<<

Under the

Physics, Engineering, and the Distortion of a World War Two Legacy

A thesis presented

by

Raphael Chayim Rosen

to

The Department of the History of Science in partial fulfillment for an honors degree in History and Science Cambridge, Massachusetts March 2006

Under the Radar

Physics, Engineering, and the Distortion of a World War Two Legacy

Raphael Chayim Rosen

Abstract

This thesis examines the historical legacy of radar development during World War II. At the M.I.T. Radiation Laboratory in 1940-1941, attempted to construct a working microwave radar set. These physicists heavily relied upon engineering skills: focusing on enhancing the efficiency and efficacy of specific components of radar technology. In the postwar era, as a result of the new celebrity of physicists, their wartime inventions, including microwave radar, became associated with physics itself, not engineering. The community of engineers lacked the authority and recognition necessary to reclaim a share of the credit for their discipline. Physicists never emphasized engineering’s importance, because they did not see it as containing independent forms of knowledge or creativity and so did not believe it had played a key role in their inventiveness. This thesis concludes with a brief investigation of engineering’s endogenous forms of knowledge and their relation to the Radiation Laboratory.

Keywords

M.I.T. Radiation Laboratory Microwave Radar World War II Engineering History of Physics Public Opinion Epistemology

Contents

Acknowledgments…………………………………………………………………..5

Introduction………………………………………………………………………… 7

One: Engineering Ascendant…………………………………………………….. 14 Irradiating the Heavens: Radar Physics in November 1940…………. 21 Targets Acquired: Radar Components in November 1940…………... 24 In “The Hour of Peril and Need”: Rad Lab Engineering…………….. 34 Physics Research at the Rad Lab: Complication and Resolution…..... 48

Two: Engineering Descendant…………………………………………………… 54 Withering Laurels: Engineering’s Immediate Postwar Reception…… 55 Popular Opinion and the Indomitable Postwar Physics……………… 66 The Dark Shadow: Engineering in Disrepute………………………... 71 The Adulterated Rad Lab Legacy……………………………………. 80

Three: Engineering Divergent……………………………………………………... 92 versus Engineer at the Rad Lab and Beyond………………. 94 Scientific Potentate, Engineering Vassal…………………………….. 100 Towards an Epistemology of Engineering…………………………… 108

Conclusion…………………………………………………………………………. 118

Appendix: Key Radar Terms Explained…………………………………………… 120

Bibliography……………………………………………………………………….. 121

Acknowledgements

“We make a living by what we get. We make a life by what we give.” -

Shawn Mullet, thank you for your invariable ebullience, your indefatigable support, and your ready willingness to assist. Your breadth of knowledge on the relevant history and guidance through it kept me on track, and, more importantly off innumerable fruitless and tortuous ones. Even brief discussions with you could transform inchoate ideas into concrete, cogent arguments. Thank you truly. Professor Sarah Jansen, for your robust confidence, your warmhearted excitement, and your insistence upon pinpointing the heart of the argument, I am deeply appreciative. Thank you for pressing me to always consider the consequential questions and to make them known. Most of all, thank you for stressing the importance of enjoying the process; you helped keep my head level. Professor Steven Shapin, thank you for helping me paddle through the unfamiliar bog of innovation studies and the literature on the epistemology of engineering. Professor Cathryn Carson and Professor David Henkin, your thoughtful recommendations were acutely helpful in clarifying my thoughts. To Professor Charles Rosenberg, Professor Daniel Kevles, Professor , Professor Gerald Holton, thank you all for sharing your insights and stories and advice with me. To Rebecca Press Schwartz, thank you for a copy of your intriguing study. Professor Steven Peter Rosen, I sincerely thank you for your open ear and candor. Our animated discussions of radar history, military secrecy, and the research process are some of my fondest memories of the thesis writing process. Thank you to the Olin foundation for your generous travel grant. Thank you to Matthew Lazen and the Harvard College Research Program for your financial support as well. Frederic Burchsted, for being superman and guiding me to terrific sources even while suffering from the flu, I am extremely grateful. Chapter two would simply not exist without your help. Your mind is the encyclopedia of encyclopedias, and my project would have perished many times over without your help. To all the archivists and librarians that assisted me, thank you for your patience and thought. Thank you Joan Gearin of the National Archives for your prompt assistance and cheerful recommendations. To Nora Murphy and Jeffrey Mifflin at M.I.T, to David Farrell and David Kessler at UC Berkeley, and to Margaret Kimball at Stanford, your assistance is also heartily appreciated. Jean-Francois Gauvin, thank you once again for your lucid criticism. This thesis was far less considered and careful without your assistance. Your tutelage over the past three years has been the brightest spot of my time studying the history of science. Peter Buck, Ben Rapoport, and especially Allie Belser, thank you for your many little recommendations. Stef Tung, many thanks for being my authority on postmodern “authors.” Sharrona Pearl, Jacob Aptekar, and Andrea Maxwell: without your help, this thesis, like the credit owed to its subject, would have been terribly muddled. My friends, thank you for your patience and understanding. Jeanette LGW, for your equanimity and compassion and love, I am and will forever be unspeakably grateful. To my family, the constancy of your love and support have been the greatest blessings I know. To the Rosen family, southwest region, thank you Michael and Debbie for your love and your humor. Eytan and Danya, thank you for being the stars of my desktop and bringing a smile to my face every morning. To Gavri and Jesse, thank you for your love and your concern. I simply could not have made the writing clean without your thorough vacuuming of it. To my Bubbie and Zaydie for always supporting me and believing in me, thank you. Zaydie, I’ve always loved your many radar and stories; maybe one day we’ll find those old manuals after all. Bubbie, your tremendous love and zeal for your family have always inspired me. Ema and Abba, for awakening in me, more than anyone else, an appreciation of the infinite blessings this life has to offer; and for breeding within me an insatiable love of truth and a sedulous work ethic, I am extremely thankful.

The help you all gave has brought this thesis to life.

This thesis is dedicated with sempiternal gratitude to all the physicists and engineers (and the methods they used) who defeated the U-boat and assisted the landings at Normandy so that the Nazi tyranny might be put to rest.

Introduction

“And yet, through the gloom and the The fate of a nation was riding that night.” -Henry Wadsworth Longfellow in “Paul Revere’s Ride”1

On January 4, 1941, history repeated itself.

Nearly two centuries earlier, Paul Revere had stood on the on bank of the Charles

River, awaiting a signal from the opposite shore in Boston. Legend has it that when two lanterns flickered in Boston’s Old North Church, Revere sped through Middlesex County sounding the alert of a British advance. That night, the nation’s survival rested in the hands of a silversmith and his steed.

Then, in 1941, atop building six of the Massachusetts Institute of Technology, a team of physicists peered across the , also anticipating a signal. On the screen of the first operational American microwave radar system, a faint arc appeared.2 It represented the dome of Boston’s Christian Science Mother Church. Their radar set had seen across the river.3 In 1941, as in 1775, the faced a grave threat from

1 Henry Wadsworth Longfellow, “Paul Revere’s Ride,” in The Poetical Works of Longfellow, ed. George Monteiro (Boston: Houghton Mifflin Company, 1975), 208, lines 77-78. 2 Technically, a microwave radar system had been up and running in Tuxedo Park, New York, in the summer of 1940, but this system used the Doppler method of detection, not the pulse method which would predominate during the war. , Tuxedo Park: A Wall Street Tycoon and the Secret Palace of Science that Changed the Course of World War II, (New York: Simon & Schuster, 2002): 177-178. 3 Coordination Committee Series 1, Box 49a, interview with Alvarez and Box 49b, interview with Van Voorhis. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. In 1998, a New York Times article would confuse this radar success with another one on February 7, 1941. Interviewing former Rad Lab physicists, the article identified the first building spotted on the radar as the Christian Science Temple, but then averred that the Rad Lab cabled the message, “We have seen Mary Baker Eddy with one eye.” Philip Hilts “Last Rites for a Plywood Palace That Was a Rock of Science,” New York Times, march 31, 1998, pg. F4. Mary Baker Eddy is the founder of Christian Science, but the expression ‘one eye’ refers to using a single system, which the Rad Lab physicists did not achieve in crudest form until January formidable and militant forces.4 For America and her allies, a great hope lay in a system that could detect enemy planes. That hope was realized in microwave radar.

Radar—Radio Detection And Ranging—bounces short radio waves off a plane or and detects the reflection of that signal; based on the time it takes to receive the reflection and the direction from which it returns, an object’s location can be pinpointed.5

Mist, haze, darkness, and fog do not impede the accuracy of these waves. Of the different kinds of radar, microwave radar, consisting of waves only ten centimeters in length, can detect planes and with much greater precision than longer-wave radar.

Microwave radar’s importance is not easily overstated. The aphorism of World

War II would maintain that the nuclear bomb ended the war, but microwave radar won it.6 As Nobel laureate explained his decision not to work on the nuclear bomb to his friend Robert Oppenheimer, “I’m very serious about this war. We could lose

10, 1941, at which point they cabled M.I.T. Radiation Laboratory director Lee Alvin DuBridge in Washington to let him know that they had “succeeded with one eye.” Conant, 218. 4 American physicists, as compared to the population in general, were acutely concerned with events unfolding inside . By 1941, numerous American physicists had already taken their European refugee colleagues into their own campuses and homes. Their colleagues testified to the fact that one could only conduct science in Germany if you were not Jewish, and a country that hampered the free practice of its physicists especially alarmed scientists. John. S. Rigden and I. I. Rabi, “Introduction” in Radar in World War II, (Los Angeles: American Institute of Physics/Tomash Publishers, 1987), xix. See also: Ernest C. Pollard, Radiation: One Story of the M.I.T. Radiation Laboratory (Durham, N.C.: The Woodburn Press, 1982), 2-3. 5 Technical details of radar will only be considered as they are necessary to understand the kind of work done. For an electromagnetic and mechanical comprehension of radar itself try first and foremost, the 28 volumes on microwave radar techniques published by McGraw-Hill in 1947. For example, on radar in navigation see: John Hall, ed. Radar Aids to Navigation (New York: McGraw-Hill Book Company Inc., 1947). For details on theory see especially Nathan Marcuvitz, Waveguide Handbook (: Peter Peregrinus Ltd., 1986). For radar signals see Charles Cook and Marvin Bernfeld, Radar Signals: An Introduction to Theory and Application (New York: Academic Press, 1967). For a hybrid treatment of both radar history and technique see S. S. Swords, Technical History of the Beginnings of Radar, (London: Peter Peregrinus Ltd., 1986), which is particularly concerned with British systems but does not treat advanced microwave radar. The word radar itself matches its own form: a pulse is emitted, then a pulse is returned, just as the word itself is a palindrome. 6 This phrase has become so common that it is ubiquitous in the World War II radar literature, but it is sometimes attributed personally to M.I.T. Radiation Laboratory director Lee DuBridge. See: Daniel Kevles, The Physicists (New York: Vintage Books, 1979), 308. it with insufficient radar.”7 By war’s end, the radar industry had outgrown the pre-war automobile industry, and the cost of the armed forces’ radar equipment by July 1945 totaled $3 billion—far ahead of the $2.2 billion .8

This thesis will concern itself with the confused legacy of microwave radar.9 For the last six decades, histories have predominantly construed the development of microwave radar during World War II as an achievement of physics. The M.I.T.

Radiation Laboratory, or the Rad Lab, as it came to be known, was the heart of this nationwide radar enterprise, the place that “brought all existing ideas to bear on each new” radar challenge, and physicists dominated its ranks.10 As such, their methodology has received the credit for its achievements. Closer investigation, however, reveals that the techniques of engineering played a critical role in birthing microwave radar as well.

Physics is the investigation of nature’s most fundamental laws. Engineering, though sharing knowledge and methods with physics, is nevertheless fundamentally

7 A Walk through the 20th Century with Bill Moyers; Episode 15, I. I. Rabi: Man of the century, VHS, created and developed by the Corporation for Entertainment and Learning Inc., and Bill Moyers (Washington, D.C.: Public Broadcasting Service, 1988). See also: James Gleick, “Columbia Lauds Rabi As Its “Brilliant Jewel,” New York Times, November 21, 1985, pg. B1; Richard Rhodes, The Making of the Atomic Bomb (New York: Simon & Schuster, 1986), 452. 8 The comparison with the automobile industry is taken from Lee A. DuBridge, “History and Activities of the Radiation Laboratory of the Massachusetts Institute of Technology,” Review of Scientific Instruments 17, no. 1 (January 1946): 3. The cost of radar is from: James Phinney Baxter, 3rd, Scientists Against Time (Boston: Little, Brown and Company, 1946), 142. A full half of the total cost was for equipment purchased based on engineering research conducted in the M.I.T Radiation Laboratory. The cost of the Manhattan Project taken from: and others, Critical Assembly: A Technical History of Los Alamos during the Oppenheimer Years, 1943-1945 (Cambridge, England: Cambridge University Press, 1993), 406. , after touring the nuclear research centers in the United States during that war, remarked to that the making of the nuclear bomb, “couldn’t be done without turning the whole country into a factory.” Rhodes, 500. Radar similarly consumed the energies of huge parts of the country, both in industry and research manpower. 9 Wall Street tycoon and amateur physicist Alfred Loomis, who oversaw the M.I.T. Radiation Laboratory, remarked to the Rad Lab’s official historian during the war, “The study of the history of science often should concern itself with the history of the financial backing and the patronage of science,” but nothing else. Loomis told Guerlac that his job was “not to interpret or to try to prove anything.” My apologies, Alfred. Coordination Committee Series 1, Box 49a, interview with Alvarez M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 10 On the centrality of the Radiation Laboratory to radar development see: DuBridge, “History,” 5. concerned with the creation of machines (we will reexamine these definitions later).11

While the physics underlying wave transmission greatly aided the understanding of radar, and while a confluence of engineering and physics techniques helped produced it, it is engineering know-how that, above all else, proved crucial at the Radiation Laboratory.

And yet, the importance of these engineering skills has almost entirely unrecognized since they were employed. This thesis revisits the Rad Lab legacy and explores how and why engineering was erased from it.

Chapter one investigates the early research conducted at the M.I.T. Rad Lab, and demonstrates that engineering skills predominated. The organization of the laboratory as well as the physicists’ notebooks themselves reveal that the Rad Lab’s primary focus lay with the improvement of already existing pieces of technology. This is a task of engineering, not of physics. The practice of engineering therefore deserves a great deal of credit for birthing the radar sets that defeated the German submarine and secured beachheads at Normandy and Iwo Jima.12

Chapter two explores the way in which, after the war ended, credit for the Rad

Lab’s achievements was distorted. Though engineering skills initially received praise in the mainstream press for radar, physics skills were soon regarded as more important. This distortion of credit resulted from that fact that in the postwar era, physicists, who had designed both radar and the nuclear bomb, gained enormous recognition and authority, and, in the public spotlight, the practice of physics became associated with the physicists’

11E. A. Guillemin, World War II era M.I.T. professor of electrical communications, laconically phrased it, “Physics deals with investigation into laws of nature, engineering with their application to the benefit of society.” “Longhairs and Short Waves,” Fortune, November 1945, 169. See also: Walter Vincenti, “Control-Volume Analysis: A Difference in Thinking between Engineering and Physics,” Technology and Culture 23, no. 2 (April 1982): 146. 12 Engineering also created weapons that complicated the war and enhanced its brutality in the first place. Such is the concern of a different thesis, however, though this issue will crop up here as well. wartime inventions. The community of American engineers, meanwhile, lacked both the

prestige and the visibility that physicists boasted, and thus could not reclaim their

discipline’s radar legacy.

The final chapter will illuminate the epistemological issues underlying this

misallocation of credit. By retelling the Rad Lab story with a focus on the relationships

between physicists and engineers, I will argue that the reason physicists did not associate

their wartime ingenuity with engineering while in the public spotlight, was because they

viewed it as a discipline unequal to their own. The philosophy of the physics community

regarded engineering as manufacturing. While physicists admitted that their radar work

involved engineering, they imputed their creative breakthroughs to their physics

background, believing engineering to be completely devoid of knowledge. Here the

definition of engineering as simply the construction of artifacts will be re-examined. Only

by comprehending engineering’s endogenous modes of knowledge can the discipline be

appreciated as its own creative source. Only then, can its role in the be

fully understood.

Much prior historical work has dealt with American science in the second World

War.13 Many historians have examined the military’s employment of civilian scientists in

the development of radar and other technologies.14 Others have analyzed the

13 The literature is immense, some of the most famous and important include: Baxter; Conant; Louis Brown, A Radar History of World War II: Technical and Military Imperatives (Bristol, UK: Institute of Physics Publishing, 1999); Robert Buderi, The Invention that Changed the World (New York: Simon & Schuster, 1996); Henry Guerlac, Radar in World War II (Los Angeles: American Institute of Physics/Tomash Publishers, 1987); Hoddeson et. al.; Rhodes; Tom Shachtman, Terrors and Marvels: How Science and Technology Changed the Character and Outcome of World War II (New York: William Morrow, 2002). 14Previous American wartime ventures typically featured a laboratory administered by the military and working directly for the military. In the 1930’s, the British had developed radar using laboratories run largely by civilian scientists. American Edward Bowles noted that in dealing with the military, when one proposed a new idea the administrative brass would “open a file cabinet and show you a piece of paper.” transformation of the relationship between state and science as a result of the predominance of government funding for research; still others have paid particular attention to the wartime integration of physics and engineering.15 Finally, historians of

World War II have devoted much attention to the nuclear bomb and its postwar legacy.16

This thesis will move away from this existing literature, discussing the bomb and government funding only in the context of the American public’s postwar opinions about physics.17 It will focus instead on clarifying engineering’s role in radar, and how public opinion and the internal philosophy of the physics community transformed the legacy of that role.

The significance of this historical misattribution of credit extends well beyond

1945. Physicists’ and engineers’ contributions to the World War II effort fundamentally and permanently altered the relationship between the nation and the science and engineering communities.18 The defense and welfare of nations now depend intimately

The Rad Lab would end up being led entirely by civilian administrators. Buderi, 45. See also Kevles, 303; Brown; Jeremy Bernstein, “Profiles: I. I. Rabi,” The New Yorker 51, October 20, 1975: 49-50; A. Hunter Dupree, “The Great Instauration of 1940,” in The Twentieth-Century Sciences: Studies in the Biography of Ideas, ed. Gerald Holton (New York: W. W. Norton and Company Inc., 1972), 443-468; Joel Genuth, “Microwave Radar, the Atomic Bomb, and the Background to U.S. Research Priorities During World War II,” Science, Technology, and Human Values 13, no. 3/4 (1988): 276-289; 15 Paul Forman especially addresses these issues of funding in his postwar study: Paul Forman, "Behind Quantum Electronics: National Security as Basis for Physical Research in the United States, 1940-1960,” Historical Studies in the Physical and Biological Sciences 18 (1987): 149-229. See also: M. Fortun and S. S. Schweber, “Scientists and the Legacy of World War II: The case of Operations Research,” Social Studies of Science 23 (1993): 595-642; Carroll W. Pursell, Jr., “Science and Government Agencies,” in Science and Society in the United States, eds. David Van Tassel and Michael Hall (Homewood, IL: The Dorsey Press, 1966). On the integration of physics and engineering research during the war see: Peter Galison, Image and Logic: A Material Culture of Microphysics (Chicago: University of Chicago Press, 1997); Hoddeson et. al; Lillian Hoddeson “Research on crystal rectifiers during World War II and the invention of the transistor,” History and Technology 11 (1994): 121-130. 16 For what is often considered the seminal treatment see Rhodes. Rhodes includes one of the most thorough bibliographies as well. 17 As the Manhattan Project accelerated in 1943, Los Alamos drafted many M.I.T. Radiation Laboratory physicists including and , but only well after the first microwave radar sets had already been developed and deployed. The proper assignment of credit for developing radar does not require consideration of the nuclear bomb project. Pollard, 64. 18 Fortun and Schweber, 595. upon technological advancement.19 Issues of social policy and national security, such as

nuclear safety, toxic waste disposal, genetic engineering, and alternative fuel sources

increasingly rely on comprehension of technology and engineering.20 Yet, the great

preponderance of Americans know little about engineers and their practice.21 The

question of credit for the Rad Lab’s achievements therefore assumes contemporary

importance: the public failure to appreciate engineering’s past successes weakens

mainstream understanding of the invaluable role engineering plays in modern life. As

long as Americans fail to recognize engineering’s import, the country runs the risk of

misjudging technology’s tangible benefits and risks, and its capacity to improve or

worsen national stability and public safety.

In 1943, Radio News inveighed against self-styled “prophets” who claimed credit

for single-handedly developing radar.22 The Radio News article insisted that all

individuals and their practices behind “the development of this highly important weapon

should receive proper credit,” and concluded by asking its readers, “Don’t you agree?”23

Responding to that question, I begin by investigating precisely how microwave radar

came to be.

19 In the words of two of the first political scientists to look at the issue of government and science in 1960: “Both security and welfare depend less upon the sheer productive capacity and natural resources and more upon scientific research and development. Increasingly, science plays a role in the economy, in military preparedness, and even in the quest for prestige.” J. Stefan Dupré and Sanford A. Lakoff, Science and the Nation (Englewood Cliffs, N.J.: Prentice Hall, Inc., 1962), 171. 20 Support Organizations for the Engineering Community (Washington D.C.: National Academy Press, 1985), 49. 21 Ibid. 22 “Who Invented Radar?” Radio News 30, July 1943, 72. 23 Ibid.

Chapter One

Engineering Ascendant

“We had the objective of putting together some parts that had already been ordered.” -Ernest Pollard on his initial research at the Radiation Laboratory24

“The U-boat alone can win this war,” boasted German Admiral Karl Dönitz in

1940.25 Though the Allies had deployed sound-ranging methods—sonar—to neutralize

the Unterseeboot during the First World War, in the ensuing decades Germany had begun employing chemicals that effervesced in sea water, creating bubbles that confounded

sonar systems.26 In 1940, with and Poland overrun; with the Molotov-Ribbentrop

Soviet-Nazi non-aggression pact holding firm; with the German U-boat fleet rebuilt and

sonar nearly obsolete; Dönitz’s confidence appeared well founded. By 1941, U-boats

were sinking 16,000 tons of Allied shipping each day.27

By 1942, the situation had changed. Armed with a new technology—microwave radar—United States Army B-18 bombers began to hunt and destroy U-boats. Allied

convoy losses dropped with celerity—faster than the U-boats maneuvering. In May 1943

alone, the Allies sank 42 U-boats, more than they had in all of 1941.28

24 Pollard, 47. 25 Baxter, 37. 26 Kevles, 305. 27 “Radar: The Technique,” Fortune 32, October 1945, 198. See also: Dexter Masters, “We Outsmarted Them on Radar,” Saturday Evening Post 218, September 8, 1945, 110. 28 Baxter, 45. See also: G. Pascal Zachary, Endless Frontier: , Engineer of the American Century (New York: The Free Press, 1997), 171. In all of 1940, the Allies had liquidated only 22 U-boats, in all of 1941, 35. Baxter, 48. With U-boat losses suddenly driven to “impossible heights” in 1943, Germany withdrew its fleet from the North Atlantic.29 himself credited microwave radar as the Allied invention that alone had crippled Germany’s submarines.30 Admiral

Dönitz, formerly so optimistic, now admitted that, “the methods of radio-location that the

Allies have introduced have conquered the U-boat menace;” the enemy “has torn our sole offensive weapon in the war against the Anglo-Saxons from our hands.”31 Dönitz stressed the importance of technology in achieving the submarine’s götterdämmerung: “It is essential that we make good our scientific disparity,” to counter “the modern battle weapon—detection.”32 Here the distinction between microwave and longer-wave radar proved crucial. The Germans could detect longer radar waves, those that were a meter and a half in length, but could never counter, nor did they develop, microwave radar.33

29 “Impossible heights,” taken from Kevles, 315; On the withdrawal of the fleet see Zachary, 171. 30 Lee DuBridge, “The Birth of Two Miracles,” California Institute Forum 6 (1949): 6. Vannevar Bush’s prescient vision in a 1939 letter to former President Hoover had come to fruition, both through the defensive meaning originally intended (as was the case in the ) and in its offensive version (against the U-boat): “The whole world situation would be much altered if there was an effective defense against bombing by aircraft. There are promising devices, not now being developed to my knowledge, which warrant intense effort. This would be true even if the promise of success were small, and I believe it is certainly not negligible.” Conant, 163. 31 “The methods” taken from Conant, 253; “offensive weapon” is from Baxter, 46. 32 Baxter, 46. After the war, Vannevar Bush, director of the Office of Scientific Research and Development, would affirm, “the bitter and dangerous battle against the U-boat was a battle of scientific techniques.” See Vannevar Bush Science the Endless Frontier, (Washington: United States Government Printing Office, 1945), 1. “Radar… was… the decisive weapon in the defeat of the U-boat.” Henry Guerlac and Marie Boas, “The Radar War Against the U-Boat,” Military Affairs 14, no. 2 (Summer 1950): 100. See also: Zachary, 166. Co-inventor of the magnetron insisted, “If we are asked the question— ‘What scientific development contributed most to the winning of the war?’—we should all undoubtedly answer ‘Radar.’” John T Randall, “Radar and the Magnetron.” Journal of the Royal Society of Arts 94: 4715 (April 12, 1946): 303. The British called radar RDF—radio detection finding—until adopting the name American term ‘radar’ in 1943. 33Wavelength of German radar see Kevles, 315 and Guerlac and Boas, 102. On the Germans never developing microwave radar see Vannevar Bush, Pieces of the Action (New York: William Morrow & Co., 1970), 82. British destroyers were already equipped with long-wave (1.5 meter) radar in November 1940. By October 1942, however, the Germans began attaching receivers to their U-boats that could detect the 1.5 meter waves of the British airplanes, often alerting the submarine before the plane even detected it. Part of the reason the Third Reich never built receivers for microwave radar was because they had been unable to generate enough energy in the microwave range (the problem solved by Randall and Boot) and never anticipated that the Allies might have succeeded. Guerlac and Boas, 102, 107. The Allied microwave radar sets that extirpated the U-boat from the Atlantic— opening the ocean to the shipment of materiel and men—emerged from a secret research center at the Massachusetts Institute of Technology: the Radiation Laboratory.34 By war’s end, the Rad Lab proudly stood as the largest laboratory in the world, employing

4000 people, among them one-fifth of America’s leading physicists.35

An analysis of the research conducted at such a lab far exceeds the scope of this study; however, the early days of the Rad Lab reveal an institution strikingly dissimilar from the behemoth that it had become by war’s end. The laboratory opened on Armistice

Day, November 11, 1940 and in December of that year, only 39 scientists worked there; the entire personnel list fit onto a single page.36 By July 1941, however, the Lab had

34 The name was a bit of linguistic legerdemain. Radar is based on the principle of using electromagnetic radiation to detect objects, but the Berkeley Radiation Laboratory, from which the M.I.T. lab’s name was derived, took its name because it worked on , a field in which nuclear radiation is a common byproduct. The name M.I.T. Radiation Laboratory therefore suggested that the physicists in Cambridge were conducting what in 1940 was considered the militarily fruitless field of nuclear physics. The name also helped explain why such a large concentration of nuclear physicists and cyclotroneers that had suddenly descended upon M.I.T. See Guerlac, Radar, 261; Kevles, 303; Buderi, 46. Another reason the Radiation Laboratory took its name from the University of California Radiation Laboratory, was because of , who had directed the Berkeley laboratory. Though Lawrence did not work at the M.I.T. Rad Lab, he was largely responsible for the recruitment of the first physicists. As testament to his efforts, his students and protégés from Berkeley suggested the name, Radiation Laboratory. Conant, 213. See also John Rigden, Rabi: Scientist and Citizen (New York: Basic Books, 1987), 134. M.I.T. was chosen as the site because it was one of the only locations that could be used to gather academic scientists without attracting attention, and research could start away because the university was willing to advance them the money. M.I.T. also already had an independent microwave research program being funded by Alfred Loomis. Conant, 205. The gendered use of “men” refers to the fact that the soldiers who fought were all male. 35 Massachusetts Institute of Technology, Five Years at the Radiation Laboratory: Presented to the members of the Radiation Laboratory by the Massachusetts Institute of Technology, Cambridge, 1946 (Andover, MA: The Andover Press, Ltd., 1947), 6. 36 The official task of the laboratory when it was set up according to contract NDCrc-53, was to “devote suitable laboratory space up to 11,000 sq. ft. net working space, with necessary laboratory facilities and supplementary services with skilled personnel. To conduct with utmost secrecy and dispatch experimental investigations including assembly and tests… relating to certain… radio equipment and methods.” Coordination Committee Series 1, Box 21a. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. When the Microwave Committee, section D-1 of the National Defense Research Committee, met months earlier under Alfred Loomis to set up the Lab, at that meeting on July 14, 1940, their stated goal was: “to organize and coordinate research, invention, and development as to obtain the most effective military application of in minimum time.” Conant, 169. Personnel information from: Coordination Committee Series 1, Boxes 21a and 25. M.I.T. Rad Lab, RG 227. Office of Scientific mushroomed to include 141 scientists and 54 technicians, and by January 1942, the

personnel list totaled ten pages and featured numerous new project groups such as Radio

Frequency and Navigation.37

The early days of the Rad Lab then—from its opening on November 11, 1940

until June 1941—not only provide a manageable period for study, but also enjoy title to

decisive developments in microwave radar: notably, the first microwave radar to fit

inside a plane and the first successful microwave radar submarine tracking on March 27,

1941.

At first glance, it would seem that airborne microwave radar, the kind of radar

that could fit into a plane and detect submarines, owes its birth to physics, because

academic physicists dominated the ranks of the first Rad Lab recruits.38 Almost every

member of the laboratory’s executive Steering committee held a physics doctorate as did

nearly every second-rank leader.39 The laboratory was “a physicist’s world, run for, and

as completely as possible, by physicists,” quite simply, “a physicist’s show.”40

By contrast, during the Rad Lab’s infancy, engineers were few in number.41

Vannevar Bush, director of Office of Scientific Research and Development that oversaw

Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 37 By July 1941, there were also 15 consulting scientists, two guards, and an office staff of 30. By July 1942, the personnel list was 16 pages long. Coordination Committee Series 1, Boxes 21a and 25. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 38 Conant, 278. 39 “Longhairs and Short Waves,” 169. 40 “A physicist’s world,” from Guerlac, Radar, 297. “A physicist’s show,” from Coordination Committee Series 1, Box 49. Guerlac interview of Vannevar Bush, 1944. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. Eight Rad Lab researchers (at least) would later win Nobel prizes, seven in physics: Norman Ramsey, , Luis Alvarez, Hans Bethe, , Edward Purcell, Isidor Isaac Rabi. Edwin McMillan would win in chemistry. 41 William Tuller and William Hall, two of the earliest Rad Lab members, were both electrical engineers. Galison, Image, 817. American civilian science and engineering during World War II, had feared that this dearth of engineers would hinder the production of a workable radar. As radar’s military achievements accrued, though, physicist Lee DuBridge, the Rad Lab’s director, coolly answered Bush, “You see we did not need engineers.”42 It appeared that physics carried the day.

Numerous scholars, however, have countered that radar work at the Rad Lab required first and foremost the expertise of engineering, the creation and study of human- made artifacts, in addition to physics, the study of nature.43 At the Rad Lab, these

“scientific stars were expected to… get to work as an engineering team.”44 Academic physicists “discovered their latent talents as engineers.”45 If ever “there was a place where the latent engineer in the scientist burst out it was at the M.I.T. Radiation

Laboratory.”46

The very assignment of credit to engineering or physics practice, however, must factor in the fact that physics and engineering, despite unmistakable differences in each

42 Bush, Pieces, 138. Earlier in 1940, Robert Millikan, the senior statesman of American physics and 1923 physics Nobel laureate, had cautioned that it was, “a mistake… to concentrate fifty prima donnas in physics at any one spot.” Conant, 219. 43 Engineer and head of the Office of Scientific Research and Development (OSRD), Vannevar Bush commented, centimeter radar was “to a very large extent an engineering job. Many a physicist became an engineer in the process, and a very effective engineer.” Vannevar Bush, Box 23, “What is an engineer,” Vannevar Bush Papers, MC 78. Institute Archives and Special Collections, M.I.T. Libraries. Cambridge, Massachusetts. Lillian Hoddeson adds, “during the war scientists tended to shift their interest from theories of general, abstract, perfect objects to theories of particular, real, imperfect materials.” Hoddeson, “Research,” 129. Daniel Kevles writes that in retrospect much of the work done at the Rad Lab “we can recognize as engineering” today even though it was done by physicists. Daniel Kevles, personal correspondence with the author, 22 February 2006. 44 Buderi, 48. 45 Silvan Schweber, “The Empiricist Temper Regnant: Theoretical Physics in the United States, 1920- 1950,” Historical Studies in the Physical Sciences 17, no. 1 (1986): 93. 46 Brown, 167. Brown later explains, “Radar was derived… from physics, but the link was through a path of that had already converted the basic ideas about electromagnetic fields and the motion of in vacuum and conductors into design elements.” Brown, 433. discipline’s aspirations, do not exist in different spheres.47 Physicists and engineers, especially since World War II, frequently work in symbiosis.48 To observe a laboratory and determine whether the practices of physics or engineering are at work is no trivial task.49 Indeed, both are usually evident.

Evaluating how the skills of each discipline contributed to the development of microwave radar, therefore requires a careful consideration of the Rad Lab physicists’ work: their aims and methods.50 Despite engineering and physics’ being intertwined, many scholars have argued, as this thesis will, that they nevertheless involve crucial

47 Historian Otto Mayr has suggested that drawing such a line at all is impossible, and that the only worthy task is to study what people in the past have envisioned the science-technology relationship to be. Otto Mayr, “The Science-Technology Relationship as a Historiographic Problem,” Technology and Culture 17, no. 4 (October 1976): 670. In the 1920’s, academic physicists became worried that their students were ceasing to be useful and that an education in “classical physics and [with] a ,” was “being usurped by the engineering faculties, whose graduates more and more resembled the product of physics departments.” Spencer Weart, “The Rise of ‘Prostituted’ Physics,” Nature 262, (1976): 16. It has been argued that engineers and scientists intentionally blur their interdisciplinary distinctions in order to earn either monetary support (scientists) or prestige (engineers). Ronald Kline, “Construing ‘Technology’ as ‘Applied Science’: Public Rhetoric of Scientists and Engineers in the United States, 1880-1945,” Isis 86, no. 2 (June 1995): 194-221. 48 After World War II, “heterogeneous professionals—physicists, chemists, [and] electrical engineers,” worked together, Fortun and Schweber, 607. A preliminary Allied report on future German research from June 4, 1945 noted that, “science and engineering are commonly interwoven.” Ernest Lawrence Papers, “Preliminary report of Project 3 Committee on German Research and Engineering,” Series 5, Reel 40, Frame 089152 MSS 72/117c. The Bancroft Library, University of California at Berkeley, Berkeley, California. The fields of radar engineering and radar physics are especially close: “radio engineering had been intimately associated with science from the outset.” Edwin Layton Jr., The Revolt of the Engineers (Cleveland, OH: The Press of Case Western Reserve University, 1971), 251. See also: Fortun and Schweber, 595-642; Kendall A. Birr, “Science in American Industry,” in Science and Society in the United States, eds. David Van Tassel and Michael G. Hall (Homewood, IL: The Dorsey Press, 1966), 35-81; Edwin Layton Jr., “Conditions of Technological Development,” in Science Technology and Society, eds. Ina Spiegel-Rösing and Derek de Solla Price (London: Sage Publications, 1977), 197-221; Galison, Image; Hoddeson et. al. An interesting source on the matter also worth considering is: C. P. Snow, The Two Cultures and the Scientific Revolution (New York: Cambridge University Press, 1959). 49 Mayr, 667. 50 One frequent, but inappropriate line along which the methods of physics and engineering are divided is pragmatism. Historian Silvan Schweber shows how American physicists, to a far greater extent than their European counterparts, were inclined towards the pragmatic. Schweber, “The Empiricist,” 55-98. See also: John C. Slater, “Quantum Physics in America Between the Wars,” Physics Today 21, no. 1 (January 1968): 43-51; Albert E. Moyer, American Physics in Transition: A History of Conceptual Change in the Late Nineteenth Century (Los Angeles: Tomash Publishers, 1983). The American pragmatic streak is old. As Alexis de Tocqueville observed, “In America, the purely practical part of the sciences is cultivated admirably, and people attend carefully to the theoretical portion immediately necessary to application.” Alexis de Tocqueville, Democracy in America, trans. Harvey C. Mansfield and Delba Winthrop (Chicago: University of Chicago Press, 2000), 434. disparities in thinking about problems and diverge in culture, content, and method.51

Each discipline has its own schools and departments, methodologies and pedagogies.

Most notably, improving the design and construction of a specific piece of technology is the hallmark of the engineer.

As this chapter will demonstrate, though Rad Lab physicists relied upon a confluence of engineering and physics skills, the development of airborne microwave radar primarily required enhancing the efficiency of particular machines. As a result, the technology owes its success overwhelmingly to engineering methods. Section I describes how, when Rad Lab physicists arrived in Cambridge, they knew almost nothing about radar, but, fortuitously, much of radar’s underlying physics was already sufficiently understood. Section II reveals, how, still more auspiciously, nearly all of radar’s components had been invented by November 1940. The task remaining then was simply one of improving and integrating already extant materials, not innovating new physics.

This task required the knowledge and expertise of the engineer.

The content of Rad Lab work itself exhibits this central importance of engineering skill. Section III analyzes the laboratory notebooks of a half-dozen leading Rad Lab physicists during the laboratory’s early months. Engineering strategies and methodologies came to dominate the practices of the physicists. Nevertheless, the

51 Vincenti, “Control-Volume,” 165. See also: E.T. Layton Jr. “Technology as Knowledge,” Technology and Culture 15 (1974): 31-41; E.T. Layton Jr. “American Ideologies of Science and Engineering,” Technology and Culture 17, no. 4 (October 1976): 688-701; H. Skolimowski, “The Structure of Thinking in Technology,” Technology and Culture 7 (1966): 371-83; Günter Küppers, “On the Relation between Technology and Science,” in The Dynamics of Science and Technology, eds. Wolfgang Krohn, Edwin T. Layton Jr. and Peter Weingart (Dordrecht, Holland: D. Reidel Publishing Inc., 1978), 113-136; Karl-Heinz Manegold, “Technology Academised: Education and Training of the Engineering the Nineteenth Century,” in The Dynamics of Science and Technology, eds. Wolfgang Krohn, Edwin T. Layton Jr. and Peter Weingart (Dordrecht, Holland: D. Reidel Publishing Inc., 1978), 137-158; Peter Weingart, “The Relation Between Science and Technology—A Sociological Explanation,” in The Dynamics of Science and Technology, eds. Wolfgang Krohn, Edwin T. Layton Jr. and Peter Weingart (Dordrecht, Holland: D. Reidel Publishing Inc., 1978.), 251-286. methods and goals of physics creep up as well. Section IV treats some of the contributions of physics to the Rad Lab, but concludes that given the blitzing speed with

which airborne microwave radar was developed, engineering skills deserve the bulk of

the credit for creating it (and indeed, all microwave radar). It was engineering that

extirpated the Unterseeboot, engineering that neutralized the Axis powers’ sole offensive

weapon.

I. Irradiating the Heavens: Radar Physics in

November 1940

As early as September 1940, British scientists understood numerous aspects of

radar physics. Robert Watson-, the self-titled father of radar, together with Arnold

“Skip” Wilkins, considered the most elementary radar physics question of all: if an electromagnetic signal is sent into space, can one detect its reflection off of an aircraft? In answering this question, Watson-Watt and Wilkins approximated an airplane as a wire half the length of the incoming electromagnetic beam’s waves.52

The details of their approximation bear the hallmark of physics tactics: it strips

the plane of wings and wingspan, dismisses its bearing and altitude, ignores the type of

metal doing the reflecting, and disregards the process of generating and accurately

detecting the waves . In short, they simplify the problem as much as possible.

Watson-Watt and Wilkins were asking a question about fundamental electromagnetic

theory: would the wire (the airplane), when excited by an incoming electromagnetic

wave, reradiate a sufficient amount of energy for detection? Much to their delight, they

found that it would.

52 Brown, 51. A second British achievement in radar physics came with the advent of the cavity

magnetron, the first device to generate waves of sufficient strength for powering a

microwave radar set. John Randall, inventor of the first phosphors used in fluorescent

lamps, teamed up with Henry Boot, lecturer in radiophysics at the University of

Birmingham, to build it.53 Under the directorship of physicist Marc Oliphant and as part

of a consortium of scientists and electrical engineers organized by the British Admiralty,

the two physicists ignored the existing engineering literature on wave generator

technologies, and instead examined the fundamental physics of electromagnetic wave

propagation.54 While vacationing in Aberystwyth, Wales, Randall picked up a book by

one , the first scientist to intentionally demonstrate electromagnetic

radiation, and together with Boot, Randall reviewed the absolute essentials of

electromagnetic wave production.55 Like Watson-Watt and Wilkins’s, Randall and

Boot’s physics-guided efforts met with a dramatic success: the creation of the cavity

magnetron. The British had asked and answered these basic physics questions in advance

of November 1940, when the Rad Lab opened.

Like British scientists, Americans had investigated radar physics as early as the

1920’s. In 1925, and of George Washington University used

both physics and engineering principles to develop pulse ranging: a method of emitting

an electromagnetic wave in microsecond bursts.56 The extremely short time it takes for

53 Ibid., 151. 54 The oversight of the British Admiralty can be found in Stephen Peter Rosen, Winning the Next War (Ithaca, NY: Cornell University Press, 1991), 235-6; The shunning of the literature on existing technologies is from Guerlac, Radar, 226. 55 Hugh Aldersey-Williams, Findings: Hidden Stories in First-hand Accounts of Scientific Discovery (Norwich: Lulox books, 2005), 107. 56Merle Tuve, interestingly enough, was boyhood friends with chief Rad Lab recruiter Ernest Lawrence in South Dakota. They played with short wave radios as children. Luis Alvarez, Box 78, “Ernest Orlando the pulse to return allowed for precise determination of a target’s range.57 In the mid-

1930’s, George Southworth of Bell Laboratories and Wilmer Barrow of M.I.T., completed their theory on the physics of —hollow metal pipes indispensable for the transmission of microwaves.58

Of course, none of this is to suggest that by 1940 scientists had closed the canon on the physics underlying radar. In 1942, Rad Lab Group 42, the theory group, produced monthly reports on the theory underlying nearly every aspect of radar: magnetrons and other resonant cavities, waveguides, antennas, transmission lines, crystals, and transmitter-receiver (TR) boxes.59 But in 1940, though radar physics was still incomplete, it was sufficient for the Rad Lab’s construction of microwave radar.60 No more radar physics needed to be done: a working microwave radar set would fit inside a bomber on

March 10, 1941, a year before the Rad Lab’s monthly theoretical physics reports began.61

Lawrence, ‘A biographical Memoir for Publication by the National Academy of Sciences,’” Luis Alvarez Papers MSS 84/82 cz. The Bancroft Library, University of California at Berkeley, Berkeley, California. 57 Since the emitted wave “pulse” travels at the speed of light, the range of an object can be determined with great accuracy by the length of time required for the wave to return, hence pulse ranging. Tuve and Breit’s discovery did not forever seal the physics of pulse ranging, but, much as was true for the , as far workable microwave radar was concerned, the physics of pulse ranging needed no further improvement. The challenge lay with perfecting the signal emitted. Beginning in 1933, the Naval Research Laboratory took up this charge, sharpening wave pulses to the ideal length and shape. Conant, 170-1. 58 A waveguide in some ways fulfills the same role as the traditional coaxial cable that carries power to your television. Southworth and Barrow individually mastered the physics of waveguides. Burchard, 217. They solved for the waveguide’s natural vibration modes and successfully predicted the wavelengths past which the guides would be unable to transmit signals. At a 1936 joint conference of the American Physics Society and the Institute of Radio Engineers, Wilmer and Barrow resolved their patent issues. Brown, 147. 59 Coordination Committee Series 1, Box 41a. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 60 Though Rad Lab physicists did undertake theoretical studies of the magnetron (in August 1941, John C. Slater released a report on the physics of the cavity magnetron, see Section IV of this chapter), they never actually had to undergo the fatigue of drawing on basic physics to design a different microwave generator. Randall and Boot’s 10-cm microwave generator wanted only minimal refinements. Rad Lab physicists, it should be noted, would develop even higher energy wave generators in the 3-cm (X-band) and 1-cm (K- band) radar systems. This work was begun in June 1941. 61 Coordination Committee Series 1, Box 141. Luis Alvarez notebook #1. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts By November 1940, scientists had adequately mastered the theory of irradiating the heavens and listening for their reply.

II. Targets Acquired: Radar Components in November

1940

Just as important to the Rad Lab’s work as the mastery of the relevant physics was the mastery of radar’s basic technical components. These essential components were: waveguides, pulsers, wave generators, transmitters, receivers, crystal mixers, antennas, and indicator systems (See Appendix: Key Radar Terms Explained). The development of these components too, had been achieved by the time Rad Lab physicists set foot in

Cambridge.

British researchers first constructed working (longer-wave) radar prototypes in

1935.62 In May of that year, in the coastal town of Orford, a team of British physicists and engineers set out to develop a long-range radar system for the defense of the British

Isle.63 Within two months, the team of researchers had constructed a ground-based system that could track an aircraft’s height, range, and azimuth—its east vs. west, north vs. south coordinates.64

62 By no means were the British alone in their pursuit of radar. In 1939, The United States, , Germany, , France, , China, , and all had active radar development programs. Guerlac, Radar, 27. 63 The four researchers on the team were: L. H. Bainbridge, a circuit engineer, , a junior scientific officer at the Radio Department of the National Laboratory, Edward “Taffy” Bowen, a physicist, and George Willis, a technician. Buderi, 65. 64 In August 1937, Keith Wood and Gerald Touch took radar aloft and located a ship several miles away. In war games of September 4, 1937, the British achieved the first ever successful blind landing: an aircraft’s crew found its way back to land despite cloudy and inclement weather by relying on radar to locate the coastline. Buderi, 67-73. Buderi offers an in-depth chapter on British radar. He also addresses the issue of why, with so many countries concurrently working on radar, Britain managed to take the lead. On British radar work in greater detail see: Russell Burns, ed. Radar Development to 1945 (London: Peter Peregrinus Ltd., 1988); E. B. Callick, Metres to Microwaves (London: Peter Peregrinus Ltd., 1990); Denis Robinson, While the Orford system had used long-wave, not microwave radar, the extent of

England’s microwave radar progress became manifest when the British Scientific

Mission arrived in the United States in September 1940.65 Commonly known as the

Tizard Mission, the delegation aimed to share British scientific with the United

States. Among these secrets were many radar technologies previously unknown to

American scientists, including expertise about nearly every component needed to build an airborne radar set.66 Just two months later, when the Rad Lab opened, the physicists there did not have to invent radar’s key components; the British had done that for them (and, as we will see, so had other Americans). Indeed, Rad Lab physicists frequently turned to the

British radar physicist, Edward, “Taffy” Bowen, their Rad Lab scholar-in-residence, for counsel about radar components: “Bowen… practically designed all those [components] single-handed. He decided what was needed and came amazingly close.”67

“British Microwave Radar 1939-41,” Proceedings of the American Philosophical Society 127, no. 1 (February 1983): 26-31. 65 Previous British successes had not passed unnoticed. A 1938 New York Herald Tribune headline blared, “Plane Detection by Television Planned for U.S. Sub: R.C.A. Developing Device, Accidental Discovery in England, to Aid Army and Navy Air defense.” The article describes the British system, though it does not call it that. Coordination Committee Series 1, Box 49c, article from “Herald Tribune” March 20, 1938. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 66 Buderi, 31. Beyond radar components, the box included British insights into a potential bomb. 67 Coordination Committee Series 1, Box 49b, Interview with van Voorhis, March 15, 1944. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. The Rad Lab physicists thoroughly looked to the British for advice. Rabi would remark, “Everything we knew at the beginning about radar we learned from them. They were terrific. They were tremendous, both technically and what I call philosophically. How to use it. What do you need and why do you need it and how would you use when you got it? What are its limitations? What are its strengths?… The British could do wonders with a few sets.” Bernstein, 49. In designing the rotating coils for the Plan Position Indicator, Pollard and Marshall modeled their design after the British indicator model; Coordination Committee Series 1, Box 43a. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. In looking for pulsing methods in November 1940, Bainbridge notes the method “the British utilize” involving “the cutting off of the current though an inductance,” and added that “Bowen suggests 5% as a good figure;” in dealing with problems of waveguides Bainbridge records the “British method” while also proposing his own. Committee Series 1, Box 149, Bainbridge notebooks, November 19, 1940 and August 22, 1941, M.I.T. Rad Lab, RG 227. The Rad Lab physicists’ advanced starting point was not, however, purely a

product of British discoveries; American physicists had independently developed

numerous devices essential to radar. Although the Americans looked to the British for

advice, even Taffy Bowen admitted that American radar had done well for itself. The

technical expertise he encountered in receiver design, crystal mixers, and waveguides

surpassed that of Britain. When Bowen completed his circuit of American laboratories,

he sent a series of communiqués back to his countrymen to enlighten them regarding

what he had found.68

In fact, by November 1940, radar components developed by American researchers

in their own right almost covered the full diversity of the pieces needed. Naval scientists

and engineers in the 1930’s developed the duplexer: a single antenna that could transmit

pulses of waves and detect the echoes.69 produced crystal mixers, devices that

detected the received signal then passed it onto the receiver, which amplified the signal.70

By 1936, the Army’s Signal Corps Laboratory had constructed a working, long-wave radar system that could find the direction and elevation of a fleet of planes at a distance of seventy miles.71 The cathode-ray oscillograph, essential to radar because it can

visually display target locations, reached fruition a full decade before Armistice Day

Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts; Purcell notes “some British circuits,” and reproduces British circuit diagrams as a model for transistors and “range calibration;” Coordination Committee Series 1, Box 389. Purcell notebook, October 17, 1941, M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. As DuBridge summarized, “The Radiation Laboratory drew its first problems and technique directly from the British.” DuBridge, “History,” 2. With a proud twist, Ernest Pollard felt that at the Rad Lab they were going to build radar on schedule, because “we had E. G. Bowen in our midst, who knew that it could be done, or at least pretended to know, and we weren’t to be outdone by any Britisher.” Pollard, 40. 68 Buderi, 99. 69 Conant, 171. 70 Burchard, 217. 71 Conant, 170-172. 1940.72 Without the oscillograph, radar research (not to mention all branches of

electronics research), would have stalled.73

Research on radar-related components had advanced so far that by 1938, many

Americ an physicists were now no longer asking questions pertaining to invention and discovery, but to improvement of existing components. This represented an important change in the scientific and engineering communities’ approach to radar. The work of

William Hansen, a professor at Stanford and a leading American theorist on wave propagation, evinces this shift in focus. In 1937, Hansen proclaimed that his latest paper on the , a type of wave generator, offered “nothing new or deep;” rather, his aim was that his “results would be quite useful” and that “the answers to an enormous number of questions can be put on a single chart.”74 Hansen sought to simplify future researchers’

work, to increase the progress of the technology (the klystron), not the advance of

science.

Similarly, in a 1938 paper, Hansen initially solved a set of partial differential

equations for the boundary conditions of a resonant cavity, but then went a step further.

He transformed those solutions into what are known as equivalent circuit elements.75

Many engineers were far more familiar with electrical circuits than with the newer idea of wave generators, and Hansen’s use of equivalent circuit elements made resonant cavities

72 Cathode-ray oscillographs, forerunner to the modern analog oscilloscope, had served in combat as early as World War I, where they distinguished between the sound waves generated by a shell’s passing over head and the sound waves of a field gun’s muzzle, the muzzle’s boom being the noise necessary to determine the gun’s location. Kevles, 127. Between the wars the cathode ray oscillographs had improved further, and by 1930, American physicist Vladimir Zworykin and German physicist Manfred von Ardenne had each perfected the cathode ray oscillograph so that by 1935, they were ubiquitous in laboratories the world over. Brown, 35. 73 Henry Guerlac, “The Radio Background of Radar,” Journal of the 250, no. 4 (October 1950): 292. 74 William Hansen, Box 6, “klystron efficiency,” William Hansen Papers, SC 4. Special Collections and University Archives, Stanford University, Palo Alto, California. 75 William Hansen, “A Type of Electrical Resonator,” Journal of Applied Physics 9, pp. 654-663 (1938). Hansen shows the equivalency between his resonator and a transformer circuit in section III, 659-661. more accessible to them.76 Hansen was satisfied with the physics behind his discovery.

He was now shifting the focus of his work to the refinement of existing devices. Hansen was thinking in terms of machines and thinking about how to enhance those machines: he was engineering.

All across the country, scientists and engineers followed Hansen’s lead. By

Novem ber 1940, American microwave radar research did not aim to discover new phenomena about waves. It sought to improve technology, to enhance the efficiency of existing components.

The primacy of this drive to improve components is evident in the frenzied state of research by 1940. Bell Labs was developing 10-30 cm wave generators and improving waveguides and microwave electronics. RCA worked on centimeter transmitters, receivers and waveguides. led a program on microwaves and supporting electronics. C.V. Litton Engineering Laboratories conducted research on receivers.

Sperry Corporation operated wave generators at 10-40 cm, and subcontracted Stanford, led by Hansen, to improve receiver and horn design. M.I.T., under the aegis of its president Karl Compton, who was deeply concerned about the tide of the global conflict, supported a sizeable team enhancing transmitters, receivers, and radiators.77 In Tuxedo

76 Hansen proved himself an adept tinkerer. In a 1938 paper coauthored with John R. Woodyard, he showered that under “certain types of array the gain can be increased by an arrangement so that the waves going from the array elements in the direction of maximum transmission are not strictly in phase at large distances.” Emphasis in the original. Hansen and Woodyard found that though theory held that putting waves in phase increased gain, instead putting the waves out of phase increased gain over in phase waves by about 1.8. William Hansen, Box 5, “Antenna Arrays,” William Hansen Papers SC 4. Special Collections and University Archives, Stanford University, Palo Alto, California. Hansen’s discovery demonstrates his interest in very practical matters, namely the construction of improved components. These components were accepted as suitable enough and needing only improvement rather than a wide-ranging dismantling of a new antenna based on first principles. In other words, Hansen’s paper suggests that the physical theory of antennas was good enough. 77 Also, Westinghouse Electric and Federal Telegraph Company each worked on klystron wave generators in the range of 10-40 cm. Coordination Committee Series 1, Box 43b, M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region Park, New York, researchers, many from M.I.T., developed crude microwave radar sets using Doppler methods (as opposed to pulse methods), while the Army and Navy took significant strides in the—still-classified—field of pulse radar.78

As a result of this research, by the summer of 1940, three months before the youthful, eager-eyed team of physicists arrived in M.I.T. room 4-133, seven of microwave radar’s nine basic components existed in useable form.79 The pulser, the transmitter, the antenna, the mixer, the receiver, the waveguides, and the indicator (the oscilloscope), were all ready to go. All that remained for the physicists to develop were the transmitter-receiver box, which they would have to invent themselves, and a wave source generator.80

The United States Navy had long recognized the importance of developing a sufficiently powerful source of microwaves, having applied to Congress for increased

(Boston). Waltham, Massachusetts. At UC Berkeley, Lauriston Marshall and Dave Sloan produced 2.5 kilowatts of wave power on a 50 cm tube named the resnatron. Conant, 160. In a May 4, 1941 address on “Technology [Technology is the name Compton used for M.I.T.] in the service of defense,” Compton related with pleasure how M.I.T. was serving the nation in three ways (keep in mind, this is before America even entered the war). It had 70 faculty and staff serving on committees advising the government. It was conducting research aimed at improving weapons of war (i.e. Rad Lab). Lastly, M.I.T. was operating special educational programs to train engineers and technicians in using radar. Compton spoke at greatest length and with greatest pride about the Defense Training Program. Kart T. Compton, Box 2, “Lectures and Addresses,” Kart T. Compton Personal Papers, MC 416. Institute Archives and Special Collections, M.I.T. Libraries, Cambridge Massachusetts. 78 Conant, 176. For details of the Doppler work done at Loomis Laboratories in Tuxedo Park, see especially chapter 8 of Conant. On the Army and Navy work see: Conant, 170-172. 79 At beginning of lab, only five of initial group were over 40. The average age was 33. governing committee average age was 37. Coordination Committee Series 1, Box 57a. “A Short History of the Radiation Lab.” M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. For more about Room 4-133, the first room the Rad Lab worked out of see: Massachusetts Institute of Technology, Five Years, 4. 80 Recalls Edward Purcell, nobody knew we were “going to use for TR Box.” Purcell concluded simply, “We had to learn by doing.” Jim Lawson would be the first to successfully develop a TR box using a klystron buffer in January 1941. His success stemmed in large part from his engineering skill, as is discussed later. See: Rad Lab: Oral Histories documenting World War II Activities at the M.I.T. Radiation Laboratory, Principal Investigators Bryant, John, William Aspray, Andrew Goldstein, and Frederik Nebeker.” (Institute of Electrical and Electronics Engineers Inc., Piscataway, NJ; a venture of the Center for the History of Electrical Engineering, a joint venture of the IEEE, and Rutgers: The State University of New Jersey, 1993.). Available at Institute Archives and Special Collections, M.I.T. Libraries, Cambridge, Massachusetts. Bryant interviewing Edward Purcell, 246. funding on the matter in 1935. The Navy, however, could not surmount the limits

imposed by vacuum tubes. Instead of returning to first physics principles as Randall and

Boot did, the Navy dropped the project in 1937.81 As a result, despite the swift and

constantly accelerating pace of research in 1940, American scientists and engineers still

lacked one radar component the British already had: a powerful microwave generator. It

was then that the cavity magnetron entered from stage right—cruising across the Atlantic.

“The most valuable cargo ever brought to [American] shores,” “the most valuable

English scientific innovation of the Hitler war,” “the salvation of the allied cause,” and,

simply, “the breakthrough,” the cavity magnetron permitted the birth of microwave

radar.82 Randall and Boot’s magnetron single-handedly put the Allies ahead of the Axis

in radar research.83 Brought to the United States by the aforementioned Tizard scientific

mission, its remarkable power enthralled American physicists.84 As Nobel laureate and

Rad Lab physicist Luis Alvarez described, “we were… awed by the cavity magnetron.

Suddenly it was clear that microwave radar was there for the asking.”85

Microwave radar construction, long in development across the country, now seemed feasible. The physics of waveguides, among other principles, was sufficiently

81 Stephen Peter Rosen, 235-6. 82 James Phinney Baxter III, official historian of the Office of Scientific Research and Development in Buderi, 27; according English physicist Charles Percy Snow in Rhodes, 319; Taffy Bowen’s words in Conant, 191; and Alfred Loomis’s words in Ernest Lawrence, Series 19, Reel 70, Folder 9, “Correspondence with Loomis,” Ernest Lawrence Papers MSS 72/117c. The Bancroft Library, University of California at Berkeley, Berkeley, California. 83 Buderi, 28. 84 Nobel laureates Isaac Rabi and Norman Ramsey eagerly took on the challenge of working with the magnetron. The first entry of the Rad Lab notebook of physics Nobel laureate Luis Alvarez’s is a “model to find resonant frequencies of the O. Tube [the magnetron].” The “O. Tube” is the Oliphant tube, another name for the magnetron, because Randall and Boot worked in Marc Oliphant’s laboratory. Coordination Committee Series 1, Box 395: Rabi’s notebooks. Box 397: Ramsey’s notebooks. Box 141: Luis Alvarez laboratory notebook #1, November 15,1940-Feb 18-1942. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 85 Luis Alvarez, Adventures of a Physicist (New York: Basic Books, Inc., 1987): 83. understood and the key components existed in preliminary form. After receiving the magnetron, “America plainly harbored all the ingredients for making powerful microw ave .”86 Only a month after the , recruitment for the Rad Lab began in earnest.87

The Rad Lab’s purpose was to amalgamate and improve existing components rather than develop novel ones.88 This is evidenced most clearly by the absence of substantial discussion in Rad Lab documents about how to pursue radar: the decision to use existing components (e.g., transmitters, receivers, pulsers, oscillographs rather than a radically new approach), as the basis of research went forward without debate.89 As

Ernest Lawrence, Nobel laureate, inventor of the , and key Rad Lab planner, wrote in November 1940, “I have been having a rather interesting time with myself thinking about methods of detecting microwaves and come to the conclusion that the most promising attack on the problem would be to [work]… along essentially conventional lines.”90

86 Buderi, 99. 87 The details of how M.I.T. was selected and how a lab run by civilian scientists carried the day is a story that begins at least as early as Bush’s founding of the NDRC in June 1940, with the support and succor of Harvard president James Conant, M.I.T. president Karl Compton, Bell Labs president Frank Jewett, physicist Richard Tolman, and others. The story is discussed in many places, but Conant presents one of the easiest-to-read narratives: Conant 158-208. 88 The leisure literature of choice for Rad Lab physicists was detective stories. Massachusetts Institute of Technology, Five Years, 30. Perhaps they preferred these books because just as the detective works to synthesize existing clues into a logical whole, so too were they fusing existing radar pieces into an operational whole. 89 Though the research conducted here did not consider every one of the 1143 boxes of the RG 227 group in Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts, every folder containing documents from 1940-1 was looked into. No document therein discusses the methods of building radar. 90 Ernest Lawrence, Series 19, Reel 70, frame 157376, “Correspondence with Loomis,” Ernest Lawrence, MSS 72/117c. The Bancroft Library, University of California at Berkeley, Berkeley, California. Lawrence never worked at the Rad Lab, but was as involved in its founding and its initial approaches as anyone. The laboratory’s initial organization reveals this component-based orientation.

Research groups broke down along the lines of radar’s parts.91 One group dedicated itself

to magn etrons and another to antennas.92 Future physics Nobel laureate Isaac Isidor Rabi

headed up the transmitter tube group.93 Harvard physicist Kenneth Bainbridge worked on

pulse-modulators (wave generators). William Hall led research on cathode ray tubes.94

There was also a group for receivers.

This component organization shaped the way in which research proceeded during

the Rad Lab’s infancy. Luis Alvarez, who principally worked on magnetrons, maintained

that he regularly crossed “from one group to another” and thereby kept up his “scientific

independence.”95 In reality his independence appears to have extended only so far as he

was free to work on whichever components he liked. His first notebooks are organized by

separate sections for parabolas, transmitters, receivers, pulsers, and transmission-receiver

boxes.96 That is, he could work on components, but there was no research, no creativity,

that existed outside of working on those components. His is not equivalent to a

broad-based, free-wheeling investigation of natural phenomena that the phrase “scientific

independence” usually calls to mind.

91 Conant, 213. In looking at Rad Lab through the National Archives and Records Administration Northeast Region (Boston) documents, this mode of organization is patently clear. The “laboratory was designed to replicate the five-part electronic structure of radar.” Galison, Image, 817. 92 “Rad Lab: Oral Histories,” Purcell interviewed by Bryant, 242. 93 According to Rabi in picking which components to work on, they, “chose up just like a baseball team. We chose sides. What would we take?” Conant, 213. 94 Massachusetts Institute of Technology, Five Years, 12. 95 Alvarez, 89. 96 It’s also worth noting that these sections are almost entirely empty. Coordination Committee Series 1, Box 141. Luis Alvarez laboratory notebook #1. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. As Rad Lab physicist Ernest Pollard similarly recalls, working on component pieces was the only creative feature of his research.97 Pollard’s entire introduction to the

Rad Lab consisted of a terse lecture in which the directors, “informed me that we were

working on microwave radar, that it was intended to help the British in the night Battle of

Britain, [and] that we had the objective of putting together some parts that had already

been ordered.”98 As Pollard reveals, the Rad Lab charged its physicists not with the task

of innovating new physics, but with making a discrete set of already existent components work as a functioning whole—a task of engineering.

In fact, improving component pieces and assimilating them into a set was the only job that could have been asked of the Rad Lab physicists. In August 1940, even before the arrival of the magnetron, a microwave radar set at Tuxedo Park detected an airplane at a range of two miles. That radar set weighed 600 pounds, featured two antennas and required a multi-person crew.99 The work remaining to the Rad Lab therefore could only

have been increasing the range of the radar, making it easier to manufacture, and

miniaturizing the set so it would fit into a plane. Radar already existed. Given that in the

fall of 1940 bombs were raining in Britain, it seemed foolish to pursue new methods

when a satisfactory one already lay at hand. Improving radar, not inventing it, was the

challenge, and improving it meant putting the existing components together more

efficiently. The first Rad Lab report well attests to this: “the first task of the Laboratory

was to assemble a considerable stock of radio and electrical parts and apparatus with

97 Pollard also notes that he “needed” the basic block outline of radar, where the main components are laid out in a diagram, as a “table of organization.” It helped him “get his thoughts straight.” Pollard, 39. Thinking in terms of components was the norm. 98 Ibid., 47. Emphasis mine. 99 Ibid., 102. which to begin development work.”100 To improve and to combine existing equipment is to think in terms of the creation of artifacts; it is the practice of engineering, not physics.101 In the gloom of war, it was engineering that rose to the occasion.

III. In “The Hour of Peril and Need ”: Rad Lab

Engineering102

The position of William “Wild Bill” Hansen at the Rad Lab illustrates the extent to which physics had become peripheral by 1940. Hansen arrived in Cambridge in

December 1940, after the Rad Lab’s leading recruiters asked him to “give up your… pure physics research.”103 Though he was “one of the few people that had a real background or experience” in radar components, Hansen did not end up working in the Rad Lab, only lecturing on fundamental theories.104 Rad Lab director Lee DuBridge considered these

100 This first report was published in March 1941. Frequently these early technical reports highlight a single word: “adjust.” For example, in technical report R-25, one encounters statements such as, “When this control is adjusted,” or “this problem can be solved by adjusting,” or “The adjustment for producing a perfect sine wave was achieved.” It is a process of balancing an assemblage of pieces, of integrating them into a larger whole. Another word frequently emerging is “scheme.” For example, Kenneth Bainbridge, while working on a novel pulser design draws out a “Second scheme.” Scheming reflects a mission of assemblage, of fusing a multiplicity of parts into a cohesive whole—not probing virgin properties of nature. For March 1941 report see: Coordination Committee Series 1, Box 26c Report Number 1 of the Microwave Section, NDRC. March 7, 1941. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. The frequent technical reports of the Rad Lab are designated by the ‘R’-series running from R-2 to R-411. Coordination Committee Series 1, Boxes 114a, 114b, 114c, 115a. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. For Bainbridge see: Coordination Committee Series 1, Box 149. Kenneth Bainbridge Notebook. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 101 This distinction will be more thoroughly explored in chapter three. 102 Longfellow, “Paul Revere’s Ride,” 209, line 127. The full line reads, “In the hour of darkness and peril and need.” 103 William Hansen, Box 2, “Correspondence October-December 1940.” William Hansen Papers, SC 126. Special Collections and University Archives, Stanford University, Palo Alto, California. 104 “Rad Lab: Oral Histories,” Bryant interviewing Robert Pound, 230. lectures minimally important.105 When one man suggested that Hansen lecture thrice a week instead of once, DuBridge wrote, “This… is going too far. I think it was best to get the gang at work on the construction and worry about getting acquainted with the general problem afterward.”106 The actual construction came first, the theoretical underpinnings second.

What fundamental physics was Hansen lecturing about? In looking at the notes

Kenneth Bainbridge took on one of Hansen’s addresses, one is struck by the lecture’s simplicity. In locating an airplane with a microwave signal, Hansen explains why microwaves are the best choice: they provide a narrower beam, both decreasing false positives from ground or water and allowing one to concentrate greater energy onto a target, which increases the strength of the return echo. He then explains how a paraboloid receiver operates and how to calculate the energy received given an initial power from the transmitted pulse.107 Remarkably, Hansen delivered this extremely basic lecture on

October 10th, 1941, almost a full year after the Rad Lab opened, and half a year after the development of the first successful microwave radar sets. If the physics of radar sets was integral to their construction, why was Bainbridge, a leading researcher at the Rad Lab

105 Teaching about how components worked and how to construct them, however, was essential to the lab. As Rad Lab physicist Edward Purcell testified: “We made an effort to learn about [radar construction]. As a teacher, I also would like to claim that part of that atmosphere was due to the fact that a lot of teachers had been assembled. There were fellows who, once they understand anything, can’t wait until they explain it to somebody else. So there was always that kind of transmission in the hallways.” “Rad Lab: Oral Histories,” Bryant interviewing Edward Purcell, 246. 106 Coordination Committee Series 1, Box 43c, “Microwave Committee,” M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. The author of the document is unidentified, but it is almost certainly DuBridge. The letter is written to Lawrence, to whom DuBridge frequently gave updates on the progress of the Rad Lab. The document is also similar in content and style to other reports explicitly authored by DuBridge. 107 In other lectures Hansen discussed theories underlying coaxial cables, waveguides, resonators, and antennas. Coordination Committee Series 1, Box 531. Kenneth Bainbridge Notebooks. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. since its inception, taking notes on the most elementary principles eleven months into his work? Fundamental physics had little demanded his attention in the first place.

Instead, the expertise the Rad Lab physicists relied upon came from engineering, frequently engineering journals and books. Bainbridge, in his copious laboratory notebooks, records references in November and December 1940 to articles he wished to study further. While he does cite two articles from the Physical Review, and one from the

Review of Modern Physics, engineering and instrumentation journals, predominate. He cites a pair of articles each from The Proceedings of the Institute of Radio Engineers, from the Review of Scientific Instruments, and from Electrical Engineering.108 Moreover,

Bainbridge writes on November 24th that, at home he studied an engineering textbook

“on loosely coupled tuned transformers.”109 He notes further that he studied Steimmetz’s

Transient Phenomena, and refers to data on the permeability of iron and the skin effect of silicon. He turns to another engineering source, E. Peterson, for help with “coils and their applications.”110 All of these resources offer minutiae about real-world materials, information useful in the improvement of existing pieces of technology. Bainbridge was looking for help in engineering, not physics. Ernie Pollard turned with similar zest to

108 Review of Scientific Instruments could be classed as a physics journal, but in his notebooks Bainbridge cites these R.S.I. articles primarily to understand how the instruments work, not how to use those instruments to improve the conduct of an experiment; in other words, he is attempting to understand the devices of radar, not the conduct of physics experiments. The articles he cites are: on November 17th, E. Peterson, J. M. Mauley, Electrical Engineering 56, 995 (1937), R.S.I. 5 35, (1935) on stroboscopic circuit, L.R. Qualies R.S.I. 3, 85 (1932) and Electrical Engineering 3 27 (May 1931). On November 18th he lists: Schneider, Phys. Rev. 54 185, (1938) and Morgan, R.S.I. 9, 83, (1938). On November 22nd, he cites: A. Becker, Phys. Rev. 38, 2193, (Dec 1931) and Rev. Modern. Phys. 7, 95, (1935). On November 26 looks to “Coupled Circuits” Proc. IRE 8, 5, (1920). On January 3rd, Bainbridge cites H E. Kallman Proc. IRE 27, 690, (1939). Coordination Committee Series 1, Box 149. Kenneth Bainbridge notebook #1. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 109 Coordination Committee Series 1, Box 149. Kenneth Bainbridge notebook #1. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 110 Ibid. engineering journals: recording a total of five references in 1941, four to recent (1939 and

1940) articles from the Proceedings of the Institute of Radio Engineers and only one to a

theoretical waveguide piece in the Journal of Applied of Physics.111

Just as engineering journals comprised the majority of the lab’s reading,

engineering techniques fueled its research, as the laboratory notebooks and personal

reflections of six leading Rad Lab physicists reveal. The notebooks and recollections of

Luis Alvarez, Norman Ramsey, Ernest Pollard (in particular), Jerrold Zacharias, Stanley

van Voorhis, and Julian Schwinger provide clear insight into the work of the Rad Lab,

and are the focus of this section. These physicists had arrived at the Rad Lab by

December 1940, with the exception of Pollard who arrived after New Year’s in 1941, and

Schwinger who did not leave Purdue for M.I.T. until 1943. Three are physics Nobel

laureates—Alvarez, Ramsey, and Schwinger. All would earn reputations as eminent

physicists. At the Rad Lab, all of them practiced engineering.

The laboratory notebooks of Luis Alvarez expose an engineer and a tinkerer at

work.112 Throughout 1940 and all of 1941, Alvarez left the lime green pages of his

notebooks devoid of any advanced physics formalism or mathematics.113 Nothing more complex than Maxwell’s (extremely basic) equations in matter appear. No quantum

111 In January 1941, Pollard cites, Janskis Proc. IRE 27, 765, (1939) and Kallmann, “Transverse fillers” Proc. IRE 28, 306 (1940). In April he cites, Kallman Proc. IRE, (Aug 1940), p. 353, and in September Barrow Proc. IRE, 24, 1298 (1936) and Southworth, Journ. Applied Phys. 660 (1937). Coordination Committee Series 1, Box 389. Ernest Pollard notebook #1. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 112 One of the best illustrations of this tinkering strategy which affected many other besides Alvarez, can be seen in the events of February 7th, 1941. Luis Alvarez and fellow UC Berkeley colleague Lauriston Marshall among others worked desperately to make their single-antenna prototype work. One of the physicists, “operating on a hunch,” decided to detach the antenna and hold it by hand. Suddenly, they detected a plane at a distance of two miles. Conant, 220. 113 Page 75 of every Rad Lab notebook is stamped lightly with “Harvard Coöperative Society.” mechanical ideas are at work, no thermodynamics undertaken.114 All the great physics

research fronts of the previous fifty years, it seems, do not relate to his work. This is not

to deny that physics comprehension informed Alvarez’s work; indeed, as a Nobel

laureate and inventor of the bubble chamber, and as a member of both National Academy

of Sciences and the National Academy of Engineering, his work constantly integrated

engineering and physics.115

Nevertheless, Alvarez’s concern with the mundane, though important, details of

specific devices illustrate how deeply he had enmeshed himself in the engineering of radar. The vast majority of his notebooks’ pages contain circuit diagrams and geometrical analyses of waves. For example, on April 21, 1941 Alvarez draws out every wire for an amplifier, labeling the “jumper for plugs A to B,” and the “cord for the B-C-D.” He is immersed in the construction of technology. On May 21 of the same year, he drafts, in painstaking detail, the entire plan for the “Control and Monitoring [of the] new Raytheon

Pulser,” carefully drawing in every link for each of 20 separate cables. He labels what each cable measures. He even labels matters as humdrum as the “Off,” “Half On,” and

“Full On” switches and draws designs for every switch, scope and triggering pulse.116

Was Alvarez’s deep concern with the workings of technology an aberration?

Norman Ramsey’s records suggests otherwise.117 After the war, Ramsey lamented having

114 Coordination Committee Series 1, Box 141. Luis Alvarez notebook #1. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 115 Alvarez’s personal papers well attest to this integration. See Luis Alvarez, Box 76 “National Academy of Engineering” and “National Academy of Science,” Luis Alvarez Papers, MSS 84/82 cz. The Bancroft Library, University of California at Berkeley, Berkeley, California. 116 Coordination Committee Series 1, Box 141. Luis Alvarez notebook #1. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 117 So too does the experience of Alvarez’s Berkeley colleague and 1951 Chemistry Nobel Laureate, Edwin McMillan. McMillan’s notebooks overflow with detailed lists of purchased materials: polymers, rubber, spent some of the best years of his professional youth working on war-related research rather than on the “pure” physics of atomic clocks for which he would win the 1989

Nobel Prize. When asked what he would have missed had he never worked at the Rad

Lab, Ramsey replied that he “learned a lot about technology, general know-how, and the general availability of useful equipment.”118 He mentions “technology” and “equipment,” not physics methods.

Ernest Pollard provides a similar but more detailed case study in this emphasis on engineering over physics. Pollard was born in 1906 to British parents living in China, then studied physics under in the at Cambridge

University.119 A nuclear physics professor at Yale at the age of 27, Pollard built that university’s first cyclotron. After being tapped to work at the Rad Lab, Pollard moved to

Cambridge with haste.120 Once arrived, by his own account, “it took no more than one day for me to have shifted my constructive thinking from nuclear physics to microwave radar.”121

hycar sponges, cotton, fiberglass, metals, adhesives, fiber laminates. Such an assortment materials necessarily requires extensive materials expertise—and these are not the usul types of materials essential for , McMillan’s background. McMillan certainly developed an extensive knowledge of materials and material engineering while at the Rad Lab. Coordination Committee Series 1, Box 542. Edwin McMillan notebooks. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 118 Rad Lab: Oral Histories,” Bryant interviewing Norman Ramsey, 269. 119 At the time, , the gold-foil pummeling discoverer of the atomic nucleus, still directed the Cavendish, and Pollard worked with him as well. Pollard, 9. 120 Pollard left in such a hurry that when he returned to Yale five years later, he found his lab coat hanging where he had left it, and the telegram from director DuBridge still in the pocket. Pollard, 39. Pollard was not allowed to come earlier because DuBridge felt that the lab was already filling up too quickly. As DuBridge wrote Lawrence, “We have a sufficiently large group now so that all we can do is to keep them busy.” Ernest Lawrence, Series 5, Reel 40, frame 088132, “Correspondence with DuBridge,” Ernest Lawrence Papers, MSS 72/117c. The Bancroft Library, University of California at Berkeley, Berkeley, California. 121 Pollard, 39. Pollard also notes, however, that at the Rad Lab, “We had to make a transition, and it was one of the harder jobs to make.” Pollard, 36. Did it? His notebooks tell a different story. In mid-January 1941, Pollard begins his first substantial research, considering the general and theoretical “nature of tube noise” in a “Report on the Probability of Coincidence Counting.”122 His goal is to understand the probability of a random ’s position as a function of time, in the hopes of minimizing noise on the indicator screen. To achieve this, Pollard assumes that the Poisson Distribution formula governs the electrons’ behavior.123 This theoretical formula is not linked to a particular indicator technology. Unlike Alvarez, who worked on a specific Raytheon pulser, Pollard initially applies basic distribution laws of physics.

After using the methodology of physics for the problem of “tube noise,” Pollard continues his theoretical musings unabated through January 29, 1941, when he outlines in great theoretical detail a “Report on Signal to Noise Ratio.” He begins by considering an incoming wavepacket with 300 wave cycles in it. He then takes a Fourier transform—a method that tames a complicated wave by breaking it down into its underlying frequencies—to find the maximum amplitude of the wave packet as well as its peak frequency.124 Transform in hand, he then tries to analyze this arbitrarily sized wave

122Admittedly, development delays forced him to spend his first week in “dreamlike” speculation about how to make indicator screens work, and after two months he was transferred to the “roof team” where he continued work on indicators but with different research partners. Nevertheless, the important shifts in his thinking occurred independently of the timetable of this transition. “Dreamlike” from Pollard, 42; Descriptions of Pollard’s (and other’s) administrative movements can be found in: Coordination Committee Series 1, Box 43a. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. e− x x n 123 Poisson’s Distribution formula is given by: W = where x is the average number of electrons n! expected in a given interval and n is the number of electrons that actually arrive. He then employs Stirling’s formula, an approximation for large factorials given by ln(n!) ≅ n ln(n) − n , as a simplification. From this result, Pollard draws a bell-shaped curve to predict the distribution of electrons. Coordination Committee Series 1, Box 389. Ernest Pollard notebook #1. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 124 A Fourier Transform takes a function and re-expresses it as a sum of sinusoidal functions each with different “weights” given by their amplitudes. The form used by Pollard is the continuous Fourier packet’s shape—only to lose himself in algebra (as evidenced by erased, but still legible

and incorrect lines of algebra). He then attempts to solve for four general coefficients, but

fails.125 He “guesses,” but his values, when checked, do not satisfy the equation.

Recognizing his error, Pollard boxes the word “Questionable” directly beside his

inaccurate answers.126

At a loss for general solutions, Pollard abandons all theoretical speculations.

Instead of solving differential equations and carrying out derivations, he begins to draw

schematics for particular indicator circuits. The transition from equations to diagrams is visually startling, as if Pollard’s whole methodology transformed overnight. Starting on page 55, the entry after January 29th, and through his notebook’s remaining hundred

pages as well as the first ten pages of a second laboratory notebook, Pollard’s notes

effervesce with circuit drawings.127 He flatly drops all algebraic and geometrical considerations (save for a single optics equation he records on April 7th).128

Amidst the circuit drawings, one finds Pollard scribbling notes on manipulating

and improving specific component pieces and circuits. By early March, he has a complete

circuit diagram for a whole host of components, in addition to indicators, to be installed

∞ Transform, F(k) = ∫ f (x)e−2π ikx dx where f(x) is the initial function and F(k) is the Fourier Transform, a −∞ function whose value gives the amplitudes of the sinusoidal components. 125 Pollard neglects A, determines that B2=C2+D2 and solves for C and D. He cannot find values for C and D. Coordination Committee Series 1, Box 389. Ernest Pollard notebook #1. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 126 That is, his values for C and D. Ibid. 127 Ibid., Ernest Pollard notebooks #1 and #2. In September 1941, Pollard left indicators for a new project. 1 1 1 d 128 On that date, Pollard records = + + , an equations he takes from “Myers’ Electron Optics.”

F fg fs fg fs The equation defines the limited area for which a magnetic coil can focus an image. Pollard also reproduces an extremely crude graph of position versus electric field. Ibid., Ernest Pollard Notebook #1. in a B-18 bomber.129 On March 20th, 1941 he writes, “HS Sweep puts 200 cycles on the vertical deflection coil and shifts its phase to any required position.”130 The details of this

passage—Pollard attempting to improve control of a beam that sweeps out an image on

the indicator screen—are less important than its character: a manipulation of a specific

piece of equipment to make it operate with greater efficacy and dependability. His pages

shift from transcendent theoretical meandering to particular circuit designs. Working with

circuits and thinking in terms of improving them, this is engineering.

The notebooks of Jerrold Zacharias similarly exhibit the ascension of engineering

in place of the physicist’s usual methods. Zacharias’s notebooks begin with the practices

characteristic of any good experimental physicist: filling his pages with charts of values

for current, voltage, resistance, and intensities. In early March 1941, amidst this sea of

data, however, Zacharias introduces (and keeps using for three months) circuit diagrams.

These diagrams are designs for how to link specific models of parabolas (antennas) to the

radar set.131 He—like Pollard, Ramsey, and Alvarez—sought to improve existing

technology. Again, they were calling upon the skills required by engineers, not physicists.

The case of Stanley van Voorhis further reveals the primacy of engineering

practice. Van Voorhis tackled the problem of the Transmitter-Receiver box (TR box), the

box which must both transmit the radar pulse and receive it. The challenge of building a

TR box is that during transmission, the power of the emitted pulse often burns out the

receiver crystal. Van Voorhis could have tried protecting the receiver from burnout by

using physics principles of signal transmission. Instead, he applied a purely electrical

129 Ibid. 130 Ibid. 131 Coordination Committee Series 1, Box 508. Jerrold Zacharias notebooks. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. engineering method to try to disconnect the receiver from the transmitter during the outgoing pulse.132

Many other Rad Lab researchers’ laboratory notebooks reveal similar themes.

Kenneth Bainbridge’s notebooks, for example, preliminarily employed physics methodologies before switching to techniques for improving individual pulser models.133

William Higinbotham focused almost exclusively on electrical engineering: in his notebooks, only circuit diagrams and ideas about those circuits appear.134

Julian Schwinger provides a final instructive case study of engineering at the Rad

Lab. Schwinger arrived in Cambridge two years after the first operational microwave radar set reached completion. Enmeshed in the Rad Lab milieu, Schwinger assimilated a substantial portion of the “engineering culture of the Rad Lab” while abandoning the

“physicists’ abstract scattering theory of .”135 Of course, to make such a

132 Van Voorhis’s notebooks contain barely an entry per month by April 1941, but initially he made frequent entries. His notebooks are filled almost exclusively with circuit diagrams and data collection plots The circuit diagrams for van Voorhis’s proposed TR-box circuit can be found in the first few pages of Committee Series 1, Box 472. Van Voorhis Notebook #1. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 133 In a single entry from August 28, 1941, a date, albeit, outside the period of interest, Bainbridge exhibits how engineering methods supercede physics methods for solving radar problems. In dealing with a problem of a parabola’s reflection and emission, Bainbridge takes a classic physicist’s reductive approach: treating the radar’s radiation source as “a point source, uniformly radiating.” He uses the “first order Bessel Function” to manipulating the point source, writing out, “Exact integration for the equatorial plane… above formula ok,” and solving: “First root 3.83.” After finding this value, however, Bainbridge drops the theoretical framework. He only used it to find a rough answer. He next looks at the issue of a beam angle width at zero, and looks across the spectrum at specific angles in order to find out constants specifically related to the to the parabola under consideration. Coordination Committee Series 1, Box 149. Kenneth Bainbridge notebooks. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 134 Committee Series 1, Box 273. William Higinbotham notebook. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 135 Galison, Image, 821. Galison uses the notion of an anthropological “Trading Zone,” in which fields of expertise overlap so that physicists and engineers can share their skills and ideas despite differing backgrounds to explain much of the kind of work that unfolded in the Rad Lab. Physicists and engineers worked out a “powerful, locally understood language to coordinate their actions,” Galison, Image, 833. See also: David Kaiser, Drawing Theories Apart: The Dispersion of Feynman Diagrams in Postwar Physics (Chicago: University of Chicago Press, 2005), 41. transformation, Schwinger had to comprehend the underlying physics as well.

Nevertheless, by his own admission, he “first approached electromagnetic radar problems

as a nuclear physicist, but soon began to think of nuclear physics in the language of

electrical engineering.”136

To translate nuclear physics into electrical engineering, Schwinger, like Hansen in

the late 1930’s, used equivalent circuit elements. These circuit elements, when used to

calculate the property of a waveguide, for example, give the same predictions as the

mathematics of the physics theory. In short, they turn a theoretical physics problem into an electrical engineering one. As the postwar, Rad Lab-authored “Waveguide Handbook” phrased it, equivalent circuits cast “the results [of] field calculations in a conventional engineering mold from which information can be derived by standard engineering calculations.”137 Thanks to the “dominant role” Schwinger played in devising these equivalent circuits, Rad Lab researchers needed not solve Maxwell’s equations and go through the generally onerous task of applying boundary conditions.138 They could calculate based solely on electrical circuits.139 They used a “conventional engineering

mold,” not a physics one. Rad Lab researcher recalled, “the most

important lesson [my supervisor] taught me was that I could think of resonant cavities as

136 “Biography of Julian Schwinger,” in Nobel Lectures in Physics: 1963-1970, (River Edge, N. J.: World Scientific Publishing Co., 1998), 153. Forman asserts that Schwinger himself is responsible for this biography. See Paul Forman, “‘Swords into Ploughshares’: Breaking new ground with radar hardware and technique in physical research after World War II,” Review of Modern Physics 67, no. 2 (April 1995): 397- 455. 137 Marcuvitz, 104. For details on the interchangeability between circuit and field calculations made possible by equivalent circuits and “equivalent representations,” see especially Marcuvitz, 108-130. 138 The Waveguide Handbook singles out “J. Schwinger” for his “rather intensive and systematic exploitation of both the field and network aspects of microwave problems.” Ibid., vii. 139 Ibid, vii. interchangeable with distributed circuits.”140 Schwinger’s translations of microwave

components into the “language” of circuit elements aided Wiesner in developing a

successful transmitter-receiver box, and ameliorated development difficulties for

innumerable others.141

Schwinger, it should be noted, worked in the theoretical group at the Rad Lab;

yet, his work concerned itself with the engineering of microwave radar components. So

too, it would appear, did the work of nearly every Rad Lab scientist. The electrical

engineers at M.I.T. “had a characteristic way of learning about the world, and it was the

theorists who came to adopt it.”142 The construction of tangible products emerged as the goal of every researcher: so much so that one could not determine a Rad Lab member’s training background based on the work he or she produced.143 , for

example, a physical chemist from University of Pennsylvania, just happened—with

appropriate fortuity—to have a talent for circuit design.144 It may seem strange that a

chemist proved so adept at electrical engineering, but at the Rad Lab, everyone

engineered.

If the physicists were in fact engineering, one would expect that, at least initially,

they made many missteps. This is precisely what occurred. In the laboratory’s early days,

director DuBridge admitted that almost “none of the men had any experience with

140 Jerome Wiesner, “Remembering the Rad Lab and the RLE,” in Jerry Wiesner: Scientist, Statesman, Humanist; Memories and Memoirs, ed. Walter Rosenbluth, (Cambridge, MA: The M.I.T. Press, 2003), 210. 141 Wiesner’s box came well after (and improved upon) Jim Lawson’s original TR box, discussed below. 142 Galison, Image, 820. 143 Wiesner, 212. Degrees held by Rad Lab staff in 1943 were incredibly diverse. They included: , biochemistry, biology, , geology, astronomy, geo-physics, political science, nuclear physics, botany, anthropology, ceramics, architecture, chemistry, math, optics, music. Massachusetts Institute of Technology, Five Years, 32. See also: Bush, Pieces, 138. 144 Buderi, 106. microwaves or basic engineering.”145 Harold Hazen, chairman of the M.I.T. electrical

engineering department, annoyed that a nationwide conglomeration of physicists had

been chosen to work on a project his department had been struggling with for years,

recalled, “It was often ironic for our group to watch the scientists repeating the mistakes

that our people had made in the earlier years.”146 Rad Lab physicists had been charged

with a task in which they were not expert. They knew it. On the laboratory bulletin board,

someone posted a clipping that read, “What we know here is very little; but what we are

ignorant of is immense.”147

This ignorance made itself known in finished products. The first radar set,

completed January 2nd, 1941, featured misaligned frequency pulses, non-polyethylene cables that were poor electrical insulators, and mismatched waveguides that led to

(problem-causing) standing waves inside of them.148 The first successful single-antenna

set to detect an airplane, completed on February 7th, 1941, connected the transmitter-

receiver box and the antenna with crude, faulty, homemade coaxial cables. On the set’s

display, tuning knobs abounded, each of which had to be hand-adjusted in order to

maintain pulses of constant frequency.149

With similarly amateurish design skills, Rabi and Ramsey attached two tuning stubs to an instrument that would measure the cavity magnetron’s power. Rabi and

Ramsey found, however, that adjusting one tuner cancelled out the effect of the other.

145 Ibid., 47. 146 Harold Hazen, “Memories: An Informal Story of My Life and Work.” Unpublished Manuscript, 1976. Available at Institute Archives and Special Collections, M.I.T. Libraries. Cambridge, Massachusetts, 3-29 [that is, chapter 3, page 29; this is how the manuscript is organized]. Hazen also details how other M.I.T. electrical engineering professors, felt they received a short shrift because of the presence of physicists who been chosen to work on a project that they, the M.I.T. electrical engineers, had been working on for years. These other professors derived some pleasure from the engineering physicists’ missteps. 147 Pollard, 51. 148 Brown, 168. 149 Buderi, 102. They then realized that they had built the two stubs exactly half a wavelength apart so

that they naturally compensated for one another, evidence, Ramsey remarked, of “the low

level at which we started.”150

Given this low level of experience, the electrical engineering dexterity and knowledge of certain Rad Lab personnel proved essential to radar’s success. Jim Lawson built America’s first successful transmitter-receiver box in January 1940, allowing for the construction of a single antenna set on January 10th.151 An amateur radio enthusiast,

Lawson had mastered radio parts and transmission, and this radio engineering experience

played a decisive role in the transmitter-receiver breakthrough. Like van Voorhis,

Lawson applied engineering skills to the TR box problem, but Lawson’s efforts

succeeded.152 Alvarez remarked, “If we had been paid in proportion to our contributions

to the success of the first microwave radar program, Jim Lawson would have earned

more than half the monthly payroll.”153

Later on in 1941, Denis Robinson, a British electrical engineer, crossed the

Atlantic to join Taffy Bowen as a second British radar expert. Robinson was singularly

responsible for Britain’s use of crystal receivers—receivers several times stronger than

American ones.154 His engineering work with crystals significantly improved radar

detection.

Other electrical engineers also made important contributions: William Tuller and

William Hall, Rad Lab members since December 1940, proved critical to the

150 Another “goof” was the Rad Lab’s failure to create a separate transmission line group for “a surprisingly long time,” neglecting the importance of signal conduction Rad Lab: Oral Histories,” Bryant interviewing Norman Ramsey, 257. 151 Guerlac, Radar, 262. This set was still unreliable, but not because of the TR box. 152 Brown, 168. 153 Conant, 218. 154 Buderi, 119. development of receivers and indicators, respectively. Later on during the war, American

radar succeeded where British radar lagged, precisely because the Rad Lab hired “a

favorable proportion of young engineers,” who brought “the equipment [up] to practical

operational quality.”155

Indeed, every aspect of the Rad Lab’s microwave radar work required engineering

know-how. The physicists who led the laboratory looked to engineering journals and

engineering experts. They worked on improving specific circuits and pulser designs,

thereby constantly thinking in an engineering framework.

IV. Physics Research at the Rad Lab: Complication

and Resolution

The state of radar components before November 1940, the organization of the

Rad Lab, the research filling the physicists’ notebooks, even the physicists’ construction

miscalculations: all demonstrate that engineering skills powered the radar work of the

Rad Lab. Why, then, did theoretical physics work continue? In August 1941, for example, after eight months of research, John Clarke Slater completed a 193-page report on “The Theory of the Magnetron Oscillator.”156 A leading Rad Lab thinker spent eight

months working out the theory behind a technology that had already proved effective;

nothing more needed be discovered. Nor was Slater’s pursuit an exception: after the

155 Karl L. Wildes and Nilo A Lindgren, A Century of Electrical Engineering and Computer Science at M.I.T., 1882-1982 (Cambridge MA: The M.I.T. Press, 1985), 198. 156 Committee Series 1, Box 115a, Report R-171. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. On the importance of Slater’s work see Randall, 310. report was published, theoretical work on the magnetron not only continued, it

accelerated.157

The reason for this physics research is that the laboratory did not solely depend

upon engineering. In fact, several Rad Lab projects were rooted in physics skills, and

general physics principles commonly aided the laboratory’s progress. Nevertheless, as

will be shown at the end of this chapter, this physics research is only a small part of the

Rad Lab story, and, moreover, airborne microwave radar had been successfully

developed well before any of these physics contributions came along.

One example of the use of physics knowledge can be found in June 1941, in a

“Report on Pursuits” that derived general principles of detection. The

report’s derivation is conducted without regard to specific kinds of enemy planes.158 In other words, the report consists of the physics behind detection, not the engineering of components to be used for detection. This research unfolded even before the foundation of the official Rad Lab Theory Group in 1942.

Alvarez’s work on Ground Controlled Approach, also know as blind landings, is another example of a physics approach. The problem with blind landings is that a plane’s radar set detects the aircraft’s own reflection rather than the ground. Eliminating this reflection required understanding the advanced electromagnetic physics of wave propagation, reflection and attenuation, rather than engineering. This advanced physics is precisely what Alvarez called upon.159

157 “Longhairs and Short Waves,” 169. 158 Committee Series 1, Box 115a, Report R-6. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 159 Conant, 158-161. In the 1930’s, physicists worked on the problem of blind landing and navigation by radiolocation methods, but without success. As early as 1938, M.I.T.’s Papers in Physical Oceanography and Meteorology considered the problem of navigation through foul weather. In January 1940, Edward Similarly, the theoretical group of the Rad Lab used Maxwell’s electromagnetic equations to understand the byzantine geometries of certain waveguides. Unlike the case of simple cylindrical or box-shaped waveguides, solving Maxwell’s equations for these labyrinthine systems proved “one of the major projects” of the theoretical group. Though they worked for an applied purpose, the research they conducted was theoretical physics through and through.160 Many of these theoretical scientists’ laboratory notebooks abound with physics ideas and techniques.161

Physics methods made an impact on the lab’s organizational schemes as well, completely breaking with the traditional methods of engineers. M.I.T. Electrical engineer

Edward Bowles “thought that some engineering and engineering methods might well be tried in the Radiation laboratory, but it was Bowles against the field—they pasted hell out of him.”162 Bowles also attempted to open an engineering office at M.I.T., but the Rad

Lab physicists would not deign to use it.163 At the Rad Lab, physicists chose what to produce and how to work with industry in conformance with their physics background.

Bowles, William Barrow, W. Hall, and F.D. Lewis, and D.E. Kerr submitted a technical paper on blind landings to the American Institute of Electrical Engineers entitled The CAA-M.I.T. Microwave Instrument Landing System.” The blind landing problem, demanding both physics and engineering skills. For Papers in Physical, see Houghton, H.G. and W. H. Radford “On the Local Dissipation of Natural Fog,” Papers in Physical Oceanography and Meteorology, Volume VI, no. 3. Cambridge and Woods Hole, MA: 1938. Available in Coordination Committee Series 1, Box 49a. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. For Bowles, et. al. see “M.I.T. Office of the President, 1930-1959 (Compton- Killian)” Box 34, Folder 6, AC 4 Institute Archives and Special Collections, M.I.T. Libraries, Cambridge Massachusetts. 160 Albert Heins, “History of the Theoretical Group at the Radiation Laboratory” in Radar in World War II, Henry Guerlac (Los Angeles: American Institute of Physics/ Tomash Publishers, 1987), 625. 161 Norman Ramsey’s many notebooks, for example, especially after June 1941, deal with a good deal of advanced physics. Demanding integrals of sophisticated, abstract electromagnetic quantities abound. Coordination Committee Series 1, Box 141. Norman Ramsey notebooks. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 162 Vannevar Bush as quoted in Fortun and Schweber, 606. 163 Ibid. Fortun and Schweber deal at length with Operations Research and Systems Analysis during and especially after the war, and the methods these fields came to employ. Perhaps the most important outgrowth of physics expertise was gun-laying radar.

A system in which a radar set tracked one plane and it alone, gun-laying radar then

relayed the aircraft’s position to an artillery piece that automatically fired at the aircraft

even as the it changed trajectory. Such a radar system demanded keeping track of a single

target, a task to which the physicists were eminently well-suited. Many Rad Lab scientists were experimental nuclear physicists who, during the 1930’s received training in trapping subatomic particles. As a result, they had learned to develop exactingly

precise synchronized circuitry for their experiments. They used this expertise to build

range gates—a series of circuits so precisely timed that only echo pulses from a certain distance (i.e. from a single plane) could enter the radar receiver. As a result, radar sets could track one object at a time without confusion from trees, buildings, or other airplanes.164 In short, the methodology and skills of experimental and theoretical

physicists had clearly advanced the laboratory’s mission.

Nonetheless, physics’ function in the Rad Lab’s efforts was not so significant that

it should obfuscate engineering’s invaluable contributions. In evaluating all the physics

successes above, one must think about the production timetable. This point is especially

important for airborne microwave radar. Airborne microwave radar, the lab’s first radar

technology, the one that sent the German submarine to darker waters, was completed on

March 27th, 1941. On that day, Luis Alvarez writes in his notebooks that a “B-18 system

saw ships at 8 miles and submarines at 3.5 miles.”165 Yet, Slater’s report on the theory of

164 Buderi, 109-111. 165 One can track the major progress of airborne microwave radar through Luis Alvarez’s master calendar in the back of his first notebook. On January 4th, 1941, Alvarez celebrates with boxed words and exclamation points, “Whole [two antenna] set up working on roof!!” A month later on February 7th, he notes the accomplishment of the first single-antenna system: “Plane seen at two miles…. Success.” On March 10th Alvarez boxes the words, “B-18 outfit working in the air.” Three weeks pass without a stroke of his pen. Then on March 25th, he records, “B-18 saw planes on scanning scope at 3 miles.” Then of course, the magnetron came a full five months later. Theoretical night fighter detection, too,

came too late to contribute to the defeat of the U-boat. Nor did the Rad Lab’s unique,

physics-influenced manner of working with industry fully develop until well after March

1941. New technologies such as gun-laying and blind landings played vital roles in

World War II, but the achievement of submarine detection is not theirs; they came to

fruition well after the birth airborne microwave radar.

Airborne microwave radar, as explored at the beginning of this chapter, was

extremely important.166 The main request the British had made of the physicists in

Cambridge, Massachusetts, was assistance with airborne interception, so that Royal Air

Force planes could detect, at night and at close range, German aircraft. As German air

attacks slackened, however, and Hitler indefinitely postponed Operation Sea Lion—the

proposed invasion of England—Admiral Dönitz’s U-boat emerged as the most potent

threat to the Allies.167 Yet, by March 27, 1941, this problem was essentially solved.168

Radar’s range would be improved as would its precision, but on March 27th a United

States bomber detected an American submarine near New London, Connecticut at a range of over three miles. Given this timetable, physics skills and methodologies can only

claim a tiny fraction of airborne microwave radar’s glory.

Even beyond submarine detection, however, engineering played an inimitable role

in all the war’s radar creations. As World War II progressed, the Rad Lab worked less with already existent components and designed many of its own, but the task before the

on March 27th, he makes note of the submarine detection. This is the last event Alvarez ever notes in his master calendar. Committee Series 1, Box 141. Luis Alvarez notebook #1. M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 166 On radar’s primary importance, see note 32. 167 Massachusetts Institute of Technology, Five Years, 16. 168 The of nuclear bomb provides an analogous example. Once the fission bomb successfully exploded at the site on July 16, 1945 no doubt remained as to the weapon’s technical feasibility. physicists typically remained one of improving the efficiency of technology and

amalgamating the individuals pieces more effectively.169 The examples of physics

expertise mentioned in this chapter were important, but were diffused across the months

and were almost always coupled with engineering skills in order to reach fruition. The

methods of engineering constituted, by far, the most important knowledge in the lab—

certainly for airborne microwave radar, and even beyond.

Microwave radar cleared the Atlantic and overwhelmed the Axis. It saw action at

the beachhead at Anzio, in the Battle of the Philippine Sea, in the campaigns in Guam,

Iwo Jima, and Okinawa, and in the landings at Normandy, and brought down the German

V-1 buzz bombs.170 So how would a triumphant nation appraise its heroes? Would the discipline of engineering receive the laurels destined to be heaped upon it? Examining these questions, a most curious story emerges.

169 It is worth noting, as we will see in the next chapter, that almost all Rad Lab physicists understood their work on the whole as largely engineering, even if they did not readily embrace it as such. 170 Rigden and Rabi, xxii; Microwave radar produced at the Rad Lab also made obsolete all other forms of radar. Massachusetts Institute of Technology, Five Years, 6.

Chapter Two Engineering Descendant

“My intention is to tell of bodies changed To different forms.” -Ovid, Book I of the Metamorphoses171

In 1991, published the obituary of Rad Lab researcher Jacob

Millman, describing him as “an electrical engineer” and former chairman of Columbia

University’s electrical engineering department. Yet, when discussing this engineer’s contributions to radar, the Times chooses an odd phraseology: “Except for three years during World War II, when he was a scientist with the Radiation Laboratory at M.I.T., he was a professor of engineering at City College of New York from 1936-1952.”172 The choice of the word scientist—in opposition to his professorship in engineering—implies that his contributions to radar were not due to engineering expertise. Historically, this was not the case: at the Rad Lab, Millman was definitely an engineer.173 This may appear to be a trifling distinction, but as this chapter will demonstrate, public interest in, respect for, and credit afforded to engineering decreases whenever an invention is simply termed

171 Ovid, Metamorphoses, trans. Rolfe Humphries (Bloomington, IN: Indiana University Press, 1955), 3, lines 1-2. 172 “Jacob Millman. Expert on Radar, Dies at 80,” New York Times, May 24, 1991, pg. B8. 173 At the Rad Lab, Millman worked on precision aircraft detection. He improved the height-finding accuracy of radar sets with the beavertail technique, and the central question he tackled was one of electrical engineering: “the electrical characteristics suited to optimum… tracking.” He had to work within engineering specifications: a radar set that could detect 90 percent of enemy bombers within 60 miles to an accuracy of within 1000 feet at 45 miles away. This project was designated AN/CPS-4. Guerlac, Radar, 459. His second main project was also an improvement of detection range, known as the V-beam system (classified as AN/CPS-6). Millman’s title was project engineer, which does not mean project leader. That title fell to Andrew Longacre. Millman had to integrate five separate magnetrons and myriad display scopes. Guerlac, Radar, 463. an accomplishment of science rather than specified as one of engineering. In the decades

from 1945 onwards, a transfer of credit for microwave radar took place.

This chapter examines this postwar distortion of credit. Section I investigates how

popular periodicals characterized the achievement of radar once the war ended and the

technology became declassified. Though engineering was initially afforded its due, credit

soon became muddled with the discipline of physics. Section II looks at the immediate

cause of this muddling: physicists’ spectacular postwar rise to nationwide popularity and

authority. Section III describes how, by contrast, the public considered engineering a

second-rate discipline, and how in the postwar era engineers proved unable to claim

credit not only for radar but also for other achievements. Section IV will then show how,

with physicists revered by the public and trusted by the government, the achievements of

the war, which were properly associated with physicists themselves, rapidly became

affiliated with physics as well. As a result, retellings of the Rad Lab story since the war’s

conclusion have distorted the laboratory’s legacy, as section IV will also evince.

I. Withering Laurels: Engineering’s Immediate

Postwar Reception

Well before radar became declassified, the public clamored for information about

it. In July 1943, with microwave radar only beginning to evince its powers of sight, Radio

News observed a “considerable hullabaloo in the press” over the secret weapon.174

Claims and counterclaims about the weapon’s applications and functions were traded. No one had any idea what the truth was, of course, knowing only that the method had been

174 “Who Invented Radar?” 4. under development since the 1920’s and likely saw action in the Battle of Britain.175 The character of this “science”—the “revolutionary wartime science of detecting and ranging by radio”—remained an enigma.176 By early 1945, the “black secrecy” enshrouding this

“phenomenal” weapon still reigned.177

Then, in rapid succession, fell, Allied forces stormed Okinawa, the nuclear

bomb dealt a sockdolager to Nagasaki, and the United States government revealed the

secret of radar. The Joint Board on Scientific Information Policy released the official

radar report to the media on August 15, 1945—Victory over Japan day. They had it right.

Unlike prewar speculations that emphasized a new “science” of radar, the Joint Board on

Scientific Information Policy made plain that the Radiation Laboratory had served as a

“major factor in the rapid advance in radar techniques in this country.”178 The emphasis

lies squarely with the “techniques” needed to build radar equipment. Physics is only a minor part of the story. Above all else, the Rad Lab conducted “the improvement” of technology, not a query into natural principles.179

Nor did the media then distort the story. Despite the great swirling din over the

physicists who had ushered in the nuclear age, periodicals at the time almost

unconditionally identify the Rad Lab’s achievement as primarily the brainchild of

engineering shrewdness. In September 1945, The Saturday Evening Post lauds the

myriad types of radar as war-clinching “technical developments.”180 This stress on the

“technical” basis of radar casts it as an accomplishment of engineering, and like any

175 Ibid.; “Radar: Background of War’s Greatest Development,” Scientific American 169, August 1943, 78- 80. 176 “Radar: Background of War’s Greatest Development,” 78. 177 “May Lift Veil on Radar,” Business Week, February 10, 1945, 5. 178 United States, Joint Board on Information Policy, Radar: A Report on Science at War (Washington, D.C: Office of War Information, 1945), 10. For the Rad Lab’s role in the war, see principally, 9-12. 179 Ibid., 10. 180 Masters, 20. piece of specific technology, radar sets soon rolled out from “production” facilities.181

The Post article labels the Rad Lab’s earliest creations “Model-T radars,” conjuring up

Henry Ford’s famous inventive spirit and thereby further limelight wartime radar’s engineering character.182 At the Rad Lab, scientists fresh from “the academic groves”

practiced “engineering.”183 The millions who read the Saturday Evening Post, came away

having learned that the “engineering” of Rad Lab physicists had birthed microwave

radar.184

Readers of Fortune, another popular periodical, would have concurred. In a pair of articles published in October 1945, the magazine describes radar as “an adaptation of a well-known electromagnetic phenomenon.” Radar “in theory [is] simple to the point of austerity, but [in] practice is one of the most complex” inventions.185 The tremendous

achievement of radar lay in engineering: in “adaptation” and “practice.” Fortune even

contrasts “the British…‘pure’ research” with “American… engineering.”186 Even when

the magazine cites the “basic radar research” of the Radiation Laboratory, it refers not to

pure science, but rather to the basic, as in foundational, engineering of components—the

181 Ibid., 21. 182 Ibid. 183 Ibid., 110. 184 Circulation of The Saturday Evening Post climbed to 3 million in 1937 and 4 million by 1949. For more circulation data see “Saturday Evening Post” Downtown Magazine, (accessed 27 January 2006); “The Saturday Evening Post,” Wikipedia: The Free Encyclopedia, (accessed 27 January 2006). 185 “Radar: The Technique,” Fortune 32, October 1945, 139. Ernie Pollard affirmed that for Rad Lab physicists radar, “didn’t seem very complicated to us; it just seemed incredible.” Pollard, 40. 186 “Radar: The Technique,” 142. The article adds that “Americans have also made important contributions to radar research,” but this refers to the prewar research of Tuve, Southworth and others. Fortune is either mistaken or ignorant, however, because it never mentions the many practical engineering advances made by British physicists and engineers in radar during the war. “keystone” of the entire operation.187 The Radiation Laboratory is to be remembered as,

“by any standards” having “contributed most” to “radar research and engineering.”188

Time similarly underscores engineering’s primacy in the development of radar.

An August 20, 1945 article esteems engineers as the official radar experts, contending that it was American “engineers” who took the British magnetron and “perfected its radar adaptation.”189 Again appears this notion of technological “adaptation”—a term especially linked to engineering when used to describe devices. A September 1945 issue

of Time attributes American radar development to extraordinary “products,” that is,

tangible pieces of technology—not scientific inquiry.190

Other popular news and lifestyle magazines uniformly describe radar as a work of

engineering, making no mention whatsoever of it as a “science.” corroborates

the conclusions of Time.191 Life magazine, boasting four million readers in 1945, similarly celebrates the technological craftsmanship that vanquished the enemy.192

The New York Times further portrays engineering as the heart of wartime microwave radar. On August 15th, the day radar entered the public sphere, six separate

Times articles extol the technology.193 These pieces synopsize wartime radar

187 “Radar: The Industry,” Fortune 32, October 1945, 146 and 200. 188 In fairness, Bell Telephone Laboratories, but only Bell Labs, is listed alongside the Rad Lab as having “contributed most… to radar research and engineering.” “Radar: The Industry,” 200. 189 “Radar.” Time, 46, August 20, 1945, 78-82. 190 “Peacetime Radar.” Time 46, September 24, 1945, 77. 191 Newsweek spotlights the engineering of radar, contending that though German Heinrich Hertz generated electromagnetic waves in 1888 (Newsweek reduces the theory of radar to no more than electromagnetic wave transmission), “the structure of radar was developed” during the war by Americans. “Structure,” suggests an engineer is needed to make pieces work together. “Radar: All-Seeing Eye that Doomed the Enemy.” Newsweek, August 20, 1945, 40. 192 “Radar: Another of the War’s Great Secret Weapons is Revealed.” Life, August 20, 1945, 96-99. Circulation numbers taken from Gerard Piel, The Age of Science: What Scientists Learned in the 20th Century (New York: Basic Books, 2001): xv. 193 These six articles are all available from ProQuest. “Secrets of Radar Given to World,” New York Times, August 15, 1945, pg. 1; “5, 750 Canadians Helped Britain’s Radar Battle,” New York Times, August 15, 1945, pg. 14; “Radar Sank German Ships Hidden 20 Miles Away,” New York Times, August 15, 1945, pg. development as the collaboration of British and American scientists who “pooled their radar knowledge” and subsequently “applied it to all forms of equipment—planes, ships, guns, .”194 In applying their knowledge to differing technical blueprints,

“planes, ships, guns, searchlights,” Rad Lab physicists had adapted their thinking and devices to meet the constraints of human-made machines. They had “developed radar to a finer and finer degree” with every advancing stage of the war; in other words, they had perfected their devices, tinkering with mass-producible artifacts like engineers, not exploring natural principles like physicists.195 Other Times articles from the period highlight these themes as well.196

It would seem then that Americans in August and September 1945 knew that the triumph of radar owed an enormous debt to engineering skills. The media had made that plain.

Or had it? For despite the clear formulations of the Joint Board on Scientific

Information Policy; despite The Saturday Evening Post, Fortune, Newsweek, Time, Life, and the New York Times, in the war’s aftermath, a simultaneous, growing respect for physicists muddled this legacy. Despite the initial public adulation of engineering methods, the attribution of credit for radar quickly grew confused.

Only a few months after the war ended, the notion that the triumph of radar was indebted to physics research had garnered much attention in both public and academic

14; “Radar to be Used on Merchant Ships,” New York Times, August 15, 1945, pg. 14; “World’s Arsenal Revamped by War,” New York Times, August 15, 1945, pg. 14; “Asserts Radar Won Battle of Britain.” New York Times, August 15, 1945, pg. 14. 194 “World’s Arsenal Revamped by War,” pg 14. See also “Secrets of Radar Given to World,” pg 1. 195 “Asserts Radar Won Battle of Britain,” pg 14. 196 A month later, another Times article avers that “American… activities in electronics centered here on the banks of the Charles River.” The Rad Lab is cast a project in “electronics” and “radar devices,” focused on the alteration and harmonization of circuitry—an engineering enterprise. John Stuart, “Radar to Control All Traffic in Air: Devices Which Will Map Airways for Whole Country Were Developed for War Use,” New York Times, September 16, 1945, pg. 32. circles after the war.197 The first sign of already shifting perceptions came in October

1945, when Rad Lab associate director Isaac Isidor Rabi waxed philosophical for the

Atlantic Monthly on the future of physics as well as its past feats. In the article, Rabi

insists, as had the popular magazines, that radar was an accomplishment grounded

primarily in “the inheritance of technology” and the satisfaction of “material human

needs.” These aims stood in contradistinction to “the science of physics proper,” that is,

“our understanding of natural phenomena.”198 Yet, Rabi was distinguishing between physics and engineering precisely because there were many people whom he needed to

refute, individuals who, “with atomic bombs and radar in mind,” wonder “what the physicist… has been doing these past five years, if not physics.”199 Rabi’s disassociation

of physics from radar implies that many people must have given physics credit for radar’s

development. Otherwise, he would not have had to emphasize the fact that the process of uncovering nature’s truths played only a minor role in creating radar. Rabi concludes by stating that “the physicist returns from the war to cultivate his science;”200 thus, no one

should imagine—as apparently many had—that during the half-decade of conflict, the

physicist has had the opportunity to fully practice physics.

In a transformation of credit that Rabi was trying to stymie, Fortune magazine,

which in October 1945 had esteemed the “radar research and engineering” of the Rad

Lab, qualified that praise in November of the same year. In an article on the Rad Lab

197 On the scholarly front, one of the very first histories of radar, M.I.T. in World War II, was published just two years after the Rad Lab closed. In it, author John Burchard unequivocally proclaims that physics truly deserved credit for the development of radar. Burchard maintained that if one asked a “key” Rad Lab researcher to explain the laboratory’s success, he or she would answer that individuals trained in “the fundamentals of physics” were permitted to overturn traditional “engineering practice.” John Burchard, Q.E.D.: M.I.T. in World War II (New York: John Wiley & Sons Inc, 1948), 234. We will see these same (short-sighted) claims reappear in the November 1945 Fortune magazine article below. 198 I. I. Rabi, “The Physicist Returns From The War,” Atlantic Monthly 176, October 1945, 108. 199 Ibid., 108. 200 Ibid., 114. entitled “Longhairs and Short Waves,” that dealt primarily with the exploits of academic

physicists (longhairs), Fortune maintains that the “job was primarily developmental

engineering.”201 But the magazine implies that it was physicists’ departure from—rather

than adherence to—traditional engineering that made their efforts successful. Physicists

has used a “special approach,” had practiced “unorthodox engineering.”202 The article

adds, “Longhairs love to question standard… engineering practice.”203 The traditional

engineering practices of amalgamating components or working within a specific piece of

technology are replaced solely by the physicists’ conception of their own unique

engineering. The physicists are afforded an honored place in the pantheon of thinkers, elevated well above the “standard” engineer.204 The discipline of engineering, quickly, no

longer appeared as important.

Additionally, the Fortune article, like others at the time, pays special attention to

the physics of the magnetron.205 It omits the fact that Slater produced his first paper on

the magnetron half a year after radar had been used to pinpoint submarines, preferring to

venerate the magnetron’s basic physics and the intelligence of the physicists who

understood it.206

201 “Longhairs and Short Waves,” 169. 202 Ibid., 165 and 208. 203 Ibid., 169. 204 Ibid. 205 See also: Randall, 303-323. Randall pays little tribute to engineers. He does state that “In radar, and in the magnetron, there has been a welding together of fundamental ideas and technique by persons from widely different institutions.” Randall, 312. Whether “persons from widely different institutions” means physicists from academia, industry, and government or physicists and engineers from within those three institutions is unclear. -Watt, chairman of The Royal Society of Arts, takes a more diplomatic approach on radar and insists that “It was a cooperation in which the natural philosopher and the engineer in the university” worked together. He recognizes, the “happy story of full interplay between all the contributory factors necessary to the winning of the war.” Randall, 314. Watson-Watt appears to be spreading the credit around to engineers more than Randall, but the reasons for this are unclear. Perhaps Watson-Watt’s position as figurehead requires him to pay greater homage to the engineering community. 206 “Longhairs and Short Waves,” 169. Increasing respect for physicists led directly to inappropriate assignations of

credit. An October 1945 article in Aviation News opens by quoting James Conant,

Harvard president and chairman of the National Defense Research Committee: “This is a

physicists’ war.”207 The article contends that “as we go farther into the new age… the

need for this country to maintain an adequate research program,” will be more thoroughly

appreciated by the government.208 The implication is that since this is a “physicists’ war,”

the miracle of radar demands increased funding for “research programs” in physics. No

suggestion of similarly increasing levels of engineering research is made. The Aviation

News article further dedicates the bulk of its ink to Rad Lab physics, to research that

probed of the “ultra- field,” downplaying the importance of engineering.209 The agency is assigned to the physicist, “the physicist’s war,” but carries

over into the discipline: physics becomes the centerpiece of the article.

This disciplinary obfuscation would recur. In a piece published in January 1946,

Rad Lab director Lee DuBridge argues that postwar physics deserves increased funding

as a result of microwave radar’s legacy. DuBridge, like his associate director Rabi,

understands the dominance of engineering methods at the Rad Lab: “drawing heavily on

the experience of engineers… the Rad Lab physicists… set about developing [the]

components of radar and soon collected and assembled the parts of the first U.S. pulsed

microwave radar set.”210 Yet, DuBridge then heaps inordinate praise upon “high

frequency physics” in which “ a vast amount of research… had to be undertaken.”211

207 “Radiation Laboratory Record Forecasts Electronic Advances: Wartime success of cooperative scientific enterprise at M.I.T. kept nation ahead in radar research,” Aviation News 4, no. 28, October 8, 1945, 28. 208 Ibid. 209 Ibid. 210 Note how the Rad Lab’s mission is understood, as in chapter one, as an amalgamation of components. DuBridge, “History,” 3. 211 Ibid. The physicists had proven themselves worthy engineers. Ibid., 5. Certainly, physicists conducted a great deal of research into basic microwave

theory, but they conducted an enormous amount of research into every aspect of

microwave radar. The closest thing to an official history of microwave radar, Radar in

World War II, written by an historian who spent two years in the Rad Lab, dedicates a

mere eight of its 1100 pages to the theoretical group.212 The theory group did not even

come into being until 1942, well after the first microwave radar set spotted a submarine

in March 1941. “High frequency radar” physics research was undertaken, but for radar’s

chief success, it certainly did not have to be.

DuBridge concludes that the Rad Lab’s successes demonstrate that physicists’

postwar “basic research” can be “enormously more effective given adequate funds for

equipment and technical assistance.”213 He suggests the esteem physicists had earned for

their engineering work ought now to translate into “a new era of progress” in American physics, that “basic research” should receive increased funds. In other words, because it was physicists who did the work, physics research deserves funding—even though those physicists had heavily drawn from engineering techniques. By crediting physicists for their war work and then suggesting that physics research deserves funding as a result,

DuBridge insinuates that the work of the Rad Lab itself was rooted in physics. He is, like the Aviation News article, confusing the practitioners of a discipline with the discipline itself. We’ll return to the results of this association between practitioner and practice in section IV.

Almost no one in the immediate postwar period argued that physics methods alone had brought about radar. Instead, physics began receiving credit in a more oblique

212 The Theoretical Group presence in Guerlac’s Radar In World War II is principally confined to the article by Heins, 625-632. 213 DuBridge, “History,” 5. manner: by achieving postwar renown. Physicists were identified as a unique breed that

had brought about astonishing new technologies, and in doing so, attained celebrity. As

physicists returned to their laboratories to conduct physics, the natural conclusion of the

fawning public was that the skills of this adroit group must be responsible for radar: since

they were practitioners of physics, it must be physics they had practiced during the war.

DuBridge had played upon and reinforced this exact point: physicists are miracle

workers; they should continue to be funded after the war. But if they are being funded

postwar for physics, the observer infers that it must be because physics proved inordinately helpful during the war; hence, physics was responsible for the invention of radar.

II. Popular Opinion and the Indomitable Postwar Physics

During the postwar era, the stature of both physicists and their discipline surged.214 The G.I. Bill and the publicized success of both radar and the nuclear bomb

spawned rabid interest.215 One oft-cited piece of evidence is the explosion of physics

class enrollments in the decades after the second World War. From 1945-1951, the

number of physics Ph.D.’s doubled every year and eight months.216 By the mid-1950’s,

214 For the constraints of the military funding placed upon postwar physics see especially Forman, “Behind Quantum Electronics.” For the impact of radar technology on postwar physics developments see Forman “‘Swords into Ploughshares.’” For pedagogical aspects of postwar physics see: Kaiser, Drawing; David Kaiser, “Scientific Manpower, Requisitions, and the Production of American Physicists after World War II,” Historical Studies in the Physical and Biological Sciences 33 (Fall 2002): 131-159. For postwar suspicions of physicists in the context of anticommunism see: Jessica Wang, American Science in an Age of Anxiety (Chapel Hill, NC: University of North Carolina Press, 1999) and David Kaiser, “The Atomic Secret in Red Hands? American Suspicions of Theoretical Physicists During the Early Cold War,” Representations 90 (Spring 2005): 28-60. 215 Kaiser, Drawing, 15. 216 Kaiser, “Scientific Manpower,” 135. physics departments across the nation became veritable factories of human acumen:

churning out 500 doctorates every year, the same total number the United States had

produced in the sixty-five years intervening between the Civil War and the Depression.217

From 1945-1951, however, even as physics boomed twice as rapidly as the average across all other university disciplines, it outpaced engineering enrollments by just twelve percent.218 Though physics departments continuously swelled from 1945-1971—

picking up additional thrusters after the 1957 launch of the Soviet , Sputnik—

engineering departments actually grew faster than physics departments in the same

period: in fact, 23% faster.219 If anything, then, enrollment data suggests that engineering

emerged the mightiest postwar field of study.

Yet it was not enrollment figures, but rather the reverent public attention lavished

upon physicists that proved decisive for the Rad Lab legacy.220 After the war, physicists

were lionized.221 Time, Newsweek, and The New York Times may have focused their

radar articles on engineering, but they simultaneously elevated physicists, imputing to

them mystical powers of imagination. One August 1945 Time article praises radar

physicists for affording “man a sharp sixth sense which projects him into a world where

217 Ibid., 135-6. 218 Ibid., 136. 219 Increased momentum in physics departments after Sputnik is a well-observed phenomenon. Physics department enrollments tripled from the Sputnik launch until the end of the 1960’s. Kaiser, Drawing, 15. John Wheeler and other physicists became so alarmed that they proposed creating a whole new defense laboratory on the model and scale of the wartime laboratories. See Raphael Rosen, “The Argonauts Assemble: The Founding of The Jasons, America’s Remarkably Independent Advisory Group.” Unpublished Manuscript, 2005. Engineering growth outpacing physics is taken from Kaiser, “Scientific Manpower,” 158; calculations from Kaiser’s data are my own. 220 This reverence should not be mistaken for mere bewilderment. The public had long felt that physics demanded so much that it remained beyond the pale of average human understanding. In 1906, The Nation lamented that, “The passing of the scientific amateur is due largely to the fact that science has become too difficult for him.” “Exit the Amateur Scientist.” The Nation 83, August 23, 1906, 159. 221 Forman, “‘Swords into Ploughshares,’” 398. almost any fantasy is possible.”222 The article, which honors the predictions of engineers throughout, still concludes not with the thoughts of engineers but with the celestial dream of “academic scientists”: bouncing “radar echo[es] off the moon.”223 Newsweek follows the same thread, acknowledging the diversity of engineering projects while emphasizing the physicists’ “ingenious” creation of each one.224 In a similar vein, a September 1945

New York Times piece credits Rad Lab physicists for bringing about “miracles of military aviation” and achieving what had been “hitherto deemed impossible.”225

The widespread panegyric of the physicist’s imagination—“fantasy,” firing radar at the moon, “ingenious,” “miracles,” achieving the “impossible”—probably derived chiefly from the euphoria of victory, but whatever its causes, the results were unmistakable. The media, undoubtedly affected by the spectacle of the nuclear bomb, had fashioned a romantic vision of the physicist and made gifted imagination his province, and his province alone.226

With the oracular robes, came public esteem, something in which the field of engineering, despite its bloated academic enrollments, did not share. In the postwar

222 “Radar.” Time, 78. 223 Ibid., 82. 224 “Radar: All-Seeing Eye that Doomed the Enemy,” 42. 225 Stuart, September 16, 1945, pg. 32. 226 Gendered language is used here and elsewhere simply for convenience. Most physicists at the time, were men, thus ‘his’ is used. This was not the first time scientists had been transformed into oracles in the public sphere. In 1902, The Nation asserted, “something of the medieval notion of science as a variant of the black art seems to survive in certain popular modes of speaking of modern scientists. Marconi is a ‘magician’; Edison, a ‘wizard.’ These phrases… express the real mental attitude of millions of honest folk toward science…. Any miracle may come out of” science. “Popular Appreciation of Scientists.” The Nation 74, January 16, 1902, 46. The same “miracle” diction creeps up throughout popular culture in the 20th century. Einstein is portrayed as a wizard. A. Bowdoin Van Riper, Science in the Popular Culture: A Reference Guide (Westport, CT: Greenwood Press, 2002), 68. 92. After the war, this mysticism became the exclusive province of the physicist and physics, not the inventor and his engineering. In popular culture the scientist is also often considered “abstract, distant, and impractical.” See Van Riper, 25, 70-71, 102, 109-110, and 250. Even if this claim expresses a large degree of truth, however, the stereotype would only have increase the public adoration of the physicist in the case of postwar America, because radar, proximity fuses, and the nuclear bomb were anything but “abstract” and “impractical.” world, physicists wore the “tunic of Superman;” they were more affluent, more abundant, and far more renowned than they had ever been.227 Physics laboratories now boasted “a new importance.”228 Physicists were so “the vogue” that “no dinner party [was] a success without at least one” of them.229 Theologians and social scientists vied for their attention.230 They addressed women’s clubs and hobnobbed with the Washington elite.231

At Capitol Hill hearings, they indulged “fascinated legislators with a glimpse into “the most glamorous intellectual work in the nation.”232 A full 83% of Americans stated unequivocally that science made the world better off.233 The physicists had become celebrities, so much so that one physicist noted that perhaps they were “exalted rather too much by the general public.”234

227 “Tunic of Superman,” is taken from Francis Wickware, “Manhattan Project,” Life 19, August 20, 1945, 100. Abundance and affluence is from Silvan Schweber, QED: And the Men Who Made It (Princeton, NJ: Press, 1994), 144. 228 Samuel Allison, “The State of Physics; Or the Perils of Being Important,” Bulletin of the Atomic Scientists, VI (Jan 1950): 4. 229 Kevles, 375, quoting Harper’s contributor, Judge Frank. 230 Ibid., 376. 231 Allison, 2. “In these times any young man who can call himself a nuclear physicist is besieged with requests to speak before women’s clubs on the world political situation.” On spending time in Washington see Kevles, 376. 232 Robert P. Crease and Charles C. Mann, “Gambling with the Future of Physics,” The New York Times, December 5, 1982, pg. SM66. 233 Robert C. Davis, and the Institute for Social Research, University of Michigan, The Public Impact of Science in the Mass Media (Ann Arbor, MI: National Association of Science Writers, 1958), 179. The Sputnik launch had a trifling impact on the general opinions of Americans toward science. Thus the 83% value applies to postwar perspectives more generally. For a look at the historical context of this data see: Jon D. Miller and Kenneth Prewitt, The Measurement of the Attitudes of The U.S. Public Toward Organized Science, (Chicago: National Opinion Research Center for the Nation Science Foundation, 1979), 6-8. Medicine played an enormous role in maintaining scientific prestige throughout the 1950’s and 1960’s: the 1955 cure for polio doubtlessly ranked in the public mind as one of science’s truly extraordinary contributions. In 1958, the public identified improving health as science’s number one “good effect.” Davis and the Institute for Social Research, University of Michigan, 180. Into the 1970’s this trend persisted: improvements in medicine were “by far the greatest benefit that the public believes science and technology have produced.” Science Indicators. Washington D.C.: , National Science Foundation, 1976, 168 and 174. 234 On the physicist’s celebrity see: Jane Gregory and Steve Miller, Science in Public: Communication, Culture, and Credibility (New York: Plenum Press, 1998), 38. The physicist quoted is DuBridge. See DuBridge, “Birth,” 15. This attention was inspired chiefly by events other than radar, notably the nuclear bomb. After its initial blitz of publicity, radar received rapidly diminishing media attention.235 In the ensuing years, on the occasions when journalists investigated radar,

they almost always paid attention solely to .236 As Rad Lab director

DuBridge lamented about microwave radar, “never was a good news story so thoroughly

buried.”237

By contrast, the nuclear bomb captivated all. Even during radar’s brief

rendezvous with popularity, the news of the bomb sent shockwaves though the press,

easily scattering electromagnetic waves off the front pages.238 Seven out of ten

Americans discussed the nuclear bomb with friends and family.239 The exigencies of

postwar politics then kept public attention indelibly fixed on the implications of nuclear

weapons, and, with it, on physicists.

Physicists recognized that their newfound importance stemmed not from the

impressive developments of quantum theory or general relativity, but from their

profession’s military value. As Samuel Allison, the physicist who headed up the wartime

Metallurgical Laboratory, explained, “the principal motivating power behind the

enormously increased support for physics,” was not esteem for the scientific search for

truth, but “the realization that physics and physicists are important for waging war, and

since waging war is important, physics is important.” Rabi declared that physicists had

235 A. J. Meadows and M. M. Hancock-Beaulieu, eds. Front Page Physics: A century of Physics in the News (: Institute of Physics Publishing, 1994), 117 236 Ibid. 237 DuBridge, “Birth,” 3. The radar news had long been planned for release on August 15th, which just happened to end up being V-J Day. “Four thousand members of the radar laboratory at M.I.T. fumed with chagrin as they saw the story of their five years of effort…pushed back into the want-ad pages.” 238 See “Radar,” Time, 78, and the case of Life, “Radar: Another of the War’s Great Secret Weapons is Revealed,” 96. See also: Brown, 465. 239 Social Science Research Council, Committee on Social and Economic Aspects of Atomic Energy. Public Reaction to the Atomic Bomb and World Affairs (Ithaca, NY: Cornell University, 1947), 208. become “a military asset of such value that only with the assurance of peace will society

permit” them to resume the pursuit of “scientific knowledge.”240

With physicists’ new found military “value,” came responsibility and

influence.241 They ascended to key advising positions in government, participated in

defense advisory studies, and became a tight-knit group wielding extraordinary

influence.242 According to James Killian, president of M.I.T. and first Special Assistant

for Science and Technology to president Eisenhower, consistently influential science

advisors numbered no more than 200.243 Jerrold Zacharias, eminent M.I.T. physicist and

Rad Lab alumnus, remarked, “It’s just us boys…. We all know each other.”244 A full one

third of the members of the President’s Science Advisory Committee were professors

from Cambridge, Massachusetts.245

Precisely because physicists had entered the halls of power, the government

directed a disproportionate amount of federal funds to their discipline.246 The Atomic

Energy Commission disbursed money based largely on the advice of high energy

physicists. The situations was analogous, in one observer’s words, to “asking a hungry

cat to make recommendations about the disposition of some cream.”247 Physicists grew fat. They became private consultants to defense agencies. The National Academy of

Sciences flourished under opulent federal contracts, readily acquired thanks to the expert

240 Rabi, 107. 241 March W. White and William Crew, “Physicists In and Following World War II.” American Journal of Physics 18 (1950): 495. 242 Kevles, 376. For more on the exclusivity of advising in the postwar and early Cold War era see: Raphael Rosen, “The Argonauts Assemble,” (unpublished manuscript) 22-24. 243 Kevles, 394. 244 Ibid. 245 Ibid. 246 Ibid., 397. 247 Philip Abelson, “Are the Tame Cats in Charge?” Saturday Review 49, January 1, 1966, 102. advice of its members.248 With ample support, American physicists enjoyed a profusion

of new particle accelerators, leading to exceedingly high productivity.249

In the early 1960’s, the stream of glory and money became a deluge. In 1965, the

federal government showered basic research in physics with more than $320 million,

more than two and a half times the expenditure in 1959.250 In 1965, Americans ranked nuclear physicists third in occupational status, trailing only supreme court justices and

physicians, a full 12 positions higher than nuclear physicists had been in 1947.251

Between 1956 and 1965, American physicists won or shared in eight of the ten physics

Nobel prizes awarded. Moreover, physicists remained the discipline with the greatest

access to power, counting the Kennedy White House as a close ally.252 During the first

two postwar decades, such power came with great import; by 1960, one political scientist

remarked that the United States depended on science more than it had since time

immemorial.253 With physicists enshrined in power, the nation heard what they had to say, and knew about them, not the engineers.

Illimitable funding and public adulation could not, of course, last forever. In the late 1960’s and early 1970’s, nuclear détente and the accompanying defense spending reduction delivered the worst physics job shortage in United States history.254 Money

especially dried up for projects in high energy physics.255 Prestige diminished as well: in

1971, only 37% of Americans held a “very favorable” opinion of scientists.256 The New

248 Kevles, 376. 249 Ibid., 375. 250 Ibid., 387. 251 Ibid., 391. 252 Ibid., 390. 253 Dupré and Lakoff, vii. 254 Kaiser, Drawing, 345. 255 Kevles, 416. 256 Ibid., 399. York Times warned that physics’ hegemony as “glamour King” of the sciences had

reached a bloodless end and its successor, biology, had begin its ascension.257 Demand

for scientists slipped to 44% of what it had been a decade earlier.258 Physicists ceased to

be the principal governmental advisors.259 Physicists, observed the editor of Science, like the rest of scientists, were now “mortals—fairly intelligent, fairly well-meaning, but still merely mortals.”260 After more than two decades, the miracle- and fantasy-makers, the

übermenschen, had rejoined humankind.

But during two decades of predominance, the popularity of physics had

effectively buried engineering. Though well-regarded as a profession, the discipline of

engineering did not hold a fraction of the attention or prestige that physics did. For much

of the public, engineering did not even exist. With no one listening, how could it possibly

win credit for its contributions?

III. The Dark Shadow: Engineering in Disrepute

Engineers have long held second-tier status in the public eye. Even into the 21st century, the American populace holds greater reservations about the safety and value of technology and engineering than about science.261 Historically, engineers existed solely

for military ends, usually functioning as a commanding general’s lackeys.262 Civil

257 “Revolt Against Physics,” New York Times, February 12, 1968, pg. 38. 258 Kaiser, “Scientific Manpower,” 151. 259 Kevles, 424. 260 Philip Abelson, “Additional Sources of Financial and Political Support for Science,” Science, 180, no. 4083 (April 20, 1973): 259. 261 Science and Engineering Indicators (Washington D.C.: National Science Board, 2000), 8-13 [that is, chapter 8, page 13; this is the book’s numbering method]. See also: Science and Engineering Indicators (Washington D.C.: National Science Board, 1996), 7-20; Jon D. Miller, Rafael Pardo, and Fujio Niwa, Public Perceptions of Science and Technology (Bilbao: Fundación BBV, 1997, particularly chapter 5. 262 Vannevar Bush, Box 23, “What is an Engineer,” Vannevar Bush Papers, MC 78. Institute Archives and Special Collections, M.I.T. Libraries. Cambridge, Massachusetts. engineers as a distinct phenomenon did not emerge until the 18th century in Britain, with the foundation in 1770 of The Society of Civil Engineers.263 Not until the mid-19th century did the term “engineer” come to designate members of a group distinct from both

(less admired) artisans and (more admired) scientists.264 In the United States, civil engineering courses appeared in college catalogs only after the completion of the Erie

Canal in 1825.265 Yet by, 1837, when Dennis Mahan published the first American textbook based on the more scientifically rigorous French engineering style, it sold more than 15,000 copies.266 In 1870, 100 engineers graduated annually in America, but by

1914 that number had grown to 4300.267 The were 45,000 engineers in America at the dawn of the twentieth century, and 230,000 by 1930.268

As we have already seen in the case of postwar engineering enrollments, impressive numbers alone do not confer prestige. By the 1870’s, the media already regularly ascribed engineering achievements to physics—much as they would after

World War II, and much to the chagrin of physicists who did not want to be associated

263 Shortly thereafter, The Institution of Civil Engineers was inaugurated. Manegold, 138. 264 Ibid., 137-8. 265 Charles Weiner, “Science and Higher Education,” in Science and Society in the United States, eds. David Van Tassel and Michael Hall (Homewood, IL: The Dorsey Press, 1966), 167. For a fuller history of engineering in America, especially in the early 20th century see Layton Jr., Revolt. For the history of 19th century American engineers see especially: Layton Jr., “Mirror-Image Twins: The Communities of Science and Technology in 19th-Century America,” Technology and Culture 12, no. 4 (October 1971): 562-580. For a history of engineering education see: Bruce Seely, “Research, Engineering, and Science in American Engineering Colleges: 1900-1960,” Technology and Culture 34, no. 2 (April 1993): 344-386; Eugene Ferguson, Engineering and the Mind’s Eye (Cambridge, MA: The M.I.T. Press, 1992), 60-74. As for American engineering societies, The American Society of Civil Engineers was formed in 1852, The American Society of Mechanical Engineers in 1884, the American Institute of Electrical Engineers in 1884, and the American Institute of Chemical Engineers in 1908. The Society for the Promotion of Engineering Education started in 1894. Thomas Hughes, American Genesis (New York: Viking, 1989), 244-245. 266 Layton Jr., “Mirror-Image Twins,” 570. West Point was the first American engineering school. It remodeled itself on the more scientifically rigorous French polytechnic schools in 1818. 267 Hughes, 243-244. 268 When electrical engineering emerged as its own specialty in the late nineteenth century, college enrollments in the subject quickly outstripped those in both civil and mechanical engineering, which were still growing rapidly. By 1928, the new field already had fifty percent higher enrollment than either mechanical or civil engineering. Ibid., 243-244. with the business of practicality.269 During World War I, the Navy refused to consider officers with engineering backgrounds for high command.270 As chapter three will explore further, when the Rad Lab opened, the radar methods of the M.I.T. engineers passed away without use, because “engineering over the country as a whole did not have the intellectual standing that science had.”271

Despite engineering’s wartime fruits and the successful wartime collaboration of engineers and physicists, in the ensuing decades matters did not change.272 By the

1960’s, the “cloak of the high priesthood of science,” had “cast a dark shadow over the profession of engineering.”273 Though the public closely followed developments in science, it displayed little interest in engineering.274 One engineer quipped that the nation suffered from “science worship syndrome.”275 He was right: in the scientific “pecking order, engineers stood at the bottom. The highest prestige went to the more abstract and theoretical fields.”276 Many engineers referred to themselves as “just a peon” or “a

269 Henry Rowland, the Johns Hopkins physics professor lamented that he was “tired of seeing our professors degrading their chairs by the pursuit of applied science.” Kline, 199. 270 Vannevar Bush, Box 23, “What is an Engineer,” Vannevar Bush Papers, MC 78. Institute Archives and Special Collections, M.I.T. Libraries. Cambridge, Massachusetts. 271 “Headstrong scientists from elsewhere,” simply dismissed the work of the engineers. Hazen, 3-29. 272 The situation with regard to engineers in the differed. A full investigation of Soviet science and engineering in this era is well beyond the scope of this thesis, but some points are worth noting: In America, engineers tended to work fewer hours while at the university than physicists. In the Soviet Union, university engineers worked substantially more hours than physicists. These longer hours may have translated into greater respect. Also, the Soviet Union was training, in 1956, twice as many engineers as the United States even though far more Americans went to college than did Soviets. This greater concentration of engineers perhaps also led to increased respect within Soviet society. Soviet ideology’s reverence of the worker may have increased the engineer’s prestige as well. Alexander Korol, Soviet Education for Science and Technology (Cambridge MA: Technology Press of Massachusetts Institute of Technology, 1957), 260 and 401. For a case study in American scientific management and training of engineers and scientists, see: James W. Kuhn, Scientific and Managerial Manpower in Nuclear Industry (New York: Press, 1966). 273 W. R. Marshall, “Science Ain’t Everything,” Chemical Engineering Progress 60 (January 1964): 17. 274 Vannevar Bush, Box 23, “What is an Engineer,” Vannevar Bush Papers, MC 78. Institute Archives and Special Collections, M.I.T. Libraries. Cambridge, Massachusetts. 275 Marshall, 17. 276 Layton Jr., Revolt, 252. factory hand.”277 In the mid-1960’s, when nuclear physicists ranked as the third most well-respected profession, engineers ranked well below them, in the middle of the pack.278 American engineers have consistently endeavored to associate themselves with scientists and thereby raise their social status.279 Their efforts eventually appeared to pay off: by 1982, The New York Times wrote that engineering had “for all practical purposes,” become an “applied science.” Scientifically inept engineers simply went into manufacturing.280

Though the discipline of engineering may have successfully married into the science family in name, it continued to lag in status. Even after the fall of physics in the late 1960’s, early 1970’s, the public held a less favorable opinion of engineers than of scientists. From 1972 to 1974 to 1976, the public evaluation of scientists and engineers fluctuated by about ten percentage points, but the occupation of physicist consistently remained more prestigious than the occupation of engineer.281 In the 1980’s, journalists

277 John Dustin Kemper, The Engineer and His Profession, 2nd ed. (New York: Holt, Rinehart, and Winston, 1975), 90-91. 278 Physicists in general ranked much higher than engineers as well. No specific numbers are given in the papers. Vannevar Bush, Box 23, “What is an Engineer,” Vannevar Bush Papers, MC 78. Institute Archives and Special Collections, M.I.T. Libraries. Cambridge, Massachusetts. 279 Vincenti, What Engineers Know and How They Know It: Aeronautical Studies from Aeronautical History (Baltimore, MD: The Johns Hopkins University Press, 1990) 6. For example, more theoretically- inclined engineers like Josiah Gibbs and Henry Rowland identified with physics because of its status. Layton Jr., “Mirror-Image,” 579. Engineers often use the word “science” to include engineering, while scientists employ “science” solely to designate the investigation of nature. Layton Jr., “American Ideologies,” 689. 280 Edward Fiske, “Engineering School Shortcomings Lead to US Lag.” New York Times, March 28, 1982, pg. HT6. 281 Science Indicators, 1976, 169-170. Engineers, were, however, generally well-esteemed in the public eye, simply not as much as physicists. See also: Kemper, 91-92. Engineers in a 1972 survey, ranked third in prestige, behind only physicians and scientists. These are all the results of surveys and hence are imperfect measurements. The literature on science indicators is rapidly growing. Some especially useful ones include: Yehuda Elkana and others, eds. The Advent of Science Indicators (New York: John Wiley and Sons, 1978), particularly chapter 3 “Can Science by Measured,” by Gerald Holton; Gregory and Miller, particularly the end of chapter 9. For an international perspective on the public’s perceptions of science and engineering and the difficulties of evaluating it, see Meinolf Dierkes and Claudia von Grote, eds., Between Understanding and Trust: The Public, Science and Technology (London: Routledge, 2003), especially chapters 3 by Maria Eduarda Gonçalves and chapter 5 by John Durant and others. writing articles on medicine, technology, or science, turned to scientists 35 percent of the time, and to engineers just 11 percent.282

By 1990, engineering’s relative ignominy endured. Folklore studies scholars investigating the dynamics of modern scientific culture found that engineers were typified as “just ignorant in mathematical matters.”283 In popular jokes, the engineer is arithmetically inept, calculating 2 times 2 as 39.99, unable to find the square root of 4, and believing 9 to be a prime number.284 The engineer “is eager to calculate, but not to think.”285 Engineers generalize too much, and fail to check their work. Mathematicians and physicists insist, moreover, that these are not simply stereotypes, but “characteristics which truly exist.”286 In short, the engineer’s “ways of thought” are shoddy compared to the physicist’s (or the mathematician’s).287

In the decades since World War II, there are innumerable examples of this widespread disrespect for engineering resulting in a loss of recognition. Unfortunately, besides the immediate postwar flurry of articles on the Rad Lab, there is scant primary material on the public’s assignment of credit for radar (though we will examine what evidence there is in section IV). As such, it is worth considering how the public has divvied credit between engineering and physics in the case of other inventions. In doing so, one quickly discovers a pattern.

Chief among the developments for which engineering skills did not receive credit is the nuclear bomb. By 1949, it was gospel in America that “theorists had single-

282 Support Organizations for the Engineering Community, 66. 283 Carolyn Gilkey, “The Physicist, the Mathematician and the Engineer: Scientists and the Professional Slur,” Western Folklore 49, no. 2 (April 1990): 216. 284 Ibid., 216-219. 285 Ibid., 217. 286 Ibid., 220. 287 Ibid., 219-220. handedly built the bomb.”288 Almost no one appreciated the tremendous engineering labor that had gone into its creation: the public lacked an “understanding of the essential and courageous contributions of the engineers.”289

Indeed, engineering had been crucial. T-division, the theoretical physics division,

was the smallest of all divisions at Los Alamos, dwarfed by the Engineering Ordnance

Division, the largest.290 The implosion device for the bomb required

knowledge regarding the metallurgy of lenses, the wiring of precision-timed detonation

circuits, and physical measurements of reaction cross sections—far more than just

nuclear theory.291 Vannevar Bush would insist that “the really toughest parts of the job

were in the hands of chemical engineers, electrical engineers, and mechanical engineers.”292 Bush is probably overzealous, for the physicists also faced complex

assignments, and had to open whole unexplored realms of physics. Nevertheless, the

discipline of engineering did receive almost no credit for its contributions to the

Manhattan Project.

In fact, engineering methods could not receive credit for the nuclear bomb,

because the imperatives of national security deemed the large-scale engineering processes

behind the bomb so important that they had to be kept top secret. By contrast, the general

principles of nuclear physics were not regarded as perilous, and were allowed into the

public sphere, where they attracted attention to their discoverers, namely, physicists.293

288 Kaiser, “Atomic Secret,” 30. 289 Vannevar Bush, Box 23, “What is an Engineer,” Vannevar Bush Papers, MC 78. Institute Archives and Special Collections, M.I.T. Libraries. Cambridge, Massachusetts. 290 Hoddeson et. al., 93. 291 Kaiser, “Atomic Secret,” 32. 292 Vannevar Bush, Box 23, “What is an Engineer,” Vannevar Bush Papers, MC 78. Institute Archives and Special Collections, M.I.T. Libraries. Cambridge, Massachusetts. 293 In the aftermath of the war, the , the official, publicly-available report on the Manhattan Project focused solely on the ideas of physics, especially theoretical physics. These ideas were the only safe In the case of radar, however, such an argument for why physics alone received credit does not apply. The construction of radar as well as the physics underlying it was not strictly classified.294 Rather, what is important to learn from the case of the nuclear bomb is that engineering could disappear from the historical record, and no one in the public sphere would notice.295

The moon landing provides an example in which national security contingencies did not hamper the public’s awareness of engineering, and yet the discipline still went unnoticed.296 In the space program laboratories where hundreds of researchers milled about, “you would barely find a fundamental scientist among them. Instead… you would find engineers working on design, construction, and launching.”297 Yet, after the moon landing of July 21, 1969, “the press hailed it as a great scientific achievement,” crediting

Neil Armstrong’s giant leap to the genius of “NASA scientists.”298 A popular engineering

ones to publicize. Most people who read the Smyth report concluded that theorists had built the bomb. Kaiser, “Atomic Secret,” 33; see especially the work of Schwartz, who first noticed the impact of security constraints on credit due to engineers and other non-physicists who worked on the Manhattan Project: Rebecca Press Schwartz, The Making of the History of the Atomic Bomb: The Smyth Report and the Historiography of the Manhattan Project (Ph.D. dissertation, Princeton University, in preparation). 294 For radar, the methods might not have become public knowledge right away, but magazines could report on the existence of engineering divisions. Anyone can go into the National Archives today and review the methods of creation. By contrast, men with large guns would impede one’s progress during an attempt to read the notebooks of the nuclear bomb engineers at Oak Ridge. 295 “The engineers, upset at their delegation to a supporting role, could do little about it.” Schwartz, draft chapter 3, 31. Los Alamos director Robert Oppenheimer was very angry to learn that the contributions of the Ordnance and Explosives divisions were being downplayed. Schwartz, draft chapter 2, 13. 296 See, for example, “Astronauts Relax and Begin Taping Their Stories,” New York Times, July 28, 1969, pg 18. A complete survey of the post-landing periodicals has been beyond the scope of this paper, but would certainly prove a useful endeavor to a future researcher. Still another example of engineering’s being ignored is found is the reaction to Sputnik. Though Sputnik required the creation of better booster rockets and other components of space technology, American policy makers and eventually the public responded as if the Soviets had challenged American science. Eisenhower tapped Killian to be his advisor to plan a scientific, not a technological response to Sputnik. Donald E. Stokes, Pasteur’s Quadrant: Basic Science and Technological Innovation (Washington, D.C.: Brookings Institution Press, 1997), 54. 297 Bush, Pieces, 54 and Vannevar Bush, Box 23, “What is an Engineer,” Vannevar Bush Papers, MC 78. Institute Archives and Special Collections, M.I.T. Libraries. Cambridge, Massachusetts. From Bush’s October 8, 1964 address at the Worcester Polytechnic Institute Centennial Convocation. 298 Bush, Pieces, 54; Support Organizations for the Engineering Community, 50. saying at the time phrased it bluntly: “Every rocket-firing that is successful is hailed as a scientific achievement; every one that isn’t is regarded as an engineering failure.”299

A particularly intriguing postwar instance of credit denied to the engineer was the investigation of the 2500-year-old, 1000-foot tunnel on the Greek island of Samos.

Initially, historians assumed that Pythagorean principles of geometric theory provided the basis for the construction of an aqueduct though Samos’s Mount Castro.300 In 1965, however, two researchers overturned this explanation, demonstrating that “it is certainly unwise to assume, without more positive evidence, that Pythagoras had anything to do” with the construction the tunnel. Rather, they concluded, the ancient Hellene architect

Eupalinus would have had an easier time constructing his aqueduct by relying on line-of- sight measurements: “the problem facing Eupalinus was one of practical engineering, rather than one of geometrical theory.”301 Historians have displayed an unfounded proclivity for lauding “theory” over “practical engineering.”302

Thus a clear pattern surfaces. In the postwar period, the public, the government, and the scholar recognized and preferred achievements in physics to ones in

299 In both Kemper, 84 and Marshall, 19. Emphasis in original in Kemper. 300 June Goodfield and Stephen Toulmin, “How Was the Tunnel of Eupalinus Aligned?” in Philosophers and Machines, ed. Otto Mayr (New York: Science History Publications, 1976), 38. 301 Ibid., 46. “Heroes often get more credit than they deserve, and Pythagoras is no exception.” Ibid., 46. 302 Ibid. Goodfield and Toulmin certainly have their detractors. In response to their article, B. L. van der Waerden defended his initial assertions about Pythagorean theory and the Samos tunnel. B. L. van der Waerden, “Eupalinos and his Tunnel,” in Philosophers and Machines, ed. Otto Mayr (New York: Science History Publications, 1976), 48-49. Still, no one knows by what methods Eupalinus constructed his tunnel. Alfred Burns, a proponent of the Pythagorean explanation, nevertheless concluded that “there is no clear evidence” whether the tunnel was the work of engineering or Pythagorean theory: “both methods are possible.” Alfred Burns, “The Tunnel of Eupalinus and the Tunnel Problem of Hero of Alexandria,” in Philosophers and Machines, ed. Otto Mayr (New York: Science History Publications, 1976), 63. Goodfield and Toulmin’s argument thus clearly reveals that historians themselves, like van der Waerden, can be guilty of a predilection for explaining inventions theoretically rather than technically. engineering.303 Engineers did not receive even partial credit for the nuclear bomb, for the moon landing, or for myriad other contributions. That the discipline of engineering could be deprived of credit for radar is thus not surprising. Engineers were simply not as valuable as physicists.304

One final point ought to be made. Much of the public, to this day, does not appreciate the difference between science and engineering.305 Vannevar Bush, perhaps the greatest postwar expert on the relationship of science to the nation, certainly believed the public was confused.306 One sees evidence of the conflation of science and engineering everywhere from movies to opinion polls.307 In 1976, for example, among the lowest income bracket in America, 27 percent had no opinion about which was more important to control: science or technology. Twenty-six percent of rural respondents also had no opinion, accompanied by 26 percent of those over 60, and 23 percent of those who had not completed high school. The opinion poll concluded, “these groups are uncertain about the difference between science and technology.”308 The National Academy of

Engineering maintained that “most Americans have little idea of what engineers are or

303 Even as recently as 1985, the Panel on Support Organizations for the Engineering Community proclaimed, “the engineering community wants recognition for its contributions to society.” Support Organizations for the Engineering Community, 50. 304 Chemical engineer W. R. Marshall remarked in 1964 that, “the general impression of engineering has fallen” to a “low ebb.” Marshall, 19. 305 “Large segments of the public do not clearly separate science from technology.” Amitai Etzioni and Clyde Nunn, “The Public Appreciation of Science in Contemporary America,” Daedalus 103, no. 3 (Summer 1974): 195. 306 Vannevar Bush, Box 23, “What is an Engineer,” Vannevar Bush Papers, MC 78. Institute Archives and Special Collections, M.I.T. Libraries. Cambridge, Massachusetts.. 307 Christopher Frayling, Mad, Bad and Dangerous?: The Scientist and the Cinema (London: Reaktion Books, 2005), 7. Bruce Lewenstein similarly argued that the public culture does not differentiate between science and engineering. He further adds that one of the great problems facing engineers was that it was not perceived in popular culture as “a glamorous profession.” Bruce Lewenstein, “Frankenstein or Wizard: Images of Engineers in the Mass Media,” Engineering: Cornell Quarterly 24, no. 1 (1989): 40, 46. 308 Science Indicators, 1976, 182. what they actually do.”309 Moreover, and more substantially for the legacy of credit,

“journalists often consider the roles of engineers and scientists to be interchangeable.”310

The engineering community’s problem therefore was twofold: physics was vastly

more popular, and the public as well as the media did not care to tell engineering apart

from physics.

This simultaneous conflation of engineering with physics and deprivation of

credit from the engineers at the expense of physicists is not as paradoxical as it might

seem. A discipline that the public is uninterested in, and does not differentiate, cannot be recognized in its own right. In such a situation, radar, like these other technologies, could easily have been bumped from being an achievement that relied critically upon engineering, to one that solely grew out of physics. With physics exalted in the postwar era, this is precisely what unfolded.

IV. The Adulterated Rad Lab Legacy

In 1945, Lee DuBridge showed the path for things to come. Rad Lab physicists,

asserted DuBridge, believed that “the prestige” they “gained on war projects” would

“make it possible for American physics to enter upon a new era of progress and effectiveness.”311 Indeed, the government funneled incredible sums into physics research

for some 25 years. Why were these funds diverted to the study of physics? Because of the

309 Support Organizations for the Engineering Community, 49. 310 Ibid., 50. 311 DuBridge, “History,” 5. physicists wartime work. What did they do during the war? One would infer that it must

have been physics.312

Five years later, physicist Samuel Allison agreed. In a piece for the Bulletin of

Atomic Scientists, Allison soberly recalled the inventions of the second World War—

radar, proximity fuses, and the nuclear bomb—that had brought physicists so much

esteem while concomitantly inaugurating an era of weapons of mass slaughter. Allison

concluded that “no one in his right mind… would state that these things should not have

been done,” even though bringing forth these weapons, “would associate physics and war

in the popular mind.”313 As far as Allison was concerned, in the public eye, physics had

received credit for these inventions, including radar. Allison adds that after the war

physics needed to justify itself “as a worthy intellectual activity of a world at peace.”314

In the popular mind, physics had become irreversibly cast as the source of modern weaponry.

A year later, physicist and former Rad Lab member, reached an identical conclusion. Examining the postwar relationship between the military and physicists, he deemed their excessive collaboration extremely dangerous because “in the years to come, science and the scientists [risk becoming] identified in the public minds

[only] as the wizardry and wizards who… made the fantastic new weapons of mass destruction.”315 Though Condon refers here to futuristic weaponry, his statement applies equally well to the armaments of World War II: in the immediate aftermath of the war,

312 Also, as Schwartz has demonstrated, the Smyth Report necessitated emphasizing physics in the development of the nuclear bomb. It was all the public was permitted to know about. See Schwartz. 313 Allison, 2. In describing wartime physics, Allison casts every project as primarily a project of physics, with radar: “the generation, reflection, and detection of short-wave electromagnetic radiation.” 314 Ibid., 27. 315 Edward Condon, “Some Thoughts on Science in the Federal Government.” Physics Today 5 (April 1952): 9. the public identified scientists as ministers of “fantasy,” and linked their wartime creations with their science, their “wizardry.” Radar was one of these wartime achievements; it had become associated with physics.

Granting authority to inventors rather than to the discipline they practice is a common human action. As French philosopher Michel Foucault has argued in the context of analyzing history: genres and schools of philosophy, “seem relatively weak, secondary, and superimposed scansions in comparison with the solid and fundamental unit of the author.”316 His analysis is easily extended to the authorship of technology.317

The application of credit to an individual creator over a discipline is a modern instinct, but a potent one.318 As both Foucault and Roland Barthes have argued, authors “impose a limit” on potential interpretations; they delineate “the edges” of possibilities.319 A human author or authors, unlike disciplines whose boundaries are very obviously murky, allow for the creation of clear classifications.320 By virtue of being an easily identifiable actor,

316 Michel Foucault, “What is an Author?” in Textual Strategies: Perspectives in Post-Structuralist Criticism, ed. Josué V. Harari (Ithaca, NY: Cornell University Press, 1979), 141. 317 Foucault explicitly states that his analysis does not apply to modern science, for which he asserts that the discoverers receive credit only to “christen” the name of a certain theorem or effect. Foucault, 149. He does not, however, apply such to the creation of artifacts, for which it seems entirely possible for an author to take credit as an individual Indeed, Robert Watson-Watt is the self-proclaimed ‘father of radar.’ “In Britain, at least, Sir Robert Watson-Watt is widely recognized as the sole inventor of radar.” Authority is easily ascribed to an individual because this a piece of technology, and Watson-Watt managed to associate it with his name, more easily than it could be associated with, for example, say physics or engineering. John Waller, Leaps in the Dark (Oxford: Oxford University Press, 2004), 219-240. A thorough extension of Foucault’s modern “author-function” from literature to technology is well beyond the scope of this thesis. It is only posited here. In support of treating radar as a technology whose authorship fits patterns of authorship common in history or literature, following the war, great patent battles over radar ensued. It was clearly a technology for which individual actors (as well as companies) sought credit. See “Radar: The Technique,” Fortune and especially “Radar: The Industry,” Fortune. 318 The “author-function” has varied in different times and places. Only its modern, post-seventeenth century incarnation is considered in Foucault. Foucault, 158. 319 Roland Barthes, “The Death of the Author,” in Image-Music-Text, ed. and trans. Stephen Heath (London: Fontana, 1977), 147; Foucault, 147. Foucault expounds that “authorial intention” limits what can and cannot be said. Foucault, 159. 320 Foucault, 147. Historian of science Mario Biagioli argues that authorship is not simply an idea resulting from credit systems but that the idea of the author makes possible the very idea of keeping track of credit. Without the author, there can be no credit: “Authorship is both accounted and accounting.” Mario Biagioli, the author, the inventor, therefore becomes the full-blown executor of the technology.321

The skills and contexts behind the creation of, in this case, radar, are too nebulous and pass unnoticed in the modern world; only the creators, the physicists, receive attention.

Had society afforded systems of thought primacy in credit, then engineering likely would have received its fair share of microwave radar’s glory.

Though the discipline of engineering had clearly lost credit for radar by 1950, it is nevertheless possible to imagine that such credit was restored after physics’ popularity declined in the 1970’s. To resolve this query, the remainder of this section will consider the memory of the Rad Lab in the last half century. The answer that we will find is unambiguous.

In serious books focused on the Rad Lab, historians generally identify engineering as the laboratory’s central activity.322 These books tend to be monographs that thoroughly consider the actual processes of creation at the Rad Lab and, as a result, recognize the physicists’ constant use of engineering skills. These books were the ones cited in the beginning of chapter one.323

“Aporias of Scientific Authorship,” in The Science Studies Reader, ed. Mario Biagioli (New York: Routledge, 1999), 26. For more on the credibility of science in relation to the public see Bruno Latour and Steve Woolgar “The Cycle of Credibility,” in Science in Context: Readings in the Sociology of Science, eds. Barry Barnes and David Edge (Milton Keynes: Open University Press, 1982): 35-43. 321 See Barthes, 142-144, 148. Jacques Derrida argues similarly, contending that that all structures, especially philosophy, rely on “a center.” This center limits possibilities of meaning and creativity (play) in a system. In the case of radar, the authorship of the physicists, limits the other potential meanings, namely giving credit to other bodies, for example, engineering. Jacques Derrida, Writing and Difference, trans. Alan Bass (London: Routledge Classics, 2001), 352-370. One can see the connection between author and authority, in addition to linguistically, as far back as the Bible. See: Exodus 24: 12-18, 31:18, 32: 15-16, and Deuteronomy 4:13. 322 So too, do the book reviews of these books. For example, see Matthew Wald, “Jam Sessions,” New York Times, June 22, 1997, p. BR31, a book review of Buderi’s The Invention that Changed the World, where the Rad Lab is described as a “powerhouse of radar engineering.” By contrast, the review of Tuxedo Park, falls back on describing radar’s engineering as “development.” Alex Beam, “Basement Science Project,” New York Times, June 16, 2002, pg. F14. 323 Buderi’s The Invention that Changed the World and Brown’s Radar in World War II are clear examples. On the other hand, mainstream histories, including newspaper articles, do not

recognize the centrality of engineering.324 Like the incautious wader-out, most newspaper

articles going into some depth about the Rad Lab, quickly go out of their depth themselves: consistently attributing the lab’s achievements solely to physics.325 Rad Lab

engineer William Tuller’s 1954 obituary cites him for “his scientific contributions during

World War II that led to the production of many types of microwave radar,” ignoring

Tuller’s primary work as an engineer.326 In 1968, an article on the boom in science and technology along Boston’s route 128 dedicates several paragraphs to M.I.T.’s history, contending that the Institute’s emphasis on “basic-science” prepared it for the role it played in “developing radar devices at its famous Radiation Laboratory.”327 A 1982 New

York Times piece on a pioneering new particle accelerator insists that “science”—not engineering—“had helped win the war with such powerful inventions as radar and the atomic bomb,” then states: “The now demobilized scientists asked for advanced tools to

further research into the frontiers of physics. For the next two decades, they received

more or less what they requested.”328 Having equated wartime “radar” with “science,”

The Times science writers then argue that it was natural for “demobilized scientists”—

who brought forth “powerful inventions”—to receive rewards that allowed them to

324 Any periodical search for “radar” and “World War II” will yield roughly 90% of articles referring to “the development of radar during World War II.” The word “development” is a fair description, but as for what skills were used on an average day at the Rad Lab, it leaves much to the imagination. See: “Dr. Henry Straus, Radar Expert, 43: Lincoln Laboratory Physicist Dies—Aided Development of Detecting Device,” New York Times, September 24, 1957, pg. 35 and John Fenton, “M.I.T. offers Spur to Area Economy,” New York Times, January 15, 1961, pg. 72; Neil Mermelstein, “How Food Technology Covered Microwaves Over the Years,” Food Technology 51 (May 1997): 82. 325 The phrase, “incautious wader out” taken from Herman Melville, The Confidence-Man (New York: Penguin Books, 1990), 239, chap. 39. 326 “Tuller, Alexandria Executive, Feared Dead in KLM Crash,” and Times Herald, September 6, 1954, pg. 7. Tuller’s scientific research was largely irrelevant. His did engage in some “basic research,” but his primary work was engineering receivers. As the war advanced, he even worked for Raytheon corporation to help engineer 10-cm ship-searching sets for the Navy. Guerlac, Radar, 399. 327 Henry Lieberman, “Technology: Alchemy of Route 128,” New York Times, January 8, 1968, pg. 139. 328 Crease and Mann, pg. SM66. conduct basic science on physics’ “frontiers.” This is the same author-centered

connection DuBridge had first articulated: linking physicists wartime performance with

the support of postwar physics itself.

Three years later, in a tribute to Rabi, The New York Times remembers him as

“scientific director of the Radiation Laboratory,… which made the crucial achievements

of radar.”329 The article indirectly associates radar with science. It also misidentifies this

“statesmen of science” as having been “scientific director” when his title was Associate

Director and he oversaw a massive engineering enterprise.330

In 1998, when M.I.T. tore down , one of the Rad Lab’s largest

structures, a physicist interviewed in a Times article lamented, “there are two prominent

shrines of the triumph of science during the war: One is… Los Alamos;… the other is the

plywood building where radar was invented.”331 The article’s author not only makes no

comment on the remark but also himself salutes the building as “important to science.”332

This may seem like mere layman’s terms that are meant to include engineering, but the failure to make a distinction for engineering’s inimitable skills ties into to the discipline’s invisibility. Recall from section III that “journalists often consider the roles of engineers and scientists to be interchangeable,” and that, largely as a result, “most Americans have little idea of what engineers are or what they actually do.”333 This obscurity means that in

contemporary periodicals, when the Rad Lab is recalled without careful scrutiny, it is as a

329 Gleick, pg. B1. 330 Ibid. 331 Hilts, pg. F4. Emphasis mine. 332 Ibid. Emphasis mine. Scientific investigations were carried out in Building 20 after the war, but Hilts groups radar among these scientific achievements. 333 See notes 309 and 310. bastion of physics and science; thus, the public does not hear about the unique importance of engineering skills.

Many mainstream history books do little better. A recent biography of Vannevar

Bush attests, “it wasn’t just Allied equipment that helped to defeat the U-boat; it was the

minds of the scientists themselves.”334 The scientists’ “minds”—and their “methods”—

brought victory.335 Certainly, the Rad Lad physicists employed scientific expertise, but

this biography completely dispenses with the physicists’ engineering techniques.

A biography of Rabi makes still grosser errors. It depicts younger physicists’ Rad

Lab experiences as “apprenticeship[s]” in physics, not engineering.336 The biographer also incorrectly describes the development of the transmitter-receiver box as “a good example of how the Rad Lab physicists were able to apply the principles of their discipline in a systematic fashion to effect a solution.”337 The Rad Lab physicists

struggled with the problem of the TR box precisely because it demanded engineering

expertise, yet the biographer neglects Jim Lawson’s extensive amateur radio engineering

background, which had proved crucial for success.338 The problem with the TR box was

that they were physicists. The solution was that they became engineers.

Contemporary chronologies of science intended for mainstream audiences similarly distort the importance of physics. While a few encyclopedias at least partially

recognize the importance of engineering to radar, most miss this point entirely.339 The

334 Zachary, 171. 335 Ibid. 336 Rigden 166. 337 Ibid., 136. 338 The biographer mentions Lawson, but says nothing about his background or methods and clearly considers the TR box a triumph of physics. See Ibid, 136. 339 For example, The Oxford Companion to the History of Modern Science emphasizes both the scientific and engineering talent required in developing radar sets. The Oxford Companion does not stress, however, the primacy of one over the other, but unlike nearly all other encyclopedias, it recognizes the centrality of History of Science in the United States: An Encyclopedia casts radar’s World War II

development as an achievement of physics. It praises the “creativity” of the physicists, but makes no mention of their applying this inventiveness in the form of engineering skills.340 The History of Science further claims that the construction of “particular” radar

“system[s]” was irrelevant to the practices of Rad Lab physicists, even though working

within the constraints of existing components—an especially important aspect of

engineering—was central to the entire enterprise.341

Wikipedia, an online, user-edited encyclopedia provides an ideal source for

identifying mainstream conceptions of the Rad Lab. Since Wikipedia articles are written

and edited by anyone, including non-experts, it is a reasonable proxy for what the general

population accepts as historical truth. As of March 5, 2006, in describing the Rad Lab’s

research, Wikipedia states, “The lab's activities eventually encompassed physical

electronics, electromagnetic properties of matter, microwave physics, and microwave

communication principles.”342 Physics—“microwave physics,” “properties of matter,”

and “microwave communication principles”—garners almost all of the attention.

“Physical electronics” is the one subject that could be construed as engineering, and even

that is couched in murky disciplinary language.343 The Rad Lab is presented as having

engineering efforts. The Oxford Companion to the History of Modern Science, s.v. “Radar,” by Woodruff T. Sullivan III, ed. J. L. Heilbron (New York: Oxford University Press, 2003). 340 The History of Science in the United States: An Encyclopedia, s.v. “World War II and science,” by Joel Genuth, ed. Marc Rothenberg (New York: Garland, 2001). 341 Ibid. 342 “Radiation Laboratory,” Wikipedia: The Free Encyclopedia, (accessed 17 January 2006, 12 February 2006 and 5 March 2006). On all dates accessed the information had not changed. 343 “Physical electronics” could refer to tinkering with electronics or to the mastery of the theoretical, “physical,” properties of electronic circuitry. Both took place at the Rad Lab. The Wikipedia entry recognizes the “technological efforts” of the Anglo-American alliance, but this technological effort is not synonymous with engineering efforts. Radar is obviously a technology as it was actually used in war. Calling it a technological effort is essentially self-evident and by no means evinces whether engineering, physics, or zoology proved necessary for the creation of radar. Ibid. made “fundamental advances in all [four] of these fields.” Indeed, it did, but the real

progress lay in engineering, not in “fields” of science. In the article, engineering itself

receives no mention.344

The Reader’s Guide to the History of Science, a chronology, also ignores

engineering, pronouncing that the federal government “enlisted physics” for its wartime

work on radar.345 In the case of The Reader’s Guide, crediting physics is the result of

lumping together radar and the nuclear bomb, both of which are listed in the same

sentence as physics achievements. Once again, the alleged centrality of physics to the

latter spreads to perceptions of the former.

Science in the Early Twentieth Century, an encyclopedia from 2005, makes no

reference to engineering and terms radar a “scientific accomplishment,” labeling radar its

own “scientific field.”346 Most radar textbooks published in last two decades, in their

344 The article does mention that the Rad Lab produced over 100 different radar systems, though. Ibid. 345 Reader’s Guide to the History of Science, s.v. “Physics—20th Century,” by David Kaiser, ed. Arne Hessenbruch (Chicago: Fitzroy Dearborn, 2000). 346 Science in the Early Twentieth Century: An Encyclopedia, s.v. “Radar,” ed. Jacob Darwin Hamblin (Santa Barbara, CA: ABC-CLIO, 2005). brief history sections, fail to mention engineering as well.347 Even engineers reflecting on wartime microwave radar have committed the same blunder.348

Professional historians not dealing directly with radar are often no more accurate, generally lavishing disproportionate attention upon academic scientists and academic science research.349 Tellingly, certain historians of physics identify their “field with all of science.”350 Historians briefly mentioning the Rad Lab may recognize engineering, but almost invariably give primacy to physics.351 Histories of the nuclear bomb do the same.352

347 Those radar textbooks that treat the history of radar usually do so in a brief introductory paragraph. I came across no radar textbooks—save the original Rad Lab series—that make any mention of engineering’s historical importance. One relatively recent radar textbook notes that radar must be thought of not simply as a piece of technology but “a mature scientific discipline with considerable theoretical foundations.” Nadav Levanon, Radar Principles, (New York: John Wiley, 1988), vii. Another recent textbook is dedicated not the to numerous engineers or physicists who worked on radar, but solely to the most abstract theoretical founder possible, the stipulator of electromagnetic waves himself: “, whose incredible insight founded all wave-based technology after his time.” The introduction to the history focuses on Maxwell and Hertz for their work on electromagnetic wave propagation—again, two physicists who did nothing for radar development itself. All of radar is cast as a form of something “discovered.” As for the crucial breakthroughs made at the Rad Lab, the textbook notes with lackluster, “Radar development was accelerated during World War II.” Peyton Z. Peebles, Radar Principles (New York: John Wiley, 1998), v and 1. Ironically, almost all radar textbooks are stored—at Harvard, at least—in the engineering library. 348 W. R. Marshall, in attacking the predominance of physics still lauds, “the scientific achievements of World War II.” “Science,” as Marshall explicitly states in his article, does not in any way include engineering. Marshall, 17-18. 349 “The twentieth century belief that ‘Science implies the breaking of new ground,’ had made the history of science and expertise the history of research. Intellectuals, even engineers like Vannevar Bush, generally used ‘science’… [to] refer to scientific research.” Emphasis in original. David Edgerton, “‘The Linear Model’ Did not Exist: Reflections on the History and Historiography of Science and Research in Industry in the Twentieth Century,” in The Science-Industry Nexus: History, Policy, Implications eds. Karl Grandin and Nina Wormbs (New York: Watson, 2004), 44-46. 350 Nathan Reingold, “Physics and Engineering in the United States, 1945-1965, A Study of Pride and Prejudice,” in The Michelson Era in American Science 1870-1930, ed. Rita G. Lerner (New York: American Institute of Physics: 1988), 288. 351 Yale historian Daniel Kevles spotlights the general “behavior of electromagnetic radiation at the frequencies of microwave radar.” Kevles, 353. Couched in the terms of fundamental physics, this is portrait a physicist would offer. Writing a treatise on the relations of basic science and engineering, Donald Stokes, though not a historian, still produced a detailed, academic work in which he describes radar as a “scientific success of World War II.” Elsewhere in the book, Stokes had distinguished clearly between science and engineering. Stokes, 103. Other historians do pay fair attention to both physics and engineering at the Rad Lab. See, for example: Schweber, QED, 136-141. 352 Again, part of this results from the Smyth report itself and the exigencies of postwar politics. Nevertheless, even latter-day historians pay engineering less tribute than physics. The entire literature is

These historiographical errors should come as no surprise. Though periodicals in the immediate postwar period initially hailed engineering for its contributions to microwave radar, this credit soon grew muddled. Physicists, on the heels of the spectacle of nuclear success, became national celebrities. They then successfully translated their wartime laurels into expansive new programs in physics and power on policy matters. As

DuBridge, Allison, Condon and Rabi testify, the unequivocal glorification of physics and physicists and their suckling at the federal bosom ossified the link in the public mind between physicists’ wartime achievement and what would be inferred to be wartime physics. When people thought about World War II, they thought of physicists, then immediately of physics itself.

Meanwhile, engineers remained mired in public mediocrity, unidentifiable to the media and the populace and easily ignorable interlopers in policy realms. The engineering community had neither the authority nor the mechanism to correct the misinformation. Only the physicists themselves—who understood that their wartime roles had been predominantly been as engineers, and who had the public’s ear—could have set the record straight. But they raised no alarm. So the natural question is: why not?

The answer, which the final chapter will explore, lies in the internal relations of physicists and engineers. To physicists, the discipline of engineering contained no knowledge except that which it borrowed from science. It was not a creative force, but

enormous, but consider two examples. One recent history summarizes the Manhattan Project as having used “a scientific approach.” The book concludes by quoting Freeman Dyson: “It was a shared ambition to do great things together in science without any personal jealousies or squabbles over credit.” Emphasis mine. Hoddeson et. al., 417. Rhodes also closes with a homily on science, declaring that “The preeminent transnational community in our culture is science…. with the release of nuclear energy that model commonwealth decisively challenged the power of the nation-state.” Rhodes, 788. See also Kevles, 353. manufacturing rooted in scientific principles. When physicists admitted that they engineered, they meant only that they had produced.353 In thinking about their inventive breakthroughs at the Rad Lab, physicists believed their unique knowledge as physicists deserved credit for radar’s breakthroughs. They never considered the possibility that engineering skills might be responsible for their progress, never grasped that they had employed modes of knowledge far better classified under the rubric of engineering than of physics.

By considering what practitioners of physics thought of those in engineering, the crediting of physics for microwave radar can be understood at its deepest roots: blindness

about the epistemological divergences of physicists and engineers.

353 We saw this admission from DuBridge and Rabi in section I; Dexter Masters, author of The Saturday Evening Post article also explored in Section I, was another Rad Lab researcher making plain that he worked on engineering. For brief biography of Masters see: Masters, 20.

Chapter Three Engineering Divergent

“Sleep hath its own world, And a wide realm of wild reality, And dreams in their development have breath, And tears, and tortures, and the touch of joy;” -Lord Byron354

Engineering, like sleep, has is its own world, its own breath and struggles. The public’s conceptions of engineering and physics are only possible because of the internal relations of physicists and engineers themselves. That is to say, the misallocation of credit for radar reflects the desire of physicists to dissociate themselves from the practice of engineering.355 Physicists’ desire to dissociate—this distaste for engineering—emerges from their discipline’s internal philosophy: physics’ dogma has long held that engineering is not a creative endeavor endowed with its own forms of knowledge, with endemic modes of thought.

The differing uses of the term “knowledge” reflect this view of engineering as completely devoid of its own thoughts. Usually, “engineering knowledge,” is taken to mean knowledge employed by engineers, that is, information they consume. Scientific knowledge, on the other hand, is the canon of information produced by science and held by the world at large.356 The former is a means to a practical end; the latter is an end unto

354 Lord George Gordon Byron, “The Dream” in The Complete Work of Lord Byron, ed. John Galt (Paris: Baudry’s European Library, 1837), 319, lines 3-6. 355 It is worth noting, that engineers have been charged with the same offense: refusing to give credit to the knowledge of skilled workers (artisans) such as machinists, electricians, riggers, millwrights, carpenters, welders, and the like. Ferguson, 58-59. 356 Vincenti, What Engineers Know, 228. itself. Given that dichotomy, is it surprising that Rad Lab physicists chose to be known as

contributors to society rather than mere users of it?

Casting the Rad Lab’s legacy as the story of “physicists practicing engineering”

therefore misses the point: this is simultaneously the story of physicists refusing to be

regarded too much as engineers. It is the story of physicists consigning a profession to the ashcan of history.

This final chapter dives into the cinders. Section I retells the very same Rad Lab story from the beginning—only differently. Instead of focusing on methods and components as in chapter one, this section probes the relationships between engineers and physicists at the lab. Physicists’ scorn for engineering resulted from a philosophy that divided the disciplines along inappropriate and inaccurate grounds: maintaining that

engineering derives from science and, as such, contains no independent modes of

thought. Section II explores the orthodoxy of this philosophy.

Section III then dismantles it, using the work of historians of science who in the

last thirty-five years have begun to scrupulously investigate the epistemology of

engineering: what it is that engineers know and do. The idea of design—creating the

maximally efficient model of a machine—will be offered as a key example of a unique

form of engineering knowledge.

This epistemology of engineering is not merely a theoretical exercise;

engineering, it will be argued, truly does have its own indigenous systems of knowledge.

This knowledge has a crucial impact on how technology is produced and how science

itself progresses. At the Rad Lab, physicists failed to appreciate this creative force of

engineering. Only when the discipline of physics recognizes the intellectual independence of its artisanal cousin, will physicists ever admit that they engaged in engineering.357 Only then can historical credit—for microwave radar or space flight or technologies still in development—be appropriately allocated.

I. Physicist versus Engineer at the Rad Lab and

Beyond

Spearheaded by the charismatic Ernest Lawrence, the recruiting effort for the Rad

Lab zeroed in on physicists from its very inception.358 Lawrence personally selected

Bainbridge, Alvarez, Edwin McMillan, and DuBridge.359 He generated a list of invitees to the October 1940 American Physical Society’s Boston meeting and then cherry-picked from among some 600 physicists.360 Nuclear physicists organized the laboratory, and in doing so invited almost exclusively physicists to join their ranks. The very name M.I.T.

Radiation Laboratory came as tribute to Lawrence’s physicist-led Berkeley Radiation

Laboratory.361

When the physicists arrived in November 1940, M.I.T. already housed an active microwave engineering program, generously endowed by science tycoon Alfred

Loomis.362 But the Rad Lab physicists paid the department no heed. As the chair of

357 This entails recognizing engineering as more than simply a helpful discipline for the construction of technologies that advance physics (like particle accelerators), but as a creative discipline in its own right. 358 In England, Taffy Bowen’s team at the British Research Establishment (AMRE) recruited predominantly physicists as well, including: Robert Hanbury Brown, , and . Buderi, 80 and 46. See also Kevles, 303. Buderi memorializes these physicists as a “talented core of scientists.” 359 Conant, 202. Lawrence generated so much enthusiasm that by November 1940, “roughly one eminent scientist was joining the staff every day.” Edward Bowen, “The Tizard Mission to the USA and Canada,” in Radar Development to 1945, ed. Russell Burns, 296-307 (London: Peter Peregrinus Ltd., 1988), 303. 360 Conant, 202. 361 This name, of course, also had a deceptive convenience. See note 34. 362 Stuart Leslie, “Profit and Loss: The Military and M.I.T. in the Postwar Era,” Historical Studies in the Physical Sciences 21, no. 1 (1990): 62, For M.I.T. after the Rad Lab see also: Stuart Leslie, The Cold War and American Science: The Military-Academic-Industrial Complex at M.I.T. and Stanford (New York: M.I.T.’s electrical engineering department recalled, physicists “disdain[ed] to take

advantage of the work that had been done in” his department.363 For example, Edward

Bowles, an adept radar engineer, was (in the words of a fellow engineer) hampered by,

“being an engineer rather than a scientist.”364 Recall from chapter one that when Bowles

tried to explicitly encourage engineering organizational and research methods at the lab,

the physicists “pasted hell out of him.”365

The Rad Lab physicists developed a culture of self-congratulation.366 Alvarez was

not unique in contending that “the real reason” behind the Rad Lab’s success was its

nuclear physicists, who were “the best people and… adaptable to anything.”367 The

physicists delighted in creating a world for themselves, free of the expectations of any

other discipline or body, and whose residents all shared “the same kind of values”—

namely, physicists’ values.368

The Rad Lab aspired to an academic setting: familiar to physicists but not so to

the typical engineer, who frequently worked in industrial laboratories instead. The

university-like openness of the Rad Lab ideally suited the physicists.369 In the eyes of

Rad Lab physicists, because directors DuBridge, Rabi, and F. Wheeler Loomis were

Columbia University Press, 1993). On Alfred Loomis as the financier of M.I.T.’s prewar work see Conant, 129, 158. 363 Hazen, 3-29. 364 Ibid. Even so, Bowles greatly respected the nationwide patchwork of physicists who had descended upon his university. Conant, 204. 365 Fortun and Schweber, 606. 366 Associate director F. Wheeler Loomis confessed that the laboratory itself was a “buoyant, self confident organization, imbued with the idea of its own importance… doubtless to the irritation of outsiders.” Massachusetts Institute of Technology, Five Years, 4. 367 Conant, 207. See also Burchard, 234. Alvarez was especially self-confident: “very bright and ahead of many in practical thinking. I think he didn’t feel he was appreciated enough.” “Rad Lab: Oral Histories,” Pollard interviewed by Bryant, 217. 368 On “values,” see “Rad Lab: Oral Histories,” Purcell interviewed by Bryant, 246. On the freedom from expectations of other bodies or disciplines see: Guerlac, Radar, 297 and Pollard xiii and 7. 369 Buderi, 130. On the mores of academic physicists’ and its relationship to their research see: Raphael Rosen, “The Argonauts Assemble.” “good scientists,” they respected the physicist’s individual freedom and always listened to what he had to say.370 In this physicist’s conception, scientists freely crossed administrative boundaries, “unencumbered by administrative routine,” with “loose and informal” organizations.371 It was, of course, the “physicist’s world,” a physicist’s

Shangri-la: respectful bosses, freedom to innovate, and permission to cross boundaries.

These principles stood in tacit contrast to the engineering laboratories of the armed services and industry, institutions academic physicists regarded as stifling and intellectually sterile.372

As the laboratory expanded, the physicists made plainer their contempt for the engineer.373 In 1942, physicist Louis Ridenour averred, “it is questionable whether any system [that] can be properly designed by the engineers of an industrial company should be designed by the Rad Lab.”374 Though motivated by a desire for efficiency, Ridenour nevertheless demonstrates a clear antipathy for performing the task of the engineer.375

Despite being one of Alvarez’s “best people,” Ridenour apparently proved less adept at appreciating the type of work in which he engaged.

370 Pollard, 155. Gendered language again, is used for convenience, though it is true that in the Rad Lab’s earliest days there were no female scientists or engineers on staff. 371 Crossing boundaries is from Conant, 214. The unencumbered phrase is Guerlac’s: Guerlac, Radar, 265. “Loose and informal” is Burchard’s: Burchard, 235. Researchers aided one another across fields of expertise and enterprised resourcefully on their own. 372 In England, Robert Watson-Watt would not deign to allow electrical engineers into radar work because he feared that their thinking and inventions would be too conventional and stifled. Brown, 460. 373 It is worth noting that in England, according to Vannevar Bush, the engineer was a “second-class citizen compared to the scientist.” Bush, Pieces, 54. 374 Louis N. Ridenour. “Memorandum on the Place of the Radiation Laboratory in the War Effort,” Coordination Committee Series 1, Box 59, “reorganization,” M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts. 375 Ridenour did face opposition. Frederick Dellenbaugh, M.I.T. electrical engineer and vice-president of Raytheon Corporation saw a vaunted place for engineering, “Engineering in the Laboratory is a definite function…. There should be an engineering executive upon the staff of the Director to coordinate the different degrees of engineering activities.” Ibid. See also: Galison, Image and Logic, 247-250. Years after the Rad Lab closed, the physicists continued to feel contempt for the

practice of engineering. In 1949, when Lee DuBridge, by then president of the California

Institute of Technology, reminisced about the “miracles” of radar and the nuclear bomb,

he saluted especially the work of Heinrich Hertz, J. J. Thomson, Robert Millikan, Breit,

and Tuve—all theoretical physicists, declaring that radar had truly been “born in physics

laboratories before the war.”376 Even when DuBridge mentions that the physicists acted

as engineers, he minimizes the word’s importance by placing it in a list of other terms,

calling the physicists, “engineers, military strategists, salesmen, production experts.”377 A physicist could never simply be an engineer, but had to become an expert in everything; after all, physics was the discipline of supermen. DuBridge concludes, “the job was done by scientists and could probably not have been done so effectively by any other group.”378 Punt the engineer.

Ernest Pollard’s recollections also typify the sentiments of physicists when

thinking about their Rad Lab engineering work. Pollard admits that he built devices, but

takes care to distinguish his methods from those of engineers. He stresses the value of

William Hansen’s Monday evening seminars on the basic physics of electromagnetic

wave propagation, noting, “scientists value this fundamental material even when working

hard at specific applications.”379 Pollard suggests that this “fundamental material” helped

the “scientists” in ways that, presumably, it would not have aided an engineer. Recall that in chapter one, these lectures were shown to be largely irrelevant to the production of radar.

376 DuBridge, “Birth,” 4 and 13. Emphasis mine. 377 Ibid., 8. 378 Ibid., 9. 379 Pollard, 46. More explicitly, Pollard writes that the physicists’ capacity to communicate on a

“really high, difficult scientific level,” proved of “central importance” to the laboratory;

“basic knowledge,” was critical.380 His account suggests that the physicists conducted work more advanced and “fundamental” than what a similar group of non-physicist inventors could have handled—namely, engineers. “Basic scientific integrity,” Pollard avers, was essential for inventive work.381

As has been argued, in most cases and certainly for airborne microwave radar, familiarity with physical theory played only a small—and certainly not a “central”—role in the development of the technology.382 The physicists succeeded when acting as engineers. Pollard never denied engineering’s role, but he preferred to credit scientific genius.383

During the postwar era, the academic physicist’s disrespect for engineering endured. From the 1950’s until the 1970’s, Harvard University’s Division of Engineering and Applied Physics (DEAP), despite its name, consistently favored theory over practice.384 DEAP, of its own choosing, made its “approach to Mechanical

Engineering… oriented toward Engineering Analysis rather toward specific products.”385

Courses dealt “with fundamental principles rather than… specific products” and

380 Ibid., 7. 381 Ibid., 8. 382 It should be noted that the physicists’ advanced mathematics background helped them, and was something which many engineers at the time lacked. Even earlier, inventor Thomas Edison, for example, depended upon others for mathematics and saw no reason to learn it because he could always hire mathematicians if need be. Hughes, 48. 383 John Randall never deigned to speak of himself as an engineer, either though Watson-Watt thought it was obvious. Randall, 313. 384 Raphael Rosen, “The Temperature of Innovation: The History of Superconductivity at Harvard and the Problem of Theory.” (Unpublished Manuscript, 2004), 29. As Bruce Seely has argued, federal contracts in general “favored fundamental research, encouraging university engineers to move away from their practical emphasis.” Seely, 370. 385Division Committees Materials-Mechanical,” [anonymous author], UAV 362.7011. Records of the Division of Engineering and Applied Physics, Pusey Library, Harvard University, Cambridge, Massachusetts. “devices.”386 Engineering students became so loaded down with theoretical work that

they habitually complained of insufficient time to work in the laboratory—the very

nucleus of the engineer’s labor.387 Perhaps most significantly, for the first two Chairs of

this engineering division, Harvard selected theoretical physicists John van Vleck and

Harvey Brooks, who between the two of them led the DEAP for nearly three decades.388

Across the nation, the dominance of science in the postwar decades caused enginee ring departments to drop traditional craft practice.389 The National Science

Foundation, “had considerable influence in steering the research in…colleges of

engineering toward an unrealistic scientific emphasis…. The necessity to affirm that no practical applications [were] expected… sacrificed engineering research upon the altar of

opinions of scientists.”390 At Carnegie Mellon University, for example, the dean of

engineering lamented that actual construction of artifacts “must compete with the

engineering sciences [analysis] for a place in the curriculum.” At M.I.T. in the 1960’s,

recent engineering graduates avoided problems that involved hands-on craftsmanship,

tackling the much smaller subset of challenges that could be completely “solved by

analytical methods.”391

The National Academy of Sciences (NAS) provides another illustration of the postwar stifling of engineering. From its very inception, the NAS preferred “pure”

386 Ibid. For details on the DEAP and its preference for the theoretical see: Raphael Rosen, “The Temperature of Innovation.” 387 “Discussions of plans for future years,” UAV 362.7011. Records of the Division of Engineering and Applied Physics, Pusey Library, Harvard University, Cambridge, Massachusetts. 388 John van Vleck led the department from 1951-1957. Brooks was in charge from 1957-1975. For details see: UAV 362.7011, Records of the Division of Engineering and Applied Physics, Pusey Library, Harvard University, Cambridge, Massachusetts. 389 The analysis focus of other universities at the time was a widespread phenomenon. On the resultant “crisis” for engineering, see Ferguson, 159-168. 390 Marshall, 18. 391 Ferguson, 161-2. academ ic scientists to engineers.392 For years, engineers complained that the NAS

deprived their profession of “proper recognition,” allowing few of their ilk within the

Academy’s walls.393 As a result, engineers believed—accurately—that they had suffered

a substantial loss of prestige; Academy membership correlated directly with

“membership in the highest governmental advisory groups on science and

technology.”394 Only in the early 1960’s did engineers finally grow “pretty well fed up with dominance of the scientists,” demanding (and eventually receiving) their own branch in the NAS, a hundred years after NAS’s inception.395

But even the prestige of the newly-minted National Academy of Engineering suffered in comparison to that of the NAS.396 Something else continued to prevent the engineers from receiving the respect they deserved, something much older, deeper, and

more intractable.

II. Scientific Potentate, Engineering Vassal

During the 17th and 18th centuries, the engineer and the scientist each faced a

crisis. The engineer needed to differentiate himself from the average workman: among

builders, he was productive, but not unique.397 By contrast, the scientist was renowned—

often a gentleman—but society considered him of no practical use.398

392 Pursell, Jr., 235. 393 D. S. Greenberg, “The National Academy of Sciences: Profile of an Institution (III),” Nature 156, no. 3774 (April 1967): 489. 394 Ibid. 395 Ibid. 396 Ferguson, 157. 397 The guild system of the medieval era and the absence of a patent system, locked the aspiring inventor in place. Multhauf, 43. 398 Ibid. University patronage had allowed the scientist to safely ensconce himself in the

ivory tower, taking up his role principally as a natural philosopher.399 But as the 19th century dawned, the scientist increasingly needed to justify his patronage.400 Joseph

Henry, one of the founders of professional academic science in America, struggled to

find time, approval, and support for his research in natural philosophy.401

Henry’s solution was to propose that every mechanical device had its roots in

theoretical science.402 This provided the validation he needed: without his work,

advancements in technology, which the country ever more depended upon, could not

proceed.

In truth, Henry did not invent this idea. Since Francis Bacon and The Great

Instauration, technology’s basis in science had been an essential tenet of the scientific

community.403 Bacon did not divide science and technology as neatly as would be done

in the 20th century, but his utilitarian concept of science still snaked its way into the

charter of England’s Royal Society.404 The difference between pure and applied would

eventually become institutionalized in Europe and from there journey to the United

States.405 Henry stood among the principal proponents of the theory, however, and it was

not long before the “orthodox rhetoric” of the physics community maintained that basic

399 Ibid. 400 Layton Jr., “American Ideologies,” 690. 401 Ibid. 402 Idem, “Conditions,” 206. 403 Ibid. In fact, the rivalry between those who think and those who use their hands can be dated back to the Greeks, at least. See Layton Jr., “Technology as Knowledge,” 40 and especially Stokes, 27-30. 404 Stokes, 33. 405 Ibid., 34-45. science was the locomotive of all technological progress, the source of all scientific knowledge.406

There is nothing inherently problematic in a portrayal of engineers as using scientific knowledge; a problem emerges, however, when engineers are portrayed as absolutely dependent upon scientists, and—here is the crucial point—incapable of independent means of knowledge and problem-solving. Indeed, engineers themselves see no contradiction between applying science and being an independently creative force. In

2006, the National Academy of Engineering website’s answered the question “what is engineering” by defining engineering as “the ‘application of science,’ because engineers take abstract ideas and build tangible products from them.”407 Engineers, as described in chapter two, have long endeavored to affiliate themselves with scientists, partly to gain the prestige associated with them.408 Association with applied science is not damaging; in fact, only recently have engineers particularly emphasized their discipline as distinct from science (rather than from business, as they traditionally have).409 Affiliation (and application) are not synonymous with intellectual dependence. Venerating basic science

406 For more on Henry’s role see Stokes, 41 and especially Layton Jr., “American Ideologies.” In 1976 it was clear that the dogma of the centrality of basic science was “the central theme of scientific ideology to this day…. It is clearly the source of our current theory of the relations of science and technology.” Layton Jr., “American Ideologies,” 690. “Orthodox rhetoric” taken from Kevles, 390. 407 This is one of two definitions given. The other focused on the engineer’s ability to design. On the issue of design in engineering and its long legacy see Layton Jr. “American Ideologies.” The questions and answers from the National Academy of Engineering can be found at: “Frequently Asked Questions,” National Academy of Engineering, (accessed 9 February 2006). On the score of engineers taking pride in applying science, 2003 Nobel laureate in Medicine Paul Lauterbur, proclaimed in his Nobel lecture, “I’d had a largely engineering education, so I had faith that if physicists said something could be done, the engineers could do it.” Paul C. Lauterbur, “Nobel Lecture,” The Nobel Foundation (accessed 21 February 2006). 408 Vincenti, What Engineers Know, 6. See note 279. 409 Layton Jr., “American Ideologies,” 696, 698. is not problematic, but venerating basic science at the expense of depriving engineering of all agency is.410

At the end of World War II, however, the United States government accepted the idea of basic science as the only source of knowledge. Vannevar Bush, who regularly bemoaned engineering’s lack of prestige, would ironically come to bear responsibility for institutionalizing engineering’s indentured servitude to science. Responding to a request from the White House, Bush prepared a report (Science, The Endless Frontier) on sustaining the nation’s scientific program during peacetime.411 Released in July 1945, the report’s mantra held: “basic scientific research is scientific capital.”412 Bush advised the president to “strengthen the center of basic research” in the United States and that “a nation which depends upon others for its new basic scientific knowledge will be slow in its industrial progress and weak in its competitive position in world trade.”413 This report carried substantial weight in American government, because radar, the proximity fuse, and the nuclear bomb had all inflated American muscle, and each one of them had come to fruition under the auspices of Bush’s brainchild, the Office of Scientific Research and

410 The work of the engineer “is often treated as a ‘black box’ whose contents and behavior may be assumed to be common knowledge.” Layton Jr., “Conditions,” 198. See also: Lewenstein, 44-45. 411 Although Bush prepared the report at Roosevelt’s behest, he delivered it to President Truman, as Roosevelt had passed away on April 12, 1945. 412 Bush, Endless Frontier, 2. 413 Ibid., 2 and 14. Bush adds that “the flow” of “knowledge and understanding” ran only from science to the applied sciences. Bush, Endless Frontier, 7. The full ideological and social motivations behind Bush’s conclusions are beyond the scope of this thesis. Briefly, though, Bush’s position was not categorical in the sense of believing that all technological progress derived from basic science; however, the nuance he added was mostly very minor, and the holders of the purse strings paid this complexity no attention. Kline, 219. See also: Stokes, chapters 1 and 2. Edgerton contends that Bush was arguing only for “public support of basic research in universities.” Edgerton, 40-42. Edgerton’s admits readily that, ‘there is little doubt, for example, that academic research scientists have long made, and continue to make, exaggerated claims for the significance of their work for technological and economic development, and the agencies which came to fund them did the same.” Emphasis in original. Edgerton, 38. The opinions of academic research scientists, like Joseph Henry, are exactly whom this thesis is interested in. Development. As a result, the report “gained enormous cultural authority.”414 Science,

The Endless Frontier became the standard explanation for scientific progress.415 The government embraced it, funding basic science generously.416

Expectedly, postwar physicists roundly endorsed Bush’s view.417 Hans Bethe,

1967 physics Nobel laureate and Rad Lab alumnus, asseverated that the conduct of basic research when compared with applied sciences and engineering, “is intrinsically so much more difficult and so much more varied that it keeps a scientist flexible and able to tackle any problem.”418 Physicist Edward Condon proclaimed in the 1950’s that, upon

“fundamental scientific research… rests the whole structure of modern industry, agriculture, and medicine…. Future progress must be built” on basic science.419 In 1957,

Nobel laureate Hermann Muller dismissed applied science: basic science, he remarked,

414 Kline, 218. See also: Seely, 370. In 2000, four out of every five Americans endorsed government support for basic research. Science and Engineering Indicators, 2000, 8-17 - 8-18. On American support for basic research see also: Nathan Reingold, “Reflections on 200 Years of Science in the United States,” in The Sciences in the American Context: New Perspectives, ed. Nathan Reingold, (Washington, D.C.: Smithsonian Institution Press, 1979), 11 415 Layton Jr., “American Ideologies,” 688. President Kennedy told the National Academy of Sciences, “We realize now that progress in technology depends upon progress in theory.” Kevles, 390. This belief in the primacy of basic science constituted a “paradigm,” for understanding science and technology through the remainder of the 20th century. Stokes, vii and 25. Other scholars, like Layton Jr., also argue that prominence of what is known (usually derisively) as the “linear model,” in which basic science is the only source of technological innovation, was enshrined by Bush. But this argument is not without its detractors. See especially, Edgerton, 31-57; Ferguson, 158. 416 I use the term “generously” relatively. In the scope of all science funding, basic research constituted on average only five percent of the entire military research and development budget. Forman, “Behind Quantum Electronics,” 198-199. Still, the government believed in basic science as wellspring of all technology, even if it simultaneously funded fields of technological research, which, if anything was to be constructed, obviously needed support as well. After World War II, the 1947 Steelman report, directly inspired by Bush’s report, recommended the government spend at least 1% of its Gross National Product on support for basic research and development annually. Dupré and Lakoff, 12. The department of defense invested some $2.5 billion in basic research from 1945 to 1966. Layton Jr., “Mirror-Image,” 563. In 2006, the National Science Foundation had $5.5 billion to disperse annually for basic research. “About the National Science Foundation,” National Science Foundation, (accessed 10 February 2006). 417 Bush’s report was “the first important document to embody some of the thinking of the scientists” with regards to policy after the war, Dupré and Lakoff, 11. 418 Schweber, QED, 141. 419 Condon, 8-9. “feeds…the applied sciences… best when it forgets them.”420 Myriad physicists wholeheartedly expressed similar opinions.421

As a result of these sentiments, though physicists took pride in their contributions to, for example, the advancement of radar, they consistently downplayed the importance of all things technical.422 They reliably averred “new technologies are ever the result of fundamental research.”423 Just as DuBridge had asserted that microwave radar was “born in physics laboratories before the war,” others insisted that microwave radar only “burst into being because of a long background of fundamental science.”424 Physicists surrendered a huge portion of their true wartime creativity—their appropriation of engineering methodology to solve practical construction problems—so that basic physics remained the bedrock of all progress. Aware though they were that postwar physics

420 Reingold, “Pride and Prejudice,” 291. Muller won the Nobel Prize in Medicine in 1946. 421 M.I.T. president Karl Compton thought similarly. In 1946, he declared at California Institute of Technology’s commencement that universities needed to place greater emphasis on science and research to train students for “opportunities of the future rather than techniques of the past.” “Dr. DuBridge Installed: Takes Over as Head of California Institute of Technology,” New York Times, Nov 13, 1946, pg. 8. Similarly, Columbia physicist Michael Pupin put engineering in its place, “Engineering is Science’s handmaiden following after her in honor and affection, but doing the practical chores of life.” Kline, 204. Robert Millikan flatly informed a team of industrialists: “pure science begat modern industry.” Robert A. Millikan, “The Relation of Science to Industry,” Science 69, no. 1776 (January 11, 1929): 28. During the war, one Nazi professor averred “As the practice of today rests on the science of yesterday, so is the research of today the practice of tomorrow.” Professor Thiessen as quoted and translated in: Robert Merton, “Science and the Social Order,” in Merton, The Sociology of Science, ed. Norman Storer, 254-66 (Chicago: The University of Chicago Press, 1973), 257. See also: Sheila Jasanoff, “The ‘Science Wars’ and American Politics,” in Between Understanding and Trust: The Public, Science and Technology, eds. Meinolf Dierkes and Claudia von Grote, (London: Routledge, 2003), 43. 422 Ernest Rutherford, who died in 1937, but as we saw earlier, helped advise Ernest Pollard at Cambridge, was known to disrespect engineering: having “little feeling for engineering.” Snow, 34. 423 Forman, “Swords into Ploughshares,” 397. In Bush’s postwar Science, The Endless Frontier, nearly every contribution of engineering was cast as rooted in fundamental scientific inquiry. Kline, 219. 424 Vannevar Bush, The Gentleman of Culture [No Publisher] (Distributed by E. Laurence Springer of the Pingry School and John T. Connor of Merck & Co., 1959), 2. Bush maintained, “It was painstaking scientific research over many years that made radar possible.” Bush, Endless Frontier, 5. research also owed a formidable debt to wartime engineering, physicists largely would

not admit to it.425

C. P. Snow, physicist and novelist, attested to (and criticized) physicists’

widespread dismissal of engineers’ creativity. Reflecting on his own career, Snow observed, “pure scientists have by and large been dim-witted about engineers and applied science…. They [did not] recognise that many of the problems were as intellectually

exacting as pure problems…. Their instinct…was to take it for granted that applied

science was an occupation for second-rate minds.”426 Snow had had to overcome the

mores of his community, “I say this more sharply because thirty years ago, I took

precisely that line [of thought] myself.”427

Physicists’ bifurcated model of venerated basic science and plebian, “second-

rate,” engineering clearly related to the impugning of the latter.428 Physicists could be “a

little conceited when comparing engineering and physics.”429 According to one of their

own, physicists “feel that progress in engineering is limited by the work of the physicist

and that they, therefore, are the important cogs in the wheel.”430 This scorn, typically

unspoken, was widespread. The sneer of certain Rad Lab physicists towards engineering,

in the case of Ridenour, Alvarez and others is bald; the popular jokes about engineers

cited in the previous chapter evince the same haughtiness. The battle engineers

425 Forman, “Swords into Ploughshares,” 397. Charles Townes is a notable exception who publicly recognized physics’ debt to engineering. Charles Hard Townes, “Microwave Spectroscopy,” American Scientist 40 (1952): 287. 426 Snow, 33-34. 427 Ibid., 34. 428 Perhaps physicists have come to disparage engineering because they envision it as derivative of physics, or perhaps this philosophy simply reflects such a disparagement, or perhaps disparagement and philosophy are mutually reinforcing. The issue is too sweeping for full consideration here. 429 Elmer Hutchisson, “The Egg or the Chick,” Journal of Applied Physics 10, no. 3 (March 1939): 141. 430 Ibid. One of the National Science Foundation’s first reports was based upon a “technological sequence,” which strictly held a clear linear model” basic science leads to applied science which in turn leads to development. Stokes, 10-11. encountered at the National Academy of Sciences provides still further illustration: during the conflict, physicists were heard to remark that scientists received deference as government advisors because there were “generally… smarter than engineers.”431

Physicists in the National Science Foundation likewise overtly expressed hostility towards the presence of engineers in the organization.432 By 1990, engineers continued to be viewed as “taking… their knowledge from scientists.”433

In short, engineering knowledge has traditionally been discounted as irrelevant,

“intellectually uninteresting;” only recently has scholarship in the history of technology begun to candidly probe what it is that engineers know and do.434 It is not mere pedantry to argue against the claim that basic science is the font of all knowledge, because engineering’s impact upon science is empirically palpable.435 Scholars have shown that since the time of Bacon, knowledge has flowed from engineering to science.436

Engineering, it will be demonstrated, does contain its own forms of knowledge—its own breath and tears and tortures. It has its own agency, its own authority.

431 Greenberg, 489. 432 Reingold, “Pride and Prejudice,” 291. 433 Vincenti, What Engineers Know, 3. 434 “Intellectually uninteresting” taken from Vincenti, What Engineers Know, 3. Technology and engineering are not the same matter, but for the purposes of this thesis, technology ought not to be considered much broader than engineering. It is interesting to note, that during the course of my research that the bulk of the work on engineering as its own form of knowing has come in two principal spurts: for about five years in the mid-1970’s and for about another five years in the late 1980’s, early 1990’s. This observation might provide a starting point for someone wishing to write a historiography of the treatment of engineering as its own intellectual entity. 435 The history of science itself, should it adopt the view of basic science as the source of all innovation and knowledge, damages its own investigative pliability, because, if all the thoughts of engineering depend upon science then the epistemology of science subsumes all engineering knowledge as well. Vincenti, What Engineers Know, 3. 436 Stokes, 20. Technology heavily influenced the Scientific Revolution. Layton Jr., “Science as a Form of Action: The Role of the Engineering Sciences,” Technology and Culture 29, no. 1 (January 1988): 88. Scientists have tried to enumerate instances in which technology derived from basic science, citing the magnetic compass, the printing press, gunpowder, the clock, glass, the cotton gin, the steam engine, and textile machinery. Yet, in all of these cases the influence of basic science is not readily apparent, and certainly the linear model alone is wholly insufficient. Layton Jr., “Mirror-Image,” 563. III. Towards an Epistemology of Engineering

Several months before Germany invaded Poland, sparking the second World War,

Elmer Hutchisson, editor of the Journal of Applied Physics, remarked, “The physicist uses engineering to advance physics and the engineer uses physics to advance engineering.”437 Hutchisson’s observations broke rank with the orthodox view of physics by affording engineering its own agency. Science did not simply feed its knowledge to engineering; rather, the progress of each depended upon the other. Physics was a transmitter-receiver, not just a transmitter.438

After the war, a few physicists similarly recognized engineering as an independent source of innovation.439 Charles Townes, for example, inventor of the maser, noted in 1952 that, “‘pure’ physicists like to regard as typical…that pure science develops principles and ideas which are then applied to technology by other scientists and engineers.”440 Townes insisted that nevertheless, “in the case of microwave spectroscopy… the equipment and devices were developed first, and pure science owes a considerable debt to technology.”441

437 Hutchisson, 141. 438 One of the first historians to make such a proposal, as far as I have been able to tell, Hessen in 1931 who remarked in relation to the Newtonian revolution “the gigantic development of technique tremendously stimulated the development of science,” and vice versa. Hessen’s argument is strictly Marxist. B. Hessen, “The Social and Economic Roots of Newton’s ‘Principia,’” in Science at the Cross Roads, ed. Joseph Needham, 149-212 (London: Frank Cass and Company, 1971), 206. Though detailed models of the relationship between science and engineering is not the primary interest here, one model is especially worth noting. Donald Stokes, shortly before his death, argued for the importance of what he termed “use-inspired basic research.” This referred to a body of research intended both to understand basic phenomena and to apply it. This interesting hybrid of science and technology motives is exemplified in Pasteur’s work on microbiological agents which he discovered and wanted to control, or the Manhattan project physicists who wanted to understand the nucleus and weaponize it. See Stokes, especially 78-80. 439 Physicists were also lectured to about engineering’s independence. Dr. T Keith Glennan, President of the Case Institute of Technology told the American Physical Society, “We have all heard, many times, of the role of research in basic science in advancing our technologies. However, there is a less often stressed, but in my opinion, equally valid observation—that science owes a debt to technological advances which is equally great as the more commonly recognized debt of technology to science.” Marshall, 20. 440 Townes, 287. 441 Ibid. Engineering’s independent thought is further evident in the findings of Project

Hindsight, a 1966 Defense Department review of twenty major weapon systems

developed since 1945.442 The Project concluded that of some 700 total innovations, 90

percent grew out of engineering, not basic science. “The folklore of science” did not hold

up.443

Unsurprisingly, scientists lambasted the results. Letters to the editor flooded the

journal Science. A National Science Foundation counter-study concluded that basic

science and invention were much more closely linked than the Defense Department

review claimed.444 The true disciplinary source of each of the 700 innovations is less

important than the stunned reactions of the scientific community, forced to confront a

creative energy they had declared nonexistent.

Or consider the case of the Materials Advisory Board of the National Academy of

Sciences. When the Defense Department asked the NAS to investigate important

innovations in technology, the Board attempted to organize its results using a tidy, seven-

stage model of development: starting with basic science and proceeding down the ladder to technological applications. It failed. Not a single one of the historical case studies followed the linear model. Innovation predominantly flourished as a byproduct of technological activity itself.445

Findings like those of the Project Hindsight or the Materials Advisory Board are

often criticized, and not without good reason, for failing to consider the interaction of

442 As a result of the prestige of physics explored in the previous chapter, it had already taken two decades for the Defense Department to investigate whether engineering, as its own creative endeavor, not simply as applied science, might be making crucial contributions. Layton Jr., “Mirror Image,” 563. 443 As quoted in Kline, 220. Officially, out of approximately 700 key contributions and events, an astounding 91 percent resulted from technological progress. Layton Jr. “Mirror-Image,” 564. 444 Kline, 220. 445 Layton Jr. “Conditions,” 209. engineering and science; but such critiques also overlook an important point.446 True, engineering and physics are intertwined, and their interaction was critical to both the Rad

Lab and the Manhattan Project, and true, scientists sometimes pursue particular devices, while engineers may develop basic theories about materials.447 Nevertheless, the engineering and physics community each have their own textbooks, their own journals and societies, their own classroom lectures and experimental laboratories and pedagogical tools.448 As one aeronautical engineer and historian phrased it, “that there is a difference in thinking between engineering and physics could hardly be more plain.”449

Though engineers share many materials with physicists, they must still maintain their own libraries, storehouses for their unique methods.450

What specifically do engineers know and do that differs from scientists? This question has been long overlooked by historians, engineers, sociologists, and policy makers alike.451 I here offer my own attempt at a partial answer: perhaps the most important knowledge unique and indigenous to the engineer is that of design.452

446 Ibid., 207. The intertwining of engineering and physics need not be viewed as a strong one. “Technology and science… are in fact enmeshed in a symbiotic relationship—a weak mutually beneficial interaction, which looks much the same whichever way round it is considered.” Barry Barnes, “The Science-Technology Relationship: A Model and a Query,” Social Studies of Science 12, no. 1 (February 1982): 168. This thesis focuses does not go into depth about the mechanism by which physics work affected the conduct of the engineering work. As stated in the introduction, the precise mechanism of this interdisciplinary interaction is not the aim here. Engineering’s great importance is. 447 Layton Jr. “Conditions,” 210. 448 On engineering education methods emphasizing concrete problem-solving see Vincenti, What Engineers Know, 253. See also: Layton Jr., “Mirror-Image;” Layton Jr., “Conditions,” 209. 449 Vincenti, “Control,” 165. 450 Idem, What Engineers Know, 4. 451The precise comprehension of this engineering methodology requires an undertaking well beyond the scope of this investigation. The only book-length study on the subject I am aware of is Vincenti, What Engineers Know. Even this sees itself as merely an opening salvo in an enormous undertaking. To a lesser extent see also: Layton, Revolt. 452 Another native engineering form of knowledge, for example, is a concern with increased efficiency overall—“the measure of effectiveness.” That is known as praxiology, a philosophical system that “analyzes action from the point of view of efficiency.” Skolimowski, 376-377. Design is no simple matter.453 For some engineers and historians it represents the intangible “essence” of engineering, while for others it is simply part of the engineering curriculum.454 Engineers in America frequently identify the “ability to design” as a common denominator, using it as part of the qualification requirements for engineering societies, including the National Academy of Engineering.455 In this thesis, design will be employed principally to designate a unique breed of knowledge native to the engineer.456

The importance of design reflects the fact that engineers think not in abstract terms of nature, but in the concrete terms of machines: they know and ponder in the language of electric dynamos, airplane wings, and hydraulic systems. Values are entrenched not in a hierarchy of sub-atomic particles or conservation of momentum and energy laws, but within artifacts, within hardware.457 Scientists explain their forms of knowledge in reference to fundamental entities, engineers in reference to measurable

453 I have lumped together what Layton Jr. separately characterizes as engineering science and design, Layton Jr., “American Ideologies,” 699; I take them under the same umbrella here to give a full portrait of engineering knowledge without becoming mired in many technical details. Design should not be confused with tacit knowledge, but the two can overlap. Design is the art of creating something and involves aspects of tacit knowledge. But there are aspects of design that can be written down and communicated as a sort of theoretical framework, and there are aspects of tacit knowledge, for example, reading instruments, that do not relate to design. On tacit knowledge see: Harry M Collins, “The TEA Set: Tacit Knowledge and Scientific Networks,” Science Studies 4 (1974): 165-186; Harry M. Collins, “What is Tacit Knowledge,” in The Practice Turn in Contemporary Theory, eds. Theodore R. Schatzki, Karin Knorr-Cetina, & Eike von- Savigny (London: Routledge, 2001), 107-119; Donald MacKenzie & Graham Spinardi, “Tacit Knowledge, Weapons Design and the Uninvention of Nuclear Weapons,” American Journal of Sociology 101, No. 1, (1995): 44-99. See also: Ferguson, 32, 59. 454 “Essence” taken from Layton Jr., “American Ideologies,” 696. Engineer William McClellan described design as the defining trait of the “real engineer,” Layton Jr., “American Ideologies,” 696. Ferguson takes design to be the key element of an engineer’s education. When design became de-emphasized in engineering curricula after the second World War, Ferguson argues that engineers became dramatically less capable and needed to help from those with greater design expertise; see Ferguson, chapter 6. 455 “Ability to design” taken from Layton Jr. “Technology as Knowledge,” 37; design as qualifying criteria from Layton Jr., “American Ideologies,” 700. Information concerning the National Academy of Engineering from Luis Alvarez Papers, “National Academy of Engineering,” Box 76 MSS 84/82 cz. The Bancroft Library, University of California at Berkeley, Berkeley, California. 456 This is a more general use of ‘design’ than that of many scholars, but by no means a contradictory one. 457 For more on these values see: Hughes, 4-5. quantities.458 This is not a derivative manner of thought, but a divergent one. When an engineer is creating an artifact and uses existing expertise about related machines and about how to combine systems and parts, this is using engineering knowledge; this is design.

To understand this concept more clearly, let us return to the Rad Lab. There, the dovetailing of many imperfect, independent elements into a complex, working whole constituted the principal charge.459 This, it has been argued, constituted an act of

engineering. Why? The reason is because amalgamating existing pieces into a working

whole is precisely the work of design. The essential task in the early days demanded that

the physicists take components and vivify them. Initially, they were ill equipped, but as

they began to familiarize themselves with the components, with the devices themselves,

and as an increasingly large fleet of M.I.T. engineers assisted them, the physicists applied

design knowledge and made radar see.460 British microwave radar, by contrast, “suffered

from a lack of engineering personnel, the young physicists having to learn from

experience many of the practical design techniques that would have been rules of thumb

for engineers.”461 In England, they wanted for a knowledge of “practical design,” and struggled as a result.

458 Layton Jr., “Mirror-Image,” 569. 459 “So much technological work requires combining elements into a working whole in order to reach some preconceived end.” Idem, “Technology as Knowledge,” 39. Drawing the Rad Lab connection is my idea, not Layton Jr’s. 460 Wildes and Nilo, 198. The support of engineers for the project is expounded upon further throughout chapter 13 of this volume. At least as early as December 3, 1940, Lee DuBridge wanted more engineers. As he wrote to Ernest Lawrence, “It seems to me that we can make good use of men who have engineering training…. We are now rather short of engineering experience and I think several young men who have had practical experience with microwaves would be very useful.” Ernest Lawrence, Series 5, Reel 40, Frame 088133, “Correspondence with DuBridge,” Ernest Lawrence Papers, MSS 72/117c. The Bancroft Library, University of California at Berkeley, Berkeley, California. 461 Wildes and Nilo, 198. The postwar volume on microwave circuits by three Rad Lab alumni further illuminates the unique brand of engineering knowledge underlying radar. In the book’s preface, the authors discuss “underlying principles,” “generalized… concepts” and

“general methods” in the “application of microwaves.” The authors, however, are not referring to “underlying principles” of microwave physics, but to “general methods” of

“microwave devices.” As the preface proffers, “the principles discussed in this volume can be applied to microwave equipment of all kinds:” microwave equipment, not microwaves.462 Knowledge is embedded in the equipment, in the devices, providing useful information for system design.463

The example of thermodynamics again illustrates engineering knowledge’s independent character.464 In physics, the study of thermodynamics is grounded in statistical mechanics, the random bouncing of particles. In engineering, a common tool is the “thermodynamic cycle,” in which one can calculate something called “thermal efficiency.” Both “thermal efficiency” and “thermodynamic cycle,” however, are

462 C. G. Montgomery, R. H. Dicke and E. M. Purcell, Principles of Microwave Circuits (New York: McGraw-Hill Book Company, 1948), ix. 463 This volume does dedicate a substantial portion of its space to what can fairly be categorized as physics. Nevertheless, throughout the book generalized principles are always thought of in terms of real-world parameters. For example theorems are applied to T-junctions, places where waves are transmitted down only certain arms of a junction and not all of them. T-Junctions are a piece of hardware. When the junctions are considered, generalized impedance matrices may be deployed (for equal-height, symmetric branches of waveguides intersecting at 120°): ⎛ Z11 Z12 Z12 ⎞ , but again, only within the context of hardware: the T- Z = ⎜ Z Z Z ⎟ ⎜ 12 11 12 ⎟ ⎝ Z12 Z12 Z11 ⎠ junctions themselves. Ibid., 293. 464 Or consider still another example. In the 1870’s, physicist Henry Rowland and electrical engineer Francis Hopkinson made essentially the same discovery regarding electromagnetism, but they framed their findings differently. Rowland couched his discovery as relating to magnetic permeability, while Hopkinson framed his in relation to electric dynamos. In other words, Rowland thought about the relations of different parts of electromagnetic theory; Hopkinson thought about physical properties of machine (dynamos) parameters. Layton Jr., “Mirror-Image,” 577. technical concepts and rules that apply only to engines; they are rooted in machines. They

differ in content and form from the rules of statistical mechanics, a law of nature.465

Design as a kind of knowledge is further characterized by one thing purely scientific knowledge does not take into account: art.466 When British engineer, J. D.

North described the character of engineering knowledge to the Royal Aeronautical

Society in 1922, he declared, “Aeroplanes are not designed by science, but by art in spite of some pretence and humbug to the contrary. I do not mean to suggest for one moment that engineering can do without science, on the contrary it stands on scientific foundations, but there is a big gap between scientific research and the engineering product which has to be bridged by the art of the engineer.”467

Indeed, the engineer’s mode of thinking is often characterized as more like that of

an artist than a scientist.468 This makes sense given that engineering requires the selection

of the maximally efficient model from among many choices. The engineer must learn to

create the most elegant, refined product, even if another design will suffice. Like the poet searching for the most cogent expression, the engineer’s seeks the machine’s most effective design. Art becomes a critical form of what he knows: for the engineer, “the successful design of real things… will always be based more on art than on science.”469

The art of design lives within machines. Unlike physics knowledge which

describes the behavior of nature, design is a form of knowledge embedded in real world,

465 Ferguson, 11. See also: Vincenti, “Control;” Küppers, “On the Relation.” 466 In the spirit of art’s crucial importance to engineering, this thesis has included, mostly in the epigrams, its own artistic (literary) references. This is done so that form might match content not, let it be known, as some hubristic exhibition of urbanity. At least ideally speaking, theoretical physics is not a subject to be fudged into perfection. Formalism is not an art form, it is a body of knowledge. 467 Vincenti, What Engineers Know, 4. 468 Layton Jr. “Technology as Knowledge,” 36. 469 Ferguson, 194. human-made artifacts. If one is trying make a radar indicator work more effectively, for example, then it is knowledge about electronic circuit efficiency and about how display screens work—knowledge about designing machines—that is essential. Engineering’s knowledge repertoire, of which design is only the most obvious example, stands on its own, characterized by an “array of abstract concepts, independent of science.”470

It is true that production of devices and technological artifacts, of course, is the main goal of engineering. Indeed, this thesis began with that simplistic definition. But though the search for truths about the world of machines is rarely the explicit aim of engineering, engineers uncover those truths just the same, and engineering can be seen as far more than only production.471

The underlying philosophical problem is that it is easy to mistakenly classify concrete, real-world artifacts as wanting for transcendent ideas. One rarely thinks of a human-crafted object giving rise to any knowledge eclipsing the material world: “what makes a fact different from an artefact is that the former is not perceived to be man- made.”472 Scientists are not conceived of as inventing something in nature, so much as detecting something within it.473 They out intangible truths and are venerated for

470 Ibid., 11. For more on the tools of science and technology see: Bruno Latour, Science in Action (Cambridge, Massachusetts: , 1987) and Kaiser, Drawing. 471 Engineering thus uses its unique forms of knowledge for two ends: the production of artifacts and the generation of more knowledge—about machines themselves. Asymmetrically, physicists only use knowledge for one end: to generate still more knowledge. Vincenti, What Engineers Know, 226. Nor is science simply the pursuit of knowledge. This simple models harkens back to Platonic distinction between episteme (knowledge) and techne (art). In such an envisioning, the linear model quickly surfaces because at first glance it would seem absurd to think of knowledge and ideas flowing from art. Layton Jr. “Technology as Knowledge,” 40. See also: Layton Jr., “Conditions,” 209. 472 Steven Shapin, “Pump and Circumstance: Robert Boyle’s Literary Technology,” Social Studies of Science 14, no. 4 (November 1984): 507. “What men make, men may unmake, but a matter of fact is taken to be the very mirror of nature.” Shapin, 507. Robert Boyle insisted experimental disagreements be over (ethereal) findings and not (concrete) people. See: Shapin, 502, 504. 473 Biagioli, 15. Biagioli discusses how one can track the movement of credit with an item that is private property much more easily than one can with a theoretical object of science, for when a scientific observation becomes associated with a specific individual’s name, it still has not passed out of the public their powers of abstraction. Yet, engineers, too, are discoverers of truth, detecting abstract principles within concrete objects. If physics is the pursuit of truth in nature, engineering is the pursuit of truth in humankind’s creations.

That engineering deals with worldly matters makes it no less worthy a pursuit.

The study of the things in this world can protect trans-Atlantic shipping and power homes. It can also pulverize cities. The goal here is not to adjudicate on the superiority of that which is human-made or that which arises from nature, but simply to demonstrate that engineering constitutes its own world. Thus, engineering must be studied not as simple application of science, but as application of science according to engineering’s own rules: as part of the discipline’s own artistic, creative venture, producing knowledge, not just machine.

Since physicists viewed engineering as derivative and without creativity, after the war, they did not credit it in the public eye. They preferred to think of themselves as creators rather than mechanics. What physicists wished to focus on was the triumph of knowledge, and for them knowledge was not engineering’s domain. According to the scientific community’s dogma, only basic science contained abstract intelligence.

But at the Rad Lab, engineering proved itself to be a mindset, an empirical endeavor with its own exigencies and methods of thought. Its endogenous mechanisms shine through in Pollard’s notebooks when he ceases to think only as a physicist and crafts circuit diagrams like an engineer. These methods stand strong as Alvarez uses them to improve a pulser design or as Schwinger translates his theoretical physics knowledge

domain. Where the person’s particular truth claim ends and the rest of science beings does not cross over an obvious legal threshold. into electrical circuit elements. Unique engineering methodologies explain why only Jim

Lawson could create a TR box: as a radio amateur, he knew theoretical facets not just about electromagnetism but about its machinery. In the Rad Lab, even if physicists principally credited their physics background, they made innumerable breakthroughs precisely because of engineering, precisely because they embraced its indigenous forms of knowledge, thinking, not just doing, in terms of devices and artifacts.

Conclusion

“For ‘tis the sport to have the enginer Hoist with his own petar.” -William Shakespeare474

Friday night at the Commander Hotel bar, just outside Harvard Square, the Rad

Lab physicists would convene to unwind. British physicist Taffy Bowen, on loan to the

Rad Lab, remembers that the bar had “a mural around the walls which depicted various

episodes dear to the heart of all Americans—the Boston Tea Party [and] Paul Revere’s

Ride.”475 The heroes who had seen through the gloom in one generation stepped back and contemplated heroes who had done the same in another. Yet, the legacy of these 20th century heroes would not weather time as their radar had weathered space. As individuals they received recognition, but the skills they employed did not.

At the M.I.T. Radiation Laboratory, engineering methods proved critical to the birth of microwave radar. Building the first working sets had demanded only the improvement and amalgamation of already existing components. Whether tinkering with specific indicator electronics or translating microwave cavities into equivalent circuits, the Rad Lab physicists both constructed and thought like engineers.

474 William Shakespeare, Hamlet Prince of Demark, ed. Willard Farnham (New York: Penguin Books, 1970): III, iv, lines 207-208. A petar (petard) is a mine. 475 Edward G. Bowen, Radar Days (Bristol, England: J W Arrowsmith, Ltd., 1987), 182. In 2006, the Commander Hotel was the Sheraton Commander Hotel. It is interesting to note that in 1776, Colonel Henry Knox, one of George Washington’s commanders, also recognized the importance of wartime research, “there shall be one or more capital laboratories erected at distance from the seat of war…. At the same place a sufficient number of able artificers [shall] be employed to make carriages for cannon, of all sorts and sizes, ammunition wagons, tumbrels, harness, etc.” G. M. Barnes, “Ordinance Research: the Department’s Use of Technical and Scientific Facilities,” Army Ordinance 21, no. 122 (September-October 1940): 110. Despite a brief period of recognition after the war, the discipline of engineering

lost credit to the discipline of physics. The postwar celebrity of physicists led to the association of the physicists’ invention of microwave radar with physics itself. The engineering community, with neither clout nor recognition, could not claim its due.

Though the physicists could have spoken up for the primacy of engineering methods at the Rad Lab, they did not. According to the orthodox philosophy of the physics community, engineering was manufacturing. Unlike basic science, it contained no original ideas of its own.

But this philosophy is demonstrably false. Throughout the 20th century, engineers

have regularly evinced their creativity.476 Moreover, their inventiveness is rooted in

engineers’ unique knowledge of design—of machines and how to craft them with utmost

efficiency. Knowledge within artifacts is the engineer’s epistemological domain.

These ideas have real world consequences. The skills employed by Pollard,

Alvarez, and others find regular practice in resolving (or potentially aggravating)

numerous contemporary issues such as nuclear safety and energy supply. Even more so in

our times than in the those of the M.I.T. Rad Lab, national security and scientific

progress depend deeply upon engineering’s independent imagination. Engineering skills

are not derivative, but inventive—not a receiver, but a transmitter-receiver. Long ago

now, microwaves echoed off Boston’s Christian Science Mother Church. It is time

someone noticed them.

476 This inventiveness is especially evident in the increased intertwining of engineering and physics, of which the Rad lab, of course, is one of the first examples. In 2004, Harvard broke ground for an enormous new center, the Laboratory for Integrated Science and Engineering. Columbia and other universities already have similar laboratories. This conjunction of physics and engineering may reflect a greater recognition by scientists that engineers have their own forms of thought that can impact new research. Appendix Key Radar Terms Explained

Microwave radar: Microwaves are more precise and easier to detect echoes from than longer waves. The first microwave radar sets used 10-centimeter long waves.

Long-wave radar: The first operational radar sets used waves that were a meter and a half or longer. The British Chain Home station which protected the Isle during the Battle of Britain used this long-wave technology.

A Step-by-Step Schematic of a (very crude) Microwave Radar Set

1. Pulser: The pulser produces large jumps of power, each “pulse” lasting roughly a microsecond (a millionth of a second).

2. Magnetron (Wave Generator): The pulses activate the magnetron, which takes the electrical pulses, and, by regulating electron flow, generates powerful microwaves.

3. Transmitter: The transmitter emits the microwaves via the antenna.

4. Antenna (Parabola): The microwaves are sent out by the antenna and the echoes are picked up in the parabola dish (your satellite dish is a parabola using the same principle).

5. TR-Box: The transmitter-receiver box takes the microwave echo from the parabola and passes it on to the receiver. The TR-box must be capable of switching from transmitting to receiving in millionths of a second.

6. Mixer: The crystal mixer then combines inputs from the TR-box with other inputs, such as information about the rotation speed of the parabola.

7. Receiver: The receiver takes the signal from the mixer and transmits it to the indicator.

8. Indicator (oscillograph and oscilloscope): The indicator transforms electrical signals from the receiver into a visual display on a screen (it’s an oscilloscope if it’s a fluorescent screen) that can be interpreted to reveal location, speed, and bearing of a detected object.

+Waveguides: Waveguides are hollow metal pipes of precise dimensions such that microwaves pass through them unimpeded: the electric and magnetic fields perpendicular to one another. The pipes connect many of the components described above.

(Sources: Brown 33-39; Pollard, 40-41; Marcuvitz 1-3) Bibliography

In general, I group sources according to the role (primary or secondary) for which each one was predominantly used. The term “secondary” source is used non-exclusively; at the end of chapter two, I used many “secondary” sources as primary ones. Encyclopedias and textbooks analyzed in chapter two are included under primary sources.

Sources are organized as follows 1. Primary Sources: Books, Journal Articles, Textbooks, and Encyclopedias 2. Archival Sources 3. Newspaper Articles 4. Popular Periodicals 5. Secondary Sources

1) Primary Sources: Books, Journal Articles, Textbooks, and Encyclopedias

Allison, Samuel K. “The State of Physics; Or the Perils of Being Important.” Bulletin of the Atomic Scientists, VI (Jan 1950): 2-4, 26-27.

Allison, director of the during World War II, reflects on the new status of physics in America and the resulting implications. He argues that physics has been cast primarily as a tool “for waging war.”

Alvarez, Luis. Adventures of a Physicist. New York: Basic Books, Inc., 1987.

The autobiography of Luis Alvarez, Nobel laureate and lifelong innovator, provides both a window into the projects and life of the Rad Lab in its early days as well as a portrait of a physicist’s reaction to many of the developments leading up to the birth of radar.

“Biography of Julian Schwinger.” In Nobel Lectures in Physics: 1963-1970, 153-4. River Edge, N. J.: World Scientific Publishing Co., 1998. [Also available online: http://nobelprize.org/physics/laureates/1965/schwinger-bio.html].

A brief biography of Schwinger from an identified author (though likely by Schwinger himself as Forman suggests in “‘Swords into Ploughshares’”) reveals how Schwinger adjusted his world view from physics to engineering at the Rad Lab.

Bowen, Edward G. Radar Days. Bristol, England: J W Arrowsmith, Ltd., 1987.

––––––––. “The Tizard Mission to the USA and Canada.” In Radar Development to 1945, edited by Russell Burns, 296-307. London: Peter Peregrinus Ltd., 1988.

Edward “Taffy” Bowen, the first British liaison to the Rad Lab, provides a look at both American and British radar systems in Radar Days. This is an especially useful guide to the events that already unfolded before the Rad Lab opened. As a result, it throws the Rad Lab’s inauguration and early work in relief. “The Tizard Mission,” is mostly a condensed version of these recollections, with a few additions.

Bush, Vannevar. The Gentleman of Culture. [No Publisher]. Distributed by E. Laurence Springer of the Pingry School and John T. Connor of Merck & Co., 1959.

––––––––. Pieces of the Action. New York: William Morrow & Co., 1970.

––––––––. Science, The Endless Frontier. Washington, D.C.: United States Government Printing Office, 1945.

In Pieces, Bush, director of wartime science as head of the OSRD, reflects on his life and what he has discovered about both teaching and management. I found the book useful both for its considerations of the engineer with respect to the physicist in social status and for its biographical recollections of wartime research. Science, The Endless Frontier contains the official recommendations of the Office of Scientific Research and Development to President Truman (initially requested by Roosevelt). It lays out the necessities of postwar American science: especially prodigious funding for basic science, which is seen as crucial to technological progress. The Gentleman of Culture is a speech delivered by Bush at the Pingry School of Elizabeth, New Jersey on December 6, 1958, and it addresses concerns of Cold War science education.

Davis, Robert C. and the Institute for Social Research, University of Michigan, The Public Impact of Science in the Mass Media. Ann Arbor, MI: National Association of Science Writers, 1958.

This 1958 report on public perceptions of science suggested both great public esteem for science as well as the fact that the Sputnik launch did not substantially affect Americans’ faith in the importance of science.

DuBridge, Lee A. “The Birth of Two Miracles.” “California Institute Forum 6 (1949): 1-15.

––––––––. “History and Activities of the Radiation Laboratory of the Massachusetts Institute of Technology.” Review of Scientific Instruments 17, no. 1 (January 1946): 1-5.

A laconic teaser anticipating the release of a fuller history (Guerlac’s Radar in World War II), “History and Activities” is the summary of the operations of the Rad Lab by its director Lee DuBridge. DuBridge places special emphasis on the contributions of physicists and the need for generous monetary support for physics in the postwar era. “Birth” is an address delivered to the Sunset Club of Los Angeles recounting the wartimes development of the nuclear bomb and radar; DuBridge recognizes the engineering nature of the project, but still exalts science and uses the history of radar as an illustration of basic science’s presence at the root of all technology.

“Frequently Asked Questions,” National Academy of Engineering, .

The National Academy of Engineering’s website reveals the discipline’s own (official) self-perception: the application of science with a commitment to ‘design.’

Getting, Ivan A. “Microwave Systems, Then and Now: Example at the 50th Reunion of the MIT Radiation Laboratory.” IEEE Transactions on Microwave Theory and Techniques 39 (December 1991): 1920-30.

A physicist and an engineer, Ivan Getting discusses his memories of the Rad Lab and the advancement of microwave technologies since the laboratory closed.

Hansen, William. “A Type of Electrical Resonator.” Journal of Applied Physics 9, (1938): 654- 663.

Hansen’s pre-Rad Lab paper reveals his attempts to make wave-cavity phenomena more accessible and easier for non-wave theory experts to apply.

Hazen, Harold. Memories: An Informal Story of My Life and Work. Unpublished Manuscript, 1976. Available at Institute Archives and Special Collections, M.I.T. Libraries. Cambridge, Massachusetts.

Hazen headed M.I.T.’s Department of Electrical Engineering when the Rad Lab was founded; his recollections demonstrate the neglect of engineering expertise in the development of radar.

Heins, Albert. “History of the Theoretical Group at the Radiation Laboratory” in Henry Guerlac Radar in World War II, 625-632. Los Angeles: American Institute of Physics/ Tomash Publishers, 1987.

In 1942, J. C. Slater and S. A. Goudsmit organized a Theory Group at the Rad Lab. Heins, a member of the group, recounts the group’s successes and provides insight into the (small but important) role played by theory at the Rad Lab.

The History of Science in the United States: An Encyclopedia. s.v. “World War II and science,” by Joel Genuth. Edited by Marc Rothenberg. New York: Garland, 2001.

This encyclopedia emphasizes physics in the Rad Lab and presents engineering practice as largely extraneous to the radar enterprise.

Hutchisson, Elmer. “The Egg or the Chick,” Journal of Applied Physics 10, no. 3 (March 1939): 141.

In a single-page editorial, Hutchisson argues that both engineering and physics draw strength from one another in equal measure and that no one field is superior.

Lauterbur, Paul C. “Nobel Lecture.” The Nobel Foundation, < http://nobelprize.org/medicine/laureates/2003/lauterbur-lecture.html>.

Nobel laureate in Medicine in 2003, Paul Lauterbur briefly reflects upon his engineering education.

Levanon, Nadav. Radar Principles. New York: John Wiley, 1988.

A textbook well described by its title, this work ignores engineering contributions to the development of radar.

Marshall, W. R. “Science Ain’t Everything.” Chemical Engineering Progress 60 (January 1964): 17-21.

Marshall gainsays the notion that physics and basic science are more important than engineering, and charges that engineering has not received nearly enough of the attention or freedom it deserves in the postwar era.

Mermelstein, Neil. “How Food Technology Covered Microwaves Over the Years.” Food Technology 51 (May 1997): 82.

This article offers minimal content, but it is a valuable resource for looking at the legacy of the Rad Lab from outside of a history of science perspective.

Miller Jon D. and Kenneth Prewitt. The Measurement of the Attitudes of The U.S. Public Toward Organized Science. Chicago: National Opinion Research Center for the Nation Science Foundation, 1979.

Miller and Prewitt review public opinion toward science in the 1970’s and compare it to conclusions researched in the 1950’s.

Millikan, Robert A. “The Relation of Science to Industry.” Science 69, no. 1776 (January 11, 1929): 27-31.

Millikan discusses the importance of “practical understandings” in science to society while concomitantly insisting upon the necessity of “pure science.”

Montgomery, C. G., R. H. Dicke, and Edward M. Purcell. Principles of Microwave Circuits. New York: McGraw-Hill Book Company, 1948.

Volume 8 of the Rad Lab series, this volume is both useful for the microwave circuit connoisseur, but proves especially interesting in displaying the forms of knowledge endemic to engineering.

The Oxford Companion to the History of Modern Science. s.v. “Radar,” by Woodruff T. Sullivan III. Edited by J. L. Heilbron. New York: Oxford University Press, 2003.

The radar effort is presented as the outgrowth of both engineering and physics expertise.

Peebles, Peyton Z. Radar Principles. New York: John Wiley, 1998.

This tome of a textbook is dedicated to James Maxwell and dedicates its historical attention to very general physics, with not a word about engineering.

Pollard, Ernest C. Radiation: One Story of the M.I.T. Radiation Laboratory. Durham, N.C.: The Woodburn Press, 1982.

Ernie Pollard worked at the Rad Lab from 1941-1945, and discusses both his memories and his thoughts on effective laboratory management, holding the Rad Lab up as a paragon.

Rabi, I. I. “The Physicist Returns From The War.” Atlantic Monthly 176, October 1945, 107-114.

The 1944 physics Nobel laureate summarizes the principles of radar and the nuclear bomb and discusses the many challenges of basic science remaining to physicists returning to the laboratory. Rabi, associate director of the Rad Lab, emphasizes that radar was principally a technical, technological achievement.

Rad Lab: Oral Histories documenting World War II Activities at the M.I.T. Radiation Laboratory. Principal Investigators Bryant, John, William Aspray, Andrew Goldstein, and Frederik Nebeker. Institute of Electrical and Electronics Engineers Inc., 1993. Piscataway, NJ: The Center for the History of Electrical Engineering, a joint venture of the IEEE, and Rutgers, The State University of New Jersey. Available at Institute Archives and Special Collections, M.I.T. Libraries [Available online: http://www.ieee.org/organizations/history_center/oral_histories.html]

This collection of oral histories offers innumerable insights into life at the Rad Lab and features discussions with, among others: Getting, Purcell, Pound, Ramsey, Pollard, and Davenport.

“Radiation Laboratory.” Wikipedia: The Free Encyclopedia, .

The international, free encyclopedia downplays to the point of oblivion the contributions of engineering to the Rad Lab.

Randall, John T. “Radar and the Magnetron.” Journal of the Royal Society of Arts 94: 4715, April 12, 1946, 303-323.

Randall provides a Briton’s interpretation of the process of radar and its legacy for England.

Reader’s Guide to the History of Science. s.v. “Physics—20th Century,” by David Kaiser. Edited by Arne Hessenbruch. Chicago: Fitzroy Dearborn, 2000.

This encyclopedia entry groups radar with proximity fuses and the nuclear bomb and lazily refers to all three of them as outgrowths of physics know how.

Rigden, John. S. and I. I. Rabi “Introduction” in Henry Guerlac Radar in World War II, xix-xxv. Los Angeles: American Institute of Physics/ Tomash Publishers, 1987

Rigden and Rabi provide the broadest historical context (war in Europe) under which the Rad Lab began and recount its primary aims at the time of its inception.

Robinson, Denis. “British Microwave Radar 1939-41.” Proceedings of the American Philosophical Society 127, no. 1 (February 1983): 26-31.

Robinson, who worked with Watson-Watt as well as at the Rad Lab, offers a primer on British microwave radar work; he highlights the pedagogical connections between Nobel laureate Ernest Rutherford and British radar pioneers.

Science and Engineering Indications. Washington D.C.: National Science Board, 2000.

––––––––. Washington D.C.: National Science Board, 1996.

Science and Engineering Indicators (Science Indicators before 1986), provide annual reviews of the American public’s estimations of the importance of science and engineering. As with any survey they are, of course, imperfect, but they do reflect important, well-researched trends in public confidence (or lack thereof) in science.

Science in the Early Twentieth Century: An Encyclopedia. s.v. “Radar.” Edited by Jacob Darwin Hamblin Santa Barbara, CA: ABC-CLIO, 2005.

This encyclopedia article ignores the role of engineering in the Rad Lab and construes it as a scientific achievement.

Science Indicators. Washington D.C.: National Science Board, National Science Foundation, 1976.

Like Science and Engineering Indicators, this provides a review of public attitudes towards Science in America.

Snow, Charles Percy. The Two Cultures and the Scientific Revolution. New York: Cambridge University Press, 1959.

In his famous examination of the divide between intellectuals in the humanities and intellectuals in the sciences, Snow analyzes the opinions of pure physicists towards applied scientists and engineers.

Social Science Research Council, Committee on Social and Economic Aspects of Atomic Energy. Public Reaction to the Atomic Bomb and World Affairs. Ithaca, NY: Cornell University, 1947.

A far-reaching report on the American public’s reaction to the nuclear bomb and its political consequences, this report says very little about perceptions of science, though the intensive surveys (Part II of the report) are more helpful in this regard.

Townes, Charles Hard. “Microwave Spectroscopy,” American Scientist 40 (1952): 270-290.

The inventor of the maser, Charles Townes, discusses the development of microwave methods in physics and mentions the debt of postwar physics to engineering.

A Walk through the 20th Century with Bill Moyers; Episode 15, I. I. Rabi: Man of the century. VHS. Created and developed by the Corporation for Entertainment and Learning Inc., and Bill Moyers. Washington, D.C.: Public Broadcasting Service, 1988.

Rabi reflects on, among other matters, the importance of the Rad Lab to the war effort and to himself.

Wiesner, Jerome. “Remembering the Rad Lab and RLE.” In Jerry Wiesner: Scientist, Statesman, Humanist; Memories and Memoirs edited by Walter Rosenbluth, 207-218. Cambridge, MA: The M.I.T. Press, 2003.

Wiesner’s recollections of his time reveal the adjustments made by a newcomer to the intellectual and organizational demands of the Rad Lab.

2) Archival Sources

Alvarez, Luis. Luis Alvarez Papers, MSS 84/82 cz. The Bancroft Library, University of California at Berkeley. Berkeley, California.

Alvarez’s correspondence with the National Academy of Engineers as well as with the National Academy of Sciences (he had membership in both groups) reveal his professional priorities. Of especial interesting historical note is the letter he wrote to his young son while on board the , just hours after the bombing of Hiroshima. [UC Berkeley’s Alvarez holdings are the same as those of the National Archives and Records Administration-Pacific Region, Record Group 434].

Bush, Vannevar. Vannevar Bush Papers, MC 78. Institute Archives and Special Collections, M.I.T. Libraries. Cambridge, Massachusetts.

Bush’s papers are especially enlightening for their analysis of the state of the engineer in America after the war, a rarely discussed topic in the literature.

Compton, Karl. Karl T. Compton Personal Papers, MC 416. Institute Archives and Special Collections, M.I.T. Libraries. Cambridge, Massachusetts.

Karl Compton’s papers are an eclectic bunch. What emanates from his correspondence most of all is the near-universal respect afforded him by nearly all who knew him. His papers are also particularly revealing for an investigator interested in both the philosophy of science as public servant and in debates over science during the depression.

Division of Engineering and Applied Physics. Records of the Division of Engineering and Applied Physics, UAV 362.7011 Pusey Library, Harvard University. Cambridge, Massachusetts.

The Division’s records reveal a marked bias in favor of theoretical approaches to experimental and engineering ones. The papers of Harvey Brooks (HUGFP 128.13), leader of the DEAP for two decades, provide a complement to the DEAP papers.

Ginzton, Edward L. Edward Ginzton papers, Special Collection 330. Special Collections and University Archives, Stanford University. Palo Alto, California.

Ginzton’s papers contain some (but not much) information relating to the developments of radar during the 1930’s and 1940’s.

Hansen, William W. William W. Hansen Papers, Special Collections 4 and 136. Special Collections and University Archives, Stanford University. Palo Alto, California.

Special Collection 4 focuses on Hansen’s research conducted at Stanford, notably his work on the klystron with the Varian brothers. His pre-Radiation Laboratory correspondence with Loomis, Bowles, Rabi and other founders of the Lab can be found in Special Collection 126.

Lawrence, Ernest Orlando. Ernest O. Lawrence Papers, MSS 72/117 c. The Bancroft Library, University of California at Berkeley. Berkeley, California.

Lawrence’s correspondence with Loomis is particularly well-detailed and reveals a great deal about their collaboration in bringing physicists to the Rad Lab.

Loomis, Alfred. Alfred Loomis Papers 1926-1975, MC 264. Institute Archives and Special Collections, M.I.T. libraries, Cambridge, Massachusetts.

Though Loomis directed the Rad Lab and the Microwave Committee, his papers at M.I.T. are very few and at best reveal a few anecdotes about his relationship with Compton and others.

Massachusetts Institute of Technology. Office of the President, 1930-1959 (Compton-Killian), AC 4. Institute Archives and Special Collections, M.I.T. libraries. Cambridge, Massachusetts.

The records of M.I.T. President Karl Compton before and during the war reveal his dedication to the Rad Lab and much of his motivation for helping to establish it.

M.I.T. Rad Lab, RG 227. Office of Scientific Research and Development, National Archives and Records Administration Northeast Region (Boston). Waltham, Massachusetts.

The Official documents of the Radiation Laboratory include laboratory notebooks of the researchers, patent files, interim reports, committee meeting minutes, project summaries, and assorted papers and photos relating to work done at the Radiation Laboratory throughout the war. The sheer volume of information (the finding aid alone is 675 pages) makes it essential to know exactly what time period and what type of information relates to one’s research.

Weisskopf, Victor. Papers, MC 572. Institute Archives and Special Collections, M.I.T. Libraries. Cambridge, Massachusetts.

Weisskopf’s papers, among other things, include interesting facts about the state of pre- war physics, especially its emphasis on the pragmatic.

3) Newspaper Articles I have only annotated the uniquely revealing or noteworthy articles. Nevertheless, every one of articles below helped provide me with a general feel for the legacy of Rad Lab.

“20 Scientists Quit Missile Project.” New York Times, December 15, 1955, pg. 1. http://www.proquest.com

A 1955 dispute over missile production led to disagreements between engineers and physicists over administrative authority. This led to the resignation of senior scientific staff in a postwar case study of scientists and engineers plainly in conflict.

“5, 750 Canadians Helped Britain’s Radar Battle,” New York Times, August 15, 1945, pg. 14. http://www.proquest.com

“Asserts Radar Won Battle of Britain.” New York Times, August 15, 1945, pg 14. http://www.proquest.com

“Astronauts Relax and Begin Taping Their Stories,” New York Times, July 28, 1969, pg 18. http://www.proquest.com

Demonstrates the prominence given to science in accounts of the moon landings.

Beam, Alex. “Basement Science Project.” New York Times, June 16, 2002, pg. F14. http://www.proquest.com

Book review of Conant’s Tuxedo Park.

Crease, Robert P. and Charles C. Mann. “Gambling with the Future of Physics.” The New York Times, December 5, 1982, pg. SM66. http://www.proquest.com

“Dr. DuBridge Installed: Takes Over as Head of California Institute of Technology.” New York Times, Nov 13, 1946, pg. 8. http://www.proquest.com

“Dr. Henry Straus, Radar Expert, 43: Lincoln Laboratory Physicist Dies—Aided Development of Detecting Device.” New York Times, September 24, 1957, pg. 35. http://www.proquest.com

Fenton, John. “MIT offers Spur to Area Economy.” New York Times, January 15, 1961, pg. 72. http://www.proquest.com

Fenton, John. “Scientists Pay Homage to MIT at 20th Anniversary of the Research Laboratory of Electronics.” New York Times, May 15, 1966, pg. 78. http://www.proquest.com

Fiske, Edward. “Engineering School Shortcomings Lead to US Lag.” New York Times, March 28, 1982, pg. HT6. http://www.proquest.com

Gleick, James. “Columbia Lauds Rabi As Its “Brilliant Jewel.” New York Times, November 21, 1985, pg. B1. http://www.proquest.com

Hilts, Philip. “Last Rites for a Plywood Palace That Was a Rock of Science.” New York Times, march 31, 1998, pg. F4. http://www.proquest.com

“Jacob Millman. Expert on Radar, Dies at 80,” New York Times, May 24, 1991, pg. B8. http://www.proquest.com

Lieberman, Henry. “Technology: Alchemy of Route 128,” New York Times, January 8, 1968, pg. 139. http://www.proquest.com

Pace, Eric. “Jerome B. Wiesner, President of M.I.T., Is Dead At 79.” New York Times, October 23, 1994, pg. 45. http://www.proquest.com

“Physicist-Engineer [Display Ad].” New York Times, January 20, 1946, pg. 61. http://www.proquest.com

“Praises U.S. Scientists: Duc de Broglie Says the World Watches Them with Eagerness.” New York Times, May 11, 1935. http://www.proquest.com “Radar Sank German Ships Hidden 20 Miles Away,” New York Times, August 15, 1945, pg. 14. http://www.proquest.com

“Radar to be Used on Merchant Ships,” New York Times, August 15, 1945, pg. 14. http://www.proquest.com

“Radio Jobs Open for War.” New York Times, November 3, 1942, pg. 20. http://www.proquest.com

The first citation of the Rad Lab in the New York Times announced the recruitment of women to the laboratory as secretaries.

Reinhold, Robert. “Dr. Vannevar Bush is Dead at 84,” The New York Times, June 30, 1974, pg 1. http://www.proquest.com

“Revolt Against Physics,” The New York Times, February 12, 1968, pg. 38. http://www.proquest.com

An editorial on the decline (though certainly not the end) of the prestige of physics, based on observations of an American Physical Society conference.

Saxon, Wolfgang. “Lee Alvin DuBridge, 92, Ex-President of Caltech.” New York Times, January 25, 1994, pg. B8. http://www.proquest.com

“Secrets of Radar Given to World.” New York Times, August 15, 1945, pg 1. http://www.proquest.com

The cover story article on V-J Day after the Joint Board on Scientific Information Policy revealed the role of radar during the war to the press.

Stuart, John. “Radar to Control All Traffic in Air: Devices Which Will Map Airways for Whole Country Were Developed for War Use.” New York Times, September 16, 1945, pg. 32. http://www.proquest.com

The second, full-length New York Times article to deal with the Rad Lab (the first, from September 9th, dealt with coastal navigation but said nothing about the actual Rad Lab work), portrays the lab as performing numerous tasks to improve electronics for radar.

Trefil, James. “Memoirs of a Master Tinkerer.” New York Times, June 7, 1987, p. BR14. http://www.proquest.com

“Tuller, Alexandria Executive, Feared Dead in KLM Crash.” The Washington Post and Times Herald, September 6, 1954, pg. 7. http://www.proquest.com

“Vital War Devices Born on Campuses.” New York Times, October 9, 1945, pg. 8. http://www.proquest.com

Wald, Matthew. “Jam Sessions.” New York Times, June 22, 1997, pg. BR31. http://www.proquest.com

Book review of Buderi’s The Invention that Changed the World. “Wiesner is Scientist-Statesman.” The Washington Post, Times Herald, January 12, 1961, pg. A2. http://www.proquest.com

“World’s Arsenal Revamped by War.” New York Times, August 15, 1945, pg 14. http://www.proquest.com

4) Popular Periodicals

“Exit the Amateur Scientist.” The Nation 83, August 23, 1906, 159-161.

A turn-of-the-century lament of the specialization of science.

“Longhairs and Short Waves,” Fortune November 1945, 163-169, 206, 208.

Investigates the importance and contributions of physicists (longhairs) at the Rad Lab.

Masters, Dexter. “We Outsmarted Them on Radar.” Saturday Evening Post 218, September 8, 1945, 20-21, 109-110.

A Rad Lab alumnus summarizes the many achievements and methods of radar for this popular magazine.

“May Lift Veil on Radar.” Business Week, February 10, 1945, 5.

Speculations about then-secret wartime radar discoveries.

“No Baldness Either.” Time 47, May 27, 1946, 63.

Several radar workers became preoccupied with health fears because of their microwave radiation exposure. They were fine.

“Peacetime Radar.” Time 46, September 24, 1945, 76-77.

Brief analysis of radar’s postwar potential that includes “more radar news” released from “the M.I.T. Radiation Laboratory.”

“Popular Appreciation of Scientists.” The Nation 74, January 16, 1902, 46-47.

Reveals some of the popular thinking concerning scientists at the turn of the century, with great similarities in certain ways to postwar perceptions of scientists.

“Radar.” Time 46, August 20, 1945, 78-82.

Intended as the cover article of the first postwar issue, the story of radar though “one of the great and lasting achievements of modern science” was relegated to p. 78 because of “more urgent” news: the atomic bomb.

“Radar: Another of the War’s Great Secret Weapons is Revealed.” Life, August 20, 1945, 96-99.

Life Magazine’s brief postwar expose on what quickly became known as the second greatest weapon of the war. The article on the atomic bomb in the same issue is more than ten times as long.

“Radar: All-Seeing Eye that Doomed the Enemy.” Newsweek, August 20, 1945, 40-42.

Newsweek Magazine’s (very brief) postwar expose on radar.

“Radar: Background of War’s Greatest Development.” Scientific American 169, August 1943, 78- 80.

In an era when no one knew how radar worked, Scientific American discusses the basic, publicly-known history of radar. Microwave radar remained a secret until after V-J Day.

“Radar: The Industry.” Fortune 32, October 1945, 146, 200-201.

A discussion of the emergent—and patent-argument riddled—radar industry after the war as well as its progenitors, including the Rad Lab.

“Radar: The Technique.” Fortune 32, October 1945, 138-145, 196, 198, 200.

A discussion of how radar works, including some history of its development.

“Radiation Laboratory Record Forecasts Electronic Advances: Wartime success of cooperative scientific enterprise at MIT kept nation ahead in radar research.” Aviation News 4, no. 28, October 8, 1945, 28.

The Rad Lab physicists are praised for their contributions to radar development and future physics research is encouraged.

“Radio, Refrigerators, and Radar.” Fortune 30, November 1944, 114.

Considers radar’s long-term, industrial marketability, especially during peacetime.

“Speaking of Pictures,” Life, October 22, 1945, 12-14.

One of the first large-scale presentations of “radar maps” to the American public.

“Who Invented Radar?” Radio News 30, July 1943, 4, 70-72.

As microwave radar secretly established its dominance over the U-boat, many individuals speculated about how specifically it operated and who had invented it.

Wickware, Francis. “Manhattan Project,” Life, 19, August 20, 1945, 92-95, 100-102.

Life’s postwar article on the nuclear bomb project, a substantially more in-depth article than the one on radar.

5) Secondary Sources

Abelson, Philip. “Additional Sources of Financial and Political Support for Science.” Science, 180, no. 4083 (April 20, 1973): 259.

Abelson, editor of Science, laments the decline of academic science’s prestige and public support, noting especially the deterioration of the former amity among government, industry, and the university.

––––––––. “Are the Tame Cats in Charge?” Saturday Review 49, January 1, 1966, 100-103.

While accepting that science and technology have greatly augmented the success of the United States as a global leader, Abelson maintains that high-energy physics has made few contributions to modern society, especially given the tremendous amount of federal investment it receives.

“About the National Science Foundation,” National Science Foundation, .

The National Science Foundation (NSF), founded in 1950, brought to fruition many of the principal ideas Vannevar Bush put forward in Science, The Endless Frontier. As of 2006, the NSF disbursed more than $5 billion of research funds annually to support “basic research.” Its website illuminates how the NSF sees its role in terms of navigating the demands of the engineering and scientific communities.

Aldersey-Williams, Hugh. Findings: Hidden Stories in First-hand Accounts of Scientific Discovery. Norwich: Lulox books, 2005.

This book traces major scientific and technological breakthroughs in each decade of the 20th century. Its brief section on transistors casts some light on Randall and Boot’s cavity magnetron research.

Armytage, Richard. A Social History of Engineering. Cambridge, MA: The Massachusetts Institute of Technology Press, 1961.

A straightforward history of British engineering, A Social History says little about engineering as a social phenomenon, despite its title. Chapter 27, “The Endless Frontier,” partially delves into engineering’s social role, mentioning Wittgenstein’s technical career.

Barnes, Barry. “On the ‘Hows’ and ‘Whys’ of Cultural Change (Response to Woolgar).” Social Studies of Science 11, no. 4 (November 1981): 481-498.

––––––––. “The Science-Technology Relationship: A Model and a Query.” Social Studies of Science 12, no. 1 (February 1982): 166-172.

In “Science-Technology Relationship,” Barnes takes the interactive (non-linear) model of the relationship between technology and science and suggests that it be applied to science’s relationship with all other fields and factors like society or government. Barnes explores this impact of culture on science more theoretically in “One the ‘Hows’ and ‘Whys,’” arguing for the relativistic nature of theory.

Barnes, G. M. “Ordinance Research: The Department’s Use of Technical and Scientific Facilities.” Army Ordinance 21, no. 122 (September-October 1940): 110-128.

Colonel Barnes made public certain information about the nation’s preparedness for the threat of war. He describes some of the technological preparations made during the American Revolutionary War.

Barthes, Roland. “The Death of the Author.” In Image-Music-Text, edited and translated by Stephen Heath, 142-148. London: Fontana, 1977.

Citing Proust and other authors who blur the relationship between author and character, Barthes maintains that the author is born with the text, and hence does not merit special consideration in the interpretation of their own texts.

Baxter, James Phinney 3rd. Scientists Against Time. Boston: Little, Brown, and Company, 1946.

The official history of the Office of Scientific Research and Development, Scientists recounts the contributions of science to the war effort. It quotes Axis sources and leaders that most other English-language radar books ignore.

Bernstein, Jeremy. “Profiles: I. I. Rabi.” The New Yorker 51, October 13 and 20, 1975.

A long-winded portrait of Rabi, Bernstein’s second article (October 20th) discusses the Rad Lab’s historical and technical roots and some of its achievements.

Biagioli, Mario. “Aporias of Scientific Authorship.” In The Science Studies Reader, edited by Mario Biagioli, 12-30. New York: Routledge, 1999.

Biagioli analyzes the two principal credit systems in society: the market system of patents and the scientific system of discovery of what are considered to be impossible-to-own facts of nature. Though his analysis does not deal with it directly, it does come into contact with some of the key issues of assigning credit to a discipline like engineering that is concerned with both artifacts and abstract knowledge.

Birr, Kendall A. “Science in American Industry.” In Science and Society in the United States, edited by David Van Tassel and Michael G. Hall, 35-81. Homewood, IL: The Dorsey Press, 1966.

A review of the growing role of American scientists in industrial laboratories, “Science in American Industry” focuses predominantly upon the 19th century and early 20th century.

Brown, Louis. A Radar History of World War II: Technical and Military Imperatives. Bristol, UK: Institute of Physics Publishing, 1999.

Brown looks at the simultaneous development of radar across the United States as well as the impact of secrecy restraints on the technology’s development. He offers a useful history of American radar research during the 1930’s.

Buderi, Robert. The Invention that Changed the World. New York: Simon & Schuster, 1996.

The first quarter of Buderi’s book gives a detailed analysis of the Rad Lab culture and inventions, as well as the Lab’s place in the larger picture of World War II. Buderi is unique among American historians in providing a detailed portrait of the history of radar research in England.

Burchard, John. Q.E.D.: M.I.T. in World War II. New York: John Wiley & Sons Inc, 1948.

One of the earliest works to treat the history of the Rad Lab, this book ecstatically praises M.I.T.’s numerous contributions to the war effort and offers a window into the scientific and technological self-congratulatory euphoria in the years after war.

Burns, Alfred. “The Tunnel of Eupalinus and the Tunnel Problem of Hero of Alexandria.” In Philosophers and Machines, edited by Otto Mayr, 50-63. New York: Science History Publications, 1976. [Originally printed under the same title in Isis 62 (1971): 172-185].

Burns’s article provides an important rebuttal to the arguments of Goodfield and Toulmin over the nature of the construction of the ancient tunnel at Samos. His conclusion that engineering methods are equally possible to mathematical-theoretical ones, however, further evinces the haste with which other scholars accepted a theoretical explanation over an engineering one.

Burns, Russell ed. Radar Development to 1945. London: Peter Peregrinus Ltd., 1988.

This collection of essays and memoirs on the history of radar development, deals especially with Britain, but also contains a detailed chronicle of the German radar program.

Byron, Lord George Gordon. The Complete Work of Lord Byron. Edited by John Galt. Paris: Baudry’s European Library, 1837.

In “The Dream,” Byron describes a unique world.

Callick, E. B. Metres to Microwaves. London: Peter Peregrinus Ltd., 1990.

A detailed history of British radar work, Metres is helpful for its detailed discussion of radar component development before the war and during its early stages.

Coben, Stanley. “The Scientific Establishment and the Transmission of Quantum Mechanics to the United States, 1919-32.” The American Historical Review 76, no. 2 (April 1971): 442-466.

A portrait depicting the rise of American Physics as much as the actual transmission of Quantum Mechanics to the United States, this article presents the basic story of some of 1920’s physics major developments and its key figures.

Collins, Harry M. “The TEA Set: Tacit Knowledge and Scientific Networks.” Science Studies 4, (1974): 165-186.

––––––––. “What is Tacit Knowledge.” In The Practice Turn in Contemporary Theory, edited by Theodore R. Schatzki, Karin Knorr-Cetina, & Eike von-Savigny, 107-119. London: Routledge, 2001.

Collins’ famous studies of tacit knowledge offer a theoretical approach to understanding the unspoken knowledge of scientific practice, especially within the laboratory. It did not prove especially necessary to this study of the Rad Lab, but undoubtedly could serve as a cooperative investigator for future research, particularly in the realm of further specifying the types of knowledge indigenous to engineers.

Conant, Jennet. Tuxedo Park: A Wall Street Tycoon and the Secret Palace of Science that Changed the Course of World War II. New York: Simon & Schuster, 2002.

Conant dedicates the last third of her book to the early work at the Radiation Laboratory. She principally relates a story of progress of inventing, also placing emphasis on the historically neglected figure Alfred Loomis, head of the wartime Microwave Committee.

Condon, Edward Uhler. “Some Thoughts on Science in the Federal Government.” Physics Today 5 (April 1952): 6-13.

When he retired from directing the National Bureau of Standards, Condon reflected on the inefficiencies of government and science policy, including the relationship of science, government, and the public.

Cook, Charles and Marvin Bernfeld. Radar Signals: An Introduction to Theory and Application. New York: Academic Press, 1967.

A solid, detailed review of the techniques and theory of radar pulses.

Derrida, Jacques. Writing and Difference. Annotated and translated by Alan Bass. London: Routledge Classics, 2001.

Derrida argues, most clearly in his essay, “Structure, Sign, and Play in the Discourse of the Human Sciences,” that most philosophies—or any structure of thought or society— have centers and that these centers limit the degrees of freedom of thought (the play) of any given system. Authors may be thought of as center texts, as non-substitutable.

Dierkes, Meinolf and Claudia von Grote, eds. Between Understanding and Trust: The Public, Science and Technology. London: Routledge, 2003.

Fretful over the decline of public trust in science, this volume analyzes the history of the dynamics between the public and scientists.

Dupré, J. Stefan and Sanford A. Lakoff. Science and the Nation. Englewood Cliffs, N.J.: Prentice Hall, Inc., 1962.

One of the first academic studies of the relationship between postwar government and postwar science, Science and the Nation looks at decision making, defense research and basic research and their significance for the United States.

Dupree, A. Hunter. “The Great Instauration of 1940.” In The Twentieth-Century Sciences: Studies in the Biography of Ideas edited by Gerald Holton, 443-468. New York: W. W. Norton and Company Inc., 1972.

Dupree analyzes the structure of American scientific organization in the form of the OSRD and its relationship to the scientific organization theories of Francis Bacon and René Descartes.

Edgerton, David. “‘The Linear Model’ Did not Exist: Reflections on the History and Historiography of Science and Research in Industry in the Twentieth Century.” In The Science- Industry Nexus: History, Policy, Implications, edited by Karl Grandin and Nina Wormbs, 31-57. New York: Watson, 2004.

Edgerton vociferously argues against the existence of a ‘linear model’ of science leading to technology in public policy. He reevaluates Bush’s Science, The Endless Frontier, and contends that the real problem plaguing the historiography of science and technology is an overemphasis on academic research more generally.

Elkana, Yehuda and others, eds. The Advent of Science Indicators. New York: John Wiley and Sons, 1978.

Dealing especially with themes of the quantitative versus the qualitative, The Advent pays special attention to the limits of science indicators and the difficulties of accurately depicting the direction of science.

Etzioni, Amitai and Clyde Nunn. “The Public Appreciation of Science in Contemporary America.” Daedalus 103, no. 3 (Summer 1974): 191-205.

A study of public attitudes towards both science and technology, Etzioni and Nunn’s report analyzes responses by education, geography, and socio-economic status.

Ferguson, Eugene S. Engineering and the Mind’s Eye. Cambridge, MA: The M.I.T. Press, 1992.

Ferguson argues that design is as essential to the work of the engineer as any mathematical analytical tool. Design went underemphasized especially after World War II. Nevertheless, it is an essential part of what the engineer knows.

Forman, Paul. "Behind Quantum Electronics: National Security as Basis for Physical Research in the United States, 1940-1960.'' Historical Studies in the Physical and Biological Sciences 18 (1987): 149-229.

–––––––. “‘Swords into ploughshares’: Breaking new ground with radar hardware and technique in physical research after World War II.” Review of Modern Physics 67, no. 2 (April 1995): 397- 455.

From his survey of the marked increase in defense funding for science following the second World War, Forman contends in “Behind” that defense funding limited the freedom of research traditionally enshrined in the academy. In “Swords,” he argues that radar hardware shaped the character of postwar physics. The direction (and magnitude) of postwar research differed in substantial ways from its state before the war; Forman concludes that technology greatly impacted the way in which physics is conducted, especially what physics is conducted.

Fortun, M. and S. S. Schweber. “Scientists and the Legacy of World War II: The case of Operations Research.” Social Studies of Science 23 (1993): 595-642. Fortun and Schweber look at Operations Research—especially management science— after World War II among both American and Britain physicists. The authors consider, in particular, matters of scientific authority and the appropriation of power.

Foucault, Michel. “What is an Author?” In Textual Strategies: Perspectives in Post-Structuralist Criticism, edited by Josué V. Harari, 141-160. Ithaca, NY: Cornell University Press, 1979.

Foucault dissects the close links between author and authority, and the constraints placed on interpretive frameworks by limiting thought to within the boundaries of authorial intention. “What is an Author” provides a framework for thinking about the primacy of the individual actor for the assignation of credit.

Frayling, Christopher. Mad, Bad and Dangerous?: The Scientist and the Cinema. London: Reaktion Books, 2005.

Though, of course, movies are deeply constrained by the need to entertain, which is not the sole aim of the media, this book provides effusive rundown of the scientist as madman and creator in popular films of the 20th century, and is a great resource for any study of the scientist in the public sphere.

Galison, Peter. Image and Logic: A Material Culture of Microphysics. Chicago: University of Chicago Press, 1997.

––––––––. "Kuhn and the quantum controversy". British Journal for the Philosophy of Science 32 (1981): 71-85.

Galison’s famous study principally explores the logic (calculating machine) and image (picture) traditions in physics. In doing so, among many ideas, Image investigates the transformation of the relationship between scientists and engineers over the course of the 20th century. Chapter nine’s section on the Rad Lab is especially useful. “Kuhn,” is a discussion of the theories of Kuhn, Klein, and others concerning the hypothesis that Planck was not first to recognize the inevitability of quantization arising from his own theory.

Genuth, Joel. “Microwave Radar, the Atomic Bomb, and the Background to U.S. Research Priorities During World War II.” Science, Technology, and Human Values 13, no. 3/4 (1988): 276-289.

A brief study of United States science and its organizational decisions during the War, focusing in particular on the experiences and penchants of Vannevar Bush.

Gilkey, Carolyn. “The Physicist, the Mathematician and the Engineer: Scientists and the Professional Slur.” Western Folklore 49, no. 2 (April 1990): 215-220.

A fascinating look at the social dynamics and standing of physicists, mathematicians, and engineers. Working from popular slurs and jokes, Gilkey argues that the practical- mindedness but asininity associated with engineers is deeply embedded in the physics community.

“Glasses,” Wikipedia: The Free Encyclopedia, .

This website features a partly speculative history of spectacles, but confirms that engineering of corrective spectacles predated Kepler’s optical theory by centuries.

Goodfield, June and Stephen Toulmin. “How Was the Tunnel of Eupalinus Aligned?” In Philosophers and Machines, edited by Otto Mayr, 38-47. New York: Science History Publications, 1976. [Originally printed under the same title in Isis 56 (1965): 46-55].

Goodfield and Toulmin refute the generally accepted, geometrical principle explanation regarding the tunnel at Samos, and instead argue that practical engineering considerations dominated the concern of the architect, Eupalinus. As a result, they demonstrate the way in which an engineering explanation can be equally viable to a theoretical (scientific) one, and yet be ignored.

Greenberg, D. S. “The National Academy of Sciences: Profile of an Institution (III).” Nature 156, no. 3774 (April 1967): 488-493.

Greenberg discusses, among other developments at the NAS, the emergence of the National Academy of Engineering and its impact on the relationship of physicists and engineers.

Gregory, Jane and Steve Miller. Science in Public: Communication, Culture, and Credibility. New York: Plenum Press, 1998.

An investigation into the rise of scientific communication, this book covers a broad range of topics in the relationship between science, the media, and the public, but I found it useful mostly for its histories of public opinion, media and scientists in the postwar era.

Guerlac, Henry. Radar in World War II. Los Angeles: American Institute of Physics/Tomash Publishers, 1987.

––––––––. “The Radio Background of Radar.” Journal of the Franklin Institute 250, no. 4 (October 1950): 285-308.

Radar is the most complete (and ponderous) detailed account of the development and deployment of radar before and during the Second World War. Guerlac worked at the Rad Lab as the official historian starting in 1942 and interviewed nearly every major administrator and scientific leader at the Rad Lab while it was still in operation. His “Radio Background of Radar,” provides a concise version of the first two hundred pages of Radar, and also discusses the types of researchers—predominantly those with experience in studying the —who pursued early radar investigations.

Guerlac, Henry and Marie Boas. “The Radar War Against the U-Boat.” Military Affairs 14, no. 2 (Summer 1950): 99-111.

One of the most laconic works on radar and submarines during the war, this is also one of the most lucid. Guerlac and Boas provide general overviews of each of seven phases of combat, with especially useful summaries of radar countermeasures and counter- countermeasures.

Hall, John, ed. Radar Aids to Navigation. New York: McGraw-Hill Book Company Inc., 1947.

Radar Aids to Navigation is one of 28 volumes published after World War II in order to disperse the many technical and theoretical discoveries of the Radiation Laboratory during the war. This volume in particular is helpful for its presentation of the concepts of radar resolution and range.

Heilbron, John L. and Bruce Wheaton. Literature on the History of Physics in the 20th Century. Berkeley, CA: Office for History of Science and Technology, 1981.

An excellent bibliographic resource. Especially useful was Section V on “Physics and Society.”

Hessen, B. “The Social and Economic Roots of Newton’s ‘Principia.’” In Science at the Cross Roads, edited by Joseph Needham, 149-212. London: Frank Cass and Company, 1971), 206.

A strict Marxist reading of the scientific revolution, “The Social and Economic Roots” is one of the first histories emphasizing the impact of technology on science.

Hoddeson, Lillian. “Research on Crystal Rectifiers during World War II and the Invention of the Transistor.” History and Technology 11 (1994): 121-130.

Hoddeson studies the impact of crystal receiver wartime research on the postwar development of the transistor. She examines how defense demands forged new relationships between engineers and scientists.

Hoddeson, Lillian, Paul Henriksen, Roger Meade, and Catherine Westfall. Critical Assembly: A Technical History of Los Alamos during the Oppenheimer Years, 1943-1945. Cambridge, England: Cambridge University Press, 1993.

Critical Assembly offers an analysis of the Manhattan Project through the special lens of the methodology of construction. Chapters 1 and 20 are particularly valuable for considering the analogous integration of skills used in radar’s development.

Hughes, Thomas P. American Genesis: A Century of Invention and Technological Enthusiasm 1870-1970. New York: Viking, 1989.

This book explores the role of technology in international development and societal transformation. It features a terse history of engineering in America and also explores some of the theoretical aspects of the impact of technology upon society and vice-versa.

Jasanoff, Sheila. “The ‘Science Wars’ and American Politics.” In Between Understanding and Trust: The Public, Science and Technology, edited by Meinolf Dierkes and Claudia von Grote, 39-59. London: Routledge, 2003.

Jasanoff probes the social and political factors constraining any objective measurement of the public comprehension of science, emphasizing “individualist” predilections in American culture.

Jungnickel, Christa and Russell McCormmach. Intellectual Mastery of Nature Vol. 2: The Now Mighty Theoretical Physics 1870-1925. Chicago: University of Chicago Press, 1986

This famous study of the rise of German theoretical physics provides descriptions of how German physics grew, descriptions useful for their insight into the tension between theoretical and experimental physics.

Kaiser, David. “The Atomic Secret in Red Hands? American Suspicions of Theoretical Physicists During the Early Cold War.” Representations 90 (Spring 2005): 28-60.

––––––––. Drawing Theories Apart: The Dispersion of Feynman Diagrams in Postwar Physics. Chicago: University of Chicago Press, 2005.

––––––––. “Nuclear Democracy: Political Engagement, Pedagogical Reform, and Particle Physics in Postwar America." Isis 93 (June 2002): 229-268.

––––––––. “Scientific Manpower, Cold War Requisitions, and the Production of American Physicists after World War II.” Historical Studies in the Physical and Biological Sciences 33 (Fall 2002): 131-159.

Evaluating the especial prosecution of theoretical physicists by the House on Un- American Activities, “Atomic Secret” offers a window into the legacy of the early Cold War and the distribution of credit for nuclear secrets—both malicious and salutary— among World War II scientists, engineers, and technicians. Drawing takes up issues of dispersion and pedagogy in physics and along the way paints a helpful portrait of the postwar physics population boom in America—a portrait emphasized in “Scientific Manpower.” “Nuclear” is the story of Geoffrey Chew and his S-Matrix theory as well as his more egalitarian approach to graduate-student education.

Kemper, John Dustin. The Engineer and His Profession. 2nd ed. New York: Holt, Rinehart, and Winston, 1975.

A guidebook to the newly-minted engineer uncovering the world of his or her discipline, The Engineer says a fair amount about the place of the engineer (both perceived and real) in society.

Kevles, Daniel. The Physicists: The History of a Scientific Community in America. New York: Vintage Books, 1979.

––––––––. Personal Correspondence. 22 February 2006.

The Physicists is Kevles’s narrative of American physics from 1865 until roughly 1970. For the study of radar, it offers a very useful background on physicists war research during World War I, as well as the development of the physics community in American between the world wars. In February 2006, Professor Kevles shared his thoughts with the author on the Rad Lab legacy.

Kline, Ronald. “Construing ‘Technology’ as ‘Applied Science’: Public Rhetoric of Scientists and Engineers in the United States, 1880-1945.” Isis 86, no. 2 (June 1995): 194-221.

Kline argues that engineers attempted to cast themselves as scientists in order to augment their status as “learned” men. Scientists emphasized the centrality of basic science in order to win support for their research.

Korol, Alexander. Soviet Education for Science and Technology. Cambridge MA: Technology Press of Massachusetts Institute of Technology, 1957.

As the title makes evident, this book is an examination of the Soviet education system, with special emphasis placed on its comparison to the American methodology of teaching science. Of particular interest is Appendix J, “Soviet References to The American System of Higher Technical Education,” in other words, a look at what the Soviet educators looked at when they looked at the United States before the United States looked back at them. The appendix is less of a mouthful.

Kuhn, James W. Scientific and Managerial Manpower in Nuclear Industry. New York: Columbia University Press, 1966.

Kuhn offers a case study on the organization of scientists and engineers during the Cold War.

Kuhn, Thomas. Black-Body Theory and the Quantum Discontinuity 1894-1912. Oxford: Oxford University Press, Inc, 1978.

––––––––. The Essential Tension: Selected Studies in Scientific Tradition and Chance. Chicago: University of Chicago Press, 1977.

Black-Body Theory is the study of Planck’s development of black-body theory and quantized energy states. Kuhn’s analysis investigates how scientists are unaware of the import and nature of their own work. The Essential Tension is a hodgepodge of essays, with “The Relations between History and History of Science,” especially informative for the historiography of science and engineering.

Küppers, Günter. “On the Relation between Technology and Science.” In The Dynamics of Science and Technology, edited by Wolfgang Krohn, Edwin T. Layton Jr. and Peter Weingart, 113-136. Dordrecht, Holland: D. Reidel Publishing Inc., 1978.

Taking up the fields of thermodynamics and fluid mechanics as his field of study, Küppers studies the production of knowledge arising from the interaction of technology and science.

Lasby, Clarence G. “Science and the Military.” In Science and Society in the United States, edited by David Van Tassel and Michael Hall, 251-282. Homewood, IL: The Dorsey Press, 1966.

Lasby places special attention on engineers in his history focused on World War II and postwar developments.

Latour, Bruno. Science in Action. Cambridge, Massachusetts: Harvard University Press, 1987.

Combining methodologies of philosophy, sociology, and history of science, Latour investigates how scientists and engineers operate, insisting on the importance of looking at practice above all else.

Latour, Bruno and Steve Woolgar. “The Cycle of Credibility.” In Science in Context: Readings in the Sociology of Science, edited by Barry Barnes and David Edge, 35-43. Milton Keynes: Open University Press, 1982. Latour and Woolgar review two other theories of scientific credibility as an economic model and break with both, concluding that the scientist’s credibility objective is to speed up the very cycle by which he or she receives credibility (and funding) in the first place.

Layton, Edwin T. Jr. “American Ideologies of Science and Engineering.” Technology and Culture 17, no. 4 (October 1976): 688-701.

Layton Jr. considers three ideologies spanning the physics-engineering spectrum: basic science, engineering science, and design. Looking in particular at the inventions of Benjamin Isherwood and Robert Thurston as well as J. A. L. Waddell and John Harrington, Layton Jr. probes the rise of engineering science and design, and analyzes the social standing concerns of the engineering community.

––––––––. “Conditions of Technological Development.” In Science Technology and Society, edited by Ina Spiegel-Rösing and Derek de Solla Price, 197-221. London: Sage Publications, 1977.

The middle sections of “Conditions” especially concern themselves with innovation and the importance of engineering as its own form of knowledge production in a terse, lucid manner. The article also highlights the importance of thinking about science and technology as communities of thinkers, not as a distinction simply between “abstract functions of knowing and doing.”

––––––––. “Mirror-Image Twins: The Communities of Science and Technology in 19th-Century America.” Technology and Culture 12, no. 4 (October 1971): 562-580.

In this seminal work on the ‘technology community,’ Layton Jr. demonstrates how the emergent engineering population structured its societies, journals and pedagogical methods very much on the mold of those of the scientific community.

––––––––. The Revolt of the Engineers. Cleveland, OH: The Press of Case Western Reserve University, 1971.

Revolt treats the American engineering community in the context of Progressivism and the Depression during the first three decades of the 20th Century. The epilogue quickly discusses the engineering community after World War II.

––––––––. “Science as a Form of Action: The Role of the Engineering Sciences.” Technology and Culture 29, no. 1 (January 1988): 82-97.

“Science as a Form of Action” seeks to refute arguments that ignore the unique nature of the engineering sciences.

––––––––. “Technology as Knowledge.” Technology and Culture 15, no. 1 (January 1974): 31- 41.

Like “Science as a Form of Action,” but coming much earlier, “Technology as Knowledge,” spotlights engineering’s own methodology, as well as its particular community and social history.

Leslie, Stuart. The Cold War and American Science: The Military-Academic-Industrial Complex at MIT and Stanford. New York: Columbia University Press, 1993.

––––––––. Profit and Loss: The Military and MIT in the Postwar Era,” Historical Studies in the Physical Sciences 21 no. 1 (1990): 59-85.

Leslie analyzes M.I.T.’s and Stanford’s expansive science budgets during the Cold War. “Profit and Loss” is a condensed version of Leslie’s analysis of M.I.T. in The Cold War; both describe the reaction of M.I.T. scientists to the foundation of the Rad Lab by scientists and bureaucrats outside of M.I.T. departments such as Loomis and DuBridge.

Lewenstein, Bruce. “Frankenstein or Wizard: Images of Engineers in the Mass Media.” Engineering: Cornell Quarterly 24, no. 1 (1989): 40-48.

Engineers can be portrayed as heroic, humdrum, or venomous, argues Lowenstein, who posits that negative images in the media of engineers result in consequences including turning potential students away from the field.

Longfellow, Henry Wadsworth. The Poetical Works of Longfellow. Edited by George Monteiro. Boston: Houghton Mifflin Company, 1975.

Contains “Paul Revere’s Ride,” “The Jewish Cemetery at Newport,” and other classics.

Lowood, Henry. “A Race on the Edge of Time. Radar—The Decisive Weapon of World War II.” The Journal of Military History 54, no. 4 (October 1990): 519-520.

In an acerbic review of David E. Fisher’s radar history, Lowood emphasizes historical elements easily overlooked: radar countermeasures, the administration of large-scale projects, and Operations Research.

MacKenzie, Donald. & Graham Spinardi. “Tacit Knowledge, Weapons Design and the Uninvention of Nuclear Weapons.” American Journal of Sociology 101, No. 1, (1995): 44-99.

MacKenzie and Spinardi propose that tacit knowledge is so essential to development that the silence of a generation of physicists could lead to the “uninvention” of nuclear technology.

Manegold, Karl-Heinz. “Technology Academised: Education and Training of the Engineering the Nineteenth Century.” In The Dynamics of Science and Technology, edited by Wolfgang Krohn, Edwin T. Layton Jr. and Peter Weingart, 137-158. Dordrecht, Holland: D. Reidel Publishing Inc., 1978.

Manegold traces the emergence of the engineer as a unique profession and entity in nineteenth century Germany.

Marcuvitz, Nathan. Waveguide Handbook. London: Peter Peregrinus Ltd., 1986.

An updated version of the 1951 original that was part of the 28-volume McGraw-Hill postwar series of Rad Lab books, Waveguide covers the mathematical basics of waveguides. It explains the computational usefulness of equivalent circuits: the translation of complicated electromagnetic field conditions into circuit diagrams. Massachusetts Institute of Technology. Five Years at the Radiation Laboratory: Presented to the members of the Radiation Laboratory by the Massachusetts Institute of Technology, Cambridge, 1946. Andover, MA: The Andover Press, Ltd., 1947.

As M.I.T.’s self-congratulatory anthology of Rad Lab activities during the war, this collection keeps track of who worked in each department at what time and reveals how many at M.I.T. viewed the university and the Rad Lab’s accomplishment after the end of World War II.

Mayr, Otto. “The Science-Technology Relationship as a Historiographic Problem.” Technology and Culture 17, no. 4 (October 1976): 663-673.

Mayr details the traditional questions and models put forward by experts of all stripes who deal with the liaisons between science and technology. He offers examples which delineate the different kinds of knowledge between technologists and scientists.

Meadows, A. J. and M. M. Hancock-Beaulieu, eds. Front Page Physics: A century of Physics in the News. Philadelphia: Institute of Physics Publishing, 1994.

A conglomeration of science stories that received prominent news attention, Front Page Physics provides some useful summaries about the reception of radar and the atomic bomb in the postwar period.

Melville, Herman. The Confidence-Man. New York: Penguin Books, 1990.

Home to terrific allegories.

Merton, Robert. The Sociology of Science. Chicago: The University of Chicago Press, 1973.

A collection of Merton’s essays on science’s role within the fabric of society. The Sociology of Science includes a chapter on the “Interactions of Science and Military Technique,” which analyzes the interests of scientists in developing military technology.

Miller, Jon D., Rafael Pardo, and Fujio Niwa. Public Perceptions of Science and Technology. Bilbao: Fundación BBV, 1997.

A study of public perceptions in the United States, European Union, Canada and Japan, Public Perceptions also poses fundamental methodological questions about the ideal way to conduct surveys of public opinion of science.

Moyer, Albert E. American Physics in Transition: A History of Conceptual Change in the Late Nineteenth Century. Los Angeles: Tomash Publishers, 1983.

Moyer recounts the theoretical outlook of leading American physicists from 1875-1900, and their reactions to the increasing number of issues found with the ether and traditional mechanical notions. The book provides a limited but useful prelude to the Rad Lab era, focusing on physicists one or two generations younger than the physicists considered here.

Multhauf, Robert P. “The Scientist and the ‘Improver’ of Technology.” Technology and Culture 1, no. 1 (Winter 1959): 38-47. Multhauf’s article contains useful presentations of some of the status problems stalking scientists and engineers since the 19th century.

Piel, Gerard. The Age of Science: What Scientists Learned in the 20th Century. New York: Basic Books, 2001.

The editor of Scientific American reviews the principal discoveries of 20th century science, but says little about the wartime work of scientists.

Ovid. Metamorphoses. Translated by Rolfe Humphries. Bloomington, IN: Indiana University Press, 1955.

Many stories of bodies changed, though none about engineering specifically

Pope, Alexander. The Works of Alexander Pope. Edited by Andrew Crozier. Hertfordshire: Wordsworth Editions Ltd., 1995.

The collected works of this witty British poet, who wrote, in “The First Epistle of the Second Book of Horace,” lines 300-301: “Who pants for glory finds but short repose/ A breath revives him, or a breath o’erthrows.”

Pursell, Carroll W., Jr. “Science and Government Agencies.” In Science and Society in the United States, edited David Van Tassel and Michael Hall, 223-249. Homewood, IL: The Dorsey Press, 1966.

Pursell traces the history of government support to science. Especially helpful is the paper’s background on 19th century relations between government and science, including the foundations and nurturing of such institutions as the National Academy of Sciences.

Reingold, Nathan. “Physics and Engineering in the United States, 1945-1965, A Study of Pride and Prejudice.” In The Michelson Era in American Science 1870-1930, edited by Rita G. Lerner, 288-298. New York: American Institute of Physics: 1988.

––––––––. “Reflections on 200 Years of Science in the United States.” In The Sciences in the American Context: New Perspectives, edited by Nathan Reingold, 9-20. Washington, D.C.: Smithsonian Institution Press, 1979.

“Reflections,” American bicentennial review covers the history of science in terms of its influence and its practitioners throughout American history. It provides data about and commentary on the status of American science in the 19th and 20th centuries. “Physics and Engineering” studies the relations between the two disciplines as well as their concerns in the postwar period.

Rhodes, Richard. The Making of the Atomic Bomb. New York: Simon & Schuster, 1986.

An account of the development of the fission bomb, as well as a biography of the numerous inventors who made possible the invention—from its beginnings through Hiroshima.

Rigden, John. Rabi: Scientist and Citizen. New York: Basic Books, 1987.

Chapter Ten deals explicitly with Rabi’s work at the Rad Lab. It is mostly simply a rehashing of the basic, chiefly entertaining stories of the Rad Lab with little insight into the work conducted there.

Rosen, Raphael C. “The Argonauts Assemble: The Founding of The Jasons, America’s Remarkably Independent Advisory Group.” Unpublished Manuscript, 2005.

––––––––. “The Temperature of Innovation: The History of Superconductivity at Harvard and the Problem of Theory.” Unpublished Manuscript, 2004.

“Argonauts” includes an examination of postwar physics elitism and physicists’ influence in government. “Temperature” details the problems of engineer shortages.

Rosen, Stephen Peter. Winning the Next War. Ithaca, NY: Cornell University Press, 1991.

Rosen studies at the mechanics of war time innovation and development. He takes the position that in terms of radar, the key innovations were made by scientists working under the navy’s command—well before the Rad Lab started.

Rothschild, Emma. “Continuing Education.” In Jerry Wiesner: Scientist, Statesman, Humanist; Memories and Memoirs, edited by Walter Rosenbluth, 157-164. Cambridge, MA: The M.I.T. Press, 2003.

Rothschild’s essay highlights the values and attitudes of Wiesner, and how they related to his research at the Rad Lab and afterwards.

Saad, T. A. “The Story of the M.I.T. Radiation Laboratory.” IEEE Aerospace and Electronic Systems Magazine 5 (October 1990): 46-51.

Saad relates the basics of the Rad Lab story in this journal for the Institute of Electrical and Electronics engineers.

“Saturday Evening Post,” Downtown Magazine, .

“Saturday Evening Post,” Wikipedia: The Free Encyclopedia, .

Two online references concerning the history of this popular magazine, including circulation numbers.

Schwartz, Rebecca Press. The Making of the History of the Atomic Bomb: The Smyth Report and the Historiography of the Manhattan Project. Ph.D. dissertation, Princeton University (in preparation).

This dissertation argues that the reason engineering, chemistry and other fields received little approbation for their contributions to the Manhattan project is precisely because they had played such an important role. The secrecy constraints of the Smyth Report demanded that the theoretical physics of the nuclear bomb was one of the weapon’s only safe facets, able to be divulged to the public.

Schweber, Silvan. “The Empiricist Temper Regnant: Theoretical Physics in the United States, 1920-1950” Historical Studies in the Physical Sciences 17, no. 1 (1986): 55-98

––––––––. QED: And the Men Who Made It. Princeton, NJ: Princeton University Press, 1994.

An analysis of the relationships between American theorists and experimentalists, Schweber’s “Empiricist” explores the unique circumstances that brought these two groups into an unusual proximity. His QED picks up where “Empiricist” leaves off and details the process by which the pragmatism of American theorists defined the very character of Quantum Electrodynamics after World War II. Both pieces are useful navigation aids for comprehending the philosophical and epistemological mores (or lack thereof) of the Rad Lab physicists.

Schwinger, Julian. “Biography.” In Nobel Lectures in Physics: 1963-1970. River Edge, N.J.: World Scientific, 1998.

Schwinger recalls the developments of his career leading up to the development of quantum electrodynamics, and in so doing comments on his work at the Rad Lab.

Seely, Bruce. “Research, Engineering, and Science in American Engineering Colleges: 1900- 1960.” Technology and Culture 34, no. 2 (April 1993): 344-386.

This study of engineering pedagogy contends that after World War II, American engineering schools shifted to an increasingly fundamental focus in order to win more federal contracts.

Shachtman, Tom. Terrors and Marvels: How Science and Technology Changed the Character and Outcome of World War II. New York: William Morrow, 2002.

This history of weaponry touches on the development of some less-frequently discussed World War II inventions, such as poison gas, cryptography devices, and primitive heat- seeking technologies.

Shakespeare, William. Hamlet Prince of Demark. Edited by Willard Farnham. New York: Penguin Books, 1970.

In this play, in argument with his mother, Hamlet makes reference to military engineers.

Shapin, Steven. “Pump and Circumstance: Robert Boyle’s Literary Technology.” Social Studies of Science 14, no. 4 (November 1984): 481-520.

By investigating Robert Boyle’s 17th century methods of dispersing knowledge about the famous air pump, Shapin unravels the very nature of a scientific fact and how such facts become something incontrovertible to the public. Moreover, he demonstrates the methods by which Boyle tried to build an experimentalist community

Skolimowski, H. “The Structure of Thinking in Technology.” Technology and Culture 7 (1966): 371-83.

Skolimowski explores the epistemology of engineering, discussing praxiology, a study of the engineer’s creativity from the point of view of efficiency. Slater, John C. “Quantum Physics in America Between the Wars.” Physics Today 21, no. 1 (January 1968): 43-51.

Slater’s article is mostly a history of technical developments, but includes a few general observations about the state of American physics in the 1920’s and 1930’s.

Sopka, Katherine. Quantum Physics in America: The Years through 1935. Los Angeles: Tomash Publishers/American Institute of Physics, 1988.

Sopka’s account of developments largely in 1920’s American physics focuses upon the growing contributions of American physicists to international science. The book also underscores the rise of theoretical physics in the United States and its connection to the advent of new problems spawned by Quantum Mechanics.

Stokes, Donald E. Pasteur’s Quadrant: Basic Science and Technological Innovation. Washington, D.C.: Brookings Institution Press, 1997.

Stokes dispenses with the traditional linear perspective of scientific discovery in which basic science lies at the opposite end of the spectrum from applied technology, and instead proposes a two-dimensional model in which science is pursued both for the sake of fundamental knowledge and for the sake of harnessing that knowledge. This new realm is one of four quadrants of science divided up by the two questions: is it searching for fundamental understanding and is it considering the applications of such an understanding. From this framework, Stokes proposes a new post-Cold war paradigm in which scientists emphasize this new quadrant—Pasteur’s quadrant—for science and technology funding in America. As opposed to the dominant school of thought that esteems basic science as the root of all progress, Stokes proposes a unique model more attuned to the actual processes of scientific advancement.

Support Organizations for the Engineering Community. Washington D.C.: National Academy Press, 1985.

Written to help the engineering community achieve a more respectable position in the public eye, this book is one of the few that looks at the world and the media from the perspective of the engineering community’s needs.

Swords, S. S. Technical History of the Beginnings of Radar. London: Peter Peregrinus Ltd., 1986.

Though Swords’s book does not consider radar after the invention of the cavity magnetron, it is still a resource of some technical depth as well as historical breadth— covering the United States, Japan, Germany, France, and especially Britain. de Tocqueville, Alexis. Democracy in America. Translated by Harvey C. Mansfield and Delba Winthrop. Chicago: University of Chicago Press, 2000.

An enduring classic in history and political science, de Tocqueville’s brief but trenchant insights into early American science and its democratic incubator remain a valuable perspective on the idiosyncrasies of scientific pursuits, both theoretical and pragmatic in the United States.

United States, Joint Board on Information Policy. Radar: A Report on Science at War. Washington, D.C: Office of War Information, 1945.

Released to the press on Tuesday, August 14, 1945, just after the war’s resolution, this report is the first public document on radar. It is a useful tool for understanding the radar’s immediate postwar legacy. van der Waerden, B. L. “Eupalinos and his Tunnel.” In Philosophers and Machines, edited by Otto Mayr, 48-49. New York: Science History Publications, 1976.

In this article, van der Waerden discusses how he believes that the famous Greek tunnel at Samos was the work of Pythagorean theory.

Van Riper, A. Bowdoin. Science in the Popular Culture: A Reference Guide. Westport, CT: Greenwood Press, 2002.

Treating technologies, scientists, and natural disasters in movies and books, Van Riper teases out the chief negative and positive popular stereotypes about modern scientists. van Vleck, John H. “American Physics Comes of Age.” Physics Today 17, no. 6 (June 1964): 21- 26.

The first American to earn a Ph.D. for purely theoretical work on quantum mechanics, van Vleck reviews, at a very superficial level, the transformation in American physics during quantum mechanics’ development.

Vincenti, Walter G. “Control-Volume Analysis: A Difference in Thinking between Engineering and Physics.” Technology and Culture 23, no. 2 (April 1982): 145-174.

––––––––. What Engineers Know and How They Know It: Aeronautical Studies from Aeronautical History. Baltimore, MD: The Johns Hopkins University Press, 1990.

By comparing the methodological approaches to thermodynamics and fluid mechanics of engineers and physicists, Vincenti, an aeronautical engineer himself, argues in “Control- Volume” that the physical and economic concerns of engineering can lead to very different technical considerations than those of physics. What Engineers Know takes five case studies from the history of aeronautical engineering in an effort to formulate an epistemology of engineering.

Waller, John. Leaps in the Dark. Oxford: Oxford University Press, 2004.

Waller features a chapter on Robert Watson-Watt and the historiography of how he and circumstance made him into the “father of radar.”

Wang, Jessica. American Science in an Age of Anxiety. Chapel Hill, NC: University of North Carolina Press, 1999.

Wang studies the suspicions of anticommunism that penetrated American science, especially the physics community, in the late 1940’s and early Cold War.

Weart, Spencer. “The Physics Business in America, 1919-1940: A Statistical Reconnaissance.” In The Sciences in the American Context: New Perspectives, edited by Nathan Reingold, 295-358. Washington, D.C.: Smithsonian Institution Press, 1979.

––––––––. “The Rise of ‘Prostituted’ Physics.” Nature 262, (1976): 13-17.

In “Business,” Weart considers, above all else, the growth of American physics between the two World Wars. His paper is especially useful as a rich statistical source on the geography, salaries, institutions, societies, and papers of American physicists of the 1920’s and 1930’s. “The Rise” details the relationship between academic science and industry from the late 19th century until the second world war.

Weiner, Charles. “A New Site for the Seminar: The Refugees and American Physics in the Thirties.” In The Intellectual Migration, edited by Donald Fleming and Bernard Bailyn, 190-234. Cambridge, MA: Harvard University Press, 1968.

––––––––. “Physics in the .” Physics Today 23, no. 10 (October 1970): 31-38.

––––––––. “Science and Higher Education.” In Science and Society in the United States, edited by David Van Tassel and Michael Hall, 163-189. Homewood, IL: The Dorsey Press, 1966.

“A New Site” examines the dispersion of continental European physicists in the 1930’s as well as the self-improvement movement of American physics in that epoch. The title of “Physics in the Great Depression,” encapsulates the entire content: in an effort similar to that present in “A New Site,” Weiner writes a history of the changes in American physics in this era. “Science and Higher Education” contains especially valuable information on the history of engineering in America.

Weingart, Peter. “The Relation Between Science and Technology—A Sociological Explanation.” In The Dynamics of Science and Technology, edited by Krohn, Wolfgang, Edwin T. Layton Jr. and Peter Weingart, 251-286. Dordrecht, Holland: D. Reidel Publishing Inc., 1978.

Weingart applies sociology to the study of the relationship between science and technology, breaking with previous explanations of this relationship.

White, March W. and William Crew. “Physicists In and Following World War II.” American Journal of Physics 18 (1950): 487-495.

White and Crew have compiled a statistical analysis of the wartime work and postwar dispersion of physicists in America; it is minimally useful for comparing physics and engineering.

Wildes, Karl L. and Nilo A Lindgren. A Century of Electrical Engineering and Computer Science at MIT, 1882-1982. Cambridge MA: The M.I.T. Press, 1985.

This history contains valuable chapters on the war years, that pay special attention to engineering contributions.

Zachary, G. Pascal. Endless Frontier: Vannevar Bush, Engineer of the American Century. New York: The Free Press, 1997.

This biography of Bush contains a history of the Radiation Laboratory concerned principally with martial results and the cooperation between armed forces leaders and scientists.