishing itself as the standard
bearer of American physics in the eyes of the public and federal funders. In the
THE PUBLICATION PROBLEM 83

January 1949 issue, George Gamow published his philosophical suspicions


that physics was converging on an ultimate set of theories and concepts. “If
and when all the laws governing physical phenomena are finally discovered,
and all the empirical constants occurring in these laws are finally expressed
through the four independent basic constants,” he speculated, “we will be
able to say that physical science has reached an end, that no excitement is
left in further explorations, and that all that remains to a physicist is either
tedious work on minor details or the self-educational study and adoration of
the magnificence of the completed system.”17 Although Gamow took care to
note that he was making no bold predictions, he did record his instinct that
such a theory of the micro-scale was within grasp.
Gamow’s suggestion that a self-consistent theory of elementary particles
would render the rest of physics uninteresting provoked some indignant let-
ters to the editor. Representative of these complaints, two long examples of
which were accorded unprecedented column inches in the March issue, was
Richard C. Raymond’s. The Pennsylvania State College (later University)
thermodynamicist derided Gamow’s “unbridled speculation,” and took him
to task for failing to appreciate the complexity of the physical world.18 By
1949, the editorial courting of diverse demographics was showing results,
and a vocal portion of Physics Today’s readership found the grand specula-
tions of nuclear and particle physicists outré.19
Nevertheless, contrary to the journal’s aims of unifying the community,
some found its scrupulous attention to breadth alienating. George R. Har-
rison, then dean of the School of Science at MIT, cautioned Harnwell when
the latter assumed control of an ad hoc committee to review the structure and
effectiveness of Physics Today: “There is a species of nuclear physicist who
has no use whatever for Physics Today, and who does not hesitate to make his
views known. Listening to such people I would have thought that we should
give up the Journal long since, but always when I have come to this conclu-
sion I have found in other walks of life a set of opposite views to counter-
balance them.”20 Along similar lines, Samuel Goudsmit, poised to take over
the editorship of the Physical Review, commented in response to Harnwell’s
invitation to serve on the magazine’s Governing Board: “It is obvious that
Physics Today is meant for the broader group of non-academic colleagues
with whom I have too little contact to know their needs,”21 and John Van
Vleck replied that “some less pretentious publication, such as an appendix to
the Physical Review, would be adequate for my own personal needs.”22
The unifying mission of Physics Today ran abruptly into the realities of
84 SOLID STATE INSURRECTION

its fragmentation. Those who perceived themselves as occupying the hard


core of American physics, nuclear physicists in particular, had little use for a
publication that was making active overtures to those engaged in far-ranging
applications of physics. Much like solid state physics, Physics Today assumed
the desirability of political unity and encountered pushback from the bloc of
physicists who retained a firm commitment to conceptual unity.
The magazine, despite being spurned by some, found an enthusiastic au-
dience among a wide range of physicists. Given its commitment to breadth
and efforts to instill unity, however haltingly, it is unsurprising that many early
Physics Today articles confronted demographic issues. Not least among them
was the question of industrial physics and its place. The second issue con-
tained an apology for industrial research, which acknowledged that industry
was pushing the boundaries of physics. “How can one measure the comfort
of a floor?” asked Howard A. Robinson: “In years gone by this would not
have been a proper question to ask a physicist, but in the past decade . . .
physicists . . . have discovered, somewhat to their amazement, that these
border-line problems in which an individual is part of the measuring system
can sometimes be solved. Thus the broadening of physics to include phys-
iological manifestations is now well established.”23 Robinson articulated the
new orthodoxy among the industrial set that physics could no longer be con-
tained within a traditional academic definition of pure science, or the newly
popular category of basic research.
Broadening into industrial areas was attributable in part to the legacy
of war research. Even though APS membership was expanding at record
rates during and after the war, PhD production had stalled. Vannevar Bush
lamented in the very first issue of Physics Today that “we foolishly ceased to
train [physicists] during the war.”24 The lack of traditionally trained PhD
physicists created a problem for academic programs looking to expand their
ranks or replace retiring faculty. Frederick Seitz, then at the Carnegie Insti-
tute of Technology, put the problem to William Shockley of Bell Labs early
in 1945:
Graduate education was stopped cold in the winter of 1940–41. Moreover,
many of the men who would have entered graduate school then will probably
never do so. If the present situation lasts another two years, there will be a
missing generation covering a range between seven and ten years. In a recent
survey the American Institute of Physics has decided that no less than 2000
Ph.D. physicists, who would have been created had the educational situation
THE PUBLICATION PROBLEM 85

continued as of 1939, will never receive complete graduate training. These are
just the men you would look forward to hiring. It seems to me there are three
choices:
(a) Hiring Ph.D.’s who will be thirty or over when they join your staff.
(b) Hiring men who have not had formal graduate training but who have
received an apprenticeship like at the Radiation Laboratory.
(c) Waiting until a new group comes along in 1950 or later.25

Seitz was sour on each of these options. He harbored a prejudice that physi-
cists over thirty had lost too many of their most creative years, regarded those
without doctorates as risky investments, and dismissed the third option as
“the worst of the three prospects.” He concluded pessimistically that “good
physicists will be difficult to obtain in the immediate post-war period and
that you will have to be willing to make some concessions.”26 The physics
community’s growth, although robust, proceeded along nontraditional lines,
challenged the prewar professional status quo, and upset long-standing train-
ing and hiring practices.
Where academic institutions and quasi-academic research labs like Bell
saw concessions, other areas of industry saw opportunity. A wide range of
industrial interests proved more than willing to hire physicists who had cut
their teeth on war work, however unorthodox their training. The Radiation
Laboratory (Rad Lab) at MIT and Harvard’s Radio Research Laboratory
(RRL) had been particularly rigorous proving grounds for young solid state
physicists who would otherwise have been occupied by their graduate edu-
cation. An RRL administrative report boasted that “the requirements of RRL
were far more stringent than those of even a peacetime industrial firm.”27 The
success that both the MIT Rad Lab and the RRL enjoyed bringing new tech-
nologies into the field conditioned the expectations for lab-to-marketplace
turnaround in postwar industry and exposed areas of research that were ripe
for industrial exploitation.28
A survey conducted by the AIP in 1954 showed the proportion of physi-
cists employed in industry gaining on the proportion employed in academia,
with 42.0 percent still within the academy and 35.8 percent in industrial
positions.29 Industries with direct interests in physical research, such as
communications, atomic power, instrument and electrical component devel-
opment, and aviation, employed the preponderance of industrial physicists;
however, physicists also found homes in less obvious venues, like the textile,
86 SOLID STATE INSURRECTION

petroleum, plastics, and photography industries, which had previously been


dominated by chemists.
Not only had the size of the Physical Society and the publication load of
the Physical Review ballooned by the mid-1950s, the physics community had
assumed a distinctly different complexion. The growth of American physics,
in population, publication volume, funding, and scope was unprecedented,
to the point of creating cost, backlog, and relevance problems for its core
publications. Furthermore, topical specialization and institutional reorienta-
tion within the community generated friction that threatened the discipline’s
unity. When existing journals became saturated, facing the prospect of a
growing backlog and damaging publication delays, these pressures prompted
physicists to consider how a response to the publication problem might also
be used to address the instability they saw in the professional sphere. The re-
sponse from solid state physicists, discussed below, exposes some of the fault
lines created by physics’ expansion into new professional and topical areas
that could not be easily accommodated by traditional practices.

BACKGROUND TO THE PUBLICATION PROBLEM:


THE TRAJECTORY OF FREDERICK SEITZ
The task of mapping out possible responses to the publication problem was
taken up by Frederick Seitz. A native San Franciscan, Seitz earned a phys-
ics and mathematics education at Stanford University, including a semester’s
intermezzo at the California Institute of Technology, before relocating to
Princeton for graduate school, where he joined in the first wave of Americans
trained specifically in the physics of solids. Seitz would go on to become,
through the central position he would occupy in the advisory apparatus of
American science, a force shaping solid state’s institutional evolution. The
intellectual oeuvre in which the mindset he brought to this position evolved
is therefore worth exploring in some detail.
While at Princeton, Seitz would form key professional connections and
take on an intellectual approach that informed his later scientific work and
institutional maneuvering. He encountered both Roman Smoluchowski,
who visited Princeton in the mid-1930s before emigrating from Poland in
1939, and William Shockley, who was John Slater’s graduate student at MIT.
Shockley was also a Californian and the two shared a cross-country road
trip in Shockley’s DeSoto convertible to begin the 1932–33 academic year.30
Shockley, Seitz, and Smoluchowski would go on to form half of the group of
THE PUBLICATION PROBLEM 87

six, which founded the American Physical Society’s Division of Solid State
Physics, as discussed in chapter 2.
Seitz’s training was just as important as the personal connections he made.
Through the mid-1930s, three centers had emerged for aspiring physicists
interested in solids. John Slater was lured from Harvard to MIT in 1930,
where incoming president Karl T. Compton gave him free rein to expand
the physics department in accordance with his vision. John Van Vleck, after
stints at Minnesota and Wisconsin, returned to Harvard, where both he and
Slater had earned their doctorates, in 1934.31 Finally, Eugene Wigner secured
a permanent position at Princeton in 1938 in part at the urging of Van Vleck,
but had, with the exception of a visiting stint at Wisconsin in 1937–38, held
down one temporary appointment or another at Princeton since 1931.32 It
was in this latter capacity that Wigner oversaw Seitz’s doctoral work, which
Seitz later remembered as “one of the most remarkable experiences of my
life.”33
Wigner, a Hungarian émigré, was an exception within this group. Slater
and Van Vleck had both been trained at Harvard, learning their quantum
mechanics from Edwin Kemble, who, in the 1920s, offered the first intensive
training in quantum theory available in the United States. Their experience
at the vanguard of quantum physics in the United States led Slater and Van
Vleck to see themselves as carrying the torch for American physics.34 Van
Vleck, in 1971, would bridle at an offhand suggestion that Slater was an heir
to the British tradition. The Belgian physicist Léon Rosenfeld, in a historical
overview of atomic theory, emphasized the formative nature of Slater’s post-
doctoral visit to the Cavendish laboratory, referring to him as “a physicist ed-
ucated in the British and American tradition.”35 Van Vleck sent Slater a copy
of the article, along with an expressive note: “I am usually something of an
Anglophile but the reference to your training . . . rather made my blood boil.
I’ll grant you that Slater is an English name but what the author says makes
about as much sense as it would be to say that I am Dutch-trained because
my name is Van Vleck.”36 Slater was equally eager to distinguish American
and European physical traditions, penning a Physics Today editorial in 1968
in which he attacked the conventional wisdom that American physics in the
1930s was dragged reluctantly into modernity by the influx of European émi-
grés.37 The hesitancy Van Vleck and Slater exhibited to sully their work with
what they saw as baser pursuits can be better understood within the context
of the pride they both took in representing American physics, and American
88 SOLID STATE INSURRECTION

theory in particular. Van Vleck’s resistance to divisions was one manifestation


of this phenomenon.
Wigner, in contrast, counted himself among “the Martians,” the group
of Hungarian scientists that also included Theodore von Kármán, John von
Neumann, Leo Szilard, and Edward Teller, who were chased from Central
Europe by Hitler’s rise.38 Wigner arrived in the United States with a back-
ground in chemical engineering, which he had studied at the Technische
Hochschule in Berlin, earning a doctorate in 1925. He recalled that his chem-
ical education, which was more theory-oriented in Berlin than similar train-
ing in the United States would have been, “came in handy many times in my
life in physics.”39 Indeed, Wigner’s career was characterized by remarkable
topical breadth, which often involved flirtations with chemical phenomena.
Moreover, Wigner came from a European tradition that was not shy when it
came to talking about the reality of physical microstructures, even if only pro-
visionally. He was immediately taken, for instance, with the discovery of the
neutron, and wasted no time employing this new tool to better understand
nuclear masses.40
Slater offers a compelling contrast on this score. He was famously embit-
tered by his experience as a postdoc at Niels Bohr’s Institute for Theoretical
Physics, snapping at Thomas Kuhn in an interview that he “never had any
respect for those people [Bohr and Hendrik Kramers],” after his experience
in Copenhagen. Although Slater later recanted, claiming that his differences
with Bohr were scientific rather than personal, this unguarded remark indi-
cates the lingering psychological influence Copenhagen had over Slater, neg-
ative though it might have been. While Slater was in Copenhagen, Bohr and
Kramers seized on his idea that light–matter interactions could be described
in terms of a suite of “virtual oscillators” that determined the probabilities of
allowed quantum transitions as a way to escape the quantization of light that
most others had accepted on the basis of Arthur Holly Compton’s explana-
tion of his eponymous effect. The result was the short-lived Bohr–Kramers–
Slater (BKS) theory, which denied light quanta at the expense of rejecting the
strict conservation of energy, which it treated as a statistical phenomenon.
By his own account, Slater felt as though he had been hijacked in service of
an agenda that was not his own.41 He came back to the United States having
been convinced that “Bohr was fundamentally of a mystical turn of mind and
I’m fundamentally of a matter-of-fact turn of mind.”42
The commitment to a calculation-based style Slater brought to solid
state can be traced in part to his experience abroad. He was repulsed by the
THE PUBLICATION PROBLEM 89

speculative approach he saw in Bohr, having been convinced by the BKS


experience that pursuing research on the basis of deeply held metaphysi-
cal prejudice was fruitless. Although Slater’s group at MIT, like Wigner’s at
Princeton, bled over into chemistry, it did so for different reasons: Wigner
was willing to employ approximation strategies gleaned from chemistry and
maintained a chemist’s willingness to provisionally commit to expedient on-
tological assumptions. Slater, on the other hand, realized that there was little
instrumental difference between cranking out wave functions for molecules
and cranking out wave functions for solids. Slater’s approach to solid state
and molecular theory, which, with the advent of electronic computers, would
involve throwing more and more computing power at ab initio calculations,
bore little resemblance to Wigner’s, which used the phenomenological fea-
tures of solids, rather than quantum mechanical first principles, as a concep-
tual starting point.
Wigner’s approach had a clear influence in the young Seitz, whose thesis,
“On the Constitution of Metallic Sodium,” would remain his most influential
intellectual contribution to physics. Published jointly with Wigner, it estab-
lished what became known as the Wigner–Seitz method for describing the
properties of metals. This approach negotiated between the two most preva-
lent alternatives at the time. The first, championed by Felix Bloch and Léon
Brillouin among others, was the free electron model, which aimed to describe
conduction and ignored chemical properties, which were the products of va-
lence, by modeling the interactions between free conduction electrons and
lattice vibrations in crystalline metals. The second, backed largely by Slater,
aimed to describe a wider range of chemical and mechanical properties by
calculating the influence of valence forces on metals.43 The former method
offered a ready tool with which to confront electrical conduction, but the sim-
plifications it introduced made it ill-suited for much else—for instance, it only
worked for ideal metals—and led many physicists to view it as unsatisfactory
as a result. On the other hand, Slater’s approach required laborious calcula-
tions, and, in an era before computing power made them tractable, could be
faulted for being too cumbersome to be practically useful.44
Seitz and Wigner sought out a middle ground between these two meth-
ods. Wigner, drawing from his training as a chemical engineer, was sensitive
to the notion that a theory of metals should describe more than electrical
conduction. The two set out a method of approximation that struck a balance
between solvability and verisimilitude, which would allow the free electron
model to be applied to real, not just ideal metals. The use of creative approx-
90 SOLID STATE INSURRECTION

imation methods to simultaneously simplify calculations and capture struc-


tural features of complex systems would become a signature of solid state
physicists’ approach.45
The influence of the 1933 paper in which Wigner and Seitz published
their method, which rapidly led to extensions and applications by Slater, Van
Vleck, Nevill Mott, and others, fueled Seitz’s early career. Through the re-
mainder of the 1930s, he explored positions in both industry and academia,
spending two years at the University of Rochester before finishing out the
decade at General Electric. Moreover, the conceptual approach embodied in
Seitz’s thesis was mirrored in his approach as an institution builder. Seitz
repeatedly sought to combine the diverse approaches endemic to solid state
into a cohesive whole. This often required sacrificing strict conceptual or
methodological continuity while uniting diverse approaches under a single
institutional banner.
In 1940, Seitz released his textbook Modern Theory of Solids. It was the
first comprehensive textbook devoted to the topic, and it would cement his
influence over the field for decades. It reinforced the approach to solid state
theory Seitz had acquired from Wigner, which emphasized creative approx-
imation and attention to properties that were traditionally considered chem-
ical alongside those more commonly found in physics training. Second, and
perhaps more important, it marked the beginning of his gradual transition
from practitioner to administrator. Through the 1940s and the early 1950s,
until shortly after his arrival at the University of Illinois, Seitz remained an
active research physicist. In the 1950s, he devoted increasing measures of his
time to assorted advisory committees and governing boards, many of which
would make crucial decisions about solid state and its direction.
Modern Theory of Solids begins by striking the topically ecumenical note
Seitz had inherited from Wigner. Seitz expressed his hope that the book
would serve the needs of three types of reader: “First, of course, students of
physics and chemistry who desire to learn some details of a particular branch
of physics that has general use; second, experimental physicists and chemists,
and engineers and metallurgists with mathematical leanings who are inter-
ested in keeping an eye on a field of physics that is of possible value to them;
and third, theoretical physicists of various stages of development who are
interested in the present status of that phase of solid bodies that deals with
electronic structure.”46 This statement expresses an early vision for the per-
missive scope of what would become solid state physics: a field, firmly within
physics, that was nonetheless broadly conversant with a variety of neighbor-
THE PUBLICATION PROBLEM 91

ing disciplines. The influence of this vision is discernible in Seitz’s actions


in the face of the institutional pressures that shook solid state in the 1950s.
By the time the publication problem began to weigh on physicists’ minds,
Seitz had become one of the most astute institutional animals in the solid state
community. He developed a facility for navigating the labyrinth of advisory
committees, society councils, and journal boards that gave him a low-angle
view of the growing field. In 1949, Wheeler Loomis lured him to the Univer-
sity of Illinois at Urbana-Champaign, beginning a process that would pro-
duce one of the most influential centers of solid state physics in the country,
second, perhaps, only to Bell Labs. From the center of the country, in an
emerging center of solid state research, Seitz began to parlay his integration
with disciplinary governance mechanisms into influence over the field’s di-
rection. When physics journals began to stagger under pressures of growing
backlogs, rising costs, and increasing specialization, he was therefore in a po-
sition to direct the response of solid state physicists.

COPING WITH THE PUBLICATION PROBLEM


Signs of trouble with the Physical Review appeared in the late 1940s. Up
until 1948, the journal turned a profit through subscription fees and page
charges. At the January 1948 council meeting, however, the editor John Tate
and society treasurer George Pegram made it clear that “the period in which
the Physical Review returned a net profit to the Society from subscriptions of
non-members has come to its end.” Failing to break even was not an immedi-
ate hardship, however, as long as other AIP journals were still profitable and
the society enjoyed a “handsome surplus,” which would sustain its flagship
publication for some time.47
The publication problem was on the radar, but would generate more heat
than light through the next few years. Tate’s death in 1950 and the transfer
of editorial operations to Brookhaven National Laboratory produced logisti-
cal issues aplenty to keep all concerned occupied as the journal grew thicker
and slipped further into the red. By 1953, the financial situation had become
pressing. Pegram’s report on the society’s financial situation warned that “the
Society may have done a little better than ‘break even’ during 1952, even
without taking into account the $20,000 donated by the National Science
Foundation to assist in meeting the deficit of The Physical Review; and that
he expects that in 1953 the margin of income over expenditures may attain
$30,000.”48
By the mid-1950s, increasing publication costs and delays spurred action.
92 SOLID STATE INSURRECTION

Both the American Physical Society and the American Institute of Physics
put institutional machinery into motion to address it. At the April 1955 APS
council meeting, Samuel Goudsmit, Tate’s successor as the Physical Review’s
managing editor, provided “a lengthy report on the situation ensuing from
the interminable expansion of The Physical Review.” In accordance with the
“ominous” financial prospects such expansion brought about, the council
approved steep hikes in both page charges and subscription rates. A motion
to split the journal, proposed to gauge opinion rather than to spur action,
was defeated, and the council also ruled unfavorably on a proposal that the
APS take over the Journal of Chemical Physics from the AIP. Nonetheless,
the prospect of major restructuring loomed. Goudsmit recorded his strong
feelings “that the American Institute of Physics should enlarge its journals.”49
The AIP had similar inclinations. Institute director Henry A. Barton com-
mented in March of 1955: “Pressure for publication of research results in
certain fields has again come to the point of severe strain,” and although he
did not promote any specific solutions, he assured his readers that “the In-
stitute stands ready to help study such problems and continually investigates
proposed ways of reducing publishing costs.”50 Barton and the AIP Govern-
ing Board, at their March meeting, appointed a joint AIP-APS committee to
generate recommendations for easing the publication burden. Demographic
changes complicated the committee’s mission, particularly the increasing im-
portance of industrial physics. Seitz made the observation, common by that
point, that “industrial organizations which were uninterested in physicists
prior to 1940 are now eagerly attempting to hire Ph.D.’s.”51 Solid state in
particular thrived on the growth of physics in industry, and the separate inter-
ests and professional challenges that drove industrial physicists contributed
to the professional instability solid state experienced amid the publication
crunch.
Alan T. Waterman, director of the National Science Foundation, singled
out the Physical Review as one site where diversification within physics could
be identified. Replying to Karl Darrow’s request for funds to support the
AIP’s publication study, Waterman reported hearing “statements to the effect
that probably no single individual is interested in more than one-tenth of the
contents of the Review.” He further suggested that if this really was the case:
“It may eventually be desirable or even necessary to restrict publication in
the journals of wide circulation to papers of more general interest. Questions
such as this could be studied objectively. Perhaps the recent vote with the
Physical Society on the desirability of splitting the Review has already shed
THE PUBLICATION PROBLEM 93

some light on the question.”52 The field of physics was becoming compart-
mentalized and the growth of topical enclaves put pressure on a journal struc-
ture that was conceived for a small community with few internal divisions.
These considerations motivated the second circular letter—the first being
“The Present War Is a Physicist’s War” distributed by the group of six—that
would bear heavily on the fate of solid state physics. In March 1955, on behalf
of the AIP-APS joint committee, Seitz circulated a questionnaire to selected
solid state and chemical physicists asking if they would welcome an exodus
of solid state publication from the Physical Review to the Journal of Chem-
ical Physics (JCP), which the committee tentatively proposed renaming the
Journal of Solid State and Chemical Physics.53 The JCP was published by
the AIP, but, as the Division of Chemical Physics was being formed in 1949,
a few members of the APS began advocating for the society to take it over.54
The idea had been bandied about for several years, but failed to produce
any substantial changes. Having been a regular element of council meeting
discussions, however, the notion of acquiring the JCP, not just for chemical
physics, but for solid state as well, was a logical option to pursue.
At the time, the authorship of the JCP was composed principally of chem-
ists.55 The field known as chemical physics—as distinguished from physical
chemistry—was conceived and operated as an interdisciplinary field, but it
was populated predominantly by those trained in chemistry, even though
they often published in physics journals, and the chemical physics graduate
programs across the United States tended to be housed in chemistry depart-
ments.56 Colocating chemical and solid state physics publications in JCP
would therefore necessitate a much closer relationship between the solid state
and chemistry communities than the names alone would suggest. With that
consideration in mind, Seitz advanced the suggestion cautiously:
It is the writer’s opinion that this transformation would inevitably make the
journal less valuable to the chemists who do not participate actively in the APS
or AIP and hence would act to the disadvantage of this important segment of
the scientific world. For this reason the change would probably not be justified
unless a great majority of the solid state physicists would be willing to use the
transformed journal as their principal outlet for publication, leaving the Phys-
ical Review in the main to the nuclear physicists and diverse minorities which
would not feel at home in the revised journal.57

The enclosed survey asked those interested to indicate, (a) their field (solid
state, chemical physics, or other), (b) whether they favored, did not favor, or
94 SOLID STATE INSURRECTION

were agnostic about the proposal, and (c) their willingness to publish in the
revised journal.
Responses were mixed, although tilted distinctly against the proposal.58
Harvard’s Harvey Brooks replied: “While I can see some virtue in a closer re-
lation between Chemical Physics and Solid State Physics, shotgun marriages
of this sort are usually not very successful,”59 and concluded that since the
interests of solid state physics cleaved more closely to the topics covered by
the Physical Review, a forced exodus in the direction of chemistry would be
inadvisable. William Shockley, on the other hand, favored the proposal, com-
menting: “Solid state physics papers are now too diffuse a component of the
Phys. Rev.”60
Voices favoring and opposing the proposal shared a concern for bound-
ary issues, but differed on how to navigate them. George E. Pake, head of
the physics department at Washington University in St. Louis, neglected to
identify himself either as a solid state or as a chemical physicist. Instead he
pointed to magnetic resonance as his primary research interest, suggesting
that it bridged the divide. In favoring the proposal, Pake maintained that
“structure of matter physics and chemical physics do not have a readily dis-
cerned boundary between them.”61 Walter Kohn, then of the Carnegie Insti-
tute, held the opposite view. In his eyes, “Solid state physics has closer ties to
other branches of physics than to chemistry and would be damaged if these
ties were weakened.”62
The difference between Pake’s view and Kohn’s fell along topical lines
and their disagreement is emblematic of a clear split within the pool of re-
actions to the proposal Seitz was able to assemble. Pake, an experimentalist
who helped develop early nuclear magnetic resonance techniques, saw appli-
cations of those techniques flow smoothly from solids to molecules, with little
practical or conceptual difference. Nuclear magnetic resonance formed what
Cyrus Mody calls an “instrumental community,” a community organized
around specific instrumental practices and committed to their instrumental
uses, wherever those uses led.63 Kohn, on the other hand, was a theorist who
had made his career up to that point in semiconductor physics. His research
wrestled with the foundational issues quantum mechanics faced when ap-
plied to complex systems, and he was therefore less inclined to think that he
had much to gain from a closer association with chemistry.64
Similarly, Brookhaven National Laboratory’s Hillard B. Huntington, a
theorist focusing on metallic lattice structures and dynamics worried about
too close an association with chemistry, responding: “I don’t believe that
THE PUBLICATION PROBLEM 95

solid state physics and chemical physics will be compatible bedfellows in


such a close union.”65 Conyers Herring, founder of Bell Laboratories’ theo-
retical physics division, was concerned that “this move would tend to widen
the gulfs between solid state physics and fundamental physics, on the one
hand, and chemistry, on the other.”66 Columbia’s Shirley L. Quimby sup-
ported the proposed merger. He articulated his instinct that those “engaged
in experimental research . . . will favor the proposed journal and patronize
it.”67 Several other self-identified chemical physicists also expressed a will-
ingness to publish alongside solid state physicists, as long as the latter did not
displace chemically oriented work.
This wide range of responses reflects the parochial interests of the respon-
dents. Spencer Weart has argued that the diverse conceptual scope of solid
state physics made it susceptible to the formation of smaller internal social
structures, each of which developed its own set of values and interests.68
Questions about how solid state, as a field, should govern its publication
practices did not simply generate opposing camps, one in favor of a closer
alliance with neighboring disciplines and one opposed. Rather, individual
subgroups fell on one side or the other of this divide on the basis of highly
local considerations. If a research program happened to enjoy a close and
mutually constructive relationship with chemistry, then its members would
be favorably disposed to publishing alongside chemists. Those representing
other groups saw little to gain from such crossover and were bemused and
alarmed at the suggestion that they should conceive of themselves as engaged
in an interdisciplinary undertaking.
The responses to Seitz’s circular expose some of the developments that,
even into the mid-1950s when the field was well established, strained the
stitching of a patchwork solid state community as research-based subgroups
formed and developed clearer perspectives on their interests. The most ve-
hement opposition to the proposal came from solid state theorists, especially
those in the influential semiconductor group, who were busy adapting the
methods of quantum mechanics to complex systems and saw little profit in
distancing their work from the core publication outlet of the physics commu-
nity. The case in favor of the proposal was carried mostly by those experimen-
talists who identified relevance for their work to both chemical and solid state
problems. Chemists and chemical physicists also lent their support, both be-
cause of the experimental connections and because techniques developed in
the context of solid state theory were relevant for theoretical chemistry.69 The
breadth of the solid state enterprise meant that the connections researchers
96 SOLID STATE INSURRECTION

drew to related fields depended strongly on the type of work in which they
were engaged. The theory/experiment division is one dimension of this ef-
fect, but topical focus was also a factor. Views about how to structure solid
state publications therefore reflected convictions about how the discipline
should be organized.
This pattern of responses raises the question of why the committee pro-
duced a proposal that was so evidently out of step with the desires of those
who self-identified as solid state physicists and felt strongly enough to re-
spond to the survey. Why did the committee, after reviewing the facts on the
ground, craft a proposal that was so widely panned? The answer can be found
by examining the emergence of one consolidated bloc within solid state that
was vocal, but not necessarily representative of the whole. As indicated above,
physicists as a whole, and solid state researchers in particular, were increas-
ingly topically diverse and industrial. Solid state physicists in industry were
collaborating regularly with chemists and engineers. At the same time, how-
ever, several cohesive research programs were developing within a field that
had hitherto been without a clear center. The proposal that solid state form
an alliance with chemistry touched a nerve with members of these groups.
Those interested in the electrical and magnetic properties of matter
showed particular resistance. Notably, this was the same area where John
Van Vleck had made his most important contributions. It had also produced
some of the most technologically relevant discoveries, such as the transis-
tor. As the area that had the most quickly and successfully adopted quantum
methods, it was also the area whose practitioners were best able to claim that
they occupied intellectual frontiers of physics. This group, as a result, was
positioned to exert considerable influence on the professionalization process.
Its representatives, like Kohn, were more inclined to see solid state as a tra-
ditional physical subfield than as an interdisciplinary synthesis and therefore
opposed too close a marriage between solid state and chemistry, metallurgy,
or engineering. A decision about how, if at all, to restructure the publishing
patterns of solid state physicists would therefore be a test of their status and
influence over the direction of the field.
Just as Seitz was getting a sense of how the journal infrastructure in the
United States would shape the future of solid state physics, he was blindsided
by Harvey Brooks, the Harvard physicist and student of John Van Vleck who
had earlier expressed skepticism about the wisdom of a closer alliance be-
tween solid state and chemistry. Seitz and the AIP had been deliberately test-
ing the waters before acting on the publication problem, and so were taken
THE PUBLICATION PROBLEM 97

aback upon learning that Brooks had, without consulting the movers and
shakers at the AIP or the APS, cofounded the International Journal of the
Physics and Chemistry of Solids in partnership with Pergamon Press. As the
title indicated, this new journal was an international effort, publishing arti-
cles in Russian, French, German, or English, and seeking to meet the per-
ceived need within the global community “to encourage greater interchange
between physicists and chemists interested in solids.”70 The foreword to the
first issue, published in September 1956, began: “The emergence of solid-
state physics as a recognized specialty of physics has taken place over a period
of many years. A more recent development, stimulated partly by the growth of
industrial interest in the field, has been the growing realization of the common
interests of physicists and chemists in the problems of solids.”71 The journal
met a clear demand within the solid state community. But that community
was already highly heterogeneous, and it was not yet clear to its leaders that
this particular constituency should govern the direction of publication within
the field.
The journal’s sudden appearance therefore preempted the AIP’s efforts
to manage the publication problem domestically. Seitz, while the survey of
the community was in progress, had initiated discussions with the Academic
Press about the possibility of founding a new journal, with an audience to be
determined by whatever needs the AIP-APS committee identified, the scope
of which would be tuned so as to siphon an appropriate publication load from
the Physical Review. Brooks, by acting outside of the powerful institutional
mechanisms the AIP and APS were erecting, limited their ability to scale their
response to the publication problem by selecting a considered topic and vol-
ume for the new journal. Brooks lent support to an interdisciplinary journal
and had thereby decided to favor a closer association with both industry and
chemistry, just as Seitz was getting a sense that this was precisely what the
most vocal constituency within solid state did not want.
Brooks was cowed when he learned that he had upset the AIP apple cart.
He avoided extending the issue even so far as his secretary, and self-typed a
long, effusively apologetic letter to Seitz describing how he had succumbed to
a hard sell from General Electric’s J. Herbert Hollomon and Kevin Maxwell,
the director of Pergamon Press’s international division. “As I think back over
the history of this matter,” Brooks wrote, “I realized that my behavior has
been somewhat inexplicable and not to my credit, and indeed in retrospect
I feel quite unhappy about my actions.” He further expressed concern “with
the fact that in this matter I have behaved with a degree of irresponsibility
98 SOLID STATE INSURRECTION

which is a matter of great regret to me, and I feel that I have not dealt fairly
or honestly with you either in your capacity as chairman of the institute, or
as representing Academic Press, or as a friend.” Nonetheless, Brooks main-
tained: “The job itself [as editor of the new journal] is a worth while one in
my opinion, and I do not mean to imply by my present regrets that I have any
hesitation in being associated with it other than the question of whether I can
do a good job.”72
Seitz replied pointedly to Brooks’s contrite missive, outlining how the
appearance of the new international publication outlet disrupted the AIP’s
ability to mount a measured response to distinctly national challenges: “Until
the character of the new journal is clearly established, it will have the effect of
pre-empting the position for any other journal that might be contemplated.
About half the interest of any new journal would be in the field yours will
cover. Another individual might hesitate to accept the editorship at this time.
I find it very hard to decide whether this is good or bad for American physics
as a whole.”73 Brooks unwittingly trammeled Seitz’s best-laid plans, but his
journal was an honest response to widespread demand. Its impact was not to
prevent the AIP from responding to the publication problem, but to render
a summary decision on how the problem would be addressed. It was a re-
sponse, although perhaps not the precise response that would have emerged
from a more deliberative process.
For better or worse, the AIP and the APS could now focus their respective
responses to the publication problem more narrowly. The new international
journal would not bear enough of the national publication output to ade-
quately address the glut, which consumed the APS in the mid-1950s to the
extent that committees on its various aspects proliferated. These included, as
of 1956, the “Standing Committee to consider such publication-problems as
Managing Editor does not accept as lyin[g] in his province,” the “Committee
to consider a proposal of National Science Foundation regarding publica-
tions in physics,” and “Committee to study all aspects of the problem of pub-
lications of American physics.” By 1957, the former two had, in the fanciful
phrasing of Karl Darrow, reached “what some nineteenth-century statesman
called a condition of innocuous desuetude.”74 Three factors led to reduced
urgency of the publication problem. The emergence of new outlets—Brooks’s
journal, along with a smattering of other privately funded physics journals—
was one. Second, as increases in page charges and subscription fees kicked
in and officials cracked down on loopholes—such as librarians joining the
APS to get its journals at member rates rather than library rates—publication
THE PUBLICATION PROBLEM 99

operations inched back toward solvency. Finally, the Physical Review gained
a pressure valve in the form of Physical Review Letters, which launched in the
middle of 1958 with funding from the National Science Foundation.
The fast-publishing journal for short pieces, previously accommodated as
letters to the editor in the Physical Review, satisfied the demand for a quick-to-
print outlet that could protect priority claims for important new research and
relieved the Physical Review of a substantial volume of contributions. In part
due to the relief Physical Review Letters provided, the Physical Review had
nearly cleared out its backlog by 1960. A relieved Samuel Goudsmit reported
to the January council meeting that his long-suffering journal was “well on the
way to catching up.”75 The shortening of the Physical Review’s turnaround
time took considerable oomph out of the pressures favoring a topical realign-
ment of the society’s publication structure. For the time being, all topics of
physics would remain aligned with the field’s flagship journal.
The upshot was that, despite a great deal of hand-wringing and the ex-
istence of several seemingly viable plans that would have given solid state
physicists new publication homes, the bulk of the American solid state com-
munity continued publishing in the core journals. A combination of small,
specialist journals springing up on their own initiative, a return to solvency
for the Physical Review, and strong opposition from a small but vocal bloc of
solid state physicists ensured that solid state would remain firmly established
as a subfield of physics and avoid any organizational commitment to the rela-
tionships it often built informally with related fields.

COMMITTING TO PHYSICS
The challenges of a crowded publication landscape did not evaporate once
solid state’s helmsmen resolved to steer safely inside the boundaries of the
physics community, but the disciplinary identity crisis did subside, for a time.
Solid state would effectively get a dedicated journal in 1970, when the Physi-
cal Review split into four separate sections, with Physical Review B dedicated
to solid state.76 By that time, solid state’s position within physics was more
stable than it had been in the mid-1950s and the subdivision of the journal,
which had become a simple necessity based on the volume of articles Physical
Review was publishing, did not raise questions about the field’s elemental
identity.
Solid state physics in the 1950s was analogous to a disorganized system
beginning to self-organize. The ecumenical spirit with which it was founded
in the 1940s left it unusually susceptible to the formation of interest groups,
100 SOLID STATE INSURRECTION

particularly those that naturally grew around research programs, and which
shared few strong intellectual connections with other such groups that also
formed under the auspices of solid state. These interest groups developed
differing visions for the future of the field. As new professional challenges
emerged during the 1950s, these groups were given the opportunity to nudge
solid state physics in a direction that better suited their goals.
The objection of one of these groups to a closer association with chem-
istry contributed to solid state’s avoiding steps that would have nudged its
publishing operations away from the rest of physics. It helped that the group
was well organized—more so than the field as a whole—and vocal. Their suc-
cess was due in part to the structure of the solid state community. The lack of
a commonly shared conceptual program meant that solid state was grouping
into smaller, weakly interacting communities built around specific research
programs. The thrust of the whole solid state confederation could be shifted
substantially if only one of these groups, or a small subset of them, chose
to speak up. In this case, the cadre of solid state physicists who had built
a cooperative network centered on the electromagnetic properties of solids
mustered an emphatic response to an active question of disciplinary policy.
Even though this group did not necessarily represent solid state physicists as
a whole, they made enough of an impression on those responsible for making
the decisions that they were able to guide the field in the direction that best
suited their own interests. Their cause was aided by the timely appearance
of several small journals that relieved some of the pressure on APS and AIP
publishing operations and reduced the impetus for sweeping changes in pub-
lishing patterns.
Two factors are particularly notable about this episode. The first is that
it resulted from a delicate series of contingencies. Seitz and his publications
committee were in a position to exercise considerable sway over how com-
munity dynamics within solid state evolved. Their research indicated a field
that, by and large, would welcome official recognition of the close association
between solid state physics and chemistry that they saw on the ground, par-
ticularly in industrial laboratories. After Harvey Brooks unwittingly threw a
wrench in the works, their power was curtailed substantially. A vocal minority
favoring inaction thereby gained additional weight.
The second is the particular character of that vocal minority. The co-
hesive group of researchers—and theorists in particular—emerging around
studies of the electromagnetic properties of solids began to resemble a tra-
ditional subfield on a small scale much more than solid state itself did, even
THE PUBLICATION PROBLEM 101

at low resolution. This group was committed to maintaining their enterprise


as a part of physics and resisted any efforts that would introduce ambiguity
about where solid state stood. This group was, in fact, the nucleus of what
would become “condensed matter physics,” the establishment of which is
explored in chapter 8. They were wary of the conventional basis on which
solid state was founded and saw the fractiousness that resulted as an obstacle
in their quest to garner wider recognition for their intellectual contributions
to physics.
In this sense, solid state physicists’ response to the publication problem
proved to be both edifying and destabilizing. On the one hand, it resolved
some lingering ambiguity about the boundaries of the terrain on which solid
state would pitch its oversized tent. On the other, it set the stage for a chal-
lenge to solid state’s conventional definition. By exerting their influence to
keep solid state within the established physics journals, the ascendant bloc
of condensed matter theorists established the groundwork for reorganizing
their activities around a well-defined family of conceptual approaches. The
publication problem, although it might have been a relatively minor challenge
when seen in the larger context of American physics in the 1950s, was a pre-
lude to future and more complete reorganizations of the research traditions
that made up solid state physics.
5

BIG SOLID STATE PHYSICS AT


THE NATIONAL MAGNET LABORATORY

It is preposterous . . . that the country’s only national facility for high


magnetic field research is hamstrung while millions are being spent
on redundant facilities in other scientific disciplines.
—BENJAMIN LAX, 1967

The National Magnet Laboratory (NML), established in 1960, was solid


state’s answer to the large-scale particle accelerators that became the defin-
itive research instruments of high energy physics. The NML, designed to
produce very high magnetic field strengths in order to study the magnetic
properties of matter, was the brainchild of Francis Bitter. A background in
both metallurgy and the quantum theory of magnetism predisposed Bitter to
a vision of solid state physics as a field dedicated to the search for fundamen-
tal knowledge. That vision would guide NML, even after Bitter’s poor health
prevented him from directing the facility he had designed, but the lab would
also face considerable pressure, both from within and without, to make its
programming more directly answerable to short-term technical needs.
The NML was born into the era of big physics. Following the Second
World War, physicists from the whole sweep of subject specialties and in-
stitutional settings had acclimated to lavish government spending. The Na-
tional Science Foundation (NSF), which dispensed its first grants in 1950,
represented a stable commitment to federal support for research. The Atomic
Energy Commission (AEC) ensured a ready source of support for nuclear
physics. And military organizations like the Department of Defense (DOD)
and the Office of Naval Research had overflowing coffers, a vague but pow-
erful conviction in the merits of opening them to scientific research, and few
BIG SOLID STATE PHYSICS AT THE NATIONAL MAGNET LABORATORY 103

guidelines for how they were permitted to dispense funding. The unprec-
edented munificence of these organizations permitted scientists to imagine
research on a new scale, and to regard it as normal.
The NML was the first large facility to support a significant focus on solid
state research, but others followed. Later in the 1960s, the AEC founded nu-
clear reactor facilities at the National Laboratories optimized to produce high
neutron flux. The High Flux Beam Reactor at Brookhaven National Labora-
tory and the High Flux Isotope Reactor at Oak Ridge National Laboratory,
which, like the NML, were multiuser facilities, supported neutron diffraction
research that became critical for the study of materials.1 Brookhaven under-
took the National Synchrotron Light Source in the 1970s, which took what
had previously been a nuisance for high-energy accelerator designers, syn-
chrotron radiation, and harnessed it to enable precise X-ray and ultraviolet
scattering experiments in a wide variety of fields, but particularly in the study
of materials.2 The NML anticipated the style of big science conducted in
both high-flux research reactors and synchrotron sources, which emphasized
service to outside users and the flexibility to adapt to the needs of various
research programs, making it an early example of trends culminating in what
Robert P. Crease and Catherine Westfall call the “new big science” of the late
twentieth century.3
But unlike high-flux reactors or synchrotron sources, which were quickly
forced to compete with similar facilities for users, the NML remained unique
for some time as a large facility dedicated to high magnetic fields. Magnetism
was among the largest interest groups within solid state physics, and it inter-
acted robustly with neighboring fields. The American Institute of Electrical
Engineers organized a series of well-attended conferences on magnetism and
magnetic materials, beginning in 1955, which the American Physical So-
ciety (APS) cosponsored and which many solid state physicists attended.4
The momentum behind magnetism research ensured robust demand for the
NML’s services and, in the eyes of its administrators, set it apart from the par-
ticle accelerators that were proliferating around the same time.
Aside from its place in the well-known story of Cold War big science,
the NML also features in the related, but less well understood story of how
tightening science budgets reshaped the research landscape in the 1960s.5 A
focus on federal spending for social programs as part of Lyndon Johnson’s
Great Society legislation, the war in Vietnam, and the growth of a more in-
tricate bureaucracy within the funding organizations conspired to make
funding scarcer and to require greater accountability from grant recipients.
104 SOLID STATE INSURRECTION

Many facilities and research programs that had grown optimistically on the
nourishment of relatively unfettered government support had to adapt, quite
abruptly, to leaner times.
Federal belt-tightening affected some areas of physics more than others.
Particle physics saw this in the 1960s with the ribbon cutting for the Alternat-
ing Gradient Synchrotron at Brookhaven National Laboratory, which would
remain the world’s most powerful accelerator until 1968. By then, high en-
ergy physics facilities had become important for sustaining the country’s in-
ternational scientific prestige. The future of high energy physics was assured
by the unflinching commitment of the AEC, along with supplementary sup-
port from NASA, the DOD, and the NSF, and it emerged from the decade
with congressional commitment to fund the National Accelerator Labora-
tory (NAL), which would become better known as Fermilab.6 Relying to a
greater extent on discretionary funding from the DOD, solid state physics
faced steeper cuts. Solid state physicists found it particularly difficult to find
support for exploratory or theoretical research, and at a time when many in
the field saw that work as critical for maintaining their intellectual standing
with respect to other subfields of physics.
Examining how a large facility dedicated principally to solid state research
responded to these pressures permits a comparison between solid state–style
big physics, and big physics as seen through particle accelerators. Both pro-
ceeded from the conviction that the new fundamental knowledge about the
physical world made possible by quantum mechanics could be accessed by
the large machines made possible by expanded federal funding. Francis Bit-
ter, along with Benjamin Lax, the NML’s first director, understood the fa-
cility as the solid state analogue of the Alternating Gradient Synchrotron at
Brookhaven National Laboratory, the Stanford Linear Accelerator, and simi-
lar high-energy facilities, and in some respects it was similar. It was driven by
the ethos of research at the extremes—the extremes of higher energy particles
in one case, and higher intensity magnetic fields in the others. Both types of
facility ostensibly existed to satisfy our elementary curiosity about the physi-
cal world. But internal tensions over the NML’s mission indicate that the pull
of applications was never far from solid state work, and that its researchers
and administrators were called to balance those missions in ways high energy
facilities were not. The financial struggles of the mid- to late 1960s are partic-
ularly revealing of how big solid state physics differed from big particle phys-
ics, and of how the pure science ideal, which lived on despite the challenges
BIG SOLID STATE PHYSICS AT THE NATIONAL MAGNET LABORATORY 105

to its dominance in the 1940s and 1950s, had to be hybridized with applied
relevance in the context of solid state research.

FRANCIS BITTER AND A VISION FOR PHYSICAL METALLURGY


In 1939, Francis Bitter was a young associate professor in MIT’s Department
of Mining and Metallurgy. His background, though, was characteristic of an
American physicist of this era—unlike many of his departmental colleagues
with backgrounds in engineering or chemistry. His commitment to physics
was secured by a predoctoral stint in Berlin in the eventful years of 1925 and
1926. Bitter recalled hearing Max Planck speak on thermodynamics, attend-
ing the colloquium at which Erwin Schrödinger introduced wave mechanics,
and teaching himself electricity and magnetism from Max Abraham’s text-
book, The Classical Theory of Electricity and Magnetism.7 Bitter earned his
PhD in physics from Columbia University in 1928. Between leaving Colum-
bia and joining MIT, he spent time at assorted and auspicious institutions. He
conducted postdoctoral research with Robert Millikan at Caltech, worked
as a research physicist for Westinghouse, and visited the Cavendish Labora-
tory on a Guggenheim Fellowship. During these appointments, his interests
evolved from his thesis work on the magnetic susceptibilities of gases to the
nature of ferromagnetism.8
Arriving at MIT fresh off his Guggenheim, Bitter was enthusiastic, and
his outlook on metallurgy bullish. He articulated a vision for metallurgy in an
unpublished document entitled “Abstract of the Present State and Possible
Developments in Physical Metallurgy.” The document articulated his hopes
that metallurgy could transcend its historical focus on classifying the specific
properties of metals and alloys and seek fundamental contributions. The vi-
sion articulated in his “Abstract” would also shape his subsequent efforts to
craft the National Magnet Laboratory’s mission. Bitter described his strategy
for molding metallurgy into fundamental science as follows:
During my brief association with the subject of metallurgy I have obtained
the impression that in this field more than any I have come into contact with,
there is now an opportunity for rapid and fundamental development through
an application of the concepts and techniques of physics and chemistry. The
achievement of such progress must come as a result of the cooperative effort of
a group of men whose chief interest it is to discover and classify the properties
of metals and alloys in all their generality with the aim of formulating physical
106 SOLID STATE INSURRECTION

laws, rather than to follow the behavior of certain special alloy systems in de-
tail with the aim of developing and understanding commercial processes.

Bitter distinguished between the engineering and the scientific components


of metallurgical research and found that through insufficient development of
the latter, the former lacked “proper help and stimulation of a fundamental
nature.”9 He described how metallurgy might position itself to make what he
deemed fundamental contributions, which involved building a robust con-
ceptual foundation rooted in physics and chemistry, plus a strategy for col-
laborating across disciplinary boundaries to borrow techniques and insights
from fields with established fundamental research programs. This grand vi-
sion for metallurgy might just as well have been a roadmap for the as-yet-
unnamed solid state physics. It called for understanding general features of
metals through theoretical physics, a focus on mechanical properties, re-
search on crystal structure, and increased understanding of phase transitions,
all of which would fall under the auspices of solid state once the field cohered
after the Second World War.
Bitter’s understanding of fundamentality was an inclusive one. Physics
had it, chemistry had it, and metallurgy could attain it by adopting the no-
bler habits of these disciplines. Bitter’s optimism for the future of metallurgy
required a two-stage process of fostering basic insights and then building a
close relationship with, while still maintaining a separation from, practical
applications. “The physicist develops the fundamental laws which the en-
gineer applies. In chemistry we have a similar situation,” he wrote, before
presenting his rhetorical call to arms, asking: “Who, in metallurgy, is doing
a similar job?”10 Bitter emphasized that constructive dialogue between ab-
stract science and its applications could generate advances in both, and that
establishing regular discourse between them was necessary to make the field
of metallurgy a fundamental science.
Bitter’s disquisition on metallurgy outlined two dimensions of fundamen-
tal research. The first was the formulation of general principles; the second,
a practical consideration, was usefulness in a wide variety of new research.
The main flaw Bitter perceived in contemporary metallurgical work was in-
sufficient emphasis on codifying regularities in the behavior of metals. The
field lacked the generalizing input of theory. The hallmarks of fundamental
disciplines, Bitter maintained, were theoretical principles that applied—and
that actually were applied—beyond narrowly defined systems. A science con-
cerned with the properties of metals and alloys, therefore, could become fun-
BIG SOLID STATE PHYSICS AT THE NATIONAL MAGNET LABORATORY 107

damental by crafting a theoretical scheme useful for describing the properties


of metals and alloys as a class of materials.
Organizing existing knowledge into a generalized scheme, though, was
not the ultimate goal of fundamental research; fundamental science also had
to prove useful in areas where knowledge was less sure-footed. Bitter empha-
sized fecundity as a marker of fundamentality. This term implies the actual,
not just the potential, generation of intellectual progeny: the general princi-
ples that satisfy Bitter’s first criterion can be deemed fundamental once they
prove their worth in a realm for which they were not specifically designed.
From this perspective, achieving fundamentality demands more than mod-
eling research on disciplines that already exhibit it; it requires building the
personal and institutional relationships through which research efforts can
exercise influence. Science becomes fundamental, in this sense, when those
relationships have provoked new research and served as a foundation for new
conclusions.
Bitter perceived this quality in physics and chemistry. Both sciences
aimed to formulate general principles, but, more important, rich interactions
between inquiry and applications drove progress in these sciences and made
them broadly applicable to fields like metallurgy and engineering. For Bitter,
the problem was not just that metallurgists wanted for a robust theory of met-
als; they were not even in dialogue with people who were working to develop
one. In the absence of such an interaction, Bitter thought, metallurgy would
be limited to cataloguing and quantifying the properties of a growing alloy
zoo. This type of research, because it did not pursue general laws or ask novel
questions, could never be fundamental. His remedy was to encourage metal-
lurgists to overcome the insularity of their field and collaborate with physi-
cists and chemists to build the bridges that would foster new thinking. Bit-
ter’s recommendations called for MIT metallurgists to change the way they
organized their department and fit within institute infrastructure, suggesting,
for example, that “one or two physicists in the metallurgy department . . .
carry out their work in close contact with the rest of the staff,” and promoting
“closer contact with [John C.] Slater’s work in Physics and with the work of
[Charles W.] MacGregor in Mechanical Engineering.”11

FOUNDING THE NATIONAL MAGNET LABORATORY


Bitter’s view of fundamental knowledge was pivotal to the founding and
growth of the National Magnet Laboratory two decades later. Bitter coordi-
nated planning for the NML, which opened in 1960. Its early history reflects
108 SOLID STATE INSURRECTION

Francis Bitter’s 1939 vision. The proposal that convinced the Air Force Of-
fice of Scientific Research to fund the lab framed its mission compatibly with
Bitter’s notion of fundamentality, its goal being “to make continuous fields
up to 250,000 gauss available for fundamental research in solid state and low
temperature physics and related fields, and to serve as a center for advancing
the art of field generation.”12 Bitter had transferred to the physics department
in 1945 and he built the NML with solid state physics, rather than metallurgy,
in mind.13
The NML would support foundational research in order to enrich the
field and make it more productive. A quote from Bitter’s laboratory dedi-
cation speech, composed long before budget concerns caused NML staff
to emphasize its practical offshoots, indicates the place he saw it occupying
within the scientific community: “The solid-state research program is being
transferred from the M.I.T. magnet laboratory to the new facility [the NML].
The aim of this program is to increase knowledge of the basic electrical, mag-
netic, optical, acoustical, and thermal properties of solids. This fundamental
information, pursued for its own sake, has and will continue to provide the
basis for the continuing development of new and improved solid-state elec-
tronic devices.”14 Concern with establishing an environment in which funda-
mental research could flourish was at the forefront of Bitter’s thinking. With
the NML, he institutionalized his convictions about fundamental research,
hoping that its structure would encourage research that could, by focusing
relentlessly on the search for general principles, serve as the basis for some-
thing more.15
Administrative responsibility was shared among the departments that
used the NML. Beyond offering a venue for research using high magnetic
fields, the facility provided an interdepartmental forum for MIT scientists
and engineers and attracted visiting researchers from other institutions, again
in accordance with Bitter’s view that fundamental work should be outward
looking. The NML was also an educational space. Bitter had admonished
the metallurgical community in 1939 that fundamental advances required
that students be trained to ask fundamental questions. In the early 1960s,
the NML promoted itself as just such an opportunity for MIT graduate stu-
dents. A 1963 brochure emphasized this aspect of the lab’s mission, tout-
ing the “opportunity to pursue extremely fundamental speculations” that
graduate students enjoyed, citing one doctoral student’s work on magnetic
field dependence of the velocity of ultrasonic waves in metals.16 By 1965,
the lab’s second full year of operation, its thirteen academic staff supported
BIG SOLID STATE PHYSICS AT THE NATIONAL MAGNET LABORATORY 109

twenty-four students, who worked alongside nine researchers from Lincoln


Laboratory—a military research lab, which MIT administered—and fifty-six
visiting scientists.17
The NML provided the space, resources, community, and pedagogical
opportunities necessary for a solid state research facility embodying Bitter’s
vision. But Bitter, who was nearing the end of his career and facing health
problems, did not take an active role in the lab’s administration. Benjamin
Lax became its first director. Lax was Hungarian-born, but had immigrated
to the United States in his youth and earned his bachelor’s degree at Cooper
Union in 1941. He was drafted while pursuing doctoral work at Brown and
arrived at MIT in 1944 as a Radiation Laboratory researcher. He stayed on
after the war, completed his PhD in 1949, and joined the Lincoln Laboratory
shortly thereafter, rising to head of the Solid State Division in 1958.
The selection of Lax as NML director represented the desire to coordi-
nate solid state research across departments. John Slater, who wielded signif-
icant influence as an Institute Professor and head of the physics department’s
solid state and molecular theory group, wrote to MIT president Julius Strat-
ton as plans for the NML were brewing: “I feel that if we seized the opportu-
nity presented by the Magnet Laboratory, if it goes through, and correlated it
with . . . work on solids in the departments of Physics, Chemistry, Electrical
Engineering, and some of that in Metallurgy, we should have the possibility
of building up a solid-state laboratory of great value not only to M.I.T. and
the educational program, but to the services and the country as a whole.”
Slater advocated grouping representatives from these departments “together
with Lax and as much of the Lincoln solid-state group as could possibly be
included, in a great cooperative organization, if possible housed together on
or close to the campus, and including . . . students, professors, and research
scientists of the Lincoln type.”18 Support from figures of Slater’s prominence
ensured that the new lab, beyond absorbing the existing magnet program Bit-
ter had established, would seek greater integration of solid state work across
MIT’s campus.
With institutional support for interdepartmental collaboration secured,
Lax carried Bitter’s ethos forward. A 1963 promotional document for the lab
asserted: “While the primary emphasis at the National Magnet Laboratory is
on the acquisition of basic knowledge concerning the structure of solids, it is
a historical fact that work of this nature has led to far reaching technological
advances.”19 Lax toed the same line in private correspondence. “I believe it is
important for us to provide in the field of basic solid state and applied phys-
110 SOLID STATE INSURRECTION

ics, centers of excellence that will contribute to education and to science in a


most effective way,” he wrote to Roman Smoluchowski in 1965.20 The argu-
ment that investment in basic research, without efforts to shape its direction
from without, was the best way to foster useful results was a standard theme
in scientific rhetoric by the 1960s, and early on it was a useful justification
for the lab.
The NML was productive. By December 1965, its thirty-eight staff mem-
bers had published sixty-one papers that calendar year, with twenty-nine
more in press, accepted, or submitted, and had collectively delivered seventy-
nine meeting or colloquium talks. These numbers more than doubled 1963
totals, far outstripping the 18.75 percent staff increase in the same interval.21
Of the sixty-one articles, book chapters, and monographs published in 1965,
more than half, thirty-seven, were published in American Institute of Physics
journals. Of these, nineteen appeared in the Physical Review or Physical Re-
view Letters, the flagship publications dedicated to basic physical research.22
These articles contributed to major contemporary solid state research trajec-
tories, for instance by examining the band structure of solids and properties
of superconductors. Although crude, these data indicate that a large propor-
tion of the lab’s output was dedicated to the type of foundational work Bitter
championed.
The facility was a strong draw for young talent within the solid state com-
munity. Lax found himself with an embarrassment of riches in the mid-1960s
and complained to the National Research Council’s Solid State Sciences
Panel that despite having “interviewed a greater fraction of first-rate young
scientists than we have throughout my entire career . . . with very few ex-
ceptions, we reluctantly turned these away.”23 Lax himself won the Oliver E.
Buckley Prize in 1960, which, although less than a decade old at that point,
was among the most prestigious accolades a solid state physicist could garner.
He would be elected to the National Academy of Sciences in 1969 largely on
the strength of his work at the NML. It was not for want of results that the Air
Force’s enthusiasm for the facility began to wane in the mid-1960s.

MID-1960S FUNDING PRESSURES


In 1965 the NML administration asked the National Science Foundation to
assume financial responsibility for visiting scientist support and a share of
both magnet maintenance and research costs from the Air Force. Funding
for the laboratory had stalled after its initial ramp-up. Widespread national
shortages in basic science funding, coinciding with the escalation of the Viet-
BIG SOLID STATE PHYSICS AT THE NATIONAL MAGNET LABORATORY 111

nam War, became a prod with which to nudge the lab toward a more explicitly
applied stance. Lax noted with some alarm the Air Force’s increased inter-
est in the applicable fruits of magnet research in 1967: “This is a complete
change from the past when NML was discouraged from including in its char-
ter an applied program.”24 Lloyd A. Wood, director of the physical sciences
division within the Air Force’s Office of Scientific Research, substantiated
this observation, writing to Lax a month later: “It is as you know becoming
more and more an issue in Washington to ‘couple’ federally supported basic
research to ‘practical’ enterprises, and a large project such as the Magnet Lab-
oratory has a great opportunity for doing this.”25
An eye on applied benefits was not incompatible with the NML’s stated
mission, which had, since Bitter’s early vision, emphasized the importance of
basic insight for technological advance. “Coupling” of basic research fund-
ing with explicitly practical considerations, however, challenged both Bitter’s
conception of fundamental research as conversant with, but independent
from, its applications and Lax’s vision for the lab. Lax resisted any reorien-
tation of the NML’s fundamental emphasis. He was happy to accommodate
an overtly applied program as long as the Air Force was willing to supply the
additional funding, but maintained: “Financially we are in no position to be-
gin such work on our own.”26 Lax also pushed to keep applied projects and
their funding insulated from the operations and basic research budgets. He
testified before Congress on the transition from Air Force to NSF funding, for
instance, that the NML “has always coupled its basic research results with the
mission-oriented agencies having the greatest interest in a particular line of
development and will continue to do so,” while qualifying that commitment
by saying, “When there is a development of special interest to an agency we
will solicit support and participation by that agency, whether it be the Air
Force or other DOD department, NASA, NIH, or the environmental agen-
cies.”27 Explicitly applied projects were fine, in Lax’s eyes, as long as they did
not divert attention or funding from the fundamental work he considered the
lab’s raison d’être.
As negotiations with the NSF continued through the mid-1960s, the
laboratory faced tight budgets and an uncertain future. The NML advisory
committee was initially agitated. In February 1966, the committee struck a
defiant tone in the face of restrictive budgets, maintaining “the strong opinion
that a moderate and orderly expansion of funding is desirable,” and further
noting: “It is discouraging and unhealthy for a Laboratory, after a vigorous
initial period of building up from zero to a viable state, to be abruptly leveled
112 SOLID STATE INSURRECTION

off by budgetary constraints, when large areas of interesting and appropriate


research remain.”28 Just over a year later, the committee was more resigned
to the difficult environment. An April 1967 report offered less resistance to
budgetary stasis, and resignedly noted: “similar budget freezes affect all solid
state physics research, if not most scientific research activities in this country
at the present time.”29
Other projects that lacked the glitter of high energy physics or the stra-
tegic immediacy of nuclear weapons did indeed face similar challenges. In
1969, the budget request from the Los Alamos Meson Physics Facility, a proj-
ect under development that the AEC classed as a medium-energy physics
project, was slashed by over two-thirds, from $15.3 million to $5 million. The
haircut prompted Louis Rosen, the project’s director, to report to Congress
the distress the project staff felt “not only because it does not have the best
effect on morale, to fluctuate so drastically during the construction of a very
complex project, but also because it does not permit us to complete the proj-
ect in an orderly, efficient, and economical way.”30
Lax, however, was not content to accept the lab’s struggles just because
they were symptomatic of larger trends. He vented his frustration to a fellow
solid state physicist, Harvard’s Nicolaas Bloembergen: “It is true that there is
a budget squeeze all throughout the country, particularly on solid state. How-
ever, as it turns out, funds have been found for other areas of physics which
are already better funded overall nationally than the solid state activities at
the universities. This, in spite of the fact that solid state constitutes by far the
largest segment of the physical society.”31 Lax was similarly candid in a May
1967 letter to National Science Foundation director Leland Haworth: “We
realize the effects of the Vietnam War and the desire of Congress to spread
federal support for research more widely. It is preposterous, however, that the
country’s only national facility for high magnetic field research is hamstrung
while millions are being spent on redundant facilities in other scientific dis-
ciplines.”32 By “redundant facilities,” Lax almost certainly had in mind the
National Accelerator Laboratory, plans for which the Atomic Energy Com-
mission had approved just months earlier. Lax, the director of a large, one-of-
a-kind solid state research facility was irked that accelerator laboratories were
proliferating while his own was being forced to curtail its programs.
High energy physics was by no means immune to the budgetary climate
of the late 1960s. Major AEC accelerator facilities were forced to reduce their
operating time in 1969 because of funding shortages, and the AEC even con-
sidered closing some lower-energy facilities. But cuts to high energy research
BIG SOLID STATE PHYSICS AT THE NATIONAL MAGNET LABORATORY 113

were substantially less steep than cuts to other areas of science, largely due
to its secure place in the AEC budget.33 Reductions in operating time and
concerns about closing smaller and lower-energy facilities did not forestall
investment at the forefront of accelerator development in the way Lax felt it
was limiting his forefront magnet facility.
In what Lax considered a serious concession, the NML’s work did shift
in a more applied direction toward the end of the 1960s. In 1968, the eigh-
teen publications by NML staff in the Journal of Applied Physics equaled the
combined total of those published in Physical Review and Physical Review
Letters.34 An April 1969 Advisory Committee report noted the addition of
a program designed to explore medical applications of magnetic fields.35 In
the early 1970s, the lab initiated more aspirational applied projects, such as
the magneplane, which endeavored to translate the NML’s know-how into
a magnet-powered railroad system. The magneplane was a particularly sore
point for Lax, who felt it epitomized the concessions the NML had made
to Vietnam-era demands for applied payouts from solid state facilities pur-
suing basic research. On more than one occasion, NML research scien-
tist Henry Kolm, who headed the project, clashed with Lax over the lab’s
mission (figure 5.1).
In a memo to Lax entitled “Magnetism Applications Projects,” Kolm de-
scribed himself as “the only strong-minded SOB who has survived in your
entourage,” and voiced his frustration with Lax’s disapproval of Kolm’s ap-
plied interests: “Our magnetism applications programs are not a concession
to expediency, an act of prostitution in the bleak years of 69 to 71. They are
a long-neglected obligation of the scientific community. They are giving new
relevance to our graduate education, revitalizing our professional stature, and
improving the survival chances of the laboratory, of MIT, and of the entire
scientific establishment.” Kolm objected to Lax’s contention that the raft
of applied projects the lab had acquired siphoned funds from its mission-
critical research. In a stark indication of the depth of their disagreement over
the NML’s mission and direction, Kolm suggested: “If you find it impossi-
ble to integrate a significant applications program into the ‘core’ work of the
laboratory in such a way that you and others do not resent its existence, then
serious consideration should be give[n] to severing it administratively . . . by
creating a new laboratory.”36
No such schism was forthcoming, but the tensions between Lax and Kolm
reveal the extent to which changes in federal science policy challenged the
NML’s mission, with help from within. Lax, who administered in accordance
114 SOLID STATE INSURRECTION

Figure 5.1. Henry Kolm, Pyotr Kapitsa, and Benjamin Lax, 1970s. Lax (right) and
Kolm (left) converse with the visiting Soviet low-temperature physicist Kapitsa, with
whom Francis Bitter had worked at the Cavendish during his time in Cambridge.
Credit: MIT Archives and Special Collections, NMLR, box 2, folder 12. Reproduced
with permission

with Bitter’s vision of fundamental research, was shaken by the need to take
on applied projects for their own sake. Kolm, representing a younger gen-
eration, was less ideologically opposed to adding applied objectives to the
laboratory’s mission. Unlike large high energy physics facilities, which were
driven by the unanimity of purpose required to construct a large facility di-
rected at a single theoretical program, the NML faced pressures on its basic
research mission on two fronts. The solid state community remained faction-
alized, and those who preferred the field to be responsive to technological
possibilities and needs fought to have their own vision reflected in its large
facilities.

CONTRASTING THE NATIONAL MAGNET LABORATORY


AND THE NATIONAL ACCELERATOR LABORATORY
By 1971 the NML—renamed the Francis Bitter National Magnet Laboratory
in November 1967, following Bitter’s death—had found more stable, if not
more generous, financial support from the NSF. Its struggles through the late
BIG SOLID STATE PHYSICS AT THE NATIONAL MAGNET LABORATORY 115

1960s and early 1970s are telling: Francis Bitter’s view of fundamentality,
although realizable in a major research laboratory, did not fare as well in the
larger funding environment. The Air Force, its enthusiasm for facilities de-
voted to nonapplied work dwindling, began transferring responsibility to a
civilian agency, limiting the lab’s expansion, well before the 1970 and 1973
Mansfield Amendments, which required Department of Defense–funded re-
search to connect clearly with short-term objectives, compelled such a trans-
fer on a larger scale.37
As the NML struggled, high-energy particle accelerators thrived, even if
some smaller particle physics facilities faced the prospect of cuts and closure.
Unlike solid state physicists, who could not justify a large facility without
invoking potential practical outcomes, particle physicists reaped large-scale
expenditures based on the promise of fundamental knowledge. Wolfgang
Panofsky could claim before Congress in 1964 that “no scientist can point
a finger at this time to the specific way in which the study of high-energy
physics can and will affect our immediate environment, our health and safety,
our productivity, or any human affairs” without jeopardizing funding for the
Stanford Linear Accelerator.38 Similarly, the future NAL director Robert R.
Wilson famously told the congressional Joint Committee on Atomic Energy
in 1969 that the proposed accelerator “has nothing to do directly with de-
fending our country except to make it worth defending,” and insisted that
the search for fundamental physical knowledge provided the same culturally
ennobling qualities as art and literature.39
This response of Wilson’s, to a question from John O. Pastore, a Demo-
cratic senator from Rhode Island, is often quoted to indicate the commitment
of high energy physics to fundamental knowledge and the disdain for milita-
rism characteristic of the field’s culture, and to illustrate a principled stand
against the encroachment of defense interest into basic physics—suggesting
that high energy physics was under fierce assault by both military and bud-
get hawks. But Pastore, a strong supporter of the NAL, asked about defense
only as an afterthought to a larger discussion about the rhetoric of justifying
hundreds of millions of dollars’ worth of expenditures on aspirational physics
programs while the country wrestled with poverty, homelessness, and hunger.
“I want to get these Congressmen off my back,” he remarked later in the hear-
ing, referring to some resistance to the project from within the appropriations
committee, on which he sat.40 Wilson’s famous quote followed an exchange
between Pastore and Paul W. McDaniel, director of the AEC’s research di-
vision, in which Pastore pressed McDaniel to clarify the merits of the NAL:
116 SOLID STATE INSURRECTION

Senator PASTORE. For the purpose of the record, are you prepared to say,
or is this a fair question, what you expect to find through the 200 Bev
[accelerator]?

Dr. McDANIEL. A better understanding of the subnuclear universe, is my


general answer.
My specific answer is we do not know what we will find, but we know
there is a wealth of information there which needs to be developed.

Senator PASTORE. And with all these other priorities of hunger, under-
feeding, underclothing, and underhousing, how do you justify $250 mil-
lion at this time for building something with which we don’t know what
we are going to find?

Dr. McDANIEL. I would simply say as an individual that is a small amount


in comparison to the total capacity of this country to feed the hungry and
to clothe the naked.

Senator PASTORE. I don’t like that answer, at all.


These are the arguments we get when we go before the Appropriations
Committee, and it is usually my responsibility to carry the ball on this. I
would like to have some definitive answers here as to priorities, because
that is going to be thrown right into my face.
Here we are. We have these Senators going all over the District of Co-
lumbia. It has been on the front pages. They are going all over the country
showing how many people are starving, how many people are hungry, how
many people live in rat-ridden houses.
Here we are, asking for $250 million to build a machine that is an ex-
perimental machine, in fundamental high energy physics, and we cannot
be told exactly what we are trying to find out through that machine.41

It was at this juncture that Wilson jumped in, citing both the cultural impor-
tance of the search for fundamental physical principles and the social benefits
that had come from nuclear power as reasons to fund forefront research in
high energy physics.
Wilson moved the committee, which was favorably disposed to his cause.
Chester E. Holifield, Democratic representative from California and chair of
the committee, gushed, “As I listened to your eloquent appeal for this, my
mind went back before the days of Enrico Fermi to a time when St. Paul stood
before King Agrippa, and King Agrippa said to St. Paul that he wanted him
BIG SOLID STATE PHYSICS AT THE NATIONAL MAGNET LABORATORY 117

to explain his belief in the Christian principles. St. Paul was so eloquent that
when he got through, King Agrippa said, ‘Almost thou persuadest me to be a
Christian.’ I am saying that, leaving out the “almost.” I am saying, ‘Thou hast
persuadest me to support this to the best of my ability.’” Pastore, again taking
the pragmatic line, quipped, “I was not worrying about Agrippa. I was a little
worried about the taxpayers a-griping,” but reaffirmed his commitment to the
project.42
Wilson’s long-term justification for the project mirrored the rhetoric Fran-
cis Bitter and Benjamin Lax had used on behalf of the NML. He reassured
the committee of his “firm expectation that technological developments will
come. Directly, but after a very long time; from the results of the research
will come new technology.”43 But whereas high energy physicists in the late
1960s could get away with reaching for the distant promise of applicable out-
comes, and providing only vague accounts of the expected intellectual out-
puts from historically large laboratories, other areas of science were pressed
to articulate more direct and immediate relevance and, if they wanted to jus-
tify nonapplied work, had to fight to keep more proximate applications at
arm’s length. For solid state physics, that meant increased pressure to tighten
its connection with technological development, even in sui generis facilities
constructed on the big science model.
By the late 1960s, science of all stripes was being asked to be more re-
sponsive to social demands. The fact that Fermilab, despite its self-confessed
remoteness from both military and social concerns, could still thrive in this
environment indicates the privileged place high energy physics had managed
to secure. High energy physicists, ambivalent over military and economic jus-
tifications for their research in the face of 1960s protest movements, widely
embraced the high-minded rhetoric of fundamentality that Wilson’s congres-
sional testimony epitomized. They were successful in casting particle physics
as “a grand cultural enterprise, elegant and profound, that deserved the sup-
port of society.”44
That avenue to funding, especially large-scale government funding, was
not available to solid state physicists, who by the 1960s were already too
closely associated with technology in the imaginations of federal funders.
The pure science vision around which Bitter had built the NML failed to fill
the lab’s coffers. Five-year plans and annual reports to the Air Force stress-
ing the NML’s fundamental contributions could not replicate the rhetorical
success of Wilson’s eloquence on behalf of the NAL. Faced with this failure,
the NML’s solid state physicists felt slighted by the comparatively more se-
118 SOLID STATE INSURRECTION

vere funding shortfalls they suffered and resented the emerging perception
in the particle physics community that fundamental knowledge could only
be derived from the ultimate constituents of matter and energy. These frus-
trations extended beyond the walls of the NML. At the dawn of the 1970s,
solid state physics was a mature discipline, confident in its ability to generate
fundamental scientific knowledge, but it was in the midst of an identity crisis
exacerbated by unfavorable comparisons to its more lauded siblings.
6

SOLID STATE AND MATERIALS SCIENCE

Departmental allegiances and power drives do not easily go in solution


and crystallize out a new university pattern of interdepartmental
cooperation.
—ARTHUR R. VON HIPPEL, 1969

In 1970, Albert M. Clogston, the director of the Physical Research Labo-


ratory at Bell Labs, became chair of the American Physical Society’s newly
formed Committee on Problems of Physics and Society. The petition that led
to this committee reached the American Physical Society (APS) council in
1969, motivated by a sense that the rapid growth and specialization of Amer-
ican physics had narrowed the focus of many physicists and made it more
difficult for the APS and its members to consider the many and wide-ranging
connections between physical research and social problems and processes.1
One of its first tasks was a report on the economic concerns of physicists,
which Clogston authored.
“Five years ago it would have seemed incredible that in 1970 American
physicists would be seriously concerned about the economic well-being of
their profession,” Clogston began his essay, introducing a clear-eyed dis-
cussion of how the funding shifts described in the previous chapter trans-
lated into concrete challenges for American physicists. New PhDs struggled
to land job offers. Many found themselves in temporary positions, without
the promise of transitioning into permanent lines. Industry, previously ea-
ger to enlist physicists in development efforts, had soured on the promise
of basic research to enhance the bottom line and was turning increasingly to
engineers.
120 SOLID STATE INSURRECTION

These circumstances revived the old identity concerns of the 1940s,


for solid state physicists in particular. To address the economic problems
brought on by funding shortfalls, Clogston maintained, “the physics commu-
nity needs to redefine in a general way what are the unique characteristics of
a physicist.” Unlike the 1940s, however, when identity concerns were marked
by ambivalence and hesitancy about how closely technological applications
should be linked to the core mission of physics, Clogston worried in 1970
that the link was not clear enough to ensure the economic well-being of the
discipline.2
Clogston had reason for concern. By the late 1960s, the luster that the
Manhattan Project had given to basic physical research had faded and fed-
eral funders were much more hardheaded about how they distributed their
largesse. The rollback of government and especially military funds reflected
growing skepticism about the return on investment from funding basic sci-
entific research. Project Hindsight, a Department of Defense (DOD) review
panel begun in 1963, assayed the efficacy of DOD-sponsored research pro-
grams. The first interim report, released in 1966, foreshadowed the final re-
port’s conclusions, observing that the “contribution from recent undirected
science to the systems we have studied appears to have been small,” and en-
couraged “alternative practices in the management of scientific research” on
the basis of the conclusion that “the length of time to utilization of scientific
findings is decreased when the scientist is working in areas related to the
problems of his sponsor.”3 The first Mansfield Amendment of 1970 was con-
sonant with the conclusions of Project Hindsight, requiring DOD-funded
research to have direct defense applications.
What could physicists offer in this environment and how could they offer
it? One answer came in the form of the new, interdisciplinary field of materi-
als science.4 In some critical respects, solid state physics can be understood
as a historical antecedent to materials science. First, solid state physics, along
with engineering, chemistry, and metallurgy, was one of the ingredients of the
disciplinary stew that became materials science. Second, like the solid state of
matter, “materials” is a category too diffuse to offer much in the way of con-
ceptual consistency. And like solid state physics, materials science came into
being in response to a contingent set of factors that had little to do with the
conceptual development of the research programs composing it. In the case
of solid state physics, those factors had to do with the professional challenges
of a growing physics community. In the case of materials science, they had to
do with the strategic, economic, and military realities of the Cold War. The
SOLID STATE AND MATERIALS SCIENCE 121

growth of materials science is therefore relevant to understand the influence


of solid state physics on the rest of Cold War science, both because the under-
standing of material systems solid state physics offered was deemed impor-
tant for meeting the development objectives toward which material science
was oriented, and because materials science followed in solid state’s footsteps
by creating a new field on the basis of local, contingent, and nonconceptual
objectives.
This chapter follows the establishment and growth of materials science,
and solid state’s place within it, in order to understand the evolution of ven-
ues and audiences for solid state physicists’ technical aspirations. Becoming
a discipline within physics required changing some long- and dearly held as-
sumptions about the identity of physics, but it also required buying into those
assumptions to a limited extent. As a subfield of physics, therefore, solid state
was not the appropriate venue for a single-minded program of technical
development—the DOD had determined as much with its less than sanguine
assessment of the practical payoffs of basic physical research. These circum-
stances were favorable for the emergence of a new forum in which solid state
physics might pursue technical development in collaboration with other
disciplines.
Materials science provided just such a venue. In the 1950s, many fed-
eral funders became convinced that neither empirically driven engineering
efforts, nor principle-based physical investigations were adequate to address
the technical challenges of Cold War development. The rhetoric that these
two components needed to be linked developed over the course of the 1950s
within the burgeoning federal advisory system. It was within committees of
the National Academy of Science, the Office of Naval Research, and others,
that the argument was developed for putting solid state physics into conver-
sation with “materials research,” which, through the late 1950s, was largely
an engineering specialty.
That argument would be operationalized most forcefully by the Advanced
Research Projects Agency (ARPA), a military research and development or-
ganization that did the most to define materials science by founding a series
of interdisciplinary laboratories on American campuses, which effected the
collaboration between these branches of science and engineering by phys-
ically colocating otherwise scattered university research groups. By 1975,
when the Committee on the Survey of Materials Science and Engineering
(COSMAT) published its sprawling report on the status and direction of
materials science, Materials and Man’s Needs, the fates of solid state phys-
122 SOLID STATE INSURRECTION

ics and materials science were intertwined, with the consequence that ma-
terials science had become a thoroughly interdisciplinary exercise, and that
solid state physics had found a further way to realize Smoluchowski’s vision
of it as an outward-looking, collaborative field attuned to its technological
potential.

MATERIALS RESEARCH BEFORE MATERIALS SCIENCE


“Materials” became a fixation of the US federal advisory system in the 1950s.
As distinguished from the more generic term “matter,” “materials” refers to
the stuff of technological development, with properties that answer some
proximate need. Those needs were many in the early Cold War. Air and
watercraft, communications apparatus, energy and weapons systems, and
consumer technologies all faced material constraints, and so research into
new and improved materials presented strategic opportunities. It was not
immediately evident to those pursuing these strategic goals, however, how
physics might be relevant to these ends.
In 1951, the National Research Council (NRC) formed a Materials Ad-
visory Board (MAB) to evaluate how advances in materials research might
address strategic bottlenecks, particularly in military development. MAB
grew from the older Minerals and Metals Advisory Board. The name change
reflected “recognition of the interrelations of the metals and nonmetals, par-
ticularly in structural applications.” But despite the broadened scope of the
new committee, MAB’s early 1950s iteration had little contact with physi-
cists. A 1954 report described the change by noting: “The Board has been
reconstituted to include materials engineers, chemists, and metallurgists
. . . to provide advisory services to the Office of the Assistant Secretary of
Defense for Research and Development and to the Administrator of the Gen-
eral Services Administration.”5 Other early uses of “materials research” are
similarly engineering-centric.
Military research organizations followed the NRC’s lead, cementing mate-
rials research as a prominent target for Cold War research and development.
The restriction to engineering began to erode toward the end of the decade as
MAB honed its mission, installed physicists in influential roles, and embraced
the relevance of basic research in a limited fashion. A 1957–58 NRC report
noted that “increased attention to materials brought about by the needs of
weapons system development has resulted in a considerable expansion of ac-
tivity for the Materials Advisory Board.”6 The expansion referenced here did
not just indicate new personnel, but also voices from new disciplinary camps.
SOLID STATE AND MATERIALS SCIENCE 123

The inclusion of physicists, most of whom hailed from the newly vibrant field
of solid state physics, coincided with a marked shift in the emphasis of mate-
rials research.
MAB’s expanded topical breadth is evident in its 1960 report, “Funda-
mental Aspects of Materials Research.” The committee included Cornell
physicist James Krumhansl as deputy chair. The presence of a physicist
among the metallurgists, chemists, and industry mavens who previously com-
posed the committee indicates MAB’s emerging preference for close connec-
tions between basic research and its applications, a position that was overt
and urgent by 1960. The committee criticized the Department of Defense’s
existing efforts to mobilize basic research to strategic ends, remarking that
“in-house basic research capability is grossly inadequate.” The committee
urged the research arms of the army, navy, and air force to sponsor “strong
centralized laboratories in which basic research, comprising the entire spec-
trum of potentially pertinent science including the materials sciences, can be
promoted.”7
These recommendations, designed to enhance “the ability to bring
knowledge to bear on the defense needs of the nation in the shortest possi-
ble time,” came to define the mission of materials science as it was imagined
within the federal advisory infrastructure.8 The concept of basic research was
appropriated in service of technological defense needs, which were not be-
ing addressed with alacrity sufficient to appease the Department of Defense
and its army of advisers. Responding to pressure to mobilize basic research
resources in order to accelerate blackboard-to-battlefield turnaround, the
NRC broadened its conception of materials science still further. In 1960, a
committee to consider the “Scope and Conduct of Materials Research” was
formed “to view the total materials research needs of the country with rela-
tion both to national defense and the public welfare more generally; to ap-
praise the adequacy of present research programs to meet those needs; to
consider the resources of personnel, facilities, and administration that are
available; and to make recommendations for the correction of deficiencies
that the Committee may identify.”9
Alongside the regular complement of engineers, chemists, and industrial-
ists, this committee included solid state physicist Frederick Seitz and metal-
lurgist Cyril Stanley Smith, the director of the University of Chicago’s Insti-
tute for the Study of Metals, a prominent center of solid state research. The
report’s recommendations reflected a closer integration between science and
engineering. It advocated centralized funding, coordination, and oversight of
124 SOLID STATE INSURRECTION

materials research as well as “strengthening the universities in their dual role


of training scientists and engineers and also doing basic research.”10 By 1960,
scientists advising the federal government regularly advocated mechanisms
to increase dialogue between those studying the properties of materials and
those implementing that knowledge in strategically relevant ways.
The advisory emphasis on materials, in particular with respect to training
and basic research, exerted its influence within American universities. The
first major textbook for materials scientists and engineers appeared in 1959:
Lawrence Van Vlack’s Elements of Materials Science aimed to synthesize
traditional engineering approaches with basic science. Van Vlack informed
his readers: “The subject matter taught in Engineering Materials courses is
changing rapidly. Formerly, this subject was taught on an empirical basis. Now,
although the science of materials is far from complete, it can be approached
from a more scientific viewpoint, because of the development of principles
which relate the properties and behavior of many materials to their struc-
tures and environments.” Nonetheless, the volume remained focused toward
the needs of engineers: “This introductory text . . . is designed for freshman
and sophomore engineering students with a background in general physics
and chemistry; it does not use the rigorous approach which is common in
solid state physics books.”11 True to Van Vlack’s description, the textbook is
light on formalism, opting to deliver content through prose, pictures, and dia-
grams, appealing to visually and mechanically oriented engineering students.
As both the trajectory of MAB and Van Vlack’s textbook indicate, use of
the term “materials science” to identify a research area, originally restricted
to engineering, began to expand by the late 1950s. MAB came to advocate
for the desirability of increased contributions from basic science to address
strategic technical goals, reflecting the federal government’s emphasis on
the challenges of material constraints in the military and space programs
that would be amplified after the Soviet Union launched Sputnik 1 in 1957.
But materials research was still firmly ensconced in an engineering tradition
through the end of the 1950s, a state of affairs that would change in the early
to mid-1960s, when materials science emerged as a sui generis interdisciplin-
ary experiment.

THE ADVANCED RESEARCH PROJECTS AGENCY AND


THE INTERDISCIPLINARY LABORATORIES
The Advanced Research Projects Agency, founded in 1958, hastened the
transition toward an interdisciplinary definition of materials science.12 One
SOLID STATE AND MATERIALS SCIENCE 125

of ARPA’s first large-scale funding initiatives supported a series of university-


hosted interdisciplinary laboratories (IDLs) dedicated to the study of ma-
terials. The IDLs prompted universities to collapse departmental divisions
within the context of materials science, creating sites where students could be
acculturated to think broadly about problems related to materials and their
limiting effect on technological development. Within these IDLs, the techni-
cal, scientific, and administrative character of materials science began to take
shape.
ARPA’s call for proposals reflected the 1950s-era advisory emphasis on
materials as a bottleneck for strategic development. “The Government,” it
insisted, “has a vital stake in the establishment of the best possible materials
research and development program. This is true because materials are a limit-
ing factor in the performance of the advanced systems and devices essential to
the operations and missions of Government agencies and departments.” The
next paragraph indicated ARPA’s intent to expand upon the materials science
concept: “In order to strengthen basic research in materials sciences . . . the
Government decided to support the establishment of a number of interdis-
ciplinary materials research laboratories in universities. The objective of this
Interdisciplinary Laboratory Program is to expand the national program of
basic research and training in the materials sciences.”13
Given ARPA’s emphasis on attacking technological limitations by training
students, it is notable that the agency chose to promote the development of
a new interdisciplinary field, rather than to support efforts in existing disci-
plines, such as solid state physics, which already maintained a similar balance
between basic and applied aims. Materials science grew from the same type of
synthesis between research on metals and nonmetals as solid state, but solid
state, which staunchly maintained its physics bona fides, was not serving
defense needs as ARPA saw them. Materials science, as it coalesced within
ARPA’s IDLs did, however, mimic the strategy solid state had pioneered
of organizing a new discipline to address contingent contemporary needs.
These needs were professional in the case of solid state and technological in
the case of materials science, but they similarly sacrificed close cohesion to
other ends.
Twelve universities won IDL contracts between 1960 and 1962. The
first three were hosted at Cornell University, the University of Pennsylvania,
and Northwestern University, with the remainder appearing in quick succes-
sion at the University of Chicago, Brown University, Harvard University, the
University of Maryland, the Massachusetts Institute of Technology, the Uni-
126 SOLID STATE INSURRECTION

versity of North Carolina, Purdue University, Stanford University, and the


University of Illinois–Urbana.14 ARPA funds prompted these institutions to
consolidate their materials research efforts, which were often scattered across
several departments and distant campus locations, in centralized “Materials
Science Centers,” generating early examples of the “center model” described
by Cyrus Mody and Hyungsub Choi.15 The Massachusetts Institute of  Tech-
nology provides an apt case study. John C. Slater spearheaded MIT’s IDL
application. Slater, like John Van Vleck, earned his PhD at Harvard under
Edwin Kemble in the early 1920s and represented the first generation of
domestically trained quantum theorists.16 By the 1950s, he was an Institute
Professor of Physics—the first at MIT to be granted this honorary title. He
devoted the bulk of his time to a research program in solid state and molecu-
lar theory, which used the latest digital computers to attempt calculations of
the properties of solids and molecules from first principles.
Slater maintained a strong commitment to a research program rooted in
the physics department. Four of the five faculty members affiliated with the
solid state and molecular theory group were physicists—Slater, László Tisza,
George F. Koster, and Michael P. Barnett—and one, Walter R. Thorson, was
a chemist. Slater’s ab initio approach used the most recent digital computing
technology to calculate wave functions for solids and molecules with as few
simplifications as possible. Known as “Slater physics” around the institute
campus, this approach proved too abstract for the teaching needs of MIT’s
engineering department, which brought Mildred Dresselhaus from Lincoln
Laboratory to develop a course in solid state theory for engineers in 1967.17
Slater was nevertheless in tune with the collaborative, interdepartmental
efforts that characterized MIT’s research culture in the 1950s. In his cor-
respondence with ARPA administrator John F. Kincaid, Slater identified
“Aeronautics and Astronautics, Chemical Engineering, Chemistry, Civil En-
gineering, Electrical Engineering, Mechanical Engineering, Metallurgy, Naval
Architecture, and Physics” as the departments in which materials research
was conducted.18 Slater, conscious of the emphasis on interdisciplinary col-
laboration within advisory circles, took pains to emphasize this aspect of
MIT’s existing research programs.
MIT’s initial proposal to ARPA was titled “The Interdisciplinary Nature
of M.I.T. Research,” and asserted that “[o]ne of the fundamental features of
our proposal relates to the way in which we expect the various disciplines to
cooperate in the research.”19 The proposal leaned on the record MIT had
established promoting connections between departments, emphasizing in
SOLID STATE AND MATERIALS SCIENCE 127

particular the Laboratory for Insulation Research (LIR), founded in 1940,


and the Research Laboratory of Electronics (RLE), founded after the Second
World War to preserve facilities, equipment, and programs associated with
wartime radar research.20 Slater was equally attuned to ARPA’s interest in
graduate training, reassuring Kincaid: “We feel that establishment of inter-
disciplinary laboratories would be the best way to encourage expansion in
graduate training and research.”21
The neat correspondence between Slater’s rhetoric and ARPA ideals
might be read as cynical kowtowing to a funder’s demands if not for the simi-
lar initiatives that had been under way at the institute prior to ARPA’s call for
proposals. In October of 1958, Slater circulated a memo reporting “a good
deal of discussion of the desirability of some mechanism for getting closer
liaison between persons in various departments of the Institute interested in
solid-state and molecular science,” and indicating broad support from “mem-
bers of the chemistry, electrical engineering, mechanical engineering, metal-
lurgy, and physics departments.”22
The commitment to cross-department collaboration at MIT is evident as
far back as the early 1940s, when the institute began mobilizing its resources
for war work. Arthur von Hippel, the force behind the Laboratory for Insula-
tion Research commented in 1942:
There is no real boundary between physics and electrical engineering. Our
field is a branch of applied physics mainly concerned still with the applica-
tions of Maxwell’s theory. While the physicist stood “clean” of such useful
tasks and strove for insight, the electrical engineer built a new economy and
talked in a new technical language appropriate for his tools. Thus the link
between the two was wearing thin, until events forced both sides into closer
co-operation. The physicist began to toss into the domain of the electrical
engineer new instruments, such as thermionic tubes and photocells, rectifi-
ers, thermistors, and fluorescent lamps, which could not be understood on
the old classical basis. And the electrical engineer replied in kind with magic
eyes, complex impedance bridges, high-frequency generators, high-voltage
machines, and magnets for cyclotrons, which revolutionized the experimental
technique of the physicist.

In consequence, according to von Hippel, “The fence between the two fields
[physics and electrical engineering] is falling into disrepair.”23 The LIR was
von Hippel’s prime example of MIT’s conviction that departmental divides
could impede progress.
128 SOLID STATE INSURRECTION

Following in that tradition, MIT scientists had been itching for a more
consolidated materials program since the pre-ARPA 1950s, when a proposal
for an expanded program of materials research was compiled at the behest of
the Atomic Energy Commission in 1956.24 The proposal recorded a total of
88,020 square feet distributed over nine departments, serving 87 academic
staff and 419 support personnel. It called for consolidating these efforts in a
100,000-square-foot Materials Research Laboratory, which would promote a
“more fundamental approach” to material limitations on development.25 By
the time of the ARPA proposal, the estimate for the size needed to accommo-
date campus-wide materials science efforts had more than tripled to 350,000
square feet, indicating both the growth of materials research at MIT and the
expansion of its scope in the intervening years.26
At MIT and elsewhere, ARPA provided the infrastructure to make con-
solidation feasible on a large scale. Slater wrote to Kincaid with a frank as-
sessment of MIT’s physical constraints: “At present all of our work in mate-
rials research is very crowded, and we could hardly expand at all in number
of students in some parts of the field without providing additional building
space. To accommodate all the work in the field, on the scale on which we
should like to operate, would require a building of approximately 350,000
square feet gross floor space. This would cost something like $14,000,000
to build.”27 When APRA funded an IDL at MIT in 1961, it included a
200,000-square-foot building.28 Although this fell short of full consolidation,
it provided the institute with a crucible large enough to cook up a stable in-
terdisciplinary field.
The physical spaces ARPA provided were disciplinary laboratories as
well as materials research laboratories.29 They hosted a nationwide exper-
iment in interdisciplinary collaboration, out of which the field of materials
science slowly emerged. A memorandum sent from ARPA to its IDLs in 1962
described the terms of the experiment: “As you know, we have undertaken
the responsibility of initiating a program in the national interest with uni-
versities for ‘basic research and graduate education’ in a somewhat loosely
defined area called material sciences. You have to a great extent defined what
is meant—at least in your university—by material sciences by listing in your
proposals to us the names of individuals you believe to be the core of the
program at your institution. The collective research interests of these individ-
uals defines in more detail material sciences.”30 For ARPA, defining materials
science was an empirical question, with the caveat that the goals of the field
were established in advance. “Materials sciences” referred to the collabora-
SOLID STATE AND MATERIALS SCIENCE 129

tions that proved productive according to the standards determined by the


IDL program’s objectives.
The inclusion of solid state physics in this disciplinary experiment un-
derwrote ARPA’s assertion that it was concerned with basic research. By
maintaining its identity within physics, solid state retained a claim, however
contentious it would become, to being a fundamental scientific discipline.
That claim made it useful to materials science, which tried in part to set down
a smooth stretch of pavement traversing the arduous path between theoretical
insight and practical mastery. ARPA’s long-term success promulgating this
picture is evident in a 1987 survey text, which represented solid state physics
as “at the core of nearly all fields, while Materials Science embraces the wider
application of the basic studies.”31
The forging of materials science within ARPA’s IDLs represented the
large-scale adoption of solid state’s heterodox approach to defining profes-
sional categories. ARPA started with a set of concrete objectives: to consol-
idate research on materials from a range of disciplinary standpoints and to
train students within this new synthesis. It then promoted the formation of
a field to address those objectives, showing no regard for whether or not the
field that emerged obeyed traditional boundaries. In doing so, it created a
space in which a portion of the solid state community could make a home and
cement the case for its utility. At the same time, it exacerbated the underlying
tensions between those who championed solid state’s technical relevance and
those who sought to position it as a fundamental field of physics.
On one hand, the alliance with materials science represented the poten-
tial of solid state physics to secure reliable, federal support by reinforcing its
relevance to the problems of the age. On the other hand, it represented the
danger of straying from physics, and thus losing a claim to the intellectual
mission that, despite the progress solid state had made in making physics as a
whole more responsive to its social milieu, continued to sit at the heart of the
American physicists’ identity.

MATERIALS AND MAN’S NEEDS


An often-repeated witticism of uncertain provenance insists that “anything
with science in its name isn’t.”32 Materials science is in some respects a con-
firming instance. The treatment of the subject in Materials and Man’s Needs,
a National Academy of Science (NAS) panel report, illustrates how. Also
known as the COSMAT (for Committee on the Survey of Materials Science
and Engineering) report, it opened with a covering letter from Melvin Calvin,
130 SOLID STATE INSURRECTION

chair of the NAS Committee on Science and Public Policy, to NAS president
Philip Handler. Calvin pointed out that, unlike NAS reports on other disci-
plines, which emphasized the potential for scientific advance in those fields,
the COSMAT report would focus on needs, and how to meet them. Materials,
as a category, were defined by their usefulness, so any field oriented around
them would have an essential engineering component.33
But even if materials science could not be properly understood as a sci-
ence, it became gradually more scientistic through the 1960s and early 1970s.
As Cyril Stanley Smith remarked in 1968, “Even in the field where he was
once supreme because he alone could make or build, the engineer is currently
losing status to the scientist.”34 The growth of “materials science,” in contrast
to “materials research,” did signal a meaningful change, namely the growing
prevalence of solid state physicists and chemists—who maintained their iden-
tity as such, even while contributing to materials science—in what had pre-
viously been an area dominated by metallurgists, technicians, and engineers.
Even in the mid-1970s, however, after the first of ARPA’s IDLs had been
open for over a decade, the direct practical payoff of bringing the basic sci-
ences into materials sciences was less than clear. The COSMAT report ac-
knowledged: “Despite the many impressive achievements of materials re-
search there is the awareness that only the surface of scientific capability has
been scratched. The majority of advances have historically been made via
the empirical approach. Most new materials or properties are arrived at or
discovered by cut-and-try methods—new chemical or alloy compositions
are prepared and characterized and their various properties are determined.
There are usually underlying rationales or phenomenological models to this
empirical approach but it is rare indeed for a new material or property to be
predicted from basic principles.” However, the report continued: “The prin-
cipal exception to this situation is in the area of single crystal materials, par-
ticularly those used in solid state electronics. On the other hand, techniques
and concepts of physical science are often essential for characterizing and
reproducing the properties of even empirically-invented materials. With elec-
tronic materials, due to the combined talents of chemists, metallurgists, phys-
icists and electrical engineers a degree of understanding has been achieved,
at least for the simpler crystals, so that material compositions having the de-
sired physical properties can often be prescribed beforehand.”35 By the mid-
1970s, in other words, the merits of adopting a principle-first approach to
materials development were still largely notional.
Materials and Man’s Needs, in fact, expresses considerable frustration
SOLID STATE AND MATERIALS SCIENCE 131

about the difficulty of translating theoretical solid state work into practical
payoffs, repeatedly citing the need for more research to bridge the gap be-
tween a broad scientific understanding of material structure and behavior on
one hand and the tools to predict usable material properties within reason-
able tolerances on the other. Aside from a few cases in which the properties
of simple crystals could be predicted in advance, theoretical understanding
was still too crude for solid state to offer much of a practical aid to the com-
plex and iterative process of materials development. This is not to say that
the technical relevance of solid state physics was oversold, simply that the
route from theoretical understanding to practical application was somewhat
more meandering than ARPA had supposed when it envisioned its IDLs as
high-output factories for strategic materials. Successes in such direct ap-
plication of theory to practice were few. These few narrow areas of success,
however, were enough to justify the systematic approach of encouraging in-
terdisciplinary collaboration around the nexus of materials that ARPA had
spearheaded in the early 1960s. The COSMAT panel recommended that
universities increase interdisciplinary activities in both research and educa-
tion, and that the federal government continue support for the materials re-
search centers that might host such activities.
Evaluating the COSMAT panel’s stance and recommendations reveals
that the materials science alliance had as much to do with institutional con-
venience as it did with improving research into practical needs. It had been
dreamed up by ARPA in order to provide answers to technical questions, and
it certainly did so. But it is not clear to what extent its answers were system-
atically better than those that engineers might have achieved working in com-
parative isolation, if they were given access to a similar degree of funding. Ma-
terials science did, however, offer mutually beneficial professional rewards for
both its old and new constituent disciplines. Materials engineers were able to
associate themselves with the status that scientific research had attained and
fend off criticisms of their empirical, phenomenological methods as inade-
quate for the task at hand. Scientists for their part, and solid state physics in
particular, found a steady stream of financial and institutional support at time
when funding for basic research was becoming harder to secure.

FRIENDS OF CONVENIENCE
By the 1960s, having lasted through the institutional growing pains of the
APS and a shakeup of the publishing landscape, solid state was securely
ensconced in the American physics community. In 1949, I. I. Rabi had
132 SOLID STATE INSURRECTION

suggested that divisions of the APS were only necessary for “‘peripheral’
fields.”36 By the end of the 1960s, both nuclear physics and particle physics
had divisions of their own. The early postwar order, in which a collection of
new interest groups orbited a core of physics that carried on the ideals of the
early twentieth century, and at some remove, was dissolving. Largely on the
success of solid state physics and related enterprises, the influence of the APS
had broadened to the point that the fields that carried on traditional pure
science ideals were on a par with the newer, more technically adventurous
subfields, in institutional if not absolute terms.
The alliance between physics and materials science would eventually
lead, in 1990, to the APS establishing a Division of Materials Physics, pro-
moting to divisional status a topical group that had formed in 1986. At that
point, 70 percent of the topical group’s 2,869 members were also members
of the Division of Condensed Matter Physics (the name for the Division of
Condensed Matter Physics from 1978 on; see chapter 8), but the emphases of
the two groups had become sufficiently different that the APS executive com-
mittee determined that a new division was justified.37 The acknowledgment
of the explicitly applied category of “materials” as a full-fledged topic of phys-
ics reflects a softening of attitudes toward applied physics that began within
the APS in the early 1970s. The APS council, responding to the challenging
funding situation in 1970, issued a statement that cast physicists as deeply
concerned with the consequences of their work: “The problems we face as a
nation call for more knowledge, not less; and better technology. Better tech-
nology must be based on more extensive understanding of scientific facts and
possibilities.”38 The society did demure in 1972 when offered charter mem-
bership in the Federation of Materials Societies, judging that organization’s
interests to be aligned more closely with engineering than with science, but
it authorized the Division of Solid State Physics (DSSP) to participate in the
federation as an observer.39 In 1974, the APS approved an IBM-sponsored
prize for new materials.40 In 1975, a new standing committee on the applica-
tions of physics convened for the first time, responding to a strong sense that
recognition and respect for applied physics, both from other physicists and
from the wider society, needed to be improved.41
The friendlier approach to technology, and active censuring of the snob-
bish attitude toward applied physics that the APS acknowledges was deeply
rooted in the culture of American physics, was in large measure a response to
financial pressures. The notional long-term relevance of basic research, which
had sufficed to justify liberal military spending on science in the two decades
SOLID STATE AND MATERIALS SCIENCE 133

or so after the Second World War, proved insufficient to justify continued


annual increases in research budgets. For most physicists, and especially for
solid state physicists, this meant reevaluating their notion of relevance and
exploring new avenues for making that relevance known.
Materials science emerged in opportune coincidence with these devel-
opments. Skepticism about the payoff from funding basic research encour-
aged the proliferation of the interdisciplinary approach to science funding
that ARPA had pioneered—if basic research typically bore fruit on the scale
of decades, perhaps that timeline could be foreshortened by forcing closer
contact between basic and applied fields. That attitude, combined with solid
state’s institutional security within physics, allowed solid state physicists to
participate in the materials science experiment en masse without jeopardiz-
ing their identity as physicists in a way that would have been difficult in the
1950s. The COSMAT report estimated that 2,211 PhD-holding solid state
physicists were involved in materials science. The same report estimated
those solid state physicists as constituting 13.6 percent of all PhD scientists
participating in materials science. Just over twice that number were members
of the DSSP that same year. The total membership of the APS in 1973 was
27,291.42 Materials science, in other words, constituted a substantial pro-
portion of what solid state physicists did by the early 1970s. Support from
ARPA, the NSF, and other federal agencies that targeted materials allowed
solid state physicists to increase their base of financial support after the initial
exuberance for funding science after the Second World War began to wane.43
This alignment ensured an abundance of support for the field in troubled
financial times and helped pull physics as a whole, by sheer force of numbers,
in more relevant directions.
The alliance likewise pulled materials research toward the sciences. Ma-
terials research in the early 1950s was a largely empirical engineering disci-
pline. But if the advisory apparatus was sour on stand-alone basic research,
it was equally skeptical of stand-alone engineering. Interdisciplinary collab-
oration was in vogue. Basic science enjoyed a unique epistemic cachet, even
in the face of concerns that it did not produce results fast enough and against
the background of Vietnam-era concerns that physics, especially, was morally
unmoored. The federal advisory system now saw physics (and solid state in
particular) as an essential component of technical development, albeit one
that was not living up to its potential. Welcoming solid state physics there-
fore helped transform materials research into materials science and thereby
elevate its standing. The alliance between solid state physics and materials
134 SOLID STATE INSURRECTION

science, that is, formed amid the rhetoric of technological relevance, but was
driven less by the demonstrated outcomes of collaborations between disci-
plines than it was by the professional advantage each could gain from the
others.
A strong physics presence in materials science was possible because solid
state physics had succeeded in establishing itself firmly within physics, and
had broadened the identity of American physics in doing so. The decision to
remain closely tied to the core concerns of physics paid off for both solid state
and physics as a whole as the ground shifted under their feet and the context
demanded greater responsiveness to immediate technical needs. Neverthe-
less, pure science remained a powerful regulative ideal, even for solid state
physics, and the technical turn of the late 1960s and early 1970s would spark
a backlash. The next two chapters consider how solid state physicists fought
to hold on to their identity as fundamental researchers in an environment
determined to understand their value in strictly technical terms.
7

RESPONSES TO THE
REDUCTIONIST WORLDVIEW

In general, our physical world, the world that affects us as human


beings, is a low-energy world, not a high-energy world.
—ALVIN WEINBERG, 1964

During the 1960s, the pure science ideal took on a particularly virulent form
within the high energy physics community. Reductionism emerged as the
dominant philosophy among those seeking to understand the phenomena ob-
served in cloud chambers, bubble chambers, and other new instruments that
rendered the invisible microworld visible. It held that all theoretical knowl-
edge about the world rested on, and was in some sense contained within, the
rules governing the elemental components of matter and energy; everything
else, logically speaking, boiled down to those basic laws and concepts.
Establishing a reductionist research program differentiated high energy
physics from nuclear and atomic physics, with which it shared a conceptual
ancestry. In the 1960s, accelerator physicists began to speak of two frontiers:
an energy frontier and an intensity frontier. Exploring the former required
building larger machines with higher beam energies, which facilitated prob-
ing the smallest constituents of matter. The latter called for accelerators that
crammed more particles into smaller spaces, generating more collisions and
more complex interactions, favoring experiments that explored particle dy-
namics. In the early 1960s, these twin frontiers were seen as complemen-
tary. Stanford experimentalist Wolfgang Panofsky, speaking before Congress
in support of funding for the Stanford Linear Accelerator (SLAC), a lower-
136 SOLID STATE INSURRECTION

energy, higher-intensity accelerator, could state that “the different types of


accelerators are not in competition with one another.”1 That would remain
true only as long as liberal federal funding was available to meet the demands
of both programs.
By 1968, when M. Stanley Livingstone, coinventor of the cyclotron, titled
his elementary particle physics textbook Particle Physics: The High-Energy
Frontier, that was no longer the case. “The high-energy frontier has become
the financial-support frontier,” Livingstone wrote, defending expenditures
on large-scale accelerators on the basis of “value to our society of this new
knowledge about nature” that discoveries at the high-energy frontier would
presumably yield.2 The particle physics community, by the late 1960s, had
bet their chips on high energies as their best chance to ensure ongoing federal
support and the international preeminence of American particle physics.3
The reductionist worldview justified pursuing the energy frontier at the
expense of the intensity frontier. It sought to bestow exclusive jurisdiction
over fundamental physics to particle physicists, in their pursuit of elementary
particles and the forces governing them, which they were willing to share with
cosmologists, who explored the very largest scales of space and time. It was
also motivated in part by the reaction of some nuclear physicists against the
militarization of their work, and represented their commitment to the nobil-
ity of physics and rejection of the ignoble purposes to which they felt it had
been put.4 Alongside its intellectual and social motivations, the reductionist
philosophy fulfilled a timely institutional function in the 1960s. As federal
funding for basic research tightened and solid state physics grew, particle
physicists faced the prospect of dividing a smaller pot of federal funds with
competing fields at a time when they had committed to pursuing the fron-
tier of higher collision energies, which demanded ever-larger and ever more-
expensive accelerators. Reductionism claimed both the mantle of basic sci-
ence and the pot of federal funds dedicated to it.
Not all solid state physicists were content to be typecast as seekers after
something other than fundamental insight. Responses to rising reductionism
among high energy physicists shows how, after establishing technically ori-
ented work as a legitimate province of American physics, solid state physicists
then sought to claim for their field a piece of the intellectual legacy of phys-
ics as well. Broadly speaking, reductionism elicited two types of responses.
The first denied the very premise that physical knowledge could be deemed
“fundamental” strictly on the basis of its intrinsic intellectual merits, without
first considering its relationship to other realms of knowledge. The second
RESPONSES TO THE REDUCTIONIST WORLDVIEW 137

accepted the premise that knowledge could be intrinsically fundamental but


denied the further premise that fundamental knowledge could only be had at
extreme scales.
The first of these views was championed by Alvin Weinberg, the director
of Oak Ridge National Laboratory. Although not a solid state physicist him-
self, Weinberg, as the director of a large, broadly invested national laboratory,
was sensitive to the possibility that reductionism, should it become a default
assumption of the physics community, could adversely affect research on me-
soscale phenomena and he felt responsible for speaking on its behalf. The
second view is most strongly associated with Philip W. Anderson, the Bell
Laboratories solid state theorist whose 1972 Science article “More Is Differ-
ent” advanced a philosophical position that would come to be known as emer-
gentism. Weinberg’s and Anderson’s views reveal two rhetorical strategies,
aside from the reductionism of high energy physics, that American physicists
used to navigate the funding bottleneck of the 1960s and 1970s. Weinberg’s
view echoed efforts solid state physicists had pioneered to reinvent the core
identity of physics in a way that made it more responsive to immediate social
demands. Anderson’s instead represented an attempt to bind solid state more
tightly to the traditional intellectual core of the physics community.

CRITERIA FOR SCIENTIFIC CHOICE


Alvin Weinberg earned his BA, MA, and PhD at the University of Chicago in
the 1930s—a heady decade for the school. He entered in 1931, the year Chi-
cago’s young, strong-willed president Robert Maynard Hutchins instituted
radical reforms to undergraduate education. Weinberg was among the cohort
to inaugurate Chicago’s broad-based, two-year liberal arts core known as the
“New Plan,” which consisted of a common curriculum of yearlong courses
in English composition, humanities, and physical, biological, and social sci-
ences, capped by comprehensive exams at the end of two years.5 He would
credit his comprehensive undergraduate education with conditioning him to
understand science in terms of how it intertwined with human affairs.6
Weinberg’s sensitivity to the symbiosis between science and society mo-
tivated his essay “Criteria for Scientific Choice.” Published in Minerva in
1963 and reprinted in Physics Today in March of the following year, the ar-
ticle argued provocatively that high energy physics, as a field of knowledge
comparatively removed both from other areas of science and from quotidian
human concerns, was less than deserving of the priority it was receiving in the
federal science budget.7 When he published “Criteria for Scientific Choice,”
138 SOLID STATE INSURRECTION

Weinberg was the director of Oak Ridge National Laboratory, a post he had
assumed in 1955. He had risen through the ranks at Oak Ridge after alight-
ing there following his services to the Manhattan Project at Chicago’s Met-
allurgical Laboratory. That so prominent a figure would question the merits
of supporting expensive accelerator research rankled in the particle physics
community and forced more careful and pointed articulations of its reduc-
tionist philosophy.
“Criteria for Scientific Choice” warned of tough choices ahead as fund-
ing for scientific research plateaued. As the director of Oak Ridge, Weinberg
would be charged with deciding which projects to pursue and which to
sideline when funding tightened. Anticipating this challenge, he proposed
criteria for determining which research could claim priority in funding bat-
tles, articulating what would become known as the Weinberg criterion: “The
word ‘fundamental’ in basic science, which is often used as a synonym for
‘important,’ can be partly paraphrased into ‘relevance to neighboring areas of
science.’ I would therefore sharpen the criterion of scientific merit by propos-
ing that, other things being equal, that field has the most scientific merit which
contributes most heavily to and illuminates most brightly its neighboring scien-
tific disciplines.”8 Conceptual fecundity was only one element of Weinberg’s
fundamentality calculus. His view included not only conceptual content but
also technological fruitfulness and social relevance as components of funda-
mentality. Tested against these criteria, Weinberg rated particle physics as
poor, writing, “I know of few discoveries in ultra-high-energy physics which
bear strongly on the rest of science.”9
Weinberg pushed this agenda vigorously at the 1964 meeting of the Amer-
ican Physical Society (APS) in Washington, DC. “Science which commands
great public support must be justified on grounds that originate outside the
particular branch of science demanding the support; it must rate high in so-
cial, technological, or scientific merit, preferably in all three,” he argued, and
challenged particle physicists “to state their case clearly, to say exactly why
it is that elementary-particle physics is as important as all our elementary-
particle physicists believe it is. To say that it is ‘fundamental’ in itself does
not answer the question, because one then has to decide what one means by
‘fundamental.’ I tried to interpret the idea ‘fundamental’ in basic science to
mean having the greatest kind of bearing on the rest of science, and even on
other human knowledge.”10 Such a stance holds clear utility for the director
of a large, broadly invested laboratory such as Oak Ridge. Weinberg valued
research programs that would promote useful connections with other enter-
RESPONSES TO THE REDUCTIONIST WORLDVIEW 139

prises within the laboratory or shore up the laboratory’s social and political
support.
Weinberg did not challenge the intrinsic intellectual merit of high energy
physics research—which even among its sharpest critics in the physics com-
munity was never questioned—but pointed out that “if one justifies a branch
of science as a means of expanding man’s cultural horizon, then one gets into
the question of other competing ways of expanding man’s cultural horizon,
like for example, an expansion of the arts and of literature and music.”11 The
argument high energy physicists advanced that the knowledge high energy
physics seeks is uniquely fundamental and so should be pursued for its in-
trinsic cultural merits raised the question of why it was deserving of funds out
of proportion with expenditures for other activities of intrinsic cultural value
but little direct practical relevance.
An approach that focused on the interconnectedness of knowledge set
Weinberg in direct opposition to the emerging orthodoxy within particle
physics. Victor Weisskopf, an MIT particle theorist, took issue, and the two
aired their disagreement in a coauthored Physics Today article entitled “Two
Open Letters,” published in June 1964. The exchange did little to forge a
common understanding, but it did record a clear exposition of each per-
spective. Weisskopf argued that “the nucleon is the basis for all matter and
therefore of all science,” whereas Weinberg retorted that he was nonetheless,
“justified in characterizing high-energy physics as ‘rather remote.’”12 Weiss-
kopf emphasized the possibility in principle of reducing higher-level laws to
lower, but Weinberg focused on the impracticality of doing physics from the
bottom up. Weisskopf claimed that theories of lower levels were privileged
because higher-level theories could be reduced to them; Weinberg allowed
any field that met his fecundity criterion to claim fundamental status. It did
not matter for Weinberg, when choosing which research to fund, whether
or not higher-level phenomena could, in principle, be explained in terms of
lower-level phenomena after the fact because, contra Weisskopf, he tested the
degree of fundamentality research exhibited by its relationship to other ar-
eas of knowledge rather than by its purported relationship to physical reality.
These differences account for why Weisskopf and Weinberg talked past each
other: they disagreed at the core about how fundamentality was derived.
The exchange with Weinberg prompted Weisskopf to develop his think-
ing further, and he again responded to “Criteria for Scientific Choice” in his
own Physics Today missive in 1967. He identified two camps in the physics
community; “intensivists,” he claimed, sought first principles above all else,
140 SOLID STATE INSURRECTION

Figure 7.1. Victor Weisskopf ’s diagram of intensive and extensive research areas. Re-
produced from Victor Weisskopf, “Nuclear Structure and Modern Research,” Physics
Today 20, no. 5 (1967), 23–26, with the permission of the American Institute of Physics

whereas “extensivists” ferreted out applications of those principles, either to


technology or to other scientific fields (figure 7.1).13 The intensivist position,
as Weisskopf described it, aptly characterized the reductionist worldview that
particle physicists adopted in the 1960s.14 He cited Weinberg as a paradigm
extensivist, and his description of extensive research captured one aspect of
the type of fundamentality Weinberg championed—the intersections between
bodies of knowledge—with the caveat that Weisskopf excluded the search for
such connections from the fundamental realm.
The competing views of fundamental science and scientific choice that
Weinberg and Weisskopf advanced are best understood as products of a shift-
ing professional landscape in the 1960s. The choices Weinberg had in mind
when he advanced his eponymous criterion were choices created by a tight-
ening funding climate, in which hard decisions had to be made about which
projects to prioritize. Weisskopf, in turn, was pushed to articulate an assump-
tion that was widespread in the high energy physics community, but which
was rarely systematically defended, because he encountered pushback in the
form of an alternative view of how to navigate the new funding landscape.
Similar concerns would motivate the best-known response to high-energy
reductionism, issued by Philip Anderson.
Figure 7.2. Philip W. Anderson, ca. 1962. Credit: AIP Emilio Segrè Visual Archives. Reproduced
with permission
142 SOLID STATE INSURRECTION

MORE IS DIFFERENT

Philip W. Anderson (figure 7.2), a solid state theorist, was perturbed both by
the financial difficulties basic solid state physics faced and by what he per-
ceived as the field’s unjustly low intellectual status. Anderson had completed
his PhD at Harvard University in 1949. He immediately joined the solid state
group at Bell Laboratories where he imbibed the spirit of the Bell system, in
which he could pursue his interests almost unencumbered.15 As a student
of John Van Vleck, who had fought against the incursion of technical inter-
ests into solid state physics, Anderson was likely predisposed to see the field
in terms of its potential to make fundamental conceptual contributions. The
environment at Bell at the time is nevertheless relevant to appreciating his
standpoint.
Bell was still buzzing from the invention of the transistor two years earlier,
which promised to revolutionize the telecommunications industry. This and
other high-profile successes attracted a slew of talented young physicists, and
Anderson found himself among a uniquely large and accomplished assem-
blage of solid state theorists and experimentalists. Stefan Machlup, a postdoc
who overlapped with Anderson, wrote to William Shockley, the codirector
of the Bell solid state division, to recount his impressions and recalled, “I
think everybody on my hall was doing exactly what he wanted to do,” and
expressed awe at the concentration of expertise, observing that although “it’s
not as ‘gemütlich’ as a college campus . . . [i]f you’ve got an obscure technical
problem, chances are that somewhere in the two-mile network of corridors
sits the expert on this particular specialty.”16
That concentration of expertise, combined with the tremendous success
of the transistor, gave Bell outsized influence over the development of solid
state physics through the postwar years.17 From 1945 to 1970—when the
Physical Review split into four sections—Bell Labs physicists were authors or
coauthors of 1,824 papers in the Physical Review and Physical Review Let-
ters. This made it far and away the leading industrial laboratory by this metric.
Over the same span, General Electric researchers placed 622 papers in these
two journals, Westinghouse 458, and IBM 367. But this output also placed
Bell among the most prolific of any contributors to the country’s most pres-
tigious journals, outpacing the large and well-funded physics departments
and research institutes of the University of Chicago (1,504), the University of
Illinois (1,475), Columbia University (1,149), Stanford University (1,010),
Harvard University (992), Cornell University (976), the University of Penn-
RESPONSES TO THE REDUCTIONIST WORLDVIEW 143

sylvania (948), and Princeton University (942). Bell’s output surpassed each
of the National Laboratories and among universities was bested only by the
University of California, Berkeley (2,649) and the Massachusetts Institute of
Technology (1,864), both of which it would have exceeded if not for contri-
butions from Berkeley’s Lawrence Radiation Laboratory (1,741) and MIT’s
Lincoln Laboratories (316).18
The translation of Bell researchers’ labors into papers as well as patents
reflects a working environment more akin to a large university department
than to most other industrial laboratories of the age. Bell hosted, for instance,
a university-style seminar series that attracted the leading solid state physi-
cists of the day and promoted a culture of intellectual exchange. In Ander-
son’s assessment: “Industry in general was still the big room of eight desks in
Westinghouse and everybody trying to figure out how the transistor works,
and nobody doing his own research. Or it was GE where there were a few
people doing their own research, but they weren’t really the high quality of
Bell Labs. . . . IBM was to become an imitator. Various things were to become
imitators. But Bell Labs was unique. With the attitude at least at this point
that you had a lot of freedom.”19 Through the 1950s and 1960s, benefiting
from Bell’s vibrant intellectual climate and considerable investigative free-
dom, Anderson conducted the research that would earn him the Nobel Prize
he shared with Van Vleck and the British physicist Nevill Mott in 1977, “for
their fundamental theoretical investigations of the electronic structure of
magnetic and disordered systems.”20
By the late 1960s, however, even denizens of the Bell oasis could detect
twinges of the concern abroad in the solid state community. In the spring
of 1967, Anderson delivered a lecture at the University of California, San
Diego, that formed the seed of his 1972 Science article, “More Is Different.”
The talk grew, as Anderson later recalled, from the simmering discontent he
perceived within the solid state community.21 In Anderson’s view, solid state
was accorded somewhat less than its due. Nuclear and particle physicists
were more highly sought-after in influential government advisory roles, many
prominent universities hired solid state faculty only as an afterthought, and
solid state physicists had trouble breaking into prestigious institutions such
as the National Academy of Sciences (NAS). The NAS member rolls validate
Anderson’s concern. Between Charles Kittel in 1957 and both Anderson
and Charles Slichter in 1967, the academy admitted six solid state physicists
compared with twenty-seven nuclear and particle physicists.22
Anderson’s entry into the dispute over the nature of fundamental physical
144 SOLID STATE INSURRECTION

knowledge did not, as might seem natural for a solid state physicist, mirror
Weinberg’s argument that fundamental research is that which exhibits broad
relevance. Instead, he rejected the narrow characterization of intensive re-
search that Weisskopf (who had served on Anderson’s doctoral examining
committee) had advanced. Anderson, however, accepted a key presumption
that Weinberg had rejected: he agreed with Weisskopf and other proponents
of reductionism that fundamentality was innate to some types of scientific
knowledge. He disagreed with them only about the realms in which it could
be found. Despite differences with his ostensible allies, he did attempt to
ground their primary conclusion—that reductionist physics should not be
funded to the detriment of other fields—on a sound philosophical basis. In
doing so he focused the debate around a single disagreement about the nature
of scientific knowledge.
“More Is Different” begins: “The reductionist hypothesis may still be a
topic for controversy among philosophers, but among the great majority of
active scientists I think it is accepted without question.”23 Some biologists,
chemists, psychologists, or even other solid state physicists might have re-
sisted this characterization, but the statement did reflect prevailing trends
in the physics community. In 1970, particle physics research received ap-
proximately four government dollars for every one spent on basic solid state
research.24 This was despite the fact that the American Physical Society’s Di-
vision of Solid State Physics remained the largest division and was over twice
as large as the Division of Particles and Fields.25 The reductionist world-
view—even if it was not, as Anderson claimed, naively accepted within the
scientific community broadly—had succeeded in shaping the way the federal
government funded physics.
Anderson opposed the reasoning, which he attributed to particle phys-
icists, “that if everything obeys the same fundamental laws, then the only
scientists who are studying anything really fundamental are those who are
working on those laws.”26 This view, according to Anderson, started from
the hypothesis that laws and concepts operating on any given level of com-
plexity could be reduced to laws and concepts at a lower level of complexity.
It thereby concluded that only research addressing the ultimate constituents
of matter and energy could be truly fundamental. The reductionist argument
built into particle physicists’ philosophy of scientific knowledge the justifica-
tion for pursuing research on progressively smaller scales, using accelerators
of progressively higher energy and greater cost. By arguing that the founda-
tional character of smaller physical scales conferred privilege upon knowl-
RESPONSES TO THE REDUCTIONIST WORLDVIEW 145

edge of the laws governing those scales, reductionism supported the view that
science funding should reflect the hierarchy that particle physicists saw in the
physical world.
In his efforts to undermine this position, Anderson avoided the Weinberg
criterion and its emphasis on conceptual, technological, and social applica-
bility. He accepted the standard reductionist premise that laws governing
higher-level phenomena could be reduced to laws governing lower-level phe-
nomena, but rejected the inverse, claiming that the particle physicists’ view
of fundamentality rested on a constructionist hypothesis, which assumed
that higher-level laws could be extrapolated from lower-level laws. With this
move, Anderson attacked the narrow definition of intensive research that un-
derwrote the reductionist scientific hierarchy, preserving the ability of solid
state physicists to claim fundamental insight, contending: “The main fallacy
of this kind of thinking is that the reductionist hypothesis does not by any
means imply a ‘constructionist’ one: The ability to reduce everything to sim-
ple fundamental laws does not imply the ability to start from those laws and
reconstruct the universe.”27 Anderson’s emergence-based view of fundamen-
tality granted the concepts and laws of solid state physics independence by
virtue of the fact that they could not realistically be derived from lower-level
concepts and laws alone.
He illustrated his point with the example of an ammonia molecule, aiming
to demonstrate construction’s failure even at the level of simple molecular
systems. Naively, ammonia should have a dipole moment, given its asymmet-
ric, pyramidal structure. A nitrogen atom forms polar covalent bonds with
three hydrogen atoms, leaving the nitrogen with a net negative and the hy-
drogen with a net positive charge. The resulting tetrahedron, though, does
not empirically behave like a dipole. In its stationary state, the molecule is in
a superposition of the left-hand and right-hand orientations; when observed
through time it undergoes a process of inversion, in which the nitrogen atom
tunnels through the plane defined by the hydrogen atoms and emerges on the
other side, several billion times per second, canceling out any dipole effect.
A broad swath of physicists would have been familiar with ammonia’s
properties; it provided the material basis for the original maser, which Charles
Townes and his research group had announced, to considerable acclaim, in
1955.28 Nitrogen inversion itself had been a familiar chemical process for
several decades. It received renewed interest in the mid-1950s when nuclear
magnetic resonance spectroscopy allowed it to be measured with accuracy
superior to that provided by older radio-frequency techniques.29 Anderson
146 SOLID STATE INSURRECTION

leveraged a familiar example to illustrate the implausibility of applying foun-


dational symmetry laws to larger systems without reference to higher-level
structure: “I would challenge you to start from the fundamental laws of quan-
tum mechanics and predict the ammonia inversion and its easily observable
properties without going through the stage of using the unsymmetrical pyra-
midal structure, even though no ‘state’ ever has that structure.”30 Construc-
tion, in other words, is impracticable: “The relationship between the system
and its parts is intellectually a one-way street. Synthesis is expected to be
all but impossible; analysis on the other hand, may not only be possible but
fruitful in all kinds of ways,” a result he hoped would undercut “the arrogance
of the particle physicist” and allow that “each level can require a whole new
conceptual structure.”31 Anderson did not draw out this statement’s implica-
tions, but it is tempting to see the tacit suggestion that a whole new concep-
tual structure would require a whole new funding structure.
The distinction between impossible in practice and impossible in princi-
ple lingers under the surface of Anderson’s analysis. He rejected the restric-
tive view of intensive research, and yet hedged when saying that the synthesis
of lower-level laws to find higher-level laws is “all but impossible” and call-
ing it an “intellectual,” rather than a physical or natural one-way street. He
claimed that the laws of solid state physics could never practically be extrap-
olated from quantum mechanics without reference to empirically established,
higher-level phenomena; he fought shy of the stronger claim that higher-level
laws could never in principle be derived from below.
Anderson does not explain the delicate dance he executes by linking the
independence of higher-level concepts to practical rather than physical con-
siderations. I propose four reasons this position is notable. The first two re-
flect Anderson’s more general focus on the practice of science. The third and
fourth draw on other elements of his context and help explain how “More
Is Different” fit within it. I do not reject the possibility that Anderson quite
straightforwardly considered the argument that new levels were independent
in practice to be the better-justified position. Rather than providing an ex-
haustive account of why Anderson held the view he did, the factors discussed
here demonstrate how richly interconnected Anderson’s position was with
his professional context. Anderson has himself acknowledged, in retrospect,
that his perspectives were shaped by conditions in the scientific community
that were unfavorable to solid state physicists: “Sociologists of science posit
that there is a personal or emotional subtext behind much scientific work,
and that its integrity is therefore necessarily compromised. I agree with the
RESPONSES TO THE REDUCTIONIST WORLDVIEW 147

first but reject the second. I think ‘More Is Different’ embodies these truths.
The article was unquestionably the result of a buildup of resentment and dis-
content on my part and among the condensed matter physicists I normally
spoke with.”32 Considering how the fine structure of Anderson’s argument
worked within the context that motivated it therefore builds upon his own
understanding of its origins.
The first reason Anderson’s focus on practical-level independence is no-
table is that it reflects his more general focus on scientific practice. Here, he
is consistent with Weinberg. Permissive views of fundamentality were uni-
formly hardheaded about the process of scientific research. The in-principle
derivability of higher-level phenomena was academic if it provided no prac-
tical directives for doing science. Anderson reprised his argument in 2001:
“A perverse reader could postulate a sufficiently brilliant genius—a super-
Einstein—who might see at least the outlines of the phenomena at the new
scale; but the fact is that neither Einstein nor Feynman succeeded in solving
superconductivity.”33 Superconductivity was a notoriously intractable the-
oretical puzzle. Failing to derive a theory of it was almost a rite of passage
for the most accomplished theoretical physicists before John Bardeen, Leon
Cooper, and Robert Schrieffer succeeded in 1957. Felix Bloch advanced the
tongue-in-cheek theorem that all theories of superconductivity can be dis-
proved, a refinement of Wolfgang Pauli’s more cutting version of the theorem:
theories of superconductivity are wrong. The core insight that led to the suc-
cessful theory, Cooper’s realization that electrons paired off at very low tem-
peratures, forming “Cooper pairs” that traveled without resistance through
superconducting materials, required thinking about the system in terms of
the relations among its components, rather than by building up from first
principles. Similarly, Anderson’s ammonia example drew its force from the
fact that anyone attempting to describe an ammonia molecule’s behavior for
the first time would, by any reasonable understanding of practice, be required
to employ higher-level concepts in addition to first principles.
Second, that necessity supplied a font of new solid state problems. In a
1999 interview with Alexei Kojevnikov, Anderson recalled being motivated
to develop his philosophical views in part by a lecture Brian Pippard, a
Cambridge solid state theorist, had given at a superconductivity conference
hosted by IBM in 1960.34 Pippard lamented a lack of compelling and accessi-
ble fundamental problems in solid state, suggesting that solutions to the most
prominent—such as superconductivity—had deprived the field of appealing
intellectual challenges for young talent. Pippard offered up the gloomy prog-
148 SOLID STATE INSURRECTION

nostication that “ten years is going to see the end of our [solid state physi-
cists’] games as pure physicists, though not as technologists,” and advocated
“a swing of emphasis now away from pure research to applications,” which
should necessitate exposing promising students to “the methods of research
in industrial laboratories.”35 Anderson, who spent the 1967–68 academic
year as Pippard’s colleague during a visiting professorship at Cambridge, de-
scribed him as “a professional pessimist.”36 The second axis of Anderson’s
argument from practice is evident in his reaction against Pippard’s gloomi-
ness about academic solid state research. “More Is Different” makes the case
for the widespread availability of academy-friendly, intellectually interesting
basic research problems in solid state physics. The practical necessity of em-
ploying higher-level concepts to describe solid state systems provided, for
Anderson, a nearly inexhaustible supply of new and interesting fundamen-
tal questions. The failure of construction ensured that solving long-standing
problems did not impoverish the field as much as Pippard supposed; sur-
prising physics could always be expected when considering the next level
of complexity. This interpretation meshes well with Anderson’s clear pref-
erence for practical considerations, because the question of in-principle in-
dependence had little bearing on whether an adequate supply of interesting
research problems would be available to slake the intellectual thirst of future
graduate students.
Third, the narrow focus on practice was expedient. A strong claim about
the nature of objective physical reality was not essential to allow solid state
research a claim to fundamental knowledge given an argument that denied the
practical possibility of synthesis. Because claims to fundamental knowledge
and financial support were correlated during this period, at least for particle
physicists, Anderson can be read as making the weakest claim necessary to
advance his position without inviting attack from those who objected to the
wholesale independence of higher levels from lower levels. As long as the case
could be made for acquiring knowledge of higher levels, questions of physical
hierarchy were merely supposition.37
Fourth and finally, the contours of Anderson’s philosophical position and
the consequences it had for fundamentality debates can be understood in
terms of changing prestige politics in the late 1960s and early 1970s. Physics
enjoyed considerable prestige following the Second World War, but as the
funding plateau Weinberg predicted arrived, prestige, like financial support,
became a limited resource, leaving physicists to carve it up among their sub-
disciplines. Separating reduction and construction severed the link between
RESPONSES TO THE REDUCTIONIST WORLDVIEW 149

physical hierarchy and intellectual hierarchy. Anderson sought to deny par-


ticle physicists an exclusive claim to fundamental knowledge, knowing, as
Weinberg did, that “fundamental” often meant the same thing as “impor-
tant.” The weaker position also avoided the question of whether a permissive
view undermined the prestige of physics more generally. Solid state might
have been struggling in this period, but physics was still firmly established
as the standard bearer for American science. A stronger view implying the
equivalence of all scientific knowledge would have been strange given the
circumstances.38 Although Anderson gives no indication that this was a con-
scious motivation, his argument does have the convenient consequence of
undermining the exclusive claim that particle physicists laid on fundamental
knowledge without similarly undermining the more general entitlement phys-
icists felt to rare levels of social approbation and federal funding. Anderson
subsequently expanded his view to encompass concepts in the social and bi-
ological as well as the physical sciences, but recalled that his initial sensitivity
to the level-dependence of concepts arose because it was “the principle by
which my own field of science arose from the underlying laws about particles
and interactions; and it was only as I broadened my perspective that I realized
how general emergence is.”39 It is therefore appropriate to understand An-
derson’s 1972 stance as primarily about physics, even though he later refined
it into a view about science in general. It is also worth noting that by the time
Anderson’s emergentism had fully matured, biology had unseated physics as
the doyen of the American sciences.
Examining these contextual pressures brings Anderson’s departure from
the Weinberg criterion into focus. Their differences correspond to a shift in
conditions within the scientific community. The funding pinch in the late
1960s and early 1970s was asymmetrical. Particle physics, the flagship enter-
prise of reductionism, enjoyed continued success in the form of new, expen-
sive facilities. Anderson recognized the growing prestige and funding gaps
between solid state and particle physics. Amid these conditions, which be-
came more acute as large government grants became more difficult to obtain,
Anderson developed a view of fundamentality that departed sharply from
fecundity arguments, although it arrived at similar conclusions. Given the un-
distinguished, impecunious position Anderson perceived solid state physics
to occupy in the late 1960s and early 1970s, the well-worn claim that research
needed only to provide a basis for further research to be fundamental would
not have met the challenge that particle physics posed. Instead, Anderson
sought the source of fundamentality in the nature of physical knowledge,
150 SOLID STATE INSURRECTION

adopting the strategy that had served particle physicists so well. “More Is Dif-
ferent” makes the best historical sense when placed against the foil of particle
physics and its strong reductionism and set within a context where physics
still dominated American science. The notion that fundamentality measured
breadth of fruitfulness implied no hierarchy and did not play favorites among
the sciences. Anderson, by accepting the innate view of fundamentality more
typical of the reductionist account, denied particle physics an exclusive claim
to the privilege physical knowledge enjoyed, as a matter of course, over chem-
ical, biological, or social scientific knowledge.

THE POWER OF PROFESSIONAL PRESTIGE


Both the rise of reductionism in the high energy physics community and re-
sponses from other physicists suggest that changing financial realities and
corresponding hierarchical shifts within the scientific community in the
1960s exerted pressure on physicists to develop the presumptions that oth-
erwise tacitly governed their practice into more fully realized philosophical
positions. Weinberg’s criteria for scientific choice presumed, correctly, that
the post–Second World War funding honeymoon would wane and that some
lines of intellectually promising research would have to be prioritized over
others. And to whatever extent Anderson’s views on fundamentality might
have been based in his research, he was moved to refine and articulate them
by professional challenges.
Those challenges, from the late 1960s on, were frequently parsed in terms
of professional prestige. Anderson’s worry that solid state physics was over-
whelmingly associated with technical applications was born out by media de-
pictions of the field in the years after “More Is Different” appeared. The New
York Times coverage of the Nobel Prize he shared with John Van Vleck and
Nevill Mott in 1977 announced that “the three winners, all theoreticians in
solid-state physics, were cited for work underlying the development of com-
puter memories, office copying machines and many other devices of mod-
ern electronics.”40 In 1973, the lede to the article reporting Leo Esaki, Ivar
Glaever, and Brian Josephson’s prize for fundamental work in superconduc-
tors and semiconductors read, “The three winners of this year’s Nobel Prize
in Physics made discoveries regarding phenomena unfamiliar to the layman,
yet vital to his television set or the computers that affect many aspects of his
life.”41 Particle physics awards received somewhat different coverage in this
era. Three years later, the Times science reporter Malcolm W. Browne, after
describing a chemistry prize relevant to the pharmaceutical industry, shifted
RESPONSES TO THE REDUCTIONIST WORLDVIEW 151

to the physics prize, awarded to Steven Weinberg, Sheldon Glashow, and Ab-
dus Salam, writing: “While such practical applications have no part in the
work of the physics prize recipients, many scientists regard their work as fun-
damental to understanding nature.”42
The persistent portrayal of solid state work as subservient to technical
aims remained a sore point for Anderson and his peers. Developing a robust
philosophical case for the field’s intellectual viability was one part of their
response. But it would not be sufficient. Addressing the tension between the
technical dimensions of solid state physics and its aspirations to intellectual
prestige would require a more thorough reimagining of its identity.
8

BECOMING CONDENSED MATTER PHYSICS

Adding a chapter so named to the conventionally labeled group of


mechanics, heat, acoustics, and so forth is, of course, a little like trying
to divide people into women, men, girls, boys, and zither players.
—DWIGHT GRAY, 1963

The second edition of the American Institute of Physics Handbook, a compre-


hensive reference volume, appeared in 1963. The revised edition included
a new section devoted to solid state physics.1 Its editor, Dwight Gray, an-
nouncing the handbook’s release in Physics Today, delivered the quip in the
epigraph above and recounted his coeditor’s droll suggestion “that perhaps
the book should contain only three major sections—Solid-State Physics,
Liquid-State Physics, and Gaseous-State Physics.” The editors resolved,
however, that “the advantages of consolidating solid state material into one
chapter outweighed the disadvantages of a somewhat untidy classification
system.”2
Through the 1960s, solid state physicists showed similar pragmatism re-
garding their field’s untidiness. Perhaps solid state was a problematic cate-
gory, but it was expedient; it created elbow room for applied physicists to
pursue technical research without threatening their identity as physicists.
But beginning in the late 1960s, that expediency began to wear thin. Indus-
trial and applied physics were by that time well-established members of the
American physics community. The professional concerns du jour among
solid state physicists centered instead on intellectual standing—the type of
concerns to which Philip Anderson was responding when he wrote “More
Is Different” in 1972. Efforts to reform solid state physics so as to empha-
BECOMING CONDENSED MATTER PHYSICS 153

size its intellectual vibrancy lent impetus to a new name, “condensed matter
physics,” which gradually gained popularity in the late 1960s and through
the 1970s. Although condensed matter physics would encompass many of
the topic areas that constituted solid state physics, its aims were substan-
tively different. Robert Proctor notes in his study of common suffixes that
“the names given to particular science fields and subfields are shaped, to a
certain extent, by ideological baggage picked up in the course of usage.”3 Far
from being innocuous, the name change reflected the ideological baggage of
“solid state physics.” It was the culmination of long-standing tensions within
the solid state community between the pro-industry agenda that motivated
Roman Smoluchowski and a desire for a conceptually coherent definition of
the discipline’s purpose and scope, which emphasized its contributions to
fundamental physical understanding.
The shift toward condensed matter and away from solid state language
comes into clearest focus through the lens of the federal advisory apparatus.
In the mid-1940s, when solid state physics emerged, the factors the physics
community examined when defining its categories were based predominantly
on internal professional concerns. Because physicists did not come to regard
centralized federal support as the norm until post–Second World War pat-
terns had stabilized, solid state physics, as incubated in the American Physi-
cal Society (APS) in the late 1940s and early 1950s, took shape in response to
the internal professional concerns of the physics community rather than re-
sponding directly to federal funding incentives. By the 1970s, the enormous
increase in government funding had changed how new disciplinary catego-
ries were constructed. As the case of the Advanced Research Projects Agency
(ARPA) and materials science shows, top-down federal incentives were po-
tent forces determining how physicists arranged their activities, and how they
talked about them. The term “condensed matter physics” originated in the
physics community, but its status as a disciplinary category was fixed by its
enshrinement as a funding category before this change was reflected in the
APS, which, by this time, had an extensive and entrenched divisional struc-
ture that was more difficult to change than it had been in the 1940s.
Three National Academy of Sciences (NAS) reports, published in 1966,
1972, and 1986, chart the rise of condensed matter physics. Examining these
against both the ideological background set out in the preceding chapters and
the institutional and conceptual evolution of the field that occurred between
the mid-1960s and the mid-1980s illustrates how condensed matter physics
emerged as an alternative to solid state and exposes the qualitative differences
154 SOLID STATE INSURRECTION

between the two categories. Historians and physicists alike commonly treat
“solid state physics” and “condensed matter physics” as effective equivalents,
distinguished only because they were preferred in different eras. Anderson,
a member of the first generation of American physicists trained in solid state
theory and an early adopter of the “condensed matter” label, assumes conti-
nuity when referring to “solid state (now ‘condensed matter’) physics.”4 Sim-
ilarly, Helge Kragh writes: “From a sociological and historical point of view,
solid state physics did not exist [in the 1930s]. It was only after World War II
that the new science of the solid bodies, later to be called condensed-matter
physics, took off.”5 These claims are not without merit. The shift from solid
state to condensed matter physics was marked by substantial continuity of
physical problems and practices; however, topical and methodological conti-
nuity do not translate unproblematically into disciplinary continuity.
This straightforward equivalence between solid state and condensed mat-
ter is sometimes complicated by appealing to condensed matter’s broader
topical scope. Walter Kohn’s historical treatment of solid state physics sug-
gests that it “was enlarged to include the study of the physical properties
of liquids and given the name ‘condensed matter physics.’”6 Volker Heine
points to polymer research as the catalyst for renaming the Cavendish Lab-
oratory’s solid state theory group “Theory of Condensed Matter.”7 Spencer
Weart takes these partial observations further by suggesting that condensed
matter resolved difficulties intrinsic to solid state: “The newly popular name
included liquids and, like ‘materials science’ in a different manner, reflected
a persistent uncertainty as to whether ‘solid-state physics’ was the best way
to group subfields.”8 Weart’s insight points toward a richer story about the
name change, which was more than either a simple rebranding or the recti-
fication of a long-standing error. Condensed matter did respond to nagging
skepticism about solid state, but the parallel growth of materials science indi-
cates that addressing these concerns was neither simple nor straightforward.
Condensed matter physics represented a resurgence of a form of the pure
science ideal within the solid state community. But it was not Rowland’s pure
science ideal. Condensed matter physicists, as denizens of the late Cold War,
championed physical knowledge as being simultaneously fundamental and
relevant to technical development. It was an audacious ambition. High en-
ergy physics had sustained its intellectual status largely by disdaining tech-
nical connections. Burton Richter’s response to a question about his work’s
possible application after being awarded the Nobel Prize, with Samuel Ting,
BECOMING CONDENSED MATTER PHYSICS 155

for the discovery of the J/Ψ meson—“The significance is that we have learned
something more about the structure of the universe. In terms of practical ap-
plication right now, it’s got none”—was a typical attitude.9 The close link
between solid state work and workaday technologies was so tight that it often
obscured conceptual accomplishments. Condensed matter physics took on
the challenge of reasserting the intellectual contributions of physical work on
complex matter while also gambling that its proximity to questions of techni-
cal interest would preserve its most stable funding streams.

THE ORIGINS OF CONDENSED MATTER


Condensed matter, like solid state, was a technical term before becoming a
disciplinary category. When physical questions did not depend on a spe-
cific state of matter, as long as it was dense enough, physicists talked about
“condensed matter” as a medium through which, for example, a muon might
travel and exhibit noteworthy behavior.10 Early uses of this type were scat-
tered, and almost exclusively by particle physicists. As a designator of a dis-
cernible field of inquiry, the term appeared first in Europe and only slowly
percolated to the United States.11 The German journal Zeitschrift für Physik
reorganized in 1962, titling its Section B “Condensed Matter.” The Physical
Review followed suit, to a limited extent, a year later, announcing in October
1963 that in the following year: “The first twelve issues of each volume will be
divided into two sections (A and B) of six issues each, appearing alternately.
Section A will be primarily devoted to the physics of atoms, molecules, and
condensed matter, and Section B will be primarily devoted to the physics of
nuclei and elementary particles.”12 But such early topical uses fell short of
becoming disciplinary designators in the United States. In the Physical Re-
view, early uses of “condensed matter” to indicate a field of study invariably
referenced Zeitschrift articles or conferences devoted to condensed matter
held in Europe, where the term held more currency.13
A more significant development was the launch of the journal Physik der
kondensierten Materie, founded in West Germany in 1962, published by
Springer-Verlag and edited by the Swiss physicist Georg Busch.14 The jour-
nal was published simultaneously as Physique de la matière condensée and
Physics of Condensed Matter, and accepted articles in German, French, or
English. The boilerplate description added in the second issue described its
scope as “relating mainly to thermal, electrical, magnetic and optical proper-
ties of solids and liquids in the broadest sense.” It explained the decision to
156 SOLID STATE INSURRECTION

cast its net beyond solids: “Inclusion of work in the physics of both solid and
the liquid phase is intended to increase closer contact between both areas and
especially to further research in the area of liquids.”15
Physics of Condensed Matter sought a broader remit than solid state phys-
ics, which in Europe tended to remain restricted to solids with regular lattice
structures. The German term festkörperphysik was little known before the
mid-1950s, when the first few professorships in the physics of solids were
awarded in Germany.16 The earliest evident published use is in the title of
a proceedings volume of a conference held in Dresden, May 9–11, 1952.17
It reached a more general audience in Die Naturwissenschaften in 1954 as
the title of a review article that synthesized mostly American and German
sources.18 Its author, Heinz Pick, alluded to the term’s recent provenance: “In
the catalogue of major fields of modern physical research, one meets increas-
ingly often with the concept solid state physics.”19 He also made observations
similar to those of his American counterparts about the category’s conceptual
consistency, remarking: “One is inclined to take such a word to designate a
clearly defined, unified field. On closer inspection, this hope turns out at first
to be entirely unconfirmed.”20 The way out of this dilemma, for Pick, was to
restrict the topical range of the field. He concluded that festkörperphysik was
conceptually consistent by virtue of revolving around the lattice structure of
crystalline solids, ignoring noncrystalline solids, superfluidity, the magnetic
susceptibility of gases, and other topics with little or nothing to do with lat-
tice structure that fell within the American solid state synthesis. By excluding
them, Pick was able to sweep troubling inconsistencies aside, even if it meant
defining the field more narrowly.
France was slower than Germany in adopting its own analogue of solid
state, but, as in the German case, the category’s amorphous nature allowed
it to bend to local priorities. Pierre Teissier has attributed the inelasticity of
French institutions after the war to the persistence of heavy-handed “feudal
regimes” that guided French research through the late 1950s. It took a new
generation of researchers, trained after the war, to dislodge this entrenched
system. When “physique du solide” (or, occasionally, physique de l’état solide)
appeared in France in the late 1950s and early 1960s, so did the equally po-
tent chimie du solide, which harnessed a long French tradition in chemistry.21
Teissier’s assessment is borne out by patterns of usage in French journals.
The first evident use of “physique du solide” in Le Journal du Physique et la
Radium was by Jacque Friedel in 1955, who, in an article on ferromagnetism,
referred to atomic and molecular orbitals as “the two fundamental approxi-
BECOMING CONDENSED MATTER PHYSICS 157

mations of solid state physics and chemistry.”22 The term would not recur in
the country’s flagship physics journal until 1959 and 1960, when it began to
appear in the names of French institutions, such as the Service de Physique du
Solide et de Résonance Magnétique in Saclay and the Laboratoire de Magné-
tisme et de Physique du Solide in Bellevue.
That Friedel would be an early adopter of solid state nomenclature is un-
surprising considering he took his PhD at the University of Bristol in 1952,
studying under Nevill Mott.23 By 1952 Mott, who identified his work as solid
state physics, was the director of a lively research group on physics of solids at
Bristol’s H. H. Willis Physical Laboratory.24 British physicists were more apt
than French or German physicists to publish in American journals or attend
conferences in the United States and the absence of a language barrier permit-
ted comparatively easy acceptance of the term in Britain.25 Lawrence Bragg
lectured at the Royal Institution on “The Physics of the Solid State” in March
1949.26 Although British institutions, like their French counterparts, were
steeped in tradition and loath to adopt new names, the establishment of the
International Journal of the Physics and Chemistry of Solids by Oxford-based
Pergamon Press in 1956 gave the moniker a stronghold on the British Isles.
Solid state physics, as a category, had a weaker hold on the European con-
tinent than it did in the Anglo-American world, in large part because its US
incarnation owed so much to the peculiar features of the relationship between
American universities and American industry. As a result, the terminology of
condensed matter physics, taken up rapidly in Europe, was slow to catch on
in the United States, even as research emphases shifted in similar directions
on both continents. The Physical Review’s partial fission in October 1963,
with one section “devoted to the physics of atoms, molecules, and condensed
matter,”27 was the only clear occurrence of “condensed matter physics” in
APS journals throughout the 1960s; discussions of reorganization in the APS
Executive Committee uniformly call section A the solid state section.28 Con-
densed matter physics language, in other words, was rare and marginal in the
1960s, an assessment borne out by a contemporary overview of the field from
the National Research Council (NRC) of the National Academy of Sciences.

THE PAKE REPORT


In 1962, Frederick Seitz was elected president of the National Academy of
Sciences. One of his first major initiatives was to commission a series of re-
ports on the current status of the scientific fields in the academy’s purview,
to be undertaken by the NRC. The committee surveying physics included
158 SOLID STATE INSURRECTION

established solid state physicists Charles Townes, Harvey Brooks, and Ro-
man Smoluchowski, along with David Pines, who had recently earned his
stripes exploring the implications of the Bardeen–Cooper–Schrieffer theory
of superconductivity. The Pake report (named for its chairman, George Pake)
was published in 1966 and identified “Solid-State (and Condensed-Matter)
Physics” as one of the primary divisions whose progress it addressed.29 The
solid state and condensed matter subcommittee inserted a footnote into a
draft of the report in April 1964 explaining the naming decision, pointing
out that around 90 percent of the field consisted of work on solids, thus using
“solid state” as a general term was good enough for government work.30
“Condensed matter” in the Pake report was both literally and figuratively
parenthetical. Despite the passing acknowledgment that it might be a more
appropriate term, the compilers referenced condensed matter only when the
phenomenon under discussion deviated too uncomfortably from the realm
of solids. They described early research in the field, for example, by slipping
seamlessly from talking about solids to invoking condensed matter when dis-
cussing superfluidity: “Until the beginning of this century . . . the science
of solids remained almost entirely empirical and descriptive. Between 1912
and the early 1930s, most of the salient properties of condensed matter, with
the striking exception of superfluidity, were understood at least qualitatively.”
Such results ensured, the report continued, slipping back into the language
of solids, that “[t]he stage was set for the beginning of solid-state physics in
its present sense.”31
“Condensed matter” papered over spots in the Pake report where the re-
strictive nature of “solid state” became too obvious for comfort. As of the
mid-1960s, the term did not presage sweeping changes to the structure and
identity of solid state physics. Though the report insisted that solid state “is
a fundamental branch of physics” and suggested that future progress in solid
state “could well turn out to be of greater significance to our knowledge of
the world than further progress in elementary-particle physics,” it saved the
greatest emphasis for solid state’s technical contributions.32
The section entitled “Intellectual Challenge” began: “In solid-state phys-
ics there is at present no clearly visible need for radically new concepts” and
made the case for conceptual importance by pointing to inchoate research
areas, such as noncrystalline solids, as the potential but unproven source of
“new concepts and principals.”33 The report reflected the pessimism at large
in the 1960s regarding solid state’s potential to make foundational intellec-
tual contributions. Brian Pippard’s recommendation that solid state physi-
BECOMING CONDENSED MATTER PHYSICS 159

cists turn away from curiosity-driven research and start training students for
careers in industry, which had so rankled Philip Anderson, was symptomatic
of the sense at the time that solid state had pushed about as far as it could into
the conceptual frontier.34 Fields such as plasma physics, which had secured
its place with an APS division of its own in 1959, benefited as physicists in-
terested in complex matter looked beyond solid state for greener intellectual
pastures.35
In a similar, if less defeatist spirit, the Pake report, although tepid about
solid state’s basic research potential, raved about its “indispensable [role] in
numerous technological developments,” boasting that “the whole [of ] com-
munications technology is being fundamentally affected by these [solid state]
developments.”36 The authors, shifting from their cautious tone when dis-
cussing solid state’s intellectual importance, emphasized the “indispensable,”
“vital,” and “essential” contributions solid state research made to technolog-
ical systems that were “totally dependent” on solid state devices and know-
how.37 Solid state was still building its intellectual portfolio, but its techno-
logical track record was strong. As long as the field justified itself primarily
on technological grounds, its most prominent research programs would be
focused around solids; nagging concerns about solid state’s appropriateness
as a category could be swept under the rug, as they had been in the 1962 AIP
Handbook.

THE BROMLEY REPORT


In the early 1970s—as dismay over the widening prestige gap between solid
state and particle physics peaked—“condensed matter” began to replace
“solid state” as the preferred term in advisory circles. The NRC report pub-
lished in 1972 included a chapter on “Physics of Condensed Matter.”38 It
eschewed the language of solid state when referencing the contemporary
field, though it retained solid state terminology for historical observations. In
all, the two terms appear with about equal frequency. The 1972 committee
contained many of the same members as the 1966 group—notably Brooks,
Townes, and Smoluchowski—but some new recruits wrote the chapter on
condensed matter physics. Among them was Morrel Cohen, a University of
Chicago physicist who was one of two Americans who had served on the ed-
itorial board of Physik der kondensierten Materie since its founding and who
had begun describing his research specialty as “physics of condensed matter”
no later than 1964.39
The shift in language was accompanied by newly potent concerns over
160 SOLID STATE INSURRECTION

funding patterns. The condensed matter chapter explained: “Our objective is


to show that basic research in physics of condensed matter, performed solely
to understand in the deepest possible way the complex behavior of solids and
liquids, has been the source of two decades of unprecedented achievement
in critical new technologies. We see no way in which these achievements
could have been planned in the past and no way in which further progress
can be programmed except by continued support of basic research.”40 Pro-
gramming—the planning and funding of research programs based on precon-
ceived technological goals—was an ever-present bugbear for solid state and
condensed matter physicists concerned for their intellectual prestige. The
formulation of research programs with specific practical outcomes in mind,
the modus operandi in materials science, threatened to undermine their in-
tellectual autonomy and motivated the condensed matter panel to state the
field’s intellectual value much more vociferously than their counterparts had
in 1966.
The 1972 panel was accordingly circumspect about solid state’s techno-
logical contributions. They asserted that solid state research had been re-
sponsible for steady innovation, but were careful to emphasize “the richness
and complexity of these events [inventions of solid state technologies], as
well as the varied motivations of the scientists and engineers involved.”41 By
emphasizing the complexity and capriciousness of the routes from basic re-
search to technological applications, the panel sought to protect federal fund-
ing for basic solid state research by undermining the federal government’s
tendency to link basic research expenditure to clearly articulated outcomes:
“The United States would not spend its research and development dollars
nearly so well if it insisted either that basic research be strongly motivated by
and directed to practical goals or that all basic research be isolated from prac-
tical considerations.”42 The shift toward condensed matter coincided with
rising concern about an environment that married research support to practi-
cal outcomes—the “coupling” that had so chagrined Benjamin Lax.
The same forces to which the 1972 NRC report reacted prompted Philip
Anderson to defend the intellectual merit of complex-systems physics in
“More Is Different” that same year. “More Is Different” introduced another
wrinkle to the name question, referring to “many-body physics.” The term
calls to mind the quantum many-body problem, the notorious intractabil-
ity of calculating a wave function for quantum systems containing three or
more interacting bodies. The impossibility of deducing the properties of
many-body systems from first principles meant, for Anderson, that the con-
BECOMING CONDENSED MATTER PHYSICS 161

cepts employed to understand complex systems were just as intellectually


valuable as those used to understand elementary particles.43 Although it con-
tinued to describe research confronting the quantum many-body problem,
“many-body physics” never threatened to supplant “solid state physics” as
a designator of the larger community, in part, no doubt, because of its nar-
row focus on theory. Anderson’s usage of the term, however, mirrors the way
condensed matter began to be used in the same era, sometimes by Anderson
himself.44 Both terms refocused attention on methodological commonalities
and thus fit more conveniently into the narrative around intellectual prestige
and research funding that condensed matter physicists preferred.
That “condensed matter” was a professionally motivated category in
the 1970s is evident when examining its published usage. Between 1970
and 1979, only 32 articles in APS journals contained “condensed matter”
in their titles or abstracts. In comparison, 121 included “solid state,” 1,019
contained “solids,” and “high energy” appeared 1,399 times.45 The move
toward condensed matter language in the NRC, then, was not a reflection of
prevailing research practices or self-identification patterns. “Solid state” and
“condensed matter” were both capacious umbrella categories that did little to
shape day-to-day research. They had greater meaning as professional group-
ings and funding designations. The terminological shift reflected a change
of professional ideology brought about as solid state physicists confronted
challenges to their intellectual prestige and funding for exploratory research.
The watershed moment for condensed matter came with a proposal at the
January 1978 council meeting to rename the Division of Solid State Physics
(DSSP) the “Division of Condensed Matter Physics.” The suggestion scan-
dalized representatives from the Division of Fluid Dynamics (DFD), who
perceived an invasion of their turf: “[François] Frenkiel expressed concern
over the overlap between the subject matter covered by a Division of Con-
densed Matter Physics and by the Division of Fluid Dynamics, and noted
that approval of such a change would force the Division of Fluid Dynam-
ics to change its name. [Tony] Maxworthy noted that the Division of Fluid
Dynamics Executive Committee expressed strong feeling against the name
change.”46 The matter was postponed until April, giving the two divisions
some time to hash out their differences. The motion succeeded at the April
meeting, but it remained controversial, passing by a vote of fifteen in favor to
seven opposed at a time when unanimity was the norm for council votes.47
The DSSP had been identified with solids for so long that the new name
stepped on the toes of other divisions. The Division of Fluid Dynamics did
162 SOLID STATE INSURRECTION

not need to change its name, but its members’ resistance shows that the shift
from solid state to condensed matter included significant reorientation of
solid state’s professional goals at a time when research on fluids was gaining
ground. As research on liquid helium, for example—one of the more colorful
feathers in the solid state cap, with its exotic properties like superfluidity—
became a more active area of research, the language of solids became corre-
spondingly more inconvenient and the observation from the Pake report that
solids accounted for upward of 90 percent of the activity in the field no longer
applied.48 The DFD understandably saw the DSSP’s efforts to bring fluids
under its aegis as imperialistic. The condensed matter partisans within the
DSSP, for their part, pursued a new conceptual focus that inconvenienced
the existing institutional structure that had grown in the days when solid
state maintained a more thoroughly industrial reputation. The DFD had been
founded in 1948, shortly after the DSSP. The renaming in the late 1970s sig-
naled a newfound commitment to advocating for condensed matter as a basic
research enterprise and, in so doing, threatened the organizations that had
filled that void in the years when solid state physics was a more thoroughly
industrial pursuit.
Solid state physics had always had a strong industrial patina. In the Na-
tional Research Council’s 1946 survey of US industrial laboratories, just be-
fore the DSSP’s founding, only Bell Laboratories counted “solid state phys-
ics” among its research specialties.49 By the 1960 edition of the same survey,
the term was not limited to the likes of Bell, which maintained a large basic
research staff. The American Machine and Foundry Company in Stamford,
Connecticut, Hughes Aircraft Company’s Newport Beach lab, and Control
Data Corporation in Minneapolis were among the dozens of laboratories that
counted solid state among their specialties, including many that listed no
physicists among their researchers.50
Solid state’s applied bent was well understood by its practitioners, even
those who preferred to nudge it away from industry. Research for the Pake
report concluded that in 1963 60 percent of funding for solid state was spent
in industry.51 The success of ARPA’s interdisciplinary laboratories (IDLs) in
the universities, though widely celebrated, generated concerns about the re-
lationship between solid state and physics as a whole. Harvey Brooks worried
in a 1964 letter to Walter Kohn that the growth of IDLs and applied physics
departments had exacerbated an existing tendency for solid state to isolate
itself. He was most concerned for solid state theorists, who, when colocated
with materials researchers, lost connections to other theoretical physicists,
BECOMING CONDENSED MATTER PHYSICS 163

hampering their ability to participate in the verbal exchanges that he identi-


fied as being critical to theory.52 By the early 1970s, condensed matter phys-
ics had begun to seem like an alternative that could address these concerns.

THE BRINKMAN REPORT


The years between the 1972 Bromley report and the next major survey, in
1986, witnessed systemic changes in the solid state and condensed matter
community. In the mid-1970s, the US Justice Department filed an antitrust
suit against AT&T, which led to a 1982 consent decree under which the
company relinquished control of its local telephone networks.53 The breakup
of the Bell System was completed in 1984. The turmoil within the company
in advance of the breakup spread to Bell Labs, whose scientists began an exo-
dus, scattering what had been a powerful, centralized cadre of semiconductor
physicists across many university departments and other research labs.54 The
industrial focus at Bell Labs had kept one of the country’s most influential
solid state physics groups focused on solids. Its researchers enjoyed the same
types of freedoms academic researchers enjoyed, but Bell was predisposed
to hire physicists whose basic interests aligned closely with areas of potential
application to telecommunications—new research areas of interest to con-
densed matter physicists, such as quantum fluids, held little interest for Bell.
The exodus from what had been the most powerful center of solid state phys-
ics therefore eased the transition to a broader conception of the field.
A nationwide shift away from solid state and toward condensed matter
that culminated with the breakup of the Bell group was also enacted through
changes in a number of major institutions and programs. The Aspen Center
for Physics, a prominent summer retreat for theoretical physicists begun in
1962, integrated condensed matter physics into its program in the late 1970s,
under the influence of Philip Anderson.55 In 1979, the University of Califor-
nia, Santa Barbara founded its Institute for Theoretical Physics with Walter
Kohn as the founding director. Kohn was a pioneer of density functional the-
ory (DFT), which models the electron structure of molecules and materials
using functions of electron density.56 The research program Kohn developed
using DFT emphasized its widest possible application to all phases of con-
densed matter—one measure of its success in this regard is that the Physical
Review article introducing the theory is one of the most highly cited physics
papers to date.57 The Santa Fe Institute, founded in 1984 and framing its
intellectual mission under the guidance of condensed matter physicist David
Pines, took the science of complexity as its mission.58
164 SOLID STATE INSURRECTION

These changes came about in large part because of theoretical devel-


opments. The quantum mechanical techniques of the 1940s, 1950s, and
1960s worked well for regular solids but were frustrated by other condensed
matter systems. DFT, with its remarkable generality, was a harbinger of fur-
ther theoretical advances in the 1970s. The British/Israeli physicist Cyril
Domb remembered the early 1970s as the moment when the study of critical
phenomena—such as phase transitions—gained theoretical respectability
through the application of renormalization group theory.59 It solidified its
professional respectability through conferences such as the Battelle Collo-
quium, which met in Geneva and Gstaad, Switzerland, in September 1970
to explore the relevance of critical phenomena to the science of materials.60
The Battelle Colloquium included Leo Kadanoff, then at Brown Univer-
sity, who in 1978 was hired by the University of Chicago to join its James
Franck Institute (JFI). The JFI offers a representative example of the insti-
tutional transitions occurring within solid state and condensed matter phys-
ics in the late 1970s and early 1980s. It had been founded after the Second
World War as the Institute for the Study of Metals (ISM), one of three re-
search institutes conceived to carry on the conceptual work of the Manhattan
Project research the university had hosted, albeit with an emphasis on basic,
nonmilitary research. Its early mission, “to pursue studies in the fundamen-
tal properties of matter as related to metals,” reflected the state of solid state
physics, incipient though it was, in the late 1940s, and resembled something
Roman Smoluchowski, who initially sought an APS division for metals phys-
icists that would encourage cross-disciplinary collaboration, might have en-
visaged.61 Originally directed by the metallurgist Cyril Stanley Smith, the
ISM took its cue from the study of the properties of uranium and plutonium
that the wartime Metallurgical Laboratory had conducted, minus the military
emphasis, and pursued the interdisciplinary study of metals.
By the time Smith resigned as director in 1957, that initial mission had
translated into a roughly equal balance between solid state physics and phys-
ical chemistry. The ISM’s research division employed nine physicists, eight
physical chemists, and two metallurgists, whose most lively research areas
included crystallography, magnetism, elastic and inelastic properties of met-
als, and low-temperature research, including superconductivity.62 It was, in
other words, firmly within the mainstream of 1950s solid state research. Its
emphasis on metals implied a research program on crystalline solids, and the
interdisciplinary constitution of the institution ensured that the physics con-
ducted there proceeded in close collaboration with chemistry and metallurgy.
BECOMING CONDENSED MATTER PHYSICS 165

The ISM changed its name in 1967 to commemorate emeritus professor


of chemistry, and 1925 physics Nobel laureate, James Franck, who had died
in 1964. But another purpose reflected an inkling that the institute’s histori-
cal focus on metals would soon prove limiting. An internal memo explained:
“One of the purposes for the change in name is to avoid a possible inhibiting
factor in promoting a new direction in the research to be carried on in this
facility. At the present time this research is generally limited to the areas of
chemical physics and solid state physics. It is believed to be desirable to ex-
pand the activities in these areas by developing research programs in the area
of chemical biology. The name of the Institute of Metals may tend to be an
inhibiting factor in promoting this development.”63 One of the motives for
considering a new emphasis on chemical biology was the 1964 dissolution
of the ISM’s sibling institute, the Institute for Radiobiology and Biophysics,
which prompted ISM researchers to think about moving into territory it had
once occupied.64
But the promised expansion into chemical biology was not forthcoming.
The JFI’s first few years under its new moniker were marked by the same
stringent budgetary environment that had led to the dissolution of the bio-
physics institute. Several retirements weakened its standing in solid state
physics and shifted its emphasis toward chemistry. When the physical chem-
ist Ole Kleppa took over as director, he placed an emphasis on recovering the
institute’s lost capacity in physics, “to achieve a somewhat closer numerical
balance between the physical and chemistry components of our faculty.”65
The sense that the institute would need to expand its topical scope none-
theless proved prescient. By the late 1970s, when the institute sought to ex-
pand its faculty, it took advantage of its new name’s flexibility to emphasize
recent theoretical developments in condensed matter physics. The addition
of Kadanoff in 1978 and Albert Libchaber in 1982 shifted the emphasis of
the JFI away from physical metallurgy and toward the study of critical phe-
nomena and nonlinear dynamics, fields in which it subsequently established
itself as a major national center.
The JFI’s trajectory illustrates the wider national context within which
condensed matter language won favor among those who sought to distance
the field from industry and reconnect with the pure science tradition that
remained a strong ideological force in American physics. The 1986 NRC
report on condensed matter physics—dubbed the Brinkman report after
committee chairman William F. Brinkman of Sandia National Laboratories—
addressed the problem of intellectual prestige by highlighting “the fact that
166 SOLID STATE INSURRECTION

condensed matter physics is the physics of systems with an enormous num-


ber of degrees of freedom.” As a consequence, the report maintained, echoing
the substance of Anderson’s 1972 arguments, “[a] high degree of creativity
is required to find conceptually, mathematically, and experimentally tractable
ways of extracting the essential features of such systems, where exact treat-
ment is an impossible task.” The Brinkman report went to great lengths to
identify condensed matter physics as a fundamental and intellectually valu-
able field of science, and thereby to distinguish it from materials science.
“Condensed-matter physics is intellectually stimulating,” the report empha-
sized, “because of the discoveries of fundamental new phenomena and states
of matter, the development of new concepts, and the opening up of new sub-
fields that have occurred continuously throughout its sixty-year history.”66
The choice of a sixty-year timeline for the field is telling of the authors’
agenda, especially since the Pake report committee thought Max von Laue’s
discovery of X-ray diffraction in 1912 a more appropriate landmark.67 Har-
kening back to 1926, the year in which quantum mechanics displaced the old
quantum theory, placed condensed matter firmly in the tradition of the quan-
tum theory of solids, which emerged in the late 1920s as some of the architects
of quantum mechanics explored the new theory’s utility for describing elec-
trons in metals and the importance of such applications for understanding the
foundations of the new theory.68
Precisely dating condensed matter to the advent of quantum mechanics
made it clear that condensed matter physics was defined, at core, by its in-
tellectual contributions to physics and that it was united around the meth-
ods required to describe the complex interactions of atoms and molecules
in close proximity. Solid state physicists before the 1980s preferred earlier
touchstones, mostly in the late nineteenth and early twentieth centuries.
Slater, for instance, launched an overview of the field in 1952, writing: “The
physics of the solid state is nothing new. In 1900 it was as well realized as
now that mechanics, heat, electricity, magnetism, optics, all have their solid-
state aspects.”69 Following such a strategy for condensed matter physics
would have made it more difficult to distinguish it from materials science.
Making that separation evident emphasized that condensed matter physi-
cists would not carry water for technological interests. The report began by
declaring, “We are not surveying materials science nor the considerable im-
pact of condensed-matter physics on technology.”70 By the mid-1980s, the
distinctions between condensed matter—the “fundamental” discipline—and
materials science—its applied cousin—were solidifying, while the fragile pro-
BECOMING CONDENSED MATTER PHYSICS 167

fessional alliance that had sustained solid state physics through the preceding
decades loosened.
Though distancing themselves from industrial associations, condensed
matter physicists notably did not seek a clean break with solid state physics,
choosing instead to emphasize the conceptual and methodological continu-
ities between the fields, a move made easier by the widespread impression
that solid state had been poorly named. Solid state’s technological accom-
plishments—such as the transistor, superconducting magnets, and magnetic
resonance imaging—remained integral to condensed matter’s rhetorical ar-
senal. Despite its protestations that it was not surveying the technological
dimension of condensed matter physics, the Brinkman report touted con-
densed matter as “the field of physics that has the greatest impact on our
daily lives through the technological developments to which it gives rise.” It
was not, however, within the job description of condensed matter physicists
to pursue those developments. It might be “a field whose health is essential
for maintaining the vital flow of new technology,” but realizing those technol-
ogies was to be left to other disciplines.71
The implied division of labor kept condensed matter physicists one de-
gree removed from actual technological applications, which fell to materials
scientists and engineers. This perspective presupposed something akin to the
linear model of innovation, namely the philosophy, unpopular with materi-
als scientists, that unfettered and unscripted basic research was the primary
wellspring of technological advance.72 Rustum Roy, cofounder of the Mate-
rials Research Society, called the notion that basic science begets innovation,
which in turn begets prosperity, “preposterous, certainly in league with per-
petual motion.”73 For condensed matter physicists, though, it promised to
realize their hopes of carving out a greater proportion of the federal budget
for fundamental research, especially when combined with reminders that
condensed matter physics was behind some of the most prominent technical
accomplishments of the preceding decades.
Condensed matter physicists, even in detailed surveys such as the 1986
NRC report, limited themselves to showing that basic scientific knowledge
was broadly relevant to existing areas of technological importance. They did
not spell out how funding for basic research would translate into technolog-
ical advances on the ground—and so in this sense they were not articulating
a strict version of the linear model.74 Most likely they held no deep commit-
ment to any particular vision of how scientific knowledge was connected to
technological development. Their real commitment was to the overall value
168 SOLID STATE INSURRECTION

of basic research; rhetoric establishing some connection to technology was a


tool suited to that goal, and if the connection was vague, all the better. The
growth of materials science into the development arm of the old solid state
constellation allowed condensed matter physicists the latitude to bracket the
question of how the raw knowledge mined by physicists would be refined
into usable technology as somebody else’s problem. Drawing too close an as-
sociation between basic research and applied goals tempted explicit links be-
tween basic research funding and applied targets, the very specter condensed
matter physicists hoped to avoid. They therefore walked a fine line, pressing
hard for intellectual prestige while still hoping to maintain a plausible claim
to technological relevance.
The demanding problems presented by the physical complexity of solids
were by no means unique to that phase of matter. The methods and concepts
developed to investigate and understand solids transitioned fluidly into the
broader category of condensed matter with the rise of new research areas.
As they did, the professional objectives to which they were turned changed
markedly. Solids were originally chosen as a disciplinary category because of
their shared relevance to industrial and academic physicists. In the 1940s,
that choice served the perceived need to bring industrial researchers into
the professional fold and to establish better lines of communication beyond
the academy. By the 1970s, condensed matter served the opposite impulse.
Solid state’s industrial past was a liability in the eyes of those who defended
the merit of its intellectual content, even if its technological accomplishments
were a rhetorical necessity. The category of condensed matter physics, by
returning to shared practices as a way of defining the field, aided physicists
interested in the basic problems posed by complex material systems in their
efforts to emphasize the intellectually challenging elements of their enter-
prise, but without also abrogating their claim to technological fertility.
Through the postwar decades, solid state physics could treat work on
nonsolids as relevant, but marginal. By the mid-1980s, quantum fluids, criti-
cal phenomena, nonlinear dynamics, complexity studies, and similar research
areas that could not be easily lumped in with solids were the most intellec-
tually exciting areas, at a time when emphasizing conceptual contributions
became one of the field’s most pressing goals. When these areas blossomed in
the early 1970s, they put the lie to the pessimistic outlook for solid state phys-
ics that Brian Pippard had articulated in 1961.75 But their new prominence
also strained the conceptual consistency of solid state physics, already a point
of concern, beyond credulity. Reframing much of what had been called solid
BECOMING CONDENSED MATTER PHYSICS 169

state physics as condensed matter physics brought these practices to the cen-
ter and emphasized the field’s claim on the intellectual status of American
physics, while also seeking to retain the reputation for technological fecun-
dity that solid state physics enjoyed.
9

MOBILIZING AGAINST MEGASCIENCE

Because it is so diverse and dispersed, small science has lacked the single
voice with which big science has been able to speak, thus permitting a
number of myths to persist: such as the myth of trickle-down technology;
the myth of the single intellectual frontier; and the myth of the non-zero-
sum gain. These imply that supporting big science is a good economic
investment that can be done without hurting small science.
—PAUL A. FLEURY, 1991

In 1982, the American Physical Society’s Division of Particles and Fields held
the first of what would become regular meetings in Snowmass, Colorado, to
plan their field’s future. The lofty elevation and rarified atmosphere of the
Rocky Mountains suited their ambitions. Attendees confronted the question
of where, when, and how to build the next-generation particle accelerator.
The massive machine they envisioned would have been the largest physical
laboratory ever constructed, and the most expensive. The reductionist ideal
that had driven high energy physics since the 1960s guided its conception
and mission. Almost a century after Henry Rowland articulated his pure sci-
ence vision, the high energy physics community was poised to push it to its
logical extremes with a taxpayer-funded effort to probe basic questions about
the structure of the universe, with scant concession to technological or eco-
nomic considerations.
The accelerator imagined at Snowmass would require so much cheap,
flat, and sparsely populated space that it would by necessity be a “machine-
in-the-desert,” as Leon Lederman—who would win the Nobel Prize in 1988
for his neutrino research—put it.1 Following Lederman’s lead, Snowmass
participants began calling this imagined machine “the Desertron.”2 News
outlets seized on the flashy name for what would eventually become the ill-
starred Superconducting Super Collider, or SSC. Time, The Economist, New
MOBILIZING AGAINST MEGASCIENCE 171

Scientist, and others reported breathlessly after Snowmass 1982 on how the
Desertron, a multibillion-dollar particle physics dream machine, would re-
shape our understanding of the universe.3 The mammoth accelerator, with a
beam pipe 54.1 miles in circumference, would eventually be sited in Waxa-
hachie, Texas, south of Dallas-Fort Worth.
From the standpoint of 1980s particle physics, an accelerator surpassing
the energies available at Fermilab’s Tevatron was necessary to provide ex-
perimental grounding for the final pieces of the standard model of particle
physics and to explore its limits. The target energy for the SSC was 20 TeV,
below which particle physicists were certain the Higgs boson, one of the last
pieces of the standard model that awaited experimental discovery, would be
found. Steven Weinberg justified the 20 TeV target to Congress with an anal-
ogy to Christopher Columbus: “It is a little bit like Columbus sailing west
from Spain. Columbus promised that he would get to the Indies if he could
sail far enough West. Well, that was wrong, but what he should have said,
which would have been correct, is that if he sailed far enough West, he would
get to the Indies unless something equally interesting got in the way.”4 The
20 TeV target guaranteed finding the Higgs, unless the machine first found
physics that broke the standard model and forced a radical reimagining of the
field. High energy physicists hoped for the Higgs, and found the possibility
that the journey to 20 TeV would be interrupted by undiscovered continents
even more tantalizing.
But where Weinberg saw Columbus, others saw Burke and Wills, the
Australian explorers who succeeded in crossing the continent from south
to north, but found little nourishment in the empty desert between the
coasts and died trying to complete the return journey.5 Some conceptions
of the standard model suggested that, although the journey to 20 TeV would
surely reveal the Higgs, little else of interest was likely to appear along the
way, and that the Desertron could therefore be expected to map out a vast,
empty conceptual desert. Stumbling upon an unadorned Higgs as an oasis
amid a wasteland came to be known as the “nightmare scenario” for particle
physics, and it appears to be playing out now at the Large Hadron Collider,
where a Higgs boson consistent with the standard model was uncovered in
2012.6 Uncertainty about the potential of a 20 TeV accelerator to do more
than find the Higgs (and perhaps the top quark, which would be detected
in 1995 at the Tevatron) opened the door to other objections, most nota-
bly from condensed matter physicists who complained that the Desertron
promised a funding desert for physicists pursuing fundamental research in
172 SOLID STATE INSURRECTION

anything other than elementary particle physics. The objections of physicists


who worried about a big-science monopoly on federal spending for basic re-
search contributed to Congress yanking the SSC’s funding, even after con-
struction on the Waxahachie site was well under way, in 1993. Unlike the
earlier internecine squabbles that had shaped post–Second World War pres-
tige hierarchies and funding dynamics, which tended to remain largely within
the physics community, these debates played out in public, where arguments
about the SSC contested much more than the merits of a single facility. The
SSC debates, as seen through the congressional record, popular books, and
op-ed columns, constituted a referendum on the future of American physics.
The referendum was high stakes. The success of a multibillion-dollar ac-
celerator, conceived at unprecedented scale, would have reaffirmed the sta-
tus of high energy physics as the marquee American science in the face of
over a decade of challenges to its privilege. The Atomic Energy Commission
(AEC), whose closed-door decision-making process favored the whims of
influential individuals, was dissolved in 1974. As Michael Riordan, Lillian
Hoddeson, and Adrienne Kolb document in Tunnel Visions, their definitive
history of the SSC, the Department of Energy, which took the AEC’s place
and approved funds through the regular congressional appropriations pro-
cesses, began a slow erosion of the influence high energy physicists held over
the federal funding of science.7 The 1970s also saw the birth of recombinant
DNA techniques and the 1980s launched with the passage of the Bayh–Dole
Act, which permitted universities and other organizations to file for patents
on intellectual property deriving from federally funded research. In conjunc-
tion, these developments encouraged the commercialization of academic mo-
lecular biology and set the groundwork for biomedicine to expand both its
public profile and its funding portfolio. Biology began to claw its way up the
institutional hierarchies of American universities with the purchase offered
by the promise of patent revenues.8 The SSC’s failure affirmed these trends,
whereas its success might have forestalled them.
The factors that led to the SSC’s demise included lack of international
collaborators, internal mismanagement, cost overruns that placed the final
estimated price tag at over $11 billion, and the end of the Cold War, with
the incentive it had provided to outcompete the Soviet Union in all things.
This case study does not offer a full accounting of these factors, but instead
examines the SSC from the perspective of solid state and condensed matter
physics.9 What did the SSC mean for condensed matter physics and what
did condensed matter physics mean for the SSC? Condensed matter physi-
MOBILIZING AGAINST MEGASCIENCE 173

cists tended to oppose the project, and they did so with greater intensity, fre-
quency, and volume through the late 1980s and early 1990s. Congressional
budget hawks, whose numbers increased after the 1992 elections, considered
the SSC profligate and found value in such testimony. High energy physicists
bristled. They felt blindsided by their colleagues’ attacks on a project they
viewed as necessary to advance fundamental knowledge, and to ensure fed-
eral support of basic research of any kind.
Vocal, public opposition to so high-profile a project from within phys-
ics was irregular. Dissent, in the unspoken rules of the physics community,
stayed in the family. Mobilizing that dissent in order to influence those hold-
ing the federal purse strings would be perceived as a betrayal. Solid state
physicists, like Benjamin Lax, might have harbored reservations about the
merits of Fermilab when it was taking shape in the late 1960s, but they were
cautious about voicing them in public. Even at the peak of his frustrations
about tepid federal support of the National Magnet Laboratory, Lax avoided
venting them in the direction of any particular project, at least in writing.
Alvin Weinberg, although his criteria for scientific choice were unfavorable
to high energy physics, also kept his critiques on the general level and did
not seek to undermine any specific funding request. So when Senator John
O. Pastore asked Robert Wilson in a congressional hearing about appropri-
ations for the National Accelerator Laboratory, “Would you say as far as you
know, the whole scientific community is behind this, without a dissent?” Wil-
son could reply, “They do not dissent to me, sir.”10 The SSC would not enjoy
the same deference. Resistance from condensed matter physicists and mate-
rials scientists who considered the SSC an extravagant toy for the amusement
of a small slice of physicists was prevalent and public.
That opposition signaled the boiling over of tensions that had strained
the American physics community for half a century. Despite the influence of
the Cold War security state, the rapid expansion of solid state and condensed
matter physics, and the increasing presence of physicists in industry, the
pure science ideal lived on in high energy physics, where it combined with
a strong reductionist philosophy. By the time of the SSC hearings, however,
condensed matter physics had synthesized a clear alternative: an embrace of
basic condensed matter research as a more probable foundation for future
technological benefits than other research as well as a good in itself. This
competing ideal proved better adapted to the context of the early 1990s. After
the demise of the SSC delivered a blow to the reductionist spin on the pure
science ideal that had sustained high energy physics for decades, the con-
174 SOLID STATE INSURRECTION

densed matter physicists’ view of physics remained as the vision that physi-
cists would have to adopt to continue federal support for their field.
The following is organized around the three myths Paul A. Fleury, a
physicist then at Bell Laboratories, identified in his congressional testimony
opposing the SSC: the myth of trickle-down technology, the myth of the
single intellectual frontier, and the myth of the non-zero-sum gain.11 These
capture the threefold objection that solid state physicists posed to the SSC
and to the rhetoric in its favor: first, that high energy physicists were claim-
ing spin-off technologies for the SSC that should more rightly be credited
to condensed matter physics; second, that high energy physics was not the
only route to fundamental knowledge; and third, that a single, multibillion-
dollar laboratory serving a small minority of physicists was not the best way
to ensure future advances of either technology or fundamental knowledge.
These objections derived from the alternative vision of physics and its place
in American society that solid state physicists had developed by the time of
the SSC debates. Analyzing these objections as they were deployed in the
emotionally charged battle over the SSC’s fate brings into focus the central
ideological disagreement that defined American physics at the end of the
twentieth century and demonstrates how solid state physics, in its first half
century as a part of American physics, shifted the field’s center.

THE MYTH OF TRICKLE-DOWN TECHNOLOGY


Spin-off claims were part of the standard rationale for the SSC. Although
high energy physicists had long invoked incidental technological develop-
ments as a fringe benefit of their research, those justifications were firmly sub-
ordinate. High energy physicists, when they suggested that spin-offs should
be factored into the value of fundamental science, traditionally took pains
to emphasize that they could not predict or guarantee specific spin-off tech-
nologies.12 “To extrapolate from the recondite topics of current fundamental
scientific investigation to a technological spin-off is to indulge in grandiose
speculation. Responsible colleagues shirk the task almost as a tradition,”
Leon Lederman wrote in a 1984 Scientific American article presenting his
arguments for supporting fundamental research to a popular audience.13
But spin-off claims would assume a new and outsized importance during the
push to get the SSC built.
By the late 1980s, the Cold War, nearing its end, no longer warranted the
hitherto potent assumption that all physics was potentially defense-relevant,
and certainly being pursued by the Soviet Union. Economic competitiveness
MOBILIZING AGAINST MEGASCIENCE 175

had taken the place of national defense as the standard by which federal ex-
penditures on science would be measured.14 Economic and technological
spin-off claims grew into a new role as an answer to this challenge. The exact-
ing demands of accelerator design and construction, SSC boosters argued,
would advance existing technologies, uncover novel applications, and gen-
erate new job-making industries. Specific promises—which some of Leder-
man’s colleagues were willing to make despite his insistence on their fruit-
lessness—included better magnet and superconducting technologies such as
those that powered MRI (magnetic resonance imaging) machines, advances
in computing, and even improved tunneling techniques that could be applied
to the construction of subway systems. Such benefits could, the argument
went, be expected to offset the up-front cost of the project in the long term.
But spin-offs, high energy physicists cautioned, were often unanticipated; un-
foreseen benefits should factor into the equation as well.
The site selection competition that ran from early 1987 to late 1988
did a considerable amount to make spin-off claims more prominent. Mich-
igan representative Milton Robert (Bob) Carr put the point succinctly in a
March 1988 House Appropriations Committee hearing. “That’s not the way
the Governors are looking at it,” he responded to Secretary of Energy John
Harrington’s insistence that the SSC was a basic research project first and
foremost, “they’re looking at the jobs and the economic development and
the spinoffs.”15 Coaxing support for siting proposals from state governments
required articulating clear and concrete benefits to the region, and these ar-
guments percolated into the overall rational for building the SSC. Through
this process, spin-offs were asked to carry a justificatory role they could not
sustain, opening a line of attack for condensed matter physicists who bridled
at the suggestion that their own field’s technical contributions were mere fall-
out from the accelerator explosion.
Legislators’ reactions to spin-off claims reveal some of their pitfalls. In a
March 5, 1986, subcommittee hearing on the Department of Energy (DOE)
budget, the chairman, Florida Democrat Don Fuqua, asked Alvin Trivelpiece,
director of the DOE’s Office of Energy Research: “There appears to be some
confusion between the Department’s High Energy Physics Program and the
Strategic Defense Initiative. We had a witness yesterday that said high energy
physics was mostly SDI. Can you kind of clarify the relationship between
those two programs?” The assertion that the DOE’s high energy physics pro-
gram was bound up in the Strategic Defense Initiative (SDI)—better known
by the sarcastic epithet “Star Wars,” the missile defense program that by the
176 SOLID STATE INSURRECTION

end of the 1980s would become a paradigm boondoggle—wrong-footed


Trivelpiece, who replied: “I have no idea what the other witness may have
had in mind. It’s difficult to imagine putting Fermilab into orbit, but perhaps
we could give a try. Sometimes some of the individuals there seem to be in
orbit, but you can hear more about that from Leon [Lederman].”16
The statement to which Fuqua referred had come from Charles J. Mankin,
the state geologist of Oklahoma, who on March 4 mistakenly conflated the
portions of the DOE budget dedicated to high energy physics and SDI in
his plea for more federal funding for fossil fuel research.17 But Trivelpiece
himself might have contributed to Fuqua’s associating SDI with the SSC. In
his own testimony leading up to Fuqua’s question, he had said: “The cur-
rent collection of activities that are in the Strategic Defense Initiative to some
extent owe their existence to the basic research that has been done in high-
energy physics, nuclear physics, and the fusion program, and basic energy
sciences in other parts of the Department. So there is a lot of contemporary
technology transfer.”18 Fuqua’s misunderstanding illustrates two important
pitfalls of spin-off claims. First, legislators often failed to distinguish between
different types of expertise when evaluating expert testimony, which made
it difficult for high energy physicists to place spin-offs in the context they
desired. Mankin, a geologist, could not be expected to offer expertise on high
energy physics spending, but through the grind of long committee hearings,
his testimony was admixed with those of other witnesses and the precise na-
ture of his expertise lost relevance. Second, the more central spin-offs be-
came, the more some legislators began to understand the principle purpose of
the SSC to be the generation of new applications, weakening the intellectual
justification for the project.
Many high energy physicists were therefore ambivalent about spin-off
claims, and became more so in the late 1980s and early 1990s as the political
heat ratcheted up and it became difficult to sustain the idea that an increas-
ingly expensive SSC was a better bet than direct investment in the techno-
logical arenas it would purportedly advance. They recognized spin-offs as
useful if presented as a bonus Congress could expect for free if it supported
the public good of fundamental research, but the more central the notion of
spin-offs became to the overall justification for the SSC, the less the project
seemed worth the price tag. Congress would have to be convinced to pay for
the goods; they were unlikely to shell out for the lagniappe.
Spin-off claims nonetheless proved hard to contain. Legislators often
pushed for specific benefits to their home states or to the nation. Pointing to
MOBILIZING AGAINST MEGASCIENCE 177

a track record of technological and medical advances linked, however weakly,


to high energy physics, whetted legislators’ appetite for pork. Many SSC ad-
vocates, including SSC director Roy Schwitters, perceived that high-minded
declarations of the ennobling nature of fundamental knowledge would not
suffice to convince legislators that the project was worthwhile. “Elementary
particle physics does not exist in isolation,” Schwitters asserted in his writ-
ten statement for a 1989 Senate subcommittee meeting on the DOE budget:
“Stimulation, information, and techniques flow both ways: from other activ-
ities toward particle physics, and from particle physics toward other activi-
ties.”19 He emphasized intellectual overlap with nuclear physics, cosmology,
and condensed matter physics, and stressed that the demands of accelerator
engineering had technological ramifications for computing, superconducting
magnets, and semiconductor devices. The cutting-edge technical needs of
accelerators, Schwitters maintained, “provide fruitful interchange with other
researchers, manufacturers, and developers of technology in fields such as
medical imagery.”20
Cosmologist George Smoot, testifying in a 1992 Senate hearing, re-
counted an anecdote about sharing a plane with a group of cataract patients
traveling for laser surgery only available in the United States, which he used
to illustrate the claim: “It just shows you do not know what development is
going to turn out to be something useful. . . . You really have to understand
the basics of all science. That is where physics, and particularly higher energy
physics comes together because physics is what we call the queen of sciences.
It is the basic underlying structure for all of sciences, the foundation every-
thing sits on. You have to understand physics to understand what is going
on.”21 Such treatment was available, he implied, and available in the United
States, at least in part because of robust support for basic research in the form
of high energy physics.
But tying the supposed generative power of high energy physics to its sta-
tus as the science of the fundamental scale proved contentious, even among
SSC advocates. Smoot, for example, testified immediately after Lederman,
who dissociated the reductionist justification from spin-off claims: “[Tech-
nological benefits] would be a crazy reason to build the SSC. We do not
build it for the spin-offs. We build it because we are humans who think and
are insatiably curious, and have an unquenchable determination to know,”
recalling Robert Wilson’s apology for the National Accelerator Laboratory
twenty-five years earlier.22 Lederman’s argument contrasts the willingness of
his younger colleagues to adopt spin-off arguments when convenient. The
178 SOLID STATE INSURRECTION

split might well have been generational. Schwitters and Smoot were both
born in the mid-1940s and earned their doctorates in the early 1970s. In
contrast, Lederman, born in 1922, along with another advocate for the un-
embellished reductionist justification, Steven Weinberg, born in 1933, were
among the generation who had overseen the articulation of the philosophy
in the 1960s. For them, stooping to arguments on the basis of technological
output weakened the justification for pursuing fundamental physics for its
own sake. The high stakes of the SSC debates influenced these two groups
differently. On one hand, it promoted an intensification of the reductionism
that had underwritten particle physics’ push to higher energies through the
1960s and 1970s. At the same time, suspicions that such a justification would
not work on its own prompted younger physicists to advance spin-off claims,
which Lederman, Weinberg, and other members of the old guard found dis-
tracting and at times counterproductive.
The persistence of spin-off claims, especially those related to supercon-
ductors, semiconductors, nuclear magnetic resonance, and other elements of
the solid state domain, stoked the fires of opposition among condensed matter
physicists. The very spin-offs that high energy physicists were claiming for
accelerator development, they insisted, were actually the result of research in
solid state physics, which nevertheless faced limited resources for basic re-
search that were likely to become leaner in the shadow of the SSC. This variety
of resistance made an impression on legislators, who responded by pushing
back on spin-off claims and asking high energy physicists to articulate specific
targeted outcomes, which was anathema to the goals of the project.
Some of the most damning testimony came from Nicolaas Bloembergen, a
Harvard-based condensed matter physicist and 1991 president of the Amer-
ican Physical Society. Bloembergen had won a share of the 1981 Nobel Prize
in Physics for his work on laser spectroscopy and had also contributed to the
early work on nuclear magnetic resonance that led to MRI techniques.23 The
MRI was probably the most-cited spin-off adduced in favor of high energy
physics research during the SSC hearings. It connected the otherwise remote
phenomenon of superconductivity, which was, after all, part of the project’s
name, to a well-known medical technology of obvious humanitarian benefit.
But Bloembergen assured the Senate: “As one of the pioneers in the field of
magnetic resonance, I can assure you that these [MRI and superconducting
magnets] are spinoffs of small-scale science, and not of the SSC.”24 He also
worked behind the scenes to try to tamp down such claims. In a sharp letter
to Fermilab’s Richard A. Carrigan Jr., Bloembergen chastised SSC support-
MOBILIZING AGAINST MEGASCIENCE 179

ers for telling Congress that MRI was a spin-off of high energy physics. Such
a claim, he wrote, was both “unwarranted and ill-advised. It completely ig-
nores the essential contributions by a very large number of physicists who
have brought MRI to fruition. MRI would be alive and well today if Fermilab
had never existed.”25 The letter found its way into the congressional record
juxtaposed with a ferocious missive by Michigan Democrat Howard Wolpe
and New York Republican Sherwood Boehlert, both members of the House
Subcommittee on Investigations and Oversight, castigating the Department
of Energy for misleading Congress about the SSC’s cost and timeline.26
Bloembergen was not the only APS president to throw the clout of his of-
fice behind his objections to the SSC. In March 1991, Cornell’s James Krum-
hansl published a piece in Physics Today based on his outgoing presidential
address at the 1990 APS meeting in Washington, in which he made veiled
references—legible to his audience as references to the SSC—to the danger
of hyperbole and failure to assign appropriate credit when evangelizing for
one’s field.27 The speech itself came only two months before Krumhansl
wrote to journalist Malcolm Browne, who himself had published an editorial
in the New York Times raising doubts “as to whether the new knowledge it
[the SSC] generates will be commensurate with the enormous cost.”28 Krum-
hansl’s letter charged that “extravagant representation to the public of the
potential fruits from the SSC was fictitious and ethically irresponsible and
that accurate acknowledgement was not given to researchers in many other
subfields of physics which were the true source of contributions from phys-
ics to medicine, technology, economics and education but imputed to par-
ticle physics.”29 Materials Research Society cofounder Rustum Roy, for his
part, invoked Alvin Weinberg’s criteria for scientific choice in 1993 to insist:
“Nothing the speculative science the SSC can discover can ever have any
impact on chemistry, biology, engineering science, materials, agriculture.”30
As the SSC’s budget ballooned in the last years of the 1980s and into
the early 1990s, skepticism about spin-off claims grew in popular forums as
well. Historian of science Alexi Assmus wrote in an op-ed for the Wall Street
Journal, “In fact, the SSC uses old technology: the name ‘superconducting’
in SSC refers to helium-cooled superconducting magnets that have been
used for more than 30 years, not to the new high-temperature superconduc-
tors that have recently been discovered by condensed matter physicists.”31 A
widely syndicated column by Los Angeles Times contributor and technology
consultant Michael Schrage skewered spin-offs in August 1992, writing that
the rosy vision of regional and national economic revitalization that often ac-
180 SOLID STATE INSURRECTION

Figure 9.1. The Supercompliant Superprovider, 1993. John Trever cartoon depicting
the disconnect between high energy physicists’ expectations and federal priorities.
Credit: Copyright 1993 by John Trever, Albuquerque Journal. Reprinted by permission

companied panegyrics for the SSC “has not a neutrino of truth to support
it.” Pork, not technological progress, was the real spin-off in Scharge’s eyes.32
Spin-off claims became more and more associated with the “quark-barrel
politics” that the New Republic had derided during the site-selection process,
and that contributed to the schadenfreude that accompanied the project’s
cancellation (figure 9.1).33
The collapse of the spin-off narrative became a convenient cudgel with
which skeptical legislators could hammer the project. Representative Vir-
ginia Smith, a Nebraska Republican, asked James F. Decker, acting director
of the Office for Energy Research, in March 1988: “I note that your budget
justification, Dr. Decker, says ‘The SSC will have discoveries, innovations
and spinoffs that will profoundly touch every American.’ I come from the sec-
ond most agricultural district in the Nation. Could you tell me just how the
SSC will profoundly touch every farmer in Western Nebraska?”34 Boehlert,
the most colorful of the SSC’s congressional opponents, was more pointed in
May 1993: “None of them [spin-off claims for the SSC] are a result of what is
going on with the SSC project, not one single one, and the SSC is not going
MOBILIZING AGAINST MEGASCIENCE 181

to feed the hungry people of Somalia, and it is not going to end the genocide
in Yugoslavia. It is just eating up more of our resources.”35
The offense condensed matter physicists expressed about spin-off claims
that trespassed on their turf both galvanized their opposition to the project
and connected effectively with legislators. Both are evident in Paul Fleury’s
identification of spin-off claims as “the myth of trickle-down technology.”
The term carried a specific ideological resonance in the early 1990s. It in-
vited comparisons between the funding structure of high energy physics and
“trickle-down economics,” a pejorative commonly hurled at the supply-side
fiscal policies of the Ronald Reagan and George H. W. Bush administrations.
The analogy implied that the interests of high energy physicists were just as
remote from the needs of technology, economy, and other areas of science as
the interests of large corporations and the wealthy magnates who ran them
were from the needs of the middle-income Americans to whom politicians
reliably pandered. Painting the SSC as remote was integral to the case against
it, and it extended also into condensed matter physicists’ objections to the
knowledge claims that high energy physicists made on behalf of the machine.

THE MYTH OF THE SINGLE INTELLECTUAL FRONTIER


Although divided about the merits and efficacy of spin-off claims, high en-
ergy physicists—and cosmologists, who were their close allies throughout
the hearings—were uniformly of the opinion that the SSC was valuable be-
cause it could offer fundamental knowledge where other facilities, and other
branches of physics, could not. They maintained the belief, summarized by
Silvan  S.  Schweber, that “elementary particle physics has a privileged po-
sition, in that the ontology of its domain and the order manifested by that
domain refer to the building blocks of the higher levels.”36 Most practicing
physicists who testified on the project’s behalf recognized the limitations of
spin-off claims, and the consequent need to push as hard as possible on the
intellectual necessity of the SSC. Reductionism, already the unifying phil-
osophical thrust of high energy physics, would be hammered to an even
sharper point in the forge of Congress, prompting condensed matter phys-
icists to deploy their emergentist philosophy in response.
High energy physicists presented the strong reductionism that was re-
quired to create a sense of urgency around the SSC to Congress wrapped
with a sense of the antiquity of the reductionist enterprise and tied with a
bow of reverent rhetoric about the importance of discovering the universe’s
182 SOLID STATE INSURRECTION

most deeply hidden secrets. Leon Lederman commonly employed the strat-
egy of casting the SSC as the culmination of a narrative beginning in An-
cient Greece: “The road from Miletis [sic] to the SSC is what philosophers
call a reductionist road. . . . Until we can complete the unification process
and make the picture mathematically whole, the question of how the world
works will not be answered.”37 Presenting the SSC as the apotheosis of a two-
millennium-old quest added to the sense that it contributed to some consti-
tutional human desire for meaning. Burton Richter promised that, should the
SSC succeed, “You’d know everything there was to know about our physical
world, and you would know much better what man’s place is in that physical
world. And that is spiritual, intellectual, what-have-you, that kind of knowl-
edge and satisfaction.”38
But in the late 1980s and early 1990s, the priestly mien high energy
physicists had adopted successfully for much of the Cold War was losing its
potency. Humor writer Dave Barry, in his syndicated column in November
1987, skewered the SSC’s cost overruns alongside the remote nature of the
physics it would pursue and the high-minded justifications physicists gave
for it. He suggested that the “giant underground racetrack for invisible par-
ticles” was conceived only after “a 400-foot-long nuclear-powered undersea
saxophone” was deemed “too practical.” “Needless to say,” Barry wrote, “the
Superconducting Super Collider concept was conceived by research scien-
tists, who are driven, as always, by a burning desire to push back the frontiers
of obtaining federal funds.”39 In defiance of such backlash, those high energy
physicists who fought shy of spin-off claims pushed to make the intellectual
case for the SSC as strong as possible, which meant embracing a strong form
of reductionism.
Steven Weinberg provided the most consistent and impassioned defense
of the strong reductionist position. The 1979 Nobel Prize had recognized
his work unifying electromagnetism and the weak nuclear force, and he was a
visible public advocate for reductionist science, writing a popular book that
expressed optimism for the culmination of physics with a unified physical
theory.40 Throughout the hearings, his broader view of science and specific
justifications for building the SSC worked in lockstep. The case he offered
August 1993 aptly summarizes his position and is notable for the direct con-
trast presented by Philip W. Anderson, who testified immediately after him.
“We are at the frontier,” Weinberg testified, “we have pushed the chain of
questions why as far as we can, and as far as we can tell we cannot make any
progress without the super collider.” He continued: “Well, who cares? You
MOBILIZING AGAINST MEGASCIENCE 183

know, there are a lot of people, a lot of Americans, a lot of members of Con-
gress who really see science only in terms of its applications. And that is a
respectable view. Not everyone is turned on by the same things. Not everyone
likes classical music. Not everyone has this hunger to know why the world is
the way it is, and we have to live with that. I find it sad, myself, but that is the
way people are.”41
By identifying those who cared only about applications as the SSC’s
main opponents, Weinberg disregarded the objection Anderson and others
mounted that particle physics was not, in fact, the only route to fundamen-
tal physical insight. Weinberg’s testimony throughout the SSC hearings as-
sumed that “without this machine we simply cannot continue the process of
uncovering nature’s fundamental laws.”42 His claim was that the United States
should fund the SSC not only because it provided fundamental knowledge
but also because it was the only route to fundamental knowledge; everything
else was derivative. By the early 1990s, the state of the art in reductionism was
substantially more virulent than it had been in the 1960s and 1970s.43 Victor
Weisskopf considered particle physics the science most fully directed toward
fundamental principles, but he saw it as occupying one extreme of a smooth
scale rather than as a categorically unique enterprise.
The psychological effectiveness of grouping fundamental science with
pro-SSC views and applied science with anti-SSC views became evident
immediately after Weinberg’s August 1993 testimony when the committee
chairman, Senator J. Bennett Johnston Jr., a Louisiana Democrat, introduced
the next witness: “a distinguished Nobel laureate. Professor Philip P. Ander-
son from the Department of  Physics, I think that is Applied Physics, at Prince-
ton University.” Departing from his prepared testimony, Anderson offered
two corrections for the record: “Senator Johnston and this committee, and
the Democratic National Committee are the only two people who give me
the middle initial ‘P’ when my actual middle initial is ‘W.’ And I receive a
lot of mail from the Democratic National Committee to Phil P. Anderson.
And I am not an applied physicist. I like to call myself a fundamental physi-
cist as well. I just am fundamental in somewhat different ways.”44 Anderson
endeavored to undermine Weinberg’s hard and fast distinction between re-
ductionist fundamental physics and applied physics, just as he had sought to
undermine Weisskopf ’s distinction between intensive and extensive research
two decades earlier. Johnston’s initial error provided him an opportune segue
into that argument.
Anderson parsed his objections to the SSC in terms of how particle phys-
184 SOLID STATE INSURRECTION

ics fit into his own broader view of science. Before the House of Representa-
tives Task Force on Defense, Foreign Policy and Space in 1991 he delivered
the same message he would give the Senate:
The standard testimony you will receive on behalf of the SSC will tell you
that in some sense elementary particle physics, high-energy physics is the
bellwether of the sciences, the one which is out there leading the pack, the one
which in some sense is investigating the “deepest” layers of reality in the world
around us and the “most fundamental” laws of physics. . . . There is at least
one other kind of frontier in the physical sciences where a lot of action—and I
would argue more action—is taking place: the frontier of looking at bigger and
more complex aggregates of matter which often behave in new ways and ac-
cording to new laws. These new laws don’t contradict the laws the elementary
particle people discover; they are simply independent of them, and I would
argue they are in no way any less or any [more] fundamental.45

Given the expense of particle physics, Anderson’s testimony continued, what


does it contribute in proportion to the funds it receives? How many scientists
does it employ? Does the knowledge it produces either motivate practical
applications or stimulate research in other areas of science? The answers to
these questions should, in his view, determine how science is funded. The
implication was clear: if fundamental knowledge could be had for a fraction
of the price of the SSC, why spend billions? As condensed matter physicist
Theodore Geballe put it, “You are exploring just as much of an unknown,
but what you find out is more relevant.”46 Whereas in the 1970s, Anderson
had distanced himself from the Weinberg criterion for fear of associating solid
state physics too closely with technology, the pressures of the SSC debates
caused him and his allies to augment their pluralistic view of fundamental
research with arguments for the proximity of their work to applications.
Weinberg and Anderson each had agendas that transcended the SSC and
framed their testimonies. Congress therefore heard competing views of sci-
entific research and its goals as much as they heard arguments for or against
the SSC. That more was at stake than a single research facility is evident from
the extent to which the SSC debate spilled over into the popular press. Both
Weinberg and Lederman published popular books espousing a reductionist
view of science.47 Weinberg took his case to the pages of the New York Times
in March 1993 as the SSC’s prospects grew dire. In an op-ed entitled “The
Answer to (Almost) Everything,” he appealed to a popular audience with an
example about the weather, writing: “Elementary particle physics is more
MOBILIZING AGAINST MEGASCIENCE 185

fundamental than, say, meteorology, not because it will help us predict the
weather, but because there are no independent principles of meteorology that
do not rest on the properties of elementary particles.”48
The op-ed angered condensed matter physicists, some of whom shot
back. Northwestern University’s Pulak Dutta retorted in a letter to the editor:
“The public relations triumph of particle physics is that it has cast itself as
the sole heir of atomic physics and quantum mechanics, and thus irrefutably
‘fundamental.’ However . . . we’ve known for some time that elementary par-
ticles are made of quarks, but that hasn’t made (and isn’t likely to make) any
difference to any other area of human activity. It just isn’t fundamental to any-
thing.”49 Dutta’s rehearsal of Alvin Weinberg’s criterion for scientific choice
in the face of Steven Weinberg’s high-proof reductionism reflected the mood
among condensed matter researchers. They viewed high energy physicists
as demanding extraordinary resources while maligning the intellectual merit
of their own endeavors and making unrealistic spin-off claims for a field with
little measurable importance to other areas of science and few prospects for
making socially or technologically useful contributions.
Though reductionist rhetoric had helped secure Fermilab’s funding in
the 1960s, it was much less effective with the 1990s Congress, in part because
the objections of condensed matter physicists provided political cover. Their
arguments were echoed in the statements of the SSC’s legislative opponents.
Sherwood Boehlert, chairman of the House Committee on Science, Space,
and Technology, opened a May 1993 hearing by rehearsing arguments solid
state physicists had presented: “My first concern is the basic question of pri-
orities. SSC supporters like to suggest that to oppose the SSC is to oppose
science. Nothing could be further from the truth. Science is not some indivis-
ible domain but is made up of separate, if related, disciplines.”50 Similarly, in
the Senate, Democrat Dale Bumpers of Arkansas brusquely dismissed the ar-
gument that particle physics and basic science were one and the same: “The
assumption that anybody who opposes this project is opposed to basic sci-
ence is a distraction and a diversion.”51 The competing view that condensed
matter physicists offered helped legislators oppose a major scientific budget
item without appearing to be anti-science.
The internecine sniping over the SSC ignited a dialogue within physics
about the unity, or disunity, of the field. In Physics Today, the widely distrib-
uted news magazine published by the American Institute of Physics, discus-
sions of the unity of physics paralleled the debate that played out in congres-
sional testimony and popular writings.52 In a letter published in March 1991,
186 SOLID STATE INSURRECTION

Lawrence Cranberg, a retired University of Virginia physicist who studied


the physics of wood as it applied to domestic energy needs, wrote that parti-
cle physicists showed little interest in work not directly related to their own.
He charged that “the quest for unity has become a specialty that narrows
so intensely the intellectual focus of its devotees that they are unwilling to
be interested in anything else in physics,” and asked, rhetorically: “Is that
what we want to encourage when we speak of ‘the unity of physics’? Or does
such ‘unity’ condemn one to a snobbish isolation from the mainstream of sci-
entific and human concerns?” Cranberg advocated accepting diversity as an
equally potent ideal.53 His criticisms of the high energy physics agenda show
that the same tensions that arose over high-stakes projects like the SSC were
much deeper, shaping discussions of the discipline’s direction throughout
the physics community.
The same issue of Physics Today featured two articles from condensed
matter physicists arguing for the importance of higher-level phenomena,
shoring up the foundations of public arguments against the SSC, and favor-
ing increased support for smaller projects. In the first, University of Chicago
theorist Leo Kadanoff wrote on self-organization in physical systems, de-
scribing how plumes develop in heated fluids and finding “many different
laws and many different levels of description.”54 Kadanoff’s focus on com-
plex phenomena complemented the perspective he had presented as a reg-
ular contributor to Physics Today’s “Reference Frame” op-ed page since the
mid-1980s. In 1986 he praised the ability to discern between practical and
impractical demands for understanding and controlling complex systems as
a key component of scientific judgment.55 In 1988 he brought his notion of
good scientific judgment to bear on contemporary trends in government sci-
ence funding: “The true value of science is in the development of beautiful
and powerful ideas. Overinvestment in big science detracts from what is re-
ally worthwhile. I do not think that the nation’s or the government’s budget
for research or for R&D is too small. It is, however, increasingly misdirected
toward grandiose projects. We physicists have a responsibility to understand
what is truly valuable in science and use this understanding to help the nation
develop and express its priorities.”56 Close links between intellectual merit
and broader relevance, of the type Anderson expressed before Congress and
Kadanoff articulated in Physics Today, became critical to the condensed mat-
ter community’s outlook on the organization of science in America in the
1980s and 1990s.
Unity, for condensed matter physicists, did not mean reduction, but
MOBILIZING AGAINST MEGASCIENCE 187

rather the relevance of knowledge at all levels to the whole of physics. Ander-
son wrote in Physics Today in July 1991: “With the maturation of physics, a
new and different set of paradigms began to develop that pointed the other
way [from reductionism], toward developing complexity out of simplicity.”57
Diversity and complexity, for solid state researchers working in the shadow
of particle physics, were critical prerequisites for unity. They indicated that
physical knowledge was interdependent rather than hierarchical, which
cemented their claim to fundamental knowledge.58 Throughout the SSC
debates, this shared belief gave condensed matter researchers a singleness
of purpose that their factionalized field had found wanting for much of its
history.
By the 1990s, solid state physicists were expressing a consistent view
of fundamental knowledge, at least for a national audience. It was similar
to what Francis Bitter and Alvin Weinberg described decades earlier. Solid
state physicists championed a permissive approach that promoted research
across disciplinary boundaries and encouraged diverse applications. With-
out the reductionist hegemony and its influence on government science fund-
ing, though, the position championed by the condensed matter community
would never have developed to the extent it did. Virulent reductionism, and
the success of its proponents, forced solid state physicists to develop their
views on why knowledge of complex systems was not less fundamental than
knowledge of simple systems, and on how money and prestige should be dis-
tributed accordingly. Anderson’s “More Is Different” responded to brewing
discontent in the solid state community while physics was at the height of
its prestige. Likewise, the resurgence of the Weinberg criterion among con-
densed matter physicists, Anderson included, during the SSC debates was
driven by angst over conditions that might allow the SSC to dominate the
funding landscape as physics as a whole watched its cultural prestige wane.

THE MYTH OF THE NON-ZERO-SUM GAIN


Even if they acknowledged that spin-offs were oversold and that the SSC was
not the only route to fundamental insight, SSC supporters could fall back on
a third, pragmatic justification: that killing the SSC would not free up funds
for other kinds of basic research. Furthermore, high energy physicists and
their allies argued, such a large federal expenditure was a rising tide. It ce-
mented a federal commitment to basic research, and was thus likely to repre-
sent a greater abundance of funds for all types of science in the long run, not
less funding. As a result, the physics community was obliged to unite behind
188 SOLID STATE INSURRECTION

the option that was on the table. This is what Fleury called “the myth of the
non-zero-sum gain.”
With respect to the billions of dollars’ worth of appropriations to fund
the SSC, the argument that freeing them up would not accrue to the benefit
of other areas of science was in some sense correct. Congressman Joe Barton
of Texas made this case in 1993 in a bid to save what would have been a
boon for his state: “I would prophesize that if we zero out the SSC . . . that
money is not going to walk on the water to other applied physics or any other
basic science, it is going to, in the best case, not be spent at all—in other
words, go to deficit reduction—or it is going to go to other projects that are
nonscientific related.”59 Unlike funding for the National Science Foundation
of the National Institutes of Health, the SSC’s appropriation was a separate
budget item. Its cancellation would not mean that those funds would be re-
distributed to other deserving scientific projects in the absence of a separate
appropriation.
However, the case against this argument was about more than the funding
earmarked for the SSC; it had to do with the effects that big science projects
made on the national research culture. The question, for condensed matter
physicists, was not whether the same $11.8 billion required for the SSC
might be carved up among other fields, it was whether megascience projects
like the SSC produced an overall damaging effect on small science. And that
question they answered in the emphatic affirmative. The concern remained
that the SSC’s long-term upkeep costs would make future funding appeals
more difficult for other projects. Both the scale of big physics projects and
the rhetoric used to justify them, condensed matter physicists charged, did
damage to other areas of physics by narrowing the definition of basic research
and optimizing physics training for particular areas. High energy physicists
had made the intellectual game zero sum by insisting on a monopoly over fun-
damental research. And the skew in the funding environment bled over into
pedagogy: teachers and advisers could be expected to devote more attention
to more fundable areas, and time in the classroom devoted to some topics was
time not exploring others.
The first half of this objection contended that the ethos behind the SSC
perpetuated the state of affairs in which condensed matter research was
viewed only in terms of applications, and again Philip Anderson was the prin-
cipal messenger. Anderson, in his opposition to the SSC, departed somewhat
from the case he had made in “More Is Different” two decades earlier. In the
MOBILIZING AGAINST MEGASCIENCE 189

early 1970s, the question of funding remained in the background. His SSC
testimony repurposed his arguments about the character of scientific knowl-
edge to underwrite a picture of how science funding should be organized.
Anderson did not oppose the SSC per se, but objected to the consolidation
of financial support for nonapplied research in big-budget particle physics in-
stallations while solid state, confined to smaller labs, pursued narrow, practi-
cal objectives and found scant opportunity for intellectual curiosity. The nar-
row focus was a consequence, he believed, of the conflation of fundamentality
with reduction. He described fundamental research in solid state as “caught
between the Scylla of the glamorous big science projects like the SSC, the ge-
nome, and the space station, and the Charybdis of the programmed research,
where you have deliverables, where you are asked to do very specific pieces
of research aimed at some very short-term goal.”60
Anderson attempted a precarious traverse of this rhetorical Strait of Mes-
sina. He advocated funding fundamental research for its own sake and re-
sisted tying funding to preconceived, near-term technological outcomes. He
opposed the SSC because, (a) condensed matter was just as fundamental as
high energy physics, and (b) funding exploratory condensed matter research
with no strings attached would, as a matter of course, produce more socially
relevant and technologically valuable results. Condensed matter physics
could boast sterling technological bona fides, but advocating funding priority
over the SSC strictly on that basis would undercut Anderson’s mission to
demonstrate solid state’s intellectual merit. Even if a technological justifica-
tion would have fallen more musically on many legislators’ ears, condensed
matter’s fight for intellectual prestige was too strong a component of its iden-
tity for Anderson to sell it short. It was a vision that could never get off the
ground in a world in which big physics and fundamental physics were synon-
ymous. The extremophilia of high energy physics and cosmology (figure 9.2)
had left condensed matter physicists, and others who worked on mesoscale
phenomena, fighting over the dregs of the federal basic research budget or
trying to make do with funding that was narrowly targeted and tied to pre-
conceived outcomes.
The question of training did not receive quite as much attention as the
question of intellectual priorities. It was nevertheless, for Anderson at least,
just as critical for assessing the influence of the SSC on the direction of phys-
ics. In his written testimony submitted to the House Subcommittee on En-
ergy Research and Development of the Committee on Energy and Natural
190 SOLID STATE INSURRECTION

Figure 9.2. Extremophilia in physics, 1988. This cartoon by S. Harris appeared along-
side an article critical of the attitudes it illustrates. Leo Kadanoff, “Cathedrals and
Other Edifices,” Physics Today 41, no. 2 (1988), 9–11. Reprinted by permission of
ScienceCartoonsPlus.com

Resources in 1987, he offered some observations on how the prominence of


high energy physics, in terms of both funding and prestige, shaped the way
physics was taught.
Particle physics, Anderson observed, “dominates many of the most pres-
tigious institutions because the spare change from large contracts in particle
MOBILIZING AGAINST MEGASCIENCE 191

physics or astrophysics can be used to support junior teaching positions, the


glamor of which attracts very able young people in spite of the insecurity of
their tenure.” In addition, because skills gained learning condensed matter,
chemical, plasma, and other complex systems physics translated more easily
into industrial settings, extreme-scale physics ended up overrepresented in
the academy. And departments that might have liked to beef up their rep-
resentation in fields like condensed matter were hampered by the scarcity
of federal funding to support research in those areas. As a result, Anderson
reported hearing “again and again from students that they just never heard
anything from their teachers about any other branches of physics until they
reached graduate school, if then.”61 The federal emphasis on big science, in
other words, was amplified by feedbacks that heightened the barriers to con-
ducting fundamental research into the properties of complex matter.
What might otherwise have been a parochial concern about the relative
status of subdisciplines within the programs that produced the country’s
physicists was projected into national-scale discourse as a question of sci-
entific staffing. Was the country producing enough physicists, and enough
of the right kinds of physicists, to address strategic national priorities? Con-
densed matter physicists were not above weaponizing nativism to make their
point. Krumhansl maintained: “We are not supporting some very capable
people. We are not training a sufficient number of Americans and we have
become dependent on a steady influx of foreign scientists. We have allowed
the funding of impressive facilities to take the support from programs at the
individual level.”62 Anderson commented that “it is beginning to be a sur-
prise to hear an American accent in the physics departments of this country
because we are having to fill in more and more with immigrants and visitors
in our junior positions.”63
Condensed matter physicists might not have been so concerned about
their field swinging back toward industry. Industry was, after all, integral to
the field’s origins and had midwifed some of its most fundamental contribu-
tions. But the industrial laboratories of the 1980s and 1990s were not the in-
dustrial laboratories of the 1950s and 1960s. Whereas the latter had bought
into farsighted, unprogrammed, long-range research as a basis for technolog-
ical development, the rise of the innovation economy sifted the “research”
out of “research and development” and focused in-house efforts on short-
term gadgeteering.64 If the Bells, GEs, and Westinghouses of the world were
not willing to support basic condensed matter research then it would have to
occur in the universities and the national laboratories, which left condensed
192 SOLID STATE INSURRECTION

matter physicists more susceptible to the whims of federal funding than they
were accustomed to and more acutely aware of the deficit they faced with
respect to high energy physics.
In one sense, the myth of the non-zero-sum gain was no myth at all: can-
celing the SSC did not, in fact, free up billions for basic condensed matter
physics. The bulk of federal funding for basic condensed matter research
came from two sources, the National Science Foundation and the DOE’s
Basic Energy Sciences (BES) budget. The total NSF research budget grew
slowly and steadily through the 1990s, just marginally ahead of inflation, and
showed no sign of responding to the SSC’s cancellation.65 Every BES bud-
get request, and every congressional appropriation, for the remainder of the
1990s would be less than the 1993 appropriation. Only in 2000 would BES
funding again reach 1993 levels in actual dollars.66
But condensed matter physicists were less concerned with the arithmetic
of federal appropriation than they were with the ethos driving it. The ques-
tion of what happened to the funds from SSC appropriations was secondary
to the effects of the dominance of big science—and the monopoly it sought
on big questions—in guiding federal funding.67 The direction and mission
of physics had been largely left to the discretion of big physicists. In 1989,
as Anderson highlighted the rise of mission-oriented research and decried
the damage it was doing to condensed matter physics, Burton Richter could
comment: “The Department of Energy, from which high-energy physics gets
almost all of it[s] money, and from which nuclear physics gets almost of its
money, has been rather good about not trying to focus research too tightly
or direct it too centrally.”68 That was the sort of control over their field con-
densed matter physicists felt they had lost while operating in the shadow of
big science, and they hoped stopping the SSC would be a first step toward
escaping from that shadow.

COMPETING PRODUCTIVITY CLAIMS


Condensed matter physicists and high energy physicists were in one sense
engaged in an intellectual property dispute throughout the SSC debates.
They were not sparring over patent rights, but they were engaged in the
type of wider discourse about intellectual property that Christine MacLeod
and Gregory Radick identify when they refer to productivity claims, which
they define as “claims for bodies of scientific knowledge as having inherently
useful offshoots,” in which “the ownership asserted is that of a theoretical-
empirical discipline over a domain of technical practice.”69 The SSC debates
MOBILIZING AGAINST MEGASCIENCE 193

offer a clear example of how this more expansive view of intellectual property
is useful. The physicists involved were never in a position to patent the tech-
nologies they wanted to claim as products of their fields, but they nonetheless
sought to establish some measure of ownership over them and to translate
that ownership into material support for further research. The differences in
how they set about that task are revealing of the distinct views of the purpose
and identity of physics they brought to the SSC debates.
The political environment surrounding the SSC encouraged productiv-
ity claims. Unlike earlier high energy physics facilities, the SSC demanded a
strong argument that it would contribute directly to national economic com-
petitiveness. SSC supporters advanced productivity claims through spin-off
rhetoric, aiming to attribute technical accomplishments in areas as diverse as
medicine, computing, construction, and manufacturing to high energy phys-
ics on the joint basis of its fundamental nature and the demands of accelerator
construction. Whereas in the 1960s and 1970s, high energy physicists had
been aggressively dismissive of connections between their work and techno-
logical development, post–Cold War circumstances gave productivity claims
new utility.
Condensed matter physicists had a longer history of making productiv-
ity claims—they suffused the National Research Council reports that traced
the field’s development—but that did not decrease their urgency in the early
1990s. It was a period of acute malaise for the field, even as it remained well
populated and rich with exciting new research topics. The collapse of Bell
was emblematic of a larger deflation of industrial research, weakening the
employment market. Former Bell Labs physicist John M. Rowell attributed
the decrease in industrial interest in research to a decrease in demand for
the knowledge that condensed matter physicists produced. Whereas indus-
trial concerns through mid-century had thrived on robust research opera-
tions, late-century corporations succeeded through commercializing existing
knowledge—every IBM had its MCI. The attitude prevailing in industry was
that “research is essential, but you’d be smart to let someone else do it.”70
Condensed matter physicists therefore had reason to worry whether their
productivity claims were potent enough, and competing claims that high en-
ergy physicists advanced on what they perceived as some of their proudest
technological accomplishments added insult to injury.
Their opposition therefore targeted the same technological outcomes that
high energy physicists claimed. If it was possible to show that other areas
were equally or more productive than high energy physics, the SSC would
194 SOLID STATE INSURRECTION

begin to seem like a poor investment. This argument alone would have pro-
vided a powerful case against the SSC in the political climate of the late 1980s
and early 1990s, and for anyone intent on simply sinking the project, it would
have been easy to stop there. But condensed matter physicists were not bent
on destruction. Their primary objective was to promote a friendlier climate
for their own research. For high energy physicists, productivity claims were a
way of making a field that legislators were inclined to find remote seem more
relevant. But condensed matter physicists faced the somewhat different chal-
lenge of reinforcing their existing, if problematic claims to technical relevance
without binding themselves too tightly to technology and undercutting their
claim to support for exploratory research. This required showing that their
intellectual standing was on a par with high energy physics, which they aimed
to do by subverting reductionism.
The justifications that condensed matter physicists had developed for
their field through the mid-1980s, which suggested that basic research could
be expected to yield fruitful results and pointed to the track record of solid
state physics doing just that, were inadequate to address the unique challenge
that late twentieth-century megascience posed. They gave no guidance for
distinguishing the potentially applicable from the eternally abstruse. Rowall
argued further that condensed matter physicists had to “face the possibility
that the changes in industrial research labs over the past 20 years constitute an
expression of dissatisfaction with our contributions over that time.”71 Making
the case against the SSC and in favor of greater support for basic condensed
matter research required moving beyond both the uncut emergentism Ander-
son had peddled in “More Is Different” and the general arguments for the fe-
cundity of basic research that marked the early rhetoric of condensed matter
physics. Finding both of these justifications inadequate for the purposes at
hand, condensed matter physicists suggested more pointedly that some areas
of fundamental inquiry were, in fact, inherently closer to useful applications
than others. Their position was a reaction both to the failure of emergentism
to make a case for their field’s economic relevance, and the failure of simple
basic research rhetoric either to explain why condensed matter physics was
a better investment than high energy physics or to advance its aspirations to
intellectual prestige.
Arguments for the importance of basic research historically rested on the
premise that it is impossible to know in advance which basic research would
turn out to be useful and which would not. In the late 1980s and early 1990s,
condensed matter physicists rejected that premise explicitly. The essence of
MOBILIZING AGAINST MEGASCIENCE 195

the resulting position can be summarized by some of the sharper criticisms


Anderson leveled at the SSC. He let his frustration show in asserting, “Sci-
ence can be fundamental without being irrelevant.”72 At another point he
would contend: “The basic science of today—except for particle physics and
much of space science—is the technology of tomorrow.”73 Basic research, that
is, does not always feed the wellspring of technology. Its fruitfulness depends
on its proximity to the domains of knowledge that most often make them-
selves useful for technical applications. A position of this sort was necessary
to distinguish a field unfamiliar for most legislators from one that was much
more visible. Refining their productivity claims in the context of the SSC
debates prompted condensed matter physicists to build a more explicit and
positive connection between their sense of their work’s intellectual impor-
tance and their convictions about its technological fruitfulness.
A focus on what might appear to be mundane substances and subjects
had not historically been an asset for solid state physics. Whereas high energy
physics and cosmology benefited from frontier rhetoric, solid state physicists
worked at the scale of the everyday, making it more difficult to claim that they
were pushing into unexplored territory.74 Arguing for the inherently greater
proximity of basic condensed matter physics to practical outcomes helped
shift these dynamics in condensed matter physicists’ favor. Their objections
to the SSC rested, in line with the three dimensions explored above, on in-
sisting that high energy physics was remote, undistinctive, and extortionate.
Anderson testified, “I think our Japanese competitors have shown, and I sin-
cerely believe, (and, in my own career, have shown by example) that the race
is much more often won by ‘doing what comes naturally,’ and looking around
at mundane subjects like glass or conducting oxides. Dollar for dollar, we in
condensed matter physics have spun off a lot more billions than they have,
and we can honestly promise to continue to do so.”75 The physics of the mun-
dane, in other words, if left to its own devices, was both intellectually valuable
and predisposed to lead in useful directions, and those benefits could be had
at a fraction of the astronomical cost of the SSC.
This line of argument was a synthesis of a half century of thinking about
solid state and condensed matter physics that brought together the various
and sometimes divergent strands of the field. Early arguments on behalf of
solid state physics had prioritized bringing industrial researchers in from the
cold. In the 1970s and 1980s, industrial research became much better inte-
grated into the infrastructure of American physics and, by the time of the SSC
debates, condensed matter physicists shifted to advancing the field’s place
196 SOLID STATE INSURRECTION

within the intellectual hierarchy of American science. The case presented in


opposition to the SSC was a way to link these strands and present an im-
age of condensed matter physics as a field that could both fuel the American
economy and advance the intellectual ambitions of American physics. By tak-
ing the unprecedented step of publicly and vociferously opposing their col-
leagues’ flagship project, those in condensed matter physics were auditioning
their own view of the field as the dominant mode of interaction between phys-
ics and American society. The SSC’s demise, in the face of gargantuan cost
overruns and management difficulties, might seem to depend little on such
opposition, which might have done little more than provide a small measure
of political cover for skeptical legislators. But it did represent the success that
solid state and condensed matter physicists had attained over a half century
of crafting a professional identity that spoke to the founding mission of the
American Physical Society while simultaneously embracing technological
relevance.

EPILOGUE: SOLID STATE AND THE NEW BIG SCIENCE


In 1967, the eminent laser physicist Arthur Schawlow wrote to Felix Bloch,
“It may be that the theory of solid and liquid states is now so complicated
that one has to allow time for a man to reach a broad perspective.”76 He re-
ferred to the training required before promising theorists of the day could be
expected to make groundbreaking contributions, but the same observation
might be applied to the history of solid state and condensed matter physics.
Growing historical perspective often prompts a search for continuities across
what were previously understood as radical shifts. Historians have revisited
the dramatic changes that occurred in the scientific revolution to show how
the supposedly new way of looking at the world that appeared in the En-
lightenment drew liberally from medieval thought and practice. They have
reexamined the quantum and relativity revolutions to emphasize their debt to
nineteenth-century physics. They have sought out the roots of big science—
often considered the product of an abrupt reconfiguration of the practice of
science during the Second World War—in the 1920s and 1930s, and some-
times earlier.77 They will no doubt look in earnest for the bridges of stability
that spanned the turbulent end to the Cold War. When they do, condensed
matter physics is certain to be among them.
The death of the SSC presents a dramatic and powerful discontinuity. “It
symbolized the end of an era for physics in the United States,” according to
Daniel Kevles’s assessment, published two years after the House of Repre-
MOBILIZING AGAINST MEGASCIENCE 197

sentatives yanked the project’s funding.78 The SSC’s demise marked the end
of the US government’s previously ironclad commitment to large-scale phys-
ics built on the reductionist program. It more broadly signaled the moment
when physics was unseated, in favor of biology, as the marquee American sci-
ence. Kevles suggested that the SSC’s demise in 1993 culminated a process
begun in the 1970s that eroded the sky-high authority and unique privilege
physics had enjoyed in American society and politics during the early Cold
War, and made it just as responsive to the prevailing political winds as any
other interest group.
How does the story of solid state physics reframe that picture? First, as
the struggles of solid state physics show, the better part of American physics
had been buffeted on the political winds that sunk the SSC for much lon-
ger, so the physics community, understood more broadly, already contained
substantial expertise navigating them. Second, mixing solid state back into
the story means that we do not need to understand the SSC in simple de-
clensionist terms. The opposition to the SSC from within physics highlights
how the SSC’s absence changed the research landscape in ways that created
new possibilities for American physics. This standpoint allows us focus on a
long-term transition through the 1980s and 1990s, which shifted major US
facilities away from high-energy accelerators and toward multiuser facilities
hosted both at universities, such as Cornell, and at national laboratories, such
as the National Synchrotron Light Source at Brookhaven, the Advanced Pho-
ton Source at Argonne, and the High Flux Isotope Reactor at Oak Ridge.79
These facilities of the “new big science,” as Robert P. Crease and Cath-
erine Westfall call it, are adaptable to multiple research programs, friendly
to outside user groups, including those from industry, and, unlike forefront
high energy facilities, are forced to compete for those users with similar in-
stallations around the world.80 They were designed with fields like solid state
physics and materials science in mind, and soon found eager constituencies
in structural biology and biomedicine, pharmaceutical research, and even art
history and archaeology. The machines of the new big science have much in
common with the National Magnet Laboratory, which similarly was rooted
in solid state physics and was therefore designed without reference to a nar-
row theoretical program, making it relevant to an assortment of external user
groups.
Seen against this backdrop, the failure of the SSC appears much less dis-
continuous. Solid state had been pulling American physics further toward
technological usefulness, industrial collaboration, and extradisciplinary rel-
198 SOLID STATE INSURRECTION

evance for decades. It was a natural next step to bring the landmark, feder-
ally funded physical research facilities along with it. The questions that arise
when we try to understand the complex research ecology that emerges within
a synchrotron radiation facility in the early twenty-first century have much in
common with the questions we have to confront when trying to understand
the complex constellation of research problems, theoretical approaches, ex-
perimental techniques, and institutional contexts that was solid state physics
in the mid-twentieth century.
American physics lived with a secret for decades: almost since its forma-
tion, it had been routinely unfaithful to its espoused ideals. But it was that
very unfaithfulness that allowed it to hold on to those ideals for so long. The
steady drumbeat of technical labor carried out by solid state physicists kept
the public and policymakers comfortable that, whatever the rhetoric issuing
from the priests of the field, it would ultimately produce something useful.
The new big science represents, in some sense, an abandonment of Henry
Rowland’s pure science ideal. But it is also a recognition of physics as a highly
heterogeneous field, and one solicitous of the society that supports it.
CONCLUSIONS

A pluralistic society rarely has consensual goals except in reaction to


some external threat (e.g., war, trade competition, waning national influ-
ence) or some internal danger (e.g., depression, insurrection, environ-
mental degradation). Lacking such challenges, each man tends to go his
own way.
—COSMAT, 1975

Henry Rowland, who guided American physics into the twentieth century,
would not have recognized it at the dawn of the twenty-first. It had grown
well over a thousandfold, had subdivided in ways he would have found ir-
rational and bizarre, and had become financially dependent on industrial
largesse in a way he would have found profoundly worrying.1 His influence
nevertheless remained. The tension between Rowland’s pure science ideal
and the pragmatism of the American context never fully resolved, and it
proved productive. Solid state physics was torn between its desire to uphold
the pure science ideal and its ability to leverage its technological relevance
through the second half of the twentieth century. Striving for purity made
solid state relevant to the conceptual core of physics; continued technological
relevance enabled the most abstract portions of physics to thrive within an
environment that might otherwise have been hostile to them. The back and
forth between the pure science ideal and the allure of technical and economic
rewards prompted physics to broaden its scope in mid-century, motivated
new approaches to fundamental knowledge, and ultimately remade the iden-
tity of American physics.
Solid state physics was such an unruly alliance that understanding it as
a whole requires examining its institutional manifestations. That standpoint
200 SOLID STATE INSURRECTION

reveals a story of how it gradually reformed the institutions of American


physics to be more conducive to its needs. The American Physical Society
(APS) developed more mechanisms to accommodate industrial and applied
physics. The Physical Review split and devoted one section to solid state
physics. Large installations like the National Magnet Laboratory dedicated
much of their research to solid state questions. These institutional develop-
ments were shaped by the ambivalence of American physics toward its pure
science mission in the face of its manifest technological relevance. They
gave the most technologically, economically, and militarily relevant portion
of physics the space to succeed, and in doing so shored up the relevance of
physics for American society.
The solid state insurrection, which mobilized against the institutions and
ideals that carried on Rowland’s legacy, proceeded on three fronts. It was a
revolt against the marginalization of applied physics, against the traditional
understanding of disciplines, and against the excesses of megascience. For
much of the later twentieth century, solid state and condensed matter physi-
cists considered themselves a persecuted plurality, fighting to bring the insti-
tutions that guided and controlled American physics, which for the most part
took nuclear and high energy physics as the default, more in line with their
interests. Like most insurgencies, the solid state insurrection was wracked
by internal disagreement and division, but it also succeeded in shifting the
center of American physics, which, by the late 1990s and early 2000s, was
more accepting of applied research, more flexible in defining its subfields,
and less beholden to very large facilities. By way of conclusion, I will review
these fronts on which the solid state insurrection advanced, each of which
shows how viewing American physics from the standpoint of solid state and
condensed matter physics can promote a more complete understanding of
the history of American physics as a whole.

APPLIED PHYSICS COMES IN FROM THE COLD


In 2014, the American Physical Society added a twelfth journal to the Physi-
cal Review family. Physical Review Applied appeared with a mandate to pub-
lish the “highest quality papers at the intersection of physics and engineer-
ing, with the goal of bridging the artificial divide between these fields.”2 As
similar changes to the publishing and professional landscape unfolded from
the 1930s through 1950s, physicists considered a number of divides to be
artificial—between metals and solids, solids and other forms of matter, indus-
CONCLUSIONS 201

try and academia, even, at times, physics and chemistry. But the line between
physics and engineering usually remained bright.3 The road that made it pos-
sible for a branch of the Physical Review, the inner sanctum of pure science in
mid-century, to declare physics and engineering of a piece was circuitous, but
it began with the first stirrings of the solid state insurrection.
The thread of physics publishing that runs through this book illustrates
the ebbs and flows of the status of applied and industrial physics in the
United States, and the manner in which solid state physics influenced them.
Journals proliferated to serve applied physicists in the first few decades of the
twentieth century. These fell into three categories. Some were launched by
professional societies with a strong applied focus, such as the Journal of the
Optical Society of America, the Journal of the Acoustical Society of America,
and the Journal of Rheology, which sought to create new communities and
forums for applied physics that the APS was not providing. Others found
their start as part of efforts to stem the influx of submissions to the Physical
Review. Technology and instrument-focused research constituted the bulk of
the flow targeted for diversion into the Journal of Applied Physics and Review
of Scientific Instruments. Following the formation of the American Institute of
Physics (AIP) in the 1930s, journals such as the Journal of Chemical Physics
and the American Journal of Physics created outlets for specialties that were
otherwise underserved. The cumulative effect was to narrow the Physical Re-
view’s focus, in tandem with the quantum revolution, which encouraged an
increase in theoretical papers, reaffirming its status as a pure science journal
and signaling the continuing marginal status of applied physics, in spite of its
growth.
The 1940s brought widespread hand-wringing about the sorry state of
the academia-industry relationship. Solid state physics established itself as a
division of the APS, and as a recognized subdiscipline of physics, in response
to those anxieties. It was not obvious that the loose constellation of activi-
ties that became solid state physics would be fully or even mostly classified
as physics. But the fact that they were had consequences for the publishing
habits of solid state physicists. Papers in the Physical Review were the coin of
the realm. The rapid and substantial increase in the population of solid state
physicists through the late 1940s and early 1950s contributed significantly to
a new wave of pressure on the APS’s flagship journal, which began to suffer
from backlogs and budget deficits.
These considerations blossomed into the publication problem of the mid-
202 SOLID STATE INSURRECTION

1950s, with the Physical Review operating at a loss and unable to promise
rapid publication. Solid state physicists seriously considered ordaining the
Journal of Chemical Physics the principal publication outlet for the field, or
creating a new journal of record for solid state physics de novo. Such propos-
als were abandoned in favor of a resolution to build solid state into a main-
stream physics subfield, which aspired to be of sufficient interest to other
physicists to merit inclusion in its general-interest journal. Solid state physics
in the 1950s, fractious and compartmentalized by research topic, inspired
some to worry that it might drift away from the core identity of physics to
explore collaborative opportunities with neighboring fields. The decision to
play the game on the terms set by the APS and the Physical Review meant
that the fate of American physics and the fate of solid state physics, applied
baggage and all, became linked. Solid state physics, in short, was the princi-
ple vector that introduced applied research into the mainstream of American
physics.
Solid state would succeed in its bid to install itself as a mainline phys-
ics subfield. The Physical Review would fragment in 1970, with one of the
four sections dedicated to solid state physics and related fields. The 1970s
and 1980s saw the rise of condensed matter physics, an evolution from solid
state physics that sought a more consistent disciplinary identity and more
traditional emphasis on pure science values. Nevertheless, proximity to tech-
nological applications remained a point of pride for what was now by far the
largest segment of American physics. By the time the Superconducting Su-
per Collider (SSC) was canceled in 1993, a significant portion of the physics
community had learned to wear their applied relevance comfortably.
The condensed matter outlook on physics was not so much the clear
victor of the SSC debates as it was the position that was left standing. The
high-powered reductionism that had infused pure science rhetoric emanating
from the particle physics community suffered a serious blow with the death of
the SSC. Even before the project’s cancellation, the likes of Steven Weinberg
and Leon Lederman were moved to distance themselves from the rampant
spin-off claims that diluted the intellectual and cultural arguments in favor
of high-price high energy physics. SSC spin-off claims, however, were pale
echoes of much stronger arguments for technological relevance that solid
state and condensed matter physicists had been making for some time. Mov-
ing into the twentieth century, these assumed a much larger role in Amer-
ican physics. The declaration in Physical Review Applied that the barriers
separating physics from engineering were artificial and worth demolishing
CONCLUSIONS 203

is one legacy of the momentum solid state physics gradually imparted to the
American physics community, nudging it away from a pure science ideal and
toward acceptance of technological relevance.

THE CHANGING SHAPE OF PHYSICAL DISCIPLINES


Physics is what physicists decide it is. American physicists themselves, how-
ever, failed to tap into that agency for much of the early twentieth century.
The dawning realization in the 1940s that disciplinary boundaries could be
gerrymandered to suit immediate professional needs gave physicists a power-
ful and flexible tool to organize a rapidly growing population and to manage
the divergent and sometimes competing interests of the groups that emerged
within it. Solid state physics was the most immediate and the most successful
result of this realization. Organizing around something other than a suppos-
edly natural category made it outré in the 1940s and 1950s, but its success
presaged further deliberate efforts to strategically engineer and sometimes
undermine disciplinary structure later in the twentieth century.
Paul Forman has identified in the late twentieth century an era of transition
from disciplinarity to antidisciplinarity, coinciding with a larger cultural shift
from modernism to postmodernism. The notion of disciplines as distinct and
formalized bodies of professional knowledge and practice, Forman argues, is
of surprisingly recent historical vintage, a product of a mid-century modern-
ist emphasis on professional discipline (in the sense of learned self-control)
and expertise. The 1960s and 1970s, though, saw the beginnings of an anti-
disciplinary reaction, which disparaged disciplines as Procrustean impedi-
ments to both epistemic and social progress.4 The rise of solid state physics
coincides with the era in which Forman locates this transition, and presents a
telling illustration of how physicists were negotiating the concept of discipli-
narity both in its heyday and through its decline.
It is an amusing irony that one of solid state’s key contributions to the
organization of American science—its self-consciously conventional outer
boundary—was the very characteristic that made it idiosyncratic when it
formed. It is also fitting that the field dedicated to studying complex phys-
ical systems was itself such a complex disciplinary assembly. Solid state pi-
oneered a new mode of disciplinary organization, one that ran counter to
the prevailing mid-century wisdom. Contemporary resistance aside, it would
soon inspire imitators. Materials science, a product of early 1960s efforts to
combine basic scientific research in the physical sciences with engineering
efforts targeted at strategic materials, was likewise a discipline by convention,
204 SOLID STATE INSURRECTION

assembled in large part to capitalize on the stated preferences of powerful


funders. A similar pattern would unfold later in the century with the advent
of nanoscience and nanotechnology, which grew not from a set of problems
or practices, but from an opportunistic reaction to federal development pri-
orities and accompanying funding opportunities.5
Solid state, as an antecedent to these and other developments, therefore
provides a powerful basis from which to understand disciplines and their
influence over scientific discourse during the twentieth century, when the
rapidly growing population of scientists favored new modes of organization.
Furthermore, the solid state example clarifies the importance of disciplinary
identity for mediating scientific authority, governing prestige disputes, and
guiding resource allocation. The lack of a clean research- or concepts-based
answer to the question “what is it that solid state physicists did?” focuses
attention on the sense of professional solidarity that unified the field for
decades.
Through the second half of the twentieth century, physics as a whole came
to resemble a constellation of loosely related interests like solid state phys-
ics much more than it resembled the conceptually coherent discipline that
physicists like John Van Vleck had sought to preserve in the face of 1940s di-
visionalization. Narratives driven by the proliferation of quantum mechanics
and efforts to hasten theoretical unification often obscure this evident state
of affairs. But the evolution of the professional infrastructure of American
physics—a story to which solid state was central—was a progression away
from the early twentieth-century Rowlandian vision of a community unified
by full-throated advocacy of pure research. The APS sprouted divisions
beginning in the early 1940s, initiating a process of specialization and com-
partmentalization that continued for the rest of the century.6 Physics Today
launched in 1948, boasting of its egalitarian editorial policy. From the 1960s
on, growing numbers of physicists participated in new interdisciplinary fields
such as materials science. All these factors indicate that physics steadily be-
came a bigger tent with thinner walls through the second half of the twentieth
century.
Understanding how and why solid state formed therefore offers a case
study, on a smaller scale, of how American physics evolved. The nature of
physics changed as membership in the disciplinary community became less
a matter of working on a small set of communally sanctioned problems and
more about training and self-identification. As with the condensed matter
movement with respect to solid state, there was some pushback against this
CONCLUSIONS 205

development within physics. The push for grand unification and the strong
reductionist impulse that drove high energy physics from the 1960s through
the 1990s, however, appears in a different light if considered as a counter-
point to a diversifying field than it does when presented as the central narra-
tive of American physics. Solid state can encourage a more complete histori-
cal understanding of American physics in the later twentieth century by itself
being a microcosm of how the larger community was structured.

MEGASCIENCE UNDERMINED
If the interests of applied science provided the impetus for the solid state in-
surrection, and the reimagining of physical disciplines was its consequence,
then ressentiment brought about by the excesses of big science was its fuel.
Prestige looms largest to those who feel they lack their fair share. Solid state
physicists spent much of the Cold War watching their opposite numbers in
nuclear and particle physics reap the social, political, and pecuniary benefits
of physics’ newfound cultural visibility. Much of that visibility came in the
form of historically large installations. As particle physicists probed smaller
and smaller elements of the atom, the machines they used to do it grew, and
the solid state community’s status consciousness heightened in direct pro-
portion. Many a solid state physicist sported a chip on the shoulder, and that
pugnacious streak flavored the field’s professional story from the 1960s on.7
The installations of big science were monumental. They were, of course,
physically large. But an enormous particle accelerator could also function, in
the words Daniel Kleppner, “as a monument to our National aspirations.”8
Kleppner, an atomic physicist, supported the SSC, although he echoed solid
state physicists’ admonitions that it should not be supported at the expense of
other areas of science. By identifying the monumental function the SSC and
other large accelerators served, Kleppner nevertheless articulated one of the
principle frustrations solid state and condensed matter physicists had with
big science. Monuments, in the popular imagination, celebrate particular
parts of an event or enterprise at the expense of others. Understanding high
energy particle accelerators as monuments to national scientific accomplish-
ment implied a particular slant on what types of scientific accomplishments
were important to the nation.
Solid state physics lacked a monument on which it could pin its aspira-
tions. A field dispersed among laboratories across academic, industrial, and
governmental institutions, supporting an abundance of research programs,
solid state did not lend itself to physics by monolithic machine. Many in the
206 SOLID STATE INSURRECTION

field responded by rebelling against monumentalism itself, suggesting that


large, sui generis facilities were a poor investment and damaging to the health
of the rest of science—the attitude that framed the scientific case against the
SSC. It would be a stretch to call the death of the SSC a victory for small
science. It did not yield increased interest in funding exploratory research
on the lab-bench scale or reduce demands for well-articulated technical or
commercial outcomes that came with much small-science funding. It did,
however, signal the closing of an era in American science in which the reduc-
tionist view of physics championed by the high energy community would
exert outsized influence on national-scale science policy.
Opposition to the funding structure implied by big science was intimately
linked to opposition to the reductionist philosophy. Although high energy
physicists often presented reductionism as the principle that had propelled
physics for millennia, dating to ancient Greek atomism, the flavor of reduc-
tionism that guided high energy research in the latter half of the twentieth
century is more properly understood as a new intellectual framework, de-
veloped from and for the research program that led to the standard model of
particle physics. The emergentism developed by Philip Anderson and other
condensed matter physicists, which maintained the conceptual indepen-
dence of concepts at higher scales of complexity, was likewise drawn from the
phenomena of condensed matter physics. But in addition to codifying con-
ceptual differences between the high- and low-energy domains, philosophi-
cal disagreements about reduction and emergence had wide-ranging conse-
quences for the self-image of physics and its organization. They contested
whether the hard core of fundamental physics should remain an exclusive
club or widen into a big tent.
Reductionism was a near necessity to underwrite the claims that large
expensive facilities were necessary to push forward the boundaries of fun-
damental knowledge. If, however, fundamental insight could be gained at
any scale of organization, then the national commitment to science as cul-
ture could reasonably be spread more widely. For this reason, philosophi-
cal arguments about the epistemic characteristics of fundamental knowledge
formed an integral component of condensed matter physicists’ case against
the SSC. Condensed matter physicists not only fought against megascience
because it sequestered resources for basic research in a few large facilities but
also because they smarted at the intellectual avarice of the reductionist move
to monopolize fundamental knowledge. American physics would not come
CONCLUSIONS 207

to embrace applied relevance as part of its identity without those areas that
boasted this relevance first demonstrating that they could contribute to the
intellectual core of physics.
The struggles of solid state physicists to carve out space for their research
and publicize their accomplishments while laboring in the shadow of big
physics also has lessons for how to tell the history of twentieth-century sci-
ence. The monuments it erected stand out in our vista of the past, but the
most imposing features of the skyline are often only dimly representative of
the surrounding terrain where most people spend most of their time. Solid
state physicists, scattered throughout the foothills of mountainous big sci-
ence installations, accounted for a much larger proportion of scientific output
in the late twentieth century. Their resentment of big science was not simply
a matter of turning up their noses at high-hanging fruit, but of advocacy for a
system of support and accountability that better reflected the demographics
of American physics.

MAKING AMERICAN PHYSICS MATTER


The story of the solid state insurrection invites revisiting standard narra-
tives of American physics in the twentieth century, many of which have been
shaped by the Inward Bound narrative, articulated in Abraham Pais’s book of
that title. Pais presents the twentieth century as a voyage into the atom, which
uncovered the deepest and most fundamental secrets of nature among the
most basic constitution of matter and revolutionized our understanding of
the universe.9 Rarely is the focus on particle physics as the basis for the rest
of the discipline so restrictive and explicit as in Inward Bound, yet particle
physics has stood in for physics in general in many of the most influential his-
toriographical movements of recent years. Historians of physics think about
the relationship between theory and experiment and the nature of physical
knowledge through the lens of bubble chambers, weak neutral currents, and
quarks, approach pedagogy through Feynman diagrams, and understand lab-
oratory culture through the stories of the Stanford Linear Accelerator, Law-
rence Berkeley Laboratory, and Fermilab.10 The cumulative result is that the
reductionist narrative that particle physics themselves favored has become
the backbone of the historiography of twentieth-century physics.
But from another perspective, the advances in nuclear and subnuclear
physics after about 1930 might seem provincial and incremental. By 1930,
with the discovery of the neutron, physicists had a clear idea of the elements
208 SOLID STATE INSURRECTION

of nuclear structure. Nuclear physics, without a doubt, proved its importance


in dramatic fashion during the Second World War. But both the Manhattan
Project and subsequent bomb and reactor work were principally engineering
enterprises. So far as the basic physics of nuclear matter is concerned, the
rest of the century succeeded only in fleshing out the picture that was largely
in place by 1930 and pushing it back one level to develop a sense of sub-
nucleonic structure—an achievement to be sure, but a parochial one, given
that the phenomena in question can only be created and observed in highly
specialized laboratories.
From this perspective, the real action of twentieth-century physics un-
folded elsewhere. Anyone walking along a black-sand beach in Hawaii, or
New Zealand, or Hong Kong can observe, without instrumentation, the
property of ferromagnetism in the magnetite deposits that give these beaches
their distinctive color. Yet that phenomenon and many other easily observed
properties of matter remained mysterious before the advent of theoretical and
experimental techniques developed within solid state physics. From this per-
spective, the most active frontier of the twentieth century was not the high-
energy frontier, but the complexity frontier. It was the demystification of the
properties of complex matter and the applications of those properties, which
remade our technological world, from home computing, stereo equipment,
and cookware to communication, transportation, medicine, and warfare.
This is a tendentious framing, but it serves a purpose: it exposes the his-
toriographical contingency of the disproportionate focus on nuclear physics,
high energy physics, and cosmology, alongside the historical contingency of
the dominance those fields assumed over public discourse about physics in
the second half of the twentieth century. Two counterfactual scenarios help
to probe that contingency further, each of which offers heuristic utility by
throwing into relief the role solid state physics played, despite lacking the
public acclaim of its sibling subfields, in securing the prominence of Ameri-
can physics.11
First, given the high degree of institutional volatility in the early post–
Second World War era, it is easy to imagine a counterfactual scenario in
which solid state physicists migrate away from physics and into chemistry,
metallurgy, and engineering, much as electrical engineering had some de-
cades earlier. It was a contingency about which the field’s founders actively
worried, and the American Physical Society council was demonstrably
squeamish about clearing the type of institutional space that would give the
society a more industrial flavor. Without solid state, which accounted for a
CONCLUSIONS 209

large proportion of the postwar population boom, physics would have stayed
smaller and grown more slowly.12 The publishing crisis of the 1950s would
have been managed by directing solid state work into the journals of other
fields. Most important, the technical accomplishments of solid state physics
would have accrued to the benefit of those other fields.
In that scenario, it is difficult to imagine how American physics would
have maintained the political will necessary to continue the megascience
projects that became its most recognizable and best-publicized accomplish-
ments. Physics would no doubt have enjoyed a period of visibility and influ-
ence on the coattails of nuclear weapons, but the half-life of that influence
would likely have been much shorter. Physics was founded, proudly, on the
fringes of American values, and in direct opposition to some of them. Solid
state was instrumental in challenging that parochialism and making American
physics more responsive to the needs of society in which it was embedded.
In contrast, many particle physicists, deeply disillusioned with weapons re-
search, sharpened their notion of scientific purity to an even finer point. Such
an attitude would have been difficult to sustain without a branch of physics
that dedicated itself to exploring the areas that were of immediate relevance
in the Cold War and supported the perception that physics was continually,
not merely occasionally, useful.
We can imagine a second counterfactual scenario in which the Manhattan
Project never acquired the scale or resources it needed to construct a working
bomb before the end of the war in the Pacific. Perhaps Leo Szilard failed to
find Albert Einstein, who never cosigned his fateful letter to Franklin Roo-
sevelt demanding a US response to the German nuclear program. Perhaps
Gregory Breit remained, unhappily, the Coordinator of Rapid Rupture and
was not able to instill the famous esprit that marked J. Robert Oppenheimer’s
directorship of the Los Alamos laboratory. Suppose, for whatever reason, the
Second World War ends without a dramatic, public demonstration of the
power of the nucleus. It is hard to imagine that nuclear weapons research
would not have continued, but it would have done so in secret, under the
guidance of a very different postwar order. Nuclear physicists would not have
been catapulted to celebrity in quite the same way and high energy physics
would thus not have benefited from the public prominence that the psycho-
logical power of nuclear weapons brought to microphysics.
In this second scenario, solid state physicists would have been well po-
sitioned to become much more politically influential in the early Cold War,
in particular on the strength of radar research, which, in the absence of the
210 SOLID STATE INSURRECTION

bomb, would likely have dominated narratives about physicists’ contribu-


tions to the war effort. Nuclear physicists, shackled both by secrecy regimes
and doubts about the success and cost effectiveness of their proposal for a
super weapon, would have faced greater difficulty gaining political and cul-
tural traction. It would be reasonable to expect a much more rapid embrace
of applied relevance from American physicists—not an insurrection but a
coup, much like the one that catapulted molecular biology to the forefront
of biology around the same time. Historians would look back on Rowland’s
ideals as quixotic, marvel that physicists were able to sustain them even into
mid-century, and find the roots of their demise in the interwar restructuring
of the journal and society infrastructure.
The purpose of a counterfactual should always be to give us a fresh per-
spective on the factual. These scenarios, which probe intuitions about what
factors shaped postwar physics, show us that the history of twentieth-century
American physics does not fit together without the story of solid state physics
contributing answers to some of the most compelling questions about it. How
and why did physics gain such prominence so quickly and maintain it for so
long? What did it mean to be a physicist in the United States and how did
that identity change over time? How do physicists negotiate what it means for
knowledge to be fundamental? These are questions that get to the core of why
the history of physics matters, and the answers to them look quite different,
as I have argued here, when taking the experiences of solid state physicists
into account.
The counterfactual exercise above also makes clear that the precise tra-
jectory that the twentieth century did take conspired to mask the significance
of solid state physics within it. High energy physics, in spite of its scorn
for the types of short-term technological ends that the security state often
demanded, was well adapted to the cultural moment of the Cold War. The
James Bond films provide an analogy that makes this compatibility clear. The
Bond series is among the longest running in cinema, and its titular charac-
ter one of the most recognizable fictional figures in Western popular culture.
But when Daniel Craig took over as the franchise’s leading man, the familiar
character took an unfamiliar turn. No longer did the debonair spy’s copious
consumption of vodka martinis merely make his aim truer as he battled out-
sized supervillains and their Rube Goldberg–esque plans for world domina-
tion. Instead, Bond was represented as a fraying alcoholic, buckling under
the emotional toll of his job as his demons threatened to undermine his fight
against enemies created by his own country’s malfeasance. Unlike the Cold
CONCLUSIONS 211

War Bond made famous by Sean Connery—the product of a world painted


in black and white—Craig’s depiction befits a world that struggles to come to
terms with stateless enemies, indiscriminate electronic surveillance, and the
morality of drone strikes.
Set against the Manichaeism of the Cold War, Connery’s Bond made
sense. Craig, in turn, plays a messy Bond for a messy world. But the dyads
of capitalism and communism, East and West, good and evil, that played so
strongly on Cold War psychology evince the masking, not the absence, of
underlying complexities.13 In physics, as in the geopolitics of the Cold War,
the messiness was there all along, but we have been at times ideologically
indisposed to notice it. The picture of physics as divided between reduction-
ist, fundamental research on one hand and derivative, applied physics on the
other was powerful to physicists—especially during the Cold War, when it
fit the spirit of the times—and it has imprinted the history of physics as well.
Telling the story of Cold War physics requires peeling back this seductive ve-
neer and confronting the messiness of Schmutzphysik. It was not a consensus
ideological vision, compatible with the demands of a Cold War context, that
made physics such a successful science; it was the ideological diversity that
allowed it to adapt to the many niches that a complex context offered.
NOTES

INTRODUCTION. What Is Solid State Physics and Why Does It Matter?

Epigraph: Gregory H. Wannier, quoted in an untitled document, 1943, CRS, folder 3.


1. Establishing Priorities in Science Funding: Hearing Before the Task Force on De-
fense, Foreign Policy and Space of the Committee on the Budget, House of Representatives,
102nd Cong. 67 (July 11 and 18, 1991) (statement of Philip W. Anderson, Ph.D., Joseph
Henry Professor, Princeton University). The terms “particle physics” and “high energy
physics” were used interchangeably around this time, and are used that way in this book.
2. The relationship between these two names, which is not unproblematic, is dis-
cussed in chapter 8.
3. Studies of discipline formation have focused principally on how communities co-
alesce around research problems and techniques, or on how practitioners become “dis-
ciplined” through pedagogical programs and professionalization procedures, both for-
mal and informal. Representative examples include: Jackson, “Chemistry as the Defining
Science,” which shows how the growth of strong laboratories created the conditions in
which consensus practices could emerge, aiding the establishment of chemistry as a dis-
cipline in nineteenth-century Germany; Long, “William McElroy,” which examines how
the suite of new problems posed by molecular genetics contributed to the emergence of
biochemistry; and Lenoir and Lécuyer, “Instrument Makers,” which emphasizes how the
distinctive set of practices that spring up around instruments and instrument making help
assemble disciplinary communities. Others have challenged the universality of this ap-
proach, especially in the realm of more recent science, when communities became larger,
more complex, and more politically engaged. See, for example, Good, “Assembly of Geo-
physics.” I follow Good in understanding disciplines as flexible tools that can be used to
solve professional and institutional problems as well as conceptual ones.
4. Hoddeson et al., Out of the Crystal Maze, viii.
5. These, for instance, were the primary categories identified by the War Policy Com-
mittee of the American Institute of Physics in the early 1940s. Klopsteg, “Work of the War
Policy Committee.”
6. Although often repeated, the origins of this apparently verbal remark are unclear. It
had entered particle physics lore by 1966 at the latest, when Murray Gell-Mann repeated
it at the 13th International Conference on High-Energy Physics. Gell-Mann, “Current
Topics in Particle Physics.”
7. Daniels and Krige, “Beyond the Reach.” I thank John Krige for sharing a preprint
of this paper with me.
8. In a similar manner, we see the history of science differently if we start from the
perspective of applied science. Johnson, “What If We Wrote?”; Mody, “Toward a History
of Science.”
214 NOTES TO PAGES 8–19

9. The most comprehensive account of the SSC’s rise and fall is Riordan, Hoddeson,
and Kolb, Tunnel Visions.
10. See Kaiser, “Atomic Secret,” for an account of how the Manhattan Project devel-
oped its popular reputation as a triumph of theoretical physics, rather than of metallurgy,
chemistry, and engineering.
11. On the postwar prestige of physicists, see Kevles, Physicists, and Brown, Dres-
den, and Hoddeson, “Pions to Quarks.” Robert W. Seidel has situated the origins of the
growing prestige of American physics slightly before the war in the success of the Law-
rence Berkeley Laboratory. Seidel, “Origins of the Lawrence Berkeley Laboratory.” On
the increased influence of physics on policymaking, see, in addition to Kevles, Physicists,
Kleinman, Politics on the Endless Frontier.
12. See Kevles, “Big Science.”
13. Missner, “Why Einstein Became Famous.”
14. Dupree, Science in the Federal Government, is the origin of a broad historiograph-
ical consensus. Studies building on Dupree’s work have explored the details of the mech-
anisms that enacted the federal government’s newly robust commitment to funding scien-
tific research. See: England, Patron for Pure Science; Thibodeau, “Science in the Federal
Government”; Kleinman, Politics on the Endless Frontier; Edgerton, “Time, Money, and
History.”
15. Westwick, National Labs; Crease, Making Physics; Seidel, “National Laborato-
ries” and “Home for Big Science.”
16. Hecht, “Atomic Hero.”
17. Doel, “Scientists as Policymakers”; Weart, Scientists in Power; Finkbeiner, Jasons.
18. Hoddeson, Kolb, and Westfall, Fermilab.
19. Wolfe, Competing with the Soviets, 43.
20. Stevens, “Fundamental Physics.”
21. Bridger, Scientists at War, 270.
22. Rowland, “Highest Aim,” 828.
23. Joas, “Campos que interagem.”
24. I largely avoid engaging with the many interesting issues in the conceptual history
of solid state and condensed matter physics, which do not present a cohesive story of
the field and have been considered piecemeal by others. For an overview of these works,
see the section “Conceptual Development of Research Programs,” in Martin, “Resource
Letter,” 91–93.
25. The fear that particle accelerators might create minuscule black holes that would
grow and engulf the planet was the source of protests directed at both the Large Hadron
Collider in Geneva and the Relativistic Heavy Ion Collider on Long Island. For an edify-
ing history of these fears, and a discussion of the legal issues surrounding injunctions filed
on the basis of them, see Johnson, “Black Hole Case.”
CHAPTER 1. THE PURE SCIENCE IDEAL AND ITS MALCONTENTS
Epigraph: Rowland, “Highest Aim,” 825.
1. Tocqueville, Democracy in America, 2:48.
2. See Warwick, Masters of Theory.
3. On physics in Prussian secondary education, see Olesko, “Physics Instruction” and
Physics as a Calling. On the Physikalisch-Technische Reichsanstalt, see Cahan, Institute
for an Empire.
4. For an overview of both the connections between science and industry in nine-
NOTES TO PAGES 19–24 215

teenth-century Europe, and the literature discussing it, see Hunt, Pursuing Power and
Light.
5. The connection between science (including physics) and technology in early
twentieth-century America has been well documented, notably in Noble, American by
Design.
6. This analysis builds implicitly on the wealth of literature that charts the transforma-
tion of the American physics community in the 1920s and the influence of scientific emi-
gration in the 1930s. Representative of the genre are: Weiner, “New Site for the Seminar”;
Coben, “Scientific Establishment”; Holton, “Formation”; Hoch, “Reception”; Stuewer,
“Nuclear Physicists”; Rider, “Alarm and Opportunity”; Schweber, “Empiricist Temper”;
and Assmus, “Americanization of Molecular Physics.”
7. Rowland, “Highest Aim,” 826. A thorough account of Rowland’s professional tra-
jectory and intellectual development can be found in Sweetnam, Command of Light. On
Edison as a foil for Rowland, see Hounshell, “Edison and the Pure Science Ideal.”
8. See Kohlstedt, Sokal, and Lewenstein, Establishment of Science.
9. Rowland “Highest Aim,” 833.
10. See Lucier, “Origins of Pure and Applied Science.”
11. For a discussion of best-science elitism and its role in early twentieth-century
American physics, see Kevles, Physicists.
12. American Physical Society, “Members. June 15, 1902,” APSM. Six other mem-
bers listed only an address, placing an upper bound on industrial membership at 6.9
percent. The remainder of members reported academic or government affiliations, and
overwhelmingly the former. For a discussion of the rise of American industrial research in
this era, see Reich, Making of American Industrial Research.
13. “List of Members of the American Physical Society together with Lists of Officers
for 1920 and Past Officers, a Geographical Index, and the Constitution and By-Laws,”
July 1920, APSM. The bulk of the remaining members were employed in secondary
education, nongovernmental, nonprofit laboratories, such as the Carnegie Institution of
Washington, or did not indicate their employment.
14. “List of APS Members,” July 1920, APSM. Maj. George O. Squier of the War
Department and George K. Burgess of the Bureau of Standards were elected members of
the council, as, it appears was Frank Jewett of the Western Electric Company. The list of
council members names “J. B. Jewett,” as an elected member, but Frank, later to become
president of Bell Laboratories, is the only Jewett in the APS member roles for 1920.
15. Lucier, “Origins of Pure and Applied Science”; Noble, American by Design.
16. Gooday, “‘Vague and Artificial.’”
17. Huxley, “Science and Practical Life,” 167.
18. Hunt, Maxwellians; Cardwell, From Watt to Clausius.
19. Carty, “Relation of Pure Science,” 512.
20. Coulter, “Role of Science,” 22.
21. Reich, Making of American Industrial Research.
22. Weart, “Physics Business.”
23. See Wise, “Ionists in Industry.”
24. A. Robert, “It Ain’t the Money,” December 2, 1944, KBP, box 10, folder Rabbi
[sic], Isidor Isaac.
25. Hyde, “Why 1916?”
26. “Anecdotal History.”
27. “Editorial.”
216 NOTES TO PAGES 24–35

28. Waterfall, interview by Lindsay.


29. Governing Board of the American Institute of Physics, Minutes of Meeting held
May 3, 1931, AIPM, accessed February 7, 2017, https://www.aip.org/history-programs
/niels-bohr-library/collections/governing-board/may-3–1931.
30. “Preliminary Report of the Policy Committee on the Reorganization of Physics,”
Addendum to the Minutes of the Joint Meeting of the Executive Committee and Policy
Committee, March 8, 1945, AIPS.
31. Pais, “‘Physical Review.’”
32. For a discussion of the growth and function of review literature in the physics
community, see Lalli, “New Scientific Journal.”
33. Richtmyer, “Editorial,” 1.
34. I use articles, rather than pages, as a measure of size with the operating assumption
that anyone reading the journal would approach each article strategically, and so that read-
ing time would be correlated more closely to number of articles than to number of pages.
35. Quoted in Duncan and Janssen, “On the Verge,” 566.
36. Van Vleck and Nier, “John Torrence Tate,” 471.
37. Tate and Van Vleck, “Editor’s Column,” 69.
38. See Ambrosio, “Historicity of Peirce’s Classification”; Csiszar, “Seriality.”
39. Whewell, Philosophy of the Inductive Sciences, 278.
40. See Schweber, “Empiricist Temper.”
CHAPTER 2. How Physics Became “What Physicists Do”
Epigraph: Buckley, “What’s in a Name?” 301.
1. A search for articles that include both “radar” in their titles and “physics,” “phys-
icist,” or “physicists” in their text between January 1940 and May 1944, when Buckley
spoke, returns only two articles from the New York Times historical archive. One is a brief
note reporting Albert Hoyt Taylor’s Medal for Merit citation; the other is a grandiloquent
feature entitled “Radar—Our Miracle Ally,” which refers to James Clerk Maxwell as a
physicist when setting down the historical background on which the “inventers and engi-
neers” responsible for radar relied. “Two Get Medals for Merit,” New York Times, March
22, 1944; Davis, “Radar.” For an overview of radar’s influence on the Second World War,
see Brown, Radar History.
2. Hull, “Outlook for the Physicist,” 66.
3. Kevles, “Cold War,” 263. The quote responds to Paul Forman’s argument that the
influx of military spending into physics during the Cold War oriented physics away from
basic research. Kevles accuses Forman of being essentialist about physics and taking the
academic, pure-science-oriented segment of the discipline as, in some sense, the “true”
physics. Forman, “Behind Quantum Electronics.”
4. William W. Hansen, letter to Daniel L. Webster, February 4, 1943, FBPS, series 1,
box 5, folder 20.
5. Kaiser, “Booms, Busts,” discusses the systemic changes brought about by the rapid
post–Second World War population boom in physics.
6. David Kaiser addresses the provenance and significance of this term in “From
Blackboard to Bombs.”
7. On the demographics of physics in the 1930s, see Weart, “Physics Business.”
8. Kaiser, “Postwar Suburbanization.”
9. Within discussions that explored the future of the American physics community,
representatives of all camps tended to map the academia/industry divide onto the distinc-
NOTES TO PAGES 35–41 217

tion between basic and applied research. This mapping assumed, as Morris Muskat, a
research physicist at Gulf Research Laboratory, observed, that industrial physicists “are
not employed to do physics research per se, but to develop applications of physical prin-
ciples and techniques. Their specific problems are generally not self-created.” As a con-
sequence, the University of Chicago’s Robert Mulliken noticed, “The academic scientists
and the industrial or applied scientists each tend to flock by themselves, because the ac-
ademic scientists sometimes lack interest in the problems of industry, and the industrial
people find the papers presented by the academic scientists too theoretical or too high-
brow.” Muskat, “Letter to the Editor,” 38; Mulliken, “Remarks on a Possible Division,”
42. Mulliken was not opposed to a spectroscopy division, but sought ways to encourage it
to serve as a center of integration rather than a source of regional identity.
10. Osgood, “Physics in 1943,” 106. The school was renamed Michigan State Uni-
versity of Agriculture and Applied Science in 1955 and dropped the prepositional phrase
in 1964.
11. Muskat, “Letter to the Editor,” 38.
12. Harnwell, “Research in Physics.” The article was based on a talk Harnwell de-
livered at the Symposium on the Role of Physics in the Postwar Period, cohosted by the
APS and the American Association of Physics Teachers at State College, Pennsylvania,
June 18, 1943.
13. Harnwell, “Research in Physics,” 232–33, 235–36.
14. Harnwell, “More Perfect Union,” 20.
15. At this time, Waterfall was consulting for the US Navy while being paid through
Columbia University. Hutchisson, a University of Pittsburgh professor and editor of the
Journal of Applied Physics, would later serve as president of the AIP.
16. Waterfall and Hutchisson, “Organization of Physics,” 408–9.
17. The IRE was one of two societies—the American Institute of Electrical Engineers
being the other—that combined to form the Institute of Electrical and Electronics Engi-
neers in 1963.
18. Goldsmith, “Comments,” 649.
19. See Rosenberg, “American Physics.”
20. Saul Dushman et al., “The Present War Is a Physicist’s War,” CRS, folder 3. For
further discussion, see chapter 3.
21. Olsen, Crittenden, and Smith, “Letter to the Editor,” 108.
22. Osgood, “Letter to the Editor,” 108.
23. When the AIP’s Governing Board held its first meeting at the Cosmos Club in
Washington, DC, in May 1931, one articulation of its mission was “to reach a wide au-
dience of men working in physics and to give them easily readable news of all kinds re-
lating to physics.” It consciously aimed to avoid the limitations of the APS and respond
to the needs of physicists wherever they might be working, and wherever their interests
might take them. Such a mission necessitated a much broader conception of “physicist”
than the more conservative voices in the APS would have considered. Governing Board
of the American Institute of Physics, Minutes of Meeting held May 3, 1931, AIPM, ac-
cessed June 19, 2018, http://www.aip.org/history-programs/niels-bohr-library/collections
/governing-board/may-3-1931.
24. This address is reprinted in Seitz, “Whither American Physics?” The January
1945 meeting of the APS hosted, and Seitz’s talk was a part of, the first symposium on the
solid state discussed in the next section. Darrow, “Symposium on the Solid State.”
25. Seitz, “Whither American Physics?” 40–41.
218 NOTES TO PAGES 42–51

26. Seitz, “Whither American Physics?” 41–42.


27. Seitz, “Whither American Physics?” 40.
28. Council of the American Physical Society, Minutes of the Meeting Held at Pitts-
burgh, June 21, 1940, APSM.
29. Council of the American Physical Society, Minutes of the Meeting Held at Balti-
more, May 1, 1942, APSM. It is not clear whether Ritchie was a member of the APS at
the time he sent this letter, since he had resigned his membership in 1940. Council of the
American Physical Society, Minutes of the Meeting Held at Philadelphia, December 26,
1940, APSM.
30. Council of the American Physical Society, Minutes of the Meeting Held at Colum-
bia University, June 5, 1943, APSM.
31. Darrow, “Current Trends,” 437–38.
32. Quoted in Darrow, “Current Trends,” 437.
33. Darrow, ironically, was employed by Bell Labs. He joined Western Electric in
1917 and continued on with the advent of the Bell System in 1925. However, he had
made a habit of accepting visiting professorships. Darrow spent a year at Stanford in the
late 1920s, completed stints at the University of Chicago and Columbia University in the
early 1930s, and at the time of these discussions about industry and academia in the early
1940s had been teaching at Smith College, where he spent the spring semesters of 1941
and 1942.
34. As a further example of the force of this policy, a division proposed in 1943 “to
be devoted to the promotion of physical principles as applied to textiles, plastics and rub-
ber,” referred to in early council meeting minutes as the “Division of Textile Physics,” was
compelled to proceed as the “Division of High-Polymer Physics.” Council of the Ameri-
can Physical Society, Minutes of the Meeting Held at Chicago, November 27, 1942, and
Minutes of the Meeting Held at Columbia University, June 5, 1943, APSM.
35. Barton, “Institute Doings,” 4.
36. Governing Board of the American Institute of Physics, Minutes of Meeting
Held March 10, 1944, AIPM, accessed June 19, 2018, http://www.aip.org/history
-programs/niels-bohr-library/collections/governing-board/march-10-1944.
37. “Report of the National Research Council,” 283.
38. Sutton, “Introductory Remarks,” 285.
39. Buckley, “What’s in a Name?” 303.
40. Buckley, “What’s in a Name?” 303.
41. Gibbs, “What Should Be the Method,” 305.
42. Kemble, “What Changes in Graduate Training?” 291.
43. Barnes, “How Can the Place?” 296.
44. Barnes, “How Can the Place?” 300–301. Sawyer was member of the University
of Michigan’s physics faculty, but at the time of this meeting, having joined the US Naval
Reserve, was directing the experimental laboratories of the Naval Proving Ground.
45. Harnwell, “Role of Organization,” 310–11.
46. Quoted in Harnwell, “Role of Organization,” 312.
47. Harnwell, “Role of Organization,” 313.
48. Darrow, “Conclusion,” 327.
49. The others were Karl T. Compton, Homer L. Dodge, Lee A. DuBridge, and Paul
E. Klopsteg.
50. “Preliminary Report of the Policy Committee on the Reorganization of Physics,”
Addendum to the Minutes of the Joint Meeting of the Executive Committee and Policy
Committee, March 8, 1945, AIPS.
NOTES TO PAGES 52–60 219

51. Forman, “Behind Quantum Electronics”; Wang, American Science; Weart, Rise of
Nuclear Fear.
52. Wildhack, “Letter to the Editor,” 271.
53. Council of the American Physical Society, Minutes of the Meeting Held at New
York, September 19, 1946, APSM.
CHAPTER 3: Balkanizing Physics
Epigraph: This excerpt from “La Marseillaise” translates: “that an impure blood
waters our furrows.” The Bureau of Standards in Washington, DC, traditionally hosted
the annual American Physical Society meetings. Their informal character encouraged
attendees to converse freely on the lawn outside. John H. Van Vleck, letter to Roman
Smoluchowski, February 26, 1944, American Physical Society, Division of Solid State
Physics, CRS, folder 1.
1. Weart, “Solid Community,” 618. Weart identifies the formation of physics in the
mid-nineteenth century as another such critical juncture.
2. Weart, “Birth of the Solid-State Physics Community,” 45.
3. Weart, “Solid Community,” 627, 628, 640.
4. Lyons, “Concerning the Division of High-Polymer Physics.” The Division of Elec-
tron and Ion Optics, formed by a group of electron microscopists, had been approved
earlier in 1943.
5. Saul Dushman et al., “The Present War Is a Physicist’s War,” CRS, folder 3. Early
discussion among the group of six and their correspondents referred to a “section”
rather than a “division” of the society. Karl K. Darrow, the APS secretary, corrected this
error once the petition came to his attention, noting: “The word . . . is Division and not
Section—our Sections are the geographically-defined groups of members such as the New
England Section.” Karl Darrow, letter to Roman Smoluchowski, March 22, 1944, CRS,
folder 1. The ease with which most physicists interchanged the two terms demonstrates
that a lack of familiarity with the internal structure of the APS, in particular with divisions
and their goals, remained widespread in the mid-1940s.
6. Roman Smoluchowski, letter to Stanley R. March, July 10, 1947, CRS, folder 4.
7. Roman Smoluchowski, letter to Sidney Siegel, December 17, 1943, CRS, folder 1.
8. Roman Smoluchowski, letter to Conyers Herring, February 15, 1944, CRS, folder
1.
9. John Van Vleck, letter to Saul Dushman, January 29, 1944, CRS, folder 1. Although
Smoluchowski was the primary architect of the effort, the letter instructed recipients to
reply to whomever among the group of six they preferred. Dushman, the first alphabet-
ically, therefore received the preponderance of the replies, which he dutifully passed on
to Smoluchowski.
10. Darrow, “How to Address the APS,” 4.
11. Van Vleck, letter to Dushman, January 29, 1944. Van Vleck, a committed Repub-
lican, in personal correspondence of this period often expressed skepticism of Franklin
Roosevelt’s policies. His reference to the New Deal may be read as derogatory.
12. See, most notably, Van Vleck, Theory of Electric and Magnetic Susceptibilities.
13. For a detailed evaluation of the contributions to the quantum revolution that
grounded Van Vleck’s standing among his peers, see Duncan and Janssen, “On the Verge.”
14. The two were close enough that Darrow felt comfortable writing Van Vleck in
Latin in order to coordinate their nomination of mutual friend Eugene Wigner for fellow-
ship in the American Philosophical Society: “Amor Germanorum, rum cum coca cola,
220 NOTES TO PAGES 60–64

et debilitas memoriae sunt radices multorum malorum. Sicut recte dixisti, Wigner non
est socius noster in Societate Philosphica Americana” [Love of the Germans, rum and
Coca-Cola, and weakness of memory are the roots of many evils. As you have correctly
said, Wigner is not our associate in the American Philosophical Society]. Karl Darrow, let-
ter to John Van Vleck, May 15, 1945, KKDP, box 19. (Translation note: Darrow uses the
genitive “Germanorum,” which literally translates as “possessed by the Germans.” Given
the context—just a week after VE Day—it appears that he would have been better served
by the dative “Germanis,” or “for the Germans.”) Van Vleck also had occasion to stay with
Darrow while the latter was a visiting professor at Smith College in 1941. After Van Vleck
departed, misplacing a pair of suspenders in the process, the two exchanged a series of
letters speculating on whether or not suspenders could be had in the Smith colors and
whether or not this reflected sartorial trends in women’s colleges. John Van Vleck, letter to
Karl Darrow, March 6, 1941; Karl Darrow, letter to Van Vleck, March 13, 1941; and John
Van Vleck, letter to Darrow, March 25, 1941, JHVVP, 1853–1981, box 9.
15. John Van Vleck, letter to Karl Darrow, September 19, 1943, KKDP, box 19, folder
1943 TUV. The reference to the real-estate business refers to the Physical Society’s pur-
chase of a building to house its operations in New York City.
16. Roman Smoluchowski, letter to John Van Vleck, February 3, 1944, CRS, folder 1.
17. Van Vleck, letter to Smoluchowski, February 26, 1944.
18. Roman Smoluchowski, letter to John Van Vleck, February 21, 1944, CRS, folder 1.
19. Frederick Seitz, letter to Karl Darrow, May 6, 1944, CRS, folder 1. These oppo-
nents of divisionalization included, at the least, Van Vleck and Eugene Wigner, the latter
of whom was famously peripatetic in his research interests and would therefore find little
to recommend a topically divided APS.
20. Karl Darrow, letter to Frederick Seitz, May 16, 1944, CRS, folder 1.
21. Frederick Seitz, letter to Karl Darrow, May 25, 1944, CRS, folder 1. The apparent
malleability of Seitz’s convictions and his adeptness at navigating the competing interests
involved carries additional significance in light of his subsequent work on behalf of cor-
porate interests in the face of the scientific consensuses on the dangers of tobacco and
anthropogenic climate change. See Oreskes and Conway, Merchants of Doubt.
22. Darrow, letter to Seitz, May 16, 1944.
23. Frederick Seitz, letter to Roman Smoluchowski, May 25, 1944, CRS, folder 1.
24. Léon Brillouin, letter to Saul Dushman, January 25, 1944, CRS, folder 1. Bril-
louin skirted the similar issue that would later dog solid state physics, namely that the
border between solids and other states of matter could be similarly fuzzy.
25. Roman Smoluchowski, letter to Sidney Siegel, December 17, 1943, CRS, folder 1.
26. Gaylord P. Harnwell, letter to Gerhard Derge, September 20, 1939, GPHP, box 6,
folder 2. Francis Bitter, a physicist hired by MIT’s Department of Mining and Metallurgy
in the mid-1930s who took steps to bend the field to a physicist’s notion of rigor, provides
another example of the position and aspirations of metallurgy in the late 1930s. By the
early 1960s, the study of metals had lost its luster in the face of “a growing trend through-
out the world towards the unified treatment of the science of metallic and non-metallic
materials.” “Finding a Forum,” 361.
27. National Academy of Sciences, Industrial Research Laboratories, 7th ed. (1946),
34. The inclusion of “solid state physics” was new in the 1946 report, having been absent
from 6th edition of 1938.
28. Labs listing solid state physics in 1950 include, in addition to Bell: the Franklin
Institute of the State of Pennsylvania; the Milwaukee Gas Specialty Company; Philips
NOTES TO PAGES 64–69 221

Laboratories, Inc. of Hudson, NY; Westinghouse Electric Corporation (“solid state elec-
tronics”), and Carl A. Zapffe’s Laboratory, of Baltimore, MD. National Research Coun-
cil, Industrial Research Laboratories, 9th ed. (1950). Information for 1956 and 1960 in
National Research Council, Industrial Research Laboratories, 10th ed. (1956) and 11th
ed. (1960).
29. Roman Smoluchowski, letter to Frederick Seitz, May 30, 1944, CRS, folder 1.
30. Frederick Seitz, letter to Roman Smoluchowski, June 14, 1944, CRS, folder 1.
31. Karl Darrow, “Memorandum of a Conversation with R. Smoluchowski—June 16,
1944,” CRS, folder 1. Darrow indicated in the memo that he had with him a copy of
Seitz’s June 14 letter to Smoluchowski, resolving the question of whether or not Smo-
luchowski had an opportunity to read it before talking the matter over with Darrow.
32. Karl Darrow, letter to Roman Smoluchowski, June 28, 1944, CRS, folder 1.
33. Dushman et al., “Physics of the Solid State.”
34. Dushman et al., “Physics of the Solid State,” 791.
35. John Van Vleck, letter to Frederick Seitz, September 9, 1944, CRS, folder 1.
36. In 1944, Journal of Applied Physics published an approximately equal number of
articles from physicists in industry and academia. These two groups together accounted
for about 40 percent of contributions each, with the remainder coming from the govern-
ment sector—mostly concentrated in an issue dedicated to naval research—along with
two articles from Soviet researchers. In contrast, during the same year the Physical Re-
view, by then considered the flagship journal of American physics, published eighty-nine
articles, six of which, less than 7 percent, came from industrial physicists. University-
affiliated physicists accounted for seventy-nine of the eighty-nine Physical Review articles,
or about 89 percent.
37. Roman Smoluchowski, letter to John Van Vleck, February 13, 1945, CRS, folder
2.
38. John Van Vleck et al., letter to Karl Darrow, January 29, 1945, CRS, folder 2.
39. Roman Smoluchowski, letter to John Van Vleck, March 15, 1945, CRS, folder 2.
40. John Van Vleck, letter to Roman Smoluchowski, March 21, 1945, CRS, folder 2.
41. Frederick Seitz, letter to Roman Smoluchowski, March 30, 1945, CRS, folder 2.
42. Seitz, letter to Smoluchowski, March 30, 1945.
43. Roman Smoluchowski, letter to Thomas Read, Saul Dushman, Sidney Siegel,
William Shockley, and Frederick Seitz, December 13, 1946, CRS, folder 4. The commit-
tee also included Karl Darrow, George Pegram, and John T. Tate. Council of the American
Physical Society, Minutes of the Meeting Held at New York, September 19, 1946, APSM.
44. Council of the American Physical Society, Minutes of the Meeting Held at Minne-
apolis, November 30, 1946, APSM.
45. Council of the American Physical Society, November 30, 1946.
46. Council of the American Physical Society, Minutes of the Meeting Held at Wash-
ington, May 2, 1947, APSM.
47. Council of the American Physical Society, May 2, 1947.
48. Council of the American Physical Society, May 2, 1947.
49. Council of the American Physical Society, Minutes of the Meeting Held at Mon-
treal, June 20, 1947, APSM. The Division of High-Polymer Physics had begun with the
policy of accepting associate members who were not members of the APS and who paid
dues directly to the division.
50. Roman Smoluchowski, letter to the Council of the American Physical Society,
September 20, 1946, CRS, folder 4.
222 NOTES TO PAGES 69–77

51. Karl K. Darrow, letter to Roman Smoluchowski, December 4, 1946, CRS, folder
4.
52. Karl K. Darrow, “Formation of a Division of Solid State Physics in the American
Physical Society,” letter to APS membership, May 1947, CRS, folder 4.
53. Darrow, “Formation of a Division.”
54. Wannier, “Statistical Problem.”
55. Van Vleck, “Survey of the Theory,” 30. On the history of the exchange concept,
see Carson, “Peculiar Notion.”
56. Bozorth and Williams, “Effect of Small Stresses.”
57. Zener, “Fracture Stress of Steel.”
58. Breck, “Catalysis.”
59. Bridgman, “Effects of High Hydrostatic Pressure.”
60. Seitz, Modern Theory of Solids. Seitz acknowledged contributions from his
physicist wife, Elizabeth Marshall Seitz, that today would likely extend as far as to merit
coauthorship.
61. Kittel, Introduction to Solid State Physics.
62. Hopfield, “Whatever Happened?” 3.
63. Dresselhaus, Mildred. 3.42J Theory of Solids, Course Notes of Randall M. Rich-
ardson, Fall 1972. Richardson—who I thank for sharing these notes with me—joins oth-
ers with whom I have spoken in recalling this course as a model of clarity, and one of the
highlights of the MIT physics curriculum.
64. Detailed accounts of the transistor’s invention and refinement can be found in
Hoddeson, “Discovery of the Point-Contact Transistor,” and Riordan, Hoddeson, and
Herring, “Invention of the Transistor.”
65. Kittel, Introduction to Solid State Physics, vii.
66. Arthur von Hippel, “New Fields for Electrical Engineering,” AvHP, box 1, folder
44.
CHAPTER 4. The Publication Problem
Epigraph: Alan T. Waterman, letter to Karl K. Darrow, July 5, 1955, Council of the
American Physical Society, Minutes of the Meeting Held at Chicago, November 25 and
26, 1955, APSM.
1. The total membership was between 230 and 240 in April of 1948, according to an
informal assay conducted at the time. Elias Burstein, letter to Karl Darrow, October 19,
1956, APSR, subgroup 2, box 14, folder 11. The APS as a whole reported a membership
of 7,649 in 1948.
2. Exact DSSP membership for 1961 is unavailable, but the division had 670 mem-
bers in 1958 and 951 by 1963. The 5 percent estimate is based on 1958 numbers, the
closest date for which both DSSP and APS membership data are available: that year, the
DSSP enrolled 670 of the Physical Society’s 13,844 members. DSSP membership num-
bers from: Sistina F. Greco, letter to Gaile Dody, March 22, 1963, APSR, subgroup 2,
box 17, folder 10; Karl Darrow, letter to John C. Slater, February 13, 1958, JCSP, folder
Darrow, Karl #3. American Physical Society membership figures from: American Physi-
cal Society, “Historical Membership Counts, 1899–2016,” accessed February 20, 2017,
http://www.aps.org/membership/statistics/index.cfm.
3. Council of the American Physical Society, Minutes of the Meeting at Oak Ridge,
March 18, 1950, APSM. The three divisions were the DSSP, the Division of High Poly-
mer Physics, and the Division of Electron Physics, which had changed its name from
NOTES TO PAGES 77–85 223

“Electron and Ion Optics” in 1948. To this day, the APS holds two major annual meet-
ings. The March meeting is dominated by solid state–style research, and the April meet-
ing by high energy, nuclear, and astrophysics.
4. Council of the American Physical Society, Minutes of the Meeting at Washington,
April 29, 1953, APSM. Van Vleck, at this point, was the past president of the APS. Iron-
ically, given his unreserved advocacy of basic research, he had been made Dean of Engi-
neering and Applied Physics at Harvard in 1951. In 1960 he would become chair of the
DSSP executive committee.
5. Frederick Seitz, letter to Elias Burstein, January 6, 1961, JHVVP, box 35, folder
American Physical Society.
6. Karl Darrow, letter to the Members of the Council of the American Physical Society,
January 18, 1960, APSM.
7. Seitz, letter to Burstein, January 6, 1961.
8. Elias Burstein, letter to DSSP Membership, undated 1956, JCSP, folder Bur-
stein, E.
9. A similar strategy was pursued, with more success, in Europe where Physica Sta-
tus Solidi was first published in 1961. For a summary of this journal’s role in the es-
tablishment of European solid state physics, and especially as a mechanism for scientific
exchange between East and West during the Cold War, see Hoffmann, “Fifty Years of
Physica Status Solidi.”
10. Data collected from the online Physical Review archive at http://prola.aps.org/.
The count includes only full articles, omitting letters, minor contributions, errata, and
editorial notes.
11. See Kaiser, “Cold War Requisitions,” on the boom in physics PhD production.
See Forman, “Behind Quantum Electronics,” on postwar funding patterns.
12. Barton, “Institute Doings,” 4.
13. Gaylord P. Harnwell, letter to Frederick Seitz, September 29, 1950, PTDR, box
3, folder Se.
14. Katcher, “Editorial,” 3.
15. Weiner, “Physics Today.”
16. The articles referenced are, respectively: Siegel and Sinnott, “El Cerrito Cyclo-
tron”; Gamow, “Reality of Neutrinos”; Tisza, “Helium”; Solomon, “Physics and Can-
cer”; Jenson, “Pigtails”; Page, “Origin of the Earth”; Iselin, “Down to the Sea.”
17. Gamow, “Any Physics Tomorrow?” 18.
18. Raymond, “Letter to the Editor,” 5.
19. High-minded speculation was becoming common in certain segments of the
physics community by the middle of the century. For an account that places them in the
context of a longer tradition of such grand theorizing, see Kragh, Higher Speculations.
20. George R. Harrison, letter to Gaylord P. Harnwell, January 20, 1950, GPHP, box
2, folder 22.
21. Sam Goudsmit, letter to Gaylord P. Harnwell, June 29, 1950, GPHP, box 2, folder
22.
22. John H. Van Vleck, letter to Gaylord P. Harnwell, June 27, 1950, GPHP, box 2,
folder 22.
23. Robinson, “Challenge of Industrial Physics,” 5.
24. Bush, “Trends in American Science,” 7.
25. Frederick Seitz, letter to William Shockley, February 28, 1945, WSP, box 1,
Shockley Correspondence, July 25, 1939–March 30, 1948, Volume I.
224 NOTES TO PAGES 85–90

26. Seitz, letter to Shockley, February 28, 1945.


27. “Origin of Radar Countermeasures (Rough Draft),” RRLR, box 4, folder Admin-
istrative Report Accumulation.
28. See Forman, “‘Swords into Ploughshares.’”
29. Henry A. Barton, “American Institute of Physics, Director’s Report for 1954,”
March 12, 1955, FSP, box 1, folder AIP Correspondence #1. These numbers indicate
not only that more people trained as physicists were taking jobs in industry, but also that
the definition of “physicist” had broadened to include scientists who might have been
otherwise classified in the 1930s or early 1940s.
30. Seitz, On the Frontier, 67. This anecdote is also recounted in Shurkin, Broken
Genius, 30–31.
31. For a detailed recounting of Van Vleck’s physical and conceptual peregrinations,
see Fellows, “J. H. Van Vleck.”
32. Wigner, Recollections, 171–79.
33. Seitz, On the Frontier, 58–59. Seitz was Wigner’s first American graduate student.
His second and third were John Bardeen and Conyers Herring. Bardeen would go on to
win two Nobel Prizes for his solid state work and Herring would found the influential
theoretical solid state physics division of Bell Labs.
34. See: Assmus, “Americanization of Molecular Physics,” esp. 22; Duncan and Jans-
sen, “On the Verge”; Schweber, “Young John Clarke Slater”; and Gavroglu and Simões,
“The Americans, the Germans, and the Beginnings.”
35. Rosenfeld, “Men and Ideas,” 77.
36. John Van Vleck, letter to John Slater, November 17, 1971, JHVVP, box 28, folder
615.
37. Slater, “Quantum Physics in America.”
38. The distinctive sound of the Hungarian accent, combined with the preternatu-
ral ability these men showed in science and mathematics, led to the nickname. Hargittai,
Martians of Science.
39. Transcript of talk entitled “My Life as a Physicist,” EPWP, box 13, folder 3.
40. Wigner, “On the Mass Defect of Helium.”
41. Slater, interview by Kuhn and Van Vleck. Nevertheless, Slater did publish a
follow-up to the BKS paper when he returned to the United States: Slater, “Quantum
Theory of Optical Phenomena.” Duncan and Janssen suggest that, despite his conster-
nation with his collaborators in Copenhagen, Slater continued to defend BKS well after
Bohr had resigned himself to the reality of light quanta. Duncan and Janssen, “On the
Verge.” Slater’s subsequent work on solid state and molecular physics, however, which
consisted mostly of conducting ab initio calculations with progressively more powerful
digital computers, is consistent with his self-reported disdain for foundational specula-
tion, which was evidently reinforced by his experience abroad. For an overview of the
Compton effect, see Stuewer, Compton Effect.
42. Slater, interview by Kuhn and Van Vleck.
43. Wigner and Seitz, “On the Constitution of Metallic Sodium.”
44. More detailed discussion of the development of the electron theory of metals can
be found in: Hoddeson and Baym, “Development of the . . . Theory of Metals”; Eckert,
“Propaganda in Science”; and Joas and Katzir, “Analogy, Extension, and Novelty.”
45. The Wigner–Seitz method begins with the lattice structure of a metal and defines
a cell by bisecting the lines between any given lattice point and its nearest neighbor. The
NOTES TO PAGES 90–93 225

points inside the polyhedron defined in this way are closer to that lattice point than any
other and a set of these polyhedrons packs to fill the entire lattice. For solids with highly
symmetrical lattice structures, each of these cells may be approximated by a sphere, allow-
ing Schrödinger’s equation for a solid to be solved relatively simply. Such cells, rendered
in reciprocal space, persist under the familiar moniker “Brillouin zones,” as a nod to his
foundational work, Brillouin, Die Quantenstatistik. See also Hoch, “Development of the
Band Theory,” which also contains a more detailed exposition of the technical features of
the Wigner–Seitz method.
46. Seitz, Modern Theory of Solids.
47. Council of the American Physical Society, Minutes of the Meeting Held at New
York, January 29, 1948, APSM.
48. Council of the American Physical Society,
Minutes of the Meeting Held at Wash-
ington, April 29, 1953, APSM.
49. Council of the American Physical Society, Minutes of the Meeting Held at Wash-
ington, April 27, 1955, APSM. A discussion of the economics of physics publishing
in this era, with an emphasis on page charges, can be found in Scheiding, “Paying for
Knowledge.”
50. Barton, “Director’s Report for 1954.”
51. Frederick Seitz, letter to Henry A. Barton, November 3, 1954, FSP, box 1, folder
AIP Correspondence #1.
52. Alan T. Waterman, letter to Karl K. Darrow, July 5, 1955, APSM.
53. The distribution list of the letter is not available, but the committee proposed con-
tacting “leading solid state physicists and chemical physicists,” naming as representative
of this group: “[LeRoy] Akper, [John] Bardeen, [Harvey] Brooks, [Conyers] Herring,
[Charles] Kittel, [Andrew W.] Lawson, [Humboldt] Levernz, [Gordon] McKay, [Freder-
ick] Seitz, [William] Shockley, [John C.] Slater, Smith, C. S. [Cyril Stanley], [Arthur H.]
Snell, [John] Van Vleck, [Eugene] Wigner.” “Recommendations of the AIP-APS Com-
mittee on Joint Publication Problems with Regard to the Journal of Chemical Physics,”
FSP, box 1, folder AIP Correspondence #1.
54. Council of the American Physical Society, Minutes of the Meeting Held at Chi-
cago, November 25 and 26, 1949, APSM.
55. The Journal of Chemical Physics was predominantly a chemical journal despite
its membership in the AIP family of publications. Of the papers it published, 65 percent
originated in chemistry departments. “Recommendations of the AIP-APS Committee on
Joint Publication Problems with Regard to the Journal of Chemical Physics,” FSP, box 1,
folder AIP Correspondence #1.
56. An illustration: in 1967 Robert Parr, then based at Johns Hopkins, who was deeply
involved with promoting chemical physics in both the American Physical Society and the
American Chemical Society (ACS), observed: “Inspection of the pages of the Journal of
Chemical Physics shows that some chemical physicists are professional chemists, some are
professional physicists. (For example, thirty-two of the papers in the April 1, 1967, issue
of Journal of Chemical Physics are by authors clearly identifiable as chemists, nine by au-
thors clearly identifiable as physicists and twenty-nine by authors not clearly identifiable
as one or the other.) The subject is truly interdisciplinary, although more chemists go
into it than do physicists.” Oral Report to the Physical Sciences Group on May 19, 1967,
RGPP, box 131, folder Chemical Physics program. By 1967, approximately 40 percent
of the members of the ACS’s Division of Physical Chemistry were also members of the
226 NOTES TO PAGES 93–98

APS. Herbert S. Gutowsky, letter to ACS membership, October 11, 1967, AIPK, folder
64:24.
57. Frederick Seitz, letter to select solid state and chemical physicists, June 1, 1955,
FSP, box 1, folder AIP—Correspondence #1.
58. The twenty-two responses preserved in the Seitz papers break down as follows: of
the solid state physicists responding, thirteen opposed the proposal, two favored it, and
three reported no strong opinion. All three chemical physicists responding were in favor.
One additional favorable vote came from a self-identified “other.” The final tally of these
reports: six in favor, thirteen opposed, three with no strong opinion.
59. Harvey Brooks, letter to Frederick Seitz, June 23, 1955, FSP, box 2, folder AIP—
Reorganization of Journals.
60. William Shockley, completed survey: “Opinion of the Recommendations Con-
cerning the Journal of Chemical Physics,” received June 16, 1955, FSP, box 2, folder
AIP—Reorganization of Journals.
61. George E. Pake, completed survey: “Opinion of the Recommendations Concern-
ing the Journal of Chemical Physics,” FSP, box 2, folder AIP—Reorganization of Journals.
Pake’s resistance to the term “solid state” is notable. Pake’s research and education were
representative of self-identified solid state physicists at the time; he had taken a PhD at
Harvard with Edward Mills Purcell and published extensively on nuclear magnetic reso-
nance and paramagnetism. His avoidance of the term, and his use of an uncommon alter-
native, “structure of matter physics,” reflect discomfort with “solid state” among portions
of the community.
62. Walter Kohn, completed survey: “Opinion of the Recommendations Concerning
the Journal of Chemical Physics,” FSP, box 2, folder AIP—Reorganization of Journals.
63. Mody, Instrumental Community.
64. It is therefore a lovely irony that Kohn would win the 1998 Nobel Prize in Chem-
istry, which he shared with John Pople. Kohn was cited for his role in developing density
functional theory (DFT), which had broad relevance for both solid state and chemical
problems. On Kohn and DFT, see Zangwill, “Education of Walter Kohn” and “Hartree
and Thomas.” On the importance of instrumental practice for early nuclear magnetic res-
onance researchers, see Lenoir and Lécuyer, “Instrument Makers.”
65. Hillard B. Huntington, completed survey: “Opinion of the Recommendations
Concerning the Journal of Chemical Physics,” FSP, box 2, folder AIP—Reorganization
of Journals.
66. Conyers Herring, completed survey: “Opinion of the Recommendations Con-
cerning the Journal of Chemical Physics,” FSP, box 2, folder AIP—Reorganization of
Journals.
67. Shirley L. Quimby, completed survey: “Opinion of the Recommendations Con-
cerning the Journal of Chemical Physics,” FSP, box 2, folder AIP—Reorganization of
Journals.
68. Weart, “Solid Community,” 652.
69. See Gavroglu and Simões, Neither Physics nor Chemistry.
70. Harvey Brooks, “Memorandum Concerning the Scope and Aims of the Inter-
national Journal of the Physics and Chemistry of Solids,” February 1956, FSP, box 11,
folder Harvey Brooks.
71. Brooks, “Foreword,” 1.
72. Harvey Brooks, letter to Frederick Seitz, November 15, 1955, FSP, box 11, folder
Harvey Brooks.
NOTES TO PAGES 98–107 227

73. Frederick Seitz, letter to Harvey Brooks, November 19, 1955, FSP, box 11, folder
Harvey Brooks.
74. Council of the American Physical Society, Part of Preliminary Agenda for Janu-
ary 29, 1957, APSM. The statesman in question was President Grover Cleveland. Safire,
“Penumbra of Desuetude.”
75. Council of the American Physical Society, Minutes of the Meeting Held at New
York, January 26, 1960, APSM.
76. Section A was devoted to atomic, molecular, and optical physics, C to nuclear
physics, and D to particle and astrophysics.
CHAPTER 5. Big Solid State Physics at the National Magnet Laboratory
Epigraph: Benjamin Lax, letter to Leland Haworth, May 10, 1967, NMLR.
1. Rush, “US Neutron Facility.”
2. Crease, “National Synchrotron Light Source.”
3. Crease and Westfall, “New Big Science.”
4. Council of the American Physical Society, Minutes of the Meeting Held at Chicago,
November 25 and 26, 1955, APSM. These conferences continue through to the present.
5. For an overview of how big science has been used as a historiographical category,
see Capshew and Rader, “Big Science.” The limitations of the big science framework are
explored in Westfall, “Rethinking Big Science.”
6. As of 1969, the AEC accounted for upward of 90 percent of funding for US high
energy physics, with contributions to the tune of 6 percent from the NSF and about 1
percent from both the DOD and NASA. AEC Authorizing Legislation, Fiscal Year 1970:
Hearings Before the Joint Committee on Atomic Energy, 91st Cong. 86 (April 17 and 18,
1969).
7. Bitter, Magnets, 55. Since Abraham’s textbook was not available in English transla-
tion at the time, Bitter most likely refers to the 1923 German edition: Abraham and Föppl,
Theorie der Elektrizität. The English translation appeared in 1937 as Abraham, Classical
Theory.
8. Bitter enjoyed excellent placement and better timing. Caltech was a leading site
of cosmic ray research in the late 1920s while Bitter was a postdoc there. His stint at
Westinghouse overlapped with one of the lab’s most productive periods of magnetron
research, which fed directly into radar work during the Second World War. Bitter also
arrived at the Cavendish immediately after Chadwick’s discovery of the neutron and the
accompanying boom in atomic theory. See: Xu and Brown, “Early History of Cosmic Ray
Research”; Stephan, “Experts at Play”; and Brown, Neutron and the Bomb. Ferromag-
netism research in the 1930s was particularly lively, and pointed to questions of founda-
tional importance for the subsequent development of solid state physics. See Keith and
Quédec, “Magnetism.”
9. Francis Bitter, “Abstract of the Present State and Possible Developments in Physi-
cal Metallurgy,” ca. 1939, FBP, box 5, folder MIT Magnet Lab.
10. Bitter, “Abstract of the Present State.”
11. Bitter, “Abstract of the Present State.” The appellation “fundamental” was com-
monly employed permissively around this time. Bitter was in accordance with the ac-
cepted usage by suggesting that sciences other than physics could be fundamental. Fred-
erick Seitz, writing just a few years later, identified “fundamental” with “pure” research,
defining it as that “which has intrinsic value as a form of culture.” He further mirrored
elements of Bitter’s definition by suggesting that “physics serves as a source of fundamen-
228 NOTES TO PAGES 108–111

tal knowledge for a majority of the most important fields of engineering.” Seitz, “Whither
American Physics?” 40. Vannevar Bush asked rhetorically in the inaugural issue of Physics
Today, “Who would have expected, looking forward from, say, 1939, to find the United
States Navy vigorously furthering a program in fundamental science, including nucleon-
ics, genetics, and mathematics?” Bush, “Trends in American Science,” 6. Clarke, “Pure
Science,” demonstrates that “fundamental research” took on a range of meanings in the
first half of the twentieth century that did not obey the basic/applied distinction. For fur-
ther discussion, see Martin, “Fundamental Disputations.”
12. “Proposal for a High Field Magnet Laboratory,” September 8, 1958, NMLR, box
1, folder 55.
13. Bitter took a leave of absence from MIT to work on degaussing naval ships during
the Second World War. During this time, his metallurgical magnetism laboratory was dis-
mantled and its resources redistributed to war work. On his return to MIT at the end of
the war, both he and the administration thought it more appropriate to reassemble the
magnetism program under the auspices of the physics department.
14. Francis Bitter, “Dedication of the National Magnet Laboratory,” April 30, 1963,
FBP, box 5, folder NML Dedication Notes.
15. The extent to which it was possible in practice for an installation such as the
NML to be truly devoted to basic research while operating on military funding is a matter
of some debate. See: Forman, “Behind Quantum Electronics”; Leslie, Cold War; Brom-
berg, “Device Physics”; and Wilson, “Consultants.” Forman and Leslie argue that military
interest diverted—or at least inclined—Cold War solid state research away from funda-
mental work. In contrast, Bromberg and Wilson argue that applied military research co-
existed, and indeed interacted constructively, with fundamental theoretical work. I do not
take a position on this debate here. Though I am sympathetic to the case Bromberg and
Wilson advance, it is enough here that a desire to pursue basic research manifested itself
in the laboratory’s goals and operations.
16. National Magnet Laboratory promotional brochure, 1963, FBP, box 5, folder
NML Dedication Notes.
17. “Visiting Scientists and Students,” 1965, NMLR, box 2, folder 17.
18. John C. Slater, letter to Julius A. Stratton, August 8, 1958, FBP box 3, folder High
Field Magnet Facility, No. 1 of 3.
19. “The National Magnet Laboratory and the Technology of the Future,” February
20, 1963, NMLR, box 3.
20. Benjamin Lax to Roman Smoluchowski, March 10, 1965, NMLR, box 3, folder 25.
21. “NML Publications Record,” December 31, 1965, NMLR, box 2, 17. From 1963
to 1965, staff increased from thirty-two to thirty-eight.
22. The remainder were spread over the Journal of Applied Physics (8), Review of Sci-
entific Instruments (4), Applied Physics Letters (2), and one each in the Journal of Chemi-
cal Physics, Journal of Mathematical Physics, Physics of Fluids, and Physics Today.
23. Benjamin Lax, letter to Roman Smoluchowski, March 10, 1965, JCSP, folder
National Academy of Science-National Research Council, Solid State Sciences Panel.
24. Benjamin Lax, letter to George H. Vineyard, March 17, 1967, NMLR, box 2,
folder 18.
25. Lloyd A. Wood, letter to Benjamin Lax, April 19, 1967, NMLR, box 2, folder 18.
26. Lax, letter to Vineyard, March 17, 1967.
27. Lax, letter to the Subcommittee on Science, Research and Development of the
NOTES TO PAGES 111–119 229

Committee on Science and Astronautics, March 5, 1971, NMLR, box 2, folder 4. Lax did
use “coupling” language here, but was careful to write that the results of basic research
could be coupled with applied questions, rather than with the planning, execution, or
funding of the research.
28. “Report of the Advisory Committee of the National Magnet Laboratory,” Febru-
ary 1966, NMLR, box 2, folder 17.
29. “Report of the Advisory Committee of the National Magnet Laboratory,” April
1967, NMLR, box 2, folder 18.
30. AEC Authorizing Legislation, 89.
31. Benjamin Lax, letter to Nicolaas Bloembergen, May 10, 1967, NMLR, box 2,
folder 18. In 1967 the Division of Solid State Physics had 1,193 members, compared to
762 in the Division of Nuclear Physics, the next-largest division. The Division of Parti-
cles and Fields held its inaugural meeting in January 1968 with a charter membership
of 551. W. V. Smith to Division of Solid State Physics Members, January 6, 1967, JCSP,
folder American Philosophical Society, #5; “Proceedings of the American Physical Soci-
ety Meeting #425,” 1967, APSM; M. Davis to Chairmen and Secretary-Treasurers of APS
Divisions, February 15, 1968, APSR, subgroup 2, box 17, folder 10.
32. Lax, letter to Haworth, May 10, 1967.
33. AEC Authorizing Legislation, 106–7.
34. “Publications of the Francis Bitter National Laboratory in 1968,” NMLR, box 2,
folder 20. In 1965, by contrast, the number of publications in the latter two journals more
than doubled those in the former.
35. “Meeting of the Advisory Committee,” April 1969, NMLR, box 2, folder 20.
36. Henry Kolm, letter to Benjamin Lax, June 2, 1973, NMLR, box 2, folder 32.
The bad blood between the two persisted, to the extent that Kolm was moved to paint
a deeply unflattering portrait of Lax on his autobiographical website. “MIT Magnet
Lab (1961–1982),” accessed May 24, 2015, http://henrykolm.weebly.com/mit-magnetic
-lab-1961–82.html. The page is now down, but can be viewed at https://web.archive.org
/web/20161103214555/http://henrykolm.weebly.com/mit-magnetic-lab-1961–82.html.
37. See Kevles, Physicists, esp. 420–21, and Asner, “Linear Model.”
38. Stanford Accelerator Power Supply: Hearing Before the Joint Committee on Atomic
Energy, 88th Cong. 23 (January 29, 1964) (statement of Dr. W. K. H. Panofsky, Director,
Stanford Linear Accelerator Center).
39. AEC Authorizing Legislation, 113.
40. AEC Authorizing Legislation, 118.
41. AEC Authorizing Legislation, 111–12.
42. AEC Authorizing Legislation, 116.
43. AEC Authorizing Legislation, 115.
44. Stevens, “Fundamental Physics,” 175.
CHAPTER 6. Solid State and Materials Science
Epigraph: Arthur von Hippel, text of a lecture given at Brown University, July 20,
1969, AvHP, box 5.
1. “Petition to the Council of the American Physical Society to Request the Forma-
tion of a New Division on the Problems of Physics and Society,” Council of the American
Physical Society, Minutes of the Meeting Held at New York, February 5, 1969, APSM.
The committee was originally proposed as a division, but the council deemed that its
230 NOTES TO PAGES 120–128

scope was too broad and that a division would limit participation. It instead recom-
mended a committee and approved a measure that would allow APS membership dues to
pay for committee members’ travel to relevant meetings. The committee was a precursor
to the Forum on Physics and Society, the history of which is addressed in more detail in
Bridger, Scientists at War.
2. Albert M. Clogston, “The American Physical Society and the Economic Concerns
of American Physicists,” March 10, 1970, in American Physical Society Council, Minutes
of the Meeting Held at Washington, DC, April 26, 1970, APSM.
3. Sherwin and Isenson, First Interim Report, 13.
4. For detailed treatments, see Bensaude-Vincent, “Construction of a Discipline” and
“Concept of Materials.”
5. National Academy of Sciences, Report, 1953–54, 60.
6. National Academy of Sciences, Report, 1957–58, 46.
7. Materials Advisory Board, “Standing Review of Department of Defense Materials
Research and Development Program,” FSP, box 1, folder Air Research and Development
Command, 1952–61 #1.
8. Materials Advisory Board, “Standing Review.”
9. National Academy of Sciences, More Effective Organization, frontmatter.
10. National Academy of Sciences, More Effective Organization, vii.
11. Van Vlack, Elements of Materials Science, vii.
12. The agency was sinusoidally forthright about its emphasis on military research;
its name vacillated between ARPA and DARPA (Defense Advanced Research Projects
Agency). To avoid confusion, I refer to it as ARPA throughout. It was founded as ARPA
in 1958, changed its name to DARPA in 1972, dropped the “D” in 1993, and restored it
in 1996.
13. “Interdisciplinary Laboratories for Basic Research in Materials Sciences,” JCSP,
folder MIT. Dept. of Physics #39.
14. National Academy of Sciences, Advancing Materials Research, 36.
15. Mody and Choi, “From Materials Science to Nanotechnology.”
16. See Schweber, “Empiricist Temper,” on the American style of theory that grew
largely from the school Kemble established.
17. Dresselhaus, interview by Martin, and 3.42J Theory of Solids, Course Notes of
Randall M. Richardson, Fall 1972.
18. John C. Slater, letter to John Kincaid, May 6, 1959, JCSP, folder Kincaid, John
F. #1.
19. “The Interdisciplinary Nature of M.I.T. Research,” JCSP, folder Proposal for a
Materials Center at M.I.T., 1960.
20. Arthur von Hippel, who established the Laboratory for Insulation Research, re-
called choosing an abstruse name as “a camouflage trick . . . to avoid stepping on sensitive
toes by encroaching on the entrenched interests of physicists, chemists, and metallurgists
in the materials field.” Arthur von Hippel, interview by Z. Malek, September 1969, AvHP,
box 1, folder 16.
21. Slater, letter to Kincaid, May 6, 1959.
22. John C. Slater, untitled memorandum, JCSP, folder M.I.T. Dept. of Physics #10.
23. Arthur von Hippel, “New Fields for Electrical Engineering,” AvHP, box 1, folder
44. This was a piece von Hippel prepared for the April 1942 edition of The Tech Engi-
neering News, a periodical published by MIT undergraduates.
NOTES TO PAGES 128–000 231

24. “Proposal for an Expanded Program of Materials Research at the Massachusetts


Institute of Technology, July 12, 1956,” JCSP, folder MIT Materials Research #1. The
AEC made a similar push within the national laboratories, as discussed in Westwick, Na-
tional Labs, 257–58.
25. “Materials Research Program (ca. 1956),” JCSP, folder MIT. Materials Research
#1.
26. Slater, letter to Kincaid, May 6, 1959.
27. John C. Slater, letter to John Kincaid, April 30, 1959, JCSP, folder Kincaid, John
F. #1.
28. John C. Slater, “On the MIT Materials Center,” ca. 1960, JCSP, folder Slater, J. C.
On the MIT Materials Center.
29. Knowles and Leslie, “‘Industrial Versailles,’” argue that the campuses at indus-
trial laboratories such as Bell, General Motors, and IBM mimicked what the architect
Eero Saarinen supposed to be the university model of organizing research, namely a linear
model, in which basic research preceded industrial applications. The rhetoric around
MIT’s IDL reveals similar goals by suggesting that placing basic research in physics and
chemistry alongside materials engineering fields would help to advance ARPA’s technical
aims.
30. ARPA, “Administrative Memo #1,” July 20, 1962, JCSP, folder MIT. Dept. of
Physics #138.
31. Weaire, Solid State Science, x.
32. The earliest use I have found is from 1979, when the computer scientist Gerald
Weinberg attributed the quip to mathematician Frank Harary. Weinberg, Introduction to
General Systems Thinking, 25.
33. Melvin Calvin, letter to Philip Handler, December 1975, in COSMAT, Materials
and Man’s Needs: Summary Report, v.
34. Smith, “Matter versus Materials.”
35. COSMAT, Materials and Man’s Needs, 2:307.
36. Council of the American Physical Society, Minutes of the Meeting Held at Chi-
cago, November 25 and 26, 1949, APSM.
37. American Physical Society Executive Committee, Minutes of the Meeting Held at
Flat Rock, NC, June 18–20, 1990, APSM.
38. “Statement of the Council of the American Physical Society,” April 28, 1970, in
American Physical Society Council, Minutes of the Meeting Held at Washington, DC,
April 26, 1970, APSM.
39. The charter societies included the American Ceramic Society, the American
Chemical Society, the American Institute of Chemical Engineers, the American Society
of Metals, the American Society of Non-Destructive Testing, the Institute of Electrical
and Electronic Engineers, the Society of Manufacturing Engineers, the Society of Plastics
Engineers, the American Society of Mechanical Engineers, and the Metallurgical Society
of AIME. The APS would eventually join the federation in 1995.
40. Council of the American Physical Society, Minutes of the Meeting Held at Chi-
cago, February 3, 1974, APSM.
41. “APS Ad Hoc Committee on Applied Physics,” in Council of the American Phys-
ical Society, Minutes of the Meeting Held at New York, October 25, 1974, APSM.
42. COSMAT, Materials and Man’s Needs: Summary Report, 140.
43. Groenewegen and Peters, “Emergence and Change.”
232 NOTES TO PAGES 136–143

CHAPTER 7. Responses to the Reductionist Worldview


Epigraph: Yang et al., “High-Energy Physics,” 52–53.
1. Stanford Accelerator Power Supply: Hearing Before the Joint Committee on Atomic
Energy, 88th Cong. 23 (January 29, 1964) (statement of Dr. W. K. H Panofsky, Director,
Stanford Linear Accelerator Center).
2. Livingstone, Particle Physics, 6–7. For detailed historical background, see Heilbron
and Seidel, Lawrence and His Laboratory.
3. Through the 1960s and 1970s, the United States and the Soviet Union, in a micro-
cosm of the arms race and the space race, competed to build higher-energy accelerators,
providing an additional incentive for the US particle physics community to pursue the
energy frontier over the intensity frontier. See also Seidel, “Accelerating Science.”
4. See Stevens, “Fundamental Physics.”
5. Boyer, University of Chicago, esp. ch. 4.
6. Weinberg, First Nuclear Era, 3.
7. Weinberg, “Criteria for Scientific Choice,” Minerva; Weinberg, “Criteria for Scien-
tific Choice,” Physics Today, 45.
8. Weinberg, “Criteria for Scientific Choice,” Physics Today, 45.
9. Weinberg, “Criteria for Scientific Choice,” Physics Today, 47.
10. Yang et al., “High-Energy Physics,” 52, 57.
11. Yang et al., “High-Energy Physics,” 55.
12. Weisskopf and Weinberg, “Two Open Letters,” 47. The quark model was pro-
posed by Murray Gell-Mann and George Zweig in 1964 but did not gain experimental
traction until 1967, so Weisskopf ’s identification of nucleons as the basis for all matter
is, in content and rhetoric, equivalent to later claims of the same status for the standard
model and its imagined successors.
13. Weisskopf, “Nuclear Structure,” 24.
14. The assumption that general principles are worked out in one realm before being
applied in another is also prominent within the history of physics. It has contributed to
the impression that solid state physics was a field devoted to mere applications of more
fundamental work. This assumption is challenged in James and Joas, “Subsequent and
Subsidiary?” James and Joas argue that so-called applications of early quantum mechan-
ics made many essential contributions to the foundations of the theory.
15. Hoddeson, “Roots of Solid-State Research,” describes the process by which a
basic research ethos took root at Bell Labs and chronicles how the establishment of solid
state research in this context led to the laissez-faire approach distinctive to the major
American industrial laboratories of the time.
16. Stefan Machlup, letter to William Shockley, May 30, 1955, WSP, box 2, folder
Correspondence 1954 and 1955.
17. On the role of the transistor and integrated circuit in reshaping the American
research landscape, see Mody, Long Arm of Moore’s Law.
18. Numbers obtained by searching for each institution’s full name, in quotes, within
the “Affiliation” field in the APS journals database at https://journals.aps.org/search, ac-
cessed December 21, 2017.
19. Anderson, interview by Kojevnikov.
20. “The Nobel Prize in Physics 1977,” Nobelprize.org, accessed January 7, 2011,
http://www.nobelprize.org/nobel_prizes/physics/laureates/1977/.
NOTES TO PAGES 143–151 233

21. Anderson, “More Is Different—One More Time.”


22. National Academy of Sciences, “Members,” accessed January 4, 2018, http://
www.nasonline.org/member-directory/?referrer=http://nas.nasonline.org/site/Dir
/1742171349?pg=srch&view=basic. Data taken from both current and deceased member
rolls. The National Academy of Sciences database does not indicate deceased members’
sections. They were counted only if they could be identified as unambiguous examples of
solid state, nuclear, or particle physicists.
23. Anderson, “More Is Different,” 393.
24. National Academy of Sciences, Physics in Perspective, 129, 453. Tables I.6 and
IV.1 of this report show $211.7 million in total expenditure for particle physics versus
$56 million for basic condensed matter research.
25. Untitled document, APSR, subgroup 2, box 17, folder 10. Data taken from num-
bers collected following a 1968 membership drive.
26. Anderson, “More Is Different,” 393.
27. Anderson, “More Is Different,” 393.
28. Gordon, Zeiger, and Townes, “Maser.”
29. See Roberts, Nuclear Magnetic Resonance, 74–76.
30. Anderson, “More Is Different,” 394.
31. Anderson, “More Is Different,” 396. This argument has parallels familiar to phi-
losophers of biology; it is a common antireductionist argument that biological processes
do not make sense in terms of genes alone, and that reference to the organismal level,
at least, is required to explain phenomena such as differential fitness. This position is
astutely summarized by the humorist Douglas Adams, who observes: “If you try to take a
cat apart to see how it works, the first thing you have on your hands is a nonworking cat.”
Adams, Salmon of Doubt, 135–36.
32. Anderson, “One More Time,” 1.
33. Anderson, “One More Time,” 4. See also Schmalian, “Failed Theories of
Superconductivity.”
34. Anderson, interview by Kojevnikov.
35. Pippard, “Cat and the Cream,” 40–41.
36. Anderson, interview by Kojevnikov.
37. Cat, “Physicists’ Debates on Unification,” makes a similar observation when as-
sessing Anderson’s position in terms of the unity of physics. Anderson, according to Cat,
saw physics as methodologically (rather than ontologically) unified. If physics could be
unified by methodology, rather than by the reduction to a single set of laws and con-
cepts, then fundamental physical knowledge need not be restricted to the lowest level of
complexity.
38. Cat’s analysis of Anderson’s position in terms of methodological unity is use-
ful when considering the relationship between physics and other sciences. By avoiding
strong claims about the ultimate nature of reality, Anderson shifted the focus to method-
ology as a basis by which physics could be at once internally unified and delineated from
other sciences. Cat, “Physicists’ Debates on Unification,” 99.
39. Anderson, More and Different, 135.
40. “2 From U.S. Among 4 Nobel Science Winners,” New York Times, October 12,
1977.
41. Sullivan, “Physics Prize,” 1.
42. Browne, “Nobel Prizes Are Awarded,” 1.
234 NOTES TO PAGES 152–156

CHAPTER 8. Becoming Condensed Matter Physics


Epigraph: Gray, “New AIP Handbook,” 41.
1. Gray and Billings, American Institute of Physics Handbook.
2. Gray, “New AIP Handbook,” 41.
3. Proctor, “‘-Logos,’ ‘-Ismos,’ and ‘-Ikos,’” 292.
4. Anderson, More and Different, 90.
5. Kragh, Quantum Generations, 366. Kragh has more recently made his own obser-
vations about the importance of names in “Naming the Big Bang.”
6. Kohn, “Essay on Condensed Matter Physics.”
7. Volker Heine, “History,” Theory of Condensed Matter, University of Cambridge,
accessed December 19, 2017, http://www.tcm.phy.cam.ac.uk/about/history/.
8. Weart, “Solid Community,” 651.
9. Quoted in Chicago Tribune, “Yanks Sweep Science Field.” For further discussion,
see Martin, “Prestige Asymmetry.”
10. As in Ferrell, Lee, and Pal, “Magnetic Quenching.”
11. Pierre Teissier notes that in 1971, when American physicists were just beginning
to use “condensed matter” to describe their own activities, Pierre-Gilles de Gennes had
already named his chair at the Collège de France the “Physics of Condensed Matter”
chair. Teissier, “Solid-State Chemistry,” 251.
12. “Important Announcement.” This was not a full-scale fission of the journal, since
it did not yet allow separate subscriptions to each section, but rather a topical grouping
within what was still a single journal. A full-scale split would come several years later in
1970.
13. For example, Guyon et al., “Tunneling into Dirty Superconductors”; Coon
and Fiske, “Josephson ac”; and McFadden, Tahir-Kheli, and Taggart, “Space-Time-
Dependent Correlation,” 854.
14. In 1975, this journal was absorbed into Zeitschrift für Physik when the latter split
into two sections, the second of which was devoted to condensed matter and general
physics.
15. Physik der Kondensierten Materie 1, no. 2 (1963), frontmatter.
16. Hoffmann, “Fifty Years of Physica Status Solidi.” A general term spanning physi-
cal, chemical, biological research, “festkörperforschung” (solid state research), had been
in circulation since the late 1940s as the name of an institute within the Academy of Sci-
ences in East Berlin. Auth, “Das Institut für Festköperphysik,” 27.
17. Höhler, Lyons, and Niekisch, Arbeitstagung Festkörperphysik. Among the partic-
ipants was Karl Wolfgang Böer, who would found the journal Physica Status Solidi in
the early 1960s shortly before immigrating to the United States to assume a post at the
University of Delaware.
18. Pick, “Festkörperphysik.” This is also the year in which the term first appeared in
Annalen der Physik.
19. “Bei der Aufzählung der großen Arbeitsgebiete der neueren physikalischen For-
schung trifft man immer häufiger auf den Begriff Festkörperphysik.” Pick, “Festkörper-
physik,” 346.
20. “Man ist geneigt, ein solches Wort als Überschrift für ein klar umrissenes, ein-
heitliches Arbeitsgebiet zu nehmen. Bei genauerem Zusehen findet sich diese Hoffnung
zunächst aber ganz und gar nicht bestätigt.” Pick, “Festkörperphysik,” 346.
21. Teissier, “Solid-State Chemistry.”
NOTES TO PAGES 157–161 235

22. “les deux approximations fondamentales de la physique du solide et de la chimie.”


Friedel, “Sur L’origine du Ferromagnétisme,” 829.
23. Friedel, interview by Aribart and Bensaude-Vincent. International contacts are
common among early adopters of the term. Russian-born French physicist Lew Kowarski,
who established strong ties with British physicists during the war, produced a UNESCO
report in 1955 that identified “physique de l’état solide” as an area relevant to a prospec-
tive nuclear research facility. Kowarski, “Les Piles de Recherches.”
24. Mott, “Mechanism of Work-Hardening,” 413. For a detailed treatment of the
development of Mott’s research school, see Keith and Hoch, “Formation of a Research
School.”
25. Between 1945 and 1965, papers with at least one author bearing a UK affiliation
(“United Kingdom,” “Great Britain,” “England,” “Scotland,” “Wales,” or “Northern Ire-
land”) appeared 989 times in American Physical Society journals, versus 121 results from
Germany and 258 from France. Data from the APS journal database at journals.aps.org.
26. “Forthcoming Events,” 420.
27. “Important Announcement,” 1.
28. It is plausible that the term was in the air at Brookhaven National Laboratory,
where managing editors Samuel Goudsmit and Simon Pasternack were based. Brook-
haven’s accelerator program provided the opportunity for them to be exposed both to the
technical term and, through communication and collaboration with international labs,
most notably CERN (European Organization for Nuclear Research), to the newly com-
mon European usage. Brookhaven’s collaborative efforts are outlined in Crease, Making
Physics.
29. National Academy of Sciences, Physics: Survey and Outlook, 67. The BCS the-
ory of superconductivity was a major triumph for solid state physics and drew a great
many talented theoreticians into the field. It was published in 1957: Bardeen, Cooper, and
Schrieffer, “Theory of Superconductivity.”
30. “NAS-NRC Physics Survey Committee, Solid State Physics and Condensed Mat-
ter,” draft, April 1964, HBP, series HUGFP 128.13, box 1, folder NAS Survey Committee
March–April 1964. The footnote was cut from the Pake report but would resurface in a
supplement that furnished more detailed reports on the subfields of physics. National
Academy of Sciences, Physics: Survey and Outlook, 143.
31. National Academy of Sciences, Physics: Survey and Outlook, 67.
32. National Academy of Sciences, Physics: Survey and Outlook, 68.
33. National Academy of Sciences, Physics: Survey and Outlook, 67–69.
34. Pippard, “Cat and the Cream,” 40–41.
35. See Weisel, “Plasma Archipelago.”
36. National Academy of Sciences, Physics: Survey and Outlook, 69.
37. National Academy of Sciences, Physics: Survey and Outlook.
38. National Academy of Sciences, Physics in Perspective.
39. Physik der kondensierten Materie 1, no. 1 (1963), frontmatter; “Authors,” 361.
40. National Academy of Sciences, Physics in Perspective, 460.
41. National Academy of Sciences, Physics in Perspective, 458.
42. National Academy of Sciences, Physics in Perspective, 459.
43. Anderson, “More Is Different” 393.
44. Common lore holds that Anderson and Volker Heine coined “condensed matter”
in 1967 while Anderson held a seasonal professorship at the University of Cambridge
and they changed the name of the solid state theory group at the Cavendish Laboratory
236 NOTES TO PAGES 161–165

to “theory of condensed matter.” Earlier occurrences of the term, in particular in the


Springer journal title, belie this simple origin story, but the adoption of the term by a ma-
jor UK research unit no doubt raised its profile in the Anglophone world.
45. American Physical Society Journals, accessed August 13, 2014, http://journals
.aps.org/search. The ratio is starker in AIP journals, with 33 instances of “condensed
matter” and 4,695 of “solid state.” The difference here is amplified by several factors,
including the applied focus of AIP journals during an era that witnessed an explosion
in topics such as solid state masers and lasers and the fact that the AIP search algorithm
includes the titles of citing articles, which generates a high rate of false positives. American
Institute of Physics Journals, accessed August 13, 2014, http://scitation.aip.org/search.
46. Minutes of the American Physical Society Council Meeting, San Francisco, Cali-
fornia, January 22, 1978, APSM.
47. Minutes of the American Physical Society Council Meeting, Washington, DC,
April 23, 1978, APSM.
48. Hopfield, “Whatever Happened?” points to the success of the BCS theory of
superconductivity as the theoretical development that encouraged physicists to see solid
state problems as general physical problems.
49. National Academy of Sciences, Industrial Research Laboratories, 7th ed. (1946).
50. National Academy of Sciences, Industrial Research Laboratories, 11th ed. (1960).
51. Walter Kohn, letter to George Pake, November 13, 1964, HPB, series HUGFP
128.13, box 1, folder NAS Survey Committee May–December 1964.
52. Harvey Brooks, letter to Walter Kohn, March 30, 1964, HBP, series HUGFP
128.13, box 1, folder NAS Survey Committee March–April 1964.
53. Brown, “AT&T and the Consent Decree.”
54. See Mody, Long Arm of Moore’s Law.
55. Turner, “Aspen Physics Turns 50.”
56. Electron density is itself represented by a function, hence the name density func-
tional theory, a functional being a function of a function. For more on Kohn and the devel-
opment and dissemination of DFT, see Zangwill “Education of Walter Kohn” and “Half
Century.”
57. Kohn and Sham, “Self-Consistent Equations.” For further discussion of this point,
see Zangwill, “Half Century.” As of January 1, 2018, journals.aip.org records 26,273
citations to the paper. Google Scholar, which crawls a wider array of sources, records
46,560.
58. A firsthand account of the origins of Santa Fe Institute is available in Cowan,
Manhattan Project, chs. 36–38.
59. Domb, “Critical Phenomena.”
60. “Battille Colloquium, Critical Phenomena: Final Program, Extended Abstracts,
and Agenda Discussion,” LKP, box 1, folder Battille Colloquium on Critical Phenomena,
Program and Abstracts, 1970.
61. University of Chicago press release, August 9, 1945, HAR, box 116, folder 4.
62. Andrew W. Lawson, “The Institute for the Study of Metals after One Decade,”
December 26, 1957, CSSP, box 2, folder 31.
63. Edward H. Levi, memo to Committee on the Budget, April 10, 1967, BAR, box
146, folder 2.
64. Sloan, “Molecularizing Chicago.”
65. Ole Kleppa, letter to Albert V. Crewe, May 23, 1974, LAR, box 173, folder 6.
NOTES TO PAGES 166–173 237

66. National Academy of Sciences, Physics through the 1990s, 3.


67. National Academy of Sciences, Physics: Survey and Outlook, 67; Physics in Per-
spective, 142.
68. See Hoddeson and Baym, “Development of . . . the Theory of Metals”; Eckert,
“Propaganda in Science”; James and Joas, “Subsequent and Subsidiary?”
69. Slater, “Solid State,” 10.
70. National Academy of Sciences, Physics through the 1990s.
71. National Academy of Sciences, Physics through the 1990s, viii, 23.
72. For an overview of the linear model, see Asner, “Linear Model,” and Godin, “Lin-
ear Model.”
73. Roy, “Funding Big Science,” 9.
74. See Edgerton, “Linear Model,” which suggests that the canonical formulation of
the linear model, in which basic research provides the primary and most immediate basis
for technical development, was rarely, if ever, defended.
75. Pippard, “Cat and the Cream.”
CHAPTER 9. Mobilizing against Megascience
Epigraph: Department of Energy’s Superconducting Super Collider Project: Hearing
Before the Subcommittee on Energy Research and Development of the Committee on Energy
and Natural Resources, United States Senate, 102nd Cong. 36 (April 16, 1991) (statement
of Dr. Paul A. Fleury, Director, Physical Research Laboratory, AT&T Bell Laboratories,
Murray Hill, NJ).
1. Lederman, “Fermilab and the Future,” 125.
2. Pondrom, “Fixed Target,” 104. See also Riordan, Hoddeson, and Kolb, Tunnel
Visions, 13–15.
3. Golden, “More Mini-Bangs”; “Hard Choices”; Herman, “Americans Want
Accelerator.”
4. Superconducting Super Collider: Hearings Before the Committee on Science, Space,
and Technology, House of Representatives, 100th Cong. 245 (April 7, 8, and 9, 1987)
(statement of Dr. Steven Weinberg, Theory Group, Physics Department, The University
of Texas at Austin).
5. For discussions of the expedition of Robert O’Hara Burke and William John Wills
from the perspective of the history of science, see Joyce and McCann, Burke and Wills.
6. Hossenfelder, “LHC ‘Nightmare Scenario.’”
7. Riordan, Hoddeson, and Kolb, Tunnel Visions, esp. ch. 1.
8. Hughes, “Making Dollars.”
9. For thorough accounts of the factors responsible for the SSC’s demise, see: Ritson,
“Demise of the Texas Supercollider,” a snap reaction that faults lack of trust in academic
expertise on the part of federal patrons and industrial managers; Kevles, Physicists, ix–xlii,
which discusses the changes wrought by the end of the Cold War and the rise of an un-
sympathetic freshman congressional delegation; Riordan, “Tale of Two Cultures,” which
examines dysfunction within the SSC’s management structure; and Hoddeson and Kolb,
“Superconducting Super Collider’s Frontier Outpost,” which exposes the early discon-
nect between the vision of the physicists designing the SSC and the inclinations of those
commanding the federal purse strings.
10. AEC Authorizing Legislation, Fiscal Year 1970: Hearings Before the Joint Com-
mittee on Atomic Energy, 91st Cong. 116 (April 17 and 18, 1969).
238 NOTES TO PAGES 174–181

11. Fleury did indeed say “non-zero-sum gain,” rather than “game,” the more com-
mon idiom, to indicate the argument that one field could enjoy outsized funding gains
without impoverishing other areas.
12. On earlier uses of spin-off claims, see Hoddeson, Kolb, and Westfall, Fermilab,
54–55.
13. Lederman, “Value of Fundamental Science,” 42.
14. For a detailed discussion of the political shifts around the end of the Cold War and
how they influenced the SSC, see Kevles, Physicists.
15. Energy and Water Development Appropriations for 1989: Hearings Before the
Subcommittee on Energy and Water Development of the Committee on Appropriations,
House of Representatives, 100th Cong. 72 (March 10, 1988).
16. Fiscal Year 1987 Department of Energy Authorization: Hearings Before the Sub-
committee on Energy Research and Production of the Committee on Science and Technology,
House of Representatives, 99th Cong. 65 (March 5, 1986). On science, politics, and SDI,
see Bridger, Scientists at War, ch. 9, and Slayton, “Discursive Choices.”
17. Fiscal Year 1987 Department of Energy Authorization, 228.
18. Fiscal Year 1987 Department of Energy Authorization, 4.
19. Proposed Fiscal Year 1990 Budget Request (DOE’s Office of Energy Research):
Hearing Before the Subcommittee on Energy Research and Development of the Committee
on Energy and Natural Resources, United States Senate, 101st Cong. 100 (February 24,
1989) (written statement of Dr. Roy F. Schwitters, Director, SSC Laboratory).
20. Proposed Fiscal Year 1990 Budget Request, 102.
21. Importance and Status of the Superconducting Super Collider: Joint Hearing Be-
fore the Committee on Energy and Natural Resources and the Subcommittee on Energy
and Water Development of the Committee on Appropriations, United States Senate, 102nd
Cong. 27 (June 30, 1992) (statement of George F. Smoot III, Scientist, Lawrence Berke-
ley Laboratory, Berkeley, CA).
22. Importance and Status of the SSC, 25–26.
23. Bloembergen, Encounters in Magnetic Resonance; Bromberg, Laser in America.
24. Department of Energy’s Superconducting Super Collider Project, 43 (statement of
Dr. Nicolaas Bloembergen, President, American Physical Society).
25. Nicolaas Bloembergen, letter to Richard A. Carrigan Jr., May 21, 1991, in Impor-
tance and Status of the SSC, 12.
26. Howard Wolpe and Sherwood Boehlert, letter to James D. Watkins, in Importance
and Status of the SSC, 5–11.
27. Krumhansl, “Unity,” 38.
28. Browne, “Big Science.”
29. Importance and Status of the SSC, 13.
30. Superconducting Super Collider: Joint Hearing Before the Committee on Energy
and Natural Resources and the Subcommittee on Energy and Water Development of the
Committee on Appropriations, United States Senate, 103rd Cong. (August 4, 1993) (writ-
ten testimony of Rustum Roy), 10.
31. Assmus, “To Most Physicists.”
32. Schrage, “Glimpses of Truth.”
33. Bazell, “Quark Barrel Politics.”
34. Energy and Water Development Appropriations for 1989, 288.
35. The Superconducting Super Collider Project: Hearing Before the Committee on Sci-
ence, Space, and Technology, House of Representatives, 103rd Cong. 101 (May 26, 1993).
NOTES TO PAGES 181–191 239

36. Schweber, “Physics, Community, and the Crisis,” 40.


37. Importance and Status of the SSC, 24 (statement of Leon M. Lederman, Director
Emeritus, Fermi National Accelerator Laboratory, Batavia, IL). The misspelling of “Mi-
letus,” the Greek city that was home to Thales and is widely considered to have been the
cradle of Greek science and philosophy, is attributable to a transcription error rather than
to Lederman himself.
38. Fiscal Year 1988 Department of Energy Authorization: Hearing Before the Sub-
committee on Energy Research and Development of the Committee on Science, Space, and
Technology, House of Representatives, 100th Cong. 143 (March 3 and 4, 1987).
39. Barry, “Searching for Particles of Logic.”
40. Weinberg, Dreams.
41. Superconducting Super Collider: Joint Hearing, 51 (statement of Dr. Steven Wein-
berg, Theory Group, Department of Physics, University of Texas at Austin).
42. Superconducting Super Collider: Joint Hearing, 58 (statement of Dr. Steven Wein-
berg, Theory Group, Department of Physics, University of Texas at Austin).
43. The extremity of the reductionism espoused in the context of the SSC debates has
been emphasized by Barbara L. Whitten, who argues that “Lederman and [Sheldon] Gla-
show, firmly entrenched in the reductionist paradigm and totally unable to hear what the
others are saying, resemble straw men constructed by feminist critics to display arrogant
androcentric science at its most glaring.” Whitten, “What Physics Is Fundamental,” 10.
44. Superconducting Super Collider: Joint Hearing, 57 (statement of Dr. Philip W.
Anderson, Department of Physics, Princeton University, Princeton, NJ).
45. Establishing Priorities in Science Funding: Hearing Before the Task Force on De-
fense, Foreign Policy and Space of the Committee on the Budget, House of Representatives,
102nd Cong. 64 (July 11 and 18, 1991) (statement of Philip W. Anderson, Ph.D., Joseph
Henry Professor, Princeton University). The word “more,” which was omitted in the tran-
script, can be reasonably inferred.
46. The SSC Project, 109.
47. Weinberg, Dreams; Lederman and Teresi, God Particle.
48. Weinberg, “Answer.”
49. Dutta, “Supercollider Is Science.”
50. Establishing Priorities in Science Funding, 2.
51. Superconducting Super Collider: Joint Hearing, 13.
52. For further discussions of the SSC debate in terms of unification, see Cat, “Phys-
icists’ Debates.”
53. Cranberg, “Paradoxical ‘Unities,’” 102.
54. Kadanoff, “Complex Structures,” 10.
55. Kadanoff, “Cathedrals and Other Edifices,” 9.
56. Kadanoff, “The Big, the Bad,” 11.
57. Anderson, “Is Complexity Physics?” 9.
58. As Cat has argued, the reductionist unity particle physicists sought was counter-
productive for the goal of methodological unity. Cat, “Physicists’ Debates.”
59. The SSC Project, 109.
60. Proposed Fiscal Year 1990 Budget Request, 134.
61. Superconducting Super Collider: Hearings, 906 (written statement of Dr. Philip
W. Anderson, Joseph Henry Professor of Physics, Princeton University, Princeton, NJ).
62. Superconducting Super Collider: Hearings, 334 (written statement of Dr. James A.
Krumhansl, Horace White Professor of Physics, Cornell University, Ithaca, NY).
240 NOTES TO PAGES 191–201

63. Superconducting Super Collider: Hearings, 906 (written statement of Dr. Philip
W. Anderson).
64. Varma, “Changing Research Cultures.”
65. National Science Foundation, “NSF Funding History by Account and FY,
Constant Dollars,” accessed April 5, 2017, https://dellweb.bfa.nsf.gov/NSFFundingby
AccountConstantDollars.pdf.
66. US Department of Energy, “BES Budget,” accessed April 5, 2017, https://science
.energy.gov/bes/about/bes-budget/.
67. Weisel, “Properties and Phenomena,” is an edifying account of the conception of
“basic research” that developed in plasma physics, which faced many of the same chal-
lenges as solid state physics, including the carefully cultivated perception that the big
questions of physics were the unique province of high energy physics and cosmology.
68. Proposed Fiscal Year 1990 Budget Request, 158 (AAAS Panel Funding and the
Academic Physical Sciences, January 17, 1989).
69. MacLeod and Radick, “Ownership in the Technosciences.”
70. Rowall, “Condensed Matter Physics.”
71. Rowall, “Condensed Matter Physics,” 45.
72. Superconducting Super Collider: Hearings, 904 (written statement of Philip
W. Anderson, Joseph Henry Professor of Physics, Princeton University, Princeton, NJ).
73. Proposed Fiscal Year 1990 Budget Request, 137 (written statement of Philip
W. Anderson, Joseph Henry Professor of Physics, Princeton University, Princeton, NJ).
74. Hoddeson and Kolb, “Superconducting Super Collider’s Frontier Outpost.” See
also Martin, “Prestige Asymmetry.”
75. Proposed Fiscal Year 1990 Budget Request, 138 (written statement of Philip
W. Anderson).
76. Arthur L. Schawlow, letter to Felix Bloch, October 18, 1967, FBPS, box 8, folder
15.
77. For argument of continuity across each of these shifts, see: Shapin, Scientific Rev-
olution; Taltavull, “Transmitting Knowledge”; and Nye, Before Big Science.
78. Kevles, Physicists, xii.
79. Doing, Velvet Revolution; Crease, “National Synchrotron Light Source”; Rush,
“US Neutron Facility”; Westfall, “Institutional Persistence.” For a perspective on the Eu-
ropean context, see Heinze, Hallonsten, and Heinecke, “From Periphery to Center” and
“Turning the Ship.”
80. Crease and Westfall, “New Big Science.”
CONCLUSIONS
Epigraph: COSMAT, Materials and Man’s Needs, 4:10.
1. The membership of the American Physical Society in 2000 was 41,570. It sur-
passed 36,000, one thousand times its charter membership, in 1985.
2. Shinbrot, “Editorial.” Physical Review Applied followed quickly on the heels of
Physical Review X, the APS contribution to online, open-access publishing, which boasts
broad coverage of “pure, applied, and interdisciplinary physics” reminiscent of the early
years of the original Physical Review. Data on the Physical Review family of journals can
be found at https://journals.aps.org/about.
3. Attentive readers will note that this is apparently contradicted by Arthur von Hip-
pel’s statement, reported in chapter 6, to the effect that the boundary between physics and
NOTES TO PAGES 203–211 241

electrical engineering was crumbling. The sentiment von Hippel expressed, although in-
dicative of a local pattern of encouraging interdisciplinary exchange at MIT, was far from
being sanctified by so mighty a church as the Physical Review in the 1940s.
4. Forman, “On the Historical Forms” and “Primacy of Science.”
5. Eisler, “‘Ennobling Unity’”; McCray, “Will Small Be Beautiful?”; Mody, Instru-
mental Community.
6. The last division to be formed in the twentieth century, representing the physics of
particle beams, was established in 1989.
7. Christian Joas observes that common anecdotes about arrogant quantum or parti-
cle physicists making snide remarks deriding “squalid state physics” (Murray Gell-Mann)
or “Schmutzphysik” (Wolfgang Pauli) are trafficked extensively by solid state physicists
themselves, indicating that an oppositional attitude was part of the field’s identity. Joas,
“Campos que interagem.”
8. Superconducting Super Collider: Hearings Before the Committee on Science, Space,
and Technology, House of Representatives, 100th Cong. 244 (April 7, 8, and 9, 1987)
(written statement of Dr. Daniel Kleppner, Lester Wolfe Professor of Physics and Associ-
ate Director of the Research Laboratory of Electronics, Massachusetts Institute of Tech-
nology, Cambridge, MA).
9. Pais, Inward Bound.
10. On the theory-experiment relationship in high energy physics, see Pickering, Con-
structing Quarks, and Galison, How Experiments End, and Image and Logic. On Feynman
diagrams, see Kaiser, Drawing Theories Apart. For accounts of large accelerator labora-
tories, see Traweek, Beamtimes and Lifetimes, Heilbron and Seidel, Lawrence and His
Laboratory, and Hoddeson, Kolb, and Westfall, Fermilab.
11. There has been a recent uptick in interest in counterfactual history of science,
seeking to revive the technique as a flexible but neglected tool for throwing light on other-
wise obscure features of the historical process. For an overview, see Radick, “Presidential
Address.”
12. On the character and consequences of the Cold War population boom, see Kai-
ser, American Physics. A summary and chapter outline are available at http://web.mit.edu
/dikaiser/www/CWB.html.
13. On how this picture of the world influenced science and technology in the Cold
War, see Edwards, Closed World.
BIBLIOGRAPHY

Abraham, Max. The Classical Theory of Electricity and Magnetism. London: Blackie
& Sons, 1937.
Abraham, Max, and August Föppl. Theorie der Elektrizität. Leipzig: B. G. Teubner,
1923.
Adams, Douglas. The Salmon of Doubt: Hitchhiking the Galaxy One Last Time. New
York: Random House, 2002.
Ambrosio, Chiara. “The Historicity of Peirce’s Classification of the Sciences.” Eu-
ropean Journal of Pragmatism and American Philosophy 8, no. 2 (2016), 9–43.
Anderson, Philip W. A Career in Theoretical Physics. 2nd ed. Princeton, NJ: World
Scientific, 2004.
Anderson, Philip W. Interview by Alexei Kojevnikov. November 23, 1999. Niels Bohr
Library and Archives, College Park, MD. http://www.aip.org/history-programs
/niels-bohr-library/oral-histories/233623.
Anderson, Philip W. “Is Complexity Physics? Is It Science? What Is It?” Physics
Today 44, no. 7 (1991), 9–11.
Anderson, Philip W. More and Different: Notes from a Thoughtful Curmudgeon. Sin-
gapore: World Scientific, 2011.
Anderson, Philip W. “More Is Different.” Science 177, no. 4047 (1972), 393–96.
Anderson, Philip W. “More Is Different—One More Time.” In More Is Different:
Fifty Years of Condensed Matter Physics, edited by N. Phuan Ong and Ravin
N. Bhatt, 1–9. Princeton, NJ: Princeton University Press, 2001.
“Anecdotal History: The Twenty-Fifth Anniversary Celebration.” Journal of the
Acoustical Society of America 26 (1954), 874–905.
Asner, Glen R. “The Linear Model of the U.S. Department of Defense, and the
Golden Age of Industrial Research.” In The Science-Industry Nexus: History,
Policy Implications, edited by Karl Grandin, Nina Wormbs, and Sven Widmalm,
3–30. Sagamore Beach, MA: Watson, 2004.
Assmus, Alexi. “The Americanization of Molecular Physics.” Historical Studies in
the Physical and Biological Sciences 23, no. 1 (1992), 1–34.
Assmus, Alexi. “To Most Physicists, SSC Isn’t Worth the Cost.” Wall Street Journal,
May 5, 1989.
244 BIBLIOGRAPHY

Auth, Joachim. “Das Institut Für Festköperphysik der Deutschen Akademie der Wis-
senschaften in den Anfangsjahren der Akademie nach 1945.” Sitzungsberichte der
Leibniz-Sozietät 15 (1996), 27.
“Authors.” IBM Journal of Research and Development 8 (1964), 361–64.
Bardeen, John, Leon N. Cooper, and J. Robert Schrieffer. “Theory of Superconduc-
tivity.” Physical Review 105, no. 5 (1957), 1175–204.
Barnes, R. Bowling. “How Can the Place of the Physicist in Industry Be Realized
More Fully?” Review of Scientific Instruments 15, no. 11 (1944), 296–301.
Barry, Dave. “Searching for Particles of Logic in the Collider.” Chicago Tribune, No-
vember 1, 1987.
Barton, Henry A. “American Institute of Physics, Report of the Director.” Physics
Today 1, no. 5 (1948), 29–31.
Barton, Henry A. “American Institute of Physics: Report of the Director for 1943.”
Review of Scientific Instruments 15, no. 5 (1944), 130–33.
Barton, Henry A. “Institute Doings.” Physics Today 1, no. 1 (1948), 4.
Bazell, Robert. “Quark Barrel Politics.” New Republic, June 22, 1987, 9–10.
Bensaude-Vincent, Bernadette. “The Concept of Materials in Historical Perspec-
tive.” NTM Zeitschrift für Geschichte der Wissenschaften, Technik und Medizin
19, no. 1 (2011), 107–23.
Bensaude-Vincent, Bernadette. “The Construction of a Discipline: Materials Science
in the United States.” Historical Studies in the Physical and Biological Sciences
32, no. 2 (2001), 223–48.
Bitter, Francis. Magnets: The Education of a Physicist. New York: Doubleday, 1959.
Bloembergen, Nicolaas, ed. Encounters in Magnetic Resonance: Selected Papers of
Nicolaas Bloembergen (with Commentary). Singapore: World Scientific, 1996.
Boyer, John W. The University of Chicago: A History. Chicago: University of Chicago
Press, 2015.
Bozorth, Richard M., and Howell J. Williams. “Effect of Small Stresses on Magnetic
Properties.” Reviews of Modern Physics 17, no. 1 (1945), 62–80.
Breck, Otto. “Catalysis: A Challenge to the Physicist.” Reviews of Modern Physics 17,
no. 1 (1945), 61–71.
Bridger, Sarah. Scientists at War: The Ethics of Cold War Weapons Research. Cam-
bridge, MA: Harvard University Press, 2015.
Bridgman, Percy W. “Effects of High Hydrostatic Pressure on the Plastic Properties
of Metals.” Reviews of Modern Physics 17, no. 1 (1945), 3–14.
Brillouin, Léon. Die Quantenstatistik und ihre Anwendung auf die Elektronentheorie
der Metalle. Berlin: Springer, 1931.
BIBLIOGRAPHY 245

Bromberg, Joan Lisa. “Device Physics vis-à-vis Fundamental Physics in Cold War
America: The Case of Quantum Optics.” Isis 97 (2006), 237–59.
Bromberg, Joan Lisa. The Laser in America. Cambridge, MA: MIT Press, 1991.
Brooks, Harvey. “Foreword.” Physics and Chemistry of Solids 1, no. 1 (1956), 1.
Brown, Andrew. The Neutron and the Bomb: A Biography of Sir James Chadwick.
Oxford: Oxford University Press, 1997.
Brown, Charles L. “AT&T and the Consent Decree.” Telecommunications Policy 7,
no. 2 (1983), 91–95.
Brown, Laurie M., Max Dresden, and Lillian Hoddeson. “Pions to Quarks: Particle
Physics in the 1950s.” In Pions to Quarks: Particle Physics in the 1950s, edited
by Laurie Brown, Max Dresden, and Lillian Hoddeson, 3–39. Cambridge: Cam-
bridge University Press, 1989.
Brown, Louis. A Radar History of World War II: Technical and Military Imperatives.
Bristol: Institute of Physics, 1999.
Browne, Malcolm W. “Big Science; Is It Worth the Price? A Periodic Look at the
Largest New Research Projects.” New York Times, May 29, 1990.
Browne, Malcolm W. “Nobel Prizes Are Awarded to 3 Physicists and 2 Chemists.”
New York Times, October 16, 1979.
Buckley, Oliver E. “What’s in a Name?” Review of Scientific Instruments 15, no. 11
(1944), 301–4.
Bush, Vannevar. “Science—The Endless Frontier: A Report to the President by Van-
nevar Bush, Director of the Office of Scientific Research and Development, July
1945.” http://www.nsf.gov/od/lpa/nsf50/vbush1945.htm.
Bush, Vannevar. “Trends in American Science.” Physics Today 1, no. 1 (1948), 5–7,
39.
Cahan, David. An Institute for an Empire: The Physicalisch-Technische Reichsanstalt
1871–1918. Cambridge: Cambridge University Press, 1989.
Capshew, James H., and Karen A. Rader. “Big Science: Price to the Present.” Osiris
7 (1992), 2–25.
Cardwell, Donald. From Watt to Clausius: The Rise of Thermodynamics in the Early
Industrial Age. Ithaca, NY: Cornell University Press, 1971.
Carson, Cathryn. “The Peculiar Notion of Exchange Forces.” Pts. 1 and 2, “I: Or-
igins in Quantum Mechanics, 1926–1928,” “II: From Nuclear Forces to QED,
1929–1950.” Studies in History and Philosophy of Modern Physics 27, no. 1
(1996), 23–45; no. 2 (1996), 99–131.
Carty, John J. “The Relation of Pure Science to Industrial Research.” Science 44
(1916), 511–18.
246 BIBLIOGRAPHY

Cat, Jordi. “The Physicists’ Debates on Unification in Physics at the End of the 20th
Century.” Historical Studies in the Physical and Biological Sciences 28, no. 2
(1998), 253–99.
Choi, Hyungsub. “The Boundaries of Industrial Research: Making Transistors at
RCA, 1948–1960.” Technology and Culture 48, no. 4 (2007), 758–82.
Clarke, Sabine. “Pure Science with a Practical Aim: The Meanings of Fundamental
Research in Britain, circa 1916–1950.” Isis 101, no. 2 (2010), 285–311.
Coben, Stanley. “The Scientific Establishment and the Transmission of Quantum
Mechanics to the United States, 1919–32.” American Historical Review 76, no.
2 (1971), 442–66.
Committee on the Survey of Materials Science and Engineering (COSMAT). Mate-
rials and Man’s Needs, 4 vols. Washington, DC: National Academy of Sciences,
1975.
Committee on the Survey of Materials Science and Engineering (COSMAT). Materi-
als and Man’s Needs: Materials Science and Engineering: Summary Report of the
Committee on the Survey of Materials Science and Engineering. Washington, DC:
National Academy of Sciences, 1975.
Compton, Karl T. “The Founding of the American Institute of Physics.” Physics To-
day 5, no. 2. (1952), 4–7.
Coon, D. D., and Milan D. Fiske. “Josephson ac and Step Structure in the Supercur-
rent Tunneling Characteristic.” Physical Review 138, no. 3A (1965), A744–46.
Cooper, Leon N., and Dmitri Feldman, eds. BCS: 50 Years. Singapore: World Scien-
tific, 2011.
Coulter, John M. “The Role of Science in Modern Civilization.” Transactions of the
Illinois State Academy of Science 11 (1918), 19–28.
Cowan, George A. Manhattan Project to the Santa Fe Institute: The Memoirs of
George A. Cowan. Albuquerque: University of New Mexico Press, 2010.
Cranberg, Lawrence. “The Paradoxical ‘Unities’ of Physics.” Physics Today 44, no. 3
(1991), 102.
Crease, Robert P. Making Physics: A Biography of Brookhaven National Laboratory,
1946–1972. Chicago: University of Chicago Press, 1999.
Crease, Robert P. “The National Synchrotron Light Source.” Pts. 1 and 2, “Part I:
Bright Idea,” “Part II: The Bakeout.” Physics in Perspective 10, no. 4 (2008),
438–67; 11, no. 1 (2009), 15–45.
Crease, Robert P., and Catherine Westfall. “The New Big Science.” Physics Today 69,
no. 5 (2016), 30–36.
Csiszar, Alex. “Seriality and the Search for Order: Scientific Print and Its Problems
during the Late Nineteenth Century.” History of Science 48 (2010), 399–434.
BIBLIOGRAPHY 247

Daniels, Mario, and John Krige. “Beyond the Reach of Regulation? ‘Basic’ and ‘Ap-
plied’ Research in Early Cold War America.” Technology and Culture 59, no. 2
(2018), 226–50.
Darrow, Karl K. “Conclusion.” Review of Scientific Instruments 15, no. 11 (1944),
327.
Darrow, Karl K. “Current Trends in the American Physical Society.” Journal of Ap-
plied Physics 14, no. 9 (1943), 437–42.
Darrow, Karl K. “How to Address the American Physical Society.” Physics Today 4,
no. 2 (1951), 4–8.
Darrow, Karl K. “Symposium on the Solid State.” Reviews of Modern Physics 17, no.
1 (1945), 2.
Davis, Harry M. “Radar—Our Miracle Ally.” New York Times, May 23, 1942.
Doel, Ronald E. “Scientists as Policymakers, Advisors, and Intelligence Agents:
Linking Contemporary Diplomatic History with the History of Contemporary
Science.” In The Historiography of Contemporary Science and Technology, edited
by Thomas Söderqvist, 215–44. Amsterdam: Harwood, 1997.
Doing, Park. The Velvet Revolution at the Synchrotron: Biology, Physics, and Change
in Science. Cambridge, MA: MIT Press: 2009.
Domb, Cyril. “Critical Phenomena: A Brief Historical Survey.” Contemporary Physics
26, no. 1 (1985), 47–72.
Donaldson, Rene, Richard Gustafson, and Frank Paige, eds. Proceedings of the 1982
DPF Summer Study on Elementary Particle Physics and Future Facilities. Wash-
ington, DC: US Department of Energy, 1983.
Dresselhaus, Mildred. Interview by Joseph D. Martin. June 24, 2014. Niels Bohr Li-
brary and Archives, College Park, MD.
Duncan, Anthony, and Michel Janssen. “On the Verge of Umdeutung in Minnesota:
Van Vleck and the Correspondence Principle.” Pts. 1 and 2. Archive for History
of Exact Sciences 61, no. 6 (2007), 553–624; 625–71.
Dupree, Hunter A. Science in the Federal Government. Baltimore: Johns Hopkins
University Press, 1957.
Dushman, Saul, Thomas A. Read, Frederick Seitz, William Shockley, Sidney Siegel,
and Roman Smoluchowski. “Physics of the Solid State.” Journal of Applied Phys-
ics 15, no. 12 (1944), 791.
Dutta, Pulak. “Supercollider Is Science out on a Limb.” New York Times, March 15,
1993.
Eckert, Michael. “Propaganda in Science: Sommerfeld and the Spread of the Elec-
tron Theory of Metals.” Historical Studies in the Physical and Biological Sciences
17, no. 2 (1987), 191–233.
248 BIBLIOGRAPHY

Eckert, Michael, and Helmut Schubert. Crystals, Electrons, Transistors: From Schol-
ar’s Study to Industrial Research, translated by Thomas Hughes. New York:
American Institute of Physics, [1986] 1990.
Edgerton, David. “The Linear Model Did Not Exist.” In The Science-Industry Nexus:
History, Policy Implications, edited by Karl Grandin, Nina Wormbs, and Sven
Widmalm, 31–58. Sagamore Beach, MA: Watson, 2004.
Edgerton, David. “Time, Money, and History.” Isis 103, no. 3 (2012), 316–27.
“Editorial.” Journal of Rheology 1, no. 1 (1929), 93–95.
Edwards, Paul N. The Closed World: Computers and the Politics of Discourse in Cold
War America. Cambridge, MA: MIT Press, 1997.
Eisler, Matthew N. “‘The Ennobling Unity of Science and Technology’: Materials
Sciences and Engineering, the Department of Energy, and the Nanotechnology
Enigma.” Minerva 51, no. 2 (2013), 225–51.
England, J. Merton. A Patron for Pure Science: The National Science Foundation’s
Formative Years, 1945–57. Washington, DC: National Science Foundation,
1983.
Fellows, Frederick Hugh. “J. H. Van Vleck: Early Life and Work of a Mathematical
Physicist.” PhD diss., University of Minnesota, 1985.
Ferrell, Richard A., Y. C. Lee, and M. K. Pal. “Magnetic Quenching of Hyperfine
Depolarization of Positive Muons.” Physical Review 118, no. 1 (1960), 317–19.
“Finding a Forum for Materials Science.” New Scientist, November 15, 1962, 361–62.
Finkbeiner, Ann. The Jasons: The Secret History of Science’s Postwar Elite. New
York: Penguin Press, 2006.
Fisher, Arthur. “The Magnetism of a Shared Facility.” Mosaic 11, no. 2 (1980), 38–45.
Forman, Paul. “Behind Quantum Electronics: National Security as a Basis for Physi-
cal Research in the United States, 1940–1960.” Historical Studies in the Physical
and Biological Sciences 18 (1987), 149–229.
Forman, Paul. “On the Historical Forms of Knowledge Production and Curation:
Modernity Entailed Disciplinarity, Postmodernity Entails Antidisciplinarity.”
Osiris 27 (2012), 56–97.
Forman, Paul. “The Primacy of Science in Modernity, of Technology in Postmoder-
nity, and of Ideology in the History of Technology.” History and Technology 23,
no. 1 (2007), 1–152.
Forman, Paul. “‘Swords into Ploughshares’: Breaking New Ground with Radar
Hardware and Technique in Physical Research after World War II.” Reviews of
Modern Physics 67, no. 2 (1995), 397–455.
Forman, Paul. “Weimar Culture, Causality, and Quantum Theory: Adaptation by
BIBLIOGRAPHY 249

German Physicists and Mathematicians to a Hostile Environment.” Historical


Studies in the Physical Sciences 3 (1971), 1–115.
“Forthcoming Events.” Nature 163, no. 4141 (1949), 419–20.
Friedel, Jacques. Interview by Hervé Aribart and Bernadette Bensaude-Vincent on
October 17, 2001. California Institute of Technology Library. http://authors
.library.caltech.edu/5456/1/hrst.mit.edu/hrs/materials/public/FriedelJacques
/Jacques_Friedel_interview.htm.
Friedel, Jacques. “Sur L’origine du Ferromagnétisme dans les Métaux se Transition.”
Journal de Physique et la Radium 16 (1955), 829–38.
Galison, Peter. How Experiments End. Chicago: University of Chicago Press, 1987.
Galison, Peter. Image and Logic: A Material Culture of Microphysics. Chicago: Uni-
versity of Chicago Press, 1997.
Galison, Peter, and David J. Stump, eds. The Disunity of Science: Boundaries, Con-
texts, and Power. Stanford, CA: Stanford University Press, 1996.
Gamow, George. “Any Physics Tomorrow?” Physics Today 2, no. 1 (1949), 17–21.
Gamow, George. “The Reality of Neutrinos.” Physics Today 1, no. 3 (1948), 4–9.
Gavroglu, Kostas, and Ana Simões. “The Americans, the Germans, and the Begin-
nings of Quantum Chemistry: The Confluence of Diverging Traditions.” His-
torical Studies in the Physical and Biological Sciences 25, no. 1 (1994), 47–110.
Gavroglu, Kostas, and Ana Simões. Neither Physics nor Chemistry: A History of
Quantum Chemistry. Cambridge, MA: MIT Press, 2012.
Geiger, Roger L. “Science, Universities, and National Defense, 1945–1970.” Osiris
7 (1992), 26–48.
Geiger, Roger L. To Advance Knowledge: The Growth of American Research Universi-
ties, 1900–1940. Oxford: Oxford University Press, 1986.
Gell-Mann, Murray. “Current Topics in Particle Physics.” In Proceedings of the
13th International Conference on High-Energy Physics, edited by Margaret H.
Alston-Garnjost, 3–9. Berkeley: University of California Press, 1967.
Gibbs, Roswell C. “What Should Be the Method of Accrediting Individuals and In-
stitutions?” Review of Scientific Instruments 15, no. 11 (1944), 305.
Godin, Benoît. “The Linear Model of Innovation: The Historical Construction of an
Analytical Framework.” Science, Technology, and Human Values 31, no. 6 (2006),
639–67.
Goldberg, Stanley, and Roger Stuewer. The Michelson Era in American Science
1870–1930: AIP Conference Proceedings. College Park, MD: American Institute
of Physics, 1988.
Golden, Frederic. “More Mini-Bangs for the Buck.” Time, July 18, 1983, 147.
250 BIBLIOGRAPHY

Goldsmith, Alfred N. “Comments on ‘Organization of Physics in America.’” Journal


of Applied Physics 15, no. 9 (1944), 649.
Good, Gregory A. “The Assembly of Geophysics: Scientific Disciplines as Frame-
works of Consensus.” Studies in History and Philosophy of Modern Physics 31,
no. 3 (2000), 259–92.
Gooday, Graeme. “‘Vague and Artificial’: The Historically Elusive Distinction be-
tween Pure and Applied Science.” Isis 103, no. 3 (2012), 546–54.
Gordon, James P., Herbert J. Zeiger, and Charles H. Townes. “The Maser—New
Type of Microwave Amplifier, Frequency Standard, and Spectrometer.” Physical
Review 99, no. 4 (1955), 1264–74.
Grantham, Todd A. “Conceptualizing the (Dis)Unity of Science.” Philosophy of Sci-
ence 71, no. 2 (2004), 133–55.
Groenewegen, Peter, and Lois Peters. “The Emergence and Change of Materials
Science and Engineering in the United States.” Science, Technology, and Human
Values 27, no. 1 (2002), 112–33.
Guyon, Étienne, Alexis Martinet, Jean Matricon, and Philip Pincus. “Tunneling into
Dirty Superconductors Near Their Upper Critical Fields.” Physical Review 138,
no. 3A (1965), A746–52.
Gray, Dwight E. “The New AIP Handbook.” Physics Today 16, no. 7 (1963), 40–42.
Gray, Dwight E., and Bruce H. Billings, eds. American Institute of Physics Handbook.
2nd ed. New York: McGraw-Hill, 1963.
“Hard Choices.” Economist, May 28, 1983, 100.
Hargittai, István. The Martians of Science: Five Physicists Who Changed the Twentieth
Century. Oxford: Oxford University Press, 2006.
Harnwell, Gaylord P. “A More Perfect Union.” Review of Scientific Instruments 15,
no. 2 (1944), 19–21.
Harnwell, Gaylord P. “Research in Physics in the Postwar Period.” Review of Scien-
tific Instruments 14, no. 8 (1943), 231–36.
Harnwell, Gaylord P. “The Role of Organization in Meeting the Needs of Physics.”
Review of Scientific Instruments 15, no. 11 (1944), 309–15.
Hartman, Paul. Cornell’s Materials Science Center: The Early Years. Ithaca, NY: Cor-
nell Center for Materials Research, 2007.
Hartman, Paul. A Memoir on the Physical Review. New York: American Institute of
Physics, 1994.
Hecht, David K. “The Atomic Hero: Robert Oppenheimer and the Making of Sci-
entific Icons in the Early Cold War.” Technology and Culture 49, no. 4 (2008),
943–66.
BIBLIOGRAPHY 251

Heilbron, John L., and Robert W. Seidel. Lawrence and His Laboratory: A History
of the Lawrence Berkeley Laboratory, vol. 1. Berkeley: University of California
Press, 1989.
Heinze, Thomas, Olof Hallonsten, and Steffi Heinecke. “From Periphery to Center:
Synchrotron Radiation at DESY.” Pts. 1 and 2, “Part I: 1962–1977,” “Part II:
1977–1993.” Historical Studies in the Natural Sciences 45, no. 3 (2015), 447–92;
no. 4 (2015), 513–48.
Heinze, Thomas, Olof Hallonsten, and Steffi Heinecke. “Turning the Ship: The
Transformation of DESY, 1993–2009.” Physics in Perspective 19, no. 4 (2017),
424–51.
Herman, Ros. “Americans Want Accelerator the Size of Belgium.” New Scientist, June
9, 1983, 679.
Herring, Conyers. “Recollections from the Early Days of Solid-State Physics.” Phys-
ics Today 46, no. 4 (1993), 26–33.
Hoch, Paul K. “The Development of the Band Theory of Solids, 1933–1960.” In
Hoddeson et al., Out of the Crystal Maze, 182–235.
Hoch, Paul K. “The Reception of Central European Refugee Physicists of the 1930s:
U.S.S.R., U.K., U.S.A.” Annals of Science 40, no. 3 (1983), 217–46.
Hoddeson, Lillian. “The Discovery of the Point-Contact Transistor.” Historical
Studies in the Physical Sciences 12, no. 1 (1981), 41–76.
Hoddeson, Lillian. “The Entry of the Quantum Theory of Solids into the Bell Tele-
phone Laboratories, 1925–40: A Case-Study of the Industrial Application of
Fundamental Science.” Minerva 18, no. 3 (1980), 422–47.
Hoddeson, Lillian Hartmann. “The Roots of Solid-State Research at Bell Labs.”
Physics Today 30, no. 3 (1977), 23–30.
Hoddeson, Lillian, and Gordon Baym. “The Development of the Quantum Mechan-
ical Electron Theory of Metals: 1900–28.” Proceedings of the Royal Society of Lon-
don A 371, no. 1744 (1980), 8–23.
Hoddeson, Lillian, Gordon Baym, and Michel Eckert. “The Development of the
Quantum-Mechanical Electron Theory of Metals: 1928–1933.” Reviews of Mod-
ern Physics 59, no. 1 (1987), 287–326.
Hoddeson, Lillian, Ernst Braun, Jürgen Teichmann, and Spencer Weart, eds. Out
of the Crystal Maze: Chapters from the History of Solid-State Physics. New York:
Oxford University Press, 1992.
Hoddeson, Lillian, and Vicki Daitch. True Genius: The Life and Science of John Bar-
deen: The Only Winner of Two Nobel Prizes in Physics. Washington, DC: Joseph
Henry Press, 2002.
252 BIBLIOGRAPHY

Hoddeson, Lillian, and Adrienne W. Kolb. “The Superconducting Super Collider’s


Frontier Outpost, 1983–1988.” Minerva 38 (2000), 271–310.
Hoddeson, Lillian, Adrienne W. Kolb, and Catherine Westfall. Fermilab: Physics, the
Frontier, and Megascience. Chicago: University of Chicago Press, 2008.
Hoddeson, Lillian, and Michael Riordan. Crystal Fire: The Birth of the Information
Age. New York: Norton, 1997.
Hoffmann, Dieter. “Fifty Years of Physica Status Solidi in Historical Perspective.”
Physica Status Solidi B 250, no. 4 (2013), 871–87.
Höhler, Gerhard, D. Lyons, and Ernst Niekisch. Arbeitstagung Festkörperphysik. Ber-
lin: Deutscher Verlag der Wissenschaften, 1952.
Holton, Gerald. “The Formation of the American Physics Community in the 1920s
and the Coming of Albert Einstein.” Minerva 19 (1981), 569–81.
Hopfield, John J. “Whatever Happened to Solid State Physics?” Annual Reviews of
Condensed Matter Physics 5 (2014), 1–13.
Hossenfelder, Sabine. “The LHC ‘Nightmare Scenario’ Has Come True.” Back Reac-
tion. August 6, 2016. https://backreaction.blogspot.co.uk/2016/08/the-lhc-night
mare-scenario-has-come-true.html.
Hounshell, David A. “Edison and the Pure Science Ideal in 19th-Century America.”
Science 207, no. 4431 (1980), 612–17.
Hughes, Sally Smith. “Making Dollars out of DNA: The First Major Patent in Bio-
technology and the Commercialization of Molecular Biology, 1974–1980.” Isis
92, no. 3 (2001), 541–75.
Hull, Albert W. “The Outlook for the Physicist and Prospective Physicist in Indus-
try.” American Journal of Physics 12 (1944), 62–70.
Hunt, Bruce J. The Maxwellians. Ithaca, NY: Cornell University Press, 1991.
Hunt, Bruce J. Pursuing Power and Light: Technology and Physics from James Watt to
Albert Einstein. Baltimore: Johns Hopkins University Press, 2010.
Huxley, Thomas H. “Science and Practical Life.” Popular Science Monthly 32 (No-
vember 1887), 166–69.
Hyde, W. Lewis. “Why 1916? A Look Back at OSA’s Roots.” Optics and Photonics
News 17, no. 1 (2006), 18–19.
“Important Announcement.” Physical Review 132, no. 1 (1963), 1.
Iselin, Columbus O’D. “Down to the Sea.” Physics Today 1, no. 4 (1948), 16–20.
Jackson, Catherine M. “Chemistry as the Defining Science: Discipline and Training
in Nineteenth-Century Chemical Laboratories.” Endeavour 35, no. 2–3 (2011),
55–62.
James, Jeremiah, and Christian Joas. “Subsequent and Subsidiary? Rethinking the
BIBLIOGRAPHY 253

Role of Applications in Establishing Quantum Mechanics.” Historical Studies in


the Natural Sciences 45, no. 5 (2015), 641–702.
Jammer, Max. The Conceptual Development of Quantum Mechanics. New York: Mc-
Graw-Hill, 1966.
Jenson, John C. “ . . . Pigtails on a Stratoliner.” Physics Today 1, no. 7 (1948), 10–16.
Joas, Christian. “Campos que interagem: Física quântica e a transferência de con-
ceitos entre física de partículas, nuclear e do estado sólido.” In Teoria quântica:
Estudos históricos e implicações culturais, edited by Olival Freire Jr., Osvaldo Pes-
soa Jr., and Joan Lisa Bromberg, 109–51. Campina Grande, Brasil: Livraria da
física, 2011.
Joas, Christian, and Michael Eckert. “Arnold Sommerfeld and Condensed Matter
Physics.” Annual Reviews of Condensed Matter Physics 8 (2017), 31–49.
Joas, Christian, and Shaul Katzir. “Analogy, Extension, and Novelty: Young
Schrödinger on Electric Phenomena in Solids.” Studies in History and Philos-
ophy of Science Part B: Studies in History and Philosophy of Modern Physics 42,
no. 1 (2011), 43–53.
Johnson, Ann. “What If We Wrote the History of Science from the Perspective of
Applied Science?” Historical Studies in the Natural Sciences 38, no. 4 (2008),
610–20.
Johnson, Eric E. “The Black Hole Case: The Injunction against the End of the
World.” Tennessee Law Review 76, no. 4 (2009), 819–908.
Jordi Taltavull, Marta. “Transmitting Knowledge across Divides: Optical Dispersion
from Classical to Quantum Physics.” Historical Studies in the Natural Sciences
45, no. 3 (2016), 313–59.
Joyce, E. Bernie, and Doug A. McCann, eds. Burke and Wills: The Scientific Legacy of
the Victorian Exploring Expedition. Collingwood, VIC: CSIRO, 2011.
Kadanoff, Leo P. “The Big, the Bad and the Beautiful.” Physics Today 41, no. 2
(1988), 9–11.
Kadanoff, Leo P. “Cathedrals and Other Edifices.” Physics Today 39, no. 11 (1986),
7–9.
Kadanoff, Leo P. “Complex Structures from Simple Systems.” Physics Today 44, no.
3 (1991), 9–11.
Kaiser, David. American Physics and the Cold War Bubble. Chicago: University of
Chicago Press, forthcoming.
Kaiser, David. “The Atomic Secret in Red Hands? American Suspicions of Theo-
retical Physicists during the Early Cold War.” Representations 90, no.1 (2005),
28–60.
254 BIBLIOGRAPHY

Kaiser, David. “Booms, Busts, and the World of Ideas: Enrollment Pressures and the
Challenge of Specialization.” Osiris 27 (2012), 276–302.
Kaiser, David. “Cold War Requisitions, Scientific Manpower, and the Production of
American Physicists after World War II.” Historical Studies in the Physical and
Biological Sciences 33, no. 1 (2002), 131–59.
Kaiser, David. Drawing Theories Apart: The Dispersion of Feynman Diagrams in
Postwar Physics. Chicago: University of Chicago Press, 2005.
Kaiser, David. “From Blackboard to Bombs.” Nature 523 (2015), 523–25.
Kaiser, David. “The Postwar Suburbanization of American Physics.” American
Quarterly 56, no. 4 (2004), 851–88.
Kaiser, David. “Training and the Generalist’s Vision in the History of Science.” Isis
96, no. 2 (2005), 244–51.
Katcher, David A. “Editorial.” Physics Today 1, no. 1 (1948), 3.
Keith, Stephen T., and Paul K. Hoch. “Formation of a Research School: Theoretical
Solid State Physics at Bristol 1930–54.” British Journal for the History of Science
19, no. 1 (1986), 19–44.
Keith, Stephen T., and Pierre Quédec. “Magnetism and Magnetic Materials.” In
Hoddeson et al., Out of the Crystal Maze, 359–442.
Kemble, Edwin C. “What Changes in Graduate Training in Physics Are Needed?”
Review of Scientific Instruments 15, no. 11 (1944), 290–93.
Kevles, Daniel J. “Big Science and Big Politics in the United States: Reflections on
the Death of the SSC and the Life of the Human Genome Project.” Historical
Studies of the Physical and Biological Sciences 27, no. 2 (1997), 269–97.
Kevles, Daniel J. “Cold War and Hot Physics: Science, Security, and the American
State, 1945–56.” Historical Studies in the Physical and Biological Sciences 20, no.
2 (1990), 239–64.
Kevles, Daniel J. The Physicists: The History of a Scientific Community in Modern
America. 3rd ed. 1971. Repr., Cambridge, MA: Harvard University Press, 1995.
Kittel, Charles. Introduction to Solid State Physics, 2nd ed. New York: Wiley, 1955.
Kleinman, Daniel Lee. Politics on the Endless Frontier: Postwar Research Policy in the
United States. Durham, NC: Duke University Press, 1995.
Klopsteg, Paul E. “The Work of the War Policy Committee of the American Institute
of Physics.” Review of Scientific Instruments 14, no. 8 (1943), 236–41.
Knowles, Scott G., and Stuart W. Leslie. “‘Industrial Versailles’: Eero Saarinen’s Cor-
porate Campuses for GM, IBM, and AT&T.” Isis 92, no. 1 (2001), 1–33.
Kohler, Robert E. Partners in Science: Foundations and Natural Scientists 1900–
1945. Chicago: University of Chicago Press, 1991.
BIBLIOGRAPHY 255

Kohlstedt, Sally Gregory, Michael Sokal, and Bruce Lewenstein. The Establishment
of Science in America. New Brunswick, NJ: Rutgers University Press, 1999.
Kohn, Walter. “An Essay on Condensed Matter Physics in the Twentieth Century.”
Reviews of Modern Physics 71, no. 2 (1999), S59–77.
Kohn, Walter, and Lu Jeu Sham. “Self-Consistent Equations Including Exchange and
Correlation Effects.” Physical Review 104, no. 4A (1965), A1133–38.
Kolb, Adrienne, and Lillian Hoddeson. “The Mirage of the ‘World Accelerator for
World Peace’ and the Origin of the SSC, 1953–1983.” Historical Studies in the
Physical and Biological Sciences 24, no. 1 (1993), 101–24.
Kowarski, Lew. “Les Piles de Recherches.” Commissariat à l’Energie Atomique, Re-
port #481, 1955. http://inis.iaea.org/search/search.aspx?orig_q=RN:38036105.
Kragh, Helge. Higher Speculations: Grand Theories and Failed Revolutions in Phys-
ics and Cosmology. Oxford: Oxford University Press, 2011.
Kragh, Helge. “Naming the Big Bang.” Historical Studies in the Natural Sciences 44,
no. 1 (2014), 3–36.
Kragh, Helge. Quantum Generations: A History of Physics in the Twentieth Century.
Princeton, NJ: Princeton University Press, 1999.
Krumhansl, James A. “Unity in the Science of Physics.” Physics Today 44, no. 3
(1991), 33–38.
Lalli, Roberto. “A New Scientific Journal Takes the Scene: The Birth of Reviews of
Modern Physics.” Annalen der Physik 526, no. 9–10 (2014), A83–87.
Lazarus, David. “Fausto Fumi and the Emergence of Solid-State Physics in Italy.” Il
Nuovo Cimento D 15, no. 2–3 (1993), 139–42.
Lederman, Leon M. “Fermilab and the Future of HEP.” In Proceedings of the 1982
DPF Summer Study on Elementary Particle Physics and Future Facilities, edited
by Rene Donaldson, Richard Gustafson, and Frank Paige, 125–27. Washington,
DC: US Department of Energy, 1983.
Lederman, Leon M. “The Value of Fundamental Science.” Scientific American 251,
no. 5 (1984), 40–47.
Lederman, Leon M., and Dick Teresi. The God Particle: If the Universe Is the Answer,
What Is the Question? New York: Houghton Mifflin, 1993.
Lenoir, Timothy, and Christophe Lécuyer. “Instrument Makers and Discipline
Builders: The Case of Nuclear Magnetic Resonance.” Perspectives on Science 3,
no. 3 (1995), 276–345.
Leslie, Stuart W. The Cold War and American Science: The Military-Industrial-
Academic Complex at MIT and Stanford. New York: Columbia University Press,
1993.
256 BIBLIOGRAPHY

Livingstone, M. Stanley. Particle Physics: The High-Energy Frontier. New York: Mc-
Graw-Hill, 1968.
Long, Tulley. “William McElroy, the McCollum-Pratt Institute, and the Transforma-
tion of Biology at Johns Hopkins, 1945–1960.” Journal of the History of Biology
42, no. 4 (2009), 765–809.
Lucier, Paul. “The Origins of Pure and Applied Science in Gilded Age America.” Isis
103, no. 3 (2012), 527–36.
Lyons, W. James. “Concerning the Division of High-Polymer Physics of the Ameri-
can Physical Society.” Journal of Applied Physics 15, no. 11 (1944), 729–30.
MacLeod, Christine, and Gregory Radick. “Claiming Ownership in the Technosci-
ences: Patents, Priority and Productivity.” Studies in the History and Philosophy
of Science Part A 44 (2013), 188–201.
Martin, Joseph D. “Fundamental Disputations: The Philosophical Debates That
Governed American Physics, 1939–1993.” Historical Studies in the Natural Sci-
ences 45, no. 5 (2015), 703–57.
Martin, Joseph D. “Prestige Asymmetry in American Physics: Aspirations, Appli-
cations, and the Purloined Letter Effect.” Science in Context 30, no. 4 (2017),
475–506.
Martin, Joseph D. “Resource Letter HCMP-1: History of Condensed Matter Phys-
ics.” American Journal of Physics 85, no. 2 (2017), 87–97.
Martin, Joseph D. “What’s in a Name Change? Solid State Physics, Condensed Mat-
ter Physics, and Materials Science.” Physics in Perspective 17, no. 1 (2015), 3–32.
Martin, Joseph D., and Michel Janssen. “Beyond the Crystal Maze: Twentieth-
Century Physics from the Vantage Point of Solid State Physics.” Historical Stud-
ies in the Natural Sciences 45, no. 5 (2015), 631–40.
McCray, W. Patrick. “Will Small Be Beautiful? Making Policies for Our Nanotech
Future.” History and Technology 21, no. 2 (2005), 177–203.
McFadden, Daniel G., Raza A. Tahir-Kheli, and G. Bruce Taggart. “Space-Time-
Dependent Correlation Functions for One-Dimensional Heisenberg Spin Sys-
tems in a Lorentzian-Gaussian Approximation.” Physical Review 185, no. 2
(1969), 854–59.
Midwinter, Charles, and Michel Janssen. “Kuhn Losses Regained: Van Vleck from
Spectra to Susceptibilities.” In Research and Pedagogy: A History of Early Quan-
tum Physics through Its Textbooks, edited by Massimiliano Badino and Jaume Na-
varro, 137–205. Berlin: Edition Open Access, 2013.
Missner, Marshall. “Why Einstein Became Famous in America.” Social Studies of Sci-
ence 15, no. 2 (1985), 267–91.
BIBLIOGRAPHY 257

Mody, Cyrus C. M. Instrumental Community: Probe Microscopy and the Path to Nan-
otechnology. Cambridge, MA: MIT Press, 2011.
Mody, Cyrus C. M. The Long Arm of Moore’s Law: Microelectronics and American
Science. Cambridge, MA: MIT Press, 2017.
Mody, Cyrus C. M. “Toward a History of Science from the Perspective of Applied
Science.” Inaugural Lecture, Maastricht University, October 6, 2017. https://cris.
maastrichtuniversity.nl/portal/files/17798700/Oratie_Mody.pdf.
Mody, Cyrus C. M., and Hyungsub Choi. “From Materials Science to Nanotech-
nology: Interdisciplinary Center Programs at Cornell University, 1960–2000.”
Historical Studies in the Natural Sciences 43, no. 2 (2013), 121–61.
Mott, Nevill F. The Beginnings of Solid State Physics: A Symposium Organized by Sir
Nevill Mott, Held 30 April–2 May 1979. London: Royal Society, 1980.
Mott, Nevill F. “The Mechanism of Work-Hardening of Metals.” Proceedings of the
Institution of Mechanical Engineers 166 (1952), 413–18.
Mulliken, Robert S. “Remarks on a Possible Division of Spectroscopy in the Amer-
ican Physical Society.” Review of Scientific Instruments 16, no. 2 (1945), 42–43.
Muskat, Morris. “Letter to the Editor.” Review of Scientific Instruments 16, no. 2
(1945), 38–39.
National Academy of Sciences. Advancing Materials Research. Washington, DC: Na-
tional Academies Press, 1987.
National Academy of Sciences. Industrial Research Laboratories of the United States.
19 editions. Washington DC: National Academy of Sciences, 1920–85.
National Academy of Sciences. More Effective Organization and Administration of
Materials Research and Development: A Report to Detlev W. Bronk, President, Na-
tional Academy of Sciences, by the Committee on Scope and Conduct of Materials
Research. Washington, DC: National Academy of Sciences, 1960.
National Academy of Sciences. Physics in Perspective, Volume II, Part A: The Core
Subfields of Physics. Washington, DC: National Academy of Sciences, 1972.
National Academy of Sciences. Physics through the 1990s: Condensed Matter Physics.
Washington, DC: National Academy of Sciences, 1986.
National Academy of Sciences. Physics: Survey and Outlook. Washington, DC: Na-
tional Academy of Sciences, 1966.
National Academy of Sciences. Report of the National Academy of Sciences, 1953–54.
Washington, DC: National Academy of Sciences, 1954.
National Academy of Sciences. Report of the National Academy of Sciences, 1957–58.
Washington, DC: National Academy of Sciences, 1958.
National Academy of Sciences. Research in Solid-State Sciences: Opportunities and
258 BIBLIOGRAPHY

Relevance to National Needs. Washington, DC: National Academy of Sciences,


1968.
National Science Foundation. “The First Annual Report of the National Science
Foundation, 1950–1951.” Washington, DC: Government Printing Office, 1951.
Nier, Alfred O. C., and John H. Van Vleck. “John Torrence Tate: 1889–1950.” In
Biographical Memoirs of the National Academy of Sciences, vol. 47, 461–84. Wash-
ington, DC: National Academies Press, 1975.
Noble, David F. American by Design. Oxford: Oxford University Press, 1977.
Nye, Mary Jo. Before Big Science: The Pursuit of Modern Chemistry and Physics,
1800–1940. Cambridge, MA: Harvard University Press, 1999.
Nye, Mary Jo. From Chemical Philosophy to Theoretical Chemistry: Dynamics of Mat-
ter and Dynamics of Disciplines, 1800–1950. Berkeley: University of California
Press, 1993.
Nye, Mary Jo., ed. The Modern Physical and Mathematical Sciences. The Cambridge
History of Science, vol. 5. Cambridge: Cambridge University Press, 1991.
Olesko, Kathryn M. Physics as a Calling: Discipline and Practice in the Königsberg
Seminar for Physics. Ithaca, NY: Cornell University Press, 1991.
Olesko, Kathryn M. “Physics Instruction in Prussian Secondary Schools before
1859.” Osiris 5 (1989), 94–120.
Olsen, Leonard O., Eugene C. Crittenden Jr., and Charles S. Smith Jr. “Letter to the
Editor.” Review of Scientific Instruments 15, no. 4 (1944), 108.
Oreskes, Naomi, and Erik M. Conway. Merchants of Doubt: How a Handful of Scien-
tists Obscured the Truth on Issues from Tobacco Smoke to Global Warming. New
York: Bloomsbury Press, 2010.
Osgood, Thomas H. “Letter to the Editor.” Review of Scientific Instruments 15, no.
4 (1944), 108.
Osgood, Thomas H. “Physics in 1943.” Journal of Applied Physics 15, no. 2 (1944),
89–107.
Page, Thornton. “The Origin of the Earth.” Physics Today 1, no. 6 (1948), 12–24.
Pais, Abraham. Inward Bound: Of Matter and Forces in the Physical World. Oxford:
Oxford University Press, 1988.
Pais, Abraham. “The ‘Physical Review’ Then and Now.” In The “Physical Review”:
The First Hundred Years, edited by H. Henry Stroke, 1–10. New York: Springer-
Verlag, 1995.
Physik der kondensierten Materie 1, no. 1 (1963).
Pick, Heinz. “Festkörperphysik.” Die Naturwissenschaften 41 (1954), 346–54.
Pickering, Andrew. Constructing Quarks: A Sociological History of Particle Physics.
Edinburgh: Edinburgh University Press, 1984.
BIBLIOGRAPHY 259

Pippard, Brian. “The Cat and the Cream.” Physics Today 14, no. 11 (1961), 38–41.
Pondrom, Lee. “Report of the Fixed Target Proton Accelerator Group.” In Proceed-
ings of the 1982 DPF Summer Study on Elementary Particle Physics and Future
Facilities, edited by Rene Donaldson, Richard Gustafson, and Frank Paige, 98–
104. Washington, DC: US Department of Energy, 1983.
Price, Derek J. de Solla. Little Science, Big Science. New York: Oxford University
Press, 1963.
Price, Don K. Government and Science. New York: New York University Press, 1954.
Proctor, Robert N. “‘-Logos,’ ‘-Ismos,’ and ‘-Ikos’: The Political Iconicity of Denom-
inative Suffixes in Science (or, Phonesthemic Tints and Taints in the Coining of
Science Domain Names).” Isis 98, no. 2 (2007), 290–309.
Radick, Gregory. “Presidential Address: Experimenting with the Scientific Past.”
British Journal for the History of Science 49, no. 2 (2016), 153–72.
Raymond, Richard C. “Letter to the Editor.” Physics Today 2, no. 3 (1949), 5, 38.
Reich, Leonard S. The Making of American Industrial Research: Science and Business
at GE and Bell, 1876–1926. New York: Cambridge University Press, 1985.
“Report of the National Research Council Conference of Physicists.” Review of Sci-
entific Instruments 15, no. 11 (1944), 283–328.
Richtmyer, Floyd K. “Editorial.” Review of Scientific Instruments 1, no. 1 (1930),
1–2.
Rider, Robin E. “Alarm and Opportunity: Emigration of Mathematicians and Physi-
cists to Britain and the United States, 1933–1945.” Historical Studies in the Phys-
ical Sciences 15, no. 1 (1984), 107–76.
Riordan, Michael. “The Demise of the Superconducting Super Collider.” Physics in
Perspective 2, no. 4 (2000), 411–25.
Riordan, Michael. “A Tale of Two Cultures: Building the Superconducting Super
Collider, 1988–1993.” Historical Studies in the Physical and Biological Sciences
32, no. 1 (2001), 125–44.
Riordan, Michael, Lillian Hoddeson, and Conyers Herring. “The Invention of the
Transistor.” Reviews of Modern Physics 71, no. 2 (1999), S336–45.
Riordan, Michael, Lillian Hoddeson, and Adrienne Kolb. Tunnel Visions: The Rise
and Fall of the Superconducting Super Collider. Chicago: University of Chicago
Press, 2015.
Ritson, David. “Demise of the Texas Supercollider.” Nature 366 (1993), 607–10.
Roberts, John D. Nuclear Magnetic Resonance: Applications to Organic Chemistry.
New York: McGraw-Hill, 1959.
Robinson, Howard A. “The Challenge of Industrial Physics.” Physics Today 1, no. 6
(1948), 4–7.
260 BIBLIOGRAPHY

Rosenberg, Robert. “American Physics and the Origins of Electrical Engineering.”


Physics Today 36, no. 10 (1983), 48–54.
Rosenfeld, Léon. “Men and Ideas in the History of Atomic Theory.” Archive for His-
tory of Exact Sciences 7, no. 2 (1971), 69–90.
Rowell, John M. “Condensed Matter Physics in a Market Economy.” Physics Today
45, no. 5 (1992), 40–47.
Rowland, Henry A. “The Highest Aim of the Physicist.” Science 10, no. 258 (1899),
825–33.
Rowland, Henry A. “A Plea for Pure Science.” Science 2, no. 29 (1883), 242–50.
Roy, Rustum. “Funding Big Science.” Physics Today 48, no. 9 (1985), 9–10.
Rush, John J. “US Neutron Facility Development in the Last Half-Century: A Cau-
tionary Tale.” Physics in Perspective 17, no. 2 (2015), 135–55.
Safire, William. “The Penumbra of Desuetude.” New York Times Magazine, October
4, 1987.
Scheiding, Tom. “Paying for Knowledge One Page at a Time: The Author Fee in
Physics in Twentieth-Century America.” Historical Studies in the Natural Sci-
ences 39, no. 2 (2009), 219–47.
Schmalian, Jörg. “Failed Theories of Superconductivity.” In BCS: 50 Years, edited by
Leon N. Cooper and Dmitri Feldman, 41–56. Singapore: World Scientific, 2011.
Schrage, Michael. “Glimpses of Truth, Not Gadgets, Are Spinoffs of Super-Collider.”
Los Angeles Times, August 6, 1992.
Schweber, Silvan S. “The Empiricist Temper Regnant: Theoretical Physics in the
United States, 1920–1950.” Historical Studies in the Physical and Biological Sci-
ences 17, no. 1 (1986), 55–98.
Schweber, Silvan S. “A Historical Perspective on the Rise of the Standard Model.” In
The Rise of the Standard Model: A History of Particle Physics from 1964 to 1979,
edited by Lillian Hoddeson, Laurie Brown, Michael Riordan, and Max Dresden,
645–84. Cambridge: Cambridge University Press, 1997.
Schweber, Silvan S. “Physics, Community, and the Crisis in Physical Theory.” Phys-
ics Today 46, no. 11 (1993), 34–40.
Schweber, Silvan S. “The Young John Clarke Slater and the Development of Quan-
tum Chemistry.” Historical Studies in the Physical and Biological Sciences 20, no.
2 (1990), 339–406.
Seidel, Robert W. “Accelerating Science: The Postwar Transformation of the Law-
rence Radiation Laboratory.” Historical Studies in the Physical Sciences 13, no.
3 (1983), 375–400.
Seidel, Robert W. “A Home for Big Science: The Atomic Energy Commission Lab-
BIBLIOGRAPHY 261

oratory System.” Historical Studies in the Physical and Biological Sciences 16


(1986), 135–75.
Seidel, Robert W. “The National Laboratories of the Atomic Energy Commission in
the Early Cold War.” Historical Studies in the Physical and Biological Sciences 32
(2001), 145–62.
Seidel, Robert W. “The Origins of the Lawrence Berkeley Laboratory.” In Big Sci-
ence: The Growth of Large Scale Research, edited by Peter Galison and Bruce
Hevly, 21–45. Stanford, CA: Stanford University Press, 1992.
Seitz, Frederick. Modern Theory of Solids. New York: McGraw-Hill, 1940.
Seitz, Frederick. On the Frontier, My Life in Science. New York: American Institute
of Physics, 1994.
Seitz, Frederick. “Solid.” Physics Today 2, no. 6 (1949), 18–22.
Seitz, Frederick. “Whither American Physics?” Review of Scientific Instruments 16,
no. 2 (1945), 39–42.
Shapin, Steven. The Scientific Revolution. Chicago: University of Chicago Press,
1996.
Sherwin, Chalmers W., and Raymond S. Isenson. First Interim Report on Project
Hindsight (Summary). Washington, DC: Office of the Director of Defense Re-
search and Engineering, 1966.
Shinbrot, Troy. “Editorial: Troy Shinbrot Introduces Physical Review Applied.” Phys-
ical Review Applied 1, no. 1 (2014), 010001–1.
Shurkin, Joel N. Broken Genius: The Rise and Fall of William Shockley, Creator of the
Electronic Age. New York: Palgrave Macmillan, 2006.
Siegel, Benjamin V., and Richard C. Sinnott. “The El Cerrito Cyclotron.” Physics
Today 1, no. 4 (1948), 1–15.
Slater, John C. Interview by Thomas S. Kuhn and John H. Van Vleck. October 3,
1963. Niels Bohr Library and Archives, College Park, MD. https://www.aip.org
/history-programs/niels-bohr-library/oral-histories/4892–1.
Slater, John C. “Quantum Physics in America between the Wars.” Physics Today 21,
no. 1 (1968), 43–53.
Slater, John C. “A Quantum Theory of Optical Phenomena.” Physical Review 25, no.
4 (1925), 395–428.
Slater, John C. “The Solid State.” Physics Today 5, no. 1 (1952), 10–15.
Slater, John C. Solid State and Molecular Theory: A Scientific Biography. New York:
Wiley, 1975.
Slayton, Rebecca. “Discursive Choices: Boycotting Star Wars between Science and
Politics.” Social Studies of Science 37, no. 1 (2007), 27–66.
262 BIBLIOGRAPHY

Sloan, Phillip R. “Molecularizing Chicago—1945–1965: The Rise, Fall, and Rebirth


of the University of Chicago Biophysics Program.” Historical Studies in the Nat-
ural Sciences 44, no. 4 (2014), 364–412.
Smith, Cyril Stanley. “Matter versus Materials: A Historical View.” Science 162, no.
3854 (1968), 637–44.
Smith, Cyril Stanley. “The Prehistory of Solid State Physics.” Physics Today 18, no.
12 (1965), 18–30.
Solomon, Arthur K. “Physics and Cancer.” Physics Today 1, no. 1 (1948), 18–20.
Staley, Richard. “Trajectories in the History and Historiography of Physics in the
Twentieth Century.” History of Science 51, no. 2 (2013), 151–78.
Stephan, Karl. “Experts at Play: Magnetron Research at Westinghouse, 1930–1934.”
Technology and Culture 42, no. 4 (2001), 737–49.
Stevens, Hallam. “Fundamental Physics and Its Justifications, 1945–1993.” Histori-
cal Studies in the Physical and Biological Sciences 34, no. 1 (2003), 151–97.
Stroke, H. Henry, ed. The Physical Review: The First Hundred Years. New York:
Springer-Verlag, 1995.
Stuewer, Roger H. The Compton Effect: Turning Point in Physics. New York: Science
History, 1975.
Stuewer, Roger H. “Nuclear Physicists in a New World: The Émigrés of the 1930s in
America.” Berichte zur Wissenschaftsgeschichte 7, no. 1 (1984), 23–40.
Sullivan, Walter. “Physics Prize Won for Research in Electronics.” New York Times,
October 24, 1973.
Sutton, Richard M. “Introductory Remarks.” Review of Scientific Instruments 15, no.
11 (1944), 284–85.
Sweetnam, George K. The Command of Light: Rowland’s School of Physics and the
Spectrum. Philadelphia: American Philosophical Society, 2000.
Tate, John T. “Let Us Go Forward. . . .” Journal of Applied Physics 8, no. 1 (1937), 1.
Tate, John T., and John H. Van Vleck. “The Editor’s Column.” Physics 1, no. 1
(1931), 69–74.
Teissier, Pierre. “Solid-State Chemistry in France: Structures and Dynamics of a Sci-
entific Community since World War II.” Historical Studies in the Natural Sci-
ences 40, no. 2 (2010), 225–58.
Thibodeau, Sharon Gibbs. “Science in the Federal Government.” Osiris 1 (1985),
81–96.
Tisza, Laszlo. “Helium, the Unruly Liquid.” Physics Today 1, no. 4 (1948), 4–9.
Tobey, Ronald C. American Ideology of National Science, 1919–1930. Pittsburgh:
University of Pittsburgh Press, 1971.
BIBLIOGRAPHY 263

Tocqueville, Alexis de. Democracy in America, vol. 2, translated by Henry Reeve.


Cambridge: Sever and Francis, 1863. https://www.gutenberg.org/files/816/816
-h/816-h.htm.
Traweek, Sharon. Beamtimes and Lifetimes: The World of High Energy Physicists.
Cambridge, MA: Harvard University Press, 1988.
Trout, J. D. “Reductionism and the Unity of Science.” In The Philosophy of Science,
edited by Richard Boyd, Philip Gasper, and J. D. Trout, 387–92. Cambridge,
MA: MIT Press, 1991.
Turner, Michael. “Aspen Physics Turns 50.” Nature 486, no. 7403 (2012), 315–17.
“2 from U.S. among 4 Nobel Science Winners.” New York Times, October 12, 1977.
“Two Get Medals for Merit.” New York Times, March 22, 1944.
Van Vlack, Lawrence H. Elements of Material Science and Engineering. Reading,
MA: Addison-Wesley, 1959.
Van Vleck, John H. “American Physics Comes of Age.” Physics Today 17, no. 6
(1964), 21–26.
Van Vleck, John H. “A Survey of the Theory of Ferromagnetism.” Reviews of Modern
Physics 17, no. 1 (1945), 27–48.
Van Vleck, John H. The Theory of Electric and Magnetic Susceptibilities. Oxford:
Clarendon Press, 1932.
Van Vleck, John H., and Alfred O. C. Nier. “John Torrence Tate: 1889–1950.” In
Biographical Memoirs of the National Academy of Sciences, vol. 47, 460–85. Wash-
ington, DC: National Academies Press, 1975.
Varma, Roli. “Changing Research Cultures in U.S. Industry.” Science, Technology,
and Human Values 25, no. 4 (2000), 395–416.
Wang, Jessica. American Science in an Age of Anxiety: Scientists, Anticommunism,
and the Cold War. Chapel Hill: University of North Carolina Press, 1999.
Wannier, Gregory H. “The Statistical Problem in Cooperative Phenomena.” Reviews
of Modern Physics 17, no. 1 (1945), 50–60.
Warwick, Andrew. Masters of Theory: Cambridge and the Rise of Mathematical Phys-
ics. Chicago: University of Chicago Press, 2003.
Waterfall, Wallace. Interview by R. Bruce Lindsay. April 17, 1964. Niels Bohr Library
and Archives, College Park, MD.
Waterfall, Wallace, and Elmer Hutchisson. “Organization of Physics in America.”
Journal of Applied Physics 15, no. 5 (1944), 407–9.
Waterfall, Wallace, and Elmer Hutchisson. “Reply to Comments.” Journal of Applied
Physics 15, no. 9 (1944), 649–50.
Weaire, Denis L., ed. Solid State Science: Past, Present and Predicted. Bristol: Adam
Hilger, 1987.
264 BIBLIOGRAPHY

Weart, Spencer R. “The Birth of the Solid-State Physics Community.” Physics Today
41, no. 7 (1988), 38–45.
Weart, Spencer R. “The Physics Business in America, 1919–1940: A Statistical Re-
connaissance.” In The Sciences in the American Context: New Perspectives, edited
by Nathan Reingold, 295–358. Princeton, NJ: Princeton University Press, 1979.
Weart, Spencer R. The Rise of Nuclear Fear. Cambridge, MA: Harvard University
Press, 2012.
Weart, Spencer R. Scientists in Power. Cambridge, MA: Harvard University Press,
1979.
Weart, Spencer R. “The Solid Community.” In Hoddeson et al., Out of the Crystal
Maze, 617–68.
Weinberg, Alvin M. “Criteria for Scientific Choice.” Minerva 1, no. 2 (1963), 159–71.
Weinberg, Alvin M. “Criteria for Scientific Choice.” Physics Today 17, no. 3 (1964),
42–48.
Weinberg, Alvin M. The First Nuclear Era: The Life and Times of a Technological
Fixer. New York: American Institute of Physics, 1994.
Weinberg, Gerald M. An Introduction to General Systems Thinking. New York: Wiley,
1975.
Weinberg, Steven. “The Answer to (Almost) Everything.” New York Times, March
8, 1993.
Weinberg, Steven. Dreams of a Final Theory. New York: Pantheon Books, 1992.
Weiner, Charles. “A New Site for the Seminar: The Refugees and American Physics
in the 1930s.” In The Intellectual Migration: Europe and America, 1930–1960,
edited by Donald H. Fleming and Bernard Bailyn, 190–234. Cambridge, MA:
Harvard University Press, 1969.
Weiner, Charles. “Physics Today and the Spirit of the Forties.” Physics Today 26, no.
5 (1973), 23–28.
Weisel, Gary J. “The Plasma Archipelago: Plasma Physics in the 1960s.” Physics in
Perspective 19, no. 3 (2017), 183–226.
Weisel, Gary J. “Properties and Phenomena: Basic Plasma Physics and Fusion Re-
search in Postwar America.” Physics in Perspective 10, no. 4 (2008), 396–437.
Weisskopf, Victor. “Nuclear Structure and Modern Research.” Physics Today 20, no.
5 (1967), 23–26.
Weisskopf, Victor, and Alvin Weinberg. “Two Open Letters.” Physics Today 17, no.
6 (1964), 46–47.
Westfall, Catherine. “A Different Laboratory Tale: Fifty Years of Mössbauer Spec-
troscopy.” Physics in Perspective 8 (2006), 189–213.
BIBLIOGRAPHY 265

Westfall, Catherine. “Institutional Persistence and the Material Transformation of the


US National Labs: The Curious Story of the Advent of the Advanced Photon
Source.” Science and Public Policy 39, no. 4 (2012), 439–49.
Westfall, Catherine. “Rethinking Big Science: Modest, Mezzo, Grand Science and
the Development of the Bevalac, 1971–1993.” Isis 94, no. 1 (2003), 30–56.
Westfall, Catherine. “Retooling for the Future: Launching the Advanced Light
Source at Lawrence’s Laboratory, 1980–1986.” Historical Studies in the Natural
Sciences 38, no. 4 (2008), 569–609.
Westwick, Peter J. The National Labs: Science in an American System, 1947–1974.
Cambridge, MA: Harvard University Press, 2003.
Whewell, William. The Philosophy of the Inductive Sciences, Founded upon Their His-
tory, vol. 2. London: John W. Parker, 1840.
Whitten, Barbara. “What Physics Is Fundamental Physics? Feminist Implications of
Physicists’ Debates over the Superconducting Supercollider.” NWSA Journal 8,
no. 2 (1996), 1–16.
Wigner, Eugene P. “On the Mass Defect of Helium.” Physical Review 43, no. 4
(1933), 252–57.
Wigner, Eugene P. The Recollections of Eugene P. Wigner as Told to Andrew Szanton.
New York: Plenum Press, 1992.
Wigner, Eugene P., and Frederick Seitz. “On the Constitution of Metallic Sodium.”
Physical Review 43, no. 10 (1933), 804–10.
Wildhack, William A. “Letter to the Editor.” Review of Scientific Instruments 15, no.
10 (1944), 271–72.
Wilson, Benjamin. “The Consultants: Nonlinear Optics and the Social World of
Cold War Science.” Historical Studies in the Natural Sciences 45, no. 5 (2015),
758–804.
Wise, George. “Ionists in Industry: Physical Chemistry at General Electric, 1900–
1915.” Isis 74, no. 1 (1983), 6–21.
Wolfe, Audra. Competing with the Soviets: Science, Technology, and the State in Cold
War. Baltimore: Johns Hopkins University Press, 2013.
Xu, Qiaozhen, and Laurie M. Brown. “The Early History of Cosmic Ray Research.”
American Journal of Physics 55, no. 1 (1987), 23–33.
Yang, Chen-Ning, Edwin L. Goldwasser, Val Fitch, Alvin M. Weinberg, Owen Cham-
berlain, George E. Pake, Maurice Goldhaber, and Milton White. “High-Energy
Physics: Round-Table Discussion.” Physics Today 17, no. 11 (1964), 50–57.
“Yanks Sweep Science Field for This Year’s Nobel Prizes.” Chicago Tribune, Octo-
ber 19, 1976.
266 BIBLIOGRAPHY

Zangwill, Andrew. “The Education of Walter Kohn and the Creation of Density Func-
tional Theory.” Archive for History of Exact Sciences 68, no. 6 (2014), 775–848.
Zangwill, Andrew. “A Half Century of Density Functional Theory.” Physics Today 68,
no. 7 (2015), 34–39.
Zangwill, Andrew. “Hartree and Thomas: The Forefathers of Density Functional
Theory.” Archive for History of Exact Sciences 67, no. 3 (2013), 331–48.
Zener, Clarence. “The Fracture Stress of Steel.” Reviews of Modern Physics 17, no. 1
(1945), 20–26.
INDEX

ab initio methods, 89, 126, 224n41 217n23; publishing operations of, 78,
Abraham, Max, 105, 227n7 80, 84, 91–93, 96–98, 100, 110, 185,
accreditation of physicists, 45–47 201, 236n45; survey conducted by,
Acoustical Society of America, 23–25, 40, 85; War Policy Committee of, 213n5
50, 201 American Philosophical Society, 45, 50,
acoustics, 7, 36, 39, 106, 152 220n14
Advanced Research Projects Agency, American Physical Society, council of,
121, 124–31, 133, 153, 162, 230n12, 41–44, 52–53, 56, 61, 65, 68–69, 76–
231n29 78, 91–93, 99, 119, 132, 161, 208;
advisory system, federal, 86, 90–91, 121– disposition toward applied research,
26, 133, 153, 159 39–45, 49–54, 65, 74–75, 200–202;
Air Force Office of Scientific Research. divisions of, 50, 56–62, 65–69, 73,
See United States Air Force 76–78, 81, 87, 112, 131–33, 144, 153,
American Association of Physics Teach- 159, 164, 170, 204, 219n5, 220n19,
ers, 40, 217n12 221n49; dominant place in American
American Chemical Society, 50, 59, 78, physics, 33–37, 217n23; as a foil for
225n56, 231n39 American pragmatism, 21–24, 31;
American culture: and big science, 188, founding of, 5, 11–14, 19–20, 196;
196–97, 200, 210; place of physics intersociety relationships of, 231n39;
within, 8–11, 174; pure science ideal journals of, 26–27, 30, 97–98, 100,
as a foil to, 18–20, relevance of phys- 157, 225n56, 235n25, 240nn1–2;
ics to, 34 meetings of, 20, 38, 41–42, 52, 56,
American Institute of Electrical Engineers, 58–59, 61–62, 64–69, 76–78, 92–93,
21–22, 107, 217n17 99, 103, 138, 161, 170, 179, 217n12,
American Institute of Mechanical Engi- 223n3; membership of, 29, 84, 86,
neers, 63–64, 231n39 215nn12–14, 218n29, 222nn1–2,
American Institute of Physics, as an 230n1; presidents of, 11, 32, 42, 44,
alternative to the American Physi- 50, 68, 178–79; and the real estate
cal Society, 34–37, 39–43; and the business, 60, 220n15
Conference of Physicists, 45, 50–53; American Physics Teacher, 25
founding of, 24; member societies of, American Society for Metals, 63–64
268 INDEX

amorphous solids, 15, 62, 64, 156, 195 Aspen Center for Physics, 163
Anderson, Philip W.: opposition to SSC, astrophysics, 191, 223n3, 227n76
3–4, 10, 182–84, 186–92, 194–95; Atomic Energy Commission, 10, 102–4,
opposition to reductionism, 15, 137; 112–13, 115, 172, 227n6, 231n24
140–52, 206, 233nn37–38; early atomic physics, 27, 73, 87, 135, 185,
adopter of condensed matter physics, 227n76
154, 235n44; and intellectual merit of
solid state physics, 159, 160–61, 166; Bardeen, John, 147, 158, 224n33,
role in the Aspen Center for Physics 225n53, 235n29
163; comparison to Douglas Adams Barnes, R. Bowling, 49
233n31 Barry, Dave, 182
applications. See applied physics Barton, Henry, 45, 59, 80, 92
applied physics: and disciplinary federal- basic research: accessibility to solid state
ism, 39–44; early twentieth century researchers, 148, 159–60, 162–64;
growth of, 24–28, 30, 33, 35; as ex- federal support for, 16, 120–25,
tensive research 140; and federal 128–33, 136, 142, 144, 172–73,
funding priorities, 175–76, 183–89, 175–78, 185, 187–89, 191–92,
194–95; as the future of solid state 194–95, 233n24; industrial support
research, 148; marginalization of, 5, for, 191–92; institutional represen-
13–16, 30, 36, 41, 52, 60, 84, 120, tation for, 40–42, 45, 63, 67, 71,
200–202, 205, 207–8; in materials 74–75; at large facilities, 14, 206;
science, 122–23, 125, 127, 129–33; and metallurgy, 106; at the National
in the media, 150–52, 154; at the Na- Magnet Laboratory, 108–11, 113–15;
tional Magnet Laboratory, 104–107, and pedagogy, 49, 128; popularity of,
109, 111–15, 117; and physics educa- 84; as a problematic category, 7, 21,
tion, 48–49; as a problematic category, 217n9, 228n11, 240n67; profitability
7, 20–22, 41, 217n9, 228n11; repre- of, 119; relationship to technical out-
sentation in the APS, 65–67, 69–75; comes, 167–68, 194–95, 203, 216n3,
routes from basic research to, 160, 228n15, 229n27, 231n29, 232n15,
162–68, 229n27, 231n29, 232n14; 237n74; and scientific merit, 138
societies and journals dedicated to, Bell Telephone Laboratories: architecture
35, 39, 240n2; in song, 23 of, 231n29; breakup of, 163, 193;
applied research. See applied physics employees of, 119, 174, 215n14,
approximation techniques, 72, 89, 90, 218n33, 224n33; establishment of,
156–57, 225n45 22; lack of gemütlichkeit, 142; meeting
Argonne National Laboratory, 197 held at, 23; Philip Anderson at, 3, 137,
ARPA. See Advanced Research Projects 142–43; represented at the Confer-
Agency ence of Physicists, 32, 46–47, 49–50;
INDEX 269

solid state research and, 63, 71, 73, Buckley, Oliver E., 32, 34, 47–48, 51
91, 95, 162, 220n28, 232n15; William Burstein, Elias, 77
Shockley at, 57, 84–85
big science: excesses of, 205–7; and fed- California, Santa Barbara, University of,
eral spending, 172, 186–89, 191–92; 163
as a historiographical category 227n5; Cambridge, University of, 18, 147–48,
limits of, 16; for solid state physics, 235n44
102–4, 117 Carnegie Institute of Technology, 57, 63,
biology: antireduction in, 233n31; ascent 69, 84, 94
to prominence, 9, 149, 172, 197; con- Cavendish Laboratory, 87, 105, 154,
trasted with physics, 6; molecular, 227n8, 235n44
210; relevance of physics for, 49, ceramics. See amorphous solids
165, 179 chemical engineering, 71, 88, 89, 126,
Bitter, Francis, 102, 105–11, 114–15, 117, 130
187, 220n26, 227–28nn8–13 chemical physics, 52, 93–95, 165, 191,
BKS theory. See Bohr–Kramers–Slater 225n56, 226n58
theory chemistry, and emergence: 144–45,
Bloch, Felix, 89, 147, 196 179; Eugene Wigner and, 88–89; in
Bloembergen, Nicolaas, 112, 178–79 France, 156–57; as a fundamental
Boehlert, Sherwood, 178–80, 185 science, 105–7; industrial relevance
Bohr–Kramers–Slater theory, 88–89, of, 191; in industry, 36; at the Institute
224n41 for the Study of Metals, 164–65; and
Bohr, Niels. See Bohr-Kramers-Slater materials science, 122–24, 126–27,
theory 130; Nobel Prize for, 150; physicists’
Bond, James, 210–11 collaboration with, 63, 73, 79, 81, 86,
Breck, Otto, 71 93–97, 100, 109, 120, 201, 208; in
Breit, Gregory, 209 physics pedagogy, 90; professional
Brillouin, Lèon, 62, 89, 220n24, 225n45 identity, 21, 32, 33, 47–49, 51–52,
Brinkman Report, The, 163–67 60–61, 213n3, 214n10, 225n53,
Bristol, University of, 157 225nn55–56, 226n58, 226n64,
Bromley Report, The, 159–63 230n20, 231n29; publications, 14,
Brookhaven National Laboratory, 91, 94, 25, 27, 92–95
103–4, 197, 235n28 Chicago, University of: faculty of, 22,
Brooks, Harvey, 94, 96–98, 100, 158–59, 159, 123, 164, 186, 217n9, 218n33;
162, 225n53 as IDL site, 125; influence on Alvin
Brown University, 27, 109, 125, 164 Weinberg, 137–38; publications from,
Browne, Malcolm, 150, 179 142
bubble chambers, 135, 207 Clogston, Albert M., 119–20
270 INDEX

Cohen, Morill, 159 crystal structure physics: as a component


Cold War: big science and, 103; demands of solid state physics, 5, 69, 156;
of the, 14–16, 52, 120–22, 172–74, quantum mechanical approaches to,
209, 216n3; end of, 172, 174, 196, 55, 62–63, 72, 89; metallurgy and,
237n9; Manichaeism of, 211; prestige 106, 164; technical relevance of,
of physics in, 4–5, 7–11, 56, 197, 205, 130–31, 156
210
Columbia University, 95, 105, 142, DARPA. See Advanced Research Projects
217n15, 218n32 Agency
Columbus, Christopher, 171 Darrow, Karl Kelcher: on absenteeism at
Committee on the Survey of Materials Sci- APS meetings, 58; and the fable of the
ence and Engineering, 121, 129–33 bell and the cat, 45; friendship with
complexity, 15, 163, 168, 187, 208 John Van Vleck, 60, 220n14; over-
Compton effect. See Compton, Arthur sight of APS divisions, 42–44, 50, 52,
Holly 57, 60–62, 64–66, 68, 75–77, 219n5,
Compton, Arthur Holly, 33, 88, 224n41 221n31; response to the publication
Compton, Karl, 87, 218n49 problem, 92, 98; visiting professor-
computers, 89, 126, 150, 175, 177, 193, ships, 218n33
208 Davisson, Clinton, 22
condensed matter physics. See solid state Defense Advanced Research Projects
physics Agency. See Advanced Research Proj-
Conference of Physicists, National Re- ects Agency
search Council. See National Research density functional theory, 163, 226n64,
Council 236n56
Congress. See United States Congress Department of Defense, 102, 104, 111,
consumer culture, 11, 122 115, 120–21, 123, 127
Cooper, Leon, 147, 158, 235n29 Department of Energy, 10, 172, 175–77,
Cornell University, 47, 72, 123, 125, 142, 179, 192
179 Derge, Gerhard, 63
COSMAT. See Committee on the Survey Disciplines: and accreditation, 45; bound-
of Materials Science and Engineering aries of, 5, 13, 33–35, 47–49, 53,
cosmology, 136, 177, 181, 189, 195, 208, 106; competition among, 14, 148,
240n67 185–87, 191–92; experimental, 69,
Cosmos Club, 24, 217n23 128–31; formation of, 7, 31, 55,
counterfactuals, 208–10 78–80, 125, 153–55, 166–68, 213n3;
Cranberg, Lawrence, 186 interaction among, 17, 91, 95–96,
Crease, Robert P., 103, 197 107, 126, 138, 164, 187, 198;
INDEX 271

and professional identity, 12, 21, source of international competitive-


99–100, 118, 120–21, 200–205, 207, ness, 174–75, 193–94
216n3; representation of, 30; unity Edison, Thomas, 19–20
of, 86 Einstein, Albert, 9, 147, 209
Division of Condensed Matter Physics, electrical engineering: at MIT 108–9,
American Physical Society, 132, 161 126–27; origins in physics, 39, 208;
Division of Electron and Ion Optics, overlap with physics, 36, 73, 130,
American Physical Society, 43, 52, 60, 240n3; place in physics pedagogy,
67–68, 219n4, 223n3 48
Division of Fluid Dynamics, American electromagnetism, 6–7, 19, 21, 29, 100,
Physical Society, 161–62 182
Division of High Polymer Physics, Amer- electronics. See electrical engineering
ican Physical Society, 52, 56, 67–69, emergence: deployed in opposition to the
218n34, 221n49 SSC, 181, 194; Philip Anderson’s
Division of Materials Physics, American support for, 15, 137, 145, 149; roots
Physical Society, 132 in condensed matter phenomena, 206
Division of Particles and Fields, American émigré physicists, 30, 86–87
Physical Society, 144, 170 engineering, accelerator: 177; at Bell Lab-
Division of Solid State Physics, American oratories, 24; chemical, 88; Manhat-
Physical Society: founding of, 13, 39, tan Project and, 8, 208, 214n10; ma-
53–54, 56, 69, 87; growth of, 75–78; terials science and, 120–24, 126–27,
integration of academic and industrial 129–20, 132–33, 231n29; physics as
researchers within, 65, 70–72, 75; glorified 48–49; professional identity
membership of, 133, 222n2, 229n31; of, 21–22; publications, 27; radar and,
name change, 161–62; participation in 216n1; relationship to physics, 14, 33,
Federation of Materials Societies, 132; 39, 41–42, 63, 73–74, 96, 106–107,
John Van Vleck as chair of, 223n4 179, 201–3, 228n11; training, 105.
Dresselhaus, Mildred, 72, 126, 222n63 See also electrical engineering
Dushman, Saul, 57, 219n9 England. See Great Britain
Dutta, Pulak, 185
federal advisory system. See advisory sys-
Eastman-Kodak, 23 tem, federal
economic relevance of physics: disdain federal funding. See funding, federal
for, 10, 13; high energy physicists’ Fermilab. See National Accelerator
ambivalence about, 117, 181; re- Laboratory
lationship to pure science ideal, 16, ferromagnetism, 68, 71, 105, 156, 208,
19–20, 196; role in sustaining pres- 227n8
tige of physics, 11, 14, 179, 200; festkörperphysik, 156, 234n16
272 INDEX

First World War, 21 Gamow, George, 83


Fletcher, Harvey, 50, 69 Gell-Mann, Murray, 12, 213n6, 232n12,
Fleury, Paul A., 174, 181, 188, 237n11 241n7
fluids, 15, 162–63, 186 General Electric: founding of research
Forman, Paul, 203, 216n3, 228n15 laboratory, 22; Frederick Seitz at,
France, 61–62, 156–57, 234n11, 235n25 90, 97; publications of, 142–43; Ro-
Francis Bitter National Magnet Labora- man Smoluchowski at, 13, 39, 56–57,
tory. See National Magnet Laboratory 69
Franck, James, 165 Germany, 18–20, 61, 155–57, 209, 213n3,
Friedel, Jacque, 156–57 235n25
fundamental physics: funding for, 171, Gibbs, Roswell C., 47–48, 51
173–74, 177–78; institutional organi- glass. See amorphous solids
zation to protect, 39, 41–42; materials Goldsmith, Alfred N., 38–39
research as, 128–29; nature of, 16, Goudsmit, Samuel, 83, 92, 99, 235n28
136–40, 142–50, 174, 210, 228n15; Gray, Dwight, 152
reductionism and, 15, 181, 183–85, Great Britain, 18–19, 21, 87, 157,
187–89, 211; relationship to appli- 235n23, 235n25
cations, 111, 114–18, 193–95; as a group of six, 13, 57–62, 65–69, 76, 86–
public good, 176–77, 227–28n11; 87, 93, 219n5
solid state physics as, 14, 95, 102, Guggenheim Foundation, 105
104–8, 153–54, 158, 164, 166–67,
199, 232n14 Hansen, William Webster, 33
funding, federal: for basic research, 187; Harnwell, Gaylord Probasco: role in
of big science, 16, 104, 173, 176, 182, establishing Physics Today, 80, 83;
191–92, 237n9; coupled to practical support for disciplinary federalism
goals, 111–13, 117, 120–21, 160, 167, 36–40, 42, 49–53, 74–75, 217n12; as
174–75; of high energy physics, 10– University of Pennsylvania president,
11, 82, 144, 172; incentives created 63–64
by, 153; influence on postwar physics, Harvard University: Edwin Kemble at,
35, 52, 149; of materials science 14, 48; George Pake at, 226n61; Harvey
129, 131, 133; through the National Brooks at, 94, 96; interdisciplinary
Science Foundation, 102; priorities, laboratory hosted at, 125; John Slater
11, 116, 137, 189, 204; for social pro- at, 126; John Van Vleck at, 58, 87,
grams, 103; tightening of, 14, 103–4, 223n4; Nicolaas Bloembergen at, 112,
136, 138 178; Philip Anderson at, 3, 142; Ra-
Fuqua, Don, 175–76 dio Research Laboratory of, 85
Herring, Conyers, 95, 224n33, 225n53
INDEX 273

Higgs boson, 171 professional representation for, 5,


high energy physics: APS division for, 13–14, 16–17, 24, 31, 36–37, 42–52,
132; as big science, 14–16, 102, 104, 54, 195, 199–201, 208; proposed
114–18, 197, 205–7; competition APS division for, 62; relationship with
between solid state physics and, 79, academia, 23, 34–37, 41–44, 47, 72,
114–18, 135–40, 143–150; derision 75, 84–86, 157, 168, 201, 217n41,
of solid state physics within, 12, 218n33; role sustaining prestige of
241n7; funding for, 11, 227n6; histo- physics, 12; roots of solid state phys-
riographical focus on, 208–10; poten- ics in, 15, 56, 58, 62–67, 69, 72, 74,
tial source of world-eating black holes, 96–97, 153, 162, 168, 198; technolo-
214n25; prestige of, 4, 159, 200; gies developed in, 11, 54, 73, 142–43
productivity claims for 3, 192–96, Institute for the Study of Metals, Univer-
202; publications, 161; and the SSC, sity of Chicago, 123, 164–65
170–90, 192–95, 197; visibility of, 8, Institute for Theoretical Physics, Copen-
10–11, 82, 112. See also reductionism hagen, 88, 224n41
Hoddeson, Lillian, 10, 172 Institute of Radio Engineers, 38–39,
House of Representatives, 3, 175, 179, 217n17
184–85, 189, 196 intellectual property, 172, 192–93
Hull, Albert W., 32, 44 interdisciplinarity: in chemical physics,
Hutchins, Robert Maynard, 137 93, 225n56; in materials science,
Hutchisson, Elmer, 37–39, 217n15 120–22, 124–26, 128, 131, 133, 204;
Huxley, Thomas Henry, 21 in metallurgy, 164; perspectives of
among solid state physicists, 95–97;
IBM, 132, 14–43, 147, 193, 231n29 as a value, 56, 240nn2–3
Illinois, University of, 90–91, 126, 142 Interdisciplinary Laboratories, 121, 125–
industry: basic research in, 119, 143; 26, 128–31, 162, 231n29
British, 19, 214–15n4; condensed integrated circuits, 11, 163
matter physics distances itself from, International Journal of the Physics and
165, 167–68; German, 19, 214–15n4; Chemistry of Solids, 97–98
growth of physics in, 19–22, 30, 35, International Union of Pure and Applied
69, 92, 152, 173, 215n12, 224n29; Physics, 78
laboratories, 22, 100, 163, 193–94,
197, 205, 231n29, 232n15; marginal- James Franck Institute, 164–65
ization of, 19, 21, 24, 28–29; pharma- Johns Hopkins University, 20, 225n56
ceutical, 150; and physics pedagogy, Journal of Applied Physics: editorial
48–49, 148, 159, 191; and physics direction of, 201, 221n36; editorials
publishing, 26–27, 142, 221n36; and letters in, 37–39, 43, 65–66;
274 INDEX

Journal of Applied Physics (cont.): editors levels of physical organization, 15, 139,
of, 37, 217n15; publications of the 144–49, 181, 186–87, 208, 233n31
National Magnet Laboratory in 113, Lincoln Laboratory, 109, 126, 143
228n22; renamed from Physics, 27 linear model of innovation, 167, 231n29,
Journal of Chemical Physics, 25, 92, 237n74
201–2, 225n25 liquid helium, 15, 81, 162, 179
Journal of Rheology, 25, 30, 201 liquids, 152, 154–56, 160, 196
Journal of the Acoustical Society of Amer- Livingstone, M. Stanley, 136
ica, 25, 201 Los Alamos, 112, 209
Journal of the Optical Society of America, low temperatures, 6, 108, 147, 164
25, 201
journals. See publishing magnetic resonance imaging, 167, 175,
178–79
Kadanoff, Leo, 164–65, 186 magnetism, 28, 60, 102, 105, 113, 164,
Katcher, David A., 80–81 166
Kelly, Mervin, 49–50 magnets, 11, 127, 167, 177, 178–79
Kemble, Edwin, 48–49, 87, 126, 230n16 Manhattan Project: Alvin Weinberg’s
Kevles, Daniel J., 32–34, 196–97, 216n3 employment with, 138; associations
Kincaid, John F., 126–28 between basic research and, 120,
Kittel, Charles, 72–73, 143, 225n53 214n10; Bern Porter’s employment
Kleppa, Ole, 165 with, 27; as engineering endeavor,
Kleppner, Daniel, 205 8–10, 208–9; institutional legacy of,
Kohn, Walter, 94, 96, 154, 162–63, 164; secrecy around, 32, 35
226n64 Mansfield Amendments, 115, 120
Kolm, Henry, 113–14, 229n36 many-body physics, 160–61
Kramers, Henrik, 88 Massachusetts Institute of Technology:
Krumhansl, James, 123, 179, 191 faculty of, 72–73, 83, 86–87, 89, 105,
139, 228n13, 240n3; laboratories of,
Laboratory for Insulation Research, 127, 14, 23, 85, 107–118, 125–26, 143;
230n20 solid state theory course, 72; students
Large Hadron Collider, 171, 214n25 of, 86, 222n63
lasers, 11, 177–78, 196, 236n45 Materials Advisory Board, 122–24. See
Lawrence Berkeley Laboratory, 81, 143, also National Research Council
207, 214n11 Materials and Man’s Needs. See Commit-
Lax, Benjamin, 104, 109–14, 117, 160, tee on the Survey of Materials Science
173, 229n36 and Engineering
Lederman, Leon, 170, 174, 176–78, 182, Materials Research Centers. See Interdis-
184, 202 ciplinary Laboratories
INDEX 275

Materials Research Society, 167–79 National Bureau of Standards, 52, 61,


materials research. See materials science 215n14
materials science: APS division for, 132; national defense, 10, 22, 120, 123, 125,
ARPA support for, 124–29; as a disci- 174–75
plinary category, 56, 120–24, 129–24, National Institutes of Health, 111, 188
153–54, 166, 197, 203; funding for, National Magnet Laboratory: as big sci-
14; practical objectives of, 132, 160, ence, 14, 102–104, 112, 115, 197,
168; not science, 129; training in, 124 200; founding of, 107–10; funding
mechanics, 7, 28, 44, 166 for, 104, 108, 110–15, 117–18, 173;
megascience, 188, 194, 200, 206, 209 research conducted at, 108–111,
Metallurgical Laboratory, University of 113–14
Chicago, 164 National Research Council: Conference of
metallurgy: collaboration with, 58, 63–64, Physicists, 32, 45–51, 80; Solid State
90, 164–65; Francis Bitter’s vision for, Sciences Panel, 110; survey of indus-
102–9, 220n26, 228n13; materials trial laboratories, 63; surveys of solid
science and, 120, 122–23, 126, 130; state and condensed matter physics,
relationship to physics, 51, 60–61, 73, 157, 159, 160–62, 165, 167, 193. See
96, 208, 214n10, 230n20; teaching also Materials Advisory Board; Na-
of, 48 tional Academy of Sciences
Michigan State University, 35, 217n10 National Science Foundation: establish-
Mody, Cyrus C. M., 94, 126 ment, 102; support of APS journals,
“More Is Different,” 15, 143–150, 152, 91–92, 98–99; support of condensed
160, 187, 188, 194 matter physics, 192; support of high
Mott, Nevill, 90, 143, 150, 157 energy physics, 104, 188, 227n6; sup-
muons, 7, 155 port of materials science, 133; sup-
Muskat, Morris, 36, 217n9 port of National Magnet Laboratory,
110–12, 114
National Academy of Sciences: biograph- national security. See national defense
ical memoirs of, 26; membership in, National Synchrotron Light Source, 103,
110, 143, 232–33n22; policy advice 197
of, 121; reports, 129–30, 153, 157. new big science, 103, 196–98
See also National Research Council New York Times, 150, 179, 184, 216n1
National Accelerator Laboratory: contri- Nobel Prize: Burton Richter, 154; Clin-
butions of, 178–79; funding for, 104, ton Davisson, 22; Isidor Isaac Rabi,
112, 185; histories of, 207; justifi- 23; James Franck, 165; John Bar-
cations for 115, 117; members of in deen, 224n33; John Van Vleck, 150;
orbit, 176; opposition to, 112, 173; Leon Lederman, 170; Nevill Mott,
Tevatron, 171 150; Nicolaas Bloembergen, 178;
276 INDEX

Nobel Prize (cont.): Philip Anderson, 143, pedagogy, 32, 35, 72, 109, 126, 199, 207
150, 183; Steven Weinberg, 182; Pegram, George, 44, 91, 221n43
Walter Kohn, 226n64 Pennsylvania, University of, 36, 63–64,
Noble, David F., 21, 215n5 125
noncrystaline solids. See amorphous Pergamon Press, 97
solids Philosophical Magazine, 26
Northwestern University, 56, 125, 185 physical chemistry. See chemical physics,
nuclear arms race, 9, 232n3 93, 164, 225n56
nuclear magnetic resonance, 11, 94, 145, Physical Review: ability to read cover-to-
178, 226n61, 226n64 cover, 24–27; editorship, 83, 92; as
nuclear physics: APS division, 132, flagship journal, 24; growth of, 79–80,
229n31; as a discipline, 55; funding 86, 201–2; narrowing of focus, 13,
for, 102, 192; overlap with high en- 26–27; profitability of, 91–92; publi-
ergy physics, 10–11, 135–36, 177; cation delays, 78, 97, 99; publications
political influence of, 4, 8–10, 143, in, 110, 113, 221n36; solid state
176, 205; and population growth, 30; physicists consider abandoning, 79,
prestige of, 81–84, 200, 207–10; pub- 93–94; subdivision of, 142, 155, 200,
lications, 26, 79, 93, 227n76; research 240n2
in 88; and undersea saxophone, 182 Physical Review Applied, 200, 202, 240n2
nuclear weapons, 32, 34, 52, 112 Physics. See Journal of Applied Physics
Physics Today: articles in, 58, 87, 137,
Oak Ridge National Laboratory, 27, 103, 139, 152, 179, 228n11; as a discus-
137–38, 197 sion forum, 36, 80–84; founding
Office of Naval Research, 102, 121 of, 51, 204; discussions of unity in,
Oppenheimer, J. Robert, 9, 81, 209 185–87
Optical Society of America, 23, 25, 40, Physik der kondensierten Materie, 155,
201 159
optics, 7, 39, 49, 108, 155, 166, 227n76 physique du solide, 155–57, 235n23
Osgood, Thomas H., 35, 40 Pick, Heinz, 156
Pines, David, 158, 163
Pais, Abraham, 24–25, 207 Pippard, Brian, 147–48, 158, 168
Pake Report, The, 157–59, 162, 166, plasma physics, 8, 159, 191, 240n67
235n30 Porter, Bernard H., 27–30
Pake, George E., 94, 158 prestige: of the APS, 77–78; discussed at
Panofsky, Wolfgang, 115, 135 the Conference of Physicists, 47–48;
particle physics. See high energy physics glut of in postwar physics, 39, 148–49,
Pastore, John O., 115–17, 173 172; of high energy physics, 4; inter-
Pauli, Wolfgang, 12, 147, 241n7 national, 104; maintenance of, 7–10;
INDEX 277

solid state physics competes for, 73, as a historiographical category, 6, 196;


150–51, 159–61, 165, 168, 187, John Slater’s contributions to, 87–89;
189–90, 194, 204–5 John Van Vleck’s contributions to, 60,
Princeton University, 4, 27, 86–87, 89, 71; as a phenomenological category, 7;
143 in the Physical Review, 26–27, 201; of
productivity claims, 192–96 solids, 72, 102, 104, 146; as a source
professional identity: and applied re- of conceptual unity, 55–56, 204
search, 27, 36, 45, 54–56, 74, 207; quantum theory, 48, 62, 126
as conceived in different branches
of physics, 4–5, 81; of engineering, Rabi, Isidor Isaac, 7, 23, 131
21; impact of APS divisions on, 65; Radiation Laboratory, MIT, 23, 85, 109.
of physics as a whole, 7, 10, 12, 19, See also Massachusetts Institute of
31–32, 44, 54–56, 121, 134, 137, Technology
193, 199, 210; of solid state physics, Radio Research Laboratory, Harvard
9, 14, 78–79, 99, 118, 120, 129, 133, University, 85. See also Harvard
151–52, 158, 196, 202, 204, 241n7 University
Project Hindsight, 120 Read, Thomas A., 57
publishing: American Institute of Phys- reductionism: feminist critique of,
ics operations, 24–27, 30, 50–51; 239n43; importance to high energy
American Physical Society operations, physics, 135–36, 140, 170, 173, 181–
25–27, 41, 43; and condensed mat- 85, 197, 202, 205, 211; relationship
ter physics, 155–57; page charges, to spin-off claims, 177–78; solid state
225n49; popular, 184–85; problems, physicists’ objections to, 15, 136–37,
14, 76–101, 131, 209; outlets for, 13, 144–45, 148–50, 185–89, 206–7,
24–27, 200–202, 221n36, 223n9, 233n31; used to justify the SSC, 3
225n55 Research Laboratory of Electronics, 127
pure science: American Physical Society Review of Scientific Instruments, 25,
as a haven for, 12, 20, 59–61, 67; as a 36–37, 201, 228n22
category, 7, 227–28n11; as an ideal, 5, Reviews of Modern Physics, 25
12–14, 16–17, 18–31, 33, 35, 76, 104, Richter, Burton, 154, 182, 192
117, 132, 134, 135, 154, 165, 170, Richtmyer, Floyd K., 25
173, 198–204; and physics pedagogy, Robinson, Howard A., 84
63; relationship to applied physics, 5, Rowland, Henry Augustus: as an advocate
40–42, 84, 148; unity and, 69–70 for pure science, 12, 19–21, 26, 35,
41, 154, 170, 198–200, 204, 210,
quantum mechanics: of complex matter, 215n7; on the study of matter, 11
15, 94–96, 160, 163–64, 166, 168, Roy, Rustum, 167, 179
232n14; of elementary particles, 185;
278 INDEX

Santa Fe Institute, 163, 236n58 leading solid state physicist, 225n53;


Sawyer, Ralph A., 49, 218n44 perspective on the history of solid
Schmutzphysik, 12, 211, 241n7 state physics, 166
Schrieffer, J. Robert, 147, 158, 235n29 Slichter, Charles, 143
Schwitters, Roy, 177–78 small science, 178, 188–89, 206
science policy, 4, 9–11, 113, 130, 184, Smith, Cyril Stanley, 123, 130, 164
198, 206 Smoluchowski, Roman: arrival in United
Second World War, 9, 31, 35, 56, 62, 196, States, 56; correspondence with Ben-
208–9 jamin Lax, 110; correspondence with
Seitz, Frederick: as corporate shill, Frederick Seitz, 86; efforts to establish
220n21; correspondence with the Division of Solid State Physics,
Gaylord Harnwell, 80; education, 13, 39, 56–58, 60–67, 69, 219n9,
224n33; member of APS council, 221n31; member of Pake Report com-
77–78; member of the group of six, mittee, 158–59; vision for American
57, 61–62, 64, 67; member of Materi- physics, 39, 58, 63–65, 122, 153, 164
als Advisory Board, 123; presidency Smoot, George, 177–78
of National Academy of Sciences, Society of Rheology, 24
157; and the publication problem, solid state electronics, 130, 150
85–91, 92–98, 100; textbook author- solid state physics: definition of, 5–7;
ship, 72–73, 222n60; views on APS diversity of, 6, 8, 11, 16, 73, 90,
divisions, 41–42; views on postwar 95–96, 187; growth of, 7, 11, 14, 16,
personnel shortage, 84–85 76–79, 92–93, 229n31; establishment
semiconductors, 11, 94–95, 150, 163, of, 56–70; and identity of American
177–78 physics, 4–12, 19, 31, 54–56, 99,
Shockley, William, 57, 84, 86, 94, 142, 121, 134, 137, 193, 199, 202; as a
225n53 name, 15, 62–65, 73, 152–54; as op-
Siegel, Sidney, 57, 63, 221n43 posed to condensed matter physics,
SLAC. See Stanford Linear Accelerator 154–69; relationship to nuclear and
Slater, John Clarke: abstract approach to high energy physics, 8–10, 55, 81–84,
physics, 72; advisor of William Shock- 93, 132, 136, 143, 176–77, 192,
ley, 86; applications of the Wigner– 200, 205–10; slurs directed at, 12,
Seitz method, 90; calculational 211, 241n7; as an unusual category,
approach to physics, 89; conflict 7, 13, 29–31, 70, 152–53. See also
with Niels Bohr, 88, 244n41; early applications
career, 87; involvement with MIT’s Soviet Union, 124, 172, 174, 221n36,
interdisciplinary laboratory, 126–128; 232n3
involvement with the National Magnet spin-offs, 15, 174–82, 185, 187, 193, 202
Laboratory 107, 109; named as a “squalid state physics,” 12, 241n7
INDEX 279

Stanford Linear Accelerator, 104, 115, United States Congress: and NAL fund-
135, 207 ing, 104; physicists’ testimony before,
Stanford University, 33, 86, 126, 142, 111–12, 115, 135, 171–74, 176, 179,
218n33 181, 184–86; and SSC funding, 3,
Star Wars, 16 171–192
stereo equipment, 11, 208 United States Department of Defense. See
Strategic Defense Initiative, 175–76 Department of Defense
Superconducting Super Collider, 3–5, United States Department of Energy. See
8–10, 15–16, 170–98, 202, 205–6, Department of Energy
239n43 United States Navy. See Office of Naval
superconductivity: industrial uses, 11; as Research
constituent field of solid state physics, unity: divisions as a threat to, 13, 42, 65,
110, 164, 167, 235n29, 236n48; in- 86; as justification for the SSC, 185–
correctness of theories of, 147; Nobel 87; political, 35, 37–38, 40, 56, 70,
Prize for, 150, 158; and magnets, 175, 73–75; intellectual, 55–56, 70; Physics
177–79 Today as catalyst for, 80–82, 84; and
superfluidity, 6, 156, 158, 162 reductionism, 233n37–38, 239n58
Sutton, Richard M., 46 University of Bristol. See Bristol, Univer-
synchrotron radiation, 103–4, 198 sity of
Szilard, Leo, 88, 209 University of California, Santa Barbara. See
California, Santa Barbara, University of
Tate, John Torrence, 26–27, 51, 79, 91–92 University of Cambridge. See Cambridge,
teaching. See pedagogy University of
technological applications. See University of Chicago. See Chicago, Uni-
applications versity of
textbooks, 72–73, 90, 124, 136, 227n7 University of Illinois. See Illinois, Univer-
textiles, 24, 85, 218n34 sity of
thermodynamics, 6–7, 19, 21, 44, 105 University of Pennsylvania. See Pennsylva-
Tocqueville, Alexis de, 18–19 nia, University of
Townes, Charles, 145, 158–59
training. See pedagogy Van Vlack, Lawrence, 124
transistor, 11, 54, 72–73, 142–43, 167, Van Vleck, John Hasbrook: education,
222n64, 232n17 87, 126; extension of Wigner-Seitz
Trivelpiece, Alvin, 175–76 method, 90; impressions of Physical
Review, 26; impressions of Physics
United States Air Force, 108, 110–11, Today, 83; as leading solid state physi-
115, 117, 123 cist, 225n53; Nobel Prize, 142–43,
280 INDEX

Van Vleck, John Hasbrook: Nobel Prize Westfall, Catherine, 10, 103, 197
(cont.), 150; opposition to APS divi- Westinghouse Electric Corporation, 57,
sions, 58–63, 65–67, 74–75, 77–78, 105, 142–43, 191, 221n28, 227n8
80, 88, 204, 219–20nn11–19, 223n4; Whewell, William, 30, 33
work on exchange interaction 71; Wigner–Seitz method, 89, 225n45
work on magnetism, 96 Wigner, Eugene, 87–90, 219–20n14,
Vietnam War, 103, 110–13, 133 220n19, 224n33
Wilson, Robert, 115–17, 173, 177
Waterfall, Wallace, 24, 37–39, 217n15 World War I. See First World War
Weart, Spencer, 55, 95, 154 World War II. See Second World War
Weinberg, Alvin, 137–140, 144–45,
147–150 X-ray diffraction, 103, 166
Weinberg criterion, 138–40, 145, 149,
184–85, 187 Zeitschrift für Physik, 155, 234n14
Weinberg, Steven, 151, 171, 173, 178–79, Zener, Clarence, 71
182–85, 187, 202
Weisskopf, Victor, 139–40, 144, 183,
232n12

You might also like