<<

University of Alberta

Radiation, Researchers, and the United States Atomic Energy Commission: Biomedical Research from the Early Twentieth Century to the Early Cold War

by

Katherine Jane Zwicker

A thesis submitted to the Faculty of Graduate Studies and Research in partial fulfillment of the requirements for the degree of

Doctor of Philosophy in History

Department of History & Classics

©Katherine Jane Zwicker Spring 2012 Edmonton, Alberta

Permission is hereby granted to the University of Alberta Libraries to reproduce single copies of this thesis and to lend or sell such copies for private, scholarly or scientific research purposes only. Where the thesis is converted to, or otherwise made available in digital form, the University of Alberta will advise potential users of the thesis of these terms.

The author reserves all other publication and other rights in association with the copyright in the thesis and, except as herein before provided, neither the thesis nor any substantial portion thereof may be printed or otherwise reproduced in any material form whatsoever without the author's prior written permission. Library and Archives Bibliotheque et Canada Archives Canada

Published Heritage Direction du Branch Patrimoine de I'edition

395 Wellington Street 395, rue Wellington Ottawa ON K1A0N4 Ottawa ON K1A 0N4 Canada Canada Your file Votre reference ISBN: 978-0-494-87892-7

Our file Notre reference ISBN: 978-0-494-87892-7

NOTICE: AVIS: The author has granted a non­ L'auteur a accorde une licence non exclusive exclusive license allowing Library and permettant a la Bibliotheque et Archives Archives Canada to reproduce, Canada de reproduire, publier, archiver, publish, archive, preserve, conserve, sauvegarder, conserver, transmettre au public communicate to the public by par telecommunication ou par I'lnternet, preter, telecommunication or on the Internet, distribuer et vendre des theses partout dans le loan, distrbute and sell theses monde, a des fins commerciales ou autres, sur worldwide, for commercial or non­ support microforme, papier, electronique et/ou commercial purposes, in microform, autres formats. paper, electronic and/or any other formats.

The author retains copyright L'auteur conserve la propriete du droit d'auteur ownership and moral rights in this et des droits moraux qui protege cette these. Ni thesis. Neither the thesis nor la these ni des extraits substantiels de celle-ci substantial extracts from it may be ne doivent etre imprimes ou autrement printed or otherwise reproduced reproduits sans son autorisation. without the author's permission.

In compliance with the Canadian Conformement a la loi canadienne sur la Privacy Act some supporting forms protection de la vie privee, quelques may have been removed from this formulaires secondaires ont ete enleves de thesis. cette these.

While these forms may be included Bien que ces formulaires aient inclus dans in the document page count, their la pagination, il n'y aura aucun contenu removal does not represent any loss manquant. of content from the thesis. Canada ABSTRACT

On the 6th and 9th of August 1945 the United States dropped atomic

bombs on the Japanese cities of Hiroshima and Nagasaki. Both cities were

utterly devastated, in part, due to the sheer blast produced by the bombs, but also

due to the radioactive fallout that contaminated each region. Biomedical

radiation researchers from around the world studied the resulting radiation

sickncss that affected atomic bomb survivors. The work of American

biomedical researchers marked a distinct phase in the history of biomedical

radiation research in the United States, because researchers had never before

been able to study the biological effects of radiation exposure amongst such a

large group of individuals. Also, American researchers' investigation of bomb

survivors was the first of many postwar initiatives of the federal government to

fund biomedical radiation research. This was, however, one chapter in a longer

history that began at the turn of the twentieth century.

This dissertation examines the development of biomedical radiation

research in the United States from the early twentieth century to the early Cold

War period. It focuses on three salient and closely related aspects of the field:

the practice of interdisciplinary collaboration and creation of hybrid expertise; the development of an influential network of biomedical radiation researchers; and the role of researchers in helping to define and respond to changing social and political priorities within the United States. It argues that the primary reason biomedical radiation research flourished throughout the century was that researchers were socially and politically responsive. They were attentive to changing social and political circumstances and promoted their research accordingly. Throughout the century they promised that their work would advance medicine, especially the diagnosis and treatment of cancer. During

World War II and later, they also accepted responsibility for investigating the hazards associated with nuclear weapons development and atomic warfare. By doing so, researchers were able to secure sufficient and, at times, abundant resources to investigate the application of radiation in biology and medicine.

The pursuit of this research within a large government-funded research enterprise blurred the lines between civilian and military spheres and helped characterize the Cold War. ACKNOWLEDGEMENTS

I want to sincerely thank my supervisor, Robert Smith, for his wonderful

guidance during my graduate studies. There were, of course, moments

throughout writing this dissertation, at which I could not see an end in sight. On

many occasions, Robert reminded me that I would not know what 1 was writing,

until I got it on the page. Perhaps a simple reassurance, but one I valued greatly.

From Robert I have learned so much about the process of research and writing

that I will continually draw on throughout my career.

I have been very fortunate to have found, not one excellent mentor, but

two. I owe a great deal of thanks to Susan Smith for her considerable investment

in my professional development. Over the years, Susan has been generous with

her time and advice and has really helped nurture my confidence as an academic.

Many of my colleagues at the University of Alberta deserve recognition.

David Marples, for instance, has been a constant support throughout my studies and, at one time, a great teammate on the soccer field. 1 have very much enjoyed being a part of the History & Classics Department, as well as the history of science and medicine communities on campus. For helping to foster these communities, thanks to Lesley Cormack, Ken Moure, Pat Prestwich, and

Andrew Ede. Roberta Lexier, Melanie Niemi-Bohun, Sharon Romeo, Robyn

Braun, Jaymie Heilman—thank you all for making work a bit more fun!

As for family and friends, I suspect my words will fail me in trying to express my gratitude to you all for your love and support over the years. To my

Edmonton, Wolfville, and Lunenburg friends, I will take you with me, wherever I go. To my ever-growing and absolutely wonderful family—well, 1 know I am one of the lucky ones to have each of you by my side. Dad, you are my rock;

Mom, my memories of you are a constant inspiration; Meredith and Phil, to me you will always be older and so much wiser and I will look to you to show me the way. Amy, I can hardly believe that there was ever a time when we wanted to go our separate ways. Through the hills and the valleys, we will walk in the best company. TABLE OF CONTENTS

Introduction 1

Chapter 1 - X-Rays, Radium, and the Early Twentieth-Century Development of Interdisciplinary Collaboration 27

Chapter 2 - Biomedicine and Bombs: Health and Safety in the 74

Chapter 3 - From War to Peace: Institutionalizing Biomedical Radiation Research within the Atomic Energy Commission 132

Chapter 4 - The Building Blocks of Research: Radioisotopes Distribution 179

Chapter 5 - The ABC's of the AEC's Biomedical Research: Fellowships and Educational Initiatives 215

Chapter 6 - Waging War on Another Front: The AEC Joins the Cancer Establishment 263

Conclusion 292

Bibliography 307 LIST OF TABLES

Table 1: Radioisotope Distribution to Non-AEC Institutions

Table 2: Diagnostic & Therapeutic Uses of Radioisotopes, 1949 LIST OF FIGURES

Figure 1: AEC Organization Chart, December 1948

Figure 2: ORINS Organizational Chart, September 1949 ABBREVIATIONS & TERMS

ACXRP - Advisory Committee on X-ray and Radium Protection, established in 1928.

ABCC - Atomic Bomb Casualty Commission, established in 1947.

AEC - Atomic Energy Commission, established 1 January 1947.

ACBM - Advisory Committee on Biology and Medicine, established in September 1947 within the AEC.

ACS - American Cancer Society, established in 1913 as the American Society for the Control of Cancer. The organization adopted its current name in 1945.

ARRS - American Roentgen Ray Society, established in 1900.

ARS - American Radium Society, established in 1916.

AUI - Associated Universities, Inc., consortium of universities, established in 1946.

BNL - Brookhaven National Laboratory, established in 1947.

C-14- carbon-14.

CWS - Chemical Warfare Service, established in 1918.

DBM - Division of Biology and Medicine, established in October 1947 within the AEC.

FBI - Federal Bureau of Investigation, established in 1908.

H-3 - hydrogen-3.

ICXRP - International Committee on X-Ray and Radium Protection, established in 1928.

Interim Medical Advisory Committee - formed in January 1947 by the AEC to advise on the future of biomedical research within the AEC.

JCAE - Joint Committee on Atomic Energy, congressional oversight committee of the AEC, established in 1947. MED - Manhattan Engineer District, established in the summer of 1942 within the Army Corps of Engineers for the development of the atomic bomb during World War II. The MED transferred authority over the nation's nuclear weapons complex to the Atomic Energy Commission as of 1 January 1947, and the MED officially ceased to exist 15 August 1947.

Medical Advisory Committee - formed by the MED in September 1946 to assess the legacy of the MED's health and safety work and advise on future activities.

Medical Board of Review - formed in May 1947 by the NAS at the request of the AEC to provide expert advice on the future of biomedical radiation research within the AEC.

MIT - Massachusetts Institute of Technology, established in 1861.

NACC - National Advisory Cancer Council of the NCI, established in 1937.

NARA Atlanta - U.S. National Archives and Records Administrations, Southeast Region, Atlanta, Georgia.

NARA College Park - U.S. National Archives and Records Administrations II, College Park, Maryland.

NAS - National Academy of Sciences, established 1863.

NCB - Naval Consulting Board, established 1915.

NCI - National Cancer Institute, established 1937.

NCRP - National Council on Radiation Protection and Measurements, the successor to the ACXRP. Reorganized and renamed as such in 1964.

NDRC - National Defense Research Committee, established in June 1940.

NRC - National Research Council, established 1916 as the research arm of the NAS.

NRU - National Research Universal reactor, Chalk River Laboratory, Canada. The Laboratory was established in 1944 and the NRU began operation in 1957.

NSF - National Science Foundation, established in 1950.

NYOO - New York Operations Office, of the AEC. ORINS - Oak Ridge Institute for Nuclear Studies, consortium of universities, established in 1946.

ORNL - Oak Ridge National Laboratory, established in 1943 as the Clinton Laboratories, and renamed in 1948.

OSRD - Office of Scientific Research and Development, established in June 1941.

P-32 - phosphorous-32.

RAEP - Rochester Atomic Energy Project, MED- and AEC-funded research project at the University of Rochester, established in 1943.

RG - Record group.

RSNA - Radiological Society of North America, established in 1915 as the Western Roentgen Society, and renamed in 1919.

Tc-99m - teehnetium-99.

UCSF - University of California, San Francisco, established in 1864.

USPHS - United States Public Health Service, established in 1798 as a network of marine hospitals, and renamed in 1902. 1

INTRODUCTION

In March 1953 Dr. Shields L. Warren delivered a speech to mark the opening of a new cancer research hospital. Warren was a leading physician and researcher with decades of experience investigating causes, diagnostic methods, and therapeutic procedures related to cancer. "We are today dedicating this

Argonne Cancer Research Hospital," he began, "because of the faith of the

American people that God in Whom we trust has created an orderly universe, has endowed man with the power to comprehend that universe, and that as man gains that knowledge he can control and remedy disorder," including finding a cure for cancer.1 Throughout his speech, Warren explained the role that this hospital would play in acquiring knowledge and remedying disorder. Its main purpose was to investigate uses of radiation to improve cancer care. He expressed his deep faith in science, which was hardly a new sentiment within either the scientific or medical communities. However, the bold message Warren delivered in 1953 reflected a growing confidence that he and many shared. This confidence stemmed from the reorganization and massive expansion of the scientific enterprise in the postwar period facilitated by federal funding, often for weapons development projects, but also for biomedical research.

This study examines the history of biomedical radiation research in the

United States from the beginning of the twentieth century through to the early

Cold War—the point at which Warren dedicated the Argonne Cancer Research

' Shields Warren, "Argonne Cancer Research Hospital, 14 March 1953," NARA College Park, RG 326, General Correspondence 1951-1958, Box 67, Organization and Management 7 - Division of Biology and Medicine, 1. 2

Hospital in Chicago, IL. It begins with the discovery of radiation and the early research conducted by scientists and physicians who sought to determine medical applications of radiation. It also encompasses the development of technologies, including accelerators and nuclear reactors, which were used to produce radioisotopes, a new source of radiation. The history of biomedical radiation research, however, entails much more than just the discovery of knowledge and technologies. In particular, this work focuses on three salient and closely related aspects of the growth of biomedical radiation research: the practice of interdisciplinary collaboration and creation of hybrid expertise or expertise that combines the knowledge and skills of multiple disciplines; the development of an influential network of biomedical radiation researchers; and the role of these researchers in helping to define and respond to changing social and political priorities within the United States.

The most significant shifts in the social and political landscape of the

United States that affected biomedical radiation research in the twentieth century were driven by World War II.2 When the United States launched the Manhattan

Project, which aimed to create an atomic bomb, the history of biomedical

" For the mobilization of science for war see, among others, Robert Buderi, The Invention That Changed the World: How a Small Group of Radar Pioneers Won the Second World War and Launched a Technological Revolution, The Sloan Technology Series (New York: Simon & Schuster, 1996); Stanley Goldberg, "Inventing a Climate of Opinion: Vannevar Bush and the Decision to Build the Bomb," /sis 83, no. 3 (1992): 429-52; Mark Harrison, "The Medicalization of War—the Militarization of Medicine," Social History of Medicine 9, no. 2 (1996): 267-76; Richard Ci. Hewlett and Oscar E. Anderson, The New World: A History of the United States Atomic Energy Commission, Volume 1, 1939-1946 (University Park, PA: Pennsylvania State University Press, 1962); Lillian Hoddeson, Critical Assembly: A Technical History of Los Alamos During the Oppenheimer Years, 1943-1945 (New York: Cambridge University Press, 1993); Jeff Hughes, The Manhattan Project: Big Science and the Atom Bomb (New York: Columbia University Press, 2002). For sources specific to wartime biomedical radiation research, see Barton C. Hacker, The Dragon's Tail: Radiation Safety in the Manhattan Project, 1942-1946 (Berkeley, CA: University of California Press, 1987). 3 radiation research and the development of wartime government- and military- funded research and development converged. Biomedical researchers were

recruited during the war to investigate the health hazards of radiation and help protect Manhattan Project workers from radiation exposure. Wartime health and safety concerns served as the primary link between biomedical radiation research and national security, between biomedical radiation researchers and the state.

When the war ended, the United States continued to produce nuclear weapons under the authority of a new government agency, the Atomic Energy

Commission (AEC). As had been the case during the Manhattan Project, the ongoing weapons development ensured that radiation safety remained a concern within the government, especially within the AEC. Weapons testing and the prospect of future nuclear warfare also fuelled their concern.3 The AEC created a biomedical program that supported ongoing research related to radiation safety, but it also supported biomedical research that was unrelated to nuclear weapons development, testing, and warfare. Researchers who had assumed responsibilities within a government and military bureaucracy and benefitted from the tools they used—namely sources of radiation and related technologies—played a significant role in the creation of a broad biomedical program within the AEC.

This study pays special attention to the creation and development of the

A EC's biomedical program during the late 1940s and the first few years of the

1950s to show the complexity of the relationships between scientists and the

' The United States' weapons testing program, especially, generated concern about radiation safety. See Barton C. Hacker, Elements of Controversy: The Atomic Energy Commission and Radiation Safety in Nuclear Weapons Testing, 1947-1974 (Berkeley, CA: University of California Press, 1994); Scott Kirsch, "Harold Knapp and the Geography of Normal Controversy: Radioiodine in the Historical Environment," Osiris 19 (2004): 167-81. 4 government, in particular, and between the social, political, and technical aspects of biomedical research. The history of the AEC's biomedical program during those years highlights the influence of the nascent Cold War on biomedical research, but, so too does it show that the development of this field of research helped create the culture of Cold War America. The boundaries between science, government, and national defense were blurred as biomedical researchers and

AEC administrators made an effort to redefine and in some cases create research programs that served various purposes and tied together numerous institutions.

QUESTIONS

The key questions this dissertation will answer pertain to the relationship between science and society. For instance, what factors were most important in fuelling the development of biomedical radiation research from the beginning of the twentieth century through to the early Cold War? More specifically, how did social, political, and economic factors shape biomedical radiation research and, in the inverse, how did this research shape society? With particular reference to the postwar development of biomedical radiation research, to what extent did researchers adapt to changing circumstances and attempt to mold their research according to perceived social and political aims? What role did they play in creating social and political agendas, not only in the postwar period, but from the start of the century? And finally, how did biomedical radiation researchers help constitute the Cold War in the United States, even when pursuing goals and continuing practices that had characterized the field since early in the twentieth century? 5

CONTEXT OF THE STUDY

The mobilization of science for war and ongoing government support of science following the war influenced the creation of a new political economy of

research. What is meant here, by political economy of research, is the supply and demand—the market—for research. That is, the social, political, and economic

factors that encouraged and supported the pursuit of scientific research, generally, or specific fields of research, in particular.4 While this study especially focuses on the expansion of biomedical radiation research during the postwar development of a new political economy of research, it aims to identify where continuities existed between early twentieth-century biomedical radiation research and wartime and postwar development of this field. The period between the discovery of radiation in the final years of the nineteenth century and the United

States' entry into World War II is important in the history of biomedical radiation research. It was during those decades that researchers established their commitment to investigating not just the properties of radiation, but also the use of radiation in biological research and for medical applications. Their pursuit of research within the institutional, social, and political context of that time resulted in particular trends and practices—most notably interdisciplinary research or collaborative research amongst physical, biological, and medical scientists.

4 My understanding and use of political economy of research is influenced by Angela N. H. Creager and Jessica Wang who both explicitly examine shifts in the postwar political economy of science. See Angela N. H. Creager, "Nuclear Medicine in the Service of Biomedicine: The U.S. Atomic Energy Commission's Radioisotope Program, 1946-1950," Journal of the History of Biology 39 (2006): 649-84; Angela N. H. Creager, "Tracing the Politics of Changing Postwar Research Practices: The Export of'American' Radioisotopes to European Biologists," Studies in the History and Philosophy of Biological and Biomedical Sciences 33 (2002): 367-88; Jessica Wang, "Liberals, the Progressive Left, and the Political Economy of Postwar American Science: The National Science Foundation Debate Revisited," Historical Studies in the Physical and Biological Sciences 26, no. 1 (1995): 139-66. 6

Interdisciplinary collaboration helped establish a network of researchers who pushed beyond traditional disciplinary boundaries to develop research programs and, over time, new biomedical disciplines.

During the first few decades of the century, X-rays and radium were studied and exploited for diagnostic and therapeutic purposes.5 Naturally occurring radioactive elements were also used in pioneering tracer studies to investigate biological processes. By the 1930s, various technological developments, especially the artificial creation of radioisotopes, expanded the tools available to researchers for such studies. The history of biomedical radiation research during this period reveals much more than technological developments and their application in research and clinical practice, however.

From this history we see that the expansion of biomedical radiation research was also driven by researchers who effectively negotiated relationships within the scientific and medical communities and beyond to gain access to resources and make their work socially relevant. This research evolved within the institutional context of universities, medical schools, and hospitals, all of which established a greater emphasis on scientific research in the early decades of the century.6

Biomedical radiation research was also influenced by the history of private

5 As we will see in Chapter One, X-rays were discovered in 1895 and radium, in 1898. Scientists and physicians began investigating the medical uses of these sources of radiation soon after their discovery. See Claudia Clark, Radium Girls: Women and Industrial Heath Reform, 1910-1935 (Chapel Hill, NC: University of North Carolina Press, 1997), 43. 6 On educational reform, the rise of research in universities, and growth of scientific medicine, Roger L. Geiger, To Advance Knowledge: The Growth of American Research Universities, 1900- 1940 (New Brunswick, NJ: Transaction Publishers, 2004); George Weisz, Divide and Conquer: A Comparative History of Medical Specialization (New York: Oxford University Press, 2006). 7

philanthropy. Foundations were an important source of patronage for American science, especially during the interwar period.7

As already mentioned, the establishment of an atomic bomb project during

World War II raised concerns within the government about radiation safety. The health hazards of radiation were not a new problem. Rather they had been

investigated for decades, largely in response to health problems that resulted from

medical and industrial uses of radiation.8 It was only during the war, though, that

the government and military assumed responsibility for radiation safety. As a

result, biomedical radiation research became part of a much larger scientific and technological research and development project, one that was overtly shaped by the social, political, and economic context of war. Wartime research and development illuminated the social and political importance of science and encouraged many individuals and institutions to reevaluate the relationship between science and society both during and following the war.

In the postwar period, individuals within the scientific community and federal government sought to shape the relationship between science and the state in ways that reflected the relationship that existed both before and during the war.

Scientists—biomedical radiation researchers included—actively tried to influence government and military policy to secure increased support for science. Their efforts were played out in relation to debates regarding postwar military funding

7 On foundations, see Robert E. Kohler, Partners in Science: Foundations and Natural Scientists, 1900-1945 (Chicago, IL: Press, 1991); Robert E. Kohler, "Science, Foundations, and American Universities in the 1920s," Osiris 3 (1987): 135-64; Nathan Reingold, Science, American Style (New Brunswick, NJ; Rutgers University Press, 1991), 190-223. K Both within the U.S. and international community, scientists, physicians, and industry cooperated to form radiation safety advisory committees in the late 1920s. See Gilbert F. Whittemore, "The National Committee on Radiation Protection, 1928-1960: From Professional Guidelines to Government Regulation," (Ph.D. diss., Harvard University, Cambridge, MA, 1983). 8 of scicnce, as well as debates over the creation of the National Science

Foundation, a new federal agency designed to support civilian science.9

While many within the scientific community and government emerged from the war experience with a strong belief that science was important to the nation's welfare, determining how science could be used best to help solve social, political, and military problems was complex. By examining the creation of the

Atomic Energy Commission's biomedical bureaucracy, policies, and programs we see how scientists and the government negotiated and cooperated to create a biomedical program that embodied goals of both scientists and the government.

For the latter, these goals were shaped by the AEC's defense agenda.

Within the institutional framework of the AEC, some elements of the

Commission's biomedical program constituted a new way of doing research, but the core objectives of the AEC's programs were, in fact, rooted in pre-existing practices of biomedical radiation research. Biomedical radiation researchers who, in the 1920s and 1930s, engaged in a process of discipline-building within local contexts, continued to do so postwar, but on a national scale. They did so within the AEC's research and development enterprise, which was vastly larger than the academic and hospital settings. Whether working in hospitals and universities prior to the war, or working under AEC contract or within AEC laboratories

" On the National Science Foundation debate which began prior to the end of the war and continued through to the creation of the agency in 1950, see J. Merton England, A Patron for Pure Science: The National Science Foundation's Formative Years. 1945-1957 (Washington, DC: National Science Foundation, 1983); Daniel J. Kevles, "The National Science Foundation and the Debate over Postwar Research Policy, 1944-1945: A Political Interpretation of Science—the Endless Frontier," Isis 68, no. 1 (1977): 4-26; Wang, "Political Economy of Postwar American Science," 139-66. 9 following the war, the role of interdisciplinary research remained important to the goal of furthering biomedical radiation research.

As indicated by Dr. Shields Warren in his dedication of the Argonne

Cancer Research Hospital, one of the most important goals shared by both the scientific community and government was scientific advancement. The motivation for scientific advancement often varied, but as the history of the

AEC's biomedical research shows, the tools used and institutions involved served as a link between science and the state. They also linked what, on the surface, seemed to be a civilian agenda of advancing science for the sake of science and public welfare, with a defense agenda of pursing scientific research for the development of military technologies and preparation for atomic warfare. In the case of biomedical radiation research, both civilian and defense goals were the responsibility of the AEC's Division of Biology and Medicine (DBM) and the

Advisory Committee on Biology and Medicine (ACBM). With the DBM,

ACBM, and the policies and programs they managed acting as a common bond between science, the state, and national security, biomedical research was part of an enterprise that merged civilian and military spheres and thereby contributed to the building of a Cold War national security state. The Argonne Cancer Research

Hospital, for instance, stood as a monument not only to the government's commitment to fighting cancer, but also the government's endeavor to build a strong scientific enterprise that could produce knowledge relevant to any challenge the nation might encounter. This research hospital, along with so many other research policies and programs, also represented the willingness of 10

researchers to work within—to become part of—a large government scientific

enterprise.

IMPORTANT THEMES & THESIS

Beyond investigating the broad issue of the relationship between science

and society, this dissertation explores three key issues. The first is the creation

and evolution of scientific disciplines. This is an issue that historians have

examined, the most relevant study being Robert Kohler's account of the creation of biochemistry as a biomedical discipline in the early twentieth century.10

Kohler presents a history of a discipline that emerged amidst reforms in both

university science departments and medical schools. His study makes important contributions to the history of biomedical disciplines in the United States, as well as the political economy of science during the first half of the twentieth century.

This dissertation builds on Kohler's work by focusing on radiation as an object of

inquiry and as a resource that, when employed, could alter existing scientific and biomedical disciplines or encourage the creation of new ones. Both before and after the war, radiation was central to the process of building disciplines and institutions.

I also draw on Timothy Lenoir's work on disciplines. Lenoir examines the formation of scientific, medical, and engineering disciplines and argues that they are institutions constructed by scientists to support the social and cultural

10 Robert E. Kohler, From Medical Chemistry to Biochemistry: The Making of a Biomedical Discipline (New York: Cambridge University Press, 1982). 11

authority of their work.11 We will see that the researchers who helped establish

and develop new biomedical disciplines organized around radiation were very

much engaged in a process of discipline-building that encompassed the pursuit of successful research, but also the marketing of research within institutional and

broad social and cultural settings.

The second key issue this dissertation examines is the categorization of science and the government and the dichotomy between civilian and military

research. Historians Angela N. H. Creager, Timothy Lenoir and Marguerite Hays,

Michael A. Dennis, and Gerald Kutcher have explored aspects of this issue both

in relation to postwar research in general, and in relation to the AEC's research specifically.12 Their contributions to the literature will be discussed in greater detail below. This study emphasizes that science was not an autonomous endeavor within society, especially in the postwar period when there was no clear divide between science and government. Many scientists became part of the government bureaucracy during and following the war and, as part of the government, it is not clear that they advanced or executed a so-called government agenda. Indeed, in the history of the AEC's biomedical research we see that those scientists who worked within the government bureaucracy strove to achieve goals

11 Timothy Lenoir, Instituting Science: The Cultural Production of Scientific Disciplines (Stanford, CA: Stanford University Press, 1997), 3 and 55. Of Creager's work on the AEC's radioisotopes program, the following article is most relevant to the categorization of civilian and defense research: Creager, "Nuclear Medicine in the Service of Biomedicine," 648-84; Michael A. Dennis, '"Our First Line of Defense:' Two University Laboratories in the Postwar American State," Jsis 85, no. 3 (1994): 427-55; Gerald Kutcher, "Cancer Therapy and Military Cold-War Research: Crossing Epistemological and Ethical Boundaries," History Workshop Journal 56 (2003): 105-30; Gerald Kutcher, Contested Medicine: Cancer Research and the Military (Chicago, IL: University of Chicago Press, 2009); Timothy Lenoir and Marguerite Hays, "The Manhattan Project for Biomedicine," in Controlling Our Destinies: Historical, Ethical, and Theological Perspectives on the Human Genome Project, ed. Phillip R. Sloan (Notre Dame, IN: University of Notre Dame Press, 2000). 12 established earlier in academic or medical settings. Regardless of what or who drove government-funded science, the end result was that research was well integrated into various aspects of society. Furthermore, the pursuit of research within a large scientific enterprise helped constitute the character of postwar

American society.

The third central issue that this dissertation examines is the creation and nature of biomedical expertise. What constituted biomedical expertise over time?

Was it defined relative to other scientific fields or in relation to medical applications? Could expertise be obtained based on entrepreneurial or administrative achievements or was it mostly a reflection of achievements in research? This issue is relevant to other historical processes that unfolded throughout the twentieth century. For instance, the professionalization of scientists through educational standards and professional societies illustrates how scientists have defined expertise both within scientific disciplines and the larger scientific community. Similarly, educational standards, professional societies, and licensing procedures have been integral to establishing medical expertise.

Throughout the twentieth century, scientific expertise was also defined by researchers' participation in government agencies. The creation of biomedical expertise was a process that, over time, was played out in the scientific and medical communities and in government bureaucracy.

This study argues that biomedical radiation research flourished throughout the century not only because researchers made significant scientific discoveries, but also because they were socially and politically responsive. That is, 13 biomedical radiation researchers were attentive to changing social and political circumstances and they promoted and adapted their research accordingly.

Throughout the century they made promises to academic, philanthropic, and federal patrons that their work would advance medicine, especially the diagnosis and treatment of cancer. During World War II and following, they also accepted responsibility for investigating the hazards associated with nuclear weapons development and atomic warfare. By doing so, researchers were able to secure sufficient and, at times, abundant resources to investigate the application of radiation in biology and medicine.

Interdisciplinary collaboration and the development of hybrid expertise were crucial to advancement in this field. Biomedical radiation research involved expensive technologies and required expertise that extended beyond the fields of biology and medicine. Physicists and engineers helped life scientists to understand the properties of radiation and the technologies they used.

Interdisciplinary collaboration and the resulting hybrid expertise not only allowed researchers to undertake research that ranged from investigations of leukemia treatments to investigations of post-atomic blast radiation sickness, they also contributed to the expansion of a professional network of researchers. An informal, yet influential network of radiation researchers developed before World

War II. Most of these researchers were mobilized during the war and emerged from their wartime work with sufficient authority to influence postwar research and development policies and programs. 14

The role of researchers in influencing federal science policy and research agendas was especially important to the development of a broad biomedical research program within the AEC. Many historians have examined the impact of the federal government on science following World War II. The postwar expansion of biomedical radiation research should be attributed to more than just the influx of government and military dollars or the necessity of defense research and development during the Cold War, though. Researchers were able to establish their pre-existing research and discipline-building goals within the

AEC's biomedical program. Beyond that, the role of scientists acting within or influencing federal agencies and the expansion of science is part of what characterized Cold War culture and the growth of government bureaucracy and military power. The history of biomedical radiation research is, therefore, one of society shaping science and science shaping society.

Furthermore, the AEC's biomedical program was often greatly influenced by researchers outside, as well as inside, of the AEC. This point helps to establish a new historiographical focus in that it offers a lab bench-up approach that draws attention to more of the historical actors and interests that shaped the federal government's postwar policies and programs. It highlights the porous boundaries between the AEC as a government agency and researchers associated with a range of institutions. Further, it contends that researchers' ability to shape the AEC's biomedical program was, at least in part, rooted in the prewar development of biomedical radiation research.

HISTORIOGRAPHY 15

Since this study spans the first half of the twentieth century it draws on

and contributes to literature related to the formation of biomedical fields, as well

as the wartime and early postwar development of science. There is limited

literature on the emergence of biomedical fields in the early twentieth century.

However, as mentioned above, Robert Kohler's work on the creation of

biochemistry as a biomedical discipline is an exemplary study of a biomedical discipline that helps to contextualize the process through which biomedical

radiation research was institutionalized within new fields including biophysics,

radiobiology, and health physics, all of which are relevant to this study.13 The

main issues related to the creation of biomedical disciplines include the relationship between biomedical research and other fields of scientific research, the increased role of research in medical education and specialization, and the relationship between biomedical research and society.14 The early development of biomedical radiation research engages with each of these issues. Thus, this study is both informed by, but also expands upon literature on, the history of physics, medical specialization, and the political economy of research which was shaped by the relationship between science, medicine, and society.

Daniel Kevles' The Physicists: The History of a Scientific Community in

Modern America is the most comprehensive history of physics in the United

11 Kohler, Medical Chemistry to Biochemistry. See also, Caroline Hannaway, ed., Biomedicine in the 20th Century: Practices, Policies, and Politics (Amsterdam, The Netherlands: IOS Press, 2008). 14 On the growing importance of research within universities, especially graduate education, see Geiger, To Advance Knowledge: The Growth of American Research Universities, 1900-1940. For an examination of the rise of research in medical education and related expansion of biomedical fields, see Kohler, Medical Chemistry to Biochemistry, Weisz, Medical Specialization. 16

States.15 In it he examines the trajectory of research problems and the rise of the physicist as a prominent figure within both the scientific community and the government-military bureaucracy that formed circa World War II. His study

illustrates that physicists operated within, and played an integral role in shaping, a changing political economy of science.

Evelyn Fox Keller's study of the influence of physics on molecular biology also examines the political economy of science during the early and mid- twentieth century specifically focusing on the relationship between disciplines.16

She notes the social authority associated with physics and that biologists were able to acquire a level of authority by adopting ideas and technologies common to physics, and by pursuing collaborative projects with physicists. Keller's evaluation of the impact of interdisciplinary collaboration helps to illuminate aspects of the relationship between physics and the pioneering biomedical radiation research pursued in the early twentieth century. Physicists shared ideas, technologies, and resources with life scientists who were interested in investigating the biological effects of radiation and the uses of radiation in biomedical research and medical practice. However, physics was not the only source of authority shared within the collaborative relationships formed between physicists and life scientists.

As is evident in the history of physicist Ernest Lawrence's Radiation

Laboratory at the University of California, Berkeley, the medical applications

15 Daniel J. Kevles, The Physicists: The History of a Scientific Community in Modern America (Cambridge, MA: Harvard University Press, 1995). Evelyn Fox Keller, "Physics and the Emergence of Molecular Biology: A History of Cognitive and Political Synergy," Journal of the History of Biology 23, no. 3 (1990): 389-409. Also, on the relationship between disciplines, see Lenoir, Instituting Science. 17

derived from radiation research attracted patronage that was central to the market

for research in the pre-World War II period. The value of biomedical research

and pursuit of medical applications within the Radiation Laboratory is well

explored by Robert Seidel and John Heilbron in their history of that laboratory.17

From Seidel's and Heilbron's study and those previously mentioned we see some of the links that encouraged collaboration across disciplines. In the case of

biomedical radiation research, radiation and related technologies served as a key

link between the physical and life sciences, as they would when scientists

partnered with the government and military both during and after the war.

There is a robust literature on the mobilization of science for World War II and the continuation of government and military funding for science following the

1 o war. The central theme of the existing literature has been the impact of government and military funding on research and development, especially in academic settings. For instance, Stuart W. Leslie's pivotal study of Stanford

University and the Massachusetts Institute of Technology (MIT) during the Cold

War shows that postwar defense contracts resulted in a considerable focus on defense problems within university science and engineering laboratories.19

17 John Heilbron and Robert Seidel, Lawrence and His Laboratoiy: A History of the Lawrence Berkeley Laboratory (Berkeley, CA: University of California Press, 1990). See also Jeffrey E. Williams, "Donner Laboratory: The Birthplace of Nuclear Medicine," The Journal of Nuclear Medicine 40, no. 1 (1999): 16N, 18N, 20N. IS For instance, on the Manhattan Project, see Hewlett and Anderson, The New World; Hughes, The Manhattan Project; Richard Rhodes. The Making of the Atomic Bomb (New York: Simon and Schuster, 1986). For the history of the development of another important wartime defense project—radar—see Buderi, The Invention That Changed the World. On the government bureaucracy created to organize wartime research and development, see Goldberg, "Inventing a Climate of Opinion," 429-52; Irvin Stewart, Organizing Scientific Research for War: The Administrative History of the Office of Scientific Research and Development (Boston, MA: Little, Brown and Company, 1948). 1' Stuart W. Leslie, The Cold War and American Science: The Military-Industrial-Academic Complex at MIT and Stanford (New York: Columbia University Press, 1993). This is a topic also 18

Briefly mentioned above, Angela N. H. Creager, Timothy Lenoir and

Marguerite Hays, Michael A. Dennis, and Gerald Kutcher have also examined

aspects of the defense-oriented, government-funded research and development

enterprise that emerged during and was maintained after World War II. Lenoir

and Hays examine the extension of the Manhattan Project's biomedical research

following the war arguing that the Manhattan Project medical officers

successfully planned for ongoing research which would be pursued in the civilian

9n sector. Similarly, Creager's studies of the AEC's radioisotopes distribution

2 1 program illustrate the government's role in bolstering civilian research. While

Lenoir, Hays, and Creager raise important points regarding the impact of wartime

and postwar funding on civilian research, Dennis examines the postwar

reorganization of science as an endeavor that aimed to reconstitute the civilian—

99 that is, establish a divide between civilian and military science. Kutcher's study

of Cold War cancer and military research, on the other hand, argues that, in the

postwar period, divisions between civilian and military research are difficult to

make. The knowledge generated from military-funded research informed

research focused on so-called civilian problems such as cancer. Thus, no line

9 "2 could be drawn between bodies of knowledge.

explored in S. S. Schweber, "Big Science in Context: Cornell and MIT," in Big Science: The Growth of Large-Scale Research, ed. Peter Galison and Bruce Hevly (Stanford, CA: Stanford University Press, 1992), 149-83. 20 Lenoir and Hays, "The Manhattan Project for Biomedicine," 29-62. 21 Angela N. H. Creager, "The Industrialization of Radioisotopes by the U.S. Atomic Energy Commission," in The Science-Industry Nexus: History, Policy, Implications, Proceedings of Nobel Symposium 123, ed. Karl Grandin, Nina Wormbs, and Sven Widmalm (Sagamore Beach, MA: Science History Publications/USA, 2004), 141-67; Creager, "Nuclear Medicine in the Service of Biomedicine," 648-84. " Dennis, "Two University Laboratories in the Postwar American State," 427-55. :,Kutcher, "Cancer Therapy and Military Cold-War Research," 105-30; Kutcher, Cancer Research and the Military. 19

This study emphasizes the importance of studying the AEC as a system

and draws on Thomas P. Hughes' systems approach. Hughes provides a model

for studying all aspects of research and development—regardless of whether they are technical, scientific, social, economic, or political—as systems of interrelated elements.24 While the focus here is the AEC's biomedical program, it is

important to acknowledge that biomedical research was neither isolated from the

AEC's primary responsibility of managing the United States' nuclear weapons,

nor from the social and political context in which the AEC operated. As part of a

larger system, the AEC's biomedical program was both influenced by other parts of the system and helped constitute the character of the entire system. This speaks to the argument that the AEC's biomedical research helped contribute to the existence and character of the Cold War.

The focus on biomedical radiation research within a larger research and development enterprise, and the role of researchers in helping to establish research policies and programs, provides a different historical narrative than that presented in official histories of the AEC.25 Written in three volumes organized chronologically by the Manhattan Project era and the 1947-1952 and 1953-1961 periods of the AEC, these official histories primarily focus on the central structure and authority of the Commission. These works establish research as an important

"4 Sec, for instance, his classic work, Thomas P. Hughes, Networks of Power: Electrification in Western Society, 1880-1930 (Baltimore, MD: The Johns Hopkins University Press, 1983); Thomas P. Hughes, Rescuing Prometheus: Four Monumental Projects That Changed the Modern WorkI (New York: Vintage Books, 1998). Hewlett and Anderson, The New World; Richard G. Hewlett and Francis Duncan, Atomic- Shield: A History of the United States Atomic Energy Commission, Volume 2, 1947-1952 (University Park, PA: Pennsylvania State University Press, 1969); Richard G. Hewlett and Jack M. Holl, Atoms for Peace and War: Eisenhower and the Atomic Energy Commission, A History of the United States Atomic Energy Commission, Volume 3, 1953-1961 (Berkeley, CA: University of California Press, 1989). 20 aspcct of the AEC's work, but they are largely concerned with physical sciences research more directly linked to weapons development.

Throughout the twentieth century radiation safety has always been an important objective of biomedical radiation researchers. The historiography on radiation safety includes Gilbert Whittemore's study of the Advisory Committee on X-ray and Radium Protection,26 as well as Barton Hacker's two volumes on

17 radiation safety in the Manhattan Project and Atomic Energy Commission."

Collectively, these histories illustrate the shift from professional regulation to government regulation of radiation. They also show how the value of radiation safety and related research shifted as the social and political context changed during World War II and following. The history of radiation safety is not examined in depth here, other than to indicate the role radiation safety played in the larger political economy of research. Similarly, this dissertation does not engage much with the growing literature on the history of human radiation experiments funded by the AEC. While human radiation experiments are certainly relevant to the history of biomedical radiation research, the ethical questions that dominate the literature are beyond the scope of this project.

SOURCES & METHODOLOGY

Following World War II, this committee changed its name to the National Committee for Radiation Protection, and in the 1960s again changed its name to National Council on Radiation Protection and Measurements. Whittemore, "The National Committee on Radiation Protection, 1928-1960," 14. Hacker, The Dragon's Tail; Hacker, Elements of Controversy. :s On human radiation experiments, see Advisory Committee on Human Radiation Experiments, The Final Report (New York: Oxford University Press, 1996); Jordan Goodman, Anthony McElligott, and Lara Marks, eds., Useful Bodies: Humans in the Service of Medical Science in the Twentieth Century (Baltimore, MD: Johns Hopkins University Press, 2003); Jonathan D. Moreno, Undue Risk: Secret State Experiments on Humans (New York: W.H. Freeman, 2000). 21

In addition to the secondary literature noted above, this study is based on a range of primary sources, including scientific reports, government documents, and oral histories. For instance, scientific papers and oral histories were used to analyze interdisciplinary collaboration and the early twentieth-century development of new biomedical disciplines. Scientific reports produced by pioneering biomedical radiation researchers illuminated the sort of research problems that served as the epistemological link for interdisciplinary collaboration.29 From the personal narratives of scientists like Hymer L. Friedell we gain insight into why and how individuals sought to pursue specific fields of research, obtain access to new technologies, and associate with particular researchers and institutions, all in an effort to advance knowledge, but also position themselves within their profession.30

The professional connections and institutional affiliations researchers made before the war, combined with their research experience, were significant factors that determined whether or not researchers were recruited for wartime research. The personal narratives of researchers, again, provide a basis for

For instance, Robley D. Evans, "The Medical Uses of Atomic Energy," The Atlantic Monthly 177, no. 1 (1946): 68-73; Henry H. Janeway, Benjamin S. Barringer, and Gioacchino Failla, Radium Therapy in Cancer at the Memorial Hospital: First Report, 1915-16 (New York: 1917); R. A. Millikan, "The Significance of Radium," Science 54, no. 1383 (1921): 1-8. ,(1 The Advisory Committee on Human Radiation Experiments (ACHRE) conducted numerous interviews with radiation researchers involved in research in the early twentieth century through to the Cold War period. This oral history project resulted in transcripts of interviews with many of the leading radiation researchers all of which are available on the Department of Energy's website at "DOE Openness: Human Radiation Experiments," (accessed 25 March 2011). In addition to various ACHRE oral histories, the transcript of an interview with Physicist Lee A. DuBridge conducted by the American Institute of Physics provided insight on interdisciplinary collaboration and the role of new technologies in advancing research practices, fields of research, and academic institutions. Oral History of Lee DuBridge by Charles Weiner on June 9, 1972. Niels Bohr Library & Archives, American Institute of Physics, College Park, MD, USA, (accessed 31 July 2009), n.p.. 22 analyzing the pattern of wartime mobilization of biomedical radiation researchers.

Oral history transcripts, combined with reports produced by the Manhattan

Engineer District (MED)—specifically biomedical researchers within the MED— help depict researchers' roles within a new scientific enterprise.31

Chapters three through six are predominantly based on records of the AEC

held at the National Archives. The AEC's "Office of the Secretary

Correspondence Files," "Research Division Correspondence Files," "Research and Development Division Correspondence Files," and "Research and Medicine

Division Correspondence Files" are particularly useful for analyzing how the

AEC's biomedical policies and programs were created and who played a key role

in that process. These record series, which contain minutes from committee meetings, personal correspondence, technical reports, and financial data related to the AEC's biomedical radiation research, were produced by AEC staff and by

IT researchers from outside of the agency. Documents written by the Advisory

Committee on Biology and Medicine, for instance, or correspondence from scientists outside of the AEC show that the AEC's biomedical program, while a

"government" program, was very much influenced by scientists who were not directly involved with this federal agency. The relationship between the

" Researchers, such as Robert Stone and Karl Morgan who, as we will see in Chapter Two were key individuals within the Manhattan Project's health and safety effort, produced reports on the activities and organization of the health and safety program. Many of these are part of the AEC collection maintained at the National Archives Southeast Region, Atlanta, GA as part of Record Group 326. The material most relevant to biomedical research during the World War II period is found in the Record Series "New York Operations Office - Research & Medicine Division Correspondence, 1945-1952." v This study is based on AEC records (Record Group 326) maintained at the National Archives in College Park, Maryland and the Southeast Regional Archives in Atlanta, Georgia. The following Record Series contain the most relevant material: "General Correspondence 1946-1951," "General Correspondence 1951-1958," and "New York Operations Office - Research & Medicine Division Correspondence, 1945-1952." 23 biomedical community and government helped create a new political economy of research that both facilitated the expansion of biomedical radiation research and was a distinguishing component of the Cold War.

OUTLINE

This dissertation is comprised of six chapters that are organized chronologically. The first chapter, "X-Rays, Radium, and the Early Twentieth-

Century Development of Interdisciplinary Collaboration," examines the development of biomedical research involving X-rays, radium, and by the 1930s, artificially produced radioisotopes. This chapter especially highlights the practice of interdisciplinary collaboration which, by the outbreak of World War II, had resulted in professional connections amongst a small, but cohesive group of biomedical researchers who were committed to exploring uses of radiation in research and as a diagnostic and therapeutic tool. These researchers were united in an informal professional network by their shared interest in investigating possible uses of new radiation-related technologies in research and medical practice. As a result of interdisciplinary collaboration we see that numerous radiation researchers redefined their own research specializations and, in the process, created or helped established a foundation for new biomedical disciplines. The fluidity of boundaries between disciplines and between science and society is a reflection of researchers' awareness of and role in shaping the social and political context in which they pursued their research.

The trend of interdisciplinary radiation research continued throughout

World War II and beyond. Thus, the second chapter, "Biomedicine and Bombs: 24

Health and Safety in the Manhattan Project," examines the continuity between

pre-war development of radiation as a research and medical tool, and wartime

research on radiation hazards. The work of biomedical researchers in the

Manhattan Project shows that although wartime goals shifted to health and safety,

there was continuity in basic research, as well as in the recruitment and practices of researchers. The researchers recruited to investigate radiation hazards and attend to radiation safety were primarily those who were already accustomed to

working collaboratively with variously trained researchers and engineers. They

were researchers who recognized the social and political value of their work and who, during the war, learned to value their work within a government and military system and relative to national security goals.

Chapters three through six cover the early postwar period—a period in which the AEC and its biomedical program were created. Collectively, these four chapters examine various elements of a postwar scientific enterprise to illuminate the complex relationships that tied science and the state together and that helped create and shape Cold War culture. The third chapter, "From War to Peace:

Institutionalizing Biomedical Radiation Research within the Atomic Energy

Commission," examines the creation of a biomedical bureaucracy within the

AEC. It especially focuses on the role of researchers in influencing the creation of the Advisory Committee on Biology and Medicine (ACBM) and Division of

Biology and Medicine (DBM). The ACBM and DBM were crucial to researchers' ability to shape the future development of biomedical radiation research. They were crucial to researchers' ability to establish biomedical 25 research as a priority not just within the AEC, but within the nation. The

American commitment to biomedical radiation research as one part of a larger scientific, technological, social, and political enterprise was an attribute of a nation preparing for and engaging in a "cold war."

The fourth chapter, "The Building Blocks of Research: Radioisotopes

Distribution," examines the creation and early development of the AEC's radioisotopes distribution program. As was the case with the ACBM and DBM, researchers both internal and external to the AEC played an important role in establishing the radioisotopes distribution program. This program provided tools for research which highlighted the need for research training and facilities, and provided resources that could be directed toward cancer research. Aside from driving other aspects of the AEC's biomedical research program, radioisotopes served the fundamental purpose of linking science and the state throughout the

Cold War. Both researchers and the state had considerable interest in radioisotopes and, thus, worked together to develop a distribution program.

Chapter five, "The ABC's of the AEC's Biomedical Research:

Fellowships and Educational Initiatives," focuses on the AEC's education and training initiatives which aimed to expand biomedical infrastructure and the number of researchers capable of working with radiation. This chapter examines the piecemeal fashion in which education and training initiatives emerged. Like radioisotopes, education and training provided common ground upon which many individuals and institutions both within and without the AEC merged scientific, social, and political goals. The education and training initiatives of the AEC 26

reveal that science and defense cultures clashed, despite the willingness of many

to cooperate in an effort to achieve common goals. These initiatives also highlight the importance of the ACBM and DBM as organizations within which science and government merged.

The final chapter, "Waging War on Another Front: The AEC Joins the

Cancer Establishment," examines the AEC's effort to advance cancer research.

Cancer was a cause that helped bridge the interests of biomedical researchers, the government, and military. Since the early decades of the century biomedical radiation researchers had pursued research with the hope that radiation might be used to increase knowledge about cancer or serve as a diagnostic and therapeutic tool. The AEC needed little encouragement to commit resources to cancer research since there was sufficient research to show that radiation was both a cause and cure of cancer. The AEC did, however, struggle to determine what sort of contribution it would make to cancer research. Rather than establishing cancer research as the primary goal of the AEC's biomedical agenda, the AEC created a cancer program that was part of a larger biomedical program. The AEC was able to maintain broad biomedical goals by focusing on basic research and providing tools and infrastructure to support cancer research. Thus, the cancer program reinforced and was, in turn, reinforced by the AEC's radioisotopes distribution program and education and training initiatives. Each of these components of the

AEC's biomedical program helped link science and the state within the new and complex research and development enterprise of the AEC. 27

CHAPTER ONE

X-RAYS, RADIUM, AND THE EARLY TWENTIETH-CENTURY DEVELOPMENT OF INTERDISCIPLINARY COLLABORATION

In 1946, physicist Robley D. Evans published an article in the popular magazine The Atlantic Monthly in which he wrote about the medical uses of atomic energy.33 "So far," Evans wrote, "the most fruitful applications of these new techniques have been the results of team work between physicists and other scientists."34 It seemed to Evans—a pioneering biomedical radiation researcher— that interdisciplinary research had been key to his own and his colleagues' success in developing medical applications of radiation. However, this approach to radiation research began decades earlier.

Throughout the early decades of the twentieth century variously trained researchers collaborated to investigate radiation. Interdisciplinary collaboration was central to the establishment of new kinds of biomedical expertise that were hybrid in nature or that combined the knowledge and skills of existing disciplines in both the physical and life sciences. Such hybrid expertise formed the foundation of new biomedical disciplines. Interdisciplinary collaboration was also central to the formation of relationships between researchers and institutions.

How did the collaborative relationships formed between individual researchers and institutions in the early twentieth century affect those individuals and institutions? Upon what were these relationships based and how did they shape the practice of research and the development of disciplines? This chapter

" Evans, "The Medical Uses of Atomic Energy," 68-73. 14 Ibid.: 73. 28 examines interdisciplinary collaboration in relation to the creation of a network of biomedical radiation researchers from approximately 1900 to 1940. It argues that the formation of professional and institutional connections provided researchers with scientific and social authority that helped them advance research goals within the political economy of early twentieth-century research, as well as influence the nature of that political economy. Interdisciplinary groups of researchers obtained materials and resources they deemed necessary for their research by responding to social concerns like cancer. However, they also generated increasing demand for biomedical radiation research throughout the early decades of the century by defining their research as a solution to the cancer problem. The relationship they established between cancer and radiation research was fundamental to the development of their research programs and, beyond that, the future development of new biomedical disciplines.

Interdisciplinary collaboration built upon late nineteenth-century developments in physics, including the discovery of X-rays and radioactive elements. As physicists explored the properties of radioactivity, physician- researchers and scientists with medical interests—those who today would be referred to as biomedical researchers—were quick to investigate its diagnostic and therapeutic benefits. The use of radiation in medical research and practice, and in various industries, encouraged collaboration amongst variously trained scientists and physician-researchers to gain a better understanding of both the medical benefits and adverse side effects of radiation. The early twentieth-century development of biomedical radiation research was significant, in part, because this research established a broad framework for investigating radiation in relation to medical applications and safety hazards, both of which gained increasing importance during the World War II and beyond.

Aside from establishing medical and safety agendas, early twentieth-century radiation research resulted in the creation of, or at least some of the building blocks for, the development of new biomedical disciplines. Furthermore, this research resulted in professional connections that shaped the wartime mobilization of biomedical radiation researchers. This chapter argues that interdisciplinary collaboration and the formation of a small, but ambitious network of biomedical radiation researchers were defining features of the early investigation of radiation—features that would have a lasting influence on the development of biomedical radiation research as social, political, and economic circumstances changed over time.

X-RAYS, RADIUM & THE ORIGINS OF RADIOLOGY, NUCLEAR MEDICINE, & BIOMEDICAL RESEARCH

The medical use of radiation dates back to 1895 when German physicist

Wilhelm Conrad Roentgen discovered X-rays. These invisible rays are able to penetrate matter such as human flesh and create an image of denser matter such as bone. This was a momentous occasion in the history of physics and one that also created new possibilities for medical research and clinical practice. To turn-of- the-century medical researchers and clinicians, the use of X-rays as a diagnostic tool looked promising. The excitement that X-rays generated was not short-lived.

Almost a century after this discovery, radiologist and former United States Army 30

Air Corps captain, Frederick J. Bonte, reflected on his own fascination with the

medical use of X-rays during an interview conducted in the 1990s for The Journal of Nuclear Medicine. He recalled that when he began his career as a radiologist in the 1940s, X-rays had already "completely altered the course of medicine" in that they allowed physicians "to make diagnoses that couldn't be made any other way."35 Bonte's enthusiasm speaks to the early recognition of and lasting impact of diagnostic uses of X-rays. Beyond this purpose, researchers and clinicians quickly decided that X-rays also had considerable potential as a therapeutic technology. Ironically, this realization came as researchers and clinicians noted the occurrence of unfortunate side effects and even mortalities caused by exposure to X-rays. They sought to harness the destructive nature of X-rays to destroy tissues, especially cancerous tumors.36

The pursuit of this goal was a process throughout which the discovery of

X-rays acquired meaning in relation to a changing political economy of scientific research. The argument made here, that biomedical radiation research and clinical applications of radiation were part of a process of discovery, stems from historian Thomas S. Kuhn's study of the structure of scientific discoveries—work that is well illustrated in Helge Kragh's and Robert W. Smith's examination of the discovery of the expanding universe.37 In their study they contend that discovery is a process that is complex, often "messy," takes place over time, and may

N.A., "Interview of Frederick J. Bonte in 'People in Nuclear Medicine: Interviews with Physicians, Scientists and Industry Leaders,'" The Journal of Nuclear Medicine 34, no. 6 (1993): 20N. For documentation of early radiation injuries, see Robert S. Stone, "The Concept of a Maximum Permissible Exposure," Radiology 51, no. 5 (1952): 640. " Thomas S. Kuhn, "Historical Structure of Scientific Discovery," Science 136, no. 3518 (1962): 76-764; Helge Krah and Robert W. Smith, "Who Discovered the Expanding Universe?," Historv of Science 41 (2003): 141-62. 31

involve many actors. Further, they argue that "What follows after the

announcement of a discovery is often as important as, and sometimes even more

-> o important than, what leads up to it." Like Kragh's and Smith's study, this dissertation considers the discovery of radiation as more than Roentgen's

identification of X-rays—that is, as more than a singular event.

The investigation and application of X-rays became part of a larger pursuit to study and develop applications for other sources of radiation. Indeed, just one year after Roentgen's discovery of X-rays, French physicist Henri Becquerel identified uranium as the first radioactive element, or an element found to emit invisible rays similar to X-rays. More radioactive elements were identified soon thereafter. For instance, French physicists Marie and Pierre Curie isolated polonium and radium in 1898 and Marie coined the term "radioactive" to describe the energy emission from these and other elements.

Like X-rays, radium proved to be a significant discovery in the history of nuclear medicine. This naturally occurring radioactive substance was thoroughly exploited for its medical uses in the early decades of the twentieth century.39

Radioactivity was further defined when Ernest Rutherford categorized radioactivity as being of three kinds of rays: alpha, beta, and gamma. In the decades that followed physicists and medical researchers investigated these forms of radiation and found that they had different uses and dangers.40

,s Smith, "Who Discovered the Expanding Universe?," 142. " In recognition of the Curies' contributions to understanding radioactivity, the "curie" was adopted as the unit of radioactivity in 1912. Clark, Radium Girls, 40-1; ACHRE, The Final Report, 3; Hacker, The Dragon's Tail, 19. 411 X-rays and gamma rays are similar in their ability to penetrate matter and, thus can affect most tissues in the body. Alpha and beta rays do not have the same ability to penetrate matter. However, they consist of high-speed particles that are relatively large in size which allows them to 32

Around the turn of the twentieth century, these developments in physics helped usher in a new era of research and clinical use of radiation in medicine.

As physician-researcher Dr. John Gofman recounted in an interview in 1994,

"Now in the very earliest years after Roentgen's discovery of the X-ray and

Curie's discovery of radium, both got into medicine very quickly. It looked promising. . . ."4I Gofman, both an M.D. and Ph.D., began his career as a medical physics researcher at the University of California, Berkeley, during World War II.

He spent much of his career there and at the Lawrence Livermore National

Laboratory. During his career he did an extensive review of early twentieth- century therapeutic uses of X-rays and radium. It seemed to Gofman that early twentieth-century physicians considered the potential medical uses of radiation to be great and were eager to use both X-rays and radium to treat any or every disease. Historian Claudia Clark drew the same conclusion arguing that, as early as 1903, X-rays and radium were being tested for the treatment of all ailments.42

At that time, physicians understood the basic physical properties of radiation, but not how it affected the body.43 It was for this reason that Gofman characterized the early twentieth century as the real era of human radiation experimentation.

Gofman made this evaluation in the mid-1990s in relation to an ongoing

causc considerable damage to matter in close proximity. If alpha and beta particles are ingested or inhaled into the body they can cause more severe damage than X-rays and gamma rays. For more information on these various types of radiation, see Clark, Radium Girls, 87; Janeway, Barringer, and Failla, Radium Therapy in Cancer, 20, 27-9. 41 Oral History of John W. Gofman, M.D. Ph. D., interview conducted December 20, 1994, as part of the Office of Human Radiation Experiments' Oral History Project, Human Radiation Studies: Remembering the Early Years. Conducted by Loretta Hefner and Karoline Gourley in San Francisco, C'A, n.p.. 42 Clark, Radium Girls, 43. 4' Charles Hayter, An Element of Hope: Radium and the Response to Cancer in Canada, 1900- 1941) (Montreal, QC: McGill-Queen's University Press, 2005), 21-22. 33

investigation of government involvement in human radiation experimentation

during World War II and following. Although Gofrnan assessed the widespread

use of X-rays and radium in the early decades of the twentieth century as being

experimental, he did insist that it "was therapy."44

Experimental medical practice was not limited to the use of radiation in

the early twentieth century. This was a time at which the American medical

profession was undergoing a transformation. Starting in the late nineteenth

century, medical advancements increasingly rooted in laboratory research rather

than clinical experience. As such, medical schools revised their curriculum to

incorporate more training in laboratory research. Historian Susan Lederer examined this transformation and the human experimentation conducted during

the late nineteenth and early twentieth centuries. She argues that while formal ethical guidelines to govern human experimentation did not exist in the form of federal legislation until the second half of the twentieth century, researchers were not entirely free to disregard ethical issues when pursuing human experimentation. As she explains, clinicians and researchers followed basic ethical standards which were informally upheld due to pressure coming from both within and without the medical profession.45 It is in light of this culture of self- regulation that Gofman qualified early twentieth-century radiation experimentation as therapeutic. This was a distinction that became more difficult to make during and following World War II when national security and medical objectives were often intertwined.

44 Oral History of John Gofman, 1994, n.p.. 45 Susan E. Lederer, Subjected to Science: Human Experimentation in America before the Second World War (Baltimore, MD: Johns Hopkins University Press, 1995), xv-xvi, 26, 72, and 138. 34

The point here is not to determine whether or not early medical

applications of radiation were experimental, therapeutic, or both. Rather, here I

seek to show that medical applications were a key factor in the development of a

biomedical radiation research community. With physicians eager to apply

radiation in medicine, but with little knowledge of the physiological effects of

radiation and few constraints on their practice, physicians' pursuit of clinical

applications of radiation provided motivation for biomedical radiation research.

This was one factor that shaped the political economy of research such that it

encouraged interdisciplinary research that crossed boundaries between the

physical and life sciences and between scientific research and medical practice.

RADIUM: SECURING A SUPPLY

Prior to World War II biomedical researchers around the world were keen

to explore the therapeutic use of radiation for a whole range of ailments, but

experimental work with radium, in particular, was limited by supply. Radium is

not an element that occurs in nature in great abundance and the cost of mining ore

from which radium could be extracted was extremely high. For American

physicians it was particularly difficult to obtain radium. From the time it was

discovered in 1898 until the early 1910s Europe had a monopoly on radium

production.46 At that time it sold for as much as $ 180,000/gram, an amount that

4<' During the first decade of the twentieth century Americans had to buy their radium from refineries in Paris, Vienna, and Germany. Some Americans, interested in breaking the European monopoly, suspected that radium could be produced at a lower cost. As we will see, various individuals and groups helped establish American sources of radium. Roger Robison, "American Radium Engenders Telecurie Therapy During World War I," Medical Physics 27, no. 6 (2000): 1212 and 15. today would be the equivalent of approximately $4,113,000/gram.47 Radium was a sought-after commodity, not only within the medical community, but also within industry. Just a few years after radium was discovered it was found to have fluorescent properties. This opened a market for producing radioluminous items such as watches. To meet both the industrial and medical demand,

American radium mining companies started to form.

A leading supplier in the United States was the Radium Chemical

Company, Inc., which formed as a subsidiary of the Standard Chemical Company.

The Radium Chemical Company produced the United States' first commercial radium in 1912 and a decade later, with the production of 18 gram/year, was able to meet the American demand for radium.48 This company did more than produce radium. With a vested interest in expanding the radium market, the Standard

Chemical Company established a biological laboratory to conduct research on the medical uses of radiation. The results of these investigations were published in the journal Radium, a journal established by the company in 1913.49

47 This conversion is based on the United States Department of Labor's Consumer Price Index Inflation Calculator. 48 R.E. Rowland, Radium in Humans: A Review of U.S. Studies (Argonne, IL: Argonne National Laboratory, 1994), 3. 49 As stated on the cover of this journal, Radium was "A monthly journal devoted to the chemistry, physics, and therapeutics of radium and radio-active substances." See Ibid., 4-5. The creation of a biological laboratory and the publication of Radium were not unlike an arrangement modeled in Paris, one of the few European centers that produced radium in the first decade of the twentieth century. A French refinery, Armet De Lisle, subsidized a radium clinic and from 1904-1914 published a free journal called Le Radium. Robison, "Telecurie Therapy," 1212-3. This was also a time at which U.S. companies were increasingly establishing laboratories to facilitate discoveries and inventions that would hopefully translate into increased profit. The first industrial laboratory established in the U.S. was created by General Electric in 1900. Du Pont and AT&T quickly followed suit opening labs of their own in 1902 and 1911 respectively. For more on the rise of industrial research labs in the U.S., see Robert Buderi, Engines of Tomorrow: How the World's Best Companies Are Using Their Research Labs to Win the Future (New York: Simon & Schuster, 2000); David A. Hounshell and Jr. John Kenly Smith, Science and Corporate Strategy: Du Pont R&D, 1902-1980 (Cambridge, UK: Cambridge University Press, 1988); Leonard S. Reich, The Making of American Industrial Research: Science and Business at GE and Belt, 1876- The medical community hardly needed any encouragement from radium

producers to exploit radium. Two physicians in particular—Drs. Howard Kelly

and James Douglas—were very much involved in securing a domestic supply of

radium. Kelly was a leading gynecological surgeon at Johns Hopkins University

who also maintained a private hospital in Philadelphia. He wanted to develop an

American radium market because his research on experimental radiation therapy

was limited by the amount of radium available to him.50

Douglas was a trained physician, but he never practiced medicine. Rather,

he became an expert in copper mining. His desire to secure a domestic radium

supply stemmed from his experience of having to purchase radium from Paris

when his daughter underwent treatments for breast cancer between 1907 and

1910. With both professional and personal motives, Kelly and Douglas

convinced the United States Congress to stimulate production of radium in the

United States. They negotiated an agreement with the federal government that

committed the Federal Bureau of Mines to mine ore that would then be delivered

to a refinery set-up and paid for by Kelly and Douglas. This operation was

established in 1913 and incorporated as the National Radium Institute.51 Thus, by

the 1910s, the supply of radium in the United States and physicians' and

researchers' access to it had become more stable. The sheer expense of radium

was still an obstacle to research and the application of radium in medicine, but

1926 (Cambridge, UK: Cambridge University Press, 1985); George Wise, Willis R. Whitney, General Electric, and the Origins of U.S. Industrial Research (New York: Columbia University Press, 1985). 50 Robison, "Telecurie Therapy," 1213. 51 The National Radium Institute created unwanted competition for private radium companies, but the operation existed only temporarily. Douglas and Kelly sold their refinery to the Radium Chemical Company in 1917. Ibid:. 1213 and 1215, 37 due to the initiatives of private companies, individual physicians and researchers, and the federal government, the obstacle was no longer as great as it once was.

RADIUM THERAPY & INTERDISCIPLINARY RESEARCH AT MEMORIAL HOSPITAL—THE MAKING OF NEW DISCIPLINES

As physicians around the country started to use radium to treat a variety of ailments, especially cancer, they sought to determine the most effective practices for radium therapy. A report published in 1917 by a leading surgeon at Memorial

Hospital in New York, the first hospital in the United States specifically devoted to cancer, attempted to do just that.52 Dr. Henry H. Janeway, along with fellow physician Dr. Benjamin S. Barringer, and physicist Gioacchino Failla, published a comprehensive report on the use of radium in cancer therapy.53 As lead author on the report, Janeway thought that the available reports describing the use of radium to treat cancer lacked the specific information practitioners needed to reproduce successful treatment practices. He deemed it to be very important to standardize the use of radium to treat malignant tumors of different kinds and located in various organs. Basing their report on the treatment of 424 malignant tumors with radium at Memorial Hospital in 1915 and 1916, Janeway explained that "it has become more and more apparent that the successful use of radium in cancer,

52 Memorial Hospital was founded in 1884 as the New York Cancer Hospital. In 1898 the hospital abandoned its specific focus on cancer due to the relative lack of cures for cancer. The hospital was renamed as the General Memorial Hospital. However, with the discovery of X-rays and radium, cancer treatment looked more promising and the hospital reestablished itself as a cancer treatment facility. At that point the hospital adopted the name of Memorial Hospital for the Study of Cancer and Allied Diseases and it is now known as the Memorial Sloan-Kettering Cancer Center. See "History & Overview," Memorial Sloan-Kettering Cancer Center, (accessed 24 June 2009); and Robison, "Telccurie Therapy," 1214. 53 The radium produced by the National Radium Institute was divided between Dr. Kelly who used his portion for his own work in radiation therapy and Dr. Douglas who gave his share to the Memorial Hospital. This supply of radium allowed for Dr. Janeway and his colleagues to pursue the work upon which this report was based. Robison, "Telecurie Therapy," 1215. 38

requires careful consideration of each particular type of the disease in each organ

as separate problems in which different methods must be devised and different

results expected."54 At the time, the hospital was just acquiring sufficient radium

to use for radium therapy and the growing supply at that hospital and elsewhere

allowed for the greater use of radium therapy in medical practice.

Although not examined in the historical literature, Janeway's and

Barringer's documentation of their clinical work with radium is an excellent

illustration of how physicians and scientists tried to improve the medical uses of

radiation. Their report served as a means to help orient practitioners to a more standardized way of interpreting and treating different forms of cancer and, thus,

helped to establish standards in radiation therapy. Failla's contribution of a

comprehensive introduction that explained the physical properties and biological

effects of radiation and radium was equally important to the effort to standardize

medical practice. Failla had been employed as a physicist at Memorial Hospital since 1915 when Janeway offered him part-time employment. Failla's

involvement in this project and other collaborative work with physicians reflected an ideal of scientific medicine—an ideal that the development of clinical practice should be built upon a foundation of basic research in the physical and biological sciences. It also reflected the aspirations of these three individuals to establish

practices that were not limited by their own personal knowledge and skills.

From this collaborative experience Failla became convinced that

investigating the medical applications of radiation was a worthy pursuit.

Furthermore, he believed that such research should encompass more than just the

M Janeway, Barringer, and Failla, Radium Therapy in Cancer, Preface and 46. expertise that he, as a junior physicist who was also trained in electrical engineering, could offer.55 When Janeway and Barringer recruited Failla to collaborate on the radium report, Failla was still pursuing his studies in physics.

He completed his Masters in Physics in 1917—the year the report was published—and then took a two-year leave of absence from his position at

Memorial Hospital to serve during World War I. Following the war, he went to

Paris to pursue a year of doctoral work with Marie Curie at the Sorbonne/6

Upon the completion of his Ph.D. in 1923 Failla returned to Memorial

Hospital where he played an important role in expanding interdisciplinary collaboration and facilitating the development of hybrid expertise.57 He established his own radiation research laboratory at the Hospital in which he employed a mix of physicists, chemists, and biologists and continued to work closely with the hospital's physicians. Due to the nature of his laboratory research, his colleagues considered him a pioneer in biophysics, specifically radiation biophysics. At the time, biophysics was an emerging discipline that applied the principles and practices of physics and chemistry to the study of biological processes or functions.

55 Edith H. Quimby, "Gioacchino Failla (1891-1961) and the Development of Radiation Biophysics," The Journal of Nuclear Medicine 6, no. 5 (1965): 377-8. M' Hymcr L. Friedell, L. D. Marinelli, and Titus C. Evans, "Gioacchino Failla, 1891-1961," Radiation Research 16, no. 5 (1962): 620. 1,7 His research continued to focus on radium therapy throughout the early 1920s. See, for instance, Gioacchino Failla, "The Absorption of Radium Radiation by Tissues," American Journal of Roentgenology 8, no. 5 (1920); Gioacchino Failla, "Dosage in Radium Therapy," American Journal of Roentgenology 8, no. 11 (1921): 674-85; Gioacchino Failla, "Ionization Measurements," American Journal of Roentgenology and Radium Therapy 10 (1923): 48-56; Gioacchino Failla and Edith H. Quimby, "The Economics of Dosimetry in Radiotherapy," American Journal of Roentgenology and Radium Therapy 10 (1923): 944-67. One of his earliest and longest standing colleagues, Edith Quimby,

captured the interdisciplinary nature of Failla's research in her attempt to offer a

definition of this field. In an article she wrote in the 1960s to establish Failla's

pioneering role in the history of radiation biophysics, Quimby argued, "A satisfactory definition of biophysics is hard to formulate, but it certainly has to do

with the use of physical tools to arrive at biological knowledge."' Failla

attempted to link physics, biology, and medicine and, in doing so, helped create

biophysics as a new field of research. He also made important contributions to

the early development of radiobiology, a field of research that developed as the study of the biological effects of radiation and that, like biophysics, was just

emerging during the 1920s.59

Failla's commitment to crossing disciplinary boundaries was evident by

the late 1920s. By then he thought it was not enough to employ biologists in his

laboratory; he also required that his physicists develop a good understanding of

biology as well.60 The interdisciplinary collaboration that he explicitly sought to

facilitate was integral to the development of X-ray and radium therapy, and would

continue to be integral to the expansion of radiation research and therapy in the

1930s after the development of the cyclotron. As Chapter Five will illustrate, ongoing interdisciplinary radiation research resulted in infrastructural changes within universities to support new fields such as biophysics, radiobiology, health

physics, and medical physics. Although some universities started to develop

5,1 Quimby, "Gioacchino Failla," 377. 54 Angela N. H. Creager and Maria Jesus Santesmases, "Radiobiology in the Atomic Age: Changing Research Practices and Policies in Comparative Perspective," .Journal of the History of Biology 39 (2006): 637; Friedell, Marinelli, and Evans, "Gioacchino Failla, 1891 -1961620-21. Quimby, "Gioacchino Failla," 378. 41

interdisciplinary radiation research programs throughout the 1930s, it was not

until after the war that these programs were formally institutionalized.

GIVING MEANING TO DISCOVERY: THE HAZARDS OF RADIATION

Aside from determining how best to apply radiation within medical practice, radiation safety concerns also helped drive interdisciplinary collaboration. With growing access to radium and X-ray equipment, both researchers and clinicians were mindful of the side effects of radiation. They generally considered commonly noted side effects like reddened skin (erythema) and hair loss (epilation) to be minor. However, worse side effects also occurred.

Researchers and clinicians recognized that radiation was not only a treatment for cancer, but also a cause of cancer. Furthermore, they determined that radiation could cause sterility and induce changes in the blood or damage blood-forming organs.61 The adverse side effects were not well understood in the early decades of the twentieth century, but did provide sufficient motivation for researchers and clinicians such as Janeway, Barringer, and Failla to seek a better understanding of radiation's hazards and clinical uses. Their goal was to understand how to achieve the desired results of shrinking or destroying cancerous tumors while preventing damage to healthy tissues in the process. Here, in fact, was a key issue: How were they to exploit the medical benefits of radiation while avoiding or minimizing the hazards of radiation? As is evident from this study, researchers and clinicians struggled with this problem throughout the century.

1,1 Hacker, The Dragon's Tail, 10; Daniel Paul Serwer, "The Rise of Radiation Protection: Sciencc, Technology and Medicine in Society, 1896-1935," (Ph.D. diss., Princeton University, Princeton, NJ, 1976), vii. 42

The federal government, historians, and journalists have given much attention to radiation safety and human radiation experiments during World War

II and the Cold War. Specifically, they have investigated the role of the federal government in radiation research that involved humans.62 The focus placed upon

World War II and after may lead to the assumption that experimentation with radiation and concern for radiation safety only began during the war. The focus of radiation research did evolve and expand during World War II and the Cold

War, but, as Dr. John Gofman implied when interviewed in the 1990s and as illustrated here in this overview, human radiation experimentation really began at the start of the century.63 So, too, did concern for radiation safety, mostly as a result of illness and death that occurred amongst researchers and clinicians who were exposed to radiation. Concern for radiation hazards or fear of radiation in the early twentieth century was, however, far less pervasive than it would be in the post-World War II era. Prior to the creation of atomic weapons, radiation generated more hope than fear amongst both the scientific/medical community and the general public.64 Not surprisingly then, the researchers investigating the medical uses of radiation and its biological effects, and the clinicians using radiation in medical practice, had fewer reservations about experimenting with radiation therapy in the early decades than they would later in the century.

A greater appreciation for the hazards of radiation developed in the 1920s, largely due to investigations of illnesses and fatalities amongst women in the

For instance, see ACHRE, The Final Report; Hacker, The Dragon 's Tail; Hacker, Elements of Controversy, Moreno, Undue Risk; Eileen Welsome, The Files: America's Secret Medical Experiments in the Cold War (New York: Dial Press, 1999). w Oral History of John Gofman, 1994, n.p.. 64 Clark, Radium Girls, 39. 43 dialpainting industry. Dialpainting, which was the practice of painting radium on watch faces to make them glow, began in 1917. Until World War II the industry employed about 4000 women, many of whom developed radiation-related illnesses due to the ingestion and inhalation of radium. Cancer of the bone was especially common because radium concentrates in bone rather than being excreted from the body. This tendency was not known at the time, but by the early 1920s it had become apparent to many physicians and government that the illnesses and deaths occurring amongst dialpainters were caused by radium poisoning. Historian Claudia Clark shows in her study of the dialpainting industry and industrial health reform that industry leaders were not as quick to acknowledge the connection between radiation exposure and illness.65 Clark's focus on the dialpainting industry as an arena from which an industrial health movement emerged, illustrates the importance of adverse effects of human exposure to radium as a means of advancing knowledge of the physiological effects of radiation.

The tragic history of many dialpainters did help advance science in that the women who developed illnesses and those who died created a group of subjects that could be studied to better determine the effects of radiation on human tissue. Researchers at the time acknowledged the very unfortunate health consequences that dialpainters suffered, but they also sought to take advantage of the situation to learn more about the effects of radiation—radium specifically—to better protect people from similar hazardous exposures. This was important research considering that the use of radium as an internal medicine continued.

65 Ibid., 86. 44

Throughout the 1920s, for instance, radium water was marketed to the public as a

medicine with great benefits. It was not until a prominent businessman developed

radium poisoning and died in 1932 due to the consumption of radium water that the popular health tonic fell out of favor.66

RADIATION SAFETY

Throughout these early decades of the century, as physicists studied the

nature of radiation and clinicians made use of radiation in medical practice, individual clinicians and researchers were their own judge of radiation safety.

Practitioners made an effort to determine radiation doses that were therapeutic, but not overly destructive. One significant challenge was to determine how best to measure dose. Variously trained scientists, physicians, and engineers or technicians not only had to devise appropriate technologies for making radiation measurements, but also had to determine how to define dose. With different professional backgrounds, they even debated whether to express dose in units common to physics or biology.67

Such issues were considered within a new institutional framework as of

1928 when the International Committee on X-Ray and Radium Protection

(ICXRP) was established. This committee was formed at the second International

Congress of Radiology held in Stockholm, Sweden.68 Lauriston Taylor, a young

M' Rowland, Radium in Humans, 7. 1,7 The roentgen was adopted in 1931 as the unit for measuring air exposure for X-rays. See Hacker, The Dragon's Tail, 11; D.C. Kocher, "Perspective on the Historical Development of Radiation Standards," Health Physics 61, no. 4 (1994): 520; Whittemore, "The National Committee on Radiation Protection, 1928-1960," 55-6. ',s The First International Congress of Radiology, held in London, England, emerged out of cooperation between British physicians and physicists who collaborated to establish radiology as a 45 physicist employed by the National Bureau of Standards represented the United

States on this international body and also took on the task of establishing a national committee—the United States Advisory Committee on X-Ray and

Radium Protection (ACXRP).69 The ACXRP was formed with equal representation of physicists, physicians, and representatives from X-ray and radium equipment manufacturers. Historian Gilbert Whittemore, whose doctoral dissertation is the best account of the history of the ACXRP, offers a detailed discussion of the Committee's membership. According to Whittemore, Taylor and others who helped him establish the ACXRP deemed interdisciplinary representation important since physicists, physicians, and equipment manufacturers were all equally at risk for radiation exposure. They also thought impartial representation was necessary for professional societies because a rivalry existed amongst the leading radiological societies formed during the first two decades of the twentieth century. These societies included the American

Roentgen Ray Society (ARRS), the Radiological Society of North America

(RSNA), and the American Radium Society (ARS).70 The careful selection of members for the ACXRP, therefore, helped facilitate interdisciplinary collaboration. As we have seen in the work at Memorial Hospital,

specialist field of medicine and sought to establish international standards of radiation measurements and protection. See Serwer, "The Rise of Radiation Protection," 214-15 and 31-32. m Both the international and American committees are now known by different names. In 1950 the international committee changed its name to the International Commission on Radiological Protection (1CRP). The U.S. committee had disbanded during WWII and in 1946 reconvened as the National Committee on Radiation Protection (NCRP). As of 1964 the NCRP renamed itself as the National Council on Radiation Protection and Measurements but continues to use the same acronym. Whittemore, "The National Committee on Radiation Protection, 1928-1960," 14. 711 Regarding the recruitment of members to the ACXRP and discussion of the rivalry amongst radiological societies, see Serwer, "The Rise of Radiation Protection," 232; Whittemore, "The National Committee on Radiation Protection, 1928-1960," 18-24. interdisciplinary collaboration was already common, but it was a trend that would become even more integral to the pursuit of radiation safety and the building of disciplines and development of clinical practices throughout the century.

Initially the ACXRP planned to make detailed recommendations that professionals could follow voluntarily. The Committee would draw up reports and then disband. In 1931 and 1932 the ACXRP did produce two reports, the first made general recommendations for working with X-rays and the second for working with radium. These recommendations were made in the spirit of occupational safety. They were meant to limit radiation exposure for physicians, nurses, and technicians, not for patients. The protection of patients, the members of the ACXRP believed, should be left to a physician's discretion.71 After the publication of these reports, the Committee did not disband as was initially planned. Throughout the early 1930s the ACXRP devised the concept of a

"tolerance dose" for external exposure to X-rays and gamma rays.72 They sought to quantify the amount of radiation to which a person could safely be exposed. To define a tolerance dose, the ACXRP drew on the experience of its own members, as well as the work of other radiation experts to devise its standards. The

Committee was comprised of a small, but expert group. For instance, one member was Gioacchino Failla who, as already mentioned, was a pioneering biophysicist and radiobiologist with well over a decade of research experience

7i Whittemore, "The National Committee on Radiation Protection, 1928-1960," 526-8. 7" In 1933 the NCRP adopted a tolerance dose of 0.1 roentgen/day. This standard was generally accepted by the members of the committee before 1933, even though it had not been formalized. It was not published until 1936. Due to the ambiguity surrounding the formal adoption of this tolerance dose, there are discrepancies in the historical literature that documents this occasion. Kocher, "Perspective on the Historical Development of Radiation Standards," 520; Stone, "The Concept of a Maximum Permissible Exposure," 641; Whittemore, "The National Committee on Radiation Protection, 1928-1960," 87. 47 and a body of knowledge shaped by his collaboration with variously trained physicians and scientists at Memorial Hospital.73 The Committee's members stayed abreast of the most recent research, such as Herman Muller's irradiation of fruit flies in the 1920s for the study of genetic injury and Robley D. Evans' investigation of radium ingestion in rats conducted during the mid-1930s.74

Animal research was useful but, as researchers believed, not an entirely accurate means for devising human standards. Other than studying the dialpainters or individuals accidently exposed to radiation, the ACXRP was limited by its inability to obtain sufficient human data. Hence, the certainty ascribed to the

ACXRP standards was, as Whittemore argues, an illusion.75 Due to this lack of certainty the term "tolerance dose" was abandoned in the early 1940s—the

Committee was not comfortable with the implication that any amount of radiation could be safely tolerated. Instead the Committee defined its standard for external radiation as "maximum permissible exposure" and for internal emitters like ingested radium as "maximum permissible body burden."76

71 The NCRP was originally comprised of: Lauriston S. Taylor, Ph.D., Chairman (National Bureau of Standards), Gioacchino Failla, Ph.D. (Radiological Society of North America), Robert R. Newell, M.D. (Radiological Society of North America), H. K. Pancoast, M.D. (American Roentgen Ray Society), J. L. Weatherwax, Ph.D. (American Roentgen Ray Society), Francis Carter Wood, M.D. (AMA), Curtis F. Burnham, M.D. (American Radium Society), W. D. Coolidge, Ph.D. (General Electric Co.), W. S. Werner, (Kelley-Koett Mfg. Co.). Whittemore, "The National Committee on Radiation Protection, 1928-1960," 19-24. 74 Ibid., 135, 140, 172,200. 1> Whittemore argues that in an environment of scientific complexity and uncertainty the revision of NCRP standards in 1936, 1949, 1956, and 1959 was a reflection of more than what was being discovered through laboratory research. Rather, the impetus for regulatory revision was driven by outside events such as the controversy regarding the dialpainters of the 1920s/early 1930s, the establishment of the AEC in 1947, and the development of the nuclear power industry and fallout controversy in the 1950s. Ibid., 1 and 5. 7(' The ACXRP standard for internal emitters was set at 0.1 microcurie of radium body burden in 1941. Kocher, "Perspective on the Historical Development of Radiation Standards," 521; Stone, "The Concept of a Maximum Permissible Exposure," 643; Whittemore, "The National Committee on Radiation Protection, 1928-1960," 174 and 209. 4K

The ACXRP was an independent professional body. It was sponsored by the National Bureau of Standards, but its recommendations were not instituted as governmental regulations. Rather they were to be adopted on a voluntary basis by medical, scientific, and industrial professionals. That the ACXRP was autonomous and its standards recommended, but not enforced, is indicative of a significant difference between pre- and post-World War II radiation safety. The war marks the federal government's entry into radiation safety. When the

ACXRP reconvened after the war as the NCRP, it remained a private organization, but it was no longer the only authoritative voice for radiation safety.

The A EC instituted formal regulations, many of which were based, at least initially, on NCRP standards. The relationship between the NCRP and government in the postwar period reflects the expansion of the network of radiation researchers and clinicians that had originated not only with radiation research and medical applications in the early twentieth century, but also with radiation safety.

BROADENING THE SCOPE OF RADIATION RESEARCH: LAWRENCE'S RAD LAB & THE CYCLOTRON DURING THE 1930s

To researchers and clinicians the importance of radiation safety was always evident, but as discussed, seemed increasingly important in the wake of the controversy surrounding the dialpainting industry. The establishment of the

ACXRP at the end of the 1920s was also well-timed relative to developments in physics that generated new opportunities for radiation research and the clinical applications of radiation. Most importantly, the invention of a new sort of particle accelerator at the University of California, Berkeley, in 1930 provided a means for artificially producing radioisotopes and radiation beams.77 The new

technology—the cyclotron—was invented by physicist Ernest O. Lawrence and

one of his graduate students, M. Stanley Livingston. It was invented at a time

when physics was flourishing throughout the country, but especially in

California.78 With this new technology, Lawrence's laboratory became a hub of

cutting-edge radiation research throughout the 1930s. Lawrence, himself, was not

very interested in investigating the use of radioisotopes and radiation beams in

medicine, but recognized the great potential to do so. As we will see, Lawrence

had compelling financial motives to promote medical research. Aside from

funding issues, Lawrence was generally interested in building a successful

laboratory and supported any lines of research that, to him, looked promising.

This was evident in Dr. John Gofman's assessment of Lawrence's nature as

laboratory director. As a contemporary of Lawrence's who first began work at

Berkeley during World War II, Gofman later recalled,

Ernest Lawrence, if he thought you were sincere and were doing worthwhile work, it didn't matter whether it was high-energy physics, low- energy physics, or medicine. If you were working in his lab and he thought you were doing something useful, there was nothing too good for you in way of facilitation of your work.79

77 Radioisotopes are unstable isotopes that both occur in nature and can also be artificially produced. Radioisotopes emit radiation as they decay or attempt to become stabilized. In the process of becoming stabilized the nucleus of a radioisotope undergoes changes that may result in a transformation into a new element. This process of decay or transmutation continues until an element with a stable nucleus is formed. See "Radioisotopes in Medicine, 1948," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945- 1952, Box 28, Information & Publications, 5-6. The acceleration of research in physics at Berkeley and the California Institute of Technology (Caltcch) in particular coincided with the growth of the high tech industry in California. See Heilbron and Seidel, Lawrence and His Laboratory, 14-15; Robert Seidel, "The Origins of the Lawrence Berkeley Laboratory," in Big Science: The Growth of Large-Scale Research, ed. Peter Galison and Bruce Hevly (Stanford, CA: Stanford University Press, 1992), 21 -45. 7" Oral History of John Gofman, 1994, n.p.. 50

This openness to various streams of research helps to explain why Lawrence, a

scientist personally interested in accelerator physics, developed a laboratory that

attracted a diverse array of researchers interested in investigating and applying

radiation in physics, chemistry, biology, and medicine.

The origins of Lawrence's laboratory date back to 1928 when Lawrence

took up an associate professorship position in the Physics Department at

Berkeley. Upon arriving at Berkeley he began building a productive laboratory

focused on particle accelerators and formally established the Radiation

Laboratory or Rad Lab in 1931. Aside from his cyclotron project which was spearheaded by Livingston, he had another graduate student, David Sloan,

working on X-ray tubes. During the first few years of the 1930s it was actually

Sloan's project that generated the most external support for the laboratory due to the medical uses of X-rays. Sloan later worked on building high energy X-ray

machines for the Institute of Cancer Research at Columbia University and the

University of California, San Francisco (UCSF) School of Medicine.80 The funds given to the laboratory for these projects helped keep the whole laboratory afloat and helped fund Livingston's continued work on the cyclotron.

Throughout the decade, the number of people working on cyclotron research and development at the Rad Lab grew. So, too, did the size of the machines they created. The original cyclotron had an accelerating chamber that measured 4 inches in diameter, but Lawrence wanted to build bigger, more

s0 Francis Carter Wood was the Director of the Institute of Cancer Research at Columbia University. He was also an original member of the ACXRP. See Heilbron and Seidel, Lawrence and His Laboratory, 89, 121-22; Whittemore, "The National Committee on Radiation Protection, 1928-1960," 19. 51

powerful versions. The cyclotrons built at the Rad Lab throughout the 1930s

measured, successively, 11 inches, 27 inches, 37 inches, and 60 inches.81 In

addition to building bigger technologies, those at the Rad Lab focused on

discovering and producing radioisotopes in the cyclotron. They aimed to produce

radioisotopes that might be used to replace radium in clinical practice since, as

«p discussed above, radium was so costly. "

Building cyclotrons was not an inexpensive endeavor, either, but

Lawrence was able to attract patronage from a variety of sources to help fund the

operation of his laboratory. The state of California was the biggest contributor to

the Rad Lab during the 1930s, with both foundations and the federal government

contributing significant funds as well.X3 Lawrence had two tactics that proved

remarkably successful for obtaining funding, at least for Depression-era America.

He was able to use offers of employment from elsewhere to demand greater

financial support from the University. Furthermore, as with Sloan's X-ray tubes,

he generated considerable interest in the cyclotron within the medical community.

By the middle of the 1930s Lawrence and others in the laboratory were especially optimistic about the use of radioisotopes of sodium, phosphorus, and iron as

replacements for radium in clinical practice, especially for the treatment of cancer. At the time, the laboratory's 27-inch cyclotron was the world's biggest

producer of radioisotopes.84

si At the end of WWII, Ernest O. Lawrence was able to secure funding for the construction of a massive new cyclotron. The 184-inch cyclotron was completed in 1946. See Heilbron and Seidel, Lawrence and His Laboratory, 93, 135, 238, 69, and 483. 82 Williams, "Donner Laboratory," 16N. 10 Throughout the decade of the 1930s the lab was funded by the state, foundations, and the federal government at a ratio of 40:38:22. Heilbron and Seidel, Lawrence and His Laboratory, 207. s4 Ibid., 188-91 and 269. 52

Ernest O. Lawrcnce c. 1935. Photo Credit: Department of Energy Office of History and Heritage Resources.

The original Rad Lab c. 1930s. Photo Credit: American Institute of Physics Center for the History of Physics.

Lawrence, whose research interests always remained in physics, was somewhat reluctant to see the focus of work in the laboratory shift towards

medical research. Yet, as Waldo E. Cohn, a biochemist who completed his graduate work at Berkeley in the 1930s recalled, Lawrence was also keen to promote the medical uses of cyclotron-produced radioisotopes and radiation beams because he recognized the financial benefits of doing so. Cohn explained in an interview conducted near the end of the century that, 53

[Ernest] Lawrence, with his cyclotron, was very anxious to get money to support not only the building of the cyclotron and its maintenance, but also to continue the running of the laboratory. Any possible use of radioactive materials, [any] practical use, he could use as a gimmick to convince donors to contribute money to his project.85

Cohn's identification of Lawrence's entrepreneurial style—his willingness to

diversify lines of research within the laboratory to attract patrons—reflects the

political economy of science which, at that time, both responded to and shaped

social and economic factors such as concern for cancer and the trends of private

philanthropy. Cohn's interview illustrates a point made by historians Heilbron

and Seidel who, in their pivotal history of the Rad Lab, argue that Lawrence was skilled at selling, and even overselling, the practical applications of the

laboratory's research.86

Historian Bruce Hevly also examines Lawrence's efforts to obtain outside

funding by linking the laboratory's technologies and research to medical applications. Hevly refers to Lawrence's success with this strategy as "a sort of

'California model' for building up a physics department." 87 Hevly argues that physicists at Stanford University pursued this same strategy in relation to a proposed supervoltage X-ray tube. Although the supervoltage tube was not built at Stanford, Hevly explains that Stanford physicists succeeded, at least partially, in their endeavor to build up their department and institution by showing the applied value of their research and technologies.88

s5 Oral History of Waldo E. Cohn, Ph. D., interview conducted January 18, 1995, as part of the Office of Human Radiation Experiments' Oral History Project, Human Radiation Studies: Remembering the Early Years. Conducted by Thomas Fisher, Jr. and Michael Yuffee in Oak Ridge, TN, n.p.. Heilbron and Seidel, Lawrenee and His Laboratory, 207-22. s7 Bruce Hevly, "Stanford's Supervoltage X-Ray Tube," Osiris 9 (1994): 85. m Ib id:. 85-100. 54

Hcvly's study shows that, as was the case at Berkeley and even Memorial

Hospital, the pursuit of research at Stanford proceeded within a political economy of research that was shaped by and helped constitute professional and institutional goals, access to technologies and financial support, and links between science and the broader society. Indeed, radiation research at Stanford, Berkeley, and

Memorial Hospital illustrates an argument made by historian Timothy Lenoir that scientific disciplines and institutions are not autonomous from external factors. In fact, external factors can be used as important sources of leverage to build disciplines• • and institutions.89

BEYOND PHYSICS: INTERDISCIPLINARY COLLABORATION

Starting early in the 1930s, Ernest Lawrence and others in the laboratory collaborated with physicians at the UCSF School of Medicine. Physicians, especially radiologists like Dr. Robert Stone, were eager to investigate the medical uses of the cyclotron. Stone had been employed at the UCSF Medical

School since 1928 as the school's first full-time radiologist. He and some of his colleagues at the medical school took patients to the cyclotron to treat tumors with neutron beams. Neutron therapy was, at that time, considered by physicians and researchers to be an experimental procedure. Although, as radiologist Dr. Hymer

L. Friedell recalled in a 1994 interview, Stone's treatment of tumors with neutrons were "cancer studies" and the patients treated were selected due to the inability to effectively treat their tumors with more conventional means.90 The point, here,

m Lenoir, Instituting Science, 17. Oral History of Hymer L. Friedell, M.D., Ph. D., interview conducted September 28, 1994, as part of the Office of Human Radiation Experiments' Oral History Project, Human Radiation 55

being that Stone's treatment of cancer patients with neutron beams was not

reckless, but rather was a reasonable course of treatment given the lack of other

treatment options.

Stone helped strengthen ties between the Rad Lab and Medical School by

facilitating communication between Lawrence and the Dean of the Medical

School, Langley Porter, to further this collaboration.91 The institutional ties

between the UCSF Medical School and the Rad Lab were sufficiently strong by

the end of the 1930s that Friedell sought out an appointment at the UCSF hospital

based on his assumption that he would be able to obtain access to the Rad Lab's

cyclotron. Friedell, who had both an M.D. and Ph.D. in radiology, left his

position as a National Cancer Institute Fellow at Memorial Hospital in New York

when he acquired his position at the UCSF hospital. Friedell had gained significant experience working with X-rays and radium during his tenure at

Memorial Hospital (1939-1940) and a previous appointment at the Chicago

Tumor Institute and he was intrigued by what he heard about the new work being done with radioisotopes at Berkeley.92 Thus, he was eager to go to California where he took up work at the hospital in San Francisco, and, as he recalled, went to Berkeley every day to work with the cyclotron.

Studies: Remembering the Early Years. Conducted by Eleanor Melamed and Dr. Darrell Fisher at Case Western Reserve University, Cleveland, OH, n.p.. " Heilbron and Seidel, Lawrence and His Laboratory, 391-93. During his tenure at Memorial Hospital Friedell worked with Failla and Quimby who, in the late 1930s, were considered leading experts in the field. Friedell's prior appointment at the Chicago Tumor Institute also allowed him to collaborate with radiologist Simeon T. Cantril and physicist Ernest O. Wollan. In Chapter Two I examine the recruitment of scientists and physicians for defense research which was overwhelmingly shaped by professional connections made prior to the war. Friedell, Cantril, and Wollan were reunited during the war to work on the same defense project. See Oral History of Hymer Friedell, 1994, n.p.. 56

Friedell continued to work at UCSF and in the Rad Lab through to 1942 when he was called up for military duty. Having been a reserve officer with the

Army for a number of years, Friedell expected to be called for active duty.

However, given his expertise in radiation research he was not recruited to fight on the battlefield. Rather, he was assigned to work on the Manhattan Project, a subject examined in greater detail in the following chapter.93 Civilian scientists with similar expertise were employed in similar ways since experience working with radiation was limited to a relatively small number of researchers across the country.

In addition to collaboration between the Rad Lab and Medical School, researchers from various scientific fields took advantage of the new research avenues created by the cyclotron. For example, Waldo Cohn who completed his

Ph.D. in biochemistry in 1938 at the University of California, Berkeley, conducted research that required use of the Rad Lab's technologies and collaboration with the laboratory's employees. His graduate research used cyclotron-produced radioisotopes as tracers to investigate metabolic and physiological processes. Reflecting on this research decades later, Cohn counted himself to be "one of the few people in the country at that time that had experience with artificial radioactive isotopes as tracers in biochemical and biological experiments."94 His use of cyclotron-produced radioisotopes greatly

As we will see in Chapter Two, Friedell was not actually called up for active duty as early as he thought he might be. He discovered that due to his teaching credentials and role at UCSF he had been placed on an "essential list" which exempted him from active duty. Upon learning this, he insisted he be taken off the list because he did not wish to receive special treatment. See Oral History of Hymer Friedell, 1994, n.p.. 1,4 Oral History of Waldo Cohn, 1995, n.p.. 57

expanded what was a succcssful, but limited line of chemical and biological

research. A Hungarian chemist Georg von Hevesy had pioneered the use of

radioactive tracers to study the behavior of non-radioactive elements in 1913. By

the early 1920s, he began studying the movement of radioactive tracers in plants

and, later in the decade, researchers adopted his technique to study blood

circulation in humans. Not every element exists in nature with a radioactive form so the creation of cyclotron-produced radioisotopes expanded the variety of

radioisotopes that could be used in research. As a result, artificially produced

radioisotopes allowed researchers like Cohn to examine a greater array of

biological processes.95

Lawrence welcomed interdisciplinary collaboration with scientists who,

like Cohn, were from fields other than physics. The Rad Lab began employing biologists, chemists, and physicians who were interested in using radioisotopes to investigate biological and chemical processes or who wished to study the biological effects of the new radioisotopes produced in the cyclotron and develop therapeutic and diagnostic applications for these radioisotopes. Dr. Joseph

Hamilton, for instance, arrived at the Rad Lab after having completed his M.D. at the UCSF Medical School in 1936. He became a longstanding employee at the laboratory and, within a few years, became the director of the medical research laboratory built to house a cyclotron specifically designated for medical purposes.

Aside from Hamilton and key among those who came to the Rad Lab to pursue

AC1IRE, The Final Report, 4-5. Cohn, like so many others who at some point worked in the Rad Lab and gained experience working with radioisotopes, went on to work for the Manhattan Engineer District, first at the in Chicago and then at Oak Ridge National Laboratory in Tennessee. 58

biological and medical research, was Lawrence's brother. Dr. John Lawrence

first visited in 1935 and left his position at Yale School of Medicine the following

year to go to the Rad Lab on a permanent basis. Like his brother he recognized

that the cyclotron created new opportunities to use radioisotopes and radiation

beams for both diagnostic and therapeutic purposes. By mid-decade, the

laboratory started to develop a clinical program that, under John Lawrence's

leadership, pioneered medical applications of the newly discovered radioisotopes and radiation beams produced by the cyclotron. These studies built upon earlier work with X-rays and radium in medicine and encouraged further collaboration between researchers and clinicians of various specialties, thus strengthening the stock of radiation researchers in the United States.

Aside from the biological tracer studies pursued by researchers like Cohn, the Rad Lab's biomedical research proceeded in two general areas: leukemia research and therapy, and neutron therapy. Trained as a hematologist and endocrinologist, John Lawrence was particularly interested in hematological work. He conducted tracer studies to investigate the metabolism of radioisotopes of iron and phosphorous and went on to use phosphorous-32 (P-32) for the treatment of leukemia and polycythemia vera.96 In an interview decades later

Gofman explained why John Lawrence thought to investigate and then employ P-

32 as a therapeutic agent. John Lawrence had found that P-32 concentrates in the bone marrow and spleen, both of which contribute to blood cell production. Since leukemia is any type of cancer that originates in bone marrow or blood-forming

Nathaniel I. and Louis R. Wasserman Berlin, "Polycythemia Vera: A Retrospective and Reprise," Journal of Laboratory and Clinical Medicine 30, no. 4 (1997): 365-73; Williams, "Donner Laboratory," 16N. 59

tissue and polycythemia vera is a disease caused by the overproduction of red

blood cells, John Lawrence believed that both might be better treated with P-32

than with X-rays and radium, as had been common practice for decades.97

Gofman provided this explanation to suggest that much of what was done at the

Rad Lab was not as experimental as it might have seemed because it built upon established practices developed for X-rays and radium.

Similarly, researchers and clinicians at the Rad Lab derived considerable optimism from early investigations of the neutron beams produced by the cyclotron. From these studies they believed that neutron beams had greater

penetrating power and, therefore, might be used with greater specificity than X- rays. Experiments conducted at the Rad Lab with animals yielded results that led researchers to believe that, relative to X-rays, neutrons were both more destructive to tumors and less harmful to healthy tissue. Researchers and clinicians sought to develop this therapy as a replacement for X-ray therapy.

They were especially eager to do so since high energy X-ray machines like the million volt machines Sloan had built for the UCSF Medical School and the

Institute of Cancer Research at Columbia University earlier in the decade had not achieved any greater results in treating cancer than low energy X-ray machines.98

Physicist Paul Aebersold who had pursued his graduate work at the Rad Lab— work that focused on using cyclotron-produced neutron beams for radiation therapy—assumed responsibility for building a treatment port for the 37-inch cyclotron so that neutron beams produced in the cyclotron could be directed at

'7 Oral History of John Gofman, 1994, n.p.. ,K Heilbron and Seidel, Lawrence and His Laboratory, 389-91. 60 tumors." He collaborated with fellow physicists like Ernest Lawrence and physicians such as John Lawrence, Robert Stone, and Joseph Hamilton, thus illustrating the integral role of interdisciplinary research to the development of neutron therapy throughout the mid-1930s. Together, variously trained researchers helped create hybrid expertise that informed the investigation of neutrons as a therapeutic tool and, ultimately, the implementation of neutron therapy at the Rad Lab in 1938.

THE RAD LAB EXPANDS: THE DEVELOPMENT OF THE DONNER & CROCKER LABORATORIES

New uses of radioisotopes and neutron therapy at the Rad Lab yielded some successes, but disappointments as well. For instance, a decade after neutron therapy was implemented, researchers were dissatisfied with the results of this treatment practice.100 The use of radioisotopes like radiophosphorous and radioiodine to target specific types of tumors, however, met researchers' optimistic expectations. John Lawrence's use of P-32 to treat polycythemia vera and one type of leukemia, myelogenous leukemia, was among the most exciting results to come out of the Rad Lab's biomedical work in the 1930s.101 Such successes were useful to attract wealthy benefactors who provided funds to build

w Aebersold completed his Ph.D. in 1939 and stayed on at the Rad Lab, becoming an integral member of the lab and one who contributed much to the lab's Manhattan Project and AEC research. See Ibid., 359-60. 100 The Rad Lab treated its first patient with neutron therapy in 1938 using the 37-inch cyclotron. The 60-inch was used for neutron therapy starting in 1939 and a regular clinical program was established in 1940. The survival statistics for those patients treated with neutron therapy during those early years were not as successful as had been expected. By 1948 radiologist Robert Stone recommended against continued neutron therapy due to little success in curing cancer and the late appearance of side effects worse than those resulting from X-rays. Essentially he argued that the negative results of neutron therapy far outweighed the positive. Neutron therapy ceased for two decades, until 1970. Ibid., 394. 101 Ibid., 359-60. 61

facilities for both a clinical program and medical research laboratory. The clinical

program and medical research laboratory were under the Rad Lab umbrella, yet,

from an operational standpoint, somewhat different entities.

Joseph Hamilton (left) consumes radiosodium while colleague R. Marshak traces its movement throughout his body, c. 1949. Photo Credit: Ernest Orlando Lawrence Berkeley National Laboratory Image Library.

Robert Stone and John Lawrence administer neutron therapy to patient Robert Penney at the 60-inch cyclotron c. 1940. Photo Credit: American Institute of Physics Center for the History of Physics.

The development of the medical research laboratory dates back to 1936 when Ernest Lawrence sought funding to build a bigger cyclotron—a 60-inch cyclotron—one that would be specifically devoted to medical research. With the help of Berkeley's President Robert Sproul, Ernest Lawrence obtained funds from various patrons to build the cyclotron. One patron, William H. Crocker who was 62 a regent of the University, donated funds to build a new laboratory to house the

medical cyclotron. Crocker donated this gift to the Rad Lab because he was very impressed with the research being conducted there. This was a time at which researchers were very excited about results obtained from animal research investigating the neutron beams produced by the 37-inch cyclotron. They were keen to pursue neutron therapy with humans.102

Ernest Lawrence's plans to build the medical cyclotron and medical research laboratory unfolded alongside the establishment and growth of a clinical program which developed under John Lawrence's leadership. The Rad Lab's clinical work which had started in the mid 1930s acquired a new home in a separate and newly constructed building, the Donner Laboratory, by the early

1940s. This building was primarily funded by William H. Donner, the president of the International Cancer Research Foundation, who visited the laboratory in

1940 after hearing about John Lawrence's research. Similarly to Crocker, Donner had great faith in what he saw developing at the Rad Lab. He embraced the optimism that the Lawrence brothers and their colleagues exuded. Donner also had personal motivations for supporting cancer research. Having lost a son to cancer, he sought to improve cancer therapy by helping to advance the work of

John Lawrence and his colleagues.103

THE RAD LAB & THE DEVELOPMENT OF CANCER RESEARCH IN THE UNITED STATES

102 Ibid., 209-10, 391. Donner donated $150,000 for the construction of the Donner Laboratory. Williams, "Donner Laboratory," 18N. 63

The development of the Rad Lab's clinical and medical research programs and creation of new laboratories in which to conduct this work coincided with increased government support of cancer research; the work pursued at the Rad

Lab and the changing landscape of cancer research and therapy were trends that fuelled each other. The federal government passed the National Cancer Institute

Act in 1937 which provided for the creation of the National Cancer Institute

(NCI) in Bethesda, Maryland. Many historians emphasize the context of the New

Deal to explain the creation of the NCI. For instance, historian James Patterson argues that during the New Deal, Americans sought to establish a healthier society and were willing to turn to the federal government for solutions to such problems as cancer. Further, he adds that a few individuals within Congress were especially committed to passing legislation aimed at addressing cancer because cancer, more than any other disease at that time, evoked feelings of immense

"dread."104 According to Patterson and others, this new institute was established with the express purpose of advancing cancer research. However, historian David

Cantor argues persuasively that, at least during the first year of the NCI's existence, its main focus was to use federal resources to provide cancer treatment to the poor. He bases his argument on the fact that over half of the NCI's initial budget was allocated to the purchase of radium to be used for routine therapy, not research.105

"M James T. Patterson, The Dread Disease: Cancer and Modern American Culture (Cambridge, MA: Harvard University Press, 1987), 115-20. 105 David Cantor, "Radium and the Origins of the National Cancer Institute," in Biomedicine in the Twentieth Century: Practices, Policies, and Politics, ed. Caroline Hannaway (Amsterdam, The Netherlands: IOS Press, 2008), 95-96. The sort of debate that evolved in relation to NCI policies regarding the Institute's support of cancer research and therapy were ongoing throughout history. 64

Although the NCI allocated a large portion of its initial budget to purchasing radium to be loaned to hospitals for therapeutic uses, the NCI did also support extramural research and would, over time, prioritize the advancement of research. The NCI's advisory council, the National Advisory Cancer Council

(NACC), organized the distribution of grants to researchers including those at the

Rad Lab. For instance, an NCI grant provided to the Rad Lab helped pay for the construction of the 60-inch cyclotron. While funded by numerous sources, the progress of construction for the new medical cyclotron was slow due to insufficient funds. Thus, Ernest Lawrence began pressing the NACC for funds in

1937, almost immediately after the NCI had been created. One of the six NACC members, Arthur H. Compton, who was a physicist at the University of Chicago, visited the Rad Lab to assess the progress of construction. Compton and

Lawrence had met in the mid-1920s when Compton was a professor and

Lawrence, a graduate student, at the University of Chicago.106 Compton did not hesitate to help secure a grant to ensure completion of the medical cyclotron.

According to one of Compton's colleagues, during Compton's visit to Berkeley on behalf of the NACC, he "had been impressed by [Lawrence]."107 With the additional financial support of the NACC grant, provided in 1938, the medical

108 cyclotron was completed in 1939. The development of biomedical research

As we will see in Chapter Six, the Atomic Energy Commission struggled to determine how best to distribute funds in an effort to combat cancer. Juan A. del Regato, Radiological Physicists (New York: American Institute of Physics, 1985), 130. 1117 This was the opinion of Juan A. del Regato who was a physician at the Chicago Tumor Institute where, as of 1937, Compton served as the Vice President and member of the Scientific Committee. See /bid., 130. ")s Heilbron and Seidel, Lawrence and His Laboratory, 207-09, 214-15, 269; Patterson, The Dread Disease, 131-3. 65 around the cyclotron facilitated correspondence and collaboration amongst researchers such as Ernest Lawrence and Compton during the 1930s. This collaboration would continue, although during the war the development of technologies and pursuit of research for the benefit of cancer and other medical research was secondary to research and development that more directly contributed to the war effort.

The level of funding the government provided for cancer research in the

1930s was minimal, at least compared to what it would become in the latter half of the twentieth century. To provide some perspective on the scope of the NCI's pre-World War II activities, historian Walter Ross notes that the NCI had a relatively small budget in its early years. The Institute's first annual budget was

$400,000 and it did not grow significantly until after World War II.109 Patterson also reports that the NCI's pre-war in-house research was "small and manageable."110 Patterson does argue, though, that the creation of the NCI, the first of the National Institutes of Health, marked the government's commitment to help fight cancer by supporting research and improving diagnostic and therapeutic practices.111 The establishment of the NCI was, indeed, a notable development in the history of cancer research. The NCI provided federal infrastructure to help foster and support research both in-house and beyond. The grant given to the Rad

Lab is an early and good example of how the NCI supported extramural cancer research.

"" Walter Sanlord Ross, Crusade: The Official Historv of the American Cancer Society (New York: Arbor House, 1987), 210-14. "" As of 1940, the NCI employed 28 researchers at its Bethesda facility. Patterson, The Dread Disease, 134. 111 For a discussion on the creation of the NCI, see Ibid., 115-21. 66

Ross argues that rather than the NCI or any other government endeavor, it

was the creation of a research program by the American Cancer Society (ACS)

that first expanded American cancer research by a significant measure. The ACS

had been in existence since 1913, but only established its research program in

1945, at which time it allocated $1 million for this purpose. This was a

significant research budget considering that as of 1945 cancer research in the

United States received about $1.5 million in funding from all other sources, both

public and private. Government funding had risen since the creation of the NCI

112 and accounted for half this amount.

As is evident in the history of the NCI and the establishment of the ACS

research program, both public and private infrastructure to support cancer

research started to expand in the late 1930s and early 1940s. The research

pioneered at the Rad Lab during the same period—investigations of the use of

radioisotopes and radiation beams produced by the cyclotron—benefitted from

this expansion and also helped drive it. The key point here is that society shaped

science, but so too did science shape society. Furthermore, researchers played a

central role in linking their research to social agendas. That they did so was a key

element for building institutional support around the interdisciplinary research

they pursued and the resulting hybrid expertise. Drawing on the work of historian

Timothy Lenoir, I suggest that those at the Rad lab were engaged in building a disciplinary program. For Lenoir, disciplinary programs are research programs around which researchers build disciplines. To do so researchers use factors both

internal and external to their research to leverage economic and cultural capital

112 Ross, Crusade, 37-38, 210-14. 67

between disciplines.113 In the case of the Rad Lab, researchers used evidence of

their clinical use of radioisotopes and neutrons which they deemed as successful,

as well as concern for cancer, to leverage resources and prestige that would help

facilitate the eventual creation of new biomedical disciplines. This did not,

however, result in a lack of resources for the physical research programs pursued at the Rad Lab. The Rad Lab received ample funding from various sources which allowed multiple research programs to flourish.

THE CYCLOTRON TAKES CENTER STAGE

The cyclotron was at the heart of the growth of the Rad Lab and provided considerable impetus for the expansion of cancer and other biomedical research.

This technology, which encouraged interdisciplinary collaboration at the Rad Lab, did the same at other institutions where cyclotrons were built. For instance, the promotion, construction, and use of a cyclotron at the University of Rochester enhanced collaboration between physicists and physician researchers who were engaged in or interested in radiation research. Physicist Lee DuBridge and Dr.

Stafford L. Warren, a radiologist at the Medical School, began collaborating shortly after DuBridge's arrival in 1934. In an interview conducted in the 1970s

DuBridge recalled that Warren reached out to DuBridge, inviting him to the

Medical School to learn about Warren's and his colleagues' work with X-rays.

Warren was keen to acquire a cyclotron at the University for medical research and

DuBridge, being well acquainted with Ernest Lawrence and the development of the cyclotron at the Rad Lab, was equally enthusiastic about the prospect of

113 Lenoir, Instituting Science, 16-17 and 55. 68

acquiring a cyclotron. There was additional interest amongst Rochester area engineers.114 At the University of Rochester, this technology opened avenues in

many fields of research just as it had at the Rad Lab. Although the Rochester cyclotron was not used for medical research to the degree intended, at least not

initially, DuBridge recalled that medical interest was important for securing

funding for the project. By the late 1930s, it was far easier to obtain financial support to construct a cyclotron for biological and medical research rather than for physics research.

Beyond interdisciplinary collaboration within institutions, and the linking of research with medical applications, cyclotrons also helped foster professional bonds between institutions. For instance, when the University of Rochester embarked on their cyclotron project in 1935 they maintained close communication with researchers at Cornell University. Cornell had preceded the

University of Rochester in building a cyclotron. The first university outside of

Berkeley to build a cyclotron, Cornell began constructing its cyclotron in 1934.

DuBridge explained that contact between the Cornell and Rochester faculty was quite extensive, ranging from joint seminars to parties. The Cornell and

Rochester cyclotrons were two amongst the twenty-two cyclotrons that existed in the United States by 1940.115 Researchers at these two universities and the others that built cyclotrons in the 1930s were also in close contact with Ernest Lawrence and his colleagues at the Rad Lab since members of the Rad Lab played an

114 DuBridge explains that support for the acquisition of a cyclotron at the University was expressed at an annual meeting of electrical engineers in the Rochester area. Oral History of Lee DuBridge by Charles Weiner on June 9, 1972, Niels Bohr Library & Archives, American Institute of Physics, College Park, MD, USA, (accesscd 31 July 2009), n.p.. "' For a complete list, see Heilbron and Seidel, Lawrence and His Laboratory, 301. 69 integral role in helping promote and build cyclotrons at other institutions.116

According to DuBridge, Ernest Lawrence was eager to facilitate the spread of this technology because Lawrence firmly believed that cyclotrons would be very important for the development of nuclear physics and other fields of research."7

BEYOND BERKELEY: PRE-WORLD WAR II RADIATION RESEARCH THROUGHOUT THE UNITED STATES

While Berkeley was certainly a locus of radiation research in the pre-

World War II period and Rad Lab researchers were pioneers in investigating the medical applications of cyclotron-produced radioisotopes and radiation beams, there were other researchers elsewhere that also conducted important biomedical radiation research. For instance, Dr. Stafford Warren who, as mentioned above was a radiologist at the University of Rochester, investigated the medical uses of

X-rays starting in the 1920s and later engaged in research with the University's

118 cyclotron. Warren and his Rochester colleagues in the Departments of

'16 The spread of the cyclotron to other institutions involved personal visits from Rad Lab members to help promote and build the technology. For instance, one of the original creators of the cyclotron, Stanley Livingston went to Cornell University in 1934 to build the first cyclotron outside of Berkeley. The following year Ernest O. Lawrence himself visited the University of Rochester to discuss the construction of a cyclotron at that university. During his visit he helped locate an ideal space on campus where the cyclotron could be located. Oral History of Lee DuBridge, 1972, n.p. 117 For instance, Lawrence made radioactive isotopes produced in the Berkeley cyclotron available to colleagues at other institutions. He provided researchers at the University of Rochester radioactive iron to be used as tracers in research investigating the metabolism of iron in relation to anemia. See Paul F. Hahn et at., "Radioactive Iron and Its Metabolism in Anemia: Its Absorption, Transportation, and Utilization," Journal of Experimental Medicine 69, no. 5 (1939): 739. "s Some of the publications emerging from Warren's and his colleagues' work include: Herman E. Pearse and Stafford L. Warren, "The Roentgenograph^ Visualization of the Arteries of the Extremities in Peripheral Vascular Disease," Annals of Surgery 94, no. 6 (1931): 1094; S. L. Warren and G. H. Whipple, "Roentgen Ray Intoxication: Intestinal Lesions and Acute Intoxication Produced by Radiation in a Variety of Animals," Journal of Experimental Medicine 38, no. 6 (1923): 741; S. L. Warren and G. H. Whipple, "Roentgen Ray Intoxication: The Path of a Beam of Hard Rays in the Living Organism," Journal of Experimental Medicine 38, no. 6 (1923): 731; Stafford L. Warren et al., "A Quantitative Method for Studying Roentgen-Ray Absorption of Tooth Slabs," American Journal of Roentgenology and Radium Therapy 31 (1934): 663. See also William F. Neuman, "An Horatio Alger Story: From War Work to Orphan Academic 70

Radiology, Physiology, and Physics were active in helping to create a new program in biophysics in 1936 that was jointly sponsored by these three departments. Biophysical research dates back earlier than the mid-1930s, but the

University of Rochester was one of the first to offer a graduate program in the field."9 Interdisciplinary radiation research was a central component of this new program, so much so that, as we will see in Chapter Five, the biophysics program was absorbed into a new interdisciplinary department of Radiation Biology following the war.

Stafford Warren pictured with a nurse and patient at a portable X-ray unit, circa 1926. Photo Credit: University of Rochester Medical Center Historical Photographs, URMC Edward G. Miner Library.

Others who developed an expertise in radiation research while pursuing various types of biomedical research included Drs. Shields L. Warren

(pathologist) and Joseph Aub (endocrinologist) at Harvard University, and Robley

D. Evans (physicist) at Massachusetts Institute of Technology (MIT). Shields

Warren's pre-World War II career encompassed positions at the Boston City

Hospital, New England Deaconess Hospital, and the Massachusetts State Tumor

Heavyweight in Three Decades—the Story of Radiation Biology and Biophysics," in To Each His Farthest Star: University of Rochester Medical Center, 1925-1975, ed. John Romano, et al. (Rochester, NY: The University of Rochester Medical Center, 1975), 281. Neuman, "Radiation Biology and Biophysics," 278-91; J. Newell Stannard, "Zealous Companions in Research: The Graduate Studies Program," in To Each His Farthest Star: University of Rochester Medical Center, 1925-1975, ed. John Romano, et al. (Rochester, NY: University of Rochester Medical Center, 1975), 130. 71

Diagnosis Service. In each of these positions he was able to pursue work related

120 to his interest in the effects of radiation on both tumors and normal tissue.

Evan's pre-war work helped establish him as an expert in radium studies. He completed his Ph.D. in Physics in 1932 at the California Institute of Technology and immediately following this began research on radium poisoning. He spent a year at Berkeley before moving east to MIT where he conducted extensive

investigations of the radium poisoning experienced by dialpainters.121 Aub collaborated with him on this work.122 With Shields Warren, Joseph Aub, and

Robley Evans at Harvard University and MIT, Cambridge, Massachusetts was another locus of radiation research similar to the Berkeley in the pre-war period.

CONCLUSION

The development of biomedical radiation research and clinical applications of radiation began shortly after the discovery of radiation around the turn of the century. This work attracted physicians, physicists, and an increasingly diverse array of researchers and technicians. As this chapter has illustrated, there are numerous early twentieth-century examples of biomedical radiation research and clinical uses of radiation that brought together interdisciplinary groups of researchers. The collaboration of Drs. Henry Janeway

1211 For instance, John Ungar Jr. and Shields Warren, "Skin Grafting as a Method of Determining the Biologic Effect of Radiation," Archives of Pathology 23 (1937): 299-306; Shields Warren, "The Distribution of Doses of Radioactive Phosphorus in Leukemic Patients," Cancer Research 3 (1943): 334-36; Shields Warren and R. F. Cowing, "The Distribution of Doses of Radioactive Phosphorus in Rodents," Journal of Laboratory and Clinical Medicine 26 (1941): 1014-16; Shields Warren and Olive Gates, "Radiation Pneumonitis: Experimental and Pathologic Observations," Archives of Pathology 30 (1940): 440-60. See also, A.M. Brues, "Shields Warren: (1898-1980)," Radiation Research 88, no. 2 (1981): 431 -32. 11 Robley D. Evans, "Radium Poisoning A Review of Present Knowledge," American Journal of Public Health and the Nation's Health 23, no. 10 (1933): 1017-23. Rowland, Radium in Humans, 28-29. 72 and Benjamin Barringer with biophysicist Gioacchino Failla starting in the 1910s is, for instance, an early example of interdisciplinary collaboration. This experience encouraged Failla to continue the trend of interdisciplinary research by establishing his own laboratory and ultimately helping to create new hybrid fields of research including radiation biophysics and radiobiology. Starting in the

1930s, Ernest O. Lawrence institutionalized this sort of collaboration by building a large and truly interdisciplinary radiation research laboratory at Berkeley.

Throughout that decade the Rad Lab grew significantly and employed radiation researchers trained in various fields. The laboratory's research was at the cutting edge of nuclear physics and biomedical radiation research and applications, not to mention, technological development. The groundbreaking biomedical research conducted at the Rad Lab led to the establishment of a new academic program in

1 7^ medical physics shortly after the end of World War II.

Biomedical radiation research in the early decades of the twentieth century was quickly translated into medical practice. Biomedical researchers ensured that research and medicine were mutually reinforcing practices. The development of

X-rays as a diagnostic tool and the use of X-rays, radium, and later cyclotron- produced radioisotopes and radiation beams to treat cancer changed medical practice considerably. For instance, the use of radiation therapy complimented and, in some cases, replaced surgical procedures commonly used to treat cancer.

Such clinical applications and related investigations of medical uses of radiation

123 David S. Jones and Robert L. Martensen, "Human Radiation Experiments and the Formation of Medical Physics at the University of California, San Francisco and Berkeley, 1937-1962," in Useful Bodies: Humans in the Service of Medical Science., ed. Jordan Goodman, Anthony McEUigott, and Lara Marks (Baltimore, MD: Johns Hopkins University Press, 2003), 81 -108. 73 played an important role in securing patronage and facilitating interdisciplinary collaboration amongst researchers. That is, biomedical radiation research was part of a political economy of research that incorporated a scientific agenda, as well as social and political factors. Within this political economy, researchers derived both scientific and social authority by collaborating with others and developing hybrid biomedical expertise. As we will see in the next chapter, interdisciplinary collaboration and hybrid expertise remained important trends for biomedical radiation researchers when navigating wartime changes in the political economy of research. 74

CHAPTER TWO

BIOMEDICINE AND BOMBS: HEALTH AND SAFETY IN THE MANHATTAN PROJECT

In 1938 the prospect of war loomed and nuclear fission was discovered in

Germany. Physicists around the world contemplated the creation of an atomic

bomb. The United States did not enter World War II when it began in 1939, but

the government and military, with encouragement from the scientific community,

did support preliminary research investigating the feasibility of building such a

weapon. The United States went to war in December 1941 and formally

embarked on an atomic bomb project code-named the Manhattan Project during

the summer of 1942. With the initiation of this project, radiation research

acquired the utmost importance. Although the government and military

prioritized research that directly contributed to the development of a weapon,

biomedical radiation research did play a role within the atomic bomb project.

This chapter examines the wartime mobilization of what was, at that point,

a network of variously trained scientists and physicians, who shared a commitment to investigating and applying radiation in biomedical research and clinical practice. By the outbreak of war, no university had yet created new

programs that provided formal institutional structure for biomedical radiation

research, but researchers at a few institutions, including the University of

Rochester and the University of California, Berkeley, had established research programs that would later serve as a foundation for new disciplines. They had engaged in an early phase of discipline-building or what historian Timothy Lenoir 75

refers to as a disciplinary program.124 Researchers at these institutions sought to

investigate particular research problems as well as situate their research programs

within the disciplinary structure of their institutions such that they served both

institutional and social needs. This was part of a process to obtain resources

necessary to sustain their work and to establish the institutional support for their

research that comes with the creation of disciplines.

How did the health and safety objectives of the Manhattan Project affect the development of this network of researchers and their objectives to advance

knowledge and establish institutional support for their work? What role did biomedical radiation researchers play, not only in meeting Manhattan Project objectives, but also in reorganizing the political economy of research? With the

mobilization of many of the nation's leading biomedical radiation researchers, their central pre-war objectives of employing radiation in biomedical research and clinical practice and obtaining support for this work was overshadowed by defense objectives. The most important of these was the need to investigate, determine, and safeguard against health hazards that might arise during the development of the atomic bomb. Biomedical radiation researchers were also responsible for determining the health consequences of fission products in case enemy nations used these highly radioactive materials as weapons on the battlefield.125 Perhaps surprisingly, investigating the health consequences of

'"4 Lenoir, Instituting Science, 55. 1:5 Attending to this responsibility differed little from protecting workers in weapons facilities at home. Both depended on obtaining a better understanding of the biological effects of the radioactivity generated by fission products. See Hacker, The Dragon's Tail, 46-47. radiation released by an atomic bomb explosion—what is now known as fallout— was not a high priority for biomedical radiation researchers during the war.126

As part of the Manhattan Project, biomedical radiation researchers worked within a vastly different context than they had prior to the war. The scientific enterprise was overwhelmingly shaped by the role of the federal government in facilitating the reorientation of research and development for defense purposes.

The history of wartime biomedical radiation research within it could be told as a story of profound change. However, this chapter argues that despite the defense objectives that drove the Manhattan Project and the relationship established between biomedical radiation research and the state, the wartime history of biomedical radiation research is, in many respects, a story of continuity and expansion. Throughout the war, the existing professional connections amongst biomedical radiation researchers were reinforced as researchers were recruited to fill the ranks of the Manhattan Project's small, but expanding health and safety operation. These connections were central to the reorganization of biomedical researchers during the war.

This chapter argues that the Manhattan Project's health and safety work engaged and reorganized a pre-existing network of biomedical radiation researchers to counter the belief that scientists worked independently before the advent of large-scale government-funded science during the war. Such an attitude l2fi As wc will see in Chapter Three, the health effects of radioactive fallout became an important area of research following the atomic bombings of Hiroshima and Nagasaki in August 1945, and even more so throughout the 1950s as the United States conducted atmospheric nuclear weapons testing. See Oral History of Hans Bethe, Frederick Reines, Robert Christy, and .1. Carson Mark, by Stanley Goldberg, interview conducted August 18, 1989, as part of the Smithsonian Videohistory Program (18), session 14, "The Manhattan Project: Collection Division 4: Los Alamos," 31; and Howard Ball, Justice Downwind: America's Atomic Testing Program in the 1950s (New York: Oxford University Press, 1986). 13. 77 was bolstered, if not fashioned, in the wake of World War II in an attempt to emphasize the success of wartime agencies in mobilizing disparate scientists.

Vannevar Bush, an engineer and science manager who played a pivotal role in organizing science for war and who is examined in greater detail below, contributed to this skewed perception. In an official history of the wartime Office of Scientific Research and Development (OSRD), Bush described the collection of scientists as hardly having any semblance of a community at all. "The scientific group, both because of the individualistic approach essential to research and because of the sufficiency of the loosest of organization for all practical purposes in normal times," Bush contended, "is much more a gathering of

197 individuals than a group in the professional or structural sense." This may be true when speaking of the scientific community as a whole before World War II, but does not accurately depict radiation research in the decades preceding the war.

Contrary to Bush's statement that an individualistic approach was "essential" to research, radiation researchers had long embraced interdisciplinary collaboration as a means to advance research and obtain resources.

In pointing out the inconsistency between Bush's judgment of scientific research as an individualistic endeavor and the reality of radiation research in the early decades of the twentieth century, this study does not intend to diminish the importance of the government and military research and development infrastructure created during the war. The agencies examined below played a crucial role in facilitating collaboration between different groups, including

1-7 Vannevar Bush, "Foreword," in Organizing Scientific Research for War: The Administrative History of the Office of Scientific Research and Development, ed. Irvin Stewart (Boston, MA: Little, Brown and Company, 1948), iv. 78

researchers, the government, military, and industry. They managed research and development projects that proceeded rapidly and on a much larger scale than had ever been attempted in the past. Furthermore, wartime agencies created new opportunities for researchers to influence government and defense policy. Indeed, the mobilization of biomedical radiation researchers allowed them to further strengthen the authority they established within the pre-war scientific community and larger society. The key point here is that both before and during the war, interdisciplinary research and resulting professional connections were integral to biomedical radiation researchers' efforts to achieve specific objectives—research and discipline-building objectives that reflected and helped constitute the political economy of research in which they worked.

GOVERNMENT-FUNDED SCIENCE IN PERSPECTIVE: A BRIEF HISTORY OF THE GOVERNMENT'S ROLE IN SCIENCE BFORE WORLD WAR II

The prioritization of defense objectives and the role of the government in facilitating this shift were central characteristics of wartime research and development. Prior to World War II, the federal government had very limited involvement in the development of biomedical radiation research. As discussed in the previous chapter, this research was driven by academic, medical, and, to a certain extent, industrial interests. The one notable exception was the National

Cancer Institute's (NCI) support of research for the development of cancer therapy. The creation of the NCI provided an avenue through which the federal government supported some biomedical radiation research. 79

Beyond biomedical radiation research, the federal government did provide some support to science, although on a vastly smaller scale than during World

War 11 and following. The comparatively modest government bureaucracy that existed to help bridge the gap between the nation's researchers and policy-makers was primarily comprised of the National Academy of Sciences (NAS) and its operating arm, the National Research Council (NRC). Perhaps not surprisingly, both the NAS and NRC were created when the nation was at war. The NAS was established in 1863 during the Civil War to advise the government on matters of science. The formation of the NRC came later in 1916 during World War I. It was established to conduct research related to the problems presented to the

NAS—problems that, at the time the NRC was created, related to war. While created by the federal government, both were always intended to operate independently from the government. Essentially, they provided a means through which to organize civilian researchers to study scientific problems and provide

1 78 information to government agencies, both civilian and defense.

Aside from the NRC, World War I prompted the government to create wartime agencies that helped facilitate the development of science for war. The most notable included the Naval Consulting Board (NCB) and Chemical Warfare

Service (CWS). The former, which was generally regarded as a failure, was established under the leadership of Thomas Edison in 1915 to investigate the possible development of defensive and offensive technologies to counteract the

128 For a comprehensive overview of the history of the NAS and NRC, see Rexmond C. Cochrane, The National Academy of Sciences: The First Hundred Years, 1863-1963 (Washington, DC: The National Academy of Sciences, 1978); Frederick W. True, A History of the First Half-Century of the National Academy of Sciences, 1863-1913 (Washington, DC: National Academy Press, 1913). 80

German submarine.129 The CWS was officially created in 1918 to centralize various initiatives for the study and development of gas warfare already underway.130 At the end of World War I most of the organizations formed to facilitate wartime research and defense were dismantled. Both the NRC and the

CWS were maintained, but neither was used to significantly expand the government's role in science. Without the threat of war, the government did not show much interest in funding or directing research and development. Thus, during the interwar years science was pursued primarily by academic and industrial researchers and the main source of external funding for scientific research came from private philanthropists, not the states or federal government.

A WARTIME ALLIANCE: SCIENCE & GOVERNMENT

During World War II the federal government created and invested tremendous funds in both civilian and defense agencies to facilitate research and technological development. This topic has been well explored in the literature, marking wartime government involvement in and support for research and development as a significant period in the history of science.131 In the context of this study, these agencies and the weapons projects they organized are important

l2'' Thomas P. Hughes, American Genesis: A Century of Invention and Technological Enthusiasm, 1870-1970 (Chicago, IL: University of Chicago Press, 1989), 118-26; Kevles, The Physicists, chapters 8-9. 1.0 Leo P. Brophy and George J. B. Fisher, The Chemical Warfare Service: Organizing for War (Washington, DC: Center of Military History, U.S. Army, 2004. First published 1959 by the Center of Military History), 3-17. 1.1 Hewlett and Anderson, The New World', Stewart, Office of Scientific Research and Development. Government involvement in science has been the focus of much work that examines the emergence of "Big Science" and the military-industrial-academic-complex. For example, see Peter Galison and Bruce Hevly, Big Science: The Growth of Large-Scale Research (Stanford, CA: Stanford University Press, 1992); Roger L. Geiger. Research and Relevant Knowledge: American Research Universities since World War II (New York: Oxford University Press, 1993); Hughes, The Manhattan Project', Leslie. The Cold War and American Science. 81 because they changed the political economy of science in which researchers operated. It was within this infrastructure that biomedical radiation researchers established themselves not only as valuable researchers capable of serving the nation's defense needs, but also as effective administrators. The main civilian agency formed during World War II to mobilize the nation's scientific and technological manpower for defense problems was the Office of Scientific

Research and Development (OSRD). Although the OSRD was established in

June 1941, before the United States entered into the war, it was not the first agency created to help organize science for war. Rather, the creation of the

OSRD was preceded by the formation of the National Defense Research

Committee (NDRC) in June 1940.

The NDRC was placed under the direction of Vannevar Bush, an electrical engineer who was, at that time, the president of the Carnegie Institution of

Washington. Bush played a significant role in convincing President Franklin D.

Roosevelt to create the NDRC and later, the OSRD, for which he was also appointed Director. Under normal circumstances Bush preferred that the federal government not interfere in civilian affairs and that government bureaucracy not expand. For instance, during the 1930s he had criticized the expansion of the government that resulted from President Roosevelt's creation of numerous New

Deal agencies. However, given the circumstances of war, Bush encouraged bureaucratization as a means of enlisting the nation's researchers and engineers for defense work. He envisioned the NDRC as a much needed bridge between civilian science and the government.132 Bush's shift in opinion regarding the

relationship between the government and the scientific enterprise hinged on his understanding of the relationship between science and society. For him, science could play an integral role in a nation preparing for and then engaged in war.

Bush was not the only scientist who sought to establish the NDRC. He had much support from leading scientists such as Karl T. Compton, physicist and

President of MIT; James B. Conant, chemist and President of Harvard University; and Frank B. Jewett, electrical engineer and President of both the NAS and Bell

Telephone Laboratories. These men were all close colleagues and each occupied a position of considerable prominence from which they could assist Bush in convincing the President to create the NDRC. Compton, Conant, and Jewett were all appointed to the NDRC along with Rear Admiral Harold G. Bowen; Brigadier

General George V. Strong; Conway Peyton Coe, attorney and Commissioner of

Patents; and Richard C. Tolman, California Institute of Technology physicist and chemist.133

To a degree, the creation of the NAS, NRC, and NDRC/OSRD during the

Civil War, World War I, and World War II, respectively, served one common purpose. Each organization provided the means to help organize civilian scientists and engineers for defense-oriented research and development.

| Bush's efforts to create not only the NDRC, but also the OSRD and to formally establish the Manhattan Project are well documented by historians Stanley Goldberg and G. Pascal Zachary. Goldberg, "Inventing a Climate of Opinion," 431, 433, 448-51; G. Pascal Zachery, Endless Frontier: Vannevar Bush, Engineer of the American Century (New York: The Free Press, 1997), 97-117, 129-33, 189-217. m The membership of the NDRC was defined so as to include the President of the NAS, Commissioner of Patents, and two individuals appointed by the Secretary of the Army and Secretary of the Navy. The four remaining members were appointed based on their scientific and technological expertise. See Stewart, Office of Scientific Research and Development, 4 and 7. 83

However, according to former Executive Secretary and later Deputy Director of the OSRD, Irvin Stewart, the scope of the NDRC and OSRD was meant to extend beyond that of the NAS and NRC. In an official history of the OSRD published shortly after the war—a time when the government's future involvement in science was a controversial topic—Stewart argued that the outbreak of World

War II had required the creation of additional governmental infrastructure for wartime research and development. Specifically, Stewart asserted that while the

NAS and NRC were created with the intention that they would respond to requests from government agencies, military services included, as of World War

II, "the number of men in the armed services capable of knowing what was needed was small."134

Stewart's message was not intended as a criticism of the military for lacking personnel with adequate foresight to plan for weapons development.

Rather, it was an acknowledgement that science had advanced significantly since

World War I. He reckoned that the state of modern science was such that

"military chieftains were not sufficiently acquainted with its possibilities to know for what they might ask with a reasonable expectation that it could be developed."135 While scientists and engineers were not generally expert in matters of national defense they did possess an important body of knowledge that equipped them to assume a more active role in directing defense research and

1,4 Stewart was the Executive Secretary of the OSRD from its inception through December 1945 at which point he assumed the newly created position of Deputy Director. Given the postwar debate regarding government involvement in research and development, Stewart's account of the OSRD was highly political. Ibid., 5 and 181. 115 Ibid. 84

development. This was certainly the case with the Manhattan Project which, from

the beginning, was driven by scientists.

SCIENCE, THE GOVERNMENT, AND THE MAKING OF THE MANHATTAN PROJECT

The creation of the NDRC and OSRD in 1940 and 1941, respectively,

formalized a relationship amongst the scientific community, government, and

military that was already being nurtured on an ad hoc basis as individuals from

these groups began to organize defense research. In terms of atomic bomb

research, scientists had taken the lead in informing the government and military of the possible military uses of nuclear fission starting in 1939. With prompting from physicists like Enrico Fermi, , Eugene Wigner, and Albert

Einstein, the government appointed an Advisory Committee on Uranium at the end of that year to pursue preliminary research on uranium fission.136 In a personal account of the Manhattan Project, University of Chicago physicist

Arthur H. Compton argued that the Advisory Committee was actually detrimental to the advancement of uranium research. He explained that the existence of this committee created an illusion that the government was prepared to further research on nuclear fission. According to Compton, though, as of late 1939 and early 1940 the government was not sufficiently well organized for or committed to pressing ahead with this research.137 The situation changed when the NDRC

'',h For a more detailed discussion of Fermi's effort to inform the Navy of the possibility of developing nuclear fission for defense purposes, or Szilard's, Wigner's, and Einstein's similar eflbrts to inform the government, see Arthur Holly Compton, Atomic Quest: A Personal Narrative (New York: Oxford University Press, 1956), 27-28; Goldberg, "Inventing a Climate of Opinion," 274. These sources document the government's support of radiation research prior to the creation of the NDRC. 117 Compton, Atomic Quest, 29-30. was created in mid-1940 and was further improved with the creation of the OSRD the following year. At that time, the Uranium Committee was renamed the S-l

Section and continued its work under the direction of the OSRD which had a greater capacity to organize research and provide the resources necessary to translate research into technological development projects.

The S-l Section oversaw various research projects focused on obtaining the amounts of fissionable materials that scientists believed were necessary for building an atomic bomb. Uranium-235 was the first material found to produce a rate of fission that could sustain an explosive chain reaction sufficient to power a bomb. Uranium-235 occurs in nature, but is very rare. It accounts for less than one percent of uranium ore and has to be separated from uranium-238, the isotope that primarily constitutes uranium ore. Starting in 1940, the S-1 Section funded research projects to investigate different methods for separating or enriching uranium. One project was located at Ernest O. Lawrence's Rad Lab at the

University of California, Berkeley, where he explored the use of cyclotrons for the electromagnetic separation of uranium. Chemist Harold Urey directed another project at Columbia University. Urey's project aimed to develop gaseous diffusion as a means of separating uranium isotopes.138

1 w The ccntrifuge was considered as another means of enriching uranium. Eger V. Murphree conducted initial research at the Standard Oil Development Company Lab in New Jersey in collaboration with Jesse W. Beams at University of Virginia. Early results led researchers to believe that other methods of uranium separation were more viable. Thermal diffusion was also investigated by the Naval Research Lab. This project was independent of the OSRD and MED. Ibid., 78; Hewlett and Anderson, The New World, 96-97; Robert S. Norris, Racing for the Bomb: General Leslie R. Groves, the Manhattan Project's Indispensable Man (South Royalton, VT: Steerforth Press, 2002), 204-7, 211 -26, and 365-72; Rhodes, The Making of the Atomic Bomb, 403-5 and 533. 86

Aside from uranium, a second element was found to be highly fissionable.

Chemist Glenn T. Seaborg, a scientist at Lawrence's Rad Lab, discovered

plutonium in February 1941. An isotope of the new element, plutonium-239, was

actually more fissionable than uranium, which prompted the S-l Section to

support research on the production of plutonium as well.139 After considering a

few different sites, a plutonium project was established in January 1942 under the

leadership of physicist Arthur H. Compton at the University of Chicago.

Columbia University physicist Enrico Fermi relocated to Chicago to join the

project where he led the effort to build the first nuclear pile or reactor140 in which

plutonium could be produced.141 The plutonium project is particularly relevant to

this study of wartime biomedical radiation research because the Manhattan

Project's health and safety program was created as part of the plutonium project.

Indeed, as Director of the plutonium project, Compton formed the Health

Division to attend to the radiation hazards related to the atomic bomb project,

especially the hazards associated with plutonium and its production. We will see

that the Health Division mobilized members of a pre-existing network of radiation

researchers and reinforced the practice of interdisciplinary collaboration both in

research and applied work.

With the NDRC and then the OSRD in charge of the atomic bomb

project's uranium and plutonium projects, as well as other defense projects, the

l"V) Hewlett and Anderson, The New World, 39-42. 1411 Nuclear piles are now referred to as reactors, but the term "reactor" did not replace the term "pile" until the early 1950s. See Leland Johnson and Daniel Schaffer, Oak Ridge National Laboratory: The First Fifty Years (Knoxville, TN: University of Tennessee Press, 1994), 12. 141 Henry De Wolf Smyth, A General Account of the Development of Methods of Using Atomic- Energy for Military Purposes under the Auspices of the United States Government, 1940-1945 (Washington, DC: United States Government Printing Office, 1945), 30. 87

NAS and NRC were displaced as the government's primary agencies for scientific

research and development. The NAS and NRC continued to contribute to the war

effort and atomic bomb development by providing expert opinion when asked to do so, but they had little authority to actually direct the course of research and development. The NDRC and OSRD, on the other hand, did.142 The history of

these new agencies and the organization of Manhattan Project research within

them, are important to this study in that this history illuminates the strong link

formed between the scientific community and the state. These agencies were the central channel through which scientists were able to directly influence science policy and even defense policy. The newly created infrastructure provided the means necessary to integrate variously trained scientists and situated in locations throughout the country into research and development projects that, due to wartime circumstances, needed to progress as quickly as possible.

Just a few years after the war ended, Bush reflected on the importance of the OSRD, highlighting its function as "the medium through which, in the main, scientists were joined in effective partnership with military men. Such a partnership," Bush reckoned, "was really a new thing in the world and was a partnership between groups which one might at first thought consider inherently incompatible."143 Bush explained the incompatibility of scientists and military men arguing that the military was so rigid and structured whereas scientists could hardly be called a group due to their loose association. It is true that the scientific

142 Historian Stanley Goldberg illustrates the greater authority rooted in the NDRC and OSRD relative to the NAS and NRC. Specifically, he argues that Vannevar Bush who, as the Director of the NDRC and then the OSRD, engineered the decision to launch a full-scale atomic bomb project, despite a lack of consensus to do so amongst a NAS committee organized to help make the decision. See Goldberg, "Inventing a Climate of Opinion," 429-52. 14 3 Bush, "Foreword," in Organizing Scientific Research for War, ix. 88

community as a whole was not cohesive and that there were incompatibilities

between the military and scientific community. Bush was also right to point to

the NDRC's and OSRD's fundamental role in helping to negotiate an effective

partnership between these seemingly incompatible groups. However, radiation

researchers were not entirely unprepared to enter into a relationship with military

men. As argued in the previous chapter, scientists who conducted radiation

research earlier in the century were in the practice of collaborating with

researchers outside of their disciplinary fields. Beyond that, researchers had long

developed relationships with various patrons, mostly in the private sector, that, at

times, required researchers to tailor their work to the interests of those patrons.

The key point here is that the mobilization of science for World War II

through agencies such as the NDRC and OSRD created a political economy in

which the federal government and military were significant consumers of

scientific knowledge. Within the large-scale, defense-oriented, and federally-

funded research and development complex, biomedical radiation researchers—

especially those who had achieved both scientific and social authority through

their pursuit of socially and politically relevant interdisciplinary research

throughout the early twentieth century—had opportunities. That is, researchers

had opportunities to pursue research and take on wartime administrative duties for

the state. Biomedical radiation researchers did not serve in the upper levels of the

OSRD, but the OSRD's defense projects created other avenues for radiation

researchers to get involved in wartime research and establish themselves as valuable contributors to the vast state-funded scientific enterprise. 89

As wc will sec, some of the leading experts in radiation research,

including physicist Ernest O. Lawrence and radiologists Dr. Robert S. Stone and

Dr. Hymer L. Friedell from Berkeley, and radiologist Dr. Stafford L. Warren and

physicist Lee A. DuBridge from the University of Rochester, took on

administrative positions that involved much work outside of the laboratory.

Leading up to World War II such scientists and physician-researchers were

primarily recognized as experts based upon their work in the laboratory and their

ability to apply their research to medical practice. Due to their administrative

contributions during the war scientists were increasingly able to extend their

expertise into political and military arenas. Indeed, the administrative role played

by some researchers—biomedical radiation researchers included—throughout the

war and following, effectively put them in a position to shape research and

development policies. As sociologist Pierre Bourdieu has argued, scientific authority is derived from more than technical competency. Social power, like that acquired by key researchers involved in wartime research, is also an important source from which scientists can draw authority.144

THE ARMY TAKES CHARGE: THE MANHATTAN ENGINEER DISTRICT (MED)

When the atomic bomb project was launched as a full-scale effort at the end of 1941, Bush moved quickly to place the project under the control of the

Army and establish an effective partnership between the scientific community and

144 Pierre Bourdieu, "The Specificity of the Scientific Field and the Social Conditions of the Progress of Reason," Social Science Information 14, no. 6 (1975): 22. military.145 The creation of an organizational structure within the Army was a

process that began in the spring of 1942 and that continued throughout the

summer months. By the fall of 1942 the Manhattan Engineer District (MED) was

established within the Army Corps of Engineers and placed under the command

of General Leslie R. Groves. The history of how the atomic bomb project was

organized within the Army Corps of Engineers has been explored elsewhere.146

What is important here is that the establishment of the MED was a step toward expanding infrastructure for development purposes. Also, the selection of Groves

reflected his Army colleagues' belief that he would be able to firmly establish

Army control over the atomic bomb project, as well as ensure that scientists

maintained sufficient influence over the future development of the project.147

The transfer of control from the OSRD to the MED occurred gradually throughout the remainder of 1942 and early months of 1943. During that period of transition, the atomic bomb project transformed from a predominantly research-oriented project to a large-scale industrial project.148 With industrialization came greater hazards. Thus, Groves expanded the MED's health and safety infrastructure. He did not, however, create the atomic bomb project's health and safety program. Rather, he built upon the work already underway at the University of Chicago's Metallurgical Laboratory.

145 Compton, Atomic Quest, 105-7; Goldberg, "Inventing a Climate of Opinion," 449; Zachery, Endless Frontier: Vannevar Bush, 203. 14(1 Hewlett and Anderson, The New World, 71-83; Norris, General Leslie R. Groves, 168-83. 147 For an in-depth history of General Leslie R. Groves' role in managing the atomic bomb project, see Norris, General Leslie R. Groves. 148 The decision to place the atomic bomb project under Army control and the development of the MED's infrastructure is examined in greater detail in Hewlett and Anderson, The New World; Norris, General Leslie R. Groves', Rhodes, The Making of the Atomic Bomb. 91

PLUTONIUM PRODUCTION & RADIATION HAZARDS AT THE METALLURGICAL LABORATORY

The Manhattan Project's health and safety program did, indeed, begin at

the laboratory that was created for the purpose of building a nuclear pile or, what

is now known as a reactor, for the production of plutonium. The University of

Chicago's Metallurgical Laboratory, which was commonly referred to as the Met

Lab, was established in January 1942. Initially this laboratory was organized with

three divisions. These divisions—Physics, Chemistry, and Engineering—were

responsible for building the pile. The Health Division was formed in the summer

of 1942 to deal specifically with radiation safety and, as such, organized a

program of clinical screening, biological research, and a range of radiation safety

practices that came to be known as health physics. In 1946 one of the pioneering

health physicists recruited to work in the Health Division, Karl Z. Morgan, wrote a brief history of the Health Division. According to Morgan, the physicists, chemists, and engineers gathered at the Met Lab were prompted to establish a

health and safety program by their immediate realization that the plutonium project would result in radiation hazards of an unprecedented scale. "It was recognized at the very beginning of these efforts," Morgan wrote, "that radiation hazards of a heretofore undreamed of magnitude would be encountered."149

Morgan elaborated on the Met Lab scientists' expectations of hazards associated with the unfolding project, explaining that their understanding was rooted in the use of X-rays and radium for medical and industrial purposes in the

14' Karl Z. Morgan, "Health Physics, Its History & Development, 31 October 1946," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 46, Medicine, Health & Safety (10/31 /1946), 1. first few decades of the century. He reported that, as of 1942 when the plutonium

project was being launched, scientists knew that only two pounds of radium had

been mined and made available for human use. They were also aware that exposure to the radiation emitted from this quantity of radium had caused more

than twenty deaths and hundreds of injuries in the United States alone.150 Met

Lab leader echoed Morgan in a personal history of the project, in which he recalled, "our physicists became worried. They knew what had happened to the early experimenters with radioactive materials. Not many of them had lived very long."151

The history of injury and death related to radium exposure was worrisome enough, but scientists anticipated that plutonium would pose a much greater hazard. Those at the Met Lab projected that the radiation generated as a result of the plutonium project would be equal to the radiation emitted by millions of pounds of radium. Based on this understanding, Morgan said that some scientists doubted whether the plutonium project should proceed at all.152 Dr. Robert S.

Stone, who, as we will see became Director of the Health Division, also emphasized the uncertainty and even fear that existed amongst Met Lab scientists.

Immediately following the war Stone reported that "some of the scientists were so fearful that they doubted whether the project could or should be prosecuted."153

The doubt and immense worries that plagued Met Lab scientists probably would have prevented the Met Lab's plutonium project from going forward had it not

150 Morgan, "Health Physics..., 1946," NARA Atlanta, Box 46, 1. 151 Compton, Atornic Quest, 177. I>: Morgan, "Health Physics..., 1946," NARA Atlanta, Box 46, 1. 15' Robert S. Stone, "Health Protection Activities of the Plutonium Project, 2 November 1945," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 38, Metallurgical Lab (11/02/1945), 2. 93 been part of the Manhattan Project. In light of the war the United States and

Allied Powers were waging against Germany and Japan, government and military officials, as well as scientists, felt the project should proceed, but proceed with the utmost caution. Compton and the team of scientists he led at the Met Lab therefore continue their work toward building a pile for plutonium production, but they also took steps toward controlling radiation hazards.

The first step was the creation of a clinical screening program in February

1942, just one month after the Met Lab had been established. Met Lab physicists helped organize this program. It was a couple of the laboratory's physicists who visited the University of Chicago's Billings Hospital to ask Dr. Leon O. Jacobson, a haematologist, to run the screening program.154 When Jacobson began screening men at the Met Lab he was not told of the nature of the work being done there. Rather he was instructed to focus on particular types of clinical testing. For instance, based on prior knowledge that radiation affects the blood,

Jacobson's screening program involved much blood work. Monitoring blood counts proved to reveal little about radiation exposure, though. Results from early screening showed that normal variations in blood counts were too great for

Jacobson to determine a standard blood count against which radiation-induced changes could be measured.155

The scientists at the Met Lab felt that a better test was needed to screen for overexposure. Indeed, as Stone explained after the war, those at the Met Lab considered it "necessary to search for some other clinical test which might give an

154 Hacker, The Dragon's Tail, 29. 155 Stone, "Health Protection Activities..., 1945," NARA Atlanta, Box 38, 3; and Stone, "The Concept of a Maximum Permissible Exposure," 648. 94

indication of incipient changes."156 Thus, two factors combined to prompt

Compton's decision to create the Health Division: the Met Lab scientists' concern

for radiation hazards and their belief that the initial screening program was

inadequate as a means of monitoring radiation exposure. For Compton, the creation of the Health Division was less a decision than it was a necessity. "There was only one thing to do," Compton recalled. He and his colleagues believed that

they had to "bring to Chicago the most competent men [they] could find in the

field of the physiological effects of ionizing radiations."157 That is, in effect, what

the Met Lab aimed to do in the summer of 1942 when forming the Health

Division. Compton and his colleagues sought out researchers they hoped could investigate radiation hazards and establish a body of knowledge from which they could devise better screening practices.

Met Lab scientists believed that this research would require the involvement of variously trained researchers and physicians, which is not surprising given the history of interdisciplinary collaboration in biomedical radiation research prior to the war. According to Met Lab radiologist Dr. Hymer

L. Friedell, Compton's second-in-command, physicist Norman Hilberry, was especially keen to involve a biophysicist in the project.158 In an interview conducted in the 1990s Friedell paraphrased what, by his account, Hilberry had said to him many years earlier: "Well, doctors don't know much about this stuff,

l5h Stone, "Health Protection Activities..., 1945," NARA Atlanta, Box 38, 3. b7 Compton, Atomic Quest, 177. I5S Arthur Compton described Hilberry as one of three individuals who were integral in helping him manage the plutonium project. Compton credited him with managing all of the administrative details of the project. See /hid., 85. 95 so we've got to get a good biophysicist."159 Friedell's recollection, while perhaps affected by the passage of time, captures the general attitude felt amongst Met

Lab scientists who believed a wide-ranging research program should be an important part of the Health Division. The first scientist recruited to the Health

Division was, in fact, a biophysicist. As we will see, Kenneth S. Cole arrived at the laboratory and helped Compton and the Met Lab scientists determine how to proceed in terms of organizing and staffing the new division.

THE STRUCTURE & WORK OF THE HEALTH DIVISION

The close connection that existed between radiation research and clinical applications of radiation throughout the early twentieth century was reflected in the structure of the Health Division. Cole, as well as two of Compton's medical colleagues, Drs. Paul Hodges and Max Cutler, helped Compton organize the

Health Division with three sections—Biological, Medical, and Health Physics.160

These were designed to encompass research and applied work. Reporting on the activities of the Health Division at the end of the war, Director Robert Stone summarized the function of these different sections. The Biological Section was primarily engaged in research investigating the biological effects of radiation.

This entailed studies of different types of radiation and exposure via different

l5<) Oral History of Hymer L. Friedell, M.D., Ph. D., interview conducted September 28, 1994, as part of the Office of Human Radiation Experiments' Oral History Project, Human Radiation Studies: Remembering the Early Years. Conducted by Eleanor Melamed and Dr. Darrell Fisher at Case Western Reserve University, Cleveland, OH, n.p.. Compton recalled that the newly recruited biophysicist Kenneth S. Cole helped Met Lab scientists -Compton included—understand the health and safety problem such that they were able to determine how to proceed. Compton then enlisted the help of Professor Paul Hodges and Dr. Max Cutler—the former, from the Department of Radiology at the University of Chicago and the latter, the Director of the Chicago Tumor Institute—to select a Director and the first few recruits to the Health Division. Compton, Atomic Quest, 177. 96 means such as inhalation and ingestion.161 For instance, Cole and some of his colleagues in the Biological Section conducted animal studies to investigate the effects of breathing xenon gas. A radioisotope of this gas was a fission product that researchers suspected would accumulate around a nuclear pile and, when inhaled, concentrate in workers' lungs or tissues. The effects of this were unknown and feared until Cole's animal studies yielded reassuring results.162 In addition to research conducted at the Met Lab, some of the research organized by the Biological Section was contracted out to researchers at universities, hospitals, and research institutes. This is an issue examined in greater detail below.

The Medical Section, Stone explained, was responsible for observing Met

Lab personnel and conducting laboratory tests to monitor their health. The

Medical Section also confronted the dilemma encountered earlier by Jacobson.

That is, routine clinical tests were not a good indication of overexposure to radiation. Jacobson did, however, continue his blood studies once he was absorbed into the Medical Section. These were part of larger survey of workers' health primarily based on the study of blood and urine.163 Clinical testing of workers' health provided researchers little useful information throughout the war, thus those in the Health Division hoped that the research conducted by both the

Medical and Biological Sections would furnish physicians with knowledge from which they could derive more appropriate tests. Like the Biological Section, the

Medical Section also organized research that was contracted out to hospitals including Memorial Hospital in New York, the Chicago Tumor Institute, and the

Stone, "Health Protection Activities..., 1945," NARA Atlanta, Box 38, 3 and 17-23. "'2 Hacker, The Dragon's Tail, 43-44. 163 Ibid., 35. 97

University of California Hospital in San Francisco. Patients at these hospitals were exposed to X-rays, sometimes in massive or multiple doses, in an effort to determine the biological effects of whole body irradiation.164 The patients used in such studies were considered terminal cancer cases and, as such, researchers seemed to have had little concern about the ethical implications of purposefully exposing them to doses of radiation that were likely harmful.

Stone's report on the Health Division described the primary responsibility of the Health Physics Section as measuring radiation exposure in terms of both the quantity and quality of radiation individuals might receive. This job required health physicists to develop new or improve existing instruments for radiation detection and design plants and equipment such that safe work environments could be maintained.165 When explaining the work of health physicists Stone captured their essential qualifications saying that the health physicist "is the person best acquainted with the behavior of radiations and of personnel.. .."I66

One body of knowledge—that dealing with radiation—does not necessarily go with the other—a familiarity with personnel, their work environments, patterns of work, and hazards they might encounter. Thus, the work of health physicists drew upon the knowledge and skills of researchers and engineers from various disciplines and applied it to the problem of radiation safety. The name adopted for the Health Physics Section illuminates the hybrid nature of so much of the

Health Division's work.

IM Ibid., 42. 165 Stone, "Health Protection Activities..., 1945," NARA Atlanta, Box 38, 2. i6h Stone, "Health Protection Activities..., 1945," NARA Atlanta, Box 38, 16. 98

While each section had different responsibilities, all participated in both research and health and safety activities that contributed to, what Stone considered, a unified program.167 For instance, health physicists assisted Cole in his xenon studies by providing instruments and performing measurements of xenon gas in work environments and that inhaled by humans.168 Also, Compton recalled that individuals in the Health Physics Section worked closely with those in the Medical Section to implement what they thought to be the best clinical procedures for monitoring radiation exposure. Although he did not relocate to the

Health Division on a full-time basis, biophysicist Gioacchino Failla helped to determine screening procedures.169 As examined in the previous chapter, Failla had established an interdisciplinary laboratory at Memorial Hospital in the 1920s.

Since that time, Failla and his colleagues were accustomed to pursuing biomedical research informed by both the physical and biological sciences and closely tied to clinical applications. The hybrid body of knowledge he had developed was advantageous given the responsibilities of health physicists and all of the Health

Division's researchers in determining the biological effects of radiation and preventing hazardous exposure. Indeed, Health Division researchers aimed to achieve a better understanding of the physical properties and biological effects of radiation in relation to exposure that might occur in laboratory and industrial settings.

The structure of the Health Division allowed for a much more extensive rcsearch-based safety program to develop than what had started as Jacobson's

"'7 Stone, "Health Protection Activities..., 1945," NARA Atlanta, Box 38, 22. ",s Hacker, The Dragon's Tail, 43-44. Compton, Atomic Quest, 178. 99 clinical screening program. With numerous studies pursued in the Health

Division or in collaboration with researchers elsewhere, researchers were

particularly concerned with determining radiation safety standards. In terms of establishing safety standards, the Health Division could rely on the United States

Advisory Committee on X-ray and Radium Protection's (ACXRP) previously set standards for external exposure to X-rays and gamma rays, and internal exposure to radium and radon gas.170 Those in the Health Division were well enough acquainted with members of the ACXRP to know that the Advisory Committee's standards were devised by highly competent experts in radiology, physics, and engineering. However, Health Division researchers still considered setting standards to be one of their most pressing problems because, as Stone explained in a paper he had published in 1952, he and his Health Division colleagues were also aware that the existing standards were not based on an extensive body of research. According to Stone, the Health Division's research sought to determine

"to what amount of ionizing radiation may a person be exposed day after day without detectable damage to himself or future generations."171 Stone's articulation of this objective implied that the lack of research that informed the

ACXRP's standards was a deficit the Health Division was determined to overcome. The scale and scope of the plutonium project, specifically, and the

Manhattan Project, generally, drove a review of the existing standards and further research. The Health Division pursued research to determine, for instance,

170 "Health Physics - History & Present Program, 16 May 1950," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 31, Organization & Management, 1; Kocher, "Perspective on the Historical Development of Radiation Standards," 520-21; Stone, "The Concept of a Maximum Permissible Exposure," 641; Whittemore, "The National Committee on Radiation Protection, 1928-1960," 87 and 209. 171 Stone, "The Concept of a Maximum Permissible Exposure," 639. 100

whether the standard that existed for ingested radium might be applied to other

internal emitters, like the newly discovered and highly toxic radioactive element

that was central to the Manhattan Project—plutonium.

STAFFING THE HEALTH DIVISION

Compton and his colleagues believed the Health Division's objectives were sufficiently diverse to require the recruitment of variously trained

researchers and clinicians who would work collaboratively to ensure the health and safety of Met Lab employees. The individuals recruited to the new division

included physicists, chemists, biologists, and physicians who, in addition to being accustomed to working with researchers from other fields, possessed hybrid expertise. Due to their collaborative work before World War II, those who staffed the Health Division had already begun to develop a broad range of knowledge and skills. This was the case with the Health Division's first scientist Kenneth S. Cole who was trained in physics, but considered himself a biophysicist. Prior to the war he had been employed in the Department of Physiology at Columbia

University where he had accrued more than two decades of biophysical research experience and was a consultant in various medical branches, including radiology.172 Between his training and work experience, Cole's expertise was much like that of fellow biophysicist Gioacchino Failla. Both Cole and Failla had obtained their Ph.D.s in physics, but pursued research in and helped develop hybrid biomedical fields such as biophysics and radiobiology.173

172 David E. Goldman, "Kenneth S. Cole, 1900-1984," Biophysical Journal 47 (1985): 859. 171 On Failla's career, see Quimby, "Gioacchino Failla," 376-82. At the Met Lab, Cole assumed leadership of the Biological Section. As

the first biomedical scientist asked to join the new division, he helped recruit

others. Chicago-based radiologists Hodges and Cutler helped him with the

recruitment process.174 They suggested Dr. Simeon T. Cantril, a radiologist, for

the leadership of the Health Division's Medical Section. Cantril came to the Met

Lab from Seattle's Swedish Hospital where he was Director of the Tumor

Institute, but he was not new to Chicago. During the 1930s he had spent five

years working at both the Michael Reese Hospital and the Chicago Tumor

Institute. 1 7^' CantriPs previous employment in Chicago played a role in his

recruitment in that he was familiar to Hodges and Cutler. When Compton was in

the midst of soliciting advice from Hodges and Cutler, Cantril was visiting

Chicago. They were quick to recommend him to Compton who reported that "we

let him leave Chicago just long enough to collect his personal belongings."176

Other Met Lab employees such as radiologist Hymer L. Friedell had worked at the Chicago Tumor Institute during the 1930s. Friedell recalled in an

interview that due to his employment at the Tumor Institute he was already acquainted with some of the scientists and physicians he encountered at the Met

Lab.177 One such prior acquaintance, aside from Cantril, was Ernest O. Wollan who was selected to head up the Health Physics Section. Wollan had earned his

Ph.D. in physics in 1929. He completed his degree at the University of Chicago

174 In addition to Cole, Hodges, and Cutler, Compton credited fellow physicist Joyce C. Stearns for helping to find the most qualified individuals to staff all the divisions of the Met Lab, the Health Division included. Stearns was director of personnel at the Met Lab. See Compton, Atomic Quest, 84. 17:1 Hacker, The Dragon's Tail, 30. I7<' Compton, Atomic Quest, 178. 177 Oral History of Hymer Friedell, 1994, n.p.; and Hacker, The Dragon's Tail, 29-32 and 46. 102 where he trained under Compton. Following that, Wollan was employed at the

Chicago Tumor Institute for a number of years throughout the 1930s. 178 Before joining the Health Division Wollan was actually recruited for the Met Lab's

Engineering Division. He was given the task of monitoring for radiation leaks and designing shielding for the pile. He relocated to the new Health Division once it was established.179

The position of Director of the Health Division was first offered to Cole sincc the division was created in the spirit of expanding biological and medical research and he was the first biomedical scientist recruited to do so. He declined the position, though, and Compton then asked Dr. Robert S. Stone to become

Director. As we saw in the previous chapter, Stone had been employed as a radiologist at the University of California, San Francisco, School of Medicine since 1928. During his first decade at UCSF he had played an integral role in helping to expand radiology as a clinical program and field of research. When the

Department of Radiology was formally established within the School of Medicine he was its first Chair, a position he held until his retirement in 1962. 1 80 His position at UCSF also afforded him ample opportunity to work with the cyclotron at Ernest Lawrence's Berkeley Rad Lab, which put him at the cutting edge of research and clinical applications of radioisotopes throughout the 1930s.

I7!i Compton, Atomic Quest, 178; Regato, Radiological Physicists, 127 and 31. I7'' 1 lacker. The Dragon's Tail, 30-31. IIW Stone has been credited with helping to transform radiology at UCSF from a small program subsumed within the Department of Surgery to a full-scale research and clinical program in its own right. Malcolm D. Jones and Glenn E. Sheline, "Robert S. Stone, Radiology: San Francisco," in University of California: In Memoriam, 1968 (Berkeley, CA: University of California Academic Senate, 1968), 125. Stone did not relocate to Chicago to take charge of the Health Division

until August 1942. However, as Friedell discovered, Stone had been secretly

visiting the Met Lab on a periodic basis since the plutonium project began in

January of that year.181 Friedell was a colleague of Stone's both at UCSF and at

the Met Lab and, like Stone, Friedell relocated to Chicago to join the Health

Division in the summer of 1942. Examined further below, Friedell's employment

with the Health Division differed from most of his Met Lab colleagues in that he served at the Met Lab as an officer with the Army. Unlike Friedell, most of the scientists who filled the ranks of the Health Division were civilian scientists.

Historian Barton Hacker argues that the pattern of recruitment at the Met

Lab depended on personal connections between colleagues and the recruitment of

graduate students.182 As already noted, some pre-existing connections were

established amongst individuals who worked at the Chicago Tumor Institute. For

instance, Friedell, Cantril, and Wollan had all known each other while employed

there. Similar connections were made in Seattle at the Swedish Hospital Tumor

Institute. There, physician-researcher James J. Nickson and physicist Herbert M.

Parker had been colleagues of Cantril's. Nickson, who worked at the Tumor

Institute in Seattle as a student, joined the Health Division as part of Cantril's

Medical Section. Parker, on the other hand, went to the Met Lab to join Wollan

in the Health Physics Section. 1 83 Ernest Lawrence's Rad Lab at Berkeley was a third, and perhaps the most notable, location at which professional connections made throughout the 1930s influenced the staffing of the Health Division.

ISI Oral History of Hymer Friedell, 1994, n.p.. Ix' Hacker, The Dragon's Tail, 29-30. 1W Compton, Atomic Quest, 178; Hacker, The Dragon's Tail, 31. 104

Hacker argues that recruitment based on such previous connections was

efficient and natural, especially given that the need for secrecy prevented those

already at the Met Lab from fully disclosing the nature of the plutonium project to

potential recruits. 1 84 The role of secrecy in shaping Health Division staffing was

an issue Friedell pondered during an interview conducted many decades after the

Manhattan Project had ended. Friedell mentioned that he had always been

puzzled as to why the University of Chicago's Radiology Department was not

asked to take responsibility for the Met Lab's health and safety program. This

was an issue that he raised when discussing the selection of Stone as Director. It

is evident in the history of recruitment presented thus far, that it was not just

Stone who went to the Health Division from elsewhere. Friedell presumed that the lack of recruitment from within the University of Chicago's Department of

Radiology may have been an effort to maintain a higher level of secrecy regarding the plutonium project.185 Regardless of whether Friedell was correct in his assumption about the role secrecy played in shaping recruitment, this study argues that the recruitment of certain individuals to the Health Division was influenced by more than just an effort to obscure any connection between radiation and the activities of the Met Lab.

Building on Hacker's argument that recruitment was shaped by previously established personal connections, this study emphasizes that these connections were formed amongst a diversely trained network of researchers and clinicians who were used to collaborating across disciplines and who had developed a type

1X4 Hacker, The Dragon '.v Tail, 30-32. 1X5 Oral History of Hymer Friedell, 1994, n.p.. 105 of expertise that was hybrid in nature. Many of those who were recruited to the

Health Division had participated in interdisciplinary research prior to the war that was oriented toward developing clinical applications of radiation. That is, they embraced interdisciplinary collaboration as a means to achieve particular ends within a particular political economy of research. The selection of Stone as

Health Division Director and his reliance upon his Rad Lab connections to staff

the Health Division helps to illustrate this point. As we will see, not just who

Health Division researchers had worked with, but what they had worked on was an important factor when staffing the Health Division.

THE MET LAB CALLS ON THE RAD LAB

Due to his close professional connection to Ernest Lawrence, Stone was well positioned to be recruited for the Manhattan Project. Ernest Lawrence was, of course, one of the leading scientists involved in Manhattan Project. During the earliest stage of the project when the feasibility of building a bomb was still uncertain, Lawrence and Compton were two of a small group of scientists who not only pursued atomic bomb research, but who also influenced policy decisions related to the project.186 Their wartime and even pre-war collaboration established a professional relationship between the two scientists that facilitated

Compton's recruitment of Lawrence's Rad Lab colleagues—Stone included—to the Met Lab.

1Sh For instance, both Lawrence and Compton were on the OSRD's S-l Section and the NAS committee formed in 1941 to evaluate the technological feasibility of building an atomic bomb. See Goldberg, "Inventing a Climate of Opinion," 438; Hewlett and Anderson, The New World, 41. 106

The selection of Stone as Director of the Health Division was also very

much based on his own personal expertise. Stone's experience as a radiologist at

UCSF and his involvement in research at the Rad Lab set him apart from many

radiologists that might otherwise have been recruited, at least according to

Compton. Based on his knowledge of the activities of the Rad Lab throughout the

1930s, Compton believed that Stone had a breadth of knowledge and collaborative experience that not all radiologists possessed. I 87 It seems natural that Compton and his Met Lab colleagues would seek out an individual with such experience given that as Director of the Health Division Stone was required to work with those in the Physics, Chemistry, and Engineering Divisions at the Met

Lab. Stone was appointed to the Laboratory and Project Councils, the expectation being that Stone would stay abreast of developments taking place in the other divisions and try to anticipate dangers rather than address them after the fact.188

Once Stone joined the Health Division, the pattern of recruitment was even more likely to depend on professional connections made within Ernest

Lawrence's Rad Lab. As already mentioned and examined in greater detail below, Friedell had worked with Stone at UCSF and the Rad Lab before he was called up for military duty and stationed at the Met Lab. Friedell was not the only

Rad Lab researcher that joined Stone in the Health Division. Former Rad Lab biophysicist Waldo Cohn attributed his own recruitment to Kenneth Cole's

Biological Section to his prior connection to the Rad Lab. In an interview conducted decades after the war, Cohn recalled that Stone relied on his colleagues

l!" Compton, Atomic Quest, 84 and 178. ISS Stone, "Health Protection Activities..., 1945," NARA Atlanta, Box 38, 1-2. 107

from Berkeley to help staff the Health Division. Cohn explained that when he

was recruited to the Met Lab in 1942 he was a few years removed from the Rad

Lab. He had left Berkeley in 1939 to accept a position at Harvard University's

Medical School, but his former colleagues Drs. John Lawrence and Joseph

Hamilton recommended him to Stone.189 Hamilton was also asked to join the

Health Division, but unlike Cohn, he did not relocate to Chicago. Rather he accepted a contract to conduct research for the Health Division and pursued this

work at his own laboratory, the Rad Lab's newly constructed Crocker Laboratory.

MOBILIZING A LABORATORY: THE RAD LAB'S DEFENSE RESEARCH

While the Rad Lab's pre-war research enterprise made it an ideal site from which to recruit radiation researchers, for the same reason, it was also an ideal site at which to conduct research. Two important projects were established at the Rad

Lab as part of the Manhattan Project, both of which drew upon the expertise of the variously trained researchers employed by or associated with the Rad Lab.190

The first was, as previously mentioned, Ernest Lawrence's investigation and development of electromagnetic separation of uranium. For Lawrence, personally, this project put him in a position to influence top policy makers

including, Vannevar Bush, James Conant, and, once the Army was involved,

IIW Oral History of Waldo E. Cohn, Ph. D., interview conducted January 18, 1995, as part of the Office of Human Radiation Experiments' Oral History Project, Human Radiation Studies: Remembering the Early Years. Conducted by Thomas Fisher, Jr. and Michael Yuffee at Colin's home. Oak Ridge, TN, n.p.. 190 John Lawrence also accepted a wartime research contract to investigate decompression sickness. This was not part of the Manhattan Project, but like Ernest Lawrence's and Joseph Hamilton's Manhattan Project research, John Lawrence's wartime research drew upon the expertise gained through interdisciplinary collaboration. It also helped to maintain connections between radiation researchers, and reinforce the scientific and social authority of Rad Lab researchers. See Williams, "Donner Laboratory," 18N. 108

General Groves too. For the rest of his laboratory, the electromagnetic separation

project generated an abundance of work and reinforced the practice of

interdisciplinary research, especially amongst physicists, chemists, and engineers.

Dr. Joseph Hamilton also conducted Manhattan Project research at the

Rad Lab—research that investigated the metabolism of radioisotopes. He did so under contract with the Met Lab's Health Division. With the Manhattan Project's various projects focused on procuring uranium-235 and plutonium-239, the

Health Division was particularly concerned with the health effects of these radioisotopes and the fission products generated during their production.191 Given his pre-war research with cyclotron-produced radioisotopes like phosphorous and

iodine, Hamilton had gained experience relevant to his wartime work. His recruitment, however, was based on more than experience alone. Stone was very familiar with Hamilton's research experience and the resources Hamilton had available to him at the Crocker Lab because the two had pursed research together years earlier.192 Years after the war, biophysicist Patricia Durbin voiced her opinion that Stone's and Hamilton's pre-war collaboration was, indeed, important to Stone's recruitment of Hamilton. Durbin, who trained under Hamilton starting in the mid 1940s and who stayed on as an employee at the Crocker Lab through to

1977, recalled in an interview that, "Stone knew Hamilton professionally. He knew he was an enthusiast in terms of the uses of radioisotopes, applications of isotopes. . . .There were detection devices available to Hamilton and the people

Hacker, Elements of Controversy, 35. 1,2 For instance, prior to the war Stone and Hamilton collaborated to investigate the metabolism of radioactive sodium in humans. See J. G. Hamilton and R. S. Stone, "Excretion of Radio-Sodium Following Intravenous in Man," Proceedings of the Society for Experimental Biology and Medicine 35 (1937): 595-98. who worked with him. Detection devices were rare. They were only present in a few laboratories in a few people's hands." Durbin continued, explaining her

perspective that based on Stone's prior relationship with Hamilton and the Rad

Lab, including the Donner and Crocker Laboratories, "It was an absolutely natural

match for Stone to go to Hamilton and request that fission products be made on the cyclotron, that their fate in [mammals] be studied in small animals, one by one, in a systematic way."193

Hamilton's group at Crocker examined the metabolism of fission products in rats, completing basic studies of many radioisotopes by 1944. Research on plutonium, though, was only just then beginning. From the time Seaborg discovered plutonium in February 1941, through the establishment of the plutonium project at the Met Lab a year later and the laboratory's ongoing effort to create a nuclear pile in which plutonium could be produced, little was known about the new element.194 At the end of the war, Stone spoke about the lack of knowledge scientists had about plutonium and other radioactive elements. He explained that most fission products were elements that had not been previously used in biology and medicine.195 Dr. John Gofman, a physician and nuclear chemist, also emphasized this point decades later during an interview. Speaking candidly, he said, "we knew damn well what we didn't know."196

' Oral History of Patricia C. Wallace Durbin, Ph. D., interview conducted November 11, 1994, as part of the Office of Human Radiation Experiments' Oral History Project, Human Radiation Studies: Remembering the Early Years. Conducted by Dr. Darrell Fisher (health physicist) and Marisa Caputo, Berkeley, CA, n.p.. '''4 Stone, "Health Protection Activities..., 1945," NARA Atlanta, Box 38, 17-23. 19:1 Stone. "Health Protection Activities..., 1945," NARA Atlanta, Box 38, 19-21. 116 Gofman worked under Seaborg, although Gofman remained at the Rad Lab whereas Seaborg went to the Met Lab to lead the Chemistry Division. See Oral History of John W. Gofman, M.D. Ph. D., interview conducted December 20, 1994, as part of the Office of Human Radiation Throughout the first two years of the plutonium project there was simply

not enough plutonium available to conduct any investigation of its biological effects. It was not until February 1944 that the pilot plant at Oak Ridge had produced enough plutonium to send Hamilton a sample for study. Hamilton's

initial investigation led him to believe that plutonium bore many similarities to

radium in that both concentrated in the bone and emitted alpha radiation. There were, however, differences found between the two elements that confounded

Hamilton's and his colleagues' attempts to determine safe limits for exposure to plutonium. For instance, the absorption and retention of plutonium and radium in different organs varied, as did the rate at which both substances were excreted.

Plutonium was found to be excreted more slowly than radium which, combined with other factors, made plutonium approximately five to ten times more toxic than radium.197

Reflecting on the Manhattan Project, Durbin considered Hamilton's research to be among the most important of any effort made towards radiation protection. She explained and AEC reports corroborate that Hamilton's metabolism studies, along with research done on uranium toxicology at the

University of Rochester, built the foundation for post-war radiation exposure standards. 198 That is, the successor to the Advisory Committee on X-ray and

Radium Protection—the National Council on Radiation Protection and

Measurements (NCRP)—devised standards for internal emitters based on the

Experiments' Oral History Project, Human Radiation Studies: Remembering the Early Years. Conducted by Loretta Hefner and Karoline Gourley, San Francisco, CA, n.p.. 1)7 For more information on the results of the plutonium studies done by Hamilton, see Stone, "Health Protection Activities..., 1945," NARA Atlanta, Box 38, 17-23; and Hacker, The Dragon's Tail, 62-63. ™ Oral History of Patricia Durbin, 1994, n.p.. 111

research Hamilton conducted during and immediately following the war. AEC

records show that Hamilton, acting on behalf of the AEC, collaborated with the

NCRP's Sub-Committee on Tolerance Doses and the National Research

Council's Radiobiology Committee in an effort to determine safe tolerance doses

for various radioisotopes. He drew on his wartime and early postwar research on

radioisotopes to provide relevant information.199

RADIATION SAFETY & THE ARMY: FROM ARMY PARTICIPATION IN THE HEALTH DIVISION TO THE MED'S MEDICAL OFFICE

As already discussed, the plutonium project began under the authority of the OSRD. At the same time that the Health Division was being formed in the summer of 1942, the Manhattan Engineer District (MED) was also created. The

MED did not immediately assume authority over the atomic bomb project's health and safety work, but the Army did have a small presence in the new division. During the early days of Health Division activity Dr. Hymer L. Friedell acted as a liaison between the Army and the Health Division. Friedell was not recruited as a civilian scientist. Rather, he was a reserve officer who, when called upon to serve, was told to report to the Met Lab.200 Friedell arrived at the Met

Lab at about the same time as Stone, in August 1942. Once there, he was

1,9 Karl Z. Morgan, "Meeting of the Sub-Committee of the National Committee on Radiation Protection Dealing with Permissible Tolerance Dose, 2 July 1948," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 39, Manager's Report - Info on Radiation History, 1-5. 200 According to Friedell, he had expected to be called up for active duty earlier than he was. A few months into 1942, Friedell realized that he had been placed on an "essential list" that exempted him from military service due to his teaching duties at UCSF. He felt obliged to fulfill his duty as a reserve officer and requested that he be removed from the list. See Oral History of Hymer Friedell, 1994, n.p.. 112

integrated into the Health Division under Stone's direction. Having worked under

Stone at UCSF, his was an easy transition.

In an interview conducted in the 1990s, Friedell described his wartime

activities as he remembered them. His first responsibility was to do "medical

surveillance." He explained that he examined numerous individuals at the Met

Lab learning of previous health conditions and even making some diagnoses

based on his observations. According to Friedell, his purpose was "to make sure

that we knew something about their physical condition. . . .We were doing

surveillance, doing basic studies, making some estimates of how we would

measure the radiation."201 His examination of Met Lab workers followed in the

same vein as Jacobson's earlier screening program. Friedell also recalled that he

visited hospitals where, for years, cancer patients had been treated with total-body

irradiation. Similarly, he visited companies that had used radium, all in an effort

to obtain information that might be pertinent to understanding the radiation

hazards created by the plutonium project.

In the absence of established practices for measuring radiation exposure,

those in the Health Division considered extensive clinical testing or surveillance, as Friedell described it, to be a necessary precaution. For instance, shortly after the war Stone explained that he and his medical colleagues had not been sure whether they should make blood counts mandatory, but had ultimately decided to do so. Stone's feeling that taking blood counts was a necessary precaution rather than a sure method to obtain desired information was evident. Indeed, Stone

201 Oral History of Hymer Friedell, 1994, n.p.. 113

admitted that even after the war he and his colleagues were not sure that they

could find anything of significance in the records they collected.202

Friedell's involvement in the Health Division at the Met Lab began in the summer of 1942 and continued for approximately one year. His tenure at the Met

Lab coincided with the transfer of control over the different projects associated

with atomic bomb development from the OSRD to MED. During that time the

Manhattan Project also transitioned from being a primarily research-oriented

endeavor pursued in academic laboratories to a large-scale industrial enterprise.

The MED undertook construction of the Clinton Engineer Works in Oak Ridge,

Tennessee, Hanford Engineer Works in Hanford, Washington, and the Los

Alamos Laboratory in Los Alamos, New Mexico in 1943. Clinton was a vast

industrial site that included electromagnetic and gaseous diffusion facilities for the separation of uranium. It also included a pilot plant for a plutonium- producing nuclear pile, known then as Clinton Laboratories, but now as Oak

Ridge National Laboratory. Clinton Laboratories was a pilot plant for the full- scale plutonium production facilities being constructed at Hanford, an industrial site even larger than Oak Ridge. The Los Alamos Laboratory was constructed specifically as a bomb laboratory. The fissionable material produced at Oak

Ridge and Hanford was sent to Los Alamos where the atomic bombs were fashioned.203 When construction of these facilities proceeded, the scale of the

Manhattan Project grew and so too did the radiation hazards.

2112 Stone, "The Concept of a Maximum Permissible Exposure," 648. "

The quantity of radioactive materials produced as a result of the scaled-up uranium separation and plutonium production facilities increased. Furthermore, the number of people involved in the Manhattan Project multiplied tremendously reaching 130,000 at the height of the war.204 Most of the individuals involved in the Manhattan Project were employed by universities, such as those already discussed or companies under contract to build and operate the weapons production facilities. Indeed, construction and operation of these facilities was contracted out to major firms including, Stone & Webster Engineering

Corporation and I. E. du Pont de Nemours and Company, which brought significantly more people into the project. As the atomic bomb project transformed into a very large-scale industrial enterprise under Army control,

General Groves decided to expand the health and radiation safety infrastructure within the MED.

STAFFORD WARREN & THE UNIVERSITY OF ROCHESTER JOIN THE FIGHT

Building upon the activities already pursued by the Health Division,

Groves authorized the creation of a Medical Office within the MED. Since

Friedell was the first and only medical officer specifically assigned to the MED, the Army consulted with him regarding the selection of an individual to be Chief of the new Medical Office. Friedell considered himself too junior to assume such a responsibility, but accepted the position of Deputy Chief.205 Dr. Stafford L.

Warren, a radiologist at the University of Rochester, asked to be Chief. Warren

JM Hughes, The Manhattan Project, 9. At the end of the war approximately 44, 000 individuals were transferred to the AEC. See Haeker, Elements of Controversy, 10-11; Hewlett and Anderson, The New World, 2. 2115 Oral History of Hymer Friedell, n.p.; and Hacker, The Dragon's Tail, 49-50. 115 was considered to be a leader—held "in high regard.. .by his colleagues in the field of radiation."206 We saw in the previous chapter that Warren had pursued research with X-rays since the 1920s and he was instrumental in helping the

University of Rochester acquire its first cyclotron. He had actively sought to collaborate with physicists at the University and individuals from the local industries to achieve this goal. Among other factors, his involvement in local industry influenced the decision Groves made to recruit Warren for the new position. The operating arm of the Rochester-based Eastman Kodak company—

Tennessee Eastman Corporation—received the contract for the electromagnetic separation plant at Oak Ridge. Eastman Kodak sought to involve Warren in their

MED work and engaged him as a consultant. Warren was first brought into MED operations at the end of 1942 as a civilian consultant. By the end of 1943, though, his position as Chief of the MED's Medical Office was finalized and the Army

?f)7 arranged for him to have a commission as Colonel in the Army Medical Corps.

Warren had been at the University of Rochester for almost twenty years when he was recruited for war work. His relationship to the University influenced the development of MED-supported radiation research at the University during the war and following. The University of Rochester accepted contracts to do research for the MED Medical Office that amounted to an extensive research program known as the Rochester Atomic Energy Project (RAEP). Warren and

Friedell began planning the University's wartime research in April 1943 and

:0h J. Lowell Orbison and James W. Bartlett, "Administration: A Means to an End," in To Each His Farthest Star: University of Rochester Medical Center, 1925-1975, ed. John Romano, et al. (Roehester, NY: The University of Rochester Medical Center, 1975), 233. :"7 Hacker, The Dragon's Tail, 49; Neuman, "Radiation Biology and Biophysics," 279-81. 116

Warren directed the program until he relocated to Oak Ridge at the end of 1943.

At that point, his colleague, radiologist Andrew H. Dowdy, took over as Chair of the Radiology Department and Director of the RAEP.20X

The creation of the RAEP was, in part, a reflection of the fact that once

Warren became Chief of the MED Medical Office, he was in a position in which he could choose to grant contracts to researchers and institutions he thought capable of conducting biomedical radiation research. Just as Stone relied on his old contacts at Berkeley when recruiting individuals to the Health Division or arranging research contracts with researchers at the Rad Lab, Warren relied on his colleagues at the University of Rochester. Beyond Warren's connection to the

University, his colleagues there did have research experience he deemed appropriate to assume responsibility for biomedical radiation research related to atomic bomb development. Indeed, Warren had not been the only physician- researcher or scientist at the University of Rochester to engage in biomedical research with X-rays and radioisotopes prior to World War II. Pre-war radiation research, especially that which made use of the cyclotron and cyclotron-produced radioisotopes, involved collaboration within and across biomedical and scientific disciplines. For instance, Dr. George H. Whipple, who was a pathologist and the founding Dean of the University's Medical School, collaborated with Warren to investigate X-ray toxicity. Whipple also collaborated with two other Rochester colleagues, biophysicist William F. Bale and physiologist Paul F. Hahn, as well as

J)S Neuman, "Radiation Biology and Biophysics," 280-81. 117

Berkeley physicist Ernest O. Lawrence, to conduct tracer studies of the

metabolism of radioactive iron.209

Stafford L. Warren c. early 1940s. Photo Credit: University of Rochester Medical Center Historical Photographs - Edward G. Miner Library.

George H. Whipple with students c. 1953. Photo Credit: University of Rochester Medical Center Historical Photographs - Edward G. Miner Library.

The research begun at the University of Rochester during the war was

primarily conducted in a "temporary" building now referred to as the B-Wing of the Annex. This was built north of the Medical School in 1943. It adjoined the

A-Wing, which was built during the previous year, for military research unrelated to the RAEP. Research for the RAEP was also carried out in various departments

Hahn et a!., "Radioactive Iron," 739; Warren and Whipple, "Intestinal Lesions and Acute Intoxication," 741; Warren and Whipple, "Hard Rays in the Living Organism," 731. 118 of the Medical School and Faculty of Science housed in existing buildings, as well as a new addition to the west end of the Rochester Medical Center—the Q-

Wing—which was constructed earlier in 1941. 210 There were more than six hundred individuals working in these facilities at the height of the war and, as we will see in a later chapter, the program continued under AEC leadership following the war.211 J. Newell Stannard, a University of Rochester physiologist and biophysicist who became the Assistant Director of Education for the Rochester

Atomic Energy Program following the war noted that the MED laboratory at

Rochester was unique. He explained that, "It differed from most of the other

Manhattan District labs, in that it was entirely biomedical and research-oriented.

It really had little to do with the actual development of the atomic bomb."2 1 2

Rochester Medical Center c. 1931. Photo Credit: University of Rochester Medical Center Historical Photographs, URMC Edward G. Miner Library.

"H) Neuman, "Radiation Biology and Biophysics," 280. 211 Andrew H. Dowdy, "Proposed Educational Program: Atomic Energy Project, The University of Rochester, 10 June 1947," NARA College Park, RG 326, General Correspondence 1946-1951, Box 79, University of Rochester (6-10-47), 1. 2I~ J. Newell Stannard, "Sketch of the Rochester Project," in ./. Newell Stannard and the University of Rochester: A Collection of Papers Presented at the 48th Annual Meeting of the Health Physics Society, ed. J. Newell Stannard (San Diego, CA: Health Physics Society, 2003), 7. 1 19

Rochester Medical Center c. 1961, edited for wartime additions. The new wing in the bottom left was the Q-Wing constructed in 1941. The building in the top center was the "temporary" building constructed in 1943. Photo Credit: University of Rochester Medical Center Historical Photographs, URMC Edward G. Miner Library.

The primary focus of RAEP researchers was to study the toxicity of uranium. When the University of Rochester's uranium toxicology research was published after the war had ended it was described as "one of the most complete and thorough toxicology studies of any one element."213 Rochester scientists and radiologists pursued other research including, but not limited to, studies of other elements like radium, plutonium, and polonium, investigations of the effects of X- rays on mice, and instrument development. Late in the war the University of

Gilbert B. Forbes, "A Half Century of Medical Science at Rochester: Selected Publications of Faculty and Students," in To Each His Farthest Star: University of Rochester Medical Center, 1925-1975, ed. John Romano, et al. (Rochester, NY: The University of Rochester Medical Center, 1975), 428: Harold C. Flodge and Carl Voegtlin, eds., Phannacology and Toxicology of Uranium Compounds, 4 vols., National Nuclear Energy Series (New York: McGraw-Hill Book Co., 1949- 1953); Stannard, "Sketch of the Rochester Project," 7-8. Rochester was one of the primary institutions that injected plutonium, uranium,

and polonium into humans to study the metabolism of these elements in the

human body, an issue examined in greater detail below.

WARREN & FRIEDELL AT THE HELM OF THE MED'S EXPANSION OF HEALTH & SAFETY

With the creation of the Medical Office and appointment of Warren as

Chief and Friedell as Deputy Chief of this new office, both men relocated to Oak

Ridge. Together, Warren and Friedell were responsible for ensuring the safety of all MED personnel or those with whom the MED had contracts. Their broad

mandate entailed that they determine the health hazards created by MED operations and institute whatever safety measures they deemed necessary to guard against them. To determine hazards the MED supported the Health Division's research program and associated research projects such as the work of Dr. Joseph

Hamilton at Berkeley and the Atomic Energy Project at the University of

Rochester. Aside from research, the MED Medical Office placed particular focus on monitoring work environments and personnel both at institutions where atomic

bomb research and development was ongoing and at the newly constructed MED sites.

Warren and Friedell helped recruit scientists and physicians to serve as health physicists at the various Manhattan Project sites. Shortly after Warren took up his position with the MED he visited the University of California, San

Francisco, to recruit an individual to conduct radiation safety work at Berkeley.

Warren asked Dr. Earl R. Miller to establish a Radiation Safety Division there.

Miller was, at the time, Chair of the Department of Radiology at UCSF. He had 121 taken over as Chair when Stone took a leave of absence to go to the Met Lab.

Warren's selection of Miller for this job was, like so many recruitment decisions, influenced by existing personal and professional connections. Miller and Stone, of course, knew each other well having worked together in the Radiology

Department at UCSF. Furthermore, as Miller recalled in an interview, he and

Warren were also well acquainted such that Miller considered Warren to be his

"close friend."214 The close proximity between the San Francisco and Berkeley campuses allowed Miller to devote almost two days of every week to radiation safety work at Berkeley while continuing his duties as Chair of the Radiology

Department. In terms of his radiation safety work, it was similar to the work done by those in the Medical and Health Physics Sections at the Health Division. He conducted routine blood testing on the Manhattan Project's Berkeley scientists, checked the radiation detection equipment used in the laboratories, and did general safety checks in any area where Manhattan Project work took place.

Miller's responsibilities for radiation safety at Berkeley were relatively small as compared to the immense health and safety challenges created by the construction of the MED's large-scale production facilities. As the vast industrial sites were constructed at Oak Ridge, Hanford, and Los Alamos, Warren and

Friedell oversaw the creation of health and safety organizations at each location.

New health and safety organizations expanded upon activities pioneered at the

Health Division, although they were given different names and their

"'4 Oral History of Earl R. Miller, M.D., interview conducted August 9 and 17, 1994, as part of the Office of Human Radiation Experiments' Oral History Project, Human Radiation Studies: Remembering the Early Years. Conducted by Anna Berge and Gregg Herken, San Rafael, CA, nP- 122 responsibilities were tailored to address specific hazards existing at each facility.

A history of wartime health physics prepared by AEC personnel in 1950 described the health and safety organizations established at the new industrial sites as such: "Each [was] the particular contractor's instrument for fulfilling his responsibility of safeguarding his employees and the people who reside in the area."215 The report explained that to staff these new organizations, the MED essentially initiated a process of duplication. That is, the MED duplicated the

"nucleus" of the Manhattan Project's health and safety infrastructure—the Met

Lab's Health Division.216

The process of duplication began in the summer of 1943. It was at that point that Oak Ridge replaced Berkeley as the centre of activity for electromagnetic separation. Aside from facilities for the electromagnetic separation of uranium, the Oak Ridge site also included a gaseous diffusion plant, also for uranium separation, and a pilot plant for plutonium production. To attend to the health and safety concerns arising from these various facilities, the MED created a Health Physics Division. The Health Physics Division was appropriately named because, like the Met Lab's Health Physics Section, the Oak

Ridge program placed greater emphasis on personnel protection than on research.

At the core of the new division were a number of individuals who transferred from the Met Lab's Health Division. Among them were Waldo E. Cohn, Herbert

M. Parker, Simeon T. Cantril, and Karl Z. Morgan. Although the Oak Ridge program was more focused on personnel protection, it included a biology research

215 "Health Physics..., 1950," NARA Atlanta, Box 31, 7. 216 "Health Physics..., 1950," NARA Atlanta, Box 31, 1. 123

group that was led by biophysicist Howard J. Curtis. Curtis1 recruitment, again,

illustrates the ongoing influence of personal and professional connections. He

was a former colleague of Kenneth Cole's at Columbia University and, thus, was

a natural choice to lead a biological research program that was modelled on

Cole's Biology Section at the Met Lab.217

In terms of personnel protection, Oak Ridge required a larger number of

health physicists than had been assembled at the Met Lab. They performed

numerous surveys of the Oak Ridge plants using approximately twenty different

instruments to detect radiation in the air and on workers' clothing. Beyond that,

they measured the scattering of radioactive materials around the nuclear piles and

created shielding to protect against it, consulted on various issues, such as

radioactive waste disposal, laundering contaminated clothing, new designs for

piles and other technologies, and much more.218 Parker was the first Director of

the Oak Ridge Health Physics Division and, under his leadership, the program

trained numerous physicists and engineers to fill the ranks of the health physics

program. Personnel protection was especially urgent when the plutonium pilot

plant started to operate in November 1943. When operation began, Oak Ridge

health physicists checked the reactor for leaks inch-by-inch.219

:'7 Hacker, The Dragon's Tail, 53. 2IS Morgan, "Health Physics..., 1946," NARA Atlanta, Box 46, 2-3. "|l> "Health Physics..., 1950," NARA Atlanta, Box 31,2. 124

Removal of radioactive material from the graphite reactor at Oak Ridge, c. 1946. Photo Credit: Oak Ridge National Laboratory Historical Photo Gallery.

Oak Ridge also trained health physicists who would be sent off to Hanford which was the MED's main plutonium production facility. Just as a number of individuals had transferred from the Met Lab's Health Division to Oak Ridge in the summer of 1943, many transferred from Oak Ridge to Hanford throughout

1944. Parker and Cantril were among the Oak Ridge employees to relocate to

Hanford. When Parker departed in the spring of 1944 Morgan took over as the

Director of the Health Physics Division at Oak Ridge. To replace Cantril, however, the MED recruited from outside its existing staff. Once again, the pattern of recruitment depended upon a professional connection. One of Cantril's former colleagues from the Seattle Tumor Institute, Dr. John E. Wirth, filled the position that Cantril left vacant at Oak Ridge.220

At Hanford, the health and safety organization was known as the Health

Instruments Division. The Hanford group was larger than both the Met Lab's

Health Division and the Health Physics Division at Oak Ridge. By the end of the war Hanford's Health Instruments Division included about 150 individuals, whereas the Met Lab's Health Division and the Health Physics Division at Oak

Cantril and Wirth worked together as radiologists at the Seattle Tumor Institute. Hacker, The Dragon's Tail, 54. 125

Ridge operated with about 30 and 70 individuals, respectively. Morgan described

Hanford's Health Instrument's Division as the least research-oriented

organization of the three. He characterized the Hanford organization as being the

"peak of perfection" regarding service problems that included personnel

protection, instrument development, and workplace/equipment monitoring.

Morgan considered Hanford's Health Instrument's Division to be at one end of a

spectrum and placed the Met Lab's far more research-oriented Health Division at

the other end. According to Morgan, the Health Physics Division at Oak Ridge

was a compromise between the two extremes.221

The bomb laboratory at Los Alamos also created a health and safety

organization in the spring of 1943 that was known as the Radiologic Safety

Group.222 Dr. Louis H. Hempelmann, Jr. was chosen to direct the program.

Hempelmann was a physician who had been working with Drs. John Lawrence

and Roberts S. Stone conducting biomedical radiation research since 1941. Like

so many others who were recruited for health and safety work, Hempelmann was

very much a part of a network of radiation researchers established before World

War II. Indeed, Hempelmann's recruitment was influenced by his prior

collaboration with Stone, the Health Division Director. He was also acquainted

221 Morgan, "Health Physics..., 1946," NARA Atlanta, Box 46, 2. "" There is some discrepancy regarding the name of the health and safety organization established at Los Alamos. An Atomic Energy Report generated in 1950 refers to the organization as the Radiologic Safety Group whereas Hacker refers to it as both the Health Group and Health Division. It seems likely that the organization was renamed as it was periodically reorganized. See "Health Physics..., 1950," NARA Atlanta, Box 31,6; Hacker, The Dragon's Tail, 60; and Hacker, Elements of Controversy, 44. 126

with J. Robert Oppenheimer, the Berkeley physicist who was selected to direct the

Los Alamos Laboratory.223

Historian Barton Hacker reports that Hempelmann planned to focus the

Radiologic Safety Group on personnel protection through close monitoring of Los

Alamos scientists and their work environment. He did not intend to conduct

research.224 That said, Hempelmann and his group relied upon and influenced

research done elsewhere, especially at the Health Division. Furthermore,

Hempelmann, Oppenheimer, and a biochemist on staff, Wright H. Langham,

played an important role in designing and gaining approval for metabolic studies

of plutonium in humans.225 The decision to proceed with human

experimentation—a decision authorized by General Groves—was made shortly

after a laboratory accident in August 1944 in which a Los Alamos scientist was

exposed to plutonium.226 Dr. Joseph Hamilton had conducted metabolic studies

of plutonium at the Crocker Laboratory earlier in 1944, but his test subjects were

animals. Despite the knowledge generated from this research, Los Alamos

researchers became increasingly concerned that they had no accurate means of

measuring plutonium retention in the human body.227

Hacker's discussion of wartime radiation research involving humans is

very brief, but his book was published prior to the creation of a government

committee in 1994 that had the express purpose of examining human radiation

experiments and collecting government documents related to this subject. The

Hacker, The Dragon's Tail, 59. Ibid.. 59-61. "5 AC'HRE, The Final Report, 140-48. I hid.. 139-42. "7 Hacker, The Dragon '.v Tail, 62-68. 127

Advisory Committee on Human Radiation Experiments (ACHRE) published a comprehensive report that documents the MED's human radiation experiments in greater detail than Hacker's. According to the Advisory Committee, following much discussion amongst the researchers at Los Alamos and the MED's central administration, including General Groves and Drs. Stafford Warren and Hymer

Friedell, a series of plutonium injections were conducted with hospital patients.

The first injection took place in April 1945 at the Oak Ridge Hospital. Between

1945 and 1947 seventeen more injections were made, with three each at the

University of Chicago and Berkeley, and eleven at the University of Rochester.

Researchers at the University of Rochester's Atomic Energy Project also injected six patients with uranium and five with polonium for similar metabolic studies.228

The ACHRE report notes that the MED's wartime radiation experiments with humans posed little risk to the subjects involved because only trace amounts of plutonium, uranium, and polonium were used. However, these experiments have come under considerable scrutiny, primarily because the patients who were used as test subjects were not informed that they were subjects of an experiment.

Informed consent was not, however, common at the time.229 That said, criticism also stemmed from the fact that the injections were in no way perceived to have any therapeutic benefit that might alleviate the illnesses or injuries for which these patients had been hospitalized. The issue of human experimentation was controversial then and remained so in the postwar years as the AEC developed its biomedical program.

^ ACHRE, The Final Report, 142-52. Ledcrer, Subjected to Science; David J. Rothman, Strangers at the Bedside: A History of How- Law and Transformed Medical Decision Making (New York: Aldine de Gruyter, 1991). These wartime experiments are relevant to this study in that they draw

attention to the nature of research conducted during the war, as compared to that

pursued throughout the earlier decades of the twentieth ccntury at places like

Memorial Hospital in the 1910s and following, or at Berkeley in the 1930s. There

are some obvious differences, however. An important difference was, of course,

that when radiation was administered to humans in an experimental context

earlier in the century it was often done so with the intent of shrinking a canccrous

tumor or treating blood conditions such as leukemia and polycythemia vera.

Beyond the specific effects on the individual subjects treated with radiation, pre­

war biomedical radiation experimentation was driven by the hope that researchers

could develop useful medical applications of radiation. During the war, however,

radiation administered to humans was not meant to benefit the subjects. Also, the

knowledge generated was meant to inform safety rather than medical practices.

Despite the different motivations for administering radiation to humans

prior to and during the war, there was some continuity between the research

pursued in these different contexts. When evaluating the risk of the proposed

experiments, researchers explicitly compared the plutonium injections to prior

research done with radium.230 This comparison shows that Manhattan Project

researchers drew on previous experience to design and execute these experiments.

Furthermore, as was the case with so much of the Manhattan Project's health and safety research, the MED tapped into the already established network of

researchers and resources that pre-existed the Manhattan Project. Indeed, of the

21,1 ACHRE, The Final Report, 143; Morgan, "Health Physics..., 1946," NARA Atlanta, Box 46, 129

eighteen plutonium injections, eleven were performed at the University of

Rochester and three at Berkeley, both of which were universities already engaged

in biomedical radiation research when they got involved in Manhattan Project

research. These universities had interdisciplinary teams of researchers in place

that included physician-investigators and biomedical scientists whose work

supported the establishment of new programs in biophysics and, after the war,

radiobiology and medical physics. The broad expertise embodied in these

institutions was an asset in terms of conducting the wartime human radiation

experiments as safely as possible and when interpreting data produced from the

experiments. Thus, the wartime human radiation experiments show how research

was redirected to address defense-oriented safety concerns during the war, but

they also show continuity in the body of knowledge researchers drew upon and

the networks of researchers and institutions involved.

CONCLUSION

Despite the new wartime defense motives that drove the Manhattan

Project's health and safety operations and the scale of that enterprise, wartime

radiation safety research built on the biomedical radiation research and clinical applications of radiation of the early twentieth century. The program that developed at the Health Division was rooted in the knowledge, emerging forms of hybrid expertise, and interdisciplinary research practices already established by an existing network of radiation researchers. Furthermore, the establishment of health and safety organizations at each of the new MED facilities illustrates how the wartime experience of radiation research and safety was one of both 130 continuity and growth within a political economy of research that was reorganized

to prioritize defense objectives. The programs established at Oak Ridge, Hanford, and Los Alamos continued and expanded the Health Division's activities at the

University of Chicago. Indeed, starting in 1942 the researchers assembled at the

Health Division identified problems and began to design research and safety protocols that established a foundation upon which those who were transferred or newly recruited to Oak Ridge, Hanford, and Los Alamos could build.

This chapter has demonstrated that the Manhattan Project tapped into a network of radiation researchers when recruiting personnel. While this network was not a formal entity, it had begun to form much earlier in the century as physicists, radiologists, biophysicists, chemists, and others engaged in interdisciplinary collaboration to push the boundaries of research and develop applications of radiation. Throughout the war, existing professional connections facilitated the recruitment of researchers who were used to working collaboratively to investigate problems that required a range of expertise.

Numerous individuals, including, but certainly not limited to, Drs. Robert Stone,

Joseph Hamilton, Stafford Warren, and Hymer Friedell, were able to adapt the experience they gained from interdisciplinary research projects throughout the

1920s and 1930s to the challenges they faced within the Manhattan Project.

Further, they were able to transfer some of the knowledge and skills they had acquired both before and during the war to the increasing number of individuals involved in the Project's health and safety operation. 131

The primary goal of the Manhattan Project was, of course, to create atomic bombs for use during the war. This goal was achieved and two bombs were dropped on Japan in early August 1945. The creation of these weapons and the conclusion of the war did not bring an end to the government's involvement in radiation research and development, however. In fact, the radiation sickness that affected atomic bomb survivors in Japan created a more urgent need to investigate the biological effects of radioactive fallout. The problem of fallout was a responsibility that the MED's successor—the Atomic Energy Commission

(AEC)—would inherit. In the following chapters we see, though, that the AEC's biomedical radiation research involved much more than investigating defense- related radiation hazards. That the AEC's biomedical program incorporated many of the aims pursued by scientists in their pre-war work was, in large part, the result of the network of radiation researchers examined here. Their wartime experience strengthened their professional connections and, perhaps more importantly, integrated them into a political economy in which they learned to pursue their research and discipline-building goals in relation to the social and political needs of the state. 132

CHAPTER THREE

FROM WAR TO PEACE: INSTITUTIONALIZING BIOMEDICAL RADIATION RESEARCH WITHIN THE ATOMIC ENERGY COMMISSION

The previous chapter examined the continuities and discontinuities between pre-war and wartime biomedical radiation research. As in so many fields of research and development, national security concerns permeated biomedical radiation research and biomedical radiation researchers were absorbed into a government-military complex created to reorganize the nation's research and development enterprise for defense purposes. Amidst this fundamental reorientation of the scientific enterprise, there was a continuous thread in basic biomedical radiation research and the network of biomedical radiation researchers that existed from the early twentieth century through World War II. Within the wartime research enterprise in which the state was a central player, the pre­ existing network of biomedical radiation researchers—especially those individuals in administrative positions—learned to work within the highly bureaucratic and goal-driven circumstances of the Manhattan Project. Their wartime experience had a lasting impact on the development of biomedical radiation research, primarily because it put researchers in a position to press their federal patrons to support biomedical research, and research related to more than just defense problems.

Following World War II, while the United States sought to demobilize its armed forces, but also prepare for the possibility of atomic warfare, researchers helped establish biomedical radiation research as a national priority. They 133 responded to and defined both national security and medical interests which, at times, seemed contradictory. Despite the gaps that existed between national security concerns and medical goals, researchers marshaled both to justify ongoing federal investment in biomedical research. Researchers who had been involved in the Manhattan Project and who had been leading radiation researchers before the war willingly worked within the boundaries of the Manhattan Engineer

District (MED) and its successor, the Atomic Energy Commission (AEC), to encourage biomedical studies of atomic bomb survivors in Japan, the creation of a broad biomedical program, and corresponding infrastructure within or funded by the AEC. This chapter therefore examines the organizational development of biomedical radiation research during the early postwar years, particularly the formation of the AEC's Division of Biology and Medicine (DBM) and Advisory

Committee on Biology and Medicine (ACBM). The DBM and ACBM, I argue, were crucial to preserving researchers' ability to influence the future development of biomedical radiation research. They were a link within a network of radiation researchers and a link between that network and the government. Furthermore, these organizational structures represented the importance of biomedical radiation research within an agency that was a significant patron of science following

World War II and central to Cold War America.

The Commission's biomedical radiation research program helped to expand the nation's infrastructure and funding for biomedical radiation research, in part because the program extended far beyond the walls of AEC laboratories.

The AEC sponsored research that was conducted in-house in its own laboratories, 134 but, as we will sec in the remaining chapters, the AEC also provided radioisotopes and funding to researchers and institutions to facilitate research and training in universities. Such support resulted in new programs and institutions. The momentum evident in the early postwar development of biomedical radiation research infrastructure, policies, and programs appears, on the surface, to have been intrinsic. However, the creation of the DBM and ACBM was a process shaped by a complex mix of objectives that researchers and administrators tried to serve as they prepared for the transition from the MED's wartime operation to the

AEC's peacetime program. Researchers played a significant role in driving and nurturing an expansion of infrastructure for biomedical radiation research within a changing social, political, economic, and scientific environment. Many scientists who had been recruited for wartime work remained in administrative positions within the AEC, if not permanently, then at least long enough to influence the organization of the Commission and its early biomedical research policies and programs. They were discipline-building, but doing so within a new institutional framework.

There is a substantial literature on the creation of the AEC, but little of it examines the efforts of biomedical radiation researchers to establish their own research agendas as priorities within the AEC.231 Accordingly, this chapter seeks to expand upon the work of historians such as Angela Creager and Karen Rader who have drawn attention to the role of government patronage in postwar

See, for example, Alice L. Buck, A History of the Atomic Energy Committee (Washington, DC: U.S. Department of Energy, 1983); Hewlett and Duncan, Atomic Shield; Hewlett and Holl, Atoms for Peace and War. 135 biological and medical sciences.*" This study furthers such work by breaking down the distinction between the government, on the one hand, and researchers or research communities, on the other—scientists were integrated into government agencies. The development of biomedical radiation research first within the MED and then within the AEC represented a merger of researchers' and the state's interests. This merger was both influenced by and helped constitute Cold War culture. Furthermore, this merger was part of the continuity and tradition of adaptation and collaboration within the history of science and not just a result of

Cold War politics and government patronage to meet national security needs.

ADVANCING RESEARCH IN THE AFTERMATH OF WAR: BIOMEDICAL RADIATION RESEARCH IN JAPAN

The destruction of Hiroshima and Nagasaki on 6 and 9 August 1945 were the culmination of the intensive effort to create atomic bombs. These attacks helped bring the war to an end and the Japanese surrender in August was made official on 2 September. While the primary goal of the Manhattan Project had been achieved and the war was over, biomedical radiation researchers still felt they had important work to do. Like most researchers involved in the Manhattan

Project or other wartime weapons projects, biomedical researchers were uncertain about the future relationship between science and government. This was an issue that would take time to settle, but in August and September 1945 some biomedical researchers were especially focused on a problem of greater immediacy. Indeed, it was not long after the bombs had been dropped that

2,2 Creager, "Industrialization of Radioisotopes"; Creager, "Nuclear Medicine in the Service of Bioniedicine"; Creager, "Export of'American' Radioisotopes"; and Karen A. Rader, "Alexander Hollaender's Postwar Vision for Biology: Oak Ridge and Beyond," Journal of the Historv of Biology 39 (2006): 685-706. 136

Manhattan Project researchers became aware that atomic bomb survivors were suffering from a range of symptoms caused by exposure to high levels of

radiation—symptoms that, at that time, were poorly understood. As historian M.

Susan Lindee has shown in her pivotal study of postwar American research in

Hiroshima, those who survived the bombings were victims of war, but, from a

research perspective, they also constituted a large group of human subjects that

many researchers felt should be studied in an effort to better understand the

effeets of total-body irradiation.233 ~

Dr. Shields L. Warren, a pathologist who later became a key member of

the AEC's biomedical program, was one such researcher who felt biomedical

researchers should go to Japan to study atomic bomb survivors. Not to be

confused with Dr. Stafford L. Warren who was a University of Rochester

radiologist and, during the war, the Chief of the MED's Medical Office, Shields

Warren joined the Naval Reserve in 1939 and conducted medical work for the

Navy during the war. He had almost no contact with the MED, but he was neither

new to radiation research nor was he oblivious to the health hazards of radiation.

As noted in Chapter One, Warren had considerable experience conducting both clinical work and research with radiation. Before the war he worked at the New

England Deaconess Hospital and was head of the Massachusetts State Tumor

Diagnosis Service where he investigated the use of radioisotopes for radiotherapy and metabolic studies. When the war ended, he drew the Navy's attention to the

M. Susan Lindee, Suffering Made Real: American Science and the Survivors a! Hiroshima (Chicago, IL: University of Chicago Press, 1994). 137

importance of conducting medical studies with bomb survivors.234 He was, at that

time, a Commander with the Navy. He encouraged officials within the Navy to

include biomedical research as part of the Naval Technical Mission which was

organized to assess the effects of the bombings. As a result, Warren was sent to

Japan to lead a small medical group consisting of 14 scientists and physicians.

Starting in Nagasaki near the end of September they spent a couple of months

measuring radiation levels in the environment with Geiger counters, examining

survivors and noting the symptoms they experienced, and collecting blood

samples for study.235

Shields Warren's group was not the only, nor was it the first, group

studying and compiling medical data on the survivors. Shortly after arriving in

Japan in late September, the Naval Technical Mission's medical team joined

forces with three other teams of researchers to form the Joint Commission for the

Investigation of the Effects of the Atomic Bomb in Japan. The other groups

involved were the General Headquarters of the U.S. Army Forces Pacific, the

Manhattan Engineer District (MED), and the Japanese Imperial Government.

General Headquarters was the first American group established in Japan. After

landing in Japan on 1 September, General Headquarters' Office of the Chief

Surgeon immediately began to organize a medical investigation of bomb survivors. General Headquarters' medical investigation was placed under the

direction of Colonel Ashley W. Oughterson who was a pathologist with the

^ Brues, "Shields Warren: (1898-1980)," 433. Oral History of Nello Pace, Ph.D., interview conducted August 16, 1994, as part of the Office of Human Radiation Experiments' Oral History Project, Human Radiation Studies: Remembering the Early Years. Conducted by Ms. Anna Berge in Berkeley, CA, 7-8. 138

Medical Corps. Almost a month had passed since the bombings, though, and

American investigators believed that the initial observations made by Japanese

investigators would be invaluable to their study.236 Thus, within days of landing

in Japan, General Headquarters made contact with the Japanese Imperial

Government and gained access to the initial medical reports. The Japanese

Government agreed to cooperate with General Headquarters and arranged for Dr.

Masao Tsuzuki to serve as a liaison to General Headquarters. In addition to

providing American researchers access to data collected in the first weeks after the bombings, Japanese cooperation also significantly expanded General

Headquarters' manpower. Tsuzuki directed approximately 90 physicians and

researchers whereas Oughterson's team only consisted of 23.237

When General Headquarters was making arrangements with Japanese

investigators and laying groundwork for a study of survivors, neither group envisioned the creation of the larger effort that formed later under the auspices of the Joint Commission. They were unaware that other American groups would arrive in Japan for the same purpose. However, as discussed above, the Navy agreed with Warren's recommendation that the Naval Technical Mission should include a medical group. The MED also organized a mission to Japan to survey the damage caused by the atomic bombs and, similarly, included a medical team.

The MED's medical group wanted to conduct a short-term study of the residual radioactivity in the two bombed cities and the medical effects experienced by survivors. The MED's medical group was placed under the direction of the Chief

2M' Ashley W. Oughterson and Shields Warren, eds.. Medical Effects of the Atomic Bomb in Japan (New York: McGraw-Hill Book Company, 1956) 7. 2,7 Ibid., 8. 139 of the MED's Medical Office, Dr. Stafford L. Warren. The MED's medical team consisted of almost 30 men.238 ' This group, like the Naval Technical Mission group agreed to join General Headquarters and the Japanese Imperial Government in a coordinated effort to conduct a more thorough study of bomb survivors.

The formation of the Joint Commission illustrates how, in the immediate postwar period, biomedical researchers continued to operate within and adapt to changing circumstances within the large government-military research enterprise established during the war. The three American medical groups involved in the

Joint Commission were part of a military operation that planned to investigate the damage caused by the atomic bombs. They surveyed the devastated cities and examined survivors to collect data on the three main types of injury caused by the bombs which they categorized as blast, burn, and ionizing radiation injuries.

There was obvious military importance to this work given that the atomic bomb was a new weapon that was far more powerful than any bomb ever used before.

Military strategists sought to gather information that would inform offensive and defensive strategies in the event that a future war might involve use of atomic bombs. Indeed, determining the impact of radiation as a weapon or how effective radiation was in causing casualties was a priority for medical investigators in

Japan. They knew then that the future design of nuclear weapons might be very different from the bombs detonated in Japan. However, they felt certain that

~w Ibid. For a full list of the personnel in each group involved in the Joint Commission, sec Appendix A, pages 431-432. 140

radiation hazards would be a problem and that their medical investigations would

provide vital information for future civilian and military defense. 239

Personnel within both the Army and Navy were eager, if not impatient, to

see for themselves the effects the new weapon had on the bombed cities in Japan.

Dr. Nello Pace, a physiologist with the Navy who went to Japan as part of the

Naval Technical Mission, described the urgency expressed by Admiral Ross

Mclntyre, Surgeon General of the Navy. In an interview conducted in the 1990s

Pace recalled that just after the atomic bomb had been used in Hiroshima and

Nagasaki Admiral Mclntyre said, '"Geez, we've got to find out something about

what happened to those people and how much radiation there is.'"240 Pace's

perception of how this high-ranking military official sought to act quickly in the

aftermath of the atomic bombings captures the novel situation that existed after

the bombs had been used. His recollection of Admiral Mclntyre's attitude

suggests a sense of unease felt by those who knew little or nothing of the

Manhattan Project during the war.241 Manhattan Project scientists had conducted

research during the war in an attempt to determine the effects of an atomic bomb

explosion, but they had mostly been concerned with determining the effects of the

bomb's blast, not the effects of fallout. As Los Alamos physicist Hans Bethe later

2y'IbicL 10 and 126-27. 240 Oral History of Nello Pace, 1994, 7. According to Susan Lindee, Mclntyre ordered the creation of a medical team as part of his Navy group after Shields Warren expressed concern that the MED had taken little initiative to study the effects of radiation. Lindee suggests that Mclntyre was "sympathetic" to Warren's suggestion. See Lindee, Suffering Made Real, 22. explained in an interview, "we figured...that people exposed to nuclear radiation

would be dead from the other effects."242

Once the new weapon had been used, individuals within different branches of the military and, similarly, within the scientific and medical community, wanted to arm themselves with an extensive knowledge of all things nuclear. To do so was, for them, important for ensuring that the nation as a whole was sufficiently knowledgeable and prepared to attend to national security concerns. Also, various branches of the military and researchers specialized in different fields were eager to understand the implications of the atomic age so they could position themselves in relation to domestic and international priorities.

That is, each branch of the military and researchers in various fields wanted to ensure that they had an important role to play in the future development of nuclear weapons andi atomic• energy.243

It was within this climate of novelty, unease, and impatience that decisions about medical investigations were made. To biomedical researchers, studies of the atomic bomb survivors were not only important to military medicine or defense-motivated health and safety. Biomedical researchers had never before been able to study the effects of total-body irradiation amongst such a large group of humans. Indeed, for ethical reasons researchers could not have purposefully exposed humans to large total-body doses of radiation since such exposure was

242 Oral History of Hans Bethe, Frederick Reines, Robert Christy, and J. Carson Mark, interview conducted August 18, 1989, as part of the Smithsonian Videohistory Program (18), session 14, "The Manhattan Project: Collection Division 4: Los Alamos." Conducted by Stanley Goldberg, 31; Ball, Justice Downwind: America's Atomic Testing Program in the 1950s, 13; Hacker, The Dragon's Tail, 76-77 and 110-11. "'41 Historian M. Susan Lindee also suggests that competition fueled the creation of what, initially, was three American medical teams. See Lindee, Suffering Made Real, 22-23. 142

thought to be very harmful. The ethical implications of investigating the health

effects of radiation exposure amongst atomic bomb survivors was an issue that

was debated amongst researchers and military and government personnel,

especially after a permanent biomedical study organized by the newly created

Atomic Bomb Casualty Commission (ABCC) replaced the Joint Commission in

1947. The Joint Commission was intended to be a temporary organization formed for the purpose of completing an initial survey of radiation hazards.244 Historian

M. Susan Lindee explains that the ABCC's policy to provide no treatment to

Japanese survivors provoked the most debate regarding ethics. This policy

reflected the dominant view held by United States government officials that,

although atomic bomb survivors were being used as research subjects, they were, first and foremost, war casualties. As Lindee argues, the United States believed that to provide war casualties treatment, might be perceived as public atonement for the decision to drop atomic bombs. The American government did not want their activity in Japan to look like a statement of apology for wrongdoing.245

Biomedical researchers found it difficult to discern the symptoms of radiation injury amidst multiple injuries, but made it a priority to determine the natural history or progression of radiation sickness. They believed that the data collected would generate knowledge relevant to civilian and defense radiation

24,1 The authors of the Joint Commission's final report in May 1946, Ashley Oughterson and Shields Warren, decided that a long-term study was necessary and that it should be organized through the National Academy of Sciences. The AEC did, however, fund the work of the ABCC and considered their patronage to be an important contribution to cancer research. See Ibid., 32. and "AEC Argonne Cancer Research Hospital, 1948," NARA College Park, RG 326, General Correspondence 1946-1951, Box 64, Research in Biological & Medical Science (6-20-47), 2-3. -4's M. Susan Lindee, "Atonement: Understanding the No-Treatment Policy of the Atomic Bomb Casualty Commission," Bulletin of the History of Medicine 68 (1994): 454-90; Lindee, Suffering Made Real, especially 117-42. 143 safety, the future use of radiation in medicine, and possible nuclear wars.246 The biomedical research conducted in Japan with atomic bomb survivors was the first postwar example of biomedical researchers helping to both define and adapt to the opportunities and limitations of the times. The circumstances in which they carried out research were very different from those in which biomedical researchers worked prior to and during the war, but the ability and willingness of researchers to adapt to what they considered to be research opportunities was not new. Biomedical researchers advocated for research that they believed to be appropriate to the changed circumstances and, thus, reinforced the relationship that they had formed with the government and military during the war.

Dr. Masao Tsuzuki and American biomedical researchers aboard a train that served as a mobile medical laboratory, Nov. 1946. Photo Credit: Atomic Bomb Casualty Commission-Radiation Effects Research Foundation Photo Album.

THE TRANSITION FROM WARTIME TO PEACETIME R&D: CREATING THE ATOMIC ENERGY COMMISSION

The MED remained in place through the end of 1946. However, the creation of a new agency to replace the MED had been a topic of discussion and debate since before the war ended. Individuals within government, military, and scientific and technical communities all tried to shape the legislative process that

~4'' Oughterson and Warren, eds.. Medical Effects of the Atomic Bomb, 126-27. 144 culminated on 1 August 1946 when President Harry S. Truman signed the Atomic

Energy Act.247 In the months that followed, the MED prepared to transfer control over its vast network of laboratories and production facilities to the AEC which was formally established on 1 January 1947. These included the laboratories now known as the national laboratories—the Met Lab's successor, Argonne in Illinois;

Oak Ridge in Tennessee; the Radiation Laboratory at the University of California,

Berkeley; Los Alamos in New Mexico; and Brookhaven in New York. The latter was not officially established until March 1947, but plans for this laboratory were initiated under MED leadership. In terms of production facilities transferred to the AEC, the in Washington state was the largest.248 Aside from laboratories and production facilities, the MED transferred personnel. At the peak of its wartime operation the MED had employed 130,000 individuals, but personnel were reduced such that at the time of transfer, 44,000 individuals joined the AEC.249

The creation of the AEC was part of a postwar trend of expanding government infrastructure for research and development. As many historians have argued, this was a pivotal moment in the history of American scientific

241 For a detailed discussion of the various renditions of legislation, see Brian Balogh, Chain Read ion: Expert Debate and Public Participation in American Commercial Nuclear Power, 1945-1975 (Cambridge, UK: Cambridge University Press, 1990); Greg Herken, The Winning Weapon: The Atomic Bomb in the Cold War, 1945-1950 (New York: Alfred A. Knopf, 1980), 117-22; Hewlett and Anderson, The New World, 482-516; Jessica Wang, American Science in an Age of Anxiety: Scientists, Anticommunism, and the Cold War (Chapel Hill, NC: University of North Carolina Press, 1999), chapter 1. :*ls In addition to the primary laboratories and production facilities listed here, the AEC also inherited a few smaller research and production facilities from its predecessor. Peter J. Westwick, The National Labs: Science in an American System, 1947-1974 (Cambridge, MA: Harvard University Press, 2003), 8-9. Hacker, Elements of Controversy, 10-11; Hewlett and Anderson, The New World, 2; Hughes, The Manhattan Project, 9. 145

2^) research, biomedical research included. ~ With the creation of agencies such as the AEC and National Science Foundation (NSF) and the reorganization and growth of others such as the National Institutes of Health (NIH), the government made a massive and long-term commitment to funding research related to both defense and civilian problems.

President Truman signs the AEC Act, August 1946. Behind him (left-right): Senators Tom Connally, Eugene Millikin, Edwin Johnson, Thomas Hart, Brien McMahon, Warren Austin, and Richard Russell. Photo Credit: Department of Energy Office of Science.

The AEC was created as a civilian agency, but this designation was misleading. The Commission had extremely significant defense responsibilities related to the management of the nation's nuclear weapons such that just months before the AEC assumed control over the MED's operations and facilities, the

MED reported: "Present plans provide for continuation of production at the present rate and for a vigorous program of research directed first at improvement of military applications and, secondly, at a utilization for civilian purposes."251 It

~50 See. for example, Geiger, Research and Relevant Knowledge: American Research Universities since World War //; Leslie, The Cold War and American Science; Westwick, The National Labs; Zachcry, Endless Frontier: Vannevar Bush. "M "Plan for Transfer of Responsibilities and Functions of the Manhattan District to the AEC, 1 November 1946," NARA College Park, RG 326, General Correspondence 1946-1951, Box 23, Organization and Functions, A-l. 146

is evident here that plans for the new AEC prioritized defense over civilian

research and development. "

Historians Mark Harrison and Michael Dennis have questioned the categorization of research as either military/defense or civilian.253 The AEC's biomedical radiation research illuminates the difficulty of making such categorizations in that it was part of a large system of government-funded research and development—an arrangement that was created in the midst of war to meet the needs of the nation at war. The determination of what constituted military/defense or civilian research is complicated when acknowledging that wartime research and development projects mobilized civilian researchers who brought to the project their knowledge, skills, and practices acquired through work at universities and other public and private institutions. Furthermore, during the Cold War, researchers worked within a culture in which the motivation for, technologies used, or funding provided for particular research projects defied simple demarcations between what constituted either civilian or military research.

As historian Gerald Kutcher argues in relation to Cold War cancer research, those engaged in research would not have considered there to be clear boundaries between medical and military research, at least not as clear as later historical constructions would suggest.254

25~ The production of nuclear bombs was a priority because as of 1946, the MED, in fact, had very few bombs stockpiled. See Herken, The Winning Weapon, 12. Dennis, "Two University Laboratories in the Postwar American State"; Mark Harrison, "Introduction - Medicine and the Management of Modern Warfare: An Introduction," in Medicine and Modern Warfare, ed. Roger Cooter, Mark Harrison, and Steve Sturdy (Atlanta, GA: Rodopi, 1999). "M Kutcher, "Cancer Therapy and Military Cold-War Research," 105-30, especially 108-9. 147

The establishment of the AEC, with its defense and civilian objectives and

ample resources and facilities, ensured that the large-scale, government-funded

scientific and technological research and development begun during the

Manhattan Project could continue unabated in the post-World War II years. The

creation of the AEC did not, however, ensure that wartime research and

development would continue unchanged. Just months after the AEC was

established, its first general manager, Carroll L. Wilson expressed his opinion that

Congress had designed the Atomic Energy Act such that it was, at least in part,

experimental. He said, "Usually, when Acts are passed they are perhaps considered to be the final word, but in view of the many unknowns of this

business, Congress formally acknowledged in the statute that it would have to be amended from time to time in light of experience and the changing needs in this

field."255 The Atomic Energy Act provided for the means through which the Act could be revised. It created the Joint Committee on Atomic Energy (JCAE) which was comprised of nine members each from the House and Senate. As

Wilson described, the members of the JCAE would be kept informed of AEC activities so that they would, it was hoped, take the initiative to recommend changes to the Act.

The origins of the JCAE are not clear in the legislative history leading to the creation of the Atomic Energy Act. Harold P. Green and Alan Rosenthal, who were both employed in the AEC's general counsel's office, suggested in their

1963 study of the JCAE that wartime arrangements influenced the design of the

255 "Transcript: First Meeting of the Medical Board of Review, June 1947," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 40, Board of Review, 5. 148 congressional committee as a joint committee. During the war the Senate and

House Appropriations and Military Affairs committees had formed an informal joint committee that cooperated to keep the Manhattan Engineer District's appropriations requests hidden from general Congressional oversight. According to Green and Rosenthal, "The proven feasibility of this arrangement and its success in preserving security during the war may have influenced the draftsmen of the 1946 act in suggesting a joint committee."256 What is clear is that members of Congress believed it to be very important that Congress play a role in the AEC.

Congressional members were concerned about the government monopoly of atomic energy and sought to ensure that, at the very least, the monopoly did not result in absolute authority resting in the executive branch.257

AEC Commissioners at Los Alamos National Laboratory, c. 1947. Left-right: Robert Bacher, David Lilienthal, Sumner Pike, William Waymack, Lewis Strauss. Photo Credit: Los Alamos National Laboratory Image Gallery.

Formal legislative change was one means of influencing the policies and programs of the AEC. Amendments to the Act were rare, though. During the remainder of the 1940s, in particular, the Commission's policies and programs

"5<' Harold P. Green and Alan Rosenthal, Government of the Atom: The Integration of Powers (New York: Atherton Press, 1963), 4. ~57 Harold P. Green and Alan Rosenthal, "Fusion of Government Power." Bulletin of the Atomic Scientists 18, no. 6 (1962): 12; Green and Rosenthal, Government of the Atom, 3. 149 were more affected by the Commissioners and general manager appointed to the

AEC as well as the researchers and administrators in charge of its various divisions." The Atomic Energy Act was sufficiently broad that the

Commissioners, who included David E. Lilienthal (chairman), Robert F. Bacher,

Lewis L. Strauss, Sumner T. Pike, and William W. Waymack, along with general manager Carroll L. Wilson, had considerable latitude to make decisions. Most of the Commissioners had experience and skills that, to President Truman's

Administration, seemed appropriate for the management of a public agency, but only one of the five was a scientist.259 Thus, they depended heavily on the advice of researchers when determining research policy. As the AEC considered how to proceed with the health and safety work begun during the Manhattan Project, the

Commissioners recognized and utilized the expertise of the researchers who had been involved in the Manhattan Project and experts outside of that enterprise.

This was certainly the case with the creation of the Division of Biology and

Medicine, the Advisory Committee on Biology and Medicine, and the biomedical program they managed.

258 Green and Rosenthal describe the JCAE's role in the atomic energy program as "passive" until late in 1949 at which point the JCAE assumed greater leadership in shaping the direction of the program. See Green and Rosenthal, "Government Power," 5. 259 Lilienthal was a lawyer and had been the chairman of the Tennessee Valley Authority. He had no scientific or technological expertise but was experienced in managing a large technical enterprise. Both Strauss and Pike were financiers who had been involved in government and military business during the war. Waymack had administrative skills honed through years as a newspaper editor and as deputy chairman of the board of the Federal Reserve Bank of Chicago. Bacher, a Los Alamos physicist, was the lone scientist amongst the group. Selected as general manager, Wilson was an engineer and former assistant to MIT President Karl T. Compton. He also helped Vannevar Bush organize the NDRC and OSRD during the war. For more information on the individuals appointed to the Commission, see Hewlett and Anderson, The New World, 621 - 22; Hewlett and Duncan, Atomic Shield, 1-6. 150

CHARTING A COURSE FOR BIOMEDICAL RESEARCH WITHIN A NEW ORGANIZATION

During the transition period between the MED and AEC, individuals from both, and others outside of these agencies, proposed that biomedical research should be amongst the most important civilian objectives of the new agency.260 It is worth noting that MED and AEC reports used the terms medicine/medical, biology/biological, and biomedicine/biomedical interchangeably. The policies and programs implemented during the early years of the AEC's history suggest that the Division of Biology and Medicine (DBM) was overwhelmingly concerned with medical or biomedical research and development. However, as the AEC's support for basic science expanded, so too did its biological research.

For instance, the DBM increased its support for research focused on ecological problems. This study remains focused on the Commission's medical or biomedical research, though, since the expansion of the DBM's program to

Oft| include more ecological research did not occur until the mid-1950s.

As the MED prepared to transfer its enterprise to the AEC there was little, if any, indication that medical and biological research would be organized

2WI See, for instance, "Program of Administration of the Atomic Energy Act of 1946," NARA College Park, RG 326, General Correspondence 1946-1951, Box 23, Atomic Energy Commission; "Plan for Transfer ..., 1946," NARA College Park, Box 23; and Stafford L. Warren, "Report of the 23-24 January 1947 Interim Medical Committee," NARA College Park, RG 326, General Correspondence 1946-1951, Box 32, Interim Medical Advisory Committee - Meetings and Agendas. "M For more information on the AEC's ecological research, see John Beatty, "Ecology and Evolutionary Biology in the War and Postwar Years: Comments and Questions," Journal of the History of Biology 21, no. 2 (1988): 259-61; Stephen Bocking, "Ecosystems, Ecologists, and the Atom: Environmental Research at Oak Ridge National Laboratory," Journal of the History of Biology 28 (1995): 1-47; Peter J. Taylor, "Technocratic Optimism, H. T. Odum, and the Partial Transformation of Ecological Metaphor after World War II," Journal of the Historv of Biology 21, no. 2 (1988): 240-41. 151 separately from all other AEC research, but that is, in fact, what happened.262 The

Commission was created with four statutory divisions: the Divisions of

Engineering, Production, Military Application, and Research. Biomedical

radiation research might have been organized within the Division of Research,

however, as plans for the AEC were put in place biology and medicine were continually treated as a separate endeavor. During the Manhattan Project,

research in medicine and biology had played an important role within the overall

project, but was peripheral to the primary objective of building bombs. In the

postwar years, biomedical radiation research became part of the AEC enterprise,

which was meant to encompass more than just bomb-building.

FIGURE 1: AEC ORGANIZATION CHART, DECEMBER 1948263

U S. ATOMIC ENERGY COMMISSION

ATOMIC ENCRSV COMMISSION

2h2 "Program of Administration... 1946," NARA College Park, Box 2, 1-9. 263 Buck, A History of the Atomic Energy Committee, 1 8. 152

Although the DBM was not one of the four statutory divisions and was not created until the fall of 1947—almost a year after the AEC was in place—the process that led to the establishment of the DBM and the ACBM began before the creation of the AEC. Here, the focus is on the role of researchers and administrators in positioning medical and biological research relative to the larger objectives of AEC business and their authority to shape government policy. The process of defining a role for biomedical radiation research within the institutional structure of the AEC was similar to and became intertwined with a process of discipline-building that was evident at a few universities prior to the war.

Those who were involved in the Manhattan Project's health and safety research were able and willing to advise the Commissioners on the type of research they thought should be pursued in the future, as well as the allocation of funds for research. In September 1946 as the transfer between the MED and AEC neared, the MED's Medical Advisory Committee prepared a report to do just that.

The Medical Advisory Committee was chaired by Dr. Stafford L. Warren and consisted of eight other members, all of whom were highly involved in the

Manhattan Project's health and safety work and who had participated in interdisciplinary biomedical research earlier in the century. For instance, the committee included, but was not limited to, individuals discussed in the previous two chapters: Warren's MED Deputy Chief of the Medical Office, Dr. Hymer L.

Friedell; the Director of the Met Lab's Health Division, Dr. Robert S. Stone; the

Director of Los Alamos' Radiologic Safety Group, Dr. Louis H. Hempelmann,

Jr.; the Director of the Rochester Atomic Energy Project, Dr. Andrew H. Dowdy; 153

and consultant to the Health Division, Gioacchino Failla.264 Their report

summarized the MED's health and safety work and emphasized that while

researchers had made good progress throughout the war, there was still much

work to be done. They also identified many important areas of research, such as

the physical measurement of various types of radiation, the biological effects of

radiation, and prevention and treatment of radiation sickness.265

Once the AEC took charge, it too formed a committee to evaluate the

Manhattan Project's health and safety research and plan for future research. The

AEC's Interim Medical Committee met from 23-24 January 1947 and was, in effect, an extension of the Medical Advisory Committee. With the addition of a

few new members, the Interim Committee, like its predecessor, was chaired by

Stafford Warren. Not surprisingly, the Interim Committee produced a report that conveyed a similar message to that of the MED's earlier report. The new committee noted the hazards associated with atomic energy, such as the danger of radiation poisoning, and recommended the continuation of research programs already in place. The Interim Committee thought, though, that it was not enough to continue the health and safety programs started by the MED. Rather, the

Committee recommended that the AEC play a role in training individuals to conduct the health and safety work of the AEC. For the Interim Committee, the lack of properly trained personnel was one of the most serious problems the AEC faced in terms of controlling radiation hazards. The Interim Committee's report attributed this problem to "the combination of a hiatus of basic science training

The remaining members were John Wirth, Andrew Dowdy, Raymond E. Zirkle, and J. H. Sterner. "Plan for Transfer1946," NARA College Park, Box 23, B-l 1. ~65 Warren, "Interim Medical Committee..., 1947," NARA College Park, 32, 3-4. 154 during the war and the uniqueness of the problems confronting the Atomic Energy

Commission."266 With the suggestion that the AEC play a role in training, the

Interim Committee encouraged more than a research program. By urging the

AEC to train new individuals to conduct research within the hybrid space of biomedical radiation research, the Interim Committee helped put in place discipline-building mechanisms within the AEC's biomedical research agenda.

Examined in greater detail in Chapter Five, the AEC's training programs factored significantly in the formation of new disciplines or sub-disciplines in academic institutions.267

The Interim Committee echoed the sentiment expressed by Robley D.

Evans a year earlier in the popular magazine The Atlantic Monthly. As we saw in

Chapter One, Evans was an MIT physicist who had been researching the effects of radium poisoning since the early 1930s. He worked on a couple of wartime research projects, one of which used radioactive tracers to investigate methods to improve blood preservation.268 He was, therefore, involved in wartime research, but further removed from the MED than those who served on the Interim

Committee. Writing soon after the war ended, Evans raised concern about "the problem of personnel." He reckoned,

So far, the most fruitful applications of these new techniques have been the results of team work between physicists and other scientists. There is a need for a number of hybrid Ph.D.'s who can bridge the gap between physics and the other scientific fields and whose breadth of knowledge fits

:w' Warren, "Interim Medical Committee..., 1947," NARA College Park, Box 32, 21. ~bl The combination of a market oriented towards weapons development and related research and the desire of researchers in academic institutions to maintain research programs allowed for the creation ofnew disciplines or sub-disciplines in many institutions. This is an issue best examined in relation to the physical sciences. See Dennis, "Two University Laboratories in the Postwar American State," 427-55; Leslie, The Cold War and American Science. :('s Evans, "The Medical Uses of Atomic Energy," 71. 155

them to act as key men in the cooperative research teams of the present and future.269

Evans' article does not relate the problem of insufficient biomedical radiation

researchers to the business of the AEC. Rather he spoke broadly about the

medical uses of radiation, excluding any mention of the government in this

research. His plea for training more researchers with hybrid expertise therefore seemed an effort to build momentum within a field of research that predated the

Manhattan Project. However, Evans surely meant for his message to reach beyond researchers to those who played a role in organizing and funding scientific

research which, in 1946, included MED administrators and scientists who shaped plans for the AEC.

Evan's reference to cooperative research teams speaks to the aims of foundation patronage for science during the interwar period. At that time, foundations mainly sought to support cooperative projects that, in effect, served as a means of building interdisciplinary networks in which knowledge, practices, and technologies were shared, and students could be trained in hybrid research practices.270 During the war years and postwar period, government agencies like the AEC supplanted foundations as the leading patrons of science. Thus, Evan's message was likely meant to influence policies and programs within such agencies so that they might play a role similar to foundations in encouraging the development of cooperative research communities.

Ibid: 73. 270 On cooperative research, see Robert E. Kohler, "Philanthropy and Science," Proceedings of the American Philosophical Society 129, no. 1 (1985): 9-10; Kohler, "Science, Foundations," 135-64, especially 39-42. 156

Evans' message was not lost on the members of the MED's Medical

Advisory Committee and the AEC's Interim Medical Committee who pressed for wider use of radiation in biology and medicine, even if unrelated to the AEC's weapons development and associated hazards. The Interim Committee insisted that in addition to attending to "the special hazards of atomic energy development...it is imperative to open new fields of application for the products of atomic energy in biology and medicine."271 The Interim Committee did not elaborate on why it was "imperative" to expand research and develop practical applications of atomic energy in biology and medicine. However, this assertion, made amidst the shifting social, political, and economic circumstances of the early postwar era, suggests that the Interim Committee hoped to take full advantage of wartime technological developments. The members of the Interim

Committee knew, for instance, that radioisotopes had been inciting biomedical radiation researchers since their discovery in the 1930s, but that cyclotrons could only produce radioisotopes in small quantities. We will see in the next chapter that biomedical researchers played an important role in arranging for the distribution of -produced radioisotopes as a solution to the radioisotopes supply problem. The Interim Committee's statement also speaks to the reality that AEC support for biomedical research and medical applications of atomic energy served as a positive public relations strategy to counter the negative image of atomic bombs. Regardless of whether the Interim Committee was concerned with utilizing wartime technologies, crafting a public relations strategy, both issues, or neither, the assertive language the committee used in the report

"7I Warren, "Interim Medical Committee..., 1947," NARA College Park, Box 32, 7. 157

reveals the committee's determination to tie the interests of biomedical radiation

researchers to those of the AEC.

The Interim Committee's report provided little in the way of specific

polices or programs for the development of atomic energy for biological and

medical purposes. The Committee did, however, identify key components for doing so. They strongly recommended and offered a very detailed plan for the creation of a biomedical advisory committee and a biomedical research division

within the AEC. The latter was referred to as both a Medical-Biological Division and Health-Safety Division throughout the report. 272 The interchangeable use of these names indicates a degree of ambiguity regarding the Committee's vision for the division. Attending to the health and safety concerns within the AEC was an essential responsibility, but how far into biomedical research the AEC would delve was still unclear.

Having emerged from the wartime research experience as leading experts in the field of radiation safety, the members of the Interim Committee used their new-found authority to make recommendations that would reinforce their own and their colleagues' role within the government-military research enterprise. To staff both the proposed biomedical advisory committee and the Medical-

Biological or Health-Safety Division, the Interim Committee generated a list of names and the professional information of numerous physicians and researchers whose experience they deemed appropriate. The individuals they recommended were trained in medicine, biology, industrial hygiene, and health physics. More importantly, they were physicians and scientists who had long been leaders in

272 Warren, "Interim Medical Committee..., 1947," NARA College Park, Box 32, 22-23. ->73 biomedical radiation research.' The Interim Committee mostly identified

individuals who had participated in the Manhattan Project, but they were also able

to suggest the names of colleagues who had not been recruited for war work.

Overall, the Interim Committee's report reflects the priorities that had

guided biological and medical research during the war and the new possibilities

created with the formation of the AEC, a peacetime, civilian agency. It illustrates

the blurring of lines between the military and civilian spheres by proposing a

division that would see research and development for both the weapons industry

and biological and medical communities organized together. Furthermore, the

Interim Committee members' specific recommendations for staffing the new division and advisory committee encouraged the ongoing presence and expansion of their professional network within government channels.

SEEKING EXPERT ADVICE OUTSIDE OF GOVERNMENT & MILITARY CHANNELS

The Interim Medical Committee did not have the last word on the organization of radiation safety and biological and medical research within the

AEC. The chairman of the AEC, David E. Lilienthal, called for the creation of a

Medical Board of Review in May 1947 to assess both the legacy of wartime health physics and the Interim Committee's proposal for a biomedical program.

He enlisted the help of Frank B. Jewett who, at that time, was the President of the

National Academies of Sciences and National Research Council (NAS/NRC), to organize the review board. Lilienthal explained to Jewett that although the

Interim Medical Committee discussed its recommendations with the AEC's

-7'1 Warren, "Interim Medical Committee..., 1947," NARA College Park, Box 32, 22-29. 159

Commissioners, the AEC was still struggling to define a biomedical policy. "The

Commission finds it very difficult at this time," Lilienthal said,

to evaluate this program until it adopts a policy in regard to the areas of research in the field of medical and biological sciences for which the Commission must assume exclusive responsibility, and the very much wider area of research in these fields in which the Commission will naturally have an interest but which should be supported by many groups....274

In this passage Lilienthal identified the need for policy to guide the Commission's

development of biological and medical research. He did so, I believe, because the

possibilities for biological and medical research were vast and the Commissioners

needed help to create a research program that was appropriate not only within the

institutional context of the AEC, but also within the larger social, political,

economic, and technical climate of the early postwar period.

The NAS/NRC wasted little time in drafting a list of qualified physicians and scientists who could serve on the Medical Board of Review and by the end of

the month—May 1947—the Commission had selected and recruited seven

individuals to do the job.275 The Commissioners hoped that by forming this committee they could obtain some objective feedback from experts in biology and

medicine who had not been "intimately engaged in the work of the Project."276

The Commissioners and general manager attended some of the Review Board's

214 "Correspondence re Board of Review, May 1947," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 40, Board of Review, 2. 275 The seven members of the Medical Board of Review included Robert F. Loeb (chairman), Herbert S. Gasser, Alan Gregg, Ernest W. Goodpasture, A. Baird Hastings, Wallace 0. Fenn, and Detlev W. Bronk. They were all highly regarded in their fields and associated with prestigious institutions such as the Rockefeller Foundation, the Rockefeller Institute for Medical Research, Harvard Medical School, and the College of Physicians and Surgeons, Columbia University. See "Correspondence re Board of Review, 1947," NARA Atlanta, Box 40, 1-25. 7<' "Correspondence re Board of Review, 1947," NARA Atlanta, Box 40, 3. 160

meetings which were held in June 1947 so they could provide its members with

information regarding the nature and scope of the health and safety problems

encountered during the Manhattan Project. They also explained the AEC's

current organization and operations. For instance, Carroll L. Wilson, the AEC's

general manager, outlined the bureaucratic structure of the new agency which, by

statute, included the five Commissioners, a General Advisory Committee, a

Military Liaison Committee, a congressional Joint Committee on Atomic Energy,

and four divisions. He provided the Review Board with a sense of the geographic

distribution of the AEC's bureaucracy with the general headquarters located in

Washington, D.C. and field offices located at the major centers of operation: New

York, Oak Ridge, Chicago, Hanford, and Los Alamos.277 Research and

development was spread amongst both industrial facilities in these major centers

and university laboratories across the country. The Medical Board of Review

considered this information along with more specific information gathered

through interviews with numerous researchers who had been involved in the

MED's health and safety work.

The Medical Board of Review process illuminates the AEC's efforts to

create a biomedical research program that was responsive to the social, political,

and economic climate. The process also allowed researchers to play an integral

role. Indeed, it provided not just its seven members, but also the AEC's

Commissioners, general manager, and a selection of researchers involved in the

MED's health and safety research, an opportunity to have in-depth discussions

211 "Transcript... Board of Review, 1947," NARA Atlanta, Box 40, 5-9. 161 about the Commission's responsibilities, priorities, and the possibilities arising

from its special resources and facilities.

An ongoing theme throughout the Medical Board of Review meetings was that of novelty. For instance, when Wilson described the laboratories built during the war and those currently being planned, he explained that while some of

the AEC's laboratories had the purpose of generating new knowledge, others were considered programmatic. By programmatic he meant that they had specific objectives such as building reactors or enriching uranium. He acknowledged that

the programmatic laboratories, especially, were a novel arrangement as they were

very large and might employ as many as 1000-2000 individuals.278 Ensuring• a sufficient supply of health physicists to attend to safety in laboratories of this size

was an obvious responsibility discussed throughout the Review Board's meetings.

The Medical Board of Review considered what the role should be of both the

AEC and universities in creating and promoting educational programs that would

train health physicists to meet the AEC's employment needs and opportunities.

This discussion stemmed from the same concerns that motivated both the MED's

Medical Advisory Committee and the AEC's Interim Medical Committee to

recommend the establishment of an AEC training program focused specifically on

radiation in the biological and medical sciences. The Medical Board of Review

reinforced the earlier suggestions that the AEC should endeavor to build a

biomedical research community that was responsive to the specific needs of the

atomic energy business.

27X "Transcript... Board of Review, 1947," NARA Atlanta, Box 40, 10-11. 162

Beyond the responsibility of radiation safety, the Board of Review and

Commissioners engaged in a discussion about the role of the Commission in basic biomedical research. The MED had recommended that 15% of the budget for proposed Medical-Biological or Health-Safety Division be devoted to basic research.279 Again, this was an issue that in 1947 emphasized the novelty of the postwar situation. The United States government did not normally support basic research, but Lilienthal suggested that the AEC do so. He wanted to establish a broad biomedical program for public relations purposes—to help mold public opinion regarding the business of the AEC. He explained that the Commission could certainly limit their biomedical work to safety concerns, but that a broader biomedical radiation research program would likely benefit the health of the nation and, therefore, generate good will amongst both the general public and

Members of Congress. He asked that the Board of Review consider this factor given that the reputation of the scientific enterprise suffered when the secretive weapons development programs were unveiled to the public at the end of the war.280 As part of the government bureaucracy Lilienthal had an interest in protecting the image of science since science and the state had become closely intertwined. More specifically, he was responsible for promoting nuclear science and embraced the optimism about biomedical radiation research that researchers had been touting since the discovery of radiation and especially radioisotopes.

Lilienthal's ambition to use biomedical research to garner support for the

AEC's operations was validated in the DBM's budget as the years passed. For

~n Westwick, The National Labs, 244-45. "Transcript... Board of Review, 1947," NARA Atlanta, Box 40. 27-30. 163

instance, after the Bureau of the Budget reviewed the AEC's proposed budget of

$15 million for basic research during the fiscal year 1948, the cuts made took a much smaller toll on the DBM than on the Division of Research. Originally, the budget had allocated $5 million to the DBM and $10 million to the Division of

Research. In the end, the Bureau of the Budget approved a budget of approximately $9.9 million of which $4.6 million was allocated to the DBM and

$5.3 million to the Division of Research.281 The DBM's budget for basic research was barely affected by the revisions, whereas the Division of Research's budget was cut almost in half. By 1949, basic biomedical research received approximately one-third of the DBM's budget. 282 This was a considerable increase from the 15% recommended by the MED. Despite the increase, the

AEC's congressional committee, the Joint Committee on Atomic Energy (JCAE), still criticized proposed budget cuts to biomedical research and continually tried to protect the biomedical budget from cuts that affected other areas of AEC

00-5 operations throughout the 1950s. The preferential treatment both the Bureau of the Budget and the JCAE gave to biomedical research illustrates the special role biomedical radiation research played in establishing a positive image of government-funded science, generally, and the AEC, specifically.

The discussion that emerged out of Lilienthal's remarks to the Medical

Board of Review regarding the importance of basic research revolved around the implications of supporting research for defense or civilian purposes. It was clear

2M "Draft Minutes: Advisory Committee for Biology and Medicine, 9-10 January 1948," NARA College Park, RG 326, General Correspondence 1946-1951, Box 31, Biology and Medicine, Advisory Committee for - Meetings and Agendas, 13. "x' Westwick, The National Labs, 252. Ibid., 253. 164 at this meeting—a meeting that preceded the creation of the DBM—that the

Commission recognized the value of categorizing biomedical radiation research as a civilian pursuit. Between defense and public relations interests, the Board of

Review faced the challenge of defining biomedical radiation research policy that served two seemingly contradictory goals. Throughout the immediate postwar and early Cold War period administrators and researchers continually refrained biomedical radiation research to serve one or both of these goals.

The various factors the Medical Board of Review considered, like the

MED's Medical Advisory Committee and the AEC's Interim Medical Committee before it, were not just scientific or technical. They also contemplated social, political, and economic factors. The Commission's and Medical Board of

Review's attention to all of these factors indicated that they did not consider science to be an enterprise that was or could be isolated from the context in which it was pursued. As examined in previous chapters, biomedical radiation research was never isolated from social, political, or economic factors, but this became more obvious to both researchers and government officials when science was mobilized for defense purposes. Wartime and postwar government- and military- funded research represented significant shifts in the political economy in which biomedical radiation research developed. Within the Cold War climate, scientific prowess was imbued with political meaning such that even medical research conducted for so-called civilian purposes contributed to the psychological contest between the United States and its Cold War opponent, the Soviet Union.

Accordingly, Commissioner Bacher informed the Medical Board of Review that 165

money was not really an issue when defining the scope of the Commission's

biomedical research. He said that neither he nor the other Commissioners had any certainty about the future financial status of the Commission. However, they suspected that for the foreseeable future the AEC would have "substantial sums as compared with other groups."284

Since money would likely not be a limiting factor, the exchange between the Board of Review and Commissioners developed into an open discussion about the scope of research that the AEC could pursue in the fields of medicine and biology. Dr. Herbert S. Gasser, who was one of the Board of Review members and a physiologist from the Rockefeller Institute, explained to the Commissioners that the use of radioisotopes in biomedical research such as investigations of cellular mechanisms was potentially limitless. "Somewhere in there," Gasser said, "is the answer to the cancer problem probably."285 He posed what he thought was a fundamental question the Commission had to answer: How far into basic biomedical research was the Commission willing to go? Did the

Commission want to be at the forefront of biomedical research regardless of how distant such research might be from the business of building bombs?

Commissioner Pike turned the question around and asked the Medical Review

Board to determine how deeply involved they, as researchers, would want the

AEC to become in basic biomedical research. This was an issue that Gasser personally felt required more time for consideration.286

2tu "Transcript...Board of Review, 1947," NARA Atlanta, Box 40, 25. :t<5 "Transcript... Board of Review, 1947," NARA Atlanta, Box 40, 32. 2Sf> "Transcript... Board of Review, 1947," NARA Atlanta, Box 40, 32-33. 166

It was possible that the Commission could provide sufficient resources for researchers to realize scientific advances that had been dreamed about, but had perhaps been hindered by lack of funds, access to technology, or lack of manpower. However, large-scale government-funded science was a relatively new arrangement that raised concerns for some scientists and administrators alike.

Secrecy and intellectual freedom were two issues that especially worried the members of the Medical Board of Review, as well as many within the AEC.

General manager Wilson admitted to the members of the Review Board that

"there has not been enough hard thinking about how security of information can be maintained without impairing freedom of research."287 Neither he nor the

Commissioners who attended the Review Board's meetings were able to guarantee that this problem would be solved. Gasser raised numerous points he thought he and his fellow board members should consider. With respect to the relationship between civilian and defense research he expressed worry that educational institutions might make "mistakes of judgment." He believed that some universities were desperate for funds and might "be inclined to accept money under conditions which would be really detrimental to their primary purpose and objectives."288 Despite his concerns, Gasser expressed optimism that the Commission could take advantage of its unique position to help finance the expansion of medical and biological training to benefit both the Commission and

"Transcript... Board of Review, 1947," NARA Atlanta, Box 40, 60. "Transcript... Board of Review, 1947," NARA Atlanta, Box 40, 49. 167 the larger medical and biological communities. "I think there is a great frontier there for [a] pioneering plan," Gasser reckoned, "if it is properly conceived."2XI)

The AEC had given Gasser and his colleagues the important task of helping to lay the foundation of a "properly conceived" plan. After contemplating the wartime history of radiation safety and the longer history of biomedical radiation research the Medical Board of Review submitted a report to the AEC that outlined numerous recommendations for research in biology and medicine.

The report affirmed the earlier suggestions made by the MED's Medical Advisory

Committee and the AEC's Interim Medical Committee in that it recommended the continuation of existing projects and an expansion of research in medicine and biology. The report made specific recommendations regarding expansion, namely that the AEC create a division and advisory board to organize medical and biological research and that the AEC, through the proposed division, take the lead in training radiation researchers and safety specialists. Regarding the latter, the report stated that "Attracting the interest of able young men in the medical and biological aspects of radiation and providing them with training appropriate to their capacities and probably future careers is the most important long term task in the Atomic Energy Commission research program."290 Here, the Medical Board of Review, like the AEC's Interim Medical Committee and the MED's Medical

Advisory Committee before it made the point that the AEC should take responsibility for more than the execution of research projects. Rather the

Medical Board of Review encouraged the AEC to play a role in developing a

28'' "Transcript...Board of Review, 1947," NARA Atlanta, Box 40, 50. 2)0 "Medical Board of Review Report, June 1947," NARA Atlanta. RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 39, Medical Policy, 9. 168 community of biomedical radiation researchers by helping to build the means necessary to train researchers.

According to the Medical Board of Review, following through on the recommendation for recruitment and training, as well as all other aspects of AEC research in biology and medicine, hinged on effective organization. In a press release prepared at the conclusion of the Board of Review's meetings, the Board praised the MED and AEC for having controlled nuclear hazards thus far and attributed this feat to the close collaboration amongst scientific and technical experts as well as cooperation between government, academic, and industrial groups. The Medical Board of Review sought to preserve what they deemed to be the successful organizational structure developed during the war. Accordingly, their report recommended that an Advisory Committee for Biology and Medicine be created as quickly as possible and that a Medical Director be appointed promptly thereafter.291

SCIENCE & SOCIETY: THE AEC AS A LINK

Although the Medical Board of Review's report offered few suggestions that were entirely new, the report and the process taken to produce it served the purpose of giving greater authority to the AEC's role in supporting basic research.

The Commissioners anticipated that the AEC's research and development programs would not be isolated from public and political scrutiny. They sought to utilize the Board's expertise as a buffer against any forthcoming criticism and political maneuvering that might jeopardize the permanence of AEC-funded

2>" "Medical Board of Review Report, 1947," NARA Atlanta, Box 39, 10-11. 169 biomedical research. In fact, Lilienthal suggested that the Board members assume the role of evangelicals in promoting the importance of such factors as permanence or continuity to the advancement of basic knowledge.292 He felt this message would be more effective if delivered by the Board of Review rather than from the Commissioners or even scientists working within the AEC system.

Lilienthal's suggestion implies that he conceived of the role of the Board of

Review as more than just an advisory board to the AEC itself. He hoped that the

Medical Board of Review would play a broader role in shaping public perceptions and, to this end, the Commission planned to make the Board's report public.293

Lilienthal's actions reflect his understanding of the relationship between science and society as one that was intricately entwined. He relied on the Medical

Board of Review to provide expert opinion that would influence the development of a research program and, beyond that, American culture. The interaction between federal bureaucrats, such as the AEC's Commissioners, and elite scientists like those who comprised the Medical Board of Review, helped create the Cold War culture in which science was held in high esteem and the expansion of federal research and development infrastructure was an asset to the nation.

In an attempt to offer specific policy recommendations the Medical Board of Review helped delineate between work for which the AEC should be solely responsible and that in which the Commission might share responsibilities with universities or other government agencies. The former included work related to

2"2 "Transcript... Board of Review, 1947," NARA Atlanta, Box 40, 53-55. 2>n "Correspondence & Report of the Medical Board of Review, 20 June 1947," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945- 1952, Box 40, Board of Review, I. radiation safety and that which used the special resources the AEC possessed. As articulated in the report:

In control of operations with potentially dangerous materials, the Atomic Energy Commission has heavy responsibilities to its employees and to any others who could suffer from its negligence or its ignorance. With an absolute monopoly of new and powerful tools for research and important knowledge, the Atomic Energy Commission has obligations to share its acquisitions with the scientific world wherever security considerations permit.". ^94

The primary responsibility defined here was informed by both defense and civilian goals and the same was true for the AEC's shared responsibilities.

According to the Medical Board of Review, "...the Commission should collaborate offering the use of its equipment, staff experience, materials such as isotopes, and the services of its staff as teachers, lecturers, and consultants."295

The report made specific recommendations for collaboration between the AEC and the U.S. Public Health Service, the Armed Forces, and universities. To categorize the AEC as distinct from the scientific community or defense research as separate from civilian research was, already at this stage, difficult. The close connections between these spheres were institutionalized within biomedical radiation research with the establishment of the Advisory Committee on Biology and Medicine and Division of Biology and Medicine.

THE AEC'S BIOMEDICAL PROGRAM, ACBM, & DBM

Following the Medical Board of Review's endorsement, the AEC created the Advisory Committee on Biology and Medicine (ACBM) in September 1947.

This committee was comprised of seven biologists and physicians and was

~'M "Medical Board of Review Report, 1947," NARA Atlanta, Box 39, 2. "c>5 "Medical Board of Review Report, 1947," NARA Atlanta, Box 39, 4. 171 chaired by Dr. Alan Gregg, the Rockefeller Foundation's Director of the Division of Medical Sciences and former member of the Medical Board of Review.296 One of the first actions taken by the ACBM was to recommend the creation of the

Division of Biology and Medicine (DBM) to coordinate the AEC's biological and medical research. Dr. Stafford L. Warren was an obvious choice to lead the DBM since he had been Chief of the MED's Medical Office, but he was not willing to commit any more than one-third of his time to a position within the AEC.

Although he remained involved in MED and AEC activities during the immediate postwar years, he also planned his return to academia and agreed to become the first dean of a new medical school established at the University of California, Los

Angeles.297 The AEC and newly formed ACBM therefore considered numerous other individuals for the position of the DBM's Director and successfully recruited Dr. Shields L. Warren. As mentioned above, Shields Warren was a pathologist who had conducted medical work for the Navy during the war and who led the Navy's medical research in Japan following the war.

Following his appointment in October 1947, Warren went to Washington to join the AEC administration. He took up the task of managing ongoing research that had been carried over from the MED and implementing new research policies and programs that had emerged out of recommendations made by the MED's Medical Advisory Board, the AEC's Interim Medical Committee, and the NAS/NRC-organized Medical Board of Review. Throughout 1948 and

Three other members of the Medical Board of Review joined Gregg on the ACBM. See Hewlett and Duncan, Atomic Shield, 113-14. J)1 "Plan for Transfer1946," NARA College Park, Box 23, B-4; and ACHRE, The Final Report, 37. 172

1949 the DBM's administrative structure expanded. The positions of Deputy

Director and Executive Officer were created in September 1948 and February

1950.298 The former assisted Shields Warren with policy and program

development whereas the latter was responsible for the implementation of policy

and managing operations. The Executive Officer coordinated the activities of the

DBM with those of other AEC divisions and operations offices and also helped

coordinate the activities within the different branches of the DBM.299 The DBM's

scientific program was organized in a similar fashion to the wartime Health

Division at the Metallurgical Laboratory with three branches in Medicine,

Biology and Biophysics.300 The Chiefs and Assistant Chiefs acted as liaisons

between the managers and researchers at the AEC's laboratories and operations

offices and the Director, Deputy Director and Executive Officer in Washington.

As the basic administrative structure was put in place throughout 1948 and

1949, Shields Warren and the Division's staff helped build a foundation for a

broad biomedical program that was truly national in scale. For instance, in the

following chapter we will see that the DBM, in coordination with the Division of

Research, managed an increasingly large-scale radioisotopes distribution program

that provided researchers from around the country and even from international

"AEC Announces Two Appointments to Division of Biology and Medicine, 26 February 1950," NARA College Park, RG 326, General Correspondence 1946-1951, Box 23, Biology and Medicine, Organization and Functions, 1; "Memorandum by the General Manager: Appointment of Deputy Director, DBM, 13 February 1951," NARA College Park, RG 326, General Correspondence 1946-1951, Box 23, Biology and Medicine, Organization and Functions, 1. 2<>,) As we will see in the following chapter, the DBM worked with other divisions on various projects and programs, most notably the radioisotopes distribution program. "Position Description—Executive Officer, DBM, n.d.," NARA College Park, RG 326, General Correspondence 1946-1951, Box 23, Biology and Medicine, Organization and Functions, 1-2. 1,10 "Memorandum from Shields Warren to Managers of Operations: Designation of Washington Program Coordination in Biology and Medicine, 21 July 1950," NARA College Park, RG 326, General Correspondence 1946-1951, Box 23, Biology and Medicine, Organization and Functions, 1-2. 173

locations access to highly sought-after radioisotopes. The radioisotopes

distribution program is an excellent illustration of how the AEC's biomedical

program helped increase technology, facilities, and funding to support biomedical

radiation research. Indeed, the number of shipments of radioisotopes sent to

researchers grew significantly year-by-year. The number of shipments in 1947,

1948, and 1949—the first three full years of operation—were 1650, 3000, and

4700, respectively.301 One of the tangible outcomes of this expansion in the

radioisotopes program was the number of papers published on biomedical

research using radioisotopes. Before the AEC began distributing radioisotopes,

biomedical research using radioisotopes and related publication was limited by the supply of radioisotopes. Publications reporting on biomedical studies using

radioisotopes increased, though, and in 1948 alone, approximately 700 papers

Of)} reporting on such research were published.

The DBM also developed a range of training programs that helped build a community of researchers able to pursue research specifically related to problems concerning the AEC and that staffed the larger biomedical community. Amongst other initiatives, the AEC provided fellowships to doctoral and medical students to pursue radiation-related studies. Examined in greater detail in Chapter Five,

101 For precise data on radioisotopes distribution, see Chapter Four or W. E. Thompson, "Oak Ridge National Laboratory Research and Radioisotope Production, n.d.," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 31, Organization & Management, 6 - Isotopes Distribution, 32. ,(L "Part III: Increased Isotope Utilization from Isotopes - A Three Year Summary of Distribution, 16 September 1949," NARA College Park, RG 326, General Correspondence 1946-1951, Box 45, Isotopes - Three Year Summary of Distribution, 19. 174

the fellowship program awarded more than 260 fellowships in 1948, the year the

program was implemented. By 1951, 650 fellowships had been awarded.303

These are just a few illustrations of how the AEC's biomedical program

helped facilitate research and provide associated infrastructure. One might attribute the expansion of the AEC's biomedical program to an intrinsic source of

momentum created by the marriage of science and national defense during World

War 11. However, momentum within biomedical radiation research was continually fuelled by researchers' efforts to shape and adapt to a range of defense and civilian objectives. Some of these stemmed from early twentieth-century endeavors to develop radiation for medical purposes or from the wartime effort to safeguard workers from the radiation hazards created within the MED. Others were derived from and contributed to Cold War politics and culture as they evolved. That researchers drew on a range of external factors to support their pursuit of various objectives, shows continuity in their efforts to institutionalize biomedical radiation research. In the postwar political economy of research, though, their efforts to build disciplines extended beyond academic institutions to encompass new federal agencies such as the AEC.

The DBM provided a structure through which researchers from within the

AEC, other government and military agencies, and countless public and private institutions could decide the future course of biomedical radiation research. The researchers who took up administrative positions within the DBM were mostly

1", Kenneth S. Pitzer, "Fellowship Program - Ratio of Fellowships in the Physical and Biological Sciences, 14 February 1949," NARA College Park, RG 326, General Correspondence 1946-1951, Box 38, 1 and 6; and "Letter to Dr. Waterman from M. W. Boyer, 5 April 1951," NARA College Park, RG 326, General Correspondence 1946-1951, Box 38, 2. 175

new to the MED/AEC enterprise. Many of those who led the MED's health and safety effort, like the physicists who had worked for the Manhattan Project,

returned to academia within a few years after the end of the war. As mentioned,

Dr. Stafford L. Warren became dean of the new medical school at UCLA. His

MED assistant, Dr. Hymer L. Friedell, took up a position at Western Reserve

University. The Health Division's Dr. Robert S. Stone returned to California to work at Berkeley.

These researchers and others who relinquished their administrative roles were not, however, cut off from their AEC successors. The close network that had helped bring researchers together at the beginning of the war still existed and would continue to influence the AEC's biomedical radiation research. Shields

Warren was well acquainted with the MED team and consulted with them often.

For instance, in 1949 he organized a DBM meeting to consider the future of the

AEC's biomedical program. He invited MED alumni, including Stafford Warren,

Friedell, Stone, Simeon T. Cantril, and Louis H. Hempelmann, all of whom had resumed academic work.304 These individuals and others from the MED continued to provide policy and program advice and also conduct research for the

DBM via contract. The network that formed early in the century and was reinforced during the war, helped bridge the gap between the government and academic spheres in the postwar years.

The ACBM also helped bridge that gap. The ACBM was an advisory committee to the AEC, not the DBM, and was not in the practice of approving or

,<>4 "Memorandum to Commissioners and General Secretary re Meeting Organized by Shields Warren, 5 June 1949," NARA College Park, RG 326, General Correspondence 1946-1951, Box 32, Biology & Medicine, 1. disapproving the DBM's programs. However, the members of the ACBM did consider it a responsibility to help the DBM identify important scientific problems and design appropriate research programs that made good use of the AEC's facilities and personnel.305 The ACBM provided an opportunity for leading researchers to help shape AEC policies and programs. As we will see in Chapter

Five, the ACBM served as an authoritative voice when the AEC faced decisions regarding its role in funding ambitious training programs at universities.

It was intended that the ACBM would be able to offer new perspectives based on wide-ranging expertise so membership was established on a rotating basis with one member being replaced every year. To ensure that the ACBM's policy advice and input on specific DBM research programs was well informed, the Committee held at least three of its five meetings at AEC research facilities every year. Within just a couple of years after it was created, the ACBM's members saw themselves as a "nexus" between the AEC and the medical and biological academic communities.306

CONCLUSION

Wartime research yielded many scientific and technological advances, especially in fields related to atomic energy. In terms of biomedical research, the

MED's health and safety program shifted the overall objectives of biomedical radiation researchers, but not drastically so. Before the war researchers were particularly interested in investigating the biological effects of radiation in an

305 "Correspondence between Lewis Strauss and Joseph Wearn, 1953," NARA College Park, RG 326, General Correspondence 1951-1958, Box 67, Organization and Management 7 - Division of Biology and Medicine, Advisory Committee, 1. "Correspondence...Strauss and Wearn, 1953," NARA College Park, Box 67, 1. 177 effort to develop clinical uses of radiation, but also in an effort to protect against harmful exposure. During the war, the latter became a more pressing concern since radiation hazards increased dramatically as the Manhattan Project progressed. This shift in priorities influenced postwar research in that radiation safety remained important to the nuclear weapons and civilian atomic energy industries managed by the AEC. However, the more significant legacy of wartime biomedical radiation research and development was organizational.

Researchers learned how to advance their own objectives in relation to national priorities and to do so within a large-scale government and military bureaucracy.

The biomedical radiation researchers who had participated in the

Manhattan Project emerged from the war with a considerable amount of authority to influence postwar demobilization. Although many MED researchers returned to academia when the war ended, they used their newly acquired authority to ensure that demobilization did not result in the dismantling of the research infrastructure created during the war. Based on the work of the MED's Medical

Advisory Committee, the AEC's Interim Medical Committee, and the

NAS/NRC's Medical Board of Review, it is evident that researchers both within the MED/AEC system and from outside of that bureaucracy sought to preserve the partnership between science and the state in the postwar period. Researchers encouraged the formation of the DBM and ACBM as the means through which government-funded biomedical radiation research could be continued and expanded. They promoted the necessity of the DBM and ACBM by defining the relevance of biomedical research to the social, political, and economic context of 178

Cold War America, Such actions were not unlike those of researchers who, a decade earlier, promoted their research in relation to the problem of cancer. For the AEC, radiation safety was an essential problem to address, but was not the only driving force behind the Commission's biomedical research. The public relations benefit derived from a broad biomedical program was also an important factor.

The following chapters examine the AEC's radioisotopes distribution program, training initiatives, and cancer program which were the central features of the AEC's biomedical radiation research. The creation and development of those programs show that the DBM and ACBM served as an important forum in which researchers and policy makers both responded to and helped define national priorities—priorities that were a mix of defense and civilian matters and not easily categorized as only one or the other. The DBM and ACBM provided the means through which the AEC could create a research community capable of catering to the Commission's specific needs and those of the larger scientific community as well. Indeed, the AEC's biomedical program and infrastructure helped academic researchers obtain research tools and funding, not only for research projects, but also for the institutionalization of hybrid disciplines rooted in the interdisciplinary research practices established earlier in the century. 179

CHAPTER 4

THE BUILDING BLOCKS OF RESEARCH: RADIOISOTOPES DISTRIBUTION

Production of tracer and therapeutic radioisotopes has been heralded as one of the great peacetime contributions of the uranium chain-reacting pile. This use of the pile will unquestionably be rich in scientific, medical, and technological applications. 307

These were the opening lines of a Manhattan Engineer District (MED) report that announced the availability of radioisotopes for distribution to researchers.

Published in Science in June 1946, this announcement publicized the launch of the MED's radioisotopes distribution program, a program that, as implied in the quotation above, aimed to make use of an MED reactor to provide materials for research that was unrelated to weapons production. Just a year after the start of the radioisotopes distribution program, an article published in Time Magazine celebrated the "important influences" that radioisotopes had already had on

IAD ^ American science. The title of that article, "Science: A Year of Isotopes," is telling. At a time when so many fields of research and new technologies were generating considerable excitement within the American scientific community and beyond, radioisotopes were a research tool that inspired great anticipation.

What role did the radioisotopes distribution program play in the broader

MED's and the Atomic Energy Commission's (AEC) biomedical research program? And, how did the distribution program affect the professional connections amongst biomedical radiation researchers and between researchers and the state? The program provided isotopes, most of which were

,(l7 N.A., "Availability of Radioactive Isotopes: Announcement from Headquarters, Manhattan Project, Washington, D.C.," Science, 14 June 1946, 697. ,(IS N.A., "Science: A Year of Isotopes," Time Magazine, 11 August 1947, 74. 180 radioisotopes—radioactive or unstable isotopes—for researchers to pursue a wide range of research in biology and medicine.309 This allowed for numerous researchers to pursue biomedical radiation research that, through the 1930s and early 1940s, was limited to only those who had access to cyclotrons. The distribution of radioisotopes also provided impetus for training more biomedical radiation researchers. By providing research tools and making obvious the need for more researchers trained to work with radioisotopes, this program generated momentum within numerous fields of biomedical research, some of which were in their infancy.

Most importantly, this chapter argues that radioisotopes and the technologies used to produce them acted as a link between academic science and medicine, on the one hand, and the national security concerns of the government and military, on the other. Historians Angela Creager and Maria Jesus

Santesmases have argued in their examination of radioisotopes and the postwar expansion of biology that the federal government was instrumental in establishing the wider use of radioisotopes in research.310 This dissertation endeavors to build on their argument by examining the various groups and individuals involved in the "government" distribution of radioisotopes. As we will see, much of the

1(l'' An isotope is a form of any chemical element. Elements are determined by the number of protons in the nucleus of an atom. For any one element, the number of protons never changes. The number of neutrons does vary, however, resulting in different isotopes of elements. Radioisotopes are isotopes that are unstable and undergo radioactive decay. See W. E. Thompson, "Oak Ridge National Laboratory Research and Radioisotope Production, n.d.," NARA Atlanta, RCi 326, New York Operations Office - Research & Medicine Division Correspondence 1945- 1952, Box 31, Organization & Management, 6 -Isotopes Distribution, 11. "" See Creager, "Nuclear Medicine in the Service of Biomedicine," 649-84; Angela N. H. Creager, "Phosphorous-32 in the Phage Group: Radioisotopes as Historical Tracers of Molecular Biology," Studies in the History and Philosophy of Biological and Biomedical Sciences 40 (2009): 29-42; Angela N. H. Creager, "Wendell Stanley's Dream of a Free-Standing Biochemistry Department at the University of California, Berkeley," Journal of the History of Biology 29, no. 3 (1996): 331-60. Creager and Santesmases, "Radiobiology in the Atomic Age," 637-47. 181 stimulus to create the distribution program came from outside of the MED and

AEC. Even those within these organizations who encouraged the creation of the distribution program did so, in large part, to extend the resources available within the MED and AEC outward, primarily to academic institutions.

The development of the distribution program preceded, but had many parallels with, the formal creation of the AEC's Advisory Committee on Biology and Medicine (ACBM) and Division of Biology and Medicine (DBM). The most significant was the role of researchers in creating and shaping MED and AEC research and development policies such that these agencies incorporated pre­ existing biomedical research priorities into national research programs that might otherwise have been entirely defense-oriented. Indeed, the distribution program marked an immediate effort to shift research away from military matters. That is, researchers sought to initiate the distribution of radioisotopes in an effort to benefit civilian research in biology, medicine, chemistry, and physics, and to do so far beyond the walls of military and government facilities.

The wartime experience of government-funded, defense-oriented research did, however, help create the climate in which researchers were able to influence the development of this program. Furthermore, the establishment of the distribution program extended the wartime practice of research funded by and organized within a government agency. As part of the MED and later the AEC, the distribution program was never isolated from those organizations' other business which was, predominantly, nuclear weapons production. This particular program, like the AEC's broader biomedical research policies and program, 182

reflected the merging of the scientific community, government, military, and

industry, all of which shaped the scientific enterprise in the postwar years.

POSTWAR RECOVERY: RESEARCHERS REDEFINE PRIORITIES

Cyclotron-produced radioisotopes had captured the attention of

biomedical researchers by the mid-1930s. However, the scarcity of cyclotrons

and their limited ability to produce radioisotopes prevented many researchers

from employing them in research. During the war, nuclear reactors, while developed to supply materials for the bomb, also produced radioisotopes and were able to do so in much greater quantities than cyclotrons. Thus, when the war ended researchers within the MED sought to make reactor-produced radioisotopes available for purposes beyond the scope of the Manhattan Project. Many

biomedical researchers who knew nothing of the project throughout the course of the war were eager to reap some of the benefits of wartime developments. They began to learn about nuclear reactors and their ability to produce radioisotopes in

August 1945 when Manhattan Project Physicist Henry De Wolf Smyth published an official account of the Manhattan Project.311 As Smyth explained three decades later, his report was meant to inform the public—scientific community included—of the civilian and military possibilities stemming from the Manhattan

Project. The MED administration prompted the publication of the report based on

,n Smyth, A General Account of the Development of Methods of Using Atomic Energy for Military Purposes under the Auspices of the United States Government, 1940-1945. 183 the premise that at the end of the war there were policy decisions to be made in which an informed public could participate.312

Within the scientific community, the so-called "Smyth Report" served as a form of publicity. It created an illusion that radioisotopes were readily available, although, at the time of publication, they were not.313 Soon thereafter, just months after the atomic bombings in Japan and the end of the war, scientists began to press government and military officials for the creation of a government- sponsored radioisotopes distribution program. Frank B. Jewett who, as we saw in the previous chapter was the president of the National Academy of Sciences

(NAS) and National Research Council (NRC) from 1939-1947, was one individual who was excited about the prospects of using reactor-produced radioisotopes for biomedical research and took steps to encourage a distribution program. Writing to the newly appointed Secretary of War Robert P. Patterson in

October 1945, Jewett explained the importance of radioisotopes to medical research—an importance, he insisted, that had been "amply demonstrated" by the late 1930s. He drew particular attention to the tracer work that researchers had accomplished with the "inadequate" supply of cyclotron-produced radioisotopes and encouraged the distribution of reactor-produced radioisotopes to further this

Henry Dc Wolf Smyth, Atomic Energy for Military Purposes: The Official Report on the Development of the Atomic Bomb under the Auspices of the United States Government, 1940-1945 (Stanford, CA: Stanford University Press, 1989), 301. 11' "Letter from Joseph G. Hamilton to Hymer L. Friedell, 30 November 1945," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence, 1945- 1952, Box 36, I.I.B. California Area, 1; and "The National Distribution of Radioisotopes from the Manhattan Engineer District, 3 January 1946," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 36, Isotopes, 2. 184

field of research.314 Among the many reasons Jewett presented to support his

request for a distribution program, he noted that reactor-produced radioisotopes

were of little use for military purposes. They were produced as byproducts during

the production of fissionable material. Thus, it seemed commonsensical to make

use of them in biomedical research.

Jewett explicitly asked that the War Department give the MED permission

to provide radioisotopes to the National Research Council (NRC) which would organize distribution to researchers, at least on an interim basis. He anticipated that a new agency would likely be established to replace and continue the work of the MED and that the new agency might organize the proposed program itself.

Jewett justified his request by explaining the history of the NRC as the operating arm of the National Academy of Sciences. Since the NRC was created in 1916 it had provided expert advice and administrative services to numerous governmental agencies. The Council also provided its services to philanthropic foundations and national societies such as the American Cancer Society (ACS). In fact, Jewett linked his request for radioisotopes distribution to cancer research and the new role the NRC was playing in advising the ACS on the creation and administration of an extensive cancer research program.315 Jewett's letter illustrated that the

NAS/NRC was well positioned to serve as an intermediary between varying interests given the Academy's and Research Council's relationship to the

114 "Letter from Frank B. Jewett to Honorable Robert P. Patterson, 18 October 1945," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 36, Requests for Radioactive Materials for Institutions, I. "5 As 1 discussed in Chapter 1, the American Cancer Society established a research program in 1945. This was a significant landmark in the history of cancer research. At that time, the ACS allocated $1 million for research which constituted approximately 40% of all funding—both public and private—for cancer research in the United States. "Letter from Jewett to Patterson, 1945," NARA Atlanta, Box 36, 1-2; and Ross, Crusade, 210-14. 185 scientific community, broadly, specific organizations such as the ACS, as well as the federal government and military. This was especially true following the war.

A biographical memoir of Jewett explains that to "a large extent the Academy, during [the war], ceased to function as a learned society in the traditional sense, and functioned in its official capacity as the top scientific advisory agency of the

Government."316 The relationship of the NAS/NRC to the military and government underlines the fact that research and development had evolved such that there was a great demand for researchers who were able to organize and administer state-funded research in all scientific fields, biomedical included.

Just weeks after Jewett offered the services of the NRC to establish a temporary isotopes distribution program, the MED's General Leslie R. Groves received a letter from John R. Dunning, a Columbia University physicist, who also encouraged the MED to establish a distribution program. Dunning and

Groves were well acquainted since Dunning had led the Manhattan Project's gaseous diffusion project at Columbia.317 He wrote to Groves on behalf of the

Radioisotopes Research Committee, a group comprised of researchers from various universities, medical schools, and other public and private institutions who engaged in biological, medical, and chemical research with radioisotopes and stable isotopes.318 Representing his fellow radiation researchers, Dunning emphasized the importance of radioisotopes to various fields of research. "The

316 Oliver E. Buckley, "Frank Baldwin Jewett, 1879-1949," in Biographical Memoirs (Washington, DC: National Academy of Sciences, 1952), 253. ,|7 Herbert L. Anderson, "John Ray Dunning, 1907-1975," in Biographical Memoirs (Washington, DC: National Academy of Sciences, 1989), 163-80. For recollections from General Groves, see page 175. ,|S "Letter from John R. Dunning to General Leslie R. Groves, 29 October 1945," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945- 1952, Box 36, Requests for Radioactive Materials for Institutions, 1. 186 application of radioisotopes in the biological-medical fields, as well as for physical, chemical, and industrial uses," Dunning argued, "have thus far only been barely touched upon." However, if the MED were to provide radioisotopes to researchers, he believed "it would be one of the government's most important contributions to science."319 Dunning identified carbon-14 (C-14) and hydrogen-

3 (H-3) as the two most important radioisotopes that the MED could distribute to researchers because these radioisotopes were extremely useful in tracer and cancer research. Raising the issue of their limited production in cyclotrons,

Dunning echoed what was becoming a very familiar argument employed to encourage the formation of an MED radioisotopes distribution program.320

Dr. Joseph G. Hamilton, director of the Crocker Laboratory at the

University of California, Berkeley delivered a similar message to Lieutenant Dr.

Hymer L. Friedell who was still serving as Deputy Chief of the MED's Medical

Office under Colonel Stafford L. Warren's leadership. Having led an extensive study of the metabolism of various radioisotopes during the war, Hamilton sought to capitalize on his MED connections to request a supply of C-14. In November

1945 he wrote to Friedell, who he knew well, not only from their wartime work, but also from their pre-war collaboration at the Berkeley Rad Lab. Hamilton told

Friedell that an increasing number of researchers had contacted him hoping to obtain C-14 for biomedical research on the metabolism of carbon in the body and biological research investigating photosynthesis in plants. However, Hamilton was forced to deny many requests because the Crocker cyclotron could not

"Letter from Dunning to Groves, 1945," NARA Atlanta, Box 36, 1. 32(1 "Letter from Dunning to Groves, 1945," NARA Atlanta, Box 36, 2. 187

produce C-14 in sufficient supply. The same was true of other radioisotopes that

were valuable in biomedical research.

Hamilton asked that the MED provide him with C-14 which he could then

distribute to researchers both at Berkeley and beyond. He emphasized that the

publication of the "Smyth Report" three months earlier had created the illusion

that radioisotopes were readily available and warned that if they were not

provided to researchers they might assume that the MED was withholding them

for administrative reasons. This, he argued, would amount to a public relations

problem.321 He did, however, make clear that he only wished to obtain C-14 from

the MED on the ad hoc basis he suggested, if to do so did not jeopardize the

future establishment of a national program for the distribution of radioisotopes.322

To him, an organized, long-term program was a more important goal than achieving immediate, but temporary access to radioisotopes.

Jewett, Dunning, and Hamilton all seemed confident that the government

would establish a radioisotopes distribution program. Their confidence reflected

their recognition of their own and their colleagues' ability to influence government policy. They were part of a community of researchers that, due to their wartime work, had established an elite status amongst their scientific peers and amongst the military and government officials with whom they worked.

They had gained access to avenues through which they could influence policy.

To do so, biomedical researchers and scientific administrators negotiated a political and cultural climate in which American citizens were wary of the atomic

"Letter from Hamilton to Friedell, 1945," NARA Atlanta, Box 36, 1. ,22 "Letter from Hamilton to Friedell, 1945," NARA Atlanta, Box 36, 1-2. 188

bomb and policymakers were eager to promote civilian uses of the newly created

Manhattan Project technologies and scientific advances. They increasingly

employed public relations rhetoric as a means to achieve certain goals when

working within the MED and later the AEC bureaucracy, predicting, for instance,

that a public relations problem would ensue if the scientific community believed

the MED was withholding radioisotopes from them. Biomedical researchers did

not, themselves, consider public relations problems to be a primary concern.

Rather, they advanced such warnings in an effort to obtain access to radioisotopes.

As these individuals advocated for a radioisotopes distribution program and assumed that such a program was inevitable, the question of when the program would or could be initiated remained. From the standpoint of late 1945,

Jewett, Dunning, and Hamilton suspected that the program would be administered by a new agency formed to succeed the MED. Dunning told Groves that he hoped the MED would not wait on the creation of the AEC to act. Jewett had expressed a similar sentiment in his letter to Secretary of War Patterson. Until a new agency was established, Hamilton offered to serve as a middle man in distributing radioisotopes to researchers. Similarly, Jewett and Dunning conveyed their own willingness to help the MED create an interim program.

They were also able to offer the service of their affiliated organizations—the NRC and the Radioisotopes Research Committee—to temporarily manage a distribution program.3i3

1-1 "Letter from Dunning to Groves, 1945," NARA Atlanta, Box 36, 1-2. 189

The assumptions these individuals made regarding the eventual establishment of a isotopes program reflect the immediate postwar environment in which scientists, as well as government and military officials, placed a great deal of value in research and development. This was also a moment at which the direction of the MED's operations was unclear. The MED was still primarily concerned with defense-oriented research and development, such as ongoing weapons production and the aftermath of the bombings in Japan. However, the end of the war initiated a transitional period throughout which researchers attempted to integrate some of their own goals into the MED's agenda. Like the researchers who advocated for biomedical research in Japan shortly after the bombings in Hiroshima and Nagasaki, those who pressed for radioisotopes distribution seized the immediate postwar period to redefine research priorities.

For them, the distance between the MED's defense priorities and their own goals of facilitating the use of radioisotopes in civilian research, was not insurmountable.

Jewett's, Dunning's, and Hamilton's correspondence to Secretary of War

Patterson and the MED's General Groves and Lieutenant Friedell used rhetoric that inspired optimism about the future of biomedical radiation research. They were also pragmatic, though, about ongoing national security objectives and the balance that might be established between defense and civilian research. The dialogue they initiated was intertwined with the broader issue of defining postwar biomedical radiation research which, as examined in the previous chapter, led to 190

the creation of the AEC's Advisory Committee on Biology and Medicine and

Division of Biology and Medicine.

The correspondence examined here helps to illuminate the various

institutions that were intricately involved in biomedical research by the end of the

war. These included, but were not limited to, the National Academy of Sciences-

National Research Council, private organizations such as the American Cancer

Society, universities, military and government agencies. While individuals

affiliated with these institutions, such as Jewett, Dunning, and Hamilton, sought

to wrest some control over radioisotopes production, they were not able to shift

complete control over this program or other biomedical programs and policies

away from the MED and later, the AEC. They worked within, and for some,

became part of the military-government bureaucracy. As plans for a distribution

program moved forward we see that radioisotopes and the technologies used to

produce them acted as a link between science, medicine, the federal government,

and military.

LOCATING A DISTRIBUTION CENTER: OAK RIDGE NATIONAL LABORATORY

Throughout the final months of 1945, as scientists and MED personnel

discussed the possibility of establishing an isotopes distribution program, they

considered different locations for the program. The selection of a location

seemed dependent on obtaining the use of a reactor capable of meeting the

production needs of the program. However, this was not a decision based simply on technological capabilities. Reactors were built at three different locations during the Manhattan Project. The first was the experimental reactor located at 19! the University of Chicago's Metallurgical Laboratory. Later in the war this reactor was moved to the nearby Argonne Laboratory. The second was the pilot- plant reactor at Clinton Laboratories in Oak Ridge, Tennessee. And finally, three full-scale reactors were constructed at the Hanford Engineer Works in Richmond,

Washington. Hereafter I refer to these facilities as Argonne, Oak Ridge, and

Hanford to avoid confusion since each site assumed different names or designations as national laboratories along varying timelines.324 Of these reactors and sites, numerous factors made the Oak Ridge reactor the likely choice for a distribution program. Oak Ridge biochemist Waldo E. Cohn justified the selection of the Oak Ridge reactor in an MED report he drafted in January 1946.

As we will see, Cohn was one of the Oak Ridge scientists who played a major role in developing the program. According to Cohn, the Argonne reactor was not suitable for the distribution program because it was a low-powered experimental reactor and was not able to produce as many radioisotopes as either the Oak Ridge or Hanford reactors. The Oak Ridge reactor was of "intermediate power" which

Cohn judged as sufficient "to permit the creation of fairly large quantities of many radioisotopes." The Hanford reactors had the greatest output in terms of sheer quantity. They could "out-produce by many fold" the Oak Ridge reactor.

However, the Hanford reactors were not able to produce as many different radioisotopes as the Oak Ridge reactor. Cohn believed, therefore, that the Oak

,24 The Argonne Laboratory developed around the original reactor constructed at the Met Lab. The Laboratory expanded following the war and was designated as the Argonne National Laboratory in 1946. Clinton Laboratories was also renamed and designated as Oak Ridge National Laboratories in 1948. The Hanford Engineer Works has acquired many names due to changes in administration and the sheer number of laboratories and production facilities located at the site. See Westwick, The National Labs, 8-9 and 37; and Thompson, "Oak Ridge...Radioisotope Production, n.d.," NARA Atlanta, Box 31,1. ' 5 "Distribution of Radioisotopes, 1946," NARA Atlanta, Box 36, 2-3. 192

Ridge reactor was the best choice of the three, as it could produce an ample

supply of the greatest variety of radioisotopes. Cohn also explained that Oak

Ridge had appropriate facilities and staff to manage both the production and

chemical separation of radioisotopes.'^26

Aside from the reactors built during the Manhattan Project, researchers

and MED administrators also considered establishing a distribution center at

Berkeley. It was, after all, a leading site of radiation research to which

researchers from other institutions looked for guidance and materials. This was

evident by the early postwar trend of researchers flocking to Joseph Hamilton in

the attempt to obtain radioisotopes from the 60-inch cyclotron at Crocker

Laboratory. Running a radioisotopes distribution program, was not a job that

researchers at Berkeley were, in fact, keen to do. Since the discovery of

radioisotopes, the Rad Lab had provided them to other Berkeley researchers as

well as to researchers affiliated with other institutions. However, the Rad Lab did so on a very small scale and to do so required considerable time and labor. Rad

Lab biophysicist and medical physicist Cornelius Tobias recalled in an interview

that,

It turned out that Ernest Lawrence. . .was not anxious to become a supply agency. We talked, you know, quite a bit about who is going to do this. Each time, you're pointing at each other, 'I don't have to do this anymore; you do it.' And it turned out that the milieu of the Lawrence Berkeley Lab was not really ready to become a supply agency.327

,26 "Distribution of Radioisotopes, 1946," NARA Atlanta, Box 36, 2-3. ,:7 Oral History of Cornelius Tobias, Ph. D., interview conducted January 16, 1995, as part of the Office of Human Radiation Experiments' Oral History Project, Human Radiation Studies: Remembering the Early Years. Conducted by Prita Pillai and Anna Berge in Eugene, OR, n.p.. 193

Tobias elaborated that Lawrence considered radioisotopes production to be an

"everyday job" that would distract from the cutting edge research he expected the

Rad Lab staff to pursue.328 Tobias himself breathed a sigh of relief when the

MED began making plans for the program to operate out of Oak Ridge. He too

felt that producing radioisotopes on a large scale was not the greatest use of the

Rad Lab's time and talent. Nor was it the sort of work that interested him. He

recalled that when cyclotrons were the only source of radioisotopes he "had to do

a lot of the chemistry, which turned out to be very time-consuming and

thankless...." Furthermore, he said, "It just didn't seem to me like it's a kind of

science [that] I, myself, would want to do. I was anxious to be at the forefront of

research."329 According to Tobias, while he and his colleagues at the Rad Lab

were in full support of the proposed radioisotopes distribution program, they did

not want to take responsibility for it.

There was little doubt that establishing and operating a radioisotopes

distribution program would be time-consuming. At the beginning of 1946, two

individuals, in particular, assumed responsibility for the program—Paul C.

Aebersold and Waldo E. Cohn. The former was very enthusiastic about, and eager to take charge of the program. The latter was less keen to devote his time to

radioisotopes production on a long-term basis since this work was outside of his specialized field of research. He was, however, willing to do so temporarily to help establish the program. Cohn, like many Manhattan Project alumni, was committed to facilitating a permanent relationship between the government and

12i< Oral History of Cornelius Tobias, 1995, n.p.. ,29 Oral History of Cornelius Tobias, 1995, n.p.. 194

scicntific enterprise even though this required him to make compromises between

his own research interests and the larger and interconnected interests of the

scientific and medical community and the state.330

As mentioned in the first chapter, both Aebersold and Cohn started their

scientific careers at Berkeley where Aebersold completed his Ph.D. in Physics

and Cohn completed his in Biochemistry. Cohn left Berkeley in 1939 to further

his research at Harvard University's Medical School before joining the Manhattan

Project. During the war he worked at both the University of Chicago's

Metallurgical Laboratory and then Oak Ridge. When Aebersold became involved

in the Manhattan Project, he remained at Ernest O. Lawrence's Rad Lab, at least

initially. His wartime work did take him to new locations, though, as he moved to

the National Bureau of Standards in Washington, D.C. and then to Los Alamos

Laboratory in New Mexico. At the end of the war Aebersold was asked to go to

Oak Ridge to become chief of the MED's Isotopes Branch which was organized within the MED's Division of Research. There he was reunited with his old Rad

Lab colleague, Cohn. From the time of Aebersold's arrival in January of 1946 they both became highly involved in launching the isotopes distribution program.

Oral History of Waldo E. Cohn, Ph. D., interview conducted January 18, 1995, as part of the Office of Human Radiation Experiments' Oral History Project, Human Radiation Studies: Remembering the Early Years. Conducted by Thomas Fisher, Jr. and Michael Yuffee in Oak Ridge, TN, n.p.. 195

Waldo E. Cohn and colleague irradiating material in the graphite reactor at Oak Ridge, c. late 1940s. Photo Credit: Oak Ridge National Laboratory Historical Image Gallery.

Paul C. Aebersold, c. early 1940s. Photo Credit: Lawrence Berkeley National Laboratory Image Library.

Under Aebersold's and Cohn's direction the distribution program was created quickly. Aebersold organized the administrative aspects of the program whereas Cohn was in charge of production. Cohn reports, though, that they were

in close collaboration and did not hesitate to offer advice to one another.

Together they promoted the value of the program they were designing to General

Leslie R. Groves and other MED administrators. As Cohn later recalled, he and

Aebersold received ample encouragement from the MED administration to forge 196

ahead with the distribution of radioisotopes. His perspective was that "Obviously,

the administration of the Project was all in favor of this because here was a [long-

term] use for the [Oak Ridge] Graphite Reactor."331 He explained that the Oak

Ridge reactor had been designed as a pilot plant primarily for research and

development purposes, whereas the three full-scale reactors completed at Hanford

were built to produce plutonium for atomic bombs. This was an important factor.

The Oak Ridge reactor was not needed to supply plutonium for weapons.

Towards the end of 1945 and beginning of 1946 when the isotopes program began

to take form, national security objectives were still the MED's priority. If

Hanford were not able to produce sufficient plutonium for bombs, the use of the

Oak Ridge reactor for the distribution program would have conflicted with the ongoing objective of producing nuclear weapons. Here we see that the isotopes program was not directly related to the MED's and AEC's defense research and development, but was not isolated from this work. The production and distribution of radioisotopes could co-exist and not conflict with defense priorities.

THE BUSINESS OF ADMINISTRATION: A JOB FOR BOTH THE MED & FOR EXPERTS

Aebersold and Cohn spent the first half of 1946 organizing the administrative framework of the distribution program and refining the processes required to produce and separate radioisotopes at Oak Ridge. The latter was work for Cohn's section of the Chemistry Division which was staffed by 15 men.332 As

Oral History of Waldo EL Cohn, 1995, n.p.. "Distribution of Radioisotopes, 1946," NARA Atlanta, Box 36, 4. 197

already mentioned, Aebersold's Isotopes Branch, which was comprised of scientists employed by the MED at Oak Ridge, was in charge of administering the

program. The staff of the Isotopes Branch grew as the program expanded. For

instance, between 1948 and 1949 the number of staff increased from 36 to 43.333

The Isotopes Branch received requests from researchers and assessed the

technical aspects associated with each request. From there, the Isotopes Branch depended on an administrative framework that was external to the MED. In his

January 1946 report Cohn outlined a proposed system of advisory committees to compliment the Isotopes Branch. It was these advisory committees, in particular,

that facilitated a merger between scientists and the government. The committees

Cohn proposed included a Policy Committee which he felt should consist of men appointed by the President of the NAS or a similar scientific body. He stressed that these individuals should represent an array of disciplines so that all disciplines which then and which may in the future use radioisotopes in research would have a voice through which to affect policy. Cohn also described an

Advisory Panel and Allocation Board. The former would function as referees in evaluating radioisotopes requests and advise the Policy Committee. He provided the names of many men who could serve on the Advisory Panel, most of which were MED alumni. Cohn envisioned the Allocation Board as an operational group that would arrange for and ensure that successful radioisotopes requests

"Atomic Energy Commission Study of Wider Use of Isotopes, 22 June 1948," NARA College Park, RG 326, General Correspondence 1946-1951, Box 45, Study of Wider Use of Isotopes, 4; and "Isotope Distribution Program, 11 October 1951," NARA College Park, RG 326, General Correspondence 1951-1958, Box 28, Isotope Program - Distribution Vol. 1, 35. 198 were prepared for shipment in accordance with the policies established by the

Policy Committee.334

The administrative guidelines Cohn laid out in his report were, for the most part, adopted when the program's committees were formed. Neither

Aebersold nor Cohn had the authority to create the committees that would administer the program. However, they had direct access to MED officials, including General Groves. Cohn recalled that although Groves sometimes had different ideas about various aspects of the radioisotopes program, he was, Cohn believed, very reasonable about most matters.335 In terms of forming the advisory committees for the distribution program Aebersold and Cohn went up the chain of command to General Groves who authorized what was, by then, a well- established method of recruitment. Groves asked NAS/NRC President Frank

Jewett to assume responsibility for assembling a group of leading radiation researchers to serve as a policy committee for the distribution program. Just as

Cohn had recommended in his report, the researchers selected to serve on the policy committee were chosen carefully to ensure broad representation of numerous disciplines. Ten individuals were selected to serve as the Interim

Committee on Isotopes Distribution Policy, two each to represent physics, chemistry, medicine, biology, and applied science.336 One of the physicists, Lee

,14 "Distribution of Radioisotopes, 1946," NARA Atlanta, Box 36, 11-16. 1,5 Oral History of Waldo E. Cohn, 1995, n.p.. 1h The full list of researchers who served on the Interim Committee on Isotopes Distribution Policy includes: representing physics Lee A. DuBridge (University of Rochester and the California Institute of Technology) and Merle A. Tuve (Carnegie Institution of Washington); representing chemistry Linus Pauling (California Institute of Technology) and Vincent du Vigneaud (Cornell University Medical College); representing medicine Cornelius P. Rhodes (director of Memorial Hospital) and Cecil J. Watson (University of Minnesota Medical School); representing biology Raymond E. Zirkle (University of Chicago) and A. Baird Hastings (Harvard 199

A. DuBridge, was appointed to serve as chair of the committee. He was not new

to the government-military science enterprise. DuBridge had been recruited to do

defense work during the war. He relocated to the Massachusetts Institute of

Technology to direct the Radiation Laboratory or Rad Lab. Not to be confused

with Ernest Lawrence's Rad Lab at Berkeley, the MIT Rad Lab was the site of

radar research and development. 337 DuBridge was also well acquainted with

Chief of the MED's Medical Office, Dr. Stafford L. Warren. He and Warren had

been colleagues since the 1930s when both were faculty at the University of

Rochester and involved in the construction and operation of that University's first cyclotron.338

The Interim Policy Committee would review existing practices and continue to define priorities. Aebersold served as its acting secretary so that he could coordinate activities and policies with the MED. From the outset, the

Isotopes Branch and the Interim Policy Committee decided to promote the use of radioisotopes for small-scale research and for therapeutic purposes, both of which they expected to lead to publishable research. The program was not intended to supply radioisotopes or stable isotopes for commercial purposes. Prioritization was necessary because, at that time, supply was limited.339

The Interim Policy Committee was also supported by two subcommittees, one that dealt with requests for radioisotopes to be supplied to researchers outside

University Medical School), and Zay Jeffries (General Electric); and representing applied sciences L. F. Curtiss (National Bureau of Standards). See, N.A., "Availability of Radioactive Isotopes," 697-98. 1,7 For further information on DuBridge's wartime work at the MIT Rad Lab, see Buderi, The Invention That Changed the World. See, Chapter One. w "Isotope Distribution, 1951," NARA College Park, Box 28, 1. 200

of MED operations and the other that dealt with requests for radioisotopes

research involving humans. The first was known as the Interim Advisory

Subcommittee on Allocation and Distribution. It acted as a gatekeeper in that it was responsible for evaluating the qualifications of the researchers and

institutions that requested radioisotopes and stable isotopes. This Subcommittee also assessed the scientific merit of proposed research. Again, these were

important functions due to the limited supply of radioisotopes available in the early days of the program. The second subcommittee was the Interim

Subcommittee on Human Application. It was charged with evaluating the individuals, institutions, and research programs associated with any research that involved human subjects.340 The members of each subcommittee were nominated by the Interim Policy Committee and appointed by General Groves. Many were individuals examined in previous chapters, such as Dr. Joseph G. Hamilton, Dr.

Hymer L. Friedell, and Gioacchino Failla.341

By the end of the war, these individuals were well known for their research expertise and also for their administrative skills. Their service on these committees did not require them to leave their academic institutions, but it did keep them well connected to government and military research and development.

The policy committees of the isotopes program provided an avenue through which prominent researchers from numerous institutions were able to collaborate in an effort to shape policies that affected many aspects of academic research,

",4° N.A., "Availability of Radioactive Isotopes," 698. 141 The Advisory Subcommittee on Allocation and Distribution included K. T. Bainbridgc (Harvard University), J. W. Kennedy (Washington University, St. Louis), J. G. Hamilton (University of California, Berkeley), and P.C. Aebersold (MED). The Subcommittee on Human Application included Andrew H. Dowdy (University of Rochester), Hymer L. Friedell (Western Reserve University), Gioacchino Failla (Columbia University). See Ibid.. 201

mcdical practice, and policies and programs related to domestic and foreign

research and development. The role of these various researchers within the MED

administration and later the AEC administration highlights the increasingly

blurred lines between "military" and "government" on the one hand and

"scientific," "medical," and "academic" on the other. For instance, the Interim

Policy Committee and its two subcommittees were military organizations when

first created and civilian government organizations after 1947, but, both before

and after the creation of the AEC, they were predominantly staffed by outside academic researchers. Furthermore, these advisory committees oversaw a government program that was intended to meet the needs of the AEC as well as all the scientists located at numerous institutions throughout the nation who sought to obtain radioisotopes from the AEC.

INSTITUTIONAL STABILITY WITHIN THE AEC: THE MEETING PLACE OF SCIENCE & GOVERNMENT

The interim administration was in place even before the MED announced the new program in June 1946, the radioisotopes that were available at that time, and the administrative procedure for making a request.342 The first request that was filled was for 1 millicurie of carbon-14 which was sent, amongst great fanfare, to a researcher at the Barnard Free Skin and Cancer Hospital in St. Louis,

Missouri on 2 August, 1946.343 Cohn characterized the early days of the distribution program as "hand-to-mouth business."344 His chemistry group worked on producing a limited selection of radioisotopes and anticipated their

142 Ibid., 697-705.. Thompson, "Oak Ridge...Radioisotope Production, n.d.," NARA Atlanta, Box 31, 11. w Oral History of Waldo E. Cohn, 1995, n.p.. 202 ability to produce a wider variety of both radioisotopes and stable isotopes. Aside from production issues, his depiction speaks to the fact that the program was launched quickly after the end of the war and during a period in which the MED was preparing to transfer control to the AEC as of 1 January 1947.

Oak Ridge physicist Eugene Wigner presenting the first shipment of radioisotopes to A. Cowdy, Director of the Barnard Free Skin & Cancer Hospital, August 1946. Photo Credit: Oak Ridge National Laboratory Historical Photo Gallery.

When the program was created there was no certainty that it would continue for any length of time or that it would continue unchanged. The longevity of the program seemed more likely, though, when the AEC formed permanent committees that replaced the interim committees at the start of 1948.

By that time the AEC's Division of Biology and Medicine (DBM) had been established and the General Manager of the AEC decided that while the distribution program would remain in the jurisdiction of the Division of Research, the two divisions should cooperate to oversee the program.345 That the DBM was asked to collaborate on the isotopes distribution program made good sense considering that the distribution program helped facilitate biological and medical research. Beyond that, the goals of the program were intricately linked to those of

,4:> "Meeting of the Advisory Committee on Isotope Distribution, 23-24 March 1950: Background Information," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 31, Organization & Management, 6 - Isotopes Distribution, 1-2. 203

the DBM-organized training and cancer programs—programs which are the subject of the two chapters that follow. As the years passed, both the DBM and the Isotopes Division, along with its advisory committees, would encounter similar policy questions regarding the role of the AEC in biomedical research.

Rather than inheriting a set of well-established policies, the AEC's newly named Advisory Committee on Isotope Distribution Policy and the

Subcommittees on Allocation and on Human Applications continued to create and refine regulations to govern the program. For instance, at a March 1949 meeting the members of the Subcommittee on Human Allocation discussed the issue of conducting tracer studies with normal adults and children. The Subcommittee determined that all studies must be conducted with animals in advance of the

Subcommittee's consideration of human research. Also, they felt that studies conducted with children should receive extra scrutiny and, in general, be discouraged.346 The Subcommittee on Human Application continued to review policy issues related to research with humans throughout the AEC's existence.

The history of these policies is, for the most part, beyond the scope of this study and has been addressed in the Final Report published by the Advisory Committee on Human Radiation Experiment (ACHRE).347 The ACHRE's report examined the system of oversight put in place to regulate risk associated with research.

What it relevant here is that the policy-making process provided ample

w> "Minutes of 13 March 1949 Meeting of Subcommittee on Human Applications," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 31, Organization & Management, 6 - Isotopes Distribution, 3-7. M7 ACHRE, The Final Report, chapter 6. 204 opportunity to blend the practices and cultures of academic and clinical research, with government business.

As already mentioned, the distribution program's advisory committees were comprised of researchers from various institutions throughout the country, including universities, hospitals, research institutions, and national laboratories.

The meetings they held were also attended by AEC administrators outside of the radioisotopes program. For instance, attendance at a March 1950 meeting of the

Advisory Committee on Isotope Distribution Policy included the members of that committee—Gioacchino Failla (chairman), Paul C. Aebersold (secretary), Henry

Borsook, Dr. Austin M. Brues, Dr. D. Harold Copp, Robley D. Evans, Dr. Hymer

L. Friedell, Dr. Albert H. Holland, Joseph W. Kennedy, Dr. Leslie F. Nims, and

P. R. Stout and C. Ernest Birchenall who were alternate members for regular members H. A. Barker and Robert F. Mehl, respectively.348 These individuals represented a range of institutions including: Columbia University, the University of California, California Institute of Technology, Massachusetts Institute of

Technology, Western Reserve University, Washington University, Carnegie

Institute of Technology, Argonne National Laboratory, Brookhaven National

Laboratory, and Oak Ridge Operations. The Committee members were joined by

Henry D. Smyth (AEC Commissioner and author of the "Smyth Report"), Carroll

L. Wilson (AEC General Manager), Kenneth S. Pitzer (Director of the AEC's

Division of Research) and Dr. Shields L. Warren (Director of the AEC's Division

,4!i Regular members were represented by alternates from the same institution. This trend emphasized the importance of institutional representation on this committee. See "United States Atomic Energy Commission, Advisory Committee on Isotope Distribution, Proposed Agenda of the Second Meeting, 23-24 March 1950," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 31, Organization & Management, 6 - Isotopes Distribution, 2. 205

of Biology and Medicine). The gathering of such a group helped facilitate

cooperation amongst different divisions within the AEC and the larger scientific

and medical research community, especially since the issues with which they dealt were not confined to any one group within or outside of the AEC.349

The blending of policies, practices, and cultures amongst researchers and government administrators was not always easy. Many researchers who sat on the advisory committees were resistant to establishing overarching rules to govern the distribution of radioisotopes. They were used to having a great deal of autonomy in their work. But others—especially AEC administrators who were

used to the bureaucratic practices of the MED and/or AEC—were keen to do just that. They were not accustomed to the sorts of practices and relationships within the scientific community whereby researchers might share or obtain such resources as radioisotopes. Such differences were evident at a meeting of the

Subcommittee on Human Applications in December 1949. Dr. Hymer Friedell, the chairman of that Subcommittee, consulted with numerous people who served on similar committees associated with other institutions, and acknowledged that few followed specific criteria for allocation. Despite his findings he opted to compromise and lent his support to the Commission's preference to determine the most specific criteria possible.350 Other members of the Subcommittee voiced opposition to establishing what they feared would be too rigid a set of rules.

Edith Quimby who was attending on behalf of her Columbia University colleague

149 "Advisory Committee... 1950," NARA Atlanta, Box 31,3. 150 "Minutes of 10 December 1949 Meeting of Subcommittee on Human Applications," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 31, Organization & Management, 6 - Isotopes Distribution, 1. 206

Gioacchino Failla was the most outspoken about this issue. In the months following that meeting both she and Failla sent letters to the Isotopes Division to reiterate this point. Quimby wrote that "there are too many individual factors to be considered. A general statement of policy is useful, but that I believe we already have."1 S 1 Another subcommittee member and medical advisor at Oak

Ridge, Dr. Albert H. Holland, sent a similar letter stating his agreement with

Quimby. " The Isotopes Division provided Quimby and all of the members of the Subcommittee an immediate response implying that the Division would try to avoid excessive standardization.353 The Division did, however, publish its first set of rules in 1951, which over time were refined and made increasingly specific.354

BUILDING MOMENTUM: PROGRAM EXPANSION & WIDER USE OF RADIOISOTOPES

The gradual creation and revision of rules to regulate the distribution of radioisotopes and stable isotopes proceeded alongside an expansion of the distribution program that was also gradual, but deliberate. One means of expansion was to extend the distribution program beyond American borders.

Following considerable controversy both within and outside of the United States, the AEC began distributing radioisotopes to researchers abroad after September

351 "Letter to Dr. Allan Lough from Edith H. Quimby, 20 January 1950," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 30, Organization & Management, 6 - Human Application, 1. "Letter to Dr. Allan Lough from Albert H. Holland, 20 February 1950," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 30, Organization & Management, 6 - Human Application, 1. 1x1 "Letter to the members of the Subcommittee on Human Application from S. Allan Lough, 18 February 1950," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 30, Organization & Management, 6 - Human Application, 1-2. The Division's attempt to accommodate Quimby and her colleagues was also evident in the proposed agenda for a meeting of the Advisory Committee on Isotope Distribution Policy held in March 1950. See "Advisory Committee... 1950," NARA Atlanta, Box 31, 6-7. •54 "Isotope Distribution, 1951NARA College Park, Box 28, 6. 207

1947.155 As Angela Creager shows in her examination of the role of radioisotopes

in postwar science, foreign distribution caused conflict between the politics of

science and politics of the Cold War.356 From the start of the distribution program

in 1946, scientists were outspoken advocates for foreign distribution as a means

of advancing scientific research because many believed that science was an

i c 7 institution that should have no geographical boundaries. Congressmen and the

Federal Bureau of Investigation, on the other hand, raised concerns about AEC

security practices. As a result, it took the better part of 1947 for the AEC's

Commissioners to approve foreign distribution.358 Political concerns slowed the

decision, but did not prevent wider distribution of radioisotopes.

Growth of the distribution program, both in the United States and abroad,

was fueled by the faith that researchers had in radioisotopes research—faith that

the AEC helped to publicize. For instance, in March 1948 Oak Ridge medical advisor, Dr. Albert H. Holland, spoke at a pharmaceutical convention to promote

the use of radioisotopes in biomedical research and in therapy. He announced

that, at that time, two radioisotopes—radiophosphorous and radioiodine—were

known to be therapeutic. He confidently projected that this number would grow and that the pharmaceutical industry could play a role in this business.359

The AEC hardly needed to seek out media attention to publicize its biomedical research. Newspapers and magazines devoted much attention to

"Appendix I-A from Isotopes, 1949," NARA College Park, Box 45, 39. ,5h Creager, "Export of'American' Radioisotopes," 367-88. 157 Ibidr. 372-74. Debate also ensued regarding the United State's image abroad, isolationist policies, the Marshall Program and other issues. See Ibid:. 373. 15'' "Radioisotopes in Medicine, 1948," NARA Atlanta, RG 326, New York Operations Office Research & Medicine Division Correspondence 1945-1952, Box 28, Information & Publications, 8 and 14. 208 nucicar issues after the war, conveying similar optimism to the public. During the summer of 1948, for example, the Washington Evening Star published a scries of articles describing the use of various radioisotopes in biological and medical research. These articles were overwhelmingly positive, expressing optimism in such headlines as: "Probing Nature's Secrets," "Key to Life's Mysteries," "Iron's

Atomic Searchlight," and "Man's Dependence on Iodine."360 These optimistic messages, just two examples among many, show that between the AEC's and media's efforts, both targeted audiences like the pharmaceutical industry and the general public were kept abreast of scientific and medical advances made possible by radioisotopes.361

Within the AEC bureaucracy optimism abounded. From the start of the distribution program, the AEC and especially the Isotopes Division sought to establish wider use of radioisotopes within the scientific and medical communities. Shortly after the AEC replaced the isotopes program's interim advisory committees with permanent ones at the start of 1948, the AEC's General

Manager Carroll Wilson asked Aebersold's Division to review the history of the program to that point and determine the conditions that limited wider distribution.362 The Isotopes Division responded with a report in the summer of

1948 that highlighted a few factors they considered to be obstacles to the future expansion of production and use of radioisotopes. First, insufficient manpower

"Progress in Isotope Research, 1948," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 46, Reports & Data Isotopes Program, 1-12. 1M See also, N.A., "Isotopes," 74. "AEC Memorandum for Information: Study of Wider Use of Isotopes, 29 July 1948," NARA College Park, RG 326, General Correspondence 1946-1951, Box 45, Study of Wider Use of Isotopes, 1. 209

was a significant problem that limited both production and use. Second, in terms

of radioisotope production, space and facilities were inadequate to accommodate

growth. And finally, to increase the use of radioisotopes amongst researchers or

within institutions new to research with radioisotopes, laboratories would have to

be converted. The cost of doing so was a huge obstacle.363

These three issues were not specific to just the isotopes distribution

program or even the AEC's other biomedical research. Rather, they were

problems that affected the overall state of biomedical radiation research. The

efforts taken to address these problems within the isotopes program would have

ripple effects elsewhere, especially in universities and in industry. As we will see

in the next chapter, the AEC created training programs both in national

laboratories and in universities to address the manpower issue. The AEC also encouraged industry to participate in various aspects of the radioisotopes business.364 For instance, the Isotopes Division worked with industry to facilitate the commercial production of radioisotope-labeled compounds. The first commercially produced compounds were sent out to researchers in July of 1948.

Starting during the winter of 1949 the AEC supplemented commercial efforts to

label compounds with radioisotopes to provide researchers access to radioisotope- labeled compounds that were not commercially available.365

"Wider Use oflsotopes, 1948," NARA College Park, Box 45, 3-4. ,w The AEC's efforts to expand industry participation in radioisotopes production and distribution increased as the program proceeded, especially throughout the 1950s. See "Isotope Distribution, 1951," NARA College Park, Box 28, 1 -57. "'5 Historians Timothy Lenoir and Marguerite Hays provide further discussion on the role of private industry in radioisotopes production and distribution. As they explain, the cost of reactors necessitated government ownership of them. However, the AEC granted contracts to commercial firms that acted as intermediate producers. See Lenoir and Hays, "The Manhattan Project for Biomedicine," in Controlling Our Destinies, 43-50; and "Appendix I-A from Isotopes - A Three 210

Approximately one year after Aebersold's division produced this initial

report, the AEC compiled data to evaluate the rate of distribution over the three years the program had been operating. The number of groups using radioisotopes in the United States increased more than 35% from 1948 to 1949 and those using radioisotopes in research conducted more studies than they had previously.

Requests for radioisotopes came from a greater variety of institutions and departments.366 There was also an increase in the AEC's distribution of stable isotopes. However, these shipments were markedly fewer than the shipments of their radioactive counterparts.

Not all of the research conducted with AEC radioisotopes was biomedical research, but the distribution program certainly helped to expand the field. From

1948 to 1949 approximately 700 papers reported on biomedical research conducted with AEC radioisotopes. These constituted approximately 38% of all publications that stemmed from research made possible by the distribution program.' Many of the biomedical studies published used radioisotopes as tracers. Tracer studies involved the use of radioisotopes to label and trace sugars, proteins, vitamins, phospholipids, hormones, nucleic acids, antibodies, dyes, drugs, organic acids, and blood cells. Papers published on diagnostic and therapeutic uses of radioisotopes were fewer because, as the AEC reported in

1949, "to date only a few applications have been found for radioisotopes as tools

Year Summary of Distribution, 16 September 1949," NARA College Park, RG 326, General Correspondence 1946-1951, Box 45, Isotopes - Three Year Summary of Distribution, 40-41. "Part 111: Increased Isotope Utilization from Isotopes A Three Year Summary of Distribution, 16 September 1949," NARA College Park, RG 326, General Correspondence 1946-1951, Box 45, Isotopes - Three Year Summary of Distribution, 19. 367 "Part III: Isotopes, 1949," NARA College Park, Box 45, 19. 211

->/; o for diagnosis and therapy." The AEC reckoned, though, that, "an increase may

be expected....Many new uses are certain to be found in diagnosis where smaller

quantities of radioactivity may be used."369 Researchers were especially

optimistic about the use of radioisotopes to diagnose and treat cancer, which is

why the AEC subsidized the cost of providing radioisotopes for cancer research starting in April 1948. A year later all Commission-produced radioisotopes were

made available to cancer researchers free of production costs. 370

TABLE 1: RADIOISOTOPE DISTRIBUTION TO NON-AEC INSTITUTIONS371

Period Number of Monthly Average % Increase Shipments of Shipments Over Previous Period

Aug. to Dec. 246 49 — 1946 1947 1652 138 200 1948 3009 250 80 1949 4715 400 60 Jan. 1950 529 529 30 Total Shipments, Aug. 10,141 1946 to Jan. 1950

,6S "Part III: Isotopes, 1949," NARA College Park, Box 45, 20. 1<,') "Part III: Isotopes, 1949," NARA College Park, Box 45, 20. ,7° "Appendix I-A from Isotopes, 1949," NARA College Park, Box 45, 40. These figures do not include shipments sent to AEC facilities. The total number of shipments sent through the end of January 1950, including those to AEC facilities, was 11,985. See Thompson, "Oak Ridge...Radioisotope Production, n.d.," NARA Atlanta, Box 31, 32. 212

TABLE 2: DIAGNOSTIC & THERAPEUTIC USES OF RADIOISOTOPES, 1949372

Radioisotope Diagnostic Application Therapeutic Application

Sodium 24 Cardiac testing including differentiation between normal and restricted blood flow, determination of heart's pumping ability, and detection of congestive heart failure. Phosphorous 32 Determination of brain Treatment of polycythemia tumor mass. vera and chronic leukemia. Iodine 131 Determination of Treatment of hyperthyroidism, thyroid hyperthyroidism, thyroid cancer and metastases. cancer and metastases. Iodine 133 (in Detection of certain brain diiodofluorescein) tumors. Strontium 89 & Beta-ray source for 90 teletherapy treatment of surface lesions. Cobalt 60 Radiation source for interstitial therapy for accessible tumors or for teletherapy for inaccessible tumors. Gold 198 Treatment of chronic leukemia and some tumors of the lymphoid system.

CONCLUSION

From the time that they were discovered in the mid-1930s, radioisotopes

were coveted due to the MED's wartime development of nuclear reactors. Thus,

at the end of the war numerous researchers sought to create and then expand an

isotopes distribution program within the MED and its successor, the AEC. Dr,

Joseph G. Hamilton at Berkeley, John R. Dunning at Columbia University, and

Frank B. Jewett at the National Academy of Sciences/National Research Council,

,7: "Part III: Isotopes, 1949," NARA College Park, Box 45, 20. 213

for instance, seized what they saw as an opportunity in the immediate postwar

climate to advocate for this program. They saw great potential in using

radioisotopes in research, even though, at that point, they could not identify many

specific applications.

By taking advantage of their authority and connections within academia

and especially within the MED and AEC, these scientists and others helped

establish a program that opened up avenues for research which had looked very

promising since the 1930s, but were limited to only a small number of

researchers. Once the program was established within the MED and then the

AEC, Paul C. Aebersold, his Isotopes Division, Waldo E. Cohn on a temporary

basis, and his Chemistry Division continued to craft a program that would support

not only the AEC's research, but also provide radioisotopes to researchers nation-

and world-wide.

The distribution program they organized did more than distribute

radioisotopes to researchers. It not only allowed, it required researchers and the government to collaborate in an effort to determine what roles each group might

play in the postwar political economy of research. The system of advisory committees was integral to the alignment of interests between the AEC and larger

research community. Indeed, the program's Advisory Committee on Isotope

Distribution Policy and Subcommittees on Allocation and Human Application drew some of the nation's leading radiation experts to participate in policy­ making. They contributed to the establishment and expansion of infrastructure that linked research communities with the government. In the process, the lines that divided the government from the research enterprise were increasingly blurred. These parties shared responsibility in applying science to the problems of the nation, regardless of whether they were national security or biomedical concerns. 215

CHAPTER 5

THE ABC'S OF THE AEC'S BIOMEDICAL RESEARCH: FELLOWSHIPS AND EDUCATIONAL INITIATIVES

The success of the radioisotopes distribution program examined in the

previous chapter depended on there being sufficient researchers, physicians, and

technicians trained to work with radioisotopes. Thus, this chapter focuses on the

Atomic Energy Commission's (AEC) education and training programs which were crcated in an effort to increase the nation's scientists and physicians trained

in all aspects of radiation research. In terms of biomedical sciences, the programs included fellowships to attract talented young individuals to fields related to atomic energy. For biomedical scientists and physicians employed at various institutions and for military personnel, the AEC offered short-term courses related to radiation, its uses and hazards. By funding these and other programs as well as associated physical infrastructure, the AEC aimed to provide access to the technologies and expertise located within the national laboratories. The AEC also sought to help independent institutions acquire similar technologies and develop comparable expertise as those located within its own facilities.

The AEC did not act alone in designing, implementing, and administering the education and training programs established within AEC facilities or those that the AEC funded elsewhere. Indeed, scientists and physicians working within academia and private organizations willingly collaborated with the AEC. For them, the AEC existed as a new source of patronage that facilitated the expansion of and, in some cases, institutionalization of hybrid programs that stemmed from pre-war interdisciplinary collaboration. Like the AEC's radioisotopes distribution 216 program, radiation-related education and training acted as a link between acadcmia and the state, and amongst science, medicine, and national security.

This chapter examines the development of the AEC's biomedical education and training programs and infrastructure which, 1 argue, were a collection of piecemeal initiatives established gradually at the encouragement of those both within and without the AEC. As we will see, partnerships formed between the AEC and the Oak Ridge Institute for Nuclear Studies (ORINS), the

University of Rochester, and the National Research Council (NRC) of the

National Academy of Sciences (NAS) resulted in educational programs such as the Rochester Atomic Energy Project (RAEP) and the AEC's fellowship program in biology and medicine. These and other programs were the products of individuals and institutions defining common goals and negotiating partnerships through which to meet those goals. The process by which these programs and infrastructure was established involved various individuals and parties who were not necessarily united by an institutional agenda, but rather by a commitment to expand the nation's expertise in the uses of radiation.

The development of the AEC's education and training programs is best understood as a two-part process. During the first phase, proposals were drafted and existing or new institutions presented themselves as able and willing partners to help the AEC achieve its educational goals. This phase began in 1946 as the

Manhattan Engineer District (MED) prepared to hand over control of the

Manhattan Project facilities to the AEC and came to an end after the creation of the AEC's Advisory Committee on Biology and Medicine (ACBM) and Division 217 of Biology and Medicine (DBM) in the fall of 1947. At that point, the AEC, with the help of others, began to make concrete steps towards bolstering the nation's radiation-related education and training opportunities. It was during this second phase, which lasted for approximately two years, that the interests of academic institutions, various other private institutions, the AEC, the government more broadly, and the military were joined such that their partnerships became a defining feature of Cold War culture.

While the various parties that collaborated in an effort to expand radiation- related biomedical education and training shared and actively sought to pursue common interests and goals, conflicts did emerge. The evolution of the AEC's educational programs and infrastructure illustrates tensions that resulted from the effort to merge science, medicine, the state, and national security in the postwar period. With regards to education and training, the most prominent tensions stemmed from the role of government, secrecy, and anti-communism in science and education.373 This chapter shows that the creation and ongoing reconfiguration of educational programs and infrastructure reflected efforts to balance shared interests and, at times, incompatibilities between science and the state. Such negotiations were not only a result of the Cold War, but were, in fact, part of what constituted the Cold War. As scientists and administrators from the

AEC and elsewhere collaborated to address what they defined as both problems and opportunities associated with biomedical education and training, they defined

Cold War policies and programs related to research and development, as well as

™ For a comprehensive analysis of the clash between Cold War secrecy and anti-communism, on the one hand, and science, on the other, see Wang, American Science in an Age of Anxiety. 218

national security. By doing so, they helped set the tone of blended civilian and

military spheres that continued to characterize American society, including

biomedical sciences, throughout the Cold War.

"THE CRITICAL SHORTAGE"

In 1949 AEC Chairman David E. Lilienthal recalled that,

Ever since the war, it has been generally recognized, and especially by such distinguished gentlemen as Dr. Conant of Harvard and Dr. Bush and others, that to carry forward the Nation's scientific development, means must be found to insure our continued scientific advance and in particular to overcome the critical shortage of trained scientists.374

Lilienthal spoke these words during his testimony before the congressional Joint

Committee on Atomic Energy (JCAE) which, in May 1949, held hearings on the

AEC's fellowship program. Science was vital to the well-being of the nation,

Lilienthal contended, and one of the most important components for scientific advancement was to increase the number of scientists. This was a problem that, from the end of the war, concerned not only the AEC, but the leaders of the scientific community.

Lilienthal invoked the authority of James B. Conant and Vannevar Bush who, due to their many academic and administrative achievements, were key and extremely influential figures in American science. Conant, for instance, was

President of Harvard University and had been chairman of the wartime National

Defense Research Committee. Bush was the president of the Carnegie Institution of Washington and, as we saw in Chapter Two, had been director of the Office of

,%1 "Atomic Energy Commission Fellowship Program, Hearings before the Joint Committee on Atomic Energy, 16-18, 23 May 1949," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 31,3. 219

Scientific Research and Development during the war.375 Both men led the ranks of scientists who advocated for government support of scientific research and education in the postwar years. Bush, in particular, helped establish a framework for what would eventually be the National Science Foundation (NSF)—the government agency created in 1950 to help achieve the postwar agenda of those scientists and policy-makers who prioritized scientific development.376 He also was the first chairman of the Joint Research and Development Board—renamed the Research and Development Board in 1947—which was created following the war to coordinate research for both the Army and Navy. But while the nation's scientific leaders were in agreement on the overall need to advance science, how this was to be achieved, how to increase the number of scientists, and what role the AEC should play in accomplishing these goals, were not at all clear.

As was the case with the creation of AEC's Division of Biology and

Medicine (DBM) and Advisory Committee on Biology and Medicine (ACBM), scientists and administrators within the MED initiated discussions regarding its successor's role in education and training that would shape the programs later established. Those within the MED were not, however, the only ones involved in the early phase of planning. Individuals within the MED were joined by scientists from private institutions in an effort to ensure that the MED's successor would

375 For an in-depth discussion of Conant's and Bush's careers, see James G. Hershbcrg, James B. Conant: Harvard to Hiroshima and the Making of the Nuclear Age (Stanford, CA: Stanford University Press, 1993); Zachery, Endless Frontier: Vannevar Bush. 376 For a history of the debate regarding postwar government support of science and the creation of the National Science Foundation, see Kevles, "The National Science Foundation," 4-26; John T. Wilson, Academic Science, Higher Education, and the Federal Government, 1950-1983 (Chicago, IL: University of Chicago Press, 1983). 220

play a role in enhancing the nation's scientific education and training, especially

that related to radiation research and clinical applications of radiation.

The broad interest in biomedical education and training was evident at a

meeting held at Oak Ridge National Laboratory (ORNL)377 in October 1946

regarding the need for training in radiobiology, a fairly new field that focused on the study of the biological effects of radiation.378 Fifty to seventy individuals attended the meeting, representing institutions including, but not limited to, the

Argonne, Brookhaven, Oak Ridge, and Los Alamos National Laboratories; United

States Public Health Service; National Cancer Institute; University of Rochester;

Columbia University; ORINS; Oak Ridge Hospital; and both the Army and the

Navy. All were in agreement regarding the need for special training facilities for radiobiology given their assessment that universities did not have sufficient programs and curricula in place to meet the AEC's radiobiology needs, specifically the Commission's radiation safety responsibilities. They generally believed that, while the nation's biomedical radiation researchers had been able to adapt their skills to meet the needs of radiation health and safety during the war, the creation of atomic weapons and plentiful supply of atomic energy necessitated an expansion of biomedical radiation research and those trained to pursue such research.379

177 At the time ORNL was still known as Clinton Laboratories, its original name. Clinton was not renamed until the start of 1948, but to avoid confusion, I will use one name—ORNL—throughout this chapter. ,7li Radiobiology began to emerge as a field of research in the 1920s, but was, as of the early postwar period, still being defined. Creager and Santesmases, "Radiobiology in the Atomic Age," 637. m Eugene P. Wignerand Paul S. Henshaw, "Report on 15 October 1946 Meeting, 23 October 1946," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 39, 1. 221

Aside from the need for biomedical scientists to attend to health and safety problems associated with the nuclear weapons industry, increased availability of radioisotopes created many promising avenues for biomedical research and medical applications, if only there were sufficient researchers and physicians to investigate their uses. Indeed, the radioisotopes distribution program had already begun at the time the meeting was held, thus, the opportunities for biomedical researchers to work with radioisotopes were greater than prior to the war when a limited number of researchers had access to radioisotopes.

Historians Angela Creager and Maria Santesmases have argued that the focus of radiobiology shifted at the start of the nuclear age. Whereas it was once a field primarily concerned with the effects of X-rays on biological processes, the creation of nuclear reactors which allowed for large-scale production of radioisotopes resulted in a more varied selection of radiation sources, the biological effects of which could and, according to those engaged in planning for biomedical education and training, should be studied. Creager and

Santesmases are correct in identifying that the focus of radiobiology shifted in that it expanded and gained new meaning relative to the postwar social and political environment in which it developed. But their argument should be expanded. In particular, the shifting scope of research in the field of radiobiology and closely related biomedical fields was very much a part of an ongoing effort to redefine the relationships amongst science, the state, and national defense, and between private institutions and the government. Researchers, as we will see,

wo Creager and Santesmases, "Radiobiology in the Atomic Age." 222 were proactive and entrepreneurial in creating or enhancing existing relationships with the MED and AEC.

The main goal for those who gathered at ORNL to discuss the need for radiobiology training was to produce a plan for a training program.381 Many of the attendees recognized that such a training program would benefit their own institutions, but conceded that they were not equipped to house such a program.

Some researchers felt their own institutions were already overextended and lacked sufficient researchers to train others. No concrete plan emerged, but the participants agreed that establishing a training program at Oak Ridge seemed the best bet. They discussed the involvement of other institutions, such as the Oak

Ridge Institute of Nuclear Studies (ORINS) and the University of Tennessee, as possible collaborators.382

The ideas suggested and discussed at this meeting helped define options for the biomedical educational agenda the AEC would soon inherit. Perhaps most importantly, the discussion closed few doors in terms of who and what institutions might be involved in the future of radiation-related education and training in the biomedical sciences. While the meeting concluded with a tentative plan for a radiobiology training program at Oak Ridge, those in attendance also agreed that the other institutions represented could and should draft proposals for a training program to be submitted to the MED's Medical Advisory Committee for consideration.

"Wigner and Henshaw Report, 1946," NARA Atlanta, Box 39, 1. "Wigner and Henshaw Report, 1946," NARA Atlanta, Box 39, 2. 223

During the final months of 1946 various proposals were drafted for

training programs in radiobiology and/or the closely related field of health

physics, most of which were located at ORNL."^83 However, James H. Lum,

Wigner's Co-Director of Research and Development at ORNL, considered most

proposals "impractical" given the "heavy and varied responsibilities" at the

laboratory.384 The one proposal that seemed feasible to him in the short-term was

from Karl Z. Morgan, the director of the Health Physics Division at Oak Ridge.

Morgan was among the scientists recruited to the Health Division at the

Metallurgical Laboratory at the start of the Manhattan Project. From there, he

relocated to Oak Ridge in 1943 as part of the new laboratory's founding group of

health physicists. Morgan was instrumental not only in laying the groundwork for an expansion of the health physics program at ORNL, but also for establishing the professional status of health physics as an independent field.

What Morgan proposed in November 1946, approximately one month after the meeting on radiobiology, was that ORNL offer a seminar in health physics during the summer of 1947. He also fully supported the idea of a broader training program as suggested by others, especially since the military and industry would, in the atomic age, require well-trained health physicists. However,

3X3 Joe W. Howland (Chief of Medical Research for the MED as of 1946), Paul Henshaw (biologist with the Metallurgical Laboratory's Health Division), Alexander Hollaender (Director of the Biology Division at Oak Ridge), and Eugene Wigner (Director of Research and Development at Oak Ridge) were among those who drafted proposals. For Henshaw's and Howland's proposals, see by Paul S. Henshaw, "Proposed Training Program in Radiobiology, 8 October 1946," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 42, 1-2; and Joe W. Howland, "Proposed Training Program for Physicians, Health Physicists, Biologist and Medical and/or Health Physics Aides, October 1946," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 42, 1-8. 3X4 "Letter from James H. Lum to Mr. A. V. Peterson, 4 December 1946," NARA Atlanta, RG 326, New York Operations Office Research & Medicine Division Correspondence 1945-1952, Box 39, I. 224

Morgan did not sec a full training program as practical given the shortage of trained persons at Oak Ridge and elsewhere. He considered himself to be one of only four individuals well enough trained to be primary instructors and counted a handful of others as qualified assistant instructors." Morgan encouraged health physicists to make it known that health physics was one of the most urgent aspects of future biomedical radiation research and work. To ensure that this message reached the Commissioners, he directed his proposal to Lilienthal as well as Wigner, Lum, and a few health physics colleagues.386

Like Morgan, the Chief of the MED's Medical Office, Dr. Stafford L.

Warren, devised a plan to make immediate use of the expertise concentrated at another location. In October 1946, just ten days after the radiobiology meeting,

Warren contacted Colonel Kenneth D. Nichols who was in charge of the MED headquarters at Oak Ridge asking for permission to train physicians and researchers through the MED's Rochester Atomic Energy Project.387 He explained that since the Manhattan Project was made public, Rochester area doctors and researchers had been contacting him to request training for use of radioisotopes in their work. Warren suggested that the most efficient way to provide such training was to send interested individuals to the University of

Rochester where the MED funded a research program focused on biomedical radiation research.

w Karl Z. Morgan, "Proposal for a Health Physics Training Program and the Creation of a Health Physics Advisory Committee, 26 November 1946," NARA Atlanta, RG 326, New York Operations Office Research & Medicine Division Correspondence 1945-1952, Box 39, 1. 3W' Morgan, "Proposal for Health Physics, 1946," NARA Atlanta, Box 39, 2. 1X7 "Memo to Colonel K. D. Nichols from Stafford L. Warren, 25 October 1946," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945- 1952. Box 42, 1. 225

Morgan and Warren were engaged in a process of discipline-building—a

proccss that took place within academic institutions, but also within the MED and,

later, the AEC bureaucracy. Both of their proposals illustrate that during the

transitional period in which the MED prepared to hand over control of MED operations to the AEC, those working within the MED administration and

facilities were eager to not only continue the biomedical work of the MED, but also make use of the facilities and expertise within the MED system to train others. Their proposals suggested small-scale and/or short-term programs, but these helped set the stage for the gradual development and implementation of

larger education and training programs in the years to come. Furthermore, they accommodated requests from individuals and groups in medicine, academia, and industry who were not part of the MED's operations that wanted to participate in the expansion of biomedical research, but needed training to do so.

The demand for MED training initiatives reflected the reality that in the postwar years, the MED and later the AEC faced the expectations of researchers, or at least their hopes, that the Manhattan Project would translate into opportunities beyond those currently available to them at their own institutions. It was not unprecedented that researchers looked to their colleagues elsewhere to gain experience in new areas of research or to gain access to new technologies.

This was certainly the case at Ernest Lawrence's Rad Lab at the University of

California, Berkeley, throughout the 1930s. Following the war, this trend continued and mostly revolved around the MED and, after the MED was dissolved in 1947, the AEC. 226

"OAK RIDGE INSTITUTE OF NUCLEAR STUDIES - A NEW PATTERN IN GRADUATE EDUCATION"388

The Oak Ridge Institute of Nuclear Studies (ORINS) was, perhaps, the

most obvious example of individuals and institutions seeking to take full

advantage of what they believed to be a great opportunity stemming from the

Manhattan Project. Whereas many of the institutions that sought to associate with

the AEC had some connection to MED operations, the ORINS was an entirely

new organization created with the express purpose of helping to organize AEC

research and development. The ORINS started to take form almost immediately

after the end of the war. In June 1946 an article published in the journal Science

announced the new organization and explained that it would "provide the formal channels for cooperative research between the universities and governmental

research and producing agencies associated with the atomic energy project at Oak

Ridge."389 The Science article underlines that the federal government, not private foundations, was the primary patron with which universities and academic

researchers would work in the postwar period. While this statement speaks to the changes taking place in scientific patronage, it also highlights the ongoing prioritization of cooperative research which sustained the policies and programs of private foundations throughout the 1930s.

1SX William G. Pollard, "The ORINS - A New Pattern in Graduate Education, Address delivered at the Academy Dinner, Meeting of the American Association for the Advancement of Science, New York City, 30 December 1949," NARA Atlanta, RG 326, New York Operations Office Research & Medicine Division Correspondence 1945-1952, Box 33, 1. ,s" The Executive Committee ORINS, "A Nuclear Research Institute at Oak Ridge," Science 103, no. 2685 (1946): 706. As examined in Chapter Three, the term "cooperative research" reflected the prominence in the 1930s of research projects that were interdisciplinary and funded by private foundations. See page 26, in particular, footnote 40. 227

The formation of the ORINS occurred during the same period in which

there were ongoing discussions amongst those within the MED regarding the

future of biomedical radiation education and training.390 The ORINS was

represented at the ORNL meeting about radiobiology and, as mentioned above, considered as a possible partner to collaborate with ORNL on the development of a radiobiology training program at the Oak Ridge facilities. The proposed partnership was exactly the sort of endeavor that the ORINS hoped to be a part of.

FIGURE 2: ORINS ORGANIZATIONAL CHART, SEPTEMBER 1949391

f»* MM* dwmii KCUU rrMMM rawm

MMCf«TR mt*i t*im* >1 UIVt H «.«NM

•*»wo em mr*jn OMtAYurrt rnaoMi wnm A, 6.

UMVIiMITT Mt«r«OM»

invtoi flKMM

•a».« **•

Oak Ridge, of course, was located in Tennessee and the ORINS was a consortium between southern universities. At the outset, 14 universities partnered

Pollard, "The ORINS - A New Pattern, 1949," NARA Atlanta, Box 33,1. ,<)l "Oak Ridge Institute of Nuclear Studies Corporation Minute Book, 1 September 1949," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 34, 3. to form ORINS. By 1949 the number of universities had risen to 24 and by 1955

it had again risen to 32. The universities involved in ORINS aimed to "[develop]

39"> ways of fulfilling the high promise inherent in the ORNL." " The universities

that founded ORINS and those that joined the organization in the years that

followed sought to coordinate collaboration between the educational institutions of the region and the research facilities at Oak Ridge. Primarily, these

universities hoped to enhance education, but they also considered partnerships with the ORNL as a means through which to bolster industrial and economic growth in the region. With this broad agenda the new organization's founding universities assembled an Executive Committee that included representatives from the academic community, ORNL and the MED, as well as from local

industries. For instance, the Committee's academic and industrial members included representatives from the University of Tennessee, Duke and Vanderbilt

Universities, Carbide & Carbon Chemicals Corporation, Tennessee Eastman

-1QT Corporation, and Monsanto Chemical Company.

The Executive Committee defined the goal of facilitating universities' access to ORNL such that students and faculty in the region might use and consider ORNL to be part of their own university's resources. However, the new organization began with no precise plan for achieving this goal.394 The Executive

Committee created various divisions, including the Medical, University Relations, and Special Training Divisions which were, as we will see below, the most

V): Pollard, "The ORINS - A New Pattern, 1949," NARA Atlanta, Box 33, 2-3; and "32 Students from 21 Countries to Receive Radioisotope Training, n.d.," NARA College Park, RG 326, General Correspondence 1951-1958, Box 30, 1. w ORINS, "A Nuclear Research Institute at Oak Ridge," 706. "w Pollard, "The ORINS - A New Pattern, 1949," NARA Atlanta, Box 33, 5. 229

instrumental divisions in developing and administering education and training

programs in the biomedical sciences. The bureaucratic structure, policies,

programs, and infrastructure were developed somewhat in service to the MED's

and AEC's educational agenda, but also helped create this agenda. Indeed, the

formation of the ORINS during the year following the war resulted from the actions of a large group of scientists and educators interested in being a part of

whatever education and training initiatives the AEC pursued. They saw the AEC as an avenue through which to obtain resources and, to ensure this, were willing to help the AEC achieve its goals and fulfill its responsibilities. The very existence of the ORINS, with its aims of advancing nuclear studies, helped ensure that there would be AEC education and training initiatives of which to be a part.

According to the Executive Committee of the ORINS, "The national stake in the future of atomic energy is a vital one, and any steps that will assure that active research is not interrupted are of the utmost importance at this time."395

The ORINS was not the only such organization created in association with one of the nation's nuclear facilities. A group of nine northeastern universities formed a similar organization, Associated Universities, Inc. (AUI), in 1946.

These universities came together with the purpose of helping to create a new nuclear laboratory in their region, Brookhaven National Laboratory (BNL), organize opportunities associated with an existing MED laboratory. The creation of AUI and BNL are best examined by historian Allan Needell who explains that, amongst other reasons, various researchers at northeastern universities pushed for

14> ORINS, "A Nuclear Research Institute at Oak Ridge," 706. 230

the creation of the new national laboratory as a regional laboratory that would

rival the Argonne National Laboratory and Berkeley Radiation Laboratory.396

Like the ORINS, the AUI defined for itself a mandate of facilitating educational programs between the participating universities and the laboratory.

Aside from that responsibility, the AUI was responsible for managing the operation of BNL. This was a responsibility the ORINS did not share given that

ORNL was managed by an industrial contractor. Despite that difference, the creation of the ORINS and AUI was driven by the common goal of strengthening the scientific infrastructure in these two regions and doing so in connection with the new government research and development agencies and facilities. The importance of these new organizations became increasingly evident in the first few years that followed their creation. By the late 1940s, both had assumed responsibility for managing various AEC-funded education and training initiatives, some of which are examined in greater detail below.

1947: THE START OF THE AEC ERA & A YEAR OF PLANNING

The AEC assumed control over the nation's atomic enterprise on 1

January 1947 and inherited the problem of education and training from the MED.

With proposals from individuals such as Morgan and Warren, and new organizations such as the ORINS eager to help establish education and training opportunities, scientists expected that the AEC would define policies regarding its role in radiation-related education, as well as initiate programs and build corresponding infrastructure. Biomedical education and training became part of

Allan A. Necdcll, "Nuclear Reactors and the Founding of Brookhaven National Laboratory," Historical Studies iti the Physical Sciences 14, no. 1 (1983): 96 and 104. 231

the broader debate amongst scientists and administrators regarding the

administrative organization of biology and medicine within the AEC. The AEC's

Interim Medical Committee and the Medical Board of Review formed by the

National Research Council (NRC) of the National Academies of Science (NAS)

considered the lack of individuals trained to control radiation hazards and apply

radiation in biology medicine to be the greatest challenge the AEC faced as it

prepared to establish a biomedical program.397 For instance, the Medical Board of

Review outlined the AEC's educational responsibilities as falling into two

streams: fellowships for pre- and post-doctoral studies and technical training in

health physics and related research. In both regards, the Board recommended that

the AEC work in concert with the NAS and, if it came into being, the proposed

National Science Foundation (NSF).398

In the months between the Interim Committee's January 1947 meetings and the Board of Review's June meetings, the Army reinforced this message and defined another educational responsibility. Colonel R. E. Duke who was Chief of the Army's Medical Corps Education and Training Program asked the AEC to take responsibility for providing training to officers in all medical and biological aspects of atomic energy. The general consensus within the Medical Corps, he wrote, was that "The importance of obtaining a large body of medical officers

"7 Stafford L. Warren, "Report of the 23-24 January 1947 Interim Medical Committee," NARA College Park, RG 326, Genera! Correspondence 1946-1951, Box 32, 21; "Medical Board of Review Report, 20 June 1947," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 39, Medical Policy, 9. , w "Medical Board of Review Report, 1947," NARA Atlanta, Box 39, 6. 232 who arc well qualified in this field cannot be overemphasized. Both the military and civilian population must be considered by the Army Medical Department."399

The Army's interest in biomedical radiation training stemmed from the need for military preparedness in the atomic age and, as such, representatives from the Army and the Navy had, since the end of the war, played a role in advocating for AEC training programs. Both armed services sought to prepare officers for the radiation hazards and associated medical problems that could arise during weapons testing and atomic warfare. The Army and the Navy sent representatives to the October 1946 meeting on radiobiology training at Oak

Ridge and Duke's letter, written in April 1947, was a more concrete contribution to the ongoing issue. In his letter he outlined a specific request from the Surgeon

General of the Army that the AEC establish an on-the-job training program to be held at AEC research facilities that could accommodate ten Army personnel/year.

The Interim Medical Committee's and the Medical Board of Review's reports, along with Duke's letter on behalf of the Army Medical Corps, reinforced the idea that the AEC must play a role in radiation-related education and training.

The AEC was better equipped to initiate the proposed fellowship program and other programs, some specifically aimed toward technical training and military personnel once the DBM and ACBM were created in the Fall of 1947; the AEC was able to delegate responsibility for biomedical education and training programs to this new division and advisory committee. That the DBM and

ACBM helped advance the process of establishing biomedical education and

W) "Letter from R. E. Duke to Director of Researcher and Development, 4 April 1947," NARA Atlanta, RG 326, New York Operations Office Research & Medicine Division Correspondence 1945-1952, Box 42, 1. 233

training from a phase dominated by planning and proposals to one in which concrete decisions were made and actions were taken is evident in the history of the Rochester Atomic Energy Project. Like the OR1NS, the University of

Rochester became an important partner to the AEC. The history of the Rochester

Atomic Energy Project in the postwar years illustrates well the actions of those outside of the Commission to shape AEC policies and programs. It provides a university-oriented look at how federal funding remade scientific and medical research following the war. This history also shows that, like the radioisotopes distribution program, the AEC's educational initiatives bound the interests of science and medicine to those of the state.

THE ROCHESTER ATOMIC ENERGY PROJECT & THE TRANSITION FROM WAR TO PEACE

As we saw in Chapter Two, the University of Rochester was the site of a

Manhattan Project biomedical research program. During the war the university received a MED contract to pursue research on the toxicology of radioactive materials, especially uranium. It was also the university at which the Chief of the

MED's Medical Office, Dr. Stafford L. Warren, was employed before he was recruited to the MED. Rochester's wartime research program was, as one faculty member recalled, a bit different from other MED laboratories in that it focused entirely on biomedical research.400 This comment reflects the fact that the

Rochester Atomic Energy Project was not overshadowed by a research and development project more closely related to the creation of atomic bombs. The

Metallurgical Laboratory's Health Division at the University of Chicago, while

40(1 Stannard, "Sketch of the Rochester Project," 7. 234 the hub of the Manhattan Project's health and safety activities, was just one part of a larger research and development project aimed at building a nuclear reactor.

Similarly, the MED funded biomedical radiation research at Ernest Lawrence's

Radiation Laboratory at the University of California, Berkeley, but the larger focus at that laboratory was the electromagnetic separation project.

Considering the wartime history of biomedical research at the University of Rochester, as well as Stafford Warren's connection to the institution, it is of little surprise that Warren had proposed the idea of providing interested

Rochester-area physicians and biomedical researchers access to radioisotopes training at that university. Warren outlined a relatively informal program that would take immediate advantage of the biomedical expertise concentrated in the

Rochester Atomic Energy Project to service the needs of a limited number of local physicians and researchers. However, soon thereafter the University of

Rochester would start vying for a more comprehensive educational program in radiation-related biomedical sciences.

In April 1947, soon after the AEC's Interim Medical Committee recommended the creation of a broad program in biomedical training, the AEC's

General Manager Carroll Wilson told Warren who had chaired the Interim

Committee that, "the Commission is interested in exploring the possibility of starting various types of training programs."401 He also called for proposals which was all the prompting needed for the University of Rochester to do so. Dr.

Andrew H. Dowdy, a radiologist at the University of Rochester's Strong

4(" Carroll Wilson quoted in Major Maxwell Dauer, "Memorandum to John A. Derry, University of Rochester Proposed Educational Program, 27 August 1947," NARA College Park, RG 326. General Correspondence 1946-1951, Box 79, 1. 235

Memorial Hospital and Director of the Rochester Atomic Energy Project, submitted the University's proposal in June 1947 on behalf of his colleagues in the School of Medicine and Dentistry and Faculty of Science.402

The University of Rochester's bid for an AEC-funded education program was an effort to ensure an ongoing relationship with the new AEC. The

Rochester Atomic Energy Project which, of course, was initiated under a MED contract, was assured funding through fiscal year 1948 because the AEC's Interim

Committee had arranged for interim funding for such contracts.4

Anxious to sustain government support, Dowdy and his colleagues pressed to transform the wartime project into a permanent one.

As Dowdy explained to the AEC, the University would organize the program through both its School of Medicine and Dentistry—specifically the

Departments of Biophysics and Pharmacology—and through the Atomic Energy

Project if the AEC maintained the contract. Instruction would be oriented towards three groups of students: graduate students, those requiring only short- term instruction on health protection techniques, and armed forces personnel.

4(L Andrew H. Dowdy, "Proposed Educational Program: Atomic Energy Project, The University of Rochester, 10 June 1947," NARA College Park, RG 326, General Correspondence 1946-1951, Box 79, 1-11. 4

protection techniques, and forty armed forces personnel requiring training in the I health problems associated with atomic energy. For the latter group, instruction could be provided in line with the requirements of the graduate program if the applicants wanted to and were qualified to complete a master's degree.

Alternatively, these students could complete a year of study which Dowdy characterized as a Certificate of Proficiency Program.405

Dowdy believed that for the success of this program, the AEC must

provide an appropriate building on campus. He recommended the construction of a new six-story building that would adjoin the School of Medicine and

Dentistry.406 The use of space within this building illustrates the range of research and groups the Rochester program sought to accommodate. The first floor was designated as a radiation therapy and tumor clinic which, as Dowdy pointed out in his proposal, would "serve the educational research activities of the Atomic

Energy Commission in much the same manner as an out-patient clinic serves a medical school, furnishing a readily available source of clinical material."407 That is, the clinic would provide researchers access to patients. Amongst other groups, the clinic would offer training opportunities for the Rochester-area physicians who had already requested such training, as well as armed forces personnel who needed clinical experience related to the biological and medical aspects of

405 Dowdy, "Educational Program, 1947," NARA College Park, Box 79, 6-7. The construction and equipment costs were estimated at $615,000. Dowdy, "Educational Program, 1947," NARA College Park, Box 79, 10-11. 107 Dowdy, "Educational Program, 1947," NARA College Park, Box 79, 10. 237 radiation. Two of the remaining five floors were designated for AEC research, one for radiological and biophysical research and the other for toxicological and pharmacological research. Two more would be set up as classrooms and laboratories for graduate students and armed forces personnel. The final floor was reserved for coordinated research. That is, Dowdy envisioned collaboration between the University and AEC on research of mutual interest and identified cancer research to be the likely focus of joint studies.408 These six floors, each with their own but related purpose, represent the merger of interests amongst the biomedical community, the government, and military.

Rochester Medical Center c. 1961, edited to indicate the AEC-funded addition. Photo Credit: University of Rochester Medical Center Historical Photographs, URMC Edward G. Miner Library.

Dowdy and his University of Rochester colleagues had already established the administrative framework necessary to ensure that coordination between the

University and the AEC could be achieved. The program would be guided by both an Educational Policy Committee and a Coordinating Committee. Here we see an example of how the Cold War transformed universities and we see how this process worked from the bottom up. The Educational Policy Committee was

408 Dowdy, "Educational Program, 1947," NARA College Park, Box 79, 11. 238

comprised of eight faculty members who represented various departments in the

Medical School and in the Faculty of Science, as well as the Deans of the

Graduate and Undergraduate Schools. The members of the Educational Policy

Committee defined for themselves the job of developing an educational program

that integrated atomic energy into fields of study including biophysics,

pharmacology, toxicology, biology, and medicine. They also sought to ensure

that the program they designed was in keeping with the existing educational

requirements for both undergraduate and graduate study at the University. The

Coordinating Committee was created with fifteen faculty members who would

advise the Director of the AEC's Atomic Energy Project as to his obligations to

the AEC; provide advice on experimental procedure, methods, and evaluation of

results; and coordinate research interests of the AEC and various departments of

the University.409

Both committees shared in the responsibility of transforming what was, at

the outset, presumed to be a short-term or wartime biomedical radiation research

program into a two-pronged program comprised of ongoing research and a complementary program of instruction. More broadly, these committees endeavored to design a comprehensive biomedical curriculum that corresponded to the interdisciplinary research pursued at the University of Rochester and elsewhere, most notably at Ernest Lawrence's Radiation Laboratory at the

University of California, Berkeley. While this research dated back to the pre-war

period, no university had instituted curricula specifically focused on the study of radiation as a research and medical tool and as a hazard for which measures of

'4"<) Dowdy, "Educational Program, 1947," NARA College Park, Box 79, 3-5. 239

protection must be devised.410 Accordingly, Dowdy stressed to the AEC that the

educational program he envisioned was new and would take time to develop.

Instructors would have to assemble course materials since no standard textbooks

existed.4"

Dowdy also had other reasons to press for rapid approval of this project.

For instance, he and others hoped that the program would be up and running in

the proposed new building by September 1948. If this were to be the case, the

construction of the new building would have to commence in the autumn of

1947.412 Perhaps the most important motivation, though, was the dwindling

number of professional staff involved in the Rochester Atomic Energy Project.

As of June 1947 when Dowdy submitted his proposal to the NYOO, the

Rochester Atomic Energy Project had 308 personnel. Of these 308,119 were

considered professional staff, 68 were technicians, 42 were administrative and

clerical staff, 23 were mechanics and engineers, and 56 were service personnel.

These numbers had declined considerably since the end of the war. Since

September 1945 the Rochester Atomic Energy Project had lost 343 personnel of

410 In the immediate postwar years the University of California, Berkeley also endeavored to establish an academic program that corresponded to the biomedical radiation research pursued at the Radiation Laboratory. Historians David S. Jones and Robert L. Martensen note that a Division of Medical Physics was established as part of the Physics Department in 1945. They argue that John Lawrence was instrumental in establishing that program as part of a larger effort to create a biomedical radiation program that was separate from the Medical School at the University of California, San Francisco. The Medical Physics program, which was established as a master's program alongside a doctoral program in Biophysics, later became part of a newly created Department of Biophysics. See David S. Jones and Robert L. Martensen, "Human Radiation Experiments and the Formation of Medical Physics at the University of California, San Francisco and Berkeley, 1937-1962, in Goodman, McElligott, and Marks, eds., Useful Bodies 87 and 98-99; and Oral History of Nello Pace, Ph.D., interview conducted August 16, 1994, as part of the Office of Human Radiation Experiments' Oral History Project, Human Radiation Studies: Remembering the Early Years. Conducted by Ms. Anna Berge in Berkeley, CA, 14 and 29-30. 411 "Letter to Mr. W.F.. Kelley from Andrew H. Dowdy, 10 June 1947," NARA College Park, RG 326, General Correspondence 1946-1951, Box 79, 3. 412 "Letter to Kelley from Dowdy, 1947," NARA College Park, Box 79, 2; and Dauer. "Memorandum to Derry, 1947," NARA College Park, Box 79, 3. 240 which 123 were professional, 90 were technicians, 55 were administrative and clcrical staff, 20 were mechanics and engineers, and 55 were service personnel.413

Many of those who left, especially the professional staff, returned to academic positions in which they enjoyed greater job security than was possible within the confines of a wartime research and development project. With regards to the personnel problem, Dowdy emphasized the need for stabilization and for work that was sufficiently interesting to not only stem the tide of terminations, but "to attract and retain desirable personnel."414 He believed these conditions must be met as soon as possible and hoped the AEC would be timely in reviewing his proposal.

In the months that followed the submission of Dowdy's proposal, Wilbur

E. Kelley, the Manager of the New York Operations Office (NYOO), played an important role in helping to relay information between Dowdy and the

Commissioners. The NYOO oversaw AEC contracts with northeastern universities. When the AEC first received Dowdy's proposal in June of 1947,

Kelley recommended that the proposed program, including the construction of the new building, be approved.415 By early August when the AEC had still not granted its approval, Kelley again tried to hasten the process. He contacted

Carroll Wilson, the General Manager of the AEC to reiterate Dowdy's message about the personnel problem and the need for prompt action. He conveyed that

413 Dowdy, "Educational Program, 1947," NARA College Park, Box 79, 1. 414 Dowdy, himself, would depart Rochester before the end of 1947 to take up a position at the new Medical School at the University of California, Los Angeles. "Letter to Kelley from Dowdy, 1947," NARA College Park, Box 79, 1. 415 Dauer, "Memorandum to Derry, 1947," NARA College Park, Box 79, 2. 241

the proposed program is planned as an integral part of the entire contract. As a result of the delay, the present research program and planning for future research, is disconnected. Personnel morale and interest are undergoing progressive deterioration and trained personnel are looking elsewhere for more definite prospects in their type of work.416

Kelley's message implied that the existing contract for the biomedical research program at the University of Rochester might be in jeopardy if the AEC did not approve the proposed educational program. His understanding stemmed from a description Dowdy had provided him which suggested that the proposed education program was completely integrated with the existing research program.

According to Dowdy,

To develop independent programs and building facilities for (1) the purposes of the Atomic Energy Commission, and (2) coordinated work in the University, would result in an unfortunate diversity of interests. It would dilute the energies and abilities of our staff to the point that one or the other projects, or both, might collapse.417

It is clear that Dowdy conceived of a program that was much broader than the wartime research program. Not only did the proposed program incorporate new educational infrastructure, including both a physical building and new curriculum, it was organized around the research interests of both the AEC and the University.

Thus, the program that Dowdy proposed combined civilian and military agendas—a blurring of lines that was very characteristic of the Cold War. The process by which Dowdy and his colleagues fashioned an enduring relationship with the AEC shows how University of Rochester faculty played an active role in creating that aspect of the Cold War.

W. E. Kelley, "Memorandum to Carroll Wilson, Educational Program Proposed by the University of Rochester, 5 August 1947," NARA College Park, RG 326, General Correspondence 1946-1951, Box 79, 1. 41' "Letter to Kelley from Dowdy, 1947," NARA College Park, Box 79, 2. The history of negotiations between the University of Rochester and the

AEC demonstrates that the creation of the ACBM in the autumn of 1947 was a turning point in the process of planning for AEC educational initiatives.

Throughout the summer of 1947, Dowdy, with the help of Kelley in the NYOO, had tried to drive the process forward. For instance, when Kelley contacted

Wilson in August he suggested that "the University is becoming uncertain over the desirability of maintaining their relationship with the Atomic Energy

Commission."418 Accordingly, he advised that the AEC approve, revise, or reject the Rochester program as quickly as possible and inform the University of a decision. Dowdy's proposal had been circulated amongst the Commissioners in

July, all of whom approved of the program. They did, however, have some reservations regarding the construction of the new building and they deferred the problem of training to the ACBM which was set to convene for its first meeting in

September 1947.419

Although the issue of biomedical education was referred to the Advisory

Committee on Biology and Medicine for "early consideration,"420 the issue did not make the agenda of the ACBM's first meeting.421 It was not until the second and third meetings in October and November that the ACBM discussed biomedical training and, specifically, the proposal from the University of

418 Kelley, "Memorandum to Wilson, 1947," NARA College Park, Box 79, 1. 4I" Dauer, "Memorandum to Derry, 1947," NARA College Park, Box 79, 2-3; and "Second Meeting of the ACBM, 1947," NARA College Park, Box 31, 5. 420 Dauer, "Memorandum to Derry, 1947," NARA College Park, Box 79, 3. 4"' The proceedings of the first meeting were largely focused on the relationship between the ACBM and the AEC and the selection of a Medical Director to take charge of the new DBM. See "Draft Minutes: First Meeting of the Advisory Committee for Biology and Medicine, 12 September 1947," NARA College Park, RG 326, General Correspondence 1946-1951, Box 31,1- 7. 243

Rochester, as well as one from Columbia University. When the issue arose at the

October meeting, Carroll Wilson suggested that the construction of new buildings on academic campuses was a policy question of special interest. Joseph Volpe, a representative of the AEC's legal staff who attended the meeting stated that there were no legal reasons that would prevent the AEC from funding the construction of new buildings. "There [is] little question," Volpe stated, "that the Commission has rather broad authority under the Atomic Energy Act to provide for a training program related to the activities authorized by the Act."422 The AEC did have choices, though, in terms of meeting its needs for training personnel and conducting research. Perhaps a more obvious course of action would have been to establish its own training programs within the national laboratories, not on university campuses. The issue was discussed in both the morning and afternoon sessions of the second meeting, but no policy decisions were made.

When the ACBM reconvened a month later, the committee discussed the proposed Rochester educational program and new building "at considerable length."423 A subcommittee formed within the ACBM to deal specifically with this issue moved that "As a step in meeting the urgent need for trained personnel, the Committee recommends support of the existing program at the University of

Rochester and the expansion proposal including the appropriation of a sum sufficient to furnish physical facilities...."424 The motion was passed unanimously. The ACBM opted to invest in a program that was already partially in place and draw on the resources—especially the organizational manpower—

4" "Second Meeting of the ACBM, 1947," NARA College Park, Box 31,8. 4"'1 "Third Meeting of the ACBM, 1947," NARA College Park, Box 31,3. 4~A "Third Meeting of the ACBM, 1947," NARA College Park, Box 31,3. 244

offered by the university. After months of correspondence between Dowdy,

Kelley, Wilson, and others, the educational program at the University of

Rochester was approved and the development of it could proceed. The AEC

announced its decision in a press release in December 1947.425

Columbia University's similar proposal, however, was subsequently

rejected. Although the proposal was submitted by Gioacchino Failla, a leading

expert in the biological and medical aspects of atomic energy and a researcher

who was well known amongst the ACBM members, support of a second training

program at Columbia seemed redundant. The ACBM rejected the Columbia

proposal based on the fact that Columbia was in the same geographical region as

Rochester and the latter had the advantage of already having a well established

program upon which to build.426 Regional considerations factored into this and

many other decisions. For instance, historian Allan Needell argues that the

creation of Brookhaven National Laboratory was driven by scientists who sought

to establish a state of the art laboratory that would rival the already established

national laboratories in other regions of the country. Region also mattered for

political reasons. The AEC, like other federal agencies, considered the political

benefits of distributing their resources in various regions.

THE EARLY DAYS OF THE AEC'S ROCHESTER ATOMIC ENERGY PROGRAM

Carroll L. Wilson, "Memorandum to W. E. Kelley: Medical Research and Training Center for University of Rochester, 28 November 1947," NARA College Park, RG 326, General Correspondence 1946-1951, Box 79, 1-4. 426 "Third Meeting of the ACBM, 1947," NARA College Park, Box 31,6; Needell, "Brookhaven National Laboratory," 96 and 104. 245

Just after the Rochester proposal was approved, Dowdy stepped down as

Director of the Rochester Atomic Energy Project. Dr. Stafford Warren, who had accepted the position of Dean at the new School of Medicine at the University of

California, Los Angeles, recruited Dowdy to take up a position as a founding member of the faculty. Dr. Henry A. Blair, a physiologist and biophysicist in the

School of Medicine and Dentistry succeeded Dowdy as the Director of the

Rochester program.427 Blair had ample support amongst his University of

Rochester colleagues; key among them was Dr. J. Newell Stannard, a physiologist and biophysicist. Stannard was a former Rochester faculty member who left the

University during the war to serve as a naval officer. He returned to Rochester in

1947 and was appointed to the position of assistant director for Education in the

Atomic Energy Project.428 Blair, Stannard, and their colleagues proceeded with the Rochester Atomic Energy Project in an effort to meet the research and training needs of the AEC and to develop an innovative biomedical curriculum that focused on using nuclear technologies and materials, many of which were newly created or discovered, in research and clinical practice.

The novelty of Rochester's program proved to be a draw for some students. For instance, after completing an undergraduate degree William J. Bair planned to pursue graduate studies in chemistry at the University of Ohio. Upon finding out about the Rochester Atomic Energy Project he instead completed his graduate studies at the University of Rochester. Working under Stannard's

427 Wilson, "Memorandum to Kelley, 1947," NARA College Park, Box 79, 3. 4"K William Bair, "Introduction," in J. Newell Stannard and the University of Rochester: A Collection of Papers Presented at the 48th Annual Meeting of the Health Physics Society, ed. J. Newell Stannard (San Diego, CA: Health Physics Society, 2003), 5. 246 supervision, Bair received the world's first ever Ph.D. in Radiation Biology in

1954.429 The Department of Radiation Biology did not exist when Dowdy first

arranged for the expanded Rochester Atomic Energy Project. It was created shortly thereafter as a result of the University's efforts to institutionalize the

hybrid nature of biomedical radiation research in the new curriculum they developed for their AEC-funded educational program. Decades later Bair

reflected on his own experience and described the University of Rochester's

program as a unique opportunity. While the curriculum developed at the

University of Rochester in association with the Rochester Atomic Energy Project constituted a unique opportunity in and of itself, the program was made more accessible by the initiation of the AEC's fellowship program. As we will see, the

fellowship program provided students throughout the country financial support to pursue studies in biomedical sciences related to radiation. Although, the AEC's

fellowships could be used to support studies at any accredited university, the fact that the University of Rochester was one of a few universities with a program specifically devoted to radiation studies, meant numerous fellowship students enrolled there.

THE AEC FELLOWSHIP PROGRAM: OPPORTUNITIES & PROBLEMS

Shortly after the University of Rochester was authorized to move forward with its AEC-funded educational program, the AEC took up the issue of fellowships in earnest. Both the AEC's Interim Medical Committee and the

4"'' Ibid., 5-6; and Oral History of William J. Bair, Ph. D., interview conducted October 14, 1994 as part of the Office of Human Radiation Experiments' Oral History Project. Human Radiation Studies: Remembering the Early Years. Conducted by Cindy Shindledecker and David Harrell at the Pacific Northwest Laboratory in Richland, WA, n.p. 247

NAS/NRC's Medical Board of Review had recommended the creation of an AEC fellowship program that would provide opportunities for pre-doctoral and post­ doctoral studies in biomedical and physical sciences related to the uses and hazards of radiation. These committees also recommended fellowships for technical training in health physics and related research. As was the case with the

University of Rochester educational program, it was not until the ACBM and

DBM were in place that the AEC made significant progress towards establishing

the recommended program. Once the ACBM endorsed the earlier recommendations in November 1947, the Director of the DBM, Dr. Shields L.

Warren, began to formulate concrete plans for fellowships.430 Since the fellowships would encompass both the physical and biomedical sciences, Warren collaborated with the Director of the Division of Research, James B. Fisk, to design the program.431

By January 1948, Warren and Fisk had obtained the Commissioners' approval of the fellowship program they had fashioned for medical, biological and physical sciences 432 As had been recommended by the Interim Medical

Committee, the Medical Board of Review, the ACBM, and finally proposed by

Warren and Fisk, the fellowship program was operated by the National Research

4311 Shields L. Warren, "Training Program for Physicians in Medical Aspects of Atomic Energy, 16 January 1948," NARA College Park, RG 326, General Correspondence 1946-1951, Box 38, 5. 411 Roy B. Snapp, "Note by the Secretary: Training Program for Physicians in Medical Aspects of Atomic Energy, 16 January 1948," NARA College Park, RG 326, General Correspondence 1946- 1951, Box 38, 1. 4,2 No specific plans were made for a formal health physics fellowship program, but some of the fellowships awarded in biology supported students in this field. See Kenneth S. Pitzer, "Fellowship Program - Ratio of Fellowships in the Physical and Biological Sciences, 14 February 1949," NARA College Park, RG 326, General Correspondence 1946-1951, Box 38, 1. 248

Council (NRC) of the National Academy of Sciences (NAS).433 The NRC was

quick to announce the program so that fellowships could be granted for the 1949

academic year. As of June 1948, only a few had been awarded, but by December

of that year the number of fellows had increased to 264. The AEC hoped to grant

350 fellowships the following year, 175 each for the biomedical and physical

sciences.434

Throughout the first few years of the fellowship program, the Directors of

the DBM and Division of Research continued to work in concert with the

Commissioners and the NRC to fine tune the program. The main policy issues

that concerned them related to the number of fellowships to be awarded; the ratio

of fellowships between the biomedical and physical sciences; the support of

fellows conducting research outside of the United States; the selection of non-

United States citizens as fellows; and the issue of security clearances.435 The

latter proved to be the most troublesome issue that plagued the program and

required numerous policy and program changes.

UNDER THE MICROSCOPE: FELLOWSHIPS & SECURITY INVESTIGATIONS

During the summer months of 1948 the Commissioners started to discuss

whether or not fellows should be subject to security investigations. The issue of

",',J Warren, "Training Program, 1948," NARA College Park, Box 38, 5; and Pitzer, "Fellowship Program, 1949," NARA College Park, Box 38, 1. 414 For 1950 and following, the AEC projected an increase of 50 additional biology and medicine fellowships to make a total of 225 granted in any one year. Pitzer, "Fellowship Program, 1949," NARA College Park, Box 38, 1 and 6. 415 For instance, see Pitzer, "Fellowship Program, 1949," NARA College Park, Box 38; Kenneth S. Pitzer and Shields L. Warren, "Award of AEC Fellowships for Study Abroad, 13 December 1948," NARA College Park, RG 326, General Correspondence 1946-1951, Box 38; Shields L. Warren, "Security Clearance of Fellows Participating in the AEC Fellowship Programs, 10 September 1948," NARA College Park, RG 326, General Correspondence 1946-1951, Box 38; and "Fellowship Program, Hearings, 1949," NARA Atlanta, Box 31. 249 security clearance for individuals involved in the atomic energy enterprise was not new, nor was it confined to the fellowship program. Fuelled by fear of communist infiltration and Soviet espionage both during and following the war, the loyalty and political affiliations of many scientists were investigated before they were granted clearance to work on weapons research and development.

Perhaps most famously, J. Robert Oppenheimer, the physicist who was the wartime director of Los Alamos and who following the war was celebrated as one of the nation's leading scientists, was the subject of ongoing security investigations. Despite his crucial contribution to the Manhattan Project and his immense scientific talent, Oppenheimer, had his security clearance revoked in

1953. An A EC hearing that ensued in 1954 exhibited the many conflicting opinions that were held within and without the AEC on the relationship between science, politics, and national security.436

By 1948 when the AEC was in the midst of creating its fellowship program, Cold War tensions between the United States and Soviet Union were such that Soviet espionage and, more broadly, the employment of communists within the AEC caused many individuals concern.437 Despite this, the

Commissioners established a security policy in September 1948 that required only those fellows who engaged in restricted research to undergo a security

'l3h For a discussion on the impact of anticommunism on the scientific enterprise and on J. Robert Oppenheimer, in particular, see Kai and Martin J. Sherwin Bird, American Prometheus: The Triumph and Tragedy of J. Robert Oppenheimer (New York: Alfred A. Knopf, 2000); Wang, American Science in an Age of Anxiety. 4,7 There is a rich literature on the start of the Cold War. For a discussion on the fear and insecurity that fuelled the conflict between the United States and Soviet Union see, for example, John Lewis Gaddis, The Cold War: A New History (New York: Penguin Books, 2005), 250

investigation. 4^8 They did so after consulting with the AEC's General Council,

the DBM, and the NRC and after reviewing the findings of security investigations completed for 45 early applicants. These were done in an effort to determine what sort of individuals were applying to the program and if they posed any threat.439

The issue of security clearance did not, however, disappear after the AEC

policy was in place. While the AEC had been wrestling with the security issue throughout the summer of 1948, so too was the Joint Committee on Atomic

Energy (JCAE). The members of the JCAE were concerned that through the fellowship program the government might provide educational opportunities and, worse, employment opportunities for Communists or those whose loyalty was suspect.440 One member, Senator Bourke B. Hickenlooper, explained that the

JCAE was especially vexed that of the 45 investigations completed early in the fellowship program, a few applicants were found to have Communist affiliations.

Unwilling to let the issue drop, the JACE decided to hold hearings to investigate the fellowship program in May 1949. 441

Throughout the JCAE hearings, Lilienthal described the Commission's responsibility to support research and training in biology, medicine, and the physical sciences and defended the Commission's policy on security. He emphasized that the fellowship program did not support "the secret side of atomic

w Roy B. Snapp, "Note by the Secretary: Security Clearance of Fellows Participating in the AEC Fellowship Programs, 10 September 1948," NARA College Park, RG 326, General Correspondence 1946-1951, Box 38, 1-2. Warren, "Security Clearance, 1948," NARA College Park, Box 38, I; and Pitzer, "Fellowship Program, 1949," NARA College Park, Box 38, 1. "Fellowship Program, Flearings, 1949," NARA Atlanta, Box 31,5. "Fellowship Program, Hearings, 1949," NARA Atlanta, Box 31.8, 251

energy development" in an attempt to allay fears that the program might facilitate

access to knowledge and practices specific to the atomic weapons industry 442 He

was joined by a few fellow Commissioners as well as DBM Director Shields

Warren and Dr. A. Newton Richards who, as President of the NAS, would speak

on behalf of both the NAS and NRC.

The detailed discussion that took place amongst all those in attendance

illuminates the tensions that arose out of the blended civilian and military spheres

of the postwar years. For instance, Lilienthal conveyed the common belief that

investigation of fellows engaged in non-secret research would

seriously threaten the interests of the fellowship program as a whole and therefore of scientific progress. More than that, the introduction of security procedures into nonsecret [sic] fields, it was feared, would establish a precedent of grave and far reaching consequence to our scientific and educational system.443

Here, Lilienthal sought to establish a boundary between civilian and military spheres wherein security investigations were acceptable only in the latter. NAS

President Richards reinforced the idea that science and education should not be tainted by national security procedures. He spoke to his extensive experience as a

teacher and what he had observed of young people, especially the sort of student

he anticipated would apply for fellowships. He considered the intellectual abilities and curiosity of such students to be commendable and hoped not to deter any bright young individuals from applying for the AEC-NRC fellowship due to the establishment of FBI investigations.444

442 "Fellowship Program, Hearings, 1949," NARA Atlanta, Box 31,2. 44' "Fellowship Program, Hearings, 1949," NARA Atlanta, Box 31,8. 444 "Fellowship Program, Hearings, 1949," NARA Atlanta, Box 31,13 The JCAE hearings examined numerous issues in great detail: the AEC's responsibilities in education and training; the implications of educating

Communists or individuals suspected of having had Communist affiliations; the nature of fellowship appointments, specifically whether they constituted government employment or the guarantee of future government employment; and much more. Without belaboring every point, the examples explained here show that the creation of AEC educational programs helped unite science, government and national defense, and did so with benefits to science, but also some drawbacks. A range of individuals from the scientific community and the AEC who, through scientific advisory boards such as the ACBM were well informed by scientific community, used postwar and early Cold War opportunities to expand government support of science. In the case of biomedical sciences, they did so hoping to resist the intrusion of national security concerns that influenced research and development more directly tied to weapons. Their ability to do so, however, proved to be limited due to the AEC's dual role in advancing atomic science, but also protecting atomic secrets.

THE FELLOWSHIP PROGRAM IN TRANSITION

Despite the arguments Lilienthal, Richards and others including DBM

Director Dr. Shields Warren mounted against security investigations, the JCAE hearings resulted in the Independent Offices Appropriations Act for fiscal year

1950. The Act, which went into effect in August 1949, required that fellows be investigated as to their "character, associations and loyalty" and granted security 253

clearance by the Federal Bureau of Investigation (FBI).445 Historian Jessica

Wang argues, and uses the fellowship hearings to illustrate, that by 1949 Cold

War ideology had effectively restricted scientists' ability to express dissent from

unfavorable policies such as security investigations. Her perceptive analysis

shows that scientists' opposition to the new security policy was stifled and that

they were forced to proceed within the confines of a national security-oriented,

Cold War state 446 Building upon her argument, but departing from it somewhat,

this study shows that the existing relationship between the AEC, NAS-NRC, and

various individuals from the larger scientific community belied an easy

interpretation of the new security policy as a triumph of one distinct group—the state—over another—the scientific community. The very existence of an

increasingly strong relationship between science and the state suggests that the scientific community could not be neatly categorized as entirely separate from the state. Following the war, the scientific community and its relationship with government agencies like the AEC, in fact, helped create the Cold War state.

The AEC's and NAS-NRC's reaction to this change in circumstances demonstrates that the partnership between science and the state was strong enough to be molded to accommodate the new circumstances. Indeed, in an effort to save the fellowship program, the AEC worked with the NRC and other groups to adapt the fellowship program in accordance with the Independent Offices

Appropriations Act, but ensure that the overall agenda to expand research

'w''' "AEC Announces Provisions for Operation of Fellowship Program During 1950-1951 Academic Year, Press Release for 16 December 1949," NARA College Park, RG 326, General Correspondence 1946-1951, Box 38, 1. 446 Wang, American Science in an Age of Anxiety, 219-52. 254 infrastructure was maintained. Since the NAS and NRC publicly opposed the new policy, the first changes made were done so in an effort to relieve the NRC from managing the program or at least the aspects that required investigation of fellows conducting non-secret work.

At this point, the value of newly created organizations such as ORINS and relationships between the AEC and universities like the University of Rochester was especially apparent. Both of these institutions, along with a few others, served as able and willing partners to help the AEC sustain the fellowship program during turbulent years.447 In 1950, the National Science Foundation

(NSF) was created as an independent federal agency that would support basic research and education in all fields of science. The NSF announced that it would initiate a fellowship program, after which the AEC decided it no longer needed to provide fellowships.448 This decision was not a reflection of waning interest in education, though.

As AEC General Manager Marion W. Boyer expressed to the NSF, the

AEC placed a great deal of value in the program.449 "It is the belief of the

Commission," Boyer wrote, "that its fellowship program will make a very substantial contribution to our national strength in atomic energy over the years ahead. Very tangible returns are already in view...."450 He helped to illustrate the

"tangible returns" by providing statistics on the articles published and degrees

447 The University of Rochester and ORINS even helped reinstate fellowships for health physics which had been discontinued. See "Fellowship Program During 1950-1951, 1949," NARA College Park, Box 38, 2-3. 44>< "Letter to Dr. Waterman from M. W. Boyer, 5 April 1951," NARA College Park, RG 326 General Correspondence 1946-1951, Box 38, 3. 44l) Marion W. Boyer succeeded Carroll L. Wilson as General Manager on 1 November 1950. Hewlett and Duncan, Atomic Shield, 666. 4511 "Letter to Waterman from Boyer, 1951," NARA College Park, Box 38, I. 255 completed by fellowship recipients. Of the 650 fellows appointed as of spring

1951, not all of whom were doctoral students, 101 had already completed Ph.D.s.

Another 114 fellows were enrolled in Ph.D. programs, awaiting completion.

Also, as of January 1951 AEC fellows had published a total of 117 scholarly articles as well as dozens of AEC reports.451 The point Boyer was making was that this was a program that should be maintained and while the AEC would willingly do so, the AEC was ready to discontinue the program if the

Commissioners were confident that the NSF program would accomplish the same goals. Thus, the AEC ceased to grant fellowships, though not until 1953. It maintained a scaled-down program to that point to ensure that existing fellowships were renewed.452

OAK RIDGE INSTITUTE FOR NUCLEAR STUDIES' EDUCATIONAL ACTIVITIES

Beyond assisting with the fellowship program, the ORINS played a significant role in helping administer the AEC's biomedical education and training programs. For instance, under AEC contract, the ORINS's Special

Training Division created a series of short-term courses related to uses of radioisotopes in research and medicine to be held at Oak Ridge National

Laboratory (ORNL). A four-week course on Radioisotopes Technique was initiated in 1948, a two-week course on Advanced Instrumentation in 1950, and a

4M "Letter to Waterman from Boyer, 1951," NARA College Park, Box 38, 2. 452 Shields L. Warren and Thomas H. Johnson, "AEC Fellowship Program: Report by the Director, Division of Biology and the Director, Division of Research, 20 December 1951," NARA College Park, RG 326, General Correspondence 1951-1958, Box 142, 3. 256 two-week course on Radioisotopes in Medicine in 1951,453 These short-term courses, which were offered multiple times each year, filled an educational niche that the fellowship program could not in that they were aimed at researchers and physicians already working in their field rather than graduate students and recent doctoral graduates. As of November 1950, before the advanced course on

Radioisotopes in Medicine had begun, nearly 600 individuals had already participated in the 4-week course on Radioisotopes Technique. More than 200 of these were physicians.454

In addition to the short-term training initiatives organized through the

Special Training Division, ORINS administered a range of educational programs through its University Relations Programs. For instance, the Graduate Training

Program offered doctoral and post-doctoral fellowship opportunities. Although seemingly similar to the AEC fellowship program, the University Relations

Division's fellowships supported graduate and post-graduate research opportunities at ORNL exclusively. AEC fellowships could be held at a variety of academic and medical institutions as well as the national laboratories.455

451 "32 Students...to Receive Radioisotope Training, n.d.," NARA College Park, Box 30, 1; Herman M. Roth and A. J. Vander Weyden, "Summary of ORINS Educational Activities, 5 December 1950," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 33, 8-11; and Oak Ridge Institute of Nuclear Studies, "Announcement of Course in Radioisotopes in Medicine, n.d.," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 33, 1. 454 Ralph T. Overman, "Letter to Dr. Gilbert Noyes, Correspondence re Advanced Medical Course in Radioisotope Therapy & Tracer Studies, November 1950," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 33, 1. 455 William G. Pollard, "Letter to Shields Warren, 21 April 1948," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 43, 1- 2; "Corporation Minute Book, 1949," NARA Atlanta, Box 34, 7-8. 257

Executive Director of ORINS William G. Pollard is pictured here (center) c. the mid-1950s during a visit made by Eleanor Roosevelt. Photo Credit: Oak Ridge National Laboratory's Historical Photo Gallery.

Starting at the end of 1948, the ORINS also developed a medical program

under contract with the AEC. This program specifically focused on advancing clinical investigations and physician training using radioactive materials to treat and better understand the nature of cancerous tumors. Over time, the program

would involve four initiatives: a directed research program, a training program, an

information program, and special projects organized with cooperating medical schools.456 To manage the program, the ORINS created a Medical Division under the leadership of Dr. Marshall Brucer, formerly a physiologist at the University of

Texas. After Brucer's appointment in December 1948, he and the small research staff with whom he worked began planning a program of research and the construction of a 30-bed clinical hospital with research facilities.457 Throughout

the 1949-1950 academic year the Medical Division was mostly concerned with developing and establishing the hospital and associated laboratories; hiring

researchers, nurses and technicians; and establishing an animal farm and

456 "Program of Research Medical Division, n.d.," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 33, 1. ',s7 Pollard, "The ORINS - A New Pattern, 1949," NARA Atlanta, Box 33, 10-11. laboratory.458 It was expected that once the proper facilities were in place, the

Division would remain active in research while also administering training opportunities for those institutions partnered with the ORINS.459

Dr. Marshall Brucer pictured with mannequins designed to evaluate instruments used to measure iodine uptake, c. 1950s. Photo Credit: Oak Ridge Associated Universities Health Physics Historical Instrumentation Museum Collection.

ORINS Medical Division's instrumentation room, c. 1950s. Photo Credit: Oak Ridge Associated Universities Health Physics Historical Instrumentation Museum Collection.

In accord with the general mission of the ORINS to facilitate the use of

Oak Ridge facilities by regional institutions, the Medical Division developed its program in cooperation with 22 participating southern medical schools. Students from these institutions could complete part of their residency training through the

ORINS medical program and participating institutions shared in the responsibility of providing patients for clinical studies.460 Due to the close cooperation with

45s "Program of Research Medical Division, n.d.," NARA Atlanta, Box 33, 1. 4?" "Corporation Minute Book, 1949," NARA Atlanta, Box 34, 9. 4W) Pollard, "The ORINS - A New Pattern, 1949," NARA Atlanta, Box 33, 10. 259 university medical schools, the AEC considered the OWNS an ideal organization to manage this program. The Commission assumed that neither an industrial contractor nor the AEC itself could administer a program that catered to medical school curricula as effectively as the ORINS which was, from the outset, a university-based organization.461 The ORINS medical program, along with the short-term courses organized by the Special Training Division and various fellowship and resident research programs managed by the University Relations

Division, highlight the role of the ORINS as an important partner to the AEC.

The ORINS helped the AEC create and administer numerous education and training programs. Together, all of these initiatives amounted to a significant effort on behalf of the AEC and ORINS to expand education and training related to atomic energy. They provided students, Oak Ridge personnel, researchers and physicians a range of opportunities through which to acquire knowledge and skills related to radiation hazards and the uses of atomic energy in research and medicine.462

CONCLUSION

As the Manhattan Engineer District prepared to transfer control over the nation's atomic enterprise to the AEC, the future of government-funded biomedical radiation research was uncertain. Even after the AEC was formally

4"' Pollard, "The ORINS A New Pattern, 1949," NARA Atlanta, Box 33, 10-11. 4',;: The ORINS' University Relations Division administered a few other educational programs including, but not limited to the Resident Graduate Program and Research Participation Program. The first facilitated government employees and contractors to work towards obtaining a M.Sc. in atomic energy-related studies at the University of Tennessee. The second provided university faculty research and training opportunities at ORNL during the summer months when they were not engaged in teaching. See, for example, "Corporation Minute Book, 1949," NARA Atlanta, Box 34, 7-9; and "Oak Ridge Educational Programs, n.d.," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 28, 1. 260 established, almost a year passed before the Division of Biology and Medicine and Advisory Committee on Biology and Medicine were created. Throughout the transitional period of 1946 and 1947, though, biomedical radiation research remained an important issue both within the MED and AEC bureaucracy and without. The development of the AEC's education and training initiatives from the end of World War II through to approximately 1950 illustrate that numerous individuals and groups were concerned about and motivated to ensure that the new agency advance biomedical radiation research by providing education and training opportunities in biomedical fields. Institutions such as the University of

Rochester, the newly formed ORINS, and the NAS-NRC not only encouraged the

AEC to create education and training programs, they actively sought to be a part of these programs.

The cooperative endeavors between these institutions were in part motivated by the AEC's responsibilities related to atomic weapons and also the agency's responsibility to develop atomic energy for civilian purposes. They were also motivated by the goals of institutions and individuals dedicated to scientific advancement and both discipline and institution building.463 Indeed, the programs described throughout this chapter illustrate the willingness of those within the scientific community to partner with the AEC to achieve the common goal of increasing the number of scientists and physicians trained to work with radiation. For instance, scientists at the University of Rochester were eager to extend and transform their wartime contract with the MED; the university sought

"tw For further discussion of academic institution building throughout the Cold War, see Dennis, "Two University Laboratories in the Postwar American State," 427-55; and Leslie, The Cold War and American Science. 261

the AEC's financial support for the creation of educational programs that

institutionalized cutting-edge biomedical radiation research begun just prior to the

start of the war. For those at the southern universities that created the ORINS

partnership, the postwar years seemed to be an excellent time to develop new

areas of research that would benefit the participating institutions and the local

economy. Both the University of Rochester and the ORINS hoped to achieve

these goals by aligning themselves with the AEC's research and development

responsibilities. At times, these institutions were the driving force behind the

development of the AEC's educational programs. However, the ACBM and

DBM, once created in the fall of 1947, were instrumental in transforming plans

into actions. The ACBM provided expert advice to the Commissioners regarding

proposed programs and the DBM helped design and carry out programs that were approved.

Like the AEC's radioisotopes program, education and training programs

linked the interests of the scientific community and the state. That the partnerships formed for educational initiatives were possible was, in part, a reflection of Cold War culture that equated national security with advanced scientific and technological expertise. These partnerships, however, also helped create that culture. Rather than return to the limited relationship that existed between academic science and the government throughout the early twentieth century, both the scientific community and the AEC facilitated the extension of the wartime relationship between science and the state into the postwar years. In education and training, as in so many of the AEC's programs and responsibilities, 262 the line between what constituted civilian and defense work was faint, although not entirely obscured. The JCAE hearings on security investigations of AEC fellows illuminated the blending of these spheres, but also attempts made by some researchers and administrators to maintain a divide.

The gradual development of plans and initiation of programs at various institutions resulted in an array of educational initiatives that, cumulatively, greatly expanded the opportunities for biomedical education and training related to atomic energy. New educational programs and infrastructure for biomedical research served as an investment to help ensure that in the future national security could be protected and medical advances made. Almost immediately following the war, the AEC made a significant contribution to biomedical radiation research by initiating the radioisotopes distribution program. The education and training programs created gradually during the late 1940s strengthened the AEC's commitment to biomedical research. Both radioisotopes distribution and expanded biomedical expertise were integral to the AEC's responsibility to attend to radiation hazards. Furthermore, they were necessary complements to the

AEC's cancer program—the final component of the AEC's biomedical initiatives and one that helped form the increasingly strong bond between biomedical sciences and the state. 263

CHAPTER 6

WAGING WAR ON ANOTHER FRONT: THE AEC JOINS THE CANCER ESTABLISHMENT

Late in his life, the radiologist Dr. Frederick Bonte recalled the

early days of his career in the 1940s when there was enormous optimism about

the future of cancer research and therapy. He had been fascinated to learn that

physicians were using radiation "directed in an intelligent way, to destroy tumors

that were not amenable to any other treatment." With radiation, he explained,

"You could destroy it and on some occasions you could cure it."4''4 Bonte was

especially hopeful about the application of radiation to the problem of cancer

because he encountered more experienced radiologists who helped instill this

hope. In 1948 he worked at Western Reserve University where Dr. Hymer

Friedell was the chief radiologist. We have seen in previous chapters that

preceding World War II Friedell had worked at the Chicago Tumor Institute,

Memorial Hospital in New York, and the University of California, San Francisco hospital. During the war he was Deputy Chief of the Manhattan Engineer

District's Medical Office. For years, then, Friedell had been well acquainted with cutting edge biomedical radiation research. Bonte recalled that "[Friedell] kept telling me that within a decade, cancer will be a thing of the past, that there will absolutely be whole families of radioactive drugs that will be selectively taken up by tumors and extirpate them."465 Friedell's confidence in the usefulness of

4W N.A., "Interview of Frederick J. Bonte in 'People in Nuclear Medicine: Interviews with Physicians, Scientists and Industry Leaders,'" The Journal of Nuclear Medicine 34, no. 6 (1993): 20N. 465 Ibid20N. 264

radiation represented the rapidly spreading sense amongst scientists and

physicians that radiation might prove to be a remarkable diagnostic and curative agent. For Friedell and other biomedical researchers, the key to uncovering the diagnostic and therapeutic uses of radiation in the fight against cancer was research.

At mid-century there were government agencies and private organizations in the United States, such as the National Cancer Institute (NCI) and American

Cancer Society (ACS), which were committed to pursuing cancer research with the hope of expanding and accelerating the pace of research. Indeed, the NCI was created in 1937 as an agency devoted to advancing cancer research. The ACS, which had been created in 1913, also launched a significant research program in

1945. Such organizations constituted one part of what is often referred to as the

Cancer Establishment. According to historian James Patterson, the Cancer

Establishment formed as a coalition of upper-middle-class groups that included these organizations, as well as the researchers they supported, physicians, journalists, and the pharmaceutical companies that helped bring new cancer treatments to the market.466 Given that the AEC's primary responsibility was to manage and develop nuclear weapons, it was not as likely a candidate as the NCI and ACS to play a significant role in the postwar Cancer Establishment. The

AEC did, however, develop a cancer program that helped establish both the growing importance of research within the Cancer Establishment and the use of radiation for research and the diagnosis and treatment of cancer.

"" Patterson, The Dread Disease, viii-ix. What role did cancer play in the AEC's biomedical program? In

particular, what were the primary objectives of the AEC's cancer program and was the problem of cancer the raison d'etre for the AEC's biomedical research?

This chapter argues that the AEC identified cancer research as a priority, but not the sole priority to be pursued at the expense of all other biomedical research.

Rather, the formal creation of a cancer program reinforced and, in turn, was reinforced by other aspects of a broad biomedical research program. That the

AEC's canccr program was established as part of a larger biomedical program, stemmed from the fact that the AEC's biomedical research evolved with input from numerous scientists and administrators both within and without the AEC.

Collectively, they pressed the AEC to facilitate the expansion of infrastructure and manpower to investigate the uses of radiation as a research, diagnostic, and therapeutic tool. Since the AEC primarily focused on directing resources and expertise toward building a foundation for radiation research, the AEC was able to contribute to cancer research while building a broad biomedical research enterprise.

The AEC's cancer program supported the long-term goal of curing cancer, but the AEC was careful not to promise too much too soon in terms of doing so.

Instead, the AEC publicized its efforts to expand the knowledge base. This strategy reflected the reality that although the AEC entered the cancer research community, it was radiation, not cancer, that drove the AEC's research.

Accordingly, the development of the Commission's radioisotopes distribution 266

program and training initiatives were integral to the AEC's endeavor to build a

foundation for cancer research.

THE RADIATION-CANCER RELATIONSHIP

Throughout the early stages of planning for an AEC biomedical program, cancer was a common topic. For instance, when John R. Dunning, a Columbia

University physicist, wrote to General Leslie R. Groves in 1945 to encourage the distribution of radioisotopes to researchers he noted the particular use of carbon-

14 (C-14) and hydrogen-3 (H-3) for cancer research.467 Similarly, when Dr.

Andrew Dowdy and his colleagues planned for the building of a medical center at

the University of Rochester they designated one of six floors as a radiation

therapy and tumor clinic and another as a floor for joint AEC-University of

Rochester studies which he suspected would likely focus on cancer, a problem of common interest.468 Cancer was certainly a problem of interest to the AEC and those such as Dunning and Dowdy who collaborated with and, at times, pressed

the AEC to support biomedical research. For them, cancer was a problem that

intersected with radioisotopes distribution and training initiatives.

The links between cancer, radioisotopes, and training had been established before the war in laboratories such as the Berkeley Radiation Laboratory where researchers pioneered new fields of research utilizing the cyclotron and the

radioisotopes and radiation beams the cyclotrons could produce. And these links

467 "Letter from John R. Dunning to General Leslie R. Groves, 29 October 1945," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945- 1952, Box 36, Requests for Radioactive Materials for Institutions, 1. 4Wi Andrew H. Dowdy, "Proposed Educational Program: Atomic Energy Project, The University of Rochester, 10 June 1947," NARA College Park, RG 326, General Correspondence 1946-1951, Box 79, 10-11. 267

built upon the relationship between radiation, cancer, and scientific and medical

specialization established even earlier in the twentieth century through the study

and use of X-rays and radium. In the context of the AEC's postwar research and development, cancer served as common ground between biomedical scientists and

the government—ground upon which they cooperated to build biomedical

infrastructure.

The AEC's interest in cancer research also stemmed from a sense of

responsibility. Dr. Shields L. Warren, Director of the Division of Biology and

Medicine, elaborated on this point in 1953, approximately five years after the

AEC launched its cancer program. In a dedication of the AEC-funded Argonne

Cancer Research Hospital he stated that "some not wholly familiar with the field of atomic energy may ask, 'Why is the Atomic Energy Commission interested in cancer?'"469 Warren explained that cancer is a specific industrial hazard of atomic energy and that the AEC, therefore, accepted responsibility for preventing cancers caused by radiation exposure. Beyond that, he expressed the AEC's belief that controlled radiation could be used to treat cancer and the AEC's commitment to developing radiation therapy. While Warren's speech explained why the AEC was involved in cancer research, it did not take into account the careful planning that was done to determine the role of the AEC in cancer research relative to its role in biomedical research generally. Decisions made by the Medical Board of Review and Advisory Committee for Biology and Medicine in 1947 and 1948 established the AEC's strategy for achieving both ends.

"Argonnc Cancer Research Hospital, 14 March 1953," NARA College Park, RG 326, General Correspondence 1951-1958, Box 67, 2. 268

THE MEDICAL BOARD OF REVIEW SETS THE TONE FOR THE AEC'S CANCER PROGRAM

As previously examined in Chapter Three, the Medical Board of Review was formed to provide the AEC advice on biomedical research and the feasibility and desirability of creating a biomedical program within the AEC. The meetings of the Medical Board of Review took place in June 1947 and involved a panel of experts as well as some of the AEC's Commissioners and staff. Throughout the course of the meetings, those assembled considered the tools the AEC possessed—namely radioisotopes and technologies that produced radiation sources—and the feasibility and responsibility of using these tools in biomedical research. Commissioner Pike explained to the members of the Medical Board of

Review that they would play an important role in determining how to make good use of radiation and the AEC's technologies in research. "My conception might be quite wrong," Pike contended, "but it has seemed to me that one of the most immediately useful things you could do here would be to find out where these new and perhaps useful things can go into the solution of questions that are already terribly bothersome—where they've gone maybe so far and may need a new line of attack or where perhaps the present line of attack could be accelerated by putting these into the picture."470

The task that Commissioner Pike laid out for the Medical Board of

Review members to define the areas of research in which the AEC's radiation- related technologies might be particularly useful in no way narrowed the possible

4711 "Transcript: First Meeting of the Medical Board of Review, June 1947," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 40,30-31. areas of research to those directly related to cancer. Cancer was, however, a topic

that arose numerous times throughout the meetings and determining effective uses of the AEC's technologies and radioactive materials in cancer research, diagnosis, and treatment, became a core objective of the AEC's cancer program as it evolved. The subject of cancer arose, for instance, when in response to

Commissioner Pike one of the members of the Medical Board of Review, Dr.

Herbert Gasser, spoke about what he considered to be the frontier of biomedical research. For him, it was the study of cellular mechanisms that seemed most

promising and would benefit from the use of the AEC's new tools. His enthusiasm about this particular field of research and his optimism that it would advance medicine was clear when he said, "I mean somewhere in there is the solution to the cancer problem probably."471 Gasser's particular mention of cancer implied the prevalence of cancer as a concern amongst biomedical researchers, but his much broader discussion of the study of cellular mechanisms conveyed his support for a biomedical research program that was not directed toward a particular disease. Indeed, for Gasser who believed the answer to the cancer problem could be found "somewhere" in the study of cellular mechanisms, narrowing the AEC's biomedical research to focus specifically on cancer was not a feasible option at that point.

General Manager Carroll Wilson pressed the Board of Review members on this exact issue. "I think there might be some question," Wilson explained, "as to whether a group such as this would like to recommend the support of a specific

",71 "Medical Board of Review, 1947," NARA Atlanta, Box 40, 31-32. 270

disease rather than a [generia]."472 Whether or not the Medical Board of Review

would like to recommend that the AEC focus specifically on cancer was not as

straightforward as it seemed. The issue was made more complex by an

appropriations bill awaiting congressional approval. The bill, which was referred

to as the Cancer Control bill, proposed to assign funds "not to exceed

$25,000,000" to the Commission for cancer research.473 At the time of the

Medical Board of Review meetings, it was unclear what amount might be

allocated to the AEC for cancer research. Nor was it clear whether such an

appropriation would be a one-time or, perhaps, an annual appropriation.474

Although the bill had only been passed in the House, and had not yet been

approved in the Senate, the members of the AEC and Medical Board of Review

engaged in a lengthy discussion regarding the implications of the proposed bill.

Wilson raised concerns about what expectations might arise from such an appropriation if the bill was approved. Specifically, he suggested that Congress might expect that the funds be spent directly on the control of cancer.

Presumably he meant that Congress might stipulate that this appropriation fund the provision of cancer treatment. Wilson worried that Congress might be wary of supporting general research even though he believed that cancer was probably not "susceptible to frontal attack," or that "control" might not be an option.475

Like Gasser, Wilson believed that to advance cancer research, the AEC would have to engage in a broad program of research that would hopefully yield

472 "Medical Board of Review, 1947," NARA Atlanta, Box 40, 56. 471 "Medical Board of Review, 1947," NARA Atlanta, Box 40, 57. 474 "Medical Board of Review, 1947," NARA Atlanta, Box 40, 56-57. 475 "Medical Board of Review, 1947," NARA Atlanta, Box 40, 58. 271

knowledge relevant to cancer. Wilson considered the timing of the Medical

Board of Review meetings to be fortunate in that he hoped the Board would guide

the AEC in using the proposed funds effectively. More specifically, he hoped that the Medical Board of Review would support the Commission in its ongoing support for the initiatives the AEC already had underway. For Wilson, radioisotopes distribution and the provision of funding for research and for helping to support education and training were cornerstones of the broad program of research he envisioned.

Despite Wilson's assumptions and worries about the expectations

Congress might have regarding the proposed cancer appropriation, a broad cancer program that primarily focused on providing research tools, facilitating education and training, and supporting extramural research was not unreasonable given the circumstances in which the proposed bill was drafted. That is, it was the AEC's broad authority and responsibilities in research and development that prompted the bill. As AEC Chairman David Lilienthal explained to the members of the

Medical Board of Review, the bill was introduced after Congressman Everett

Dirksen's failed attempt to secure a $100 million appropriation for cancer research the previous year. Compared to pre-war government funding for cancer research, SI 00 million was a staggering sum of money. For instance, when the

NCI was established in 1937 its first annual budget was $400,000. By 1945, government funding for cancer research had grown, but still only amounted to

$750,000.476 Given the level of government funding for cancer research in the mid-1940s, the requested $100 million appropriation was, perhaps, bound to fail.

*,7', Ross, Crusade, 37-38, 210-14. 272

Congressman Dirksen devised an alternative plan, though. He interpreted the

AEC's powers to be sufficiently broad such that the AEC could shoulder some responsibility for cancer research. Thus, he decided to attach a cancer appropriation to the Commission's budget.477

THE ADVISORY COMMITTEE FOR BIOLOGY AND MEDICINE PUTS CANCER IN ITS PLACE

Throughout 1947, when the issue of the cancer appropriation was not yet settled, the Medical Board of Review's consideration of the place of cancer research within the AEC's research agenda helped set the tone for the sort of program that evolved later. So too did the early deliberations of the AEC's

Advisory Committee for Biology and Medicine (ACBM). Even before the

ACBM's first meeting in September 1947, the Chairman of that committee, Dr.

Alan Gregg, confronted the issue. In July Gregg had corresponded with John

Derry who was the assistant to the AEC's General Manager to determine whether the membership of the ACBM would be defined with the express purpose of accommodating cancer research. Gregg and Derry considered whether to have a

"cancer man" appointed to the committee or, perhaps, even a few.478 Gregg noted that the AEC, the AEC's biomedical program that was then being established, and the ACBM all constituted a national program. As such, he thought regional representation was important to ensure that congressional members from across the United States supported the ACBM. However, membership that was strictly based on balanced regional representation would not allow for any special

477 The proposed appropriation would therefore be an addition to the funds made available to the AEC by the President's budget. "Medical Board of Review, 1947," NARA Atlanta, Box 40, 59. 47S "Letter from Alan Gregg to Mr. Derry re DBM Staffing, 15 July 1947," NARA College Park, RG 326, General Correspondence 1946-1951, Box 23, 1. 273 accommodation of cancer research because, as Gregg explained to Derry, the

nation's leading experts were disproportionately concentrated in the northeastern states.

Gregg personally believed that if Congress did receive an earmark for cancer research it might be preferable to have an ACBM member assigned to the specific concern of cancer. However, he offered what he thought to be a feasible alternative. The members of the ACBM, if chosen based on regional representation, could solicit advice from their colleagues engaged in cancer research. He suggested that, in many respects, this scenario would not be as ideal as having a so-called cancer man on the committee, but that it would help avoid the disadvantages of investing in one man an overwhelming authority to shape the

AEC's cancer work. Gregg's letter to Derry elaborated that if one man were selected to assume responsibility for defining cancer research, his institutional affiliation might influence him greatly. This was worrisome given that existing cancer research programs, Gregg believed, showed considerable differences of opinion as to how cancer research should be developed. Furthermore, a cancer man might favor his own institution relative to others when recommending the distribution of resources.479

Gregg's opinion seemed to hinge on whether or not Congress earmarked cancer research funds for the AEC. As the end of 1947 drew near, Congress did, in fact, approve an appropriation of $5 million for fiscal year 1948.480 This sum was a substantial reduction from Congressman Dirksen's first attempt to secure

479 "Letter from Gregg to Derry, 1947," NARA College Park, Box 23, 1-2. 4S0 Hewlett and Duncan, Atomic Shield, 252. 274

$100 million for cancer research and was only a fraction of the $25 million considered as a maximum amount to be earmarked for AEC cancer research.

However, it was still a significant appropriation for cancer research given the level of funding from the leading private and public cancer or public health agencies at the time. The AEC reported that by the end of 1947 the ACS had invested about $2 million dollars in cancer research and that the United States

Public Health Service (USPHS) had "assumed the responsibility for the support of cancer research to the amount of twelve and one-half million dollars."4*1 The

USPHS also distributed $1.5 million dollars to medical schools to help fund improved curriculum and infrastructure for cancer training.

The language of the AEC's report equated the USPHS' level of funding with that agency's responsibility for cancer research. Bearing that language in mind, the AEC assumed considerable responsibility for cancer research when

Congress allocated $5 million to the AEC specifically for cancer work. The

ACBM membership was established, though, without any special concentration in cancer. Rather the seven members represented a broad array of expertise in biology and medicine which they used to shape not just cancer research, but all

AEC biomedical research in the years to come. Indeed, the membership reflects the dominant attitude that characterized the AEC's cancer program during these early planning stages and beyond: the AEC would pay special attention to cancer, but cancer research would not be the exclusive purpose of the AEC's biomedical program. The AEC's commitment to a broad program stemmed, in part, from the

4si "Atomic Energy Commission: Argonne Cancer Research Hospital, 1 July 1948," NARA College Park, RG 326, General Correspondence 1946-1951, Box 64, 2. difficulty of determining a line between what was or was not cancer research.

Examined in greater detail below, this difficulty was sufficiently evident to prompt a policy change related to radioisotopes distribution.

THE FOUR-POINT CANCER PROGRAM TAKES SHAPE

The AEC first began making concrete arrangements for its cancer program at the third meeting of the ACBM in November 1947. The meeting was attended by members of the ACBM, Dr. Shields Warren who was the newly appointed

Director of the Division of Biology and Medicine (DBM), and both

Commissioners and staff of the AEC. At that point, they did not discuss any sort of comprehensive cancer program. Instead, Warren made a number of recommendations that would allow the AEC to make immediate allocations from the AEC's "Cancer Research fund."482 In particular, Warren recommended and the ACBM approved an allocation not to exceed $100,000 to convert buildings at

Brookhaven National Laboratory such that they could accommodate the staff of

Jackson Memorial Laboratories. The Jackson Memorial Laboratories were large research facilities located in Bar Harbor, Maine, that were known for cancer research and that had recently been destroyed by fire. Warren and the ACBM determined that the provision of temporary facilities was an appropriate use of the

AEC's cancer funds in that they would help facilitate the continuation of an established research program that involved cancer research. Similarly, Warren made three more recommendations to fund cancer projects. The first two—an allocation of $75,000 to support the work of the Atomic Bomb Casualty

482 "Draft Minutes: Third Meeting of the Advisory Committee for Biology and Medicine, 7 November 1947," NARA College Park, RG 326, General Correspondence 1946-1951, Box 31,9. 276

Commission in Japan and one of $100,000 to offset the cost of distributing

radioisotopes for cancer research—were both unanimously approved by the

ACBM. Warren's third recommendation that the AEC provide a sum of $500,000

to support the provision of beds for cancer patients at regional laboratories was not approved at that time. Rather the ACBM decided that the matter required

further consideration by both the ACBM and the directors of the regional

laboratories. '

All of these measures, including both those that were approved and the one that was deferred for later consideration, were initiatives that could stand alone as contributions to cancer research, specifically, and also to biomedical research, generally. The obvious characteristic that tied them together was the funds with which they were to be supported. They made immediate use of some of the AEC's budget that was earmarked for cancer. In terms of the support for the Atomic Bomb Casualty Commission and radioisotopes distribution, these initiatives also illustrate a basic objective to ensure that the AEC's cancer initiatives remained closely related to radioactive materials. Furthering the use of radioactive materials in cancer research seemed an obvious role for the AEC to play and one that would be consistent with Congress' express wish that the AEC use the funds earmarked for cancer research "without duplicating the research work of other public and private agencies."484

As the AEC worked towards establishing a formal cancer program that would clearly demonstrate how the AEC would make use of the funds allocated

4X1 "Third Meeting of the Advisory Committee, 1947," NARA College Park, Box 31,9-10. 4*4 "Argonne Cancer Research Hospital, 1948," NARA College Park, Box 64, 1. 277

for canccr, the agency started to carve out a niche in which it would, in the words

of a 1948 press release, "[expand] its assistance and support to the nationwide

48 S cancer research." " The program that the AEC devised sustained the already

established pattern of funding a number of projects or programs to expand

infrastructure that would, the Commission hoped, facilitate the use of radioactive

materials for the advancement of knowledge in biomedical fields. It also showed

the AEC's preference for partnering with other institutions and programs in a joint

effort to advancc cancer research rather than investing in cancer research pursued

strictly at its own facilities.

The ACBM decided upon the main tenets of a cancer program at its sixth

meeting in February 1948 and the AEC announced the program to the public in

March of that year. The program consisted of four aspects, none of which were

new initiatives, but rather stemmed from the projects recently approved or other elements of the AEC's biomedical program already underway. For instance, one part of the program had its origins in Warren's recommendation to establish beds

for cancer patients at regional laboratories. The ACBM authorized the provision of funds for this purpose and decided that the Oak Ridge Hospital would be the

first facility at which beds would be made available for cancer patients and related clinical research. This decision is what spurred the creation of the Oak Ridge

Institute for Nuclear Studies' (ORINS) Division of Medicine, a topic examined

more thoroughly in the previous chapter. Indeed, representatives from the

ORINS, Oak Ridge National Laboratory (ORNL), and regional medical schools

4S5 "Information for the Press: U.S. Atomic Energy Commission Announces Four-Point Program of Support of Cancer Research, 5 March 1948," NARA College Park, RG 326, General Correspondence 1946-1951, Box 64, 1. 278

convened for a two-day conference in March 1948 and started to establish plans

for a cancer research program operated by the ORINS.

A board of medical consultants and, by the end of 1948, the newly

appointed Chairman of the ORINS' Division of Medicine, Dr. Marshall Brucer,

led the development of the new cancer research program as a joint endeavor

amongst the ORNL, the ORINS, and the AEC. While these parties cooperated to

develop the Oak Ridge program which included a new cancer hospital, the

ACBM discussed and recommended similar plans for a cancer program and

hospital in conjunction with the Argonne National Laboratory, the University of

Chicago Medical School, and its associated clinics. The ACBM recommended

and the AEC approved plans for the Argonne Cancer Research Hospital in June

1948, including an initial budget of $1.75 million for construction costs.486 For

the new Argonne and Oak Ridge facilities, the AEC defined clinical research as

their main purpose and not routine care. Beds would be limited at such facilities

to no more than fifty and they would be utilized for patients participating in

studies. Cancer patients seeking established treatment options would not be

admitted.487 This policy illustrates that the Commission was especially

committed to research and training, both of which the AEC judged to be

fundamental aspects of understanding and eventually curing cancer. This was a

policy that caused some confusion amongst the general public in the years to

come and therefore required further clarification from the AEC.

4S(' The total cost of the new facility was projected at S3.5 million. "Argonne Cancer Research Hospital, 1948" NARA College Park, Box 64, 1 and 4. 4S7 "Draft Minutes: Sixth Meeting of the Advisory Committee for Biology and Medicine, 14 February 1948," NARA College Park, RG 326, General Correspondence 1946-1951, Box 31, 12- 13. 279

ORINS Cancer Hospital, c. 1950. Photo Credit: Oak Ridge Associated Universities.

Teletherapy unit developed by the ORINS Medical Division for radiation therapy, c. 1952. Photo Credit: Oak Ridge Associated Universities.

Another element of the cancer program was the subsidization of radioisotopes for cancer research. The distribution program had begun in the summer of 1946 and, as mentioned above, Warren had recommended and received authorization to make radioisotopes available for cancer research free of production costs at the ACBM's third meeting in November 1947. That said, the new policy of waiving radioisotope fees for cancer research had not yet been enacted when the ACBM discussed the cancer program at their February 1948 meeting. Administrative difficulties had prevented the fulfillment of the new policy to that point, but the ACBM determined that the administration of free radioisotopes for cancer could be arranged in time for the policy to go into effect as of April 1948. Initially, the AEC planned to distribute three radioisotopes to cancer researchers without charge: iodine-131, phosphorus-32, and sodium-24. 280

All of these had been investigated as research tools and therapeutic agents since the 1930s. Researchers were still required to pay handling and transportation fees, but the AEC estimated that depending on the type of investigation and radioisotopes requested the new policy would amount to a savings of approximately a few hundred dollars per request.488

When the AEC defined the cancer program early in 1948 the Commission committed to providing more than just these three radioisotopes free of production costs with the proviso that it would do so once it was "administratively feasible."489 As of February 1949, one year after the AEC began subsidizing the cost of radioisotopes production for cancer research, the AEC was able to extend this policy to all of the radioisotopes it produced.490 At that point, the AEC drew attention to the use of radioactive gold and cobalt in cancer research and treatment. Cobalt-60, especially, was found to be an appropriate replacement for cancer research and some therapies that had long required the use of radium. The advantage of using reactor-produced radioactive cobalt-60 was that it was available to researchers at the minimal cost of handling and shipping, whereas radium was a rare naturally occurring radioactive element that cost between

$ 15,000 and $20,000 per gram.491 When the AEC announced the extension of radioisotopes subsidization as part of its cancer policy, the Commission emphasized the fact that radioisotopes were mostly used in fundamental cancer research. In a statement for the press, the Commission explained that iodine-131

4Sii "Four-Point Program of Support of Cancer Research, 1948," NARA College Park, Box 64, 2. 4Kl) "Sixth Meeting of the Advisory Committee, 1948," NARA College Park, Box 31, 13-14. 4')0 "AEC to Make all Radioisotopes Available for Cancer Research Without Charge, 21 March 1949," NARA College Park, RG 326, General Correspondence 1946-1951, Box 64, 1-5. 4 ,1 "Recent Developments in Use of Radiocobalt, 25 November 1949," NARA College Park, RG 326, General Correspondence 1946-1951, Box 64, 1. 281 and phosphorous-32 were both used on a limited basis for diagnosis and treatment of thyroid cancers and leukemia, but that, for the most part, "effective

radiotherapy for cancer must wait upon further developments in basic cancer research."492 The press release continued to explain that radioisotopes were especially useful in basic research as tracers that could be used to determine which materials nourish cancers and make them grow.

Whereas the first two initiatives of the AEC's cancer program were meant to establish infrastructure and provide tools to enhance research and clinical training, the latter two simply helped fund cancer research that was external to the

Commission. Indeed, as the third element of the cancer program, the AEC would help fund research pursued by qualified researchers in areas of relevance to cancer. The AEC already supported cancer research at various facilities, including Memorial Hospital for Treatment of Cancer and Allied Diseases in New

York City, Harvard University, the University of Rochester, and University of

California. The AEC planned to increase its support to these institutions and others at which researchers pursued studies of the basic functions, growth, metabolism, diagnosis and possibly the treatment of cancer.493

Finally, as the fourth element of the cancer program, the AEC would help fund the research efforts taking place in Hiroshima and Nagasaki as part of the

Atomic Bomb Casualty Commission's activities. In terms of the AEC's support for the Atomic Bomb Casualty Commission, the ACBM identified what they considered the most important aspect of this work, which was to determine if

"1',~ "AEC to make all Radioisotopes Available, 1949," NARA College Park, Box 64, 3. 4" "Four-Point Program of Support of Cancer Research, 1948," NARA College Park, Box 64, 2. 282

exposure to radiation encouraged the development of cancer amongst the

survivors of the bombings. Such knowledge would be invaluable if the United

States were ever subject to a nuclear attack, but, as those within the AEC

believed, would also serve the general state of cancer research by helping

researchers gain a better understanding of the "natural history" of the disease

since the survivors were not to be treated.494

Again, both the third and fourth elements of the cancer program were a

continuation of the decisions made a few months earlier to assist the Jackson

Memorial Laboratories and the Atomic Bomb Casualty Commission.

Furthermore, they were also in keeping with the AEC's preference for

establishing joint research endeavors. Overall, the four-point program marked the

AEC's commitment to these initiatives and helped publicize the AEC's attitude

towards funding external research. As expressed in a press release regarding the

extension of radioisotopes subsidization, "By supporting the cancer research of

many physicians and scientists through institutional contracts, the Commission

encourages direction of the talents of skilled investigators toward the applications of atomic energy to the problem."495

THE AEC'S CANCER PROGRAM, PUBLICLY PERCEIVED

The AEC's cancer program reinforced other aspects of the AEC's biomedical research program, especially the radioisotopes distribution and education and training programs which the AEC considered to be of great

"Sixth Meeting of the Advisory Committee, 1948," NARA College Park, Box 31. 14; and "Four-Point Program of Support of Cancer Research, 1948," NARA College Park, Box 64, 2. 4,5 "AEC to make all Radioisotopes Available, 1948," NARA College Park, Box 64, 4. importance. The radioisotopes program, having been underway for a year and a

half, had been in existence long enough to elicit considerable praise from the

members of the ACBM when they discussed the four tenets of the cancer program. For instance, physicist Detlev W. Bronk who was also the chairman of the National Research Council stated, "there has been no other single development in the last year that has given such a boost to research in medicine and biology as the availability of radioactive isotopes."496 Bronk and his fellow

ACBM members spoke, as they did on many occasions, of the importance of educating the public on how radioisotopes were useful in cancer research.

Ensuring that the public understood why the AEC pursued each of the four elements of the cancer program was, indeed, an important aspect of executing the program. In Bronk's words, the Commission should take care to ensure that the public understand "that the chief value of isotopes in cancer or in any disease process will probably not be from the curative standpoint but rather that the advantage will come through basic study in which radioactive isotopes are used as tools and not as sources of therapeutic radiations."497

The AEC remained attentive to shaping the public's perception of the cancer program. Aside from the AEC's ongoing concern regarding the public understanding that radioisotopes were important tools for cancer research, the

AEC made a particular effort to explain how the agency's support of education and training—specifically that which would take place at new AEC-funded cancer facilities—would advance cancer research and benefit cancer patients. Perhaps

4% "Sixth Meeting of the Advisory Committee, 1948," NARA College Park, Box 31, 14. m "Sixth Meeting of the Advisory Committee, 1948," NARA College Park, Box 31, 14. 2X4

anticipating an onslaught of treatment requests from cancer patients, the AEC's

official announcement of its cancer program in 1948 explicitly acknowledged that

these facilities would not treat the average cancer patient, but would still benefit

them. That is, the clinical studies conducted at these cancer facilities would add

to the knowledge base and improve researchers' and physicians' understanding of

cancer. Accordingly, the clinical studies pursued at the new facilities would result

in the advancement of diagnostic and treatment methods.498

This was a point that the AEC had to emphasize to the public on numerous

occasions during the years that followed. The AEC, ORNL, and ORINS all

received letters from hopeful citizens who had heard optimistic reports about the

curative powers of radiation. For instance, in April 1950 Mr. A. Pelter, a 59-year-

old man from California with stomach cancer, sent a letter to the ORNL in which

he wrote, "I have read in the newspaper, that your plant will soon give treatments

for cancer."499 He explained his illness stating that "The pains that I have in my stomach, head and body I cannot describe, living only a hopeless life.. . .Perhaps

by taking the atomic treatments, I can be helped and able to work. Please let me

know where to go."500 Similarly, a woman from Pennsylvania by the name of

Mrs. M. S. Moon wrote on behalf of her 55-year-old husband who had recently been diagnosed with prostate cancer. In a letter sent to Oak Ridge, also in April

1950, Mrs. Moon expressed that she was "very anxious" to know if the work being done at Oak Ridge could help cure her husband of his disease. Her faith in

4W "Four-Point Program of Support of Cancer Research, 1948," NARA College Park, Box 64, 3. 4

manage the ever increasing number of letters the AEC received, the Division of

Biology drafted a form letter that could be sent out as a response. In it, the message the AEC delivered was that the ORINS' cancer program offered

"unusual opportunities for the very thorough study of possible treatments for a few specific diseases," but that the program had neither the facilities nor the

budget to operate a full scale diagnostic and treatment center.502 In response to physicians who sought to refer their patients, the AEC sent the form letter as well as a personal letter.

The ORINS' cancer program did treat patients referred by physicians outside of the cancer program, but the patients treated were very carefully chosen

based on specific diagnoses. As would be the case with the Argonne Cancer

Research Hospital once it opened in 1953, the ORINS cancer program depended on referrals from physicians at associated medical schools as a means for ensuring that each patient selected for the program suffered from a form of cancer closely related to current studies.503 At any given time, the cancer program would have a couple or perhaps a few small-scale clinical studies underway. The referral policy illuminated the AEC's inclination to establish cooperative projects that involved other institutions, in this case the ORINS which itself involved numerous

M" "Letter from Mrs. M. S. Moon to Cancer Research Department, Oak Ridge, Tennessee, 21 April 1950," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945-1952, Box 33, 1. 5,12 "Letter from Marshall Brucer to Honorable George Smithers, 10 April 1950," NARA Atlanta, RG 326, New York Operations Office - Research & Medicine Division Correspondence 1945- 1952, Box 33, 1. M)1 "Letter from Brucer to Smithers, 1950," NARA Atlanta, Box 33, 1-2. 286 southeastern universities and medical schools. It also served the political function of establishing alliances with institutions in different states—states beyond those in which AEC facilities were located.

The referral policy reflected the AEC's commitment to focusing almost entirely on research, not cancer therapy. For the most part, the radiation therapy provided to patients through the ORINS' cancer program was not routine radiation therapy. Rather, the therapeutic procedures provided were experimental—they had been tested on animal subjects and were being developed for humans. One exception to this rule was that a limited number of beds were reserved for training purposes and were used to train physicians in what were, at that point, established therapeutic uses of radioactive iodine and phosphorous.

These elements were used as therapeutic agents to treat forms of cancer as well as other illnesses.504 The use of a few beds for this purpose shows that the radioisotopes program, education and training, and the cancer program blended together. Furthermore, it suggests that the AEC's commitment to each was part of an overall commitment to expanding and improving biomedical manpower, tools, and infrastructure as a means to advance knowledge and practice.

FULFILLING THE AEC'S CANCER GOALS & BROAD BIOMEDICAL AGENDA

At the end of 1951, three years after the start of the cancer program, the

AEC's role in cancer research was still very much defined by the recommendations that the Medical Board of Review and the ACBM made in 1947

MW "Letter from Brucer to Smithers, 1950," NARA Atlanta, Box 33, 1 -2. 287 and 1948.505 The AEC felt that it had a special responsibility to support cancer research to contribute to public welfare. This was a responsibility that the AEC's

Acting Chairman Henry D. Smyth considered as great as the responsibility of building bombs, and one that the AEC could fulfill both by making use of the technologies and expertise located within AEC facilities and by providing support to other institutions. Smyth expressed his opinion to the Chairman of the ACBM,

Dr. Ernest Goodpasture. "Certainly the study and treatment of cancer is one of the most direct contributions that we can make to mankind," Smyth wrote. "It is, therefore, the intention of the Commission to continue to support research in the diagnosis and treatment of cancer."506 The Commission's ongoing support for the cancer program allowed for the continuation of the four-point program with only slight modifications in future years.

Perhaps the most notable modification was a change in policy regarding the free distribution of radioisotopes for cancer research in 1955. The AEC had already revised the subsidization policy for cancer research in 1952 by instituting a minimal charge—20% of the catalogue price—for radioisotopes used in cancer research. The policy change aimed to discourage wasteful uses of short-lived radioisotopes.507 At that point, the subsidization of radioisotopes was still limited to cancer research, but biomedical researchers had begun to press for change.

Representatives of the Heart Association wrote to the AEC to protest the

M15 "Letter from Ernest Goodpasture to H. D. Smyth, 1 December 1951," NARA College Park, RG 326, General Correspondence 1951-1958, Box 67, Organization and Management 7 - Division of Biology and Medicine, Advisory Committee, Vol. I, 1-2; and "Letter from H. D. Smyth to Ernest Goodpasture, 28 December 1951," NARA College Park, RG 326, General Correspondence 1951- 1958, Box 67, 1-2. M)h "Letter from Smyth to Goodpasture, 1951," NARA College Park, Box 67, 2. 51)7 "Proposed Subsidy of Radioisotopes Program, 9 March 1955," NARA College Park, RG 326, General Correspondence 1951-1958, Box 28, 1. 288 subsidization policy. Such letters expressed opposition to the policy on account

that radioisotopes were not provided free of charge for heart research.

The policy was further criticized during hearings on the medical uses of atomic energy held by the Subcommittee on Research and Development of the

Joint Committee on Atomic Energy (JCAE) in June 1954. Speaking during the hearings, Dr. Nicholas Werthessen "forcefully advocated" on behalf of heart researchers who sought free radioisotopes for their work.508 The JCAE later reported that they found Werthessen's testimony to be very persuasive. The members of the JCAE were especially convinced by Werthessen's description of cancer research pursued in his own laboratory. The research did not return significant results related to cancer, but rather yielded valuable information regarding the hardening of arteries. The unexpected results were certainly an asset to the development of biomedical knowledge, but they altered the nature of this research such that it was no longer eligible for the cancer subsidization. The cost of radioisotopes for Werthessen's research therefore increased fivefold.509

Werthessen warned that the subsidization policy that was applicable only to cancer research might deter researchers from investigating unexpected results— even results that seemed likely to benefit medical research and practice - if to do so resulted in a dramatic increase in expenses. Similarly, Werthessen noted that research programs not directly related to cancer could yield results that might

"Subsidy of Radioisotopes Program, 1955," NARA College Park, Box 28, 4. MW Joint Committee on Atomic Energy, "Report of the Research and Development Subcommittee: Providing Radioisotopes for Medical Research, 16 August 1954," NARA College Park, RG 326, General Correspondence 1951-1958, Box 28, 1-2. 289 advance cancer research and therefore should not be excluded from free radioisotopes distribution.510

At the conclusion of the hearings the JCAE recommended to the AEC that the AEC extend the radioisotopes subsidization policy to all medical research.

The Division of Biology and Medicine agreed and presented a report to the AEC in which the DBM proposed a new policy that subsidized radioisotope production at 50% of the cost for all biomedical and some agricultural research.5" The AEC approved the new policy which went into effect in 1955. The modification of the radioisotopes distribution policy that extended radioisotopes subsidization to cover all biomedical research rather than just cancer research was not evidence of a shift in attitude. The AEC remained as committed to cancer research as it had been in 1947 and 1948 when the AEC launched the four-point program. Rather, the revised policy reflected the reality that the categorization of research as cancer research or otherwise was often difficult to determine at the outset. Overall, the new policy which aimed to facilitate the use of radioisotopes in all biomedical research was very much rooted in the belief that the expansion of biomedical research, generally speaking, would advance cancer research.

CONCLUSION

The AEC cancer program was motivated by factors both internal and external to the Commission. Regarding the former, the Commission controlled access to reactor-produced radioisotopes and other technologies that many

510 Joint Committee on Atomic Energy, "Radioisotopes for Medical Research, 1954," NARA College Park, Box 28, 2-3. 511 "Subsidy of Radioisotopes Program, 1955," NARA College Park, Box 28, 1-4. 290

believed could be of considerable use in cancer research. Also, the AEC both

employed and consulted with biomedical researchers who had experience using

radiation in cancer research. Beyond these, external influences, especially the

congressional allotment to fund cancer research within the AEC, also encouraged

the AEC's involvement in cancer research. Shaped by these various factors, the

AEC's cancer program reflected the desire to both advance cancer research and to develop radiation as a tool for research and diagnostic and therapeutic practices.

The program also allowed the AEC to put forth a very positive image of

its research and development to the public. As is evident in the numerous comments from AEC Commissioners and members of the ACBM and DBM throughout the years, the AEC considered cancer research to be a highly valuable contribution to mankind. The AEC was careful, though, to publicize its cancer program as being very much focused on advancing knowledge. It was clear that the AEC was not interested in making promises to cure cancer quickly, even though the public relations benefits would be considerable. Instead, the AEC made a concerted effort to ensure the public that support of research would help improve researchers' and physicians' understanding of cancer and, therefore, improve diagnostic and therapeutic practices.

The AEC's cancer program sheds light on the social and political context in which research developed following the war—a social and political context characterized by government involvement in science, the authority of researchers to command resources and influence public policy, and the increasingly prominent role of science, biomedical sciences included—in domestic and 291 national security affairs. It was a time at which those researchers and administrators acting within or able to connect to the AEC bureaucracy could combine various factors to create a program focused on disease, technology, and national security. The AEC's cancer program helped to define what it was to be a nation entering into the Cold War. Whether investing in the health or security of the nation, the government supported research for the advancement of knowledge and technological development. 292

CONCLUSION

What does the history of biomedical radiation research in the United

States reveal about the relationship between science and society? How did the outbreak of World War II affect the development of biomedical radiation research? This study has examined the emergence of biomedical radiation

research at the start of the twentieth century and the development of this research

through to the early Cold War. Throughout this period, variously trained

researchers investigated and used radiation both as a research tool and research object to advance knowledge within many disciplines. These studies also served as the foundation for new disciplines that expanded the biomedical enterprise and strengthened the links between science and medicine. Biomedical radiation

research not only linked science and medicine, it linked science and the state.

Indeed, during World War II, the development of biomedical radiation research

became part of a larger government- and military-funded endeavor to develop atomic bombs. The wartime mobilization of science resulted in new technologies and the advancement of knowledge. It also initiated profound changes within the

political economy of research, including biomedical radiation research.

The history of biomedical radiation research before, during, and following

World War II shows that research developed within and influenced a social and

political context that incorporated factors both internal and external to the

scientific and medical research communities. Biomedical radiation researchers gained access to resources and institutional support based on the relationship between biomedical research and other fields of research, and also between biomedical research and the social and political significance of that research.

These were the key relationships researchers sought to manage in their efforts to build both research programs and disciplines. To do so biomedical researchers defined the advancement of knowledge, particularly knowledge of human biology, illness, and the biological effects of radiation as socially and politically important. They established for themselves the responsibility of developing medical applications of radiation, both diagnostic and therapeutic in nature, and determining radiation safety policies and procedures that were especially important within the context of nuclear weapons development, testing, and the possibility of nuclear warfare. In both medical and military contexts, researchers had to market their research to acquire necessary resources to further their investigations.

The mobilization of science for World War II was a tremendously significant episode in the history of the American scientific enterprise, biomedical sciences included. The military goals that drove wartime research and development, as well as the vast resources provided to researchers, constituted an extraordinary change in the social and political context in which researchers worked. Biomedical researchers, among others, pursued research within the changing circumstances during the war and continued to do so following the war.

They did not, however, simply strive toward achieving government and military goals. Rather, biomedical researchers sought to play a role in defining research policies and programs for the state. They sought to create a new political economy of research in which biomedical radiation research was a national 294

priority, one organized and funded by the government. As a result, biomedical

radiation researchers, like so many other researchers, became part of an ever

larger postwar scientific enterprise that was a central characteristic of the Cold

War. The boundaries between science, government, civilian, and military research were blurred as biomedical researchers partnered with or became part of government agencies such as the Atomic Energy Commission (AEC) in an effort to redefine or create research programs that expanded the biomedical enterprise.

RADIATION & EARLY TWENTIETH-CENTURY RESEARCH ACROSS THE DISCIPLINES

While this dissertation has focused largely on biomedical radiation research during the early AEC era, the prewar period is an important part of this history. The investigation of radiation as a subject of research and as a tool to be used in research and clinical applications throughout the first few decades of the twentieth century established a foundation for new biomedical disciplines such as biophysics, radiobiology, and health physics. The gradual emergence of new biomedical disciplines was influenced by a number of factors including, but not limited to, the discovery of radiation, access to sources of radiation, the availability of institutional or external research funding, and educational reforms that affected both science and medicine. This study has argued that the collaboration amongst physical scientists, life scientists, and physician- researchers to investigate radiation and its uses was also central to the process of building research programs and disciplines. The interdisciplinary collaboration that developed amongst variously trained scientists and physicians, such as biophysicist Gioacchino Failla and physicians Henry H. Janeway and Benjamin S. 295

Barringcr at Memorial Hospital in New York, was a defining feature of biomedical radiation research. In fact, the practice of interdisciplinary research helped create new areas of expertise that were hybrid in nature. It was the practice of interdisciplinary research and resulting hybrid expertise out of which biomedical radiation researchers emerged as a new kind of researcher that pioneered new biomedical fields.

Throughout the process of investigating radiation and integrating it as a tool to be used in research, researchers were influenced by factors such as access to radiation and funding. The context in which researchers pursued their studies was also very much related to social factors like concern for cancer. Indeed, the collaboration between physical scientists and life scientists and physician- researchers resulted from their shared intellectual interest in radiation, but also from an underlying effort to pursue research that was socially useful and, as a result, likely to attract funding. In part, researchers responded to a social need for cancer research. Researchers also played a role in establishing the demand for cancer research through their efforts to advance research and obtain the resources they needed to do so. That is, they defined cancer as a problem for which their research was a solution. This was evident early in the twentieth century and continued to be the case following World War II. After the war, researchers helped define the need for cancer research not only through the AEC's cancer program, but also through their efforts to establish other elements of the AEC's biomedical program. According to researchers, they needed radioisotopes and more researchers trained to work with them to pursue cancer research. 296

A NETWORK OF RESEARCHERS GOES TO WAR

The practice of interdisciplinary radiation research did more than help

establish new areas of biomedical expertise and establish links between radiation

research and the social uses of that research. Those engaged in collaborative

work built professional connections that crossed both disciplinary and

institutional boundaries. The transfer of technologies, such as cyclotrons, as well

as researchers from one institution to another created connections amongst

researchers who constituted an informal professional network. Within this

network, researchers could derive authority or define their expertise because the growing commitment to biomedical radiation research reinforced the work of individual researchers.

This professional network was more readily apparent once the Manhattan

Project commenced and biomedical radiation researchers were recruited for health and safety work. Indeed, the existing connections amongst researchers, disciplines, and institutions were reflected in the pattern of recruitment of individuals to work within the Health District at the University of Chicago's

Metallurgical Laboratory. The biomedical researchers who were recruited for war work were a diversely trained group of researchers, but they all possessed expertise related to radiation—expertise that was hybrid in nature and that stemmed from their pre-war interdisciplinary collaboration. During the war, the

Manhattan Project's biomedical researchers were able to extend their authority beyond the scientific and medical communities. Individuals such as Dr. Robert S.

Stone, Dr. Hymer L. Friedell, and Dr. Stafford L. Warren, all of whom were radiologists and had participated in interdisciplinary radiation research prior to the 297

war, assumed positions of authority within the Manhattan Engineer District

(MED).

Biomedical radiation researchers played an important role in the

Manhattan Project in that they were recruited to determine the health hazards that

would result from all aspects of atomic bomb development and ensure the safety of all those involved in the project. The literature on this topic has predominantly focused on how the MED attempted to protect the health and safety of its workers and whether or not the MED's efforts were effective.5u These are obvious and

important questions, yet they have been peripheral to this study.513 Rather, the focus here has been on how and why biomedical radiation research evolved as scientific knowledge developed and in changing social and political circumstances. The creation of a wartime partnership amongst biomedical radiation researchers and the government and military resulted in a reorganization of the political economy of research—a reorganization that preserved some existing trends common within biomedical radiation research, but that also introduced lasting changes.

The most significant outcome of the wartime mobilization of biomedical radiation researchers was the reinforcement and expansion of professional and institutional connections. Through these strengthened and expanded connections,

Sec, for instance. Hacker, The Dragon's Tail. 1 The same is true for the postwar A EC era. That is, historians and other scholars have investigated the government's role in human radiation experimentation and questioned whether or not the government maintained ethical standards for biomedical experiments involving humans. It is important to note that the issue of human radiation experimentation bears some relevance to the reorganization of the political economy of research in the postwar years, but that this study has not intended to engage in the debate regarding human radiation experimentation. For the most comprehensive overview of this issue see, ACHRE, The Final Report. This issue is also addressed in Kutcher, Cancer Research and the Military, Martensen, "Medical Physics at the University of California"; Moreno, Undue Risk. 298

researchers were able to transform biomedical expertise into entrepreneurial and administrative expertise. That is, biomedical researchers were often able to

translate their wartime responsibilities into opportunities to develop new research

programs or acquire resources for the research they deemed appropriate to the task of determining and protecting against radiation hazards.

This is evident in the history of biomedical radiation research at the

University of Rochester from the 1930s through the war. Radiologist Stafford L.

Warren acquired a particular kind of expertise from his collaboration with physicists, such as Lee A. DuBridge and their research with technologies such as the cyclotron. He was also well known to biomedical radiation researchers at other institutions, such as Ernest Lawrence's Rad Lab at the University of

California, Berkeley. All of these factors influenced his recruitment to the

Manhattan Project. Once in the position of Chief of the Manhattan Engineer

District's Medical Office, Warren was able to establish a large biomedical research program at the University of Rochester that brought more resources, including new technologies and researchers, to that university. The wartime

Rochester Atomic Energy Project was a large enough enterprise that it required the construction of new research facilities on campus. Thus, the development of biomedical radiation research at the University of Rochester from the pre-war to wartime period shows that personal and institutional connections made prior to the war helped facilitate the creation of a much larger biomedical radiation research enterprise at that university during the war. 299

BIOMEDICAL RESEARCH WITHIN THE ATOMIC ENERGY COMMISSION

During the postwar years biomedical radiation research achieved greater

institutional support through new departments and laboratories in universities and

hospitals and also within government agencies. Indeed, the creation of the AEC's

biomedical bureaucracy was a central feature of the growing field of biomedical

radiation research. The AEC's Division of Biology and Medicine (DBM) and

Advisory Committee on Biology and Medicine (ACBM) were especially integral organizational structures through which the relationship between science and the state was defined. Researchers directly associated with the DBM and ACBM, as well as those external to the AEC who communicated and collaborated with the

AEC, served as an important link between science and the state. They helped

identify or establish what would serve as other links between the scientific community and the government: radioisotopes, training initiatives, and cancer.

It was within the DBM and ACBM, in particular, and the AEC as a whole, that biomedical researchers and government officials created policies and programs informed by various scientific, technical, social, political, and economic factors. The AEC's biomedical program focused on providing radioisotopes to researchers which, as discussed throughout this study, allowed for many more researchers in many more institutions to apply radioisotopes in a broad range of research. The AEC also made a concerted effort to provide training that was complementary to the radioisotopes program. The Commission, with encouragement and often institutional cooperation from the scientific community, established programs that helped increase the number of researchers able to work 300

with radioisotopes and other sources of radiation. Finally, the AEC established cancer research as a priority, but one that was not pursued at the expense of basic

biomedical research.

Radioisotopes, training, and cancer were initiatives that were sufficiently broad to benefit both the scientific community and government. They were common ground upon which the scientific community and government defined shared goals, such as the advancement of biomedical knowledge for medical and safety purposes. The greater availability of radioisotopes as research tools, the increased number of scientists and physicians trained to work with radiation, and the pursuit of cancer research were all aspects of postwar research that were relevant to the ongoing development of atomic weapons too.

THE RELATIONSHIP BETWEEN BIOMEDICAL RADIATION RESEARCH & THE COLD WAR

The late 1940s and early 1950s was a transitional period throughout which scientists, the government, and military sought to establish some permanency in the partnership these groups had established during the war. Granted, as the history of the AEC's biomedical research shows, some of these groups, especially scientists, defined their goals much more broadly than the defense goals that dominated wartime research. The AEC was a key arena in which varying goals were pursued and the dichotomy between civilian and defense research was tenuously defined. The AEC's biomedical policies and programs reveal an effort on behalf of both researchers and government officials to establish civilian research that was separate from defense research. While the AEC's biomedical research certainly contributed to a civilian agenda of advancing biomedical 301

knowledge and medical practice, it cannot, however, be entirely separated from

the AEC's defense priorities. For instance, the radioisotopes program made use

of a reactor originally built as a prototype for reactors later built at Hanford that

would produce plutonium for bombs. Furthennore, the research made possible by

the distribution of radioisotopes was an important means through which medical

applications of radiation could be developed, but also provided the tools to

researchers who helped procure information for the military related to hazards

like radioactive fallout.

AEC-funded education and training programs, such as the University of

Rochester's graduate program in Radiation Biology, created a jagged line between civilian and military spheres. This program increased the number of

researchers able to conduct biomedical research and investigate medical uses of

radiation. However, the University of Rochester and other universities also worked in collaboration with the AEC to provide training for military personnel.

Military researchers were trained so that they could participate in research and help devise safety measures to prevent radiation exposure. Military doctors were educated such that they could understand and attend to health consequences of radiation exposure.

Even the AEC's cancer program blended civilian and defense research.

Radiation researchers had long studied the relationship between radiation exposure and cancer. They had also investigated and developed possible diagnostic and therapeutic uses of radiation in relation to cancer. During the war and following, these lines of inquiry and development acquired military 302

importance in that cancer was an industrial hazard at weapons production

facilities, a consequence of the bombings in Japan, and a threat associated with

the potential of atomic warfare in the future.

The point here is that the tools and goals associated with biomedical

radiation research cannot be easily categorized as either civilian or defensive in

nature. Also, the AEC's biomedical radiation program was both influenced by

and helped shape the political economy of research in the postwar period. Indeed,

the AEC's biomedical research was part of a system of research and development

in which the scientific community, government, and military defined goals that shaped the overall enterprise. More broadly, the AEC's biomedical radiation research represented fundamental characteristics of postwar science in the United

States. These included the federal government's and military's commitment to playing a role in research and development, as well as the scientific and medical communities' willingness to work with the government and military to achieve certain ends. These groups cooperated to create new institutional arrangements through which each group hoped to have their needs met. In this context, biomedical research—even that explicitly focused on advancing biological and medical knowledge and developing medical applications of radiation—helped constitute the Cold War. That is, such research was part of a shift that drove the ongoing partnership between science and the state following the war.

THIS STUDY & BEYOND

Overall, the development of biomedical radiation research throughout the whole historical period examined here shows that science is not isolated from society. Science is shaped by and is very much a component of society. Thus,

the examination of how and why fields of research develop in relation to social

and political factors is paramount, regardless of the historical period. When

viewed within particular social and political contexts we see that the role of

biomedical radiation researchers and individual initiatives, sometimes aimed

toward singular goals of acquiring funding or access to materials, amounted to

policies and programs that affected society at large. Indeed, the history of the

AEC's biomedical radiation research shows that researchers and government

administrators pursued various scientific and technical goals in a rather piecemeal

fashion, but that their efforts created a biomedical bureaucracy and established

other links that tied together the interests of the scientific community and the

government. In so doing, biomedical researchers marshaled resources used to

advance the state of biomedical knowledge and, ultimately medical practice.

They also, however, helped define the Cold War as a war that the United States

would fight by investing in research and development and expanding government

bureaucracy.

Perhaps the most striking outcome of the biomedical radiation research

that the AEC very much helped to expand is the prevalence of radioisotopes in

medicine today. The use of radioisotopes is well illustrated by the use of one

radioisotope alone, technetium-99 (Tc-99m), which is used in 60-70% of all

medical diagnostic tests in the United States.514 The AEC provided research

materials, training infrastructure, and an agenda that were integral to the development and growth of fields of biomedical radiation research and nuclear

Dominic Ryan, "Medical Isotopes - Sidebar," Physics in Canada 66, no. 1 (2010): 8. 304

medicine in the second half of the twentieth century. Although many researchers

who played a part in the creation of the AEC's biomedical programs felt confident

that the AEC's funding and infrastructure would result in important discoveries

related to medical uses of radiation, there was, of course, no guarantee that

radioisotopes would become so central to clinical practice. In an interview in the early 1990s Arthur M. Weis, who was a nuclear engineer, made exactly this point.

He said, "I would never have thought 30 years ago that nuclear medicine would

become a daily diagnostic procedure in every hospital."515 Weis made this statement when reflecting on his transition from the aerospace to nuclear medicine industry. He explained that at the time—around 1960—he and his colleagues had no strategic plan that would foreshadow the significance of nuclear medicine in the future. Although Weis believed that neither he nor his colleagues had a strategic plan to develop nuclear medicine, this dissertation has demonstrated that scientists who were entrepreneurial and attentive to aligning new scientific discoveries with social and political factors, helped create an environment in which biomedical radiation research flourished. They continued what, prior to World War II, had been a process of discipline-building within academic contexts, but did so on a national scale.

Had Weis or others known for certain that radioisotopes and their role in nuclear medicine would, indeed, be so significant, likely few would have imagined that the United States would play so small a role in producing them.

For decades, the National Research Universal (NRU) reactor at Canada's Chalk

5I> N.A., "Interview of Arthur M. Weis in 'People in Nuclear Medicine: Interviews with Physicians, Scientists and Industry Leaders,'" The Journal of Nuclear Medicine 34, no. 6 (1993): 30N. River Laboratory has been a leading producer of medical radioisotopes.516

Belgium, the Netherlands, South Africa, and France operate the four other main

reactors used to produce the world's supply of medical radioisotopes.517 The

United States' decline as the leading producer of medical radioisotopes and the

rise of these other nations is a shift that is yet to be examined in the literature. As

this study has shown, this historical problem is one that should be examined within the social and political context of the Cold War. Indeed, the United States'

role in producing and distributing radioisotopes at the dawn of the atomic age and eventual withdrawal from this function was very much influenced by the United

States' development of nuclear weapons and the position of scientists within the

A EC's bureaucracy.

Like the early postwar period in the United States, controversy over recent shutdowns of Canada's NRU reactor in November 2007 and May 2009 also illustrates that the development of science, medicine, and associated infrastructure in Canada is very much wrapped up in a complex system involving various interests. While the supply of medical radioisotopes was the dominant issue that captivated worldwide observers following these shutdowns, researchers tried to emphasize that the reactor also supports research in numerous fields. In an issue of the Canadian Association of Physicists' Physics in Canada dedicated exclusively to the medical radioisotopes crisis, physicist Dominic Ryan stated,

>lh Bcla Joos, "The Saga ofNRU, the Supply of Medical Isotopes, and the Future of Neutron Scattering in Canada," Physics in Canada 66, no. 1 (2010): 1; Dominic Ryan, "Medical Isotopes and the Future of Neutron Scattering in Canada," Physics in Canada 66, no. 1 (2010): 5. 517 As reported by CBC News, planned shutdowns for each of the reactors used to produce medical radioisotopes in these countries exacerbated the crisis. See CBC News, "Global Supply Under Pressure," 23 December 2010, (accessed 12 April 2011). 306

NRU is not, and never has been, just an isotope reactor. Nor is it just a development platform for [Atomic Energy of Canada Ltd.]. NRU was designed and built as a major piece of research infrastructure that has supported Canadian industry for over fifty years.518

As is evident in the title given to this issue, "The Medical Isotope Crisis and the

Broader Implications for Research," the Canadian physics community was engaged in a process aimed at working with the Canadian government to deal with the radioisotopes crisis. In fact, they sought to ensure that reactor shutdowns were perceived as more than a radioisotopes crisis because, to them, the fate of the NRU reactor was a much larger problem that affected a broad range of research. While Canadian scientists endeavored to influence decisions made about the NRU reactor, they were, on a fundamental level, trying to manage relationships within the political economy of research to ensure ongoing access to research materials and infrastructure they deemed necessary for their work.

Ryan, "Future of Neutron Scattering," 6. 307

BIBLIOGRAPHY

Archives:

In writing this dissertation I have made extensive use of the Records of the Atomic Energy Commission (record group 326) held at the National Archives and Record Administration (NARA). Specifically, I analyzed records maintained at NARA II in College Park, Maryland and the Southeast Regional Archives in Atlanta, Georgia. The record series that were most relevant to my research included the "General Correspondence 1946-1951" and "General Correspondence 1951-1958" series (both at NARA II), and the "New York Operations Office - Research & Medicine Division Correspondence, 1945-1952" series (Southeast Regional Archives).

Oral Histories:

I have also made use of oral history transcripts. Most of these oral histories were conducted by the Advisory Committee on Human Radiation Experiments (ACHRE) in the mid 1990s. The ACHRE interviewed numerous radiation researchers involved in research from the early twentieth century through the Cold War period. This oral history project resulted in transcripts of interviews with many of the nation's leading radiation researchers all of which are available on the Department of Energy's website at "DOE Openness: Human Radiation Experiments," .

In addition to the ACHRE transcripts, two other interview transcripts were particularly useful for my research:

Oral History of Lee DuBridge by Charles Weiner, interview conducted June 9, 1972, Niels Bohr Library & Archives, American Institute of Physics, College Park, MD, USA, .

Oral History of Hans Bethe, Frederick Reines, Robert Christy, and J. Carson Mark, by Stanley Goldberg, interview conducted August 18, 1989, as part of the Smithsonian Videohistory Program (18), session 14, "The Manhattan Project: Collection Division 4: Los Alamos."

Primary Sources:

Advisory Committee on Human Radiation Experiments. The Final Report. New York: Oxford University Press, 1996. 308

Brophy, Leo P., and George J. B. Fisher. The Chemical Warfare Service: Organizing for War. Washington, DC: Center of Military History, U.S. Army, 2004. First published 1959 by the Center of Military History.

Compton, Arthur Holly. Atomic Quest: A Personal Narrative. New York: Oxford University Press, 1956.

Evans, Robley D. "The Medical Uses of Atomic Energy." The Atlantic Monthly. Vol. 177, no. 1 (1946): 68-73.

. "Radium Poisoning A Review of Present Knowledge." American Journal of Public Health and the Nation's Health. Vol. 23, no. 10 (1933): 1017- 1023.

Failla, Gioacchino. "Ionization Measurements." American Journal of Roentgenology and Radium Therapy 10 (1923): 48-56.

. "Dosage in Radium Therapy." American Journal of Roentgenology. Vol. 8, no. 11 (1921): 674-685.

. "The Absorption of Radium Radiation by Tissues." American Journal of Roentgenology. Vol. 8, no. 5 (1920): 2-15.

Failla, Gioacchino, and Edith H. Quimby. "The Economics of Dosimetry in Radiotherapy." American Journal of Roentgenology and Radium Therapy 10(1923): 944-967.

Friedell, Hymer L., L. D. Marinelli, and Titus C. Evans. "Gioacchino Failla, 1891-1961." Radiation Research. Vol. 16, no. 5 (1962): 619-625.

Green, Harold P., and Alan Rosenthal. Government of the Atom: The Integration of Powers. New York: Atherton Press, 1963.

. "Fusion of Government Power." Bulletin of the Atomic Scientists. Vol. 18, no. 6 (1962): 12-16.

Hahn, Paul F., William F. Bale, Ernest O. Lawrence, and George H. Whipple. "Radioactive Iron and Its Metabolism in Anemia: Its Absorption, Transportation, and Utilization." Journal of Experimental Medicine. Vol. 69, no. 5 (1939): 739-753. 309

Hamilton, J. G., and R. S. Stone. "Excretion of Radio-Sodium Following Intravenous in Man," Proceedings of the Society for Experimental Biology and Medicine 35 (1937): 595-598.

Janeway, Henry H., Benjamin S. Barringer, and Gioacchino Failla. Radium Therapy in Cancer at the Memorial Hospital: First Report, 1915-16. New York, 1917.

Jones, Malcolm D., and Glenn E. Sheline. "Robert S. Stone, Radiology: San Francisco." In University of California: In Memoriam, 1968. Berkeley, CA: University of California Academic Senate, 1968.

Joos, Bela. "The Saga of NRU, the Supply of Medical Isotopes, and the Future of Neutron Scattering in Canada." Physics in Canada. Vol. 66, no. 1 (2010): 1-2.

Millikan, R. A. "The Significance of Radium." Science. Vol. 54, no. 1383 (1921): 1-8.

N.A. Biographical Memoirs. Washington, DC: National Academy of Sciences, N.D.

N.A. "Availability of Radioactive Isotopes: Announcement from Headquarters, Manhattan Project, Washington, D.C." Science, 14 (June 1946): 697-705.

N.A. "People in Nuclear Medicine: Interviews with Physicians, Scientists and Industry Leaders." The Journal of Nuclear Medicine. Vol. 34, no. 6 (1993): 20, 29-30, 32, 34, 36, 40, 42-43.

N.A. "Science: A Year of Isotopes." Time Magazine. August 1947, 74.

Niehoff, Richard O. "Organization and Administration of the United States Atomic Energy Commission." Public Administration Review. Vol. 8, no. 2 (Spring 1948): 91-102.

ORINS, The Executive Committee. "A Nuclear Research Institute at Oak Ridge." Science. Vol. 103, no. 2685 (1946): 705-06.

Oughterson, Ashley W., and Shields Warren, eds. Medical Effects of the Atomic Bomb in Japan. New York: McGraw-Hill Book Company, 1956. 310

Pearsc, Jr., Herman E„ and Stafford L. Warren. "The Roentgenograph Visualization of the Arteries of the Extremities in Peripheral Vascular Disease." Annals of Surgery. Vol. 94, no. 6 (1931): 1094-1102.

Quimby, Edith H. "Gioacchino Failla (1891-1961) and the Development of Radiation Biophysics." The Journal of Nuclear Medicine. Vol. 6, no. 5 (1965): 376-382.

Regato, Juan A. del. Radiological Physicists. New York: American Institute of Physics, 1985.

Ryan, Dominic. "Medical Isotopes - Sidebar." Physics in Canada. Vol. 66, no. 1 (2010): 8-9.

. "Medical Isotopes and the Future of Neutron Scattering in Canada." Physics in Canada. Vol. 66, no. 1 (2010): 5-7.

Smyth, Henry De Wolf Atomic Energy for Military Purposes: The Official Report on the Development of the Atomic Bomb under the Auspices of the United States Government, 1940-1945. Stanford, CA: Stanford University Press, 1989.

. A General Account of the Development of Methods of Using Atomic Energy for Military Purposes under the Auspices of the United States Government, 1940-1945. Washington, DC: United States Government Printing Office, 1945.

Stannard, J. Newell, ed. J. Newell Stannard and the University of Rochester: A Collection of Papers Presented at the 48th Annual Meeting of the Health Physics Society. San Diego, CA: Health Physics Society, 2003.

Stewart, Irvin. Organizing Scientific Research for War: The Administrative History of the Office of Scientific Research and Development. Boston, MA: Little, Brown and Company, 1948.

Stone, Robert S. "The Concept of a Maximum Permissible Exposure." Radiology. Vol. 51, no. 5 (1952): 639-661.

True, Frederick W. A History of the First Half Century of the National Academy of Sciences, 1863-1913. Washington, DC: National Academy Press, 1913. 311

Ungar Jr., John, and Shields Warren. "Skin Grafting as a Method of Determining the Biologic Effect of Radiation." Archives of Pathology 23 (1937): 299- 306.

Warren, S. L., and G. H. Whipple. "Roentgen Ray Intoxication: Intestinal Lesions and Acute Intoxication Produced by Radiation in a Variety of Animals." Journal of Experimental Medicine. Vol. 38, no. 6 (1923): 741-752.

. "Roentgen Ray Intoxication: The Path of a Beam of Hard Rays in the Living Organism." Journal of Experimental Medicine, Vol. 38, no. 6 (1923): 731-739.

Warren, Shields. "The Distribution of Doses of Radioactive Phosphorus in Leukemic Patients." Cancer Research 3 (1943): 334-336.

Warren, Shields, and R. F. Cowing. "The Distribution of Doses of Radioactive Phosphorus in Rodents." Journal of Laboratory and Clinical Medicine 26 (1941): 1014-1016.

Warren, Shields, and Olive Gates. "Radiation Pneumonitis: Experimental and Pathologic Observations." Archives of Pathology 30 (1940): 440-460.

Warren, Stafford L., Francis W. Bishop, Harold C. Hodge, and G. Van Huysen. "A Quantitative Method for Studying Roentgen-Ray Absorption of Tooth Slabs." American Journal of Roentgenology and Radium Therapy 31 (1934): 663.

Secondary Sources:

Avery, Donald H. The Science of War: Canadian Scientists and Allied Military Technology During the Second World War. Toronto: University of Toronto Press, 1998.

Ball, Howard. Justice Downwind: America's Atomic Testing Program in the 1950s. New York: Oxford University Press, 1986.

Balogh, Brian. Chain Reaction: Expert Debate and Public Participation in American Commercial Nuclear Power, J945-1975. Cambridge, UK: Cambridge University Press, 1990. 312

Bcatty, John. "Ecology and Evolutionary Biology in the War and Postwar Years: Comments and Questions." Journal of the History of Biology. Vol. 21, no. 2 (1988): 245-263.

Berlin, Nathaniel I. and Louis R. Wasserman. "Polycythemia Vera: A Retrospective and Reprise." Journal of Laboratory and Clinical Medicine. Vol. 30, no. 4(1997): 365-373.

Bird, Kai and Martin J. Sherwin. American Prometheus: The Triumph and Tragedy of J. Robert Oppenheimer. New York: Alfred A. Knopf, 2000.

Bocking, Stephen. "Ecosystems, Ecologists, and the Atom: Environmental Research at Oak Ridge National Laboratory." Journal of the History of Biology 28 (1995): 1-47.

Buck, Alice L. A History of the Atomic Energy Committee. Washington, DC: U.S. Department of Energy, 1983.

Bourdieu, Pierre. "The Specificity of the Scientific Field and the Social Conditions of the Progress of Reason." Social Science Information. Vol. 14, no. 6(1975): 19-47.

Brues, A. M. "Shields Warren: (1898-1980)." Radiation Research. Vol. 88, no. 2 (1981): 430-435.

Buderi, Robert. Engines of Tomorrow: How the World's Best Companies Are Using Their Research Labs to Win the Future. New York: Simon & Schuster, 2000.

. The Invention That Changed the World: How a Small Group of Radar Pioneers Won the Second World War and Launched a Technological Revolution, The Sloan Technology Series. New York: Simon & Schuster, 1996.

Chernus, Ira. Eisenhower's Atoms for Peace. College Station, TX: Texas A&M University Press, 2002.

Garfield, Gerhard H. and William M. Wiecek. Nuclear America: Military and Civilian Nuclear Power in the United States, 1940-1980. New York: Harper & Row, 1984.

Clark, Claudia. Radium Girls: Women and Industrial Heath Reform, 1910-1935. Chapel Hill, NC: University of North Carolina Press, 1997. 313

Cochrane, Rexmond C. The National Academy of Sciences: The First Hundred Years, 1863-1963. Washington, DC: The National Academy of Sciences, 1978.

Cooter, Roger, Mark Harrison and Steve Sturdy, eds. Medicine and Modern Warfare. Atlanta, GA: Rodopi, 1999.

eds. War, Medicine and Modernity. Stroud: Sutton Publishers, 1998.

Creager, Angela N. H. "Phosphorous-32 in the Phage Group: Radioisotopes as Historical Tracers of Molecular Biology." Studies in the History and Philosophy of Biological and Biomedical Sciences 40 (2009): 29-42.

"Nuclear Medicine in the Service of Biomedicine: The U.S. Atomic Energy Commission's Radioisotope Program, 1946-1950." Journal of the History of Biology 39 (2006): 649-684.

. "Tracing the Politics of Changing Postwar Research Practices: The Export of'American' Radioisotopes to European Biologists." Studies in the History and Philosophy of Biological and Biomedical Sciences 33 (2002): 367-388.

. "Wendell Stanley's Dream of a Free-Standing Biochemistry Department at the University of California, Berkeley." Journal of the History of Biology. Vol. 29, no. 3 (1996): 331-360.

Creager, Angela N. H., and Maria Jesus Santesmases. "Radiobiology in the Atomic Age: Changing Research Practices and Policies in Comparative Perspective." Journal of the History of Biology 39 (2006): 637-47.

Dennis, Michael A. '"Our First Line of Defense:' Two University Laboratories in the Postwar American State." Isis. Vol. 85, no. 3 (1994): 427-455.

England, J. Merton. A Patron for Pure Science: The National Science Foundation's Formative Years, 1945-1957. Washington, DC: National Scicnce Foundation, 1983.

Feffer, Stuart M. "Atoms, Cancer, and Politics: Supporting Atomic Scicnce at the University of Chicago, 1944-1950." Studies in the History and Philosophy of Biological and Biomedical Sciences. Vol. 22, no. 2 (1992): 233-261.

Gaddis, John Lewis. The Cold War: A New History. New York: Penguin Books, 2005. 314

Galison, Peter, and Bruce Hevly, eds. Big Science: The Growth of Large-Scale Research. Stanford, CA: Stanford University Press, 1992.

Gardner, Kirsten E. Early Detection: Women, Cancer, and Awareness Campaigns in the Twentieth-Century United States. Chapel Hill, NC: University of North Carolina Press, 2006.

Gaudilliere, Jean-Paul. "Normal Pathways: Controlling Isotopes and Building Biomedical Research in Postwar France." Journal of the History of Biology 39 (2006): 737-764.

. "Molecular Biologists, Biochemists, and Messenger RNA: The Birth of a Scientific Network." Journal of the History of Biology. Vol. 29, no. 3 (1996): 417-445.

Geiger, Roger L. To Advance Knowledge: The Growth of American Research Universities, 1900-1940. New Brunswick, NJ: Transaction Publishers, 2004.

. Research and Relevant Knowledge: American Research Universities since World War II. New York: Oxford University Press, 1993.

Goldberg, Stanley. "Inventing a Climate of Opinion: Vannevar Bush and the Decision to Build the Bomb." Isis. Vol. 83, no. 3 (1992): 429-452.

Goldman, David E. "Kenneth S. Cole, 1900-1984." Biophysical Journal 47 (1985): 859-860.

Goodman, Jordan, Anthony McElligott, and Lara Marks, eds. Useful Bodies: Humans in the Service of Medical Science in the Twentieth Century. Baltimore, MD: Johns Hopkins University Press, 2003.

Grandin, Karl and Nina Wormbs, and Sven Widmalm, eds. The Science-Industry Nexus: History, Policy, Implications. Sagamore Beach, MA: Science History Publications/USA, 2004.

Hacker, Barton C. Elements of Controversy: The Atomic Energy Commission and Radiation Safety in Nuclear Weapons Testing, 1947-1974. Berkeley, CA: University of California Press, 1994.

. The Dragon's Tail: Radiation Safety in the Manhattan Project, 1942- 1946. Berkeley, CA: University of California Press, 1987. 315

Hannaway, Caroline, ed. Biomedicine in the 20th Century: Practices, Policies, and Politics. Amsterdam, The Netherlands: IOS Press, 2008.

Hannaway, Owen. "Laboratory Design and the Aim of Science: Andreas Libavius versus Tycho Brahe." Isis. Vol. 77, no. 4 (Dec. 1986): 584-610.

Harrison, Mark. "The Medicalization of War—the Militarization of Medicine." Social History of Medicine. Vol. 9, no. 2 (1996): 267-276.

Hayter, Charles. An Element of Hope: Radium and the Response to Cancer in Canada, 1900-1940. Montreal, QC: McGill-Queen's University Press, 2005.

Heilbron, John, and Robert Seidel. Lawrence and His Laboratory: A History of the Lawrence Berkeley Laboratory. Berkeley, CA: University of California Press, 1990.

Herken, Greg. The Winning Weapon: The Atomic Bomb in the Cold War, 1945- 1950. New York: Alfred A. Knopf, 1980.

Hershberg, James G. James B. Conant: Harvard to Hiroshima and the Making of the Nuclear Age. Stanford, CA: Stanford University Press, 1993.

Hevly, Bruce. "Stanford's Supervoltage X-Ray Tube." Osiris 9 (1994): 85-100.

Hewlett, Richard G., and Oscar E. Anderson. The New World: A History of the United States Atomic Energy Commission, Volume 1, 1939-1946. University Park, PA: Pennsylvania State University Press, 1962.

Hewlett, Richard G., and Francis Duncan. Atomic Shield: A History of the United States Atomic Energy Commission, Volume 2, 1947-1952. University Park, PA: Pennsylvania State University Press, 1969.

Hewlett, Richard G., and Jack M. Holl. Atoms for Peace and War: Eisenhower and the Atomic Energy Commission, A History of the United States Atomic Energy Commission, Volume 3, 1953-1961. Berkeley, CA: University of California Press, 1989.

Hoddcson, Lillian. Critical Assembly: A Technical History of Los Alamos During the Oppenheimer Years, 1943-1945. New York: Cambridge University Press, 1993. 316

Hodge, Harold C., and Carl Voegtlin, eds. Pharmacology and Toxicology of Uranium Compounds, National Nuclear Energy Series. New York: McGraw-Hill Book Co., 1949-1953.

Holloway, David. Stalin and the Bomb: The Soviet Union and Atomic Energy, 1939-1956. New Haven, CT: Yale University Press, 1994.

Hounshell, David A., and John Kenly Smith, Jr. Science and Corporate Strategy: Du Pont R&D, 1902-1980. Cambridge, UK: Cambridge University Press, 1988.

Hughes, Jeff. The Manhattan Project: Big Science and the Atom Bomb. New York: Columbia University Press, 2002.

Hughes, Thomas P. Rescuing Prometheus: Four Monumental Projects That Changed the Modern World. New York: Vintage Books, 1998.

. American Genesis: A Century of Invention and Technological Enthusiasm, 1870-1970. Chicago, IL: University of Chicago Press, 1989.

. Networks of Power: Electrification in Western Society, 1880-1930. Baltimore, MD: The Johns Hopkins University Press, 1983.

Johnson, Leland, and Daniel Schaffer. Oak Ridge National Laboratory: The First Fifty Years. Knoxville, TN: University of Tennessee Press, 1994.

Josephson, Paul R. Red Atom: Russia's Nuclear Power Program from Stalin to Today. New York: W. H. Freeman and Company, 2000.

Keller, Evelyn Fox. "Physics and the Emergence of Molecular Biology: A History of Cognitive and Political Synergy." Journal of the History of Biology. Vol. 23, no. 3 (1990): 389-409.

Kevles, Daniel J. The Physicists: The History of a Scientific Community in Modern America. Cambridge, MA: Harvard University Press, 1995.

. "The National Science Foundation and the Debate over Postwar Research Policy, 1944-1945: A Political Interpretation of Science—the Endless Frontier." Isis. Vol. 68, no. 1 (1977): 4-26.

Kirsch, Scott. "Harold Knapp and the Geography of Normal Controversy: Radioiodine in the Historical Environment." Osiris 19 (2004): 167-181. 317

Kocher, D.C. "Perspective on the Historical Development of Radiation Standards." Health Physics. Vol. 61, no. 4 (1994): 519-527.

Kohler, Robert E. Partners in Science: Foundations and Natural Scientists, 1900- 1945. Chicago, IL: University of Chicago Press, 1991.

. "Science, Foundations, and American Universities in the 1920s." Osiris 3 (1987): 135-164.

. "Philanthropy and Science." Proceedings of the American Philosophical Society. Vol. 129, no. 1 (1985): 9-13.

. From Medical Chemistry to Biochemistry: The Making of a Biomedical Discipline. New York: Cambridge University Press, 1982.

Kraft, Alison. "Between Medicine and Industry: Medical Physics and the Rise of the Radioisotope, 1945-65." Contemporary British History. Vol. 20, no. 1 (2006): 1-35.

Kraft, Michael E. and Norman J. Vig, eds. Technology and Politics. Durham, NC: Duke University Press, 1988.

Kuhn, Thomas S. "Historical Structure of Scientific Discovery." Science. Vol. 136, no. 3518 (1962): 760-764.

Kutcher, Gerald. Contested Medicine: Cancer Research and the Military. Chicago, IL: University of Chicago Press, 2009.

. "Cancer Therapy and Military Cold-War Research: Crossing Epistemological and Ethical Boundaries." History Workshop Journal 56 (2003): 105-30.

Lederer, Susan E. Subjected to Science: Human Experimentation in America before the Second World War. Baltimore, MD: Johns Hopkins University Press, 1995.

Lenoir, Timothy. Instituting Science: The Cultural Production of Scientific Disciplines. Stanford, CA: Stanford University Press, 1997.

Lerner, Barron H. The Breast Cancer Wars: Fear, Hope, and the Pursuit of a Cure in Twentieth-Centuiy America. New York: Oxford University Press, 2001. 318

Leslie, Stuart W. The Cold War and American Science: The Military-lndustrial- Academic Complex at MIT and Stanford. New York: Columbia University Press, 1993.

Lindee, M. Susan. "Atonement: Understanding the No-Treatment Policy of the Atomic Bomb Casualty Commission." Bulletin of the History of Medicine 68 (1994): 454-490.

. Suffering Made Real: American Science and the Survivors at Hiroshima. Chicago, IL: University of Chicago Press, 1994.

Lowy, liana. Preventive Strikes: Women, Precancer, and Prophylactic Surgery. Baltimore, MD: Johns Hopkins University Press, 2010.

Marples, David R. Chernobyl and Nuclear Power in the USSR. Edmonton, AB: Canadian Institute of Ukrainian Studies, 1986.

Mendelsohn, Everett, et al. eds. Science, Technology and the Military, Volumes 1 and 2. Boston, MA: Kluwer Academic Publishers, 1988.

Moreno, Jonathan D. Undue Risk: Secret State Experiments on Humans. New York: W.H. Freeman, 2000.

Mukherjee, Siddartha. The Emperor of all Maladies: Biography of Cancer. London, UK: Harper Collins, 2010.

Needell, Allan A. "Nuclear Reactors and the Founding of Brookhaven National Laboratory." Historical Studies in the Physical Sciences. Vol. 14, no. 1 (1983): 93-122.

Norris, Robert S. Racing for the Bomb: General Leslie R. Groves, the Manhattan Project's Indispensable Man. South Royalton, VT: Steerforth Press, 2002.

Patterson, James T. The Dread Disease: Cancer and Modern American Culture. Cambridge, MA: Harvard University Press, 1987.

Rader, Karen A. "Alexander Hollaender's Postwar Vision for Biology: Oak Ridge and Beyond "Journal of the History of Biology 39 (2006): 685-706.

Rasmussen, Nicolas. "The Mid-Century Biophysics Bubble: Hiroshima and the Biological Revolution in America, Revisited." History of'Science 35 (1997): 245-293. 319

Reagan, Leslie J. "Engendering the Dread Disease: Women, Men, and Cancer." American Journal of Public Health. Vol. 87, no. 11 (November 1997): 1779-1787.

Reich, Leonard S. The Making of American Industrial Research: Science and Business at GE and Bell, 1876-1926. Cambridge, UK: Cambridge University Press, 1985.

Reingold, Nathan. Science, American Style. New Brunswick, NJ: Rutgers University Press, 1991.

Rhodes, Richard. Arsenals of Folly: The Making of the Nuclear Arms Race. New York: Simon and Schuster, 2008.

. Dark Sun: The Making of the Hydrogen Bomb. New York: Simon and Schuster, 1995.

. The Making of the Atomic Bomb. New York: Simon and Schuster, 1986.

Robison, Roger. "American Radium Engenders Telecurie Therapy During World War I." Medical Physics. Vol. 27, no. 6 (2000): 1212-1216.

Ross, Walter Sanford. Crusade: The Official History of the American Cancer Society. New York: Arbor House, 1987.

Rothman, David J. Strangers at the Bedside: A History of How Law and Bioethics Transformed Medical Decision Making. New York: Aldine de Gruyter, 1991.

Rowland, R.E. Radium in Humans: A Review of U.S. Studies. Argonne, IL: Argonne National Laboratory, 1994.

Santesmases, Maria Jesus. "Peace Propaganda and Biomedical Experimentation: Influential Uses of Radioisotopes in Endocrinology and Molecular Genetics in Spain (1947-1971)." Journal of the History of Biology 39 (2006): 765-794.

Serwer, Daniel Paul. "The Rise of Radiation Protection: Science, Technology and Medicine in Society, 1896-1935." Ph.D. diss., Princeton University, Princeton, NJ, 1976.

Sloan, Phillip R., ed. Controlling Our Destinies: Historical, Ethical, and Theological Perspectives on the Human Genome Project. Notre Dame, IN: University of Notre Dame Press, 2000. 320

Smith, Robert W. The Space Telescope: A Study of NASA, Science, Technology, and Politics. Cambridge, UK: Cambridge University Press, 1989.

Smith, Robert W. and Helge Krah. "Who Discovered the Expanding Universe?" History of Science 41 (2003): 141-62.

Smith, Susan L. Japanese American Midwives: Culture, Community, and Health Politics, 1880-1950. Champaign, IL: University of Illinois Press, 2005.

Strickland, Stephen P. Politics, Science, and Dread Disease: A Short History of United States Medical Research Policy. Cambridge, MA: Harvard University Press, 1972.

Taylor, Peter J. "Technocratic Optimism, H. T. Odum, and the Partial Transformation of Ecological Metaphor after World War II." Journal of the History of Biology. Vol. 21, no. 2 (1988): 213-244.

Walker, J. Samuel. Permissible Dose: A History of Radiation Protection in the Twentieth Century. Berkeley, CA: University of California Press, 2000.

. Containing the Atom: Nuclear Regulation in a Changing Environment, 1963-1971. Berkeley, CA: University of California Press, 1992.

and George T. Mazuzan. Controlling the Atom: The Beginning of Nuclear Regulation, 1946-62. Berkeley, CA: University of California Press, 1985.

Wang, Jessica. American Science in an Age of Anxiety: Scientists, Anticommunism, and the Cold War. Chapel Hill, NC: University of North Carolina Press, 1999.

. "Liberals, the Progressive Left, and the Political Economy of Postwar American Science: The National Science Foundation Debate Revisited." Historical Studies in the Physical and Biological Sciences. Vol. 26, no. 1 (1995): 139-166.

Weisz, George. Divide and Conquer: A Comparative History of Medical Specialization. New York: Oxford University Press, 2006.

Welsome, Eileen. The Plutonium Files: America's Secret Medical Experiments in the Cold War. New York: Dial Press, 1999.

Westwick, Peter J. The National Labs: Science in an American System, 1947- 1974. Cambridge, MA: Harvard University Press, 2003. 321

Whittemore, Gilbert F. "The National Committee on Radiation Protection, 1928- 1960: From Professional Guidelines to Government Regulation." Ph.D. diss., Harvard University, Cambridge, MA, 1983.

Williams, Jeffrey E. "Donner Laboratory: The Birthplace of Nuclear Medicine." The Journal of Nuclear Medicine. Vol.40, no. 1 (1999): 16, 18,20.

Wilson, John T. Academic Science, Higher Education, and the Federal Government, 1950-1983. Chicago, IL: University of Chicago Press, 1983.

Wise, George. Willis R. Whitney, General Electric, and the Origins of U.S. Industrial Research. New York: Columbia University Press, 1985.

Zachery, G. Pascal. Endless Frontier: Vannevar Bush, Engineer of the American Century. New York: The Free Press, 1997.

Zimmerman, Joseph F. "Regulating Atomic Energy in the American Federal System." Puhlius. Vol. 18, no. 2 (Summer 1988): 51-65.