Skip to main content
Log in

Writing the Biography of Hans Bethe: Contextual History and Paul Forman1

  • Published:
Physics in Perspective Aims and scope Submit manuscript

Abstract

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

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2

Similar content being viewed by others

Notes

  1. The sociologist Robert K. Merton observed that if two scientists arrive at similar conclusions, the more eminent or famous scientist will often receive more credit. Merton coined the term "Matthew Effect" to describe the phenomenon, from Jesus' parable of the talents as told in Matthew 25:29: "For unto every one that hath shall be given, and he shall have abundance: but from him that hath not shall be taken even that which he hath."

  2. Lebensphilosophie (“philosophy of life”) emerged in Germany during the nineteenth century as a reaction to the emphasis on science and rationalism in philosophy. It emphasized the meaning and purpose of life and had a anti-rational, Romantic component in its outlook.

References

  1. This paper is based on an invited lecture delivered in November 2011 at the annual meeting of the History of Science Society and in December 2011 on the occasion of Paul Forman’s retirement at the Smithsonian Institution. It was a privilege to have been asked to deliver the lecture honoring Forman. We have been close friends for the past thirty years. We both come out of physics and have a special relationship to physics and physicists. I admire Paul Forman and his works greatly, and he has influenced me deeply. His integrity, his comportment, and his writings made, and continue to make, clear to me the responsibilities we have as historians. This paper is dedicated to him as a token of my admiration, affection and respect. When considering what Paul Forman has accomplished as a historian of science and as a curator, and keeps on accomplishing as a historian, many commendations can be made. John Heilbron, who has known Forman since his student’s days at Berkeley did so when commenting on Forman’s oeuvre as a historian at a conference in Vancouver in 2005 honoring Forman: John L. Heilbron, “Cold War Culture, History of Science and Postmodernity: Engagement of an Intellectual in a Hostile Academic Environment.” in Cathryn Carson, Alexei Kojevnikov, and Helmuth Trischler, eds., Weimar Culture and Quantum Mechanics (Singapore: World Scientific, 2011), 2–20. The editors’ introduction to the proceedings of the conference and Heilbron’s article therein detail the magnitude of Forman’s accomplishments as a historian of science and the respect he is held in as an outstanding scholar.

  2. He had joined the department in the spring semester 1967 and completed his PhD dissertation that summer in the history department of the University of California at Berkeley; Hunter Dupree had been his thesis adviser. Paul Forman, “Weimar Culture, Causality, and Quantum Theory, 1918–1927: Adaptation by German Physicists and Mathematicians to a Hostile Intellectual Environment,” Historical Studies in the Physical Sciences 3 (1971), 1–115; “The Reception of an Acausal Quantum Mechanics in Germany and Britain,” in Seymor Mauskopf, ed., The Reception of Unconventional Science, Seymor Mauskopf, ed., [AAAS] Selected Symposium 25 ([Boulder Colo]: Westview Press, 1979), 11–50; “Kausalität, Anschaulichkeit, and Individualität; or how Cultural Values Prescribed the Character and the Lessons Ascribed to Quantum Mechanics,” in Nico Stehr and Volker Meja, eds., Society and Knowledge: Contemporary Perspectives in the Sociology of Knowledge & Science, ([New Brunswick NJ]: Transaction Books, 1984), 333–48; reprinted, 2nd revised edition (Transaction Books: New Brunswick NJ, 2005), 357–371.

  3. For the argument that Darwin had been influenced by the intellectual and political views and cultural values of powerful circles in England and Scotland, see Robert M. Young, “Malthus and the Evolutionists: The Common Context of Biological and Social Theory,” Past and Present 43 (1969), 109–45; “The Historiographic and Ideological Contexts of the Nineteenth-Century Debate on Man’s Place in Nature,” in M. Teich and R.M. Young, eds., Changing Perspectives in the History of Science (London: Heinemann, 1973), 344–438; Darwins Metaphor: Natures Place in Victorian Culture (Cambridge: Cambridge University Press, 1985). See also A. La Vergata, “Images of Darwin: A Historiographic Overview,” in D. Kohn, ed., The Darwinian Heritage (Princeton, NJ: Princeton University Press, 1985), 901–929, and Ingemar Bohlin, “Robert M. Young and Darwin Historiography,” Social Studies of Science 21 (1991), 597–648.

  4. Forman, “Behind Quantum Electronics: National Security as Basis for Physical Research in the United States, 1940–1960,” Historical Studies in the Physical Sciences, 18 (1987), 149–229.

  5. Heilbron, “Cold War Culture” (ref. 1), 17.

  6. Forman, “Recent Science: Late Modern and Post-Modern,” in Philip Mirowski and Esther-Mirjam Sent, eds., Science Bought and Sold: Rethinking the Economics of Science (Chicago: University of Chicago Press, 2002), 109–48; “The Primacy of Science in Modernity, of Technology in Postmodernity and of Ideology in the History of Technology,” History & Technology 23 (2007), 1–152; “(Re)cognizing Postmodernity: Helps for Historians—of Science Especially,” Berichte zur Wissenschaftsgeschichte 3 (2010), 157–75; and especially “On the Historical Forms of Knowledge Productions and Curation: Modernity Entailed Disciplinarity, Postmodernity Entails Antidisciplinarity,” Osiris 27 (2012), 56–100.

  7. This is abstracted from Mirowski’s “Postface” in Philip Mirowski and Dieter Plehwe, eds., The Road from Mont Pèlerin: The Making of the Neoliberal Thought Collective (Cambridge, MA: Harvard University Press, 2009). Mirowski also states that “Neoliberals see pronounced inequality of economic resources and political rights not as an unfortunate by-product of capitalism, but as a necessary functional characteristic of their ideal market system. Inequality is not only the natural state of market economies, but it is actually one of its strongest motor forces for progress. Hence the rich are not parasites, but (conveniently) a boon to humankind.” (438) See also Philip Mirowski, Science-mart: Privatizing American Science. (Cambridge, MA: Harvard University Press, 2011).

  8. Silvan S. Schweber, Nuclear Forces: The Making of the Physicist Hans Bethe (Cambridge, MA: Harvard University Press, 2012).

  9. I am aware that at most I can write succinct, contextually sensitive, narrations selected from some of the important components of his life from 1940 to his death: Los Alamos; the Shelter Island conference; his consulting for the General Electric Knolls Laboratory, Detroit Edison, and later AVCO; his involvement with H-bombs; his sabbatical in Cambridge/England during the academic year 1955–6; his shift from high energy to nuclear physics; nuclear matter; his serving on PSAC; the Nobel prize in 1967; his involvement in the Cornell student rebellion 1968–71; his becoming an astrophysicist; neutrinos and supernovae; star wars.

  10. Forman, “‘Swords into Ploughshares’: Breaking new Ground with Radar Hardware and Technique in Physical Research after World War II,” Reviews of Modern Physics 67 (1995), 397–455.

  11. Hans A. Bethe, “My Life in Astrophysics,” Annual Review of Astronomy and Astrophysics 41 (2003), 1–14.

  12. For an early example of this see Bethe, “Über die nichtstationäre Behandlung des Photoeffekts,” Annalen der Physik 4 (1930), 443–449.

  13. Bethe’s first exposure to the practice of science was in his father’s laboratory. There he became aware of the amazing diversity of animal life and learned that individual behavior can never be considered by itself but must always be seen through interactions with the environment and all the entities that make up that environment. He also first experienced science as a social activity in which his father, his father’s Assistenten, Doktoranten, and laboratory assistants were in constant interaction in a shared physical and intellectual environment. The view of science Bethe obtained in his father’s laboratory was as a practice in which knowledge is created by experiments using instruments that measure with limited accuracy, produce data that have to be analyzed statistically and interpreted with mathematical models that idealize the context in which the interactions take place. The aim of the knowledge produced in his father’s laboratory was how to understand the complexity and diversity of the biological world. There was no attempt to find an ultimate theory that would explain all biological phenomena. The mature Bethe was always skeptical of the possibility of finding final theories in physics.

  14. Szasz and Siegel (who both had studied in Göttingen) reflected Hilbert’s “modernist” views of making mathematics an autonomous discipline as well as his idealistic views of mathematics. See Jeremy Gray, “Modernism in Mathematics,” in Eleanor Robson and Jacqueline Stedall, eds., The Oxford Handbook of the History of Mathematics, (Oxford: Oxford University Press, 2009), 663–683.

  15. See Forman and Armin Hermann’s entry on Sommerfeld in the Dictionary of Scientific Biography, Charles C. Gillispie, ed. (New York: Scribner’s, 1975), 12:525–532.

  16. Timothy Lenoir, “Practical Reason and the Construction of Knowledge,” in Ernan McMullin, ed., The Social Dimensions of Science (Note Dame, IN: University of Notre Dame 1992), 158–197.

  17. See Hans-Jürgen Borchers, “Einstein’s Principle of Maximal Speed in Classical and Quantum Physics,” in Rathindra Nath Sen and Alexander Gersten, eds., Mathematical Physics Towards the 21st Century (Beer-Sheva: Ben Gurion University of the Negev Press, 1994).

  18. See in particular Karl von Meyenn, “Pauli’s Belief in Exact Symmetries,” in Manuel Garcia Doncel, Armin Hermann, Louis Michel, and Abraham Pais, eds., Symmetries in Physics (16001980) (Bellaterra [Barcelona]: Semineri d’Historia de les Ciènces, Universitat Autònoma de Barcelona, 1987), 329–360. Von Meyenn quotes a letter of Pauli to Schrödinger written on January 27, 1955: “When I consider the matter where a theory is in need of improvement, I never start from considerations about measurability but from such conclusions of the theory where the mathematics is not correct.” Von Meyenn goes on: “Behind these words is [Pauli’s] deep conviction that the mathematical structure of physical theories possesses a greater content of reality than the common intuition and direct experience.” (332)

  19. Bethe, “Quantenmechanik der Ein und Zwei-Elektronenprobleme,” in Hans Geiger and Karl Scheel, Handbuch der Physik, XXIV, Part I, Adolf Smekal, ed., Quantentheorie (Berlin: Julius Springer Verlag, 1933), 273–560; Bethe and Sommerfeld, “Elektronentheorie der Metalle,” in Geiger and Scheel, Handbuch der Physik XXIV, Part II (Berlin: Julius Springer Verlag, 1933), 333–622.

  20. Bethe, Robert F. Bacher, and Milton S. Livingston, Basic Bethe: Seminal Articles on Nuclear Physics (New York: American Institute of Physics and Tomash Publishers, 1986).

  21. Hans A. Bethe, “Theoretical Division: The Beginning,” in Theory in Action: Highlights in the Theoretical Division at Los Alamos 19432003. Volume I. Compiled by Francis H. Harlow and H. Jody Shepard. LA–14000–H. History Report. Unclassified. (Los Alamos National Laboratory: Theoretical Division, 2004), 1–5. For a more detailed account of T-Division’s involvement with computers see Bethe, “Introduction,” in Sidney Fernbach and Abraham H. Taub, eds., Computers and their Role in the Physical Sciences (New York: Gordon and Breach Science, 1970), 1–10. For the continuation of that story see William Aspray, John von Neumann and the Origins of Modern Computing (Cambridge, MA: MIT Press, 1990). It is also interesting to note that Bethe never programmed a computer to solve the complex problems he was considering.

  22. However, the Cold War context was such that Robert R. Wilson (who had severed all his ties with making atomic weapons after he left Los Alamos, where he had been the head of the Nuclear Physics Division), in 1950, when he was the director of the Newman Lab, designed a mobile electron beam gun to destroy atomic bombs after he had learned of strong focusing. See Silvan S. Schweber, “Defending against Nuclear Weapons: A 1950 Proposal,” Physics Today 60, no. 4 (2007), 36–41.

  23. See Forman, “On the Historical Forms of Knowledge Production” (ref. 6).

  24. See Paul Hartman, The Cornell Physics Department: Recollections and a History of Sorts (Ithaca, NY: n. p., 1984).

  25. Norris Bradbury, then director of Los Alamos, in 1950 asserted this when interviewed by the FBI in connection with the renewal of Bethe’s Q clearance. I am indebted for this information to Alex Wellerstein, who has studied Bethe’s FBI record.

  26. See for example Silvan S. Schweber, In the Shadow of the Bomb: Bethe, Oppenheimer, and the Moral Responsibility of the Scientist. (Princeton, NJ: Princeton University Press, 2000) for the details of the plans.

  27. Michael Gordin, Red Cloud at Dawn: Truman, Stalin, and the End of the Atomic Monopoly. (New York: Farrar, Straus, 2009).

  28. Bethe, A Theory for the Ablation of Glassy Materials, Issue 38 of Research report. Avco Manufacturing Corporation, Avco Everett Research Laboratory, 1958–30 pages; H.A. Bethe and M.D. Adams, “A Theory of the Ablation of Glassy Materials,” International Journal of Aeronautical and Space Sciences 26 (1959), 321–328.

  29. See Priscilla J. McMillan, The Ruin of J. Robert Oppenheimer, and the Birth of the Modern Arms Race (New York: Viking, 2005).

  30. See Schweber, In the Shadow of the Bomb (ref. 26).

  31. Zuoyue Wang, In Sputnik’s Shadow: The President’s Science Advisory Committee and Cold War America (New Brunswick, NJ: Rutgers University Press, 2008).

  32. Bethe, “My Life in Astrophysics,” in Gerald E. Brown and Chang-Hwan Lee, eds., Hans Bethe and His Physics (Singapore: World Scientific 2006), 27–44.

  33. The neutrino problem was concerned with the fact that there were far fewer neutrinos being emitted by the sun than solar models had predicted. See the articles on neutrinos in Brown and Lee, Hans Bethe and His Physics (ref. 32).

  34. See Jeremy Bernstein, Hans Bethe, Prophet of Energy (New York: Basic Books, 1980); Boris Ioffe, “Hans Bethe and the Global Energy Problems,” in Brown and Lee, Hans Bethe and His Physics (ref. 32), 263–272.

  35. Karin Knorr-Cetina, Epistemic Cultures: The Sciences Make Knowledge (Cambridge, MA: Harvard University Press, 1999), 1.

  36. See, e. g., David Edgerton, “Science in the United Kingdom: A Study in the Nationalization of Science,” in John Krige and Dominique Pestre, eds., Science in the Twentieth Century (Amsterdam: Harwood Academic Publishers, 1997), 759–776.

  37. See, e. g., Enrico Fermi, “Quantum Theory of Radiation,” Reviews of Modern Physics 4 (1932), 87–132.

  38. Laurie M. Brown and Helmut Rechenberg, The Origin of the Concept of Nuclear Forces (Philadelphia: Institute of Physics Pub., 1996).

  39. As a result of his researches and calculations in atomic and solid state physics, Bethe early on had recognized implicitly what Dirac would state explicitly in the first edition of his The Principles of Quantum Mechanics, namely that our representation of the physical world can be hierarchically ordered by virtue of Planck’s constant, h. Macroscopic systems, whose characteristic time T, mass M, and length L, are such that ML 2 /T ≫ h are described by classical mechanics; those for which ML 2 /T ≈ h are described by quantum mechanics. See the remarkable text on quantum mechanics, Eyvind H. Wichmann, Quantum Physics (New York: McGraw-Hill, 1971).

  40. These hierarchies are not independent: accurate measurements of atomic energy levels will reveal nuclear and subnuclear properties. Similarly, the recent startling discovery of the presence of cold dark matter—consisting of as yet undiscovered subnuclear entities—to make sense of new cosmological observational data is proof of the linkage between the various levels. But it must also be noted that these observations have not destabilized our amazingly accurate representations of the atomic world. Needless to say, the linkage of the levels is made explicit as soon as one tries to answer evolutionary questions.

  41. And more recently in terms of the standard model.

  42. See Brown and Rechenberg, The Origin of the Concept of Nuclear Forces (ref. 38).

  43. See in this connection Michelangelo De Maria, Mario Grilli, Fabio Sebastiani, eds.,The Restructuring of the Physical Sciences in Europe and the United States,194560: Proceedings of the International Conference Held in Rome, Università “La Sapienza”, 19–23 September 1988” (Singapore: World Scientific,1989); and therein, Forman, “Social Niche and Self-Image of the American Physicist” pp. 96–104.

  44. Forman, “‘Swords into Ploughshares’: Breaking New Ground with Radar Hardware and Technique in Physical Research after World War II,” Reviews of Modern Physics 67 (1995), 397–455. See also Forman, “Into Quantum Electronics: the Maser as ‘Gadget’ of Cold-War America,” in Paul Forman and José Sánchez-Ron, eds., National Military Establishments and the Advancement of Science and Technology: Studies in Twentieth Century History (Dordrecht: Kluwer Academic,.1996), 261–326.

  45. Louis N. Ridenour, Radar System Engineering (New York: McGraw-Hill, 1947).

  46. See Forman, “‘Atom Smashers: Fifty Years’—Preview of an Exhibit on the History of High Energy Accelerators,” IEEE Transactions on Nuclear Science NS–24 (1977): 1896–99.

  47. J.S. Hey, The Evolution of Radio Astronomy (New York: Science History Publications, 1973).

  48. The conference began on June 3 and ended on the 6th. See Schweber, QED and the Men who Made It (Princeton, NJ: Princeton University Press, 1994); Willis E. Lamb, and Robert C. Retherford, “Experiment to Determine the Fine Structure of the Hydrogen Atom” (Columbia University Radiation Laboratory Report, 1946), 18–26; Lamb and Retherford, “Fine Structure of the Hydrogen Atom by Microwave Method,” Physical Review 72 (1947), 241–243; John E. Nafe, Edward B. Nelson, and Isidor I. Rabi, “Hyperfine Structure of Atomic Hydrogen and Deuterium,” Physical Review 71 (1947), 914–15.

  49. I owe the notion of a “crucial calculation” to Howard Schnitzer. See Schweber, QED (ref. 48), where the idea is applied to Bethe’s calculation of the Lamb shift in hydrogen and Schwinger’s calculation of the anomalous magnetic moment of the electron using renormalization concepts. Other examples come readily to mind: Einstein’s calculation of the advance of the perihelion of Mercury using his formulation of general relativity; Pauli’s calculation of the spectrum of hydrogen using Born, Jordan, and Heisenberg’s matrix mechanics; and many other instances in modern particle and condensed matter physics.

  50. Steven Weinberg, “The Search for Unity: Notes for a History of Quantum Field Theory,” Daedalus 106, no. 4 (1977), 17–35.

  51. See Schweber, “Shelter Island Revisited,” History of Physics Newsletter 11, no. 33 (2011), 1, 8, 10–13.

  52. Max Dresden, H.A. Kramers: Between Tradition and Revolution (New York: Springer-Verlag, 1987).

  53. Bethe’s Shelter Island notes were found in 2011 in his mother’s trunk that had been stored in the basement of Bethe’s house on White Park Road in Cayuga Heights, NY. The notes can now be found in Bethe’s papers in the Rare Manuscript Division of the Cornell Library.

  54. Since libraries only maintain copies of bound books, copies of lecture notes are only to be found in the library of the institution where they were delivered or in the Nachlass of the lecturer. Dyson’s lectures were recently reissued: Freeman J. Dyson, Advanced Quantum Mechanics. Translated and transcribed by David Derbes. 2nd ed. (Hackensack, NJ: World Scientific, 2011). A first edition had been issued in 2007.

  55. Thus, the proceedings of the Pocono and Oldstone conferences, the follow-ups of the June 1947 Shelter Island conference, became available as dittoed notes within two months of when they were held in the spring of 1948 and 1949. They disseminated the lectures that Schwinger, Feynman, and Dyson had presented at them.

  56. In the United States after the war, the GI Bill allowed large numbers of young men who had served in the Armed Forces to be trained as physicists and transformed the demography of physics: there were many more physicists and they were younger.

  57. At that same time, NATO began supporting summer schools on various subjects at different places in Europe.

  58. For the diffusion of Feynman’s approach to quantum electrodynamics and of his diagrams, see David Kaiser, Drawing Theories Apart: The Dispersion of Feynman Diagrams in Postwar Physics (Chicago: University of Chicago Press, 2005). The role of lecture notes and summer school notes supplement the views expressed therein regarding textbooks in graduate education. See also Kaiser, ed., Pedagogy and the Practice of Science: Historical and Contemporary Perspectives (Cambridge, MA: MIT Press, 2005).

  59. The history of advances in theoretical physics during the twentieth century has often been written so that developments in “fundamental” theories are seen as being the most important. They thus have received a disproportionate emphasis. In addition, the Cold War—at least in the West—has had as one of its consequences that the history of physics during the second half of the century has often filtered out the Soviet and Russian contributions. The history of the solution of the phase transition problem in the late 1960s makes it clear that: a) while quantum field theory was in decline and viewed as being marginal among high energy physicists from the late 1950s to the mid-1960s, the use of field-theoretic methods was a thriving enterprise in solid state and condensed matter physics and the source of deep insights that would later be transferred to relativistic quantum field theory; and b) besides Lev Landau, Soviet theoretical physicists had made important, foundational contributions to condensed matter physics and to the unraveling of the phase transition problem.

  60. A vision now tainted for having tasted sin in building atomic weapons. That vision was spelled out in his 1953 Reith lectures, in which he quoted Bishop Sprat’s 1667 history of the Royal Society. I can readily transcribe Sprat’s statement so that it becomes Oppenheimer’s manifesto for the IAS, and for physics at the IAS: “It is to be noted that [the members of the IAS] are to freely admit Men of different religions, Countries, and Professions of Life. This they are obliged to do, or else they would come far short of the Largeness of their own Declarations. For they openly profess, not to lay the foundations of an American, British, German or Japanese mathematics or science; but a mathematics and science of Mankind.”

  61. Incidentally, in recognition of all that Bethe had accomplished in applying quantum mechanics and quantum electrodynamics to explain atomic and nuclear phenomena after the war, in the early 1950s Oppenheimer invited Bethe to join the Institute as a professor of physics. Bethe declined, feeling that his place was at Cornell. The IAS had become a finishing school for the brightest young theorists and Bethe felt that with all his other commitments he could no longer chart new directions in research for these young people to explore.

  62. Les Houches has become a year-round school, whose lectures and publications continue to influence the development of physics profoundly.

  63. Quantum mechanics had acquired a new robustness during World War II. It had explained quantitatively the properties of germanium used in radar receivers, the properties of matter at 50 million K, and could predict with fair accuracy the results of critical nuclear reactions. The teaching of quantum mechanics thus gained new importance as a result of the wartime advances. No one was able to highlight and demonstrate the new powers of quantum mechanics better than the twenty-eight-year-old Julian Schwinger, who had become a professor of physics at Harvard in the fall of 1946. His 1947 and 1948 courses on nuclear physics and quantum mechanics became legendary. Attended not only by the graduate students at Harvard, but by a large fraction of the physics community in the greater Boston area, two sets of notes of these lectures were written one by John Blatt, the other by Morton Hamermesh, and both were widely disseminated and reproduced elsewhere. They became the basis of quantum mechanics courses all over the United States. These notes are difficult to find since they were not printed, nor bound, and thus most libraries did not preserve them.

  64. See John Krige and Dominique Pestre in Peter Galison and Bruce Hevly, eds., Big Science: The Growth of Large-Scale Research (Stanford: Stanford University Press, 1992); Armin Hermann et al., eds., History of CERN (Amsterdam: North-Holland Physics Publications, 1987–1996).

  65. Forman, “On the Historical Forms of Knowledge Production” (ref. 6), 63. The valuation of disciplinarity in modernity and interdisciplinarity in postmodernity is a central concern of that article. See also the discussion of the historical development of disciplinarity and interdiciplinarity within American universities in Louis Menand, The Marketplace of Ideas (New York: W.W. Norton, 2010).

  66. See David E. Rowe, “Klein, Hilbert and the Göttingen Mathematical Tradition,” Osiris 5 (1989), 186–213; “Making Mathematics in an Oral Culture: Göttingen in the Era of Klein and Hilbert,” Science in Context 17 (2004), 85–129. Besides all the researches on infinite dimensional vector spaces carried out in Göttingen (stemming from the concerns of Hilbert and Minkowski with Boltzmann’s gas theory, lattice vibrations, and black-body radiation), after Einstein became involved with his theory of general relativity, Hilbert and Klein became deeply entangled in these activities and generated important mathematical advances, e.g. the researches of Emmy Noether. Some of the researches of Veblen and of Élie Cartan could similarly be characterized as mathematical physics.

  67. In fact, until World War II the graduate courses in classical mechanics were often offered by departments of mathematics.

  68. Think of Poincaré, Borel, Kolmogorov, Sinai, …

  69. It would be interesting to compare in detail the factors that operated in the various national settings (US, France, Soviet Union, Germany, Switzerland,…) that made possible the emergence the discipline of mathematical physics; to compare the different kinds of problems addressed in the various settings and what these reflected; to compare the status of the discipline in the differing settings; to see whether the practitioners became members of physics or mathematics departments. In addition, one can ask what made it possible for Wightman, Jost, Haag, Kastler to create their schools and for the students they trained to form a new discipline with all the accoutrements that go with it, such as professional journals and prizes. Surely in the United States the restructuring of the universities into research and teaching universities after World War II was an important factor. There was a greater emphasis on research, with lavish government support as part of its pursuance of the Cold War, and the accompanying overhead payments allowing universities to support activities and functions not directly supported by the government, such as scholarship in the arts and the humanities. Undoubtedly, during the Cold War era, national prestige and similar factors were at play in the Soviet Union and elsewhere. See, e.g., Clark Kerr, The Great Transformation in Higher Education, 1960–1980 : The Uses of the University (Albany, NY: State University of New York Press, 1991); Stuart W. Leslie, The Cold War and American Science: The Military-Industrial-Academic Complex at MIT and Stanford (New York: Columbia University Press, 1993).

  70. That mathematicians held important and influential positions in both the Soviet Academy of Sciences as well as in the Soviet Atomic Energy establishment was another important determinant of the focus of the Soviet research in this area.

  71. For valuable insights into the evolution of Russian and Soviet mathematics, see the very interesting volume A. A. Bolibruch, Yu. S. Osipov, and Ya. G. Sinai, eds., Mathematical Events of the Twentieth Century (Berlin: Springer, 2000), in particular the articles by V. I. Arnold, L. D. Faddeev, and V. S Vladimirov.

  72. Bogoliubov seems to be the exception in the Soviet Union.

  73. Wightman’s ties were with von Neumann and Wigner; von Neumann had a close working association with Hilbert during his stay in Gottingen from 1927 until 1929 as a Privatdozent there. See Steve J. Heims, John von Neumann and Norbert Wiener: From Mathematics to the Technologies of Life and Death (Cambridge, MA: MIT Press, 1980); William Aspray, John von Neumann and the Origins of Modern Computing (Cambridge, MA: MIT Press, 1990). Wigner and von Neumann were close friends since their teens, becoming close colleagues in Princeton in the early 1930s. Recall that the Institute for Advanced Study (IAS) was located on the campus of Princeton University from 1933, the date of its opening, until 1939, when Fuld Hall was completed. The IAS mathematicians had offices in Fine Hall, where the Princeton mathematics department was located. Fine Hall is the building adjacent to the Palmer Laboratory, the home of the Physics department, with open corridors to it. Jost was a student of Wentzel and Pauli and was Pauli’s successor as professor of theoretical Physics at the ETH. Both Pauli and Wentzel were students of Sommerfeld. Haag was a student of Fritz Bopp, who was a student of Heisenberg, who in turn was a student of Sommerfeld. All three at some stage gave axiomatic formulations of what they thought were the foundations of quantum field theory. Van Hove was closely associated with Weyl, von Neumann, Wigner, Bargmann, and Wightman during his stay at the IAS from 1949 to 1954. Hendrik B. G. Casimir, “Léon Charles Prudent van Hove (10 February 1924–2 September 1991),” Proceedings of the American Philosophical Society, 136, no. 4 (1992), 602–606.

  74. From the early 1930s on, Wigner was a member of both the physics and the mathematics departments at Princeton.

  75. See Arthur Wightman, “The Theory of Quantized Fields in the 50s,” in Laurie Brown, Max Dresden, and Lillian Hoddeson, eds., Pions to Quarks: Particle Physics in the 50 s (Cambridge: Cambridge University Press 1989); “The Usefulness of a General Theory of Quantized Fields,” in Yian Yu Cao, ed., Conceptual Foundations of Quantum Field Theory (Cambridge: Cambridge University Press, 1999) for a history of the developments in axiomatic and constructive field theory. See also Rudolf Haag, “Local Algebras: A Look Back at the Early Years and at Some Achievements and Missed Opportunities,” The European Physical Journal H 35 (2010), 255–261.

  76. Raymond F. Streater and Arthur S. Wightman, PCT, Spin and Statistics, and All That (New York: W. A. Benjamin, 1964); Res Jost, The General Theory of Quantized Fields ([Providence, R.I.], American Mathematical Society, 1965). David Ruelle’s book, Statistical Mechanics: Rigorous Results (New York: A.W. Benjamin, 1969), which became known as “The Book,” gives a thorough overview of the ways rigorous mathematical analyses had secured some of the foundations of statistical mechanics and had established what kinds of systems could be describe.

  77. Freeman J. Dyson and A. Lenard, “The Stability of Matter,” Journal of Mathematical Physics 8 (1967), 423–433. See also Freeman J. Dyson, “The Stability of Matter,” in M. Chrétien, E.P. Gross and S. Deser, eds., Statistical Physics, Phase Transitions, and Superfluidity (New York: Gordon and Breach, 1968), 1:179–239. See also the informative and insightful article by Larry Spruch, “Pedagogic Notes on Thomas-Fermi Theory (and on Some Improvements): Atoms, Stars, and the Stability of Bulk Matter,” Reviews of Modern Physics 63 (1991), 151–209 and the references therein to the papers of Eliot Lieb and by Walter Thirring. Also Eliot Lieb, “Thomas-Fermi and Related Theories of Atoms and Molecules,” in G. Velo and A. S. Wightman, Rigorous Atomic and Molecular Physics (New York: Plenum Press 1981), 213–301; and Walter Thirring, “The Stability of Matter,” in ibid., 309–326.

  78. See, e. g., Pierre Deligne et al., eds., Quantum Fields and Strings: A Course for Mathematicians (Providence, RI: American Mathematical Society, 1999), in particular Edward Witten’s lectures on “The Dynamics of Quantum Field Theory” (2:1119–1158).

  79. See for example Subir Sachdev, “What can Gauge-Gravity Duality Teach Us about Condensed Matter Physics?” Annual Review of Condensed Matter Physics 3 (2012): 9–33; “Strange and Stringy,” Scientific American 308, no. 12 (2013), 44–51; “The Quantum Phases of Matter,” (2011) http://arxiv.org/abs/1203.4565, last accessed April 22, 2014.

  80. Battelle Seattle Rencontres (1971). Statistical Mechanics and Mathematical Problems, ed. A. Lenard (Berlin, New York: Springer-Verlag, 1973).

  81. See Forman, “Social Niche and Self-Image of the American Physicist,” in Michelangelo De Maria et al., eds., Proceedings of the International Conference “The Restructuring of the Physical Sciences in Europe and the United States, 1945–60” (World Scientific: Singapore, 1989), 96–104.

  82. Thus, until the 1960s the general examinations for the PhD at American universities included questions on all parts of physics and students were expected to have a good grounding in all of them.

  83. Robert K. Merton, “The Matthew Effect in Science,” Science 159, no. 3810 (1968), 56–63; “The Matthew Effect in Science, II: Cumulative advantage and the symbolism of intellectual property,” Isis 79 (1988), 606–623.

  84. The wartime laboratories, the IAS, together with other postwar developments, such as the GI bill, the ONR’s and the AEC’s support of research in physics at universities, and the contract system associated with this support, were responsible for the creation in the US of a new generation of outstanding young theorists: Phillip Anderson, Geoffrey Chew, Leon Cooper, Sidney Drell, Freeman Dyson, Murray Gell-Mann, Roy Glauber, Marvin Goldberger, Walther Kohn, Norman Kroll, Francis Low, Quinn Luttinger, Marshall Rosenbluth, Arthur Wightman, …

  85. Bethe, High Energy Phenomena. A course of lectures given at Los Alamos in the spring and summer of 1952, concerning phenomena involving particles with energy in the range of hundreds of Mev. Main emphasis is placed on the properties of [pi]-mesons and on a relativistic treatment of the nucleon–nucleon interaction. (Los Alamos, 1953).

  86. Freeman Dyson, Marc Ross, Edwin E. Salpeter, Silvan S. Schweber, M. K. Sudarshan, William M. Vissher and Hans A. Bethe, “Meson-Nucleon Scattering in Tamm-Dancoff Approximation,” Physical Review 95 (1954), 1644–58.

  87. Geoffrey F. Chew and Francis Low, “Effective-range Approach to the Low-energy p-wave Pion-Nucleon Interaction,” Physical Review 101 (1956), 1570–9; “Theory of Photomeson Production at Low Energies,” Physical Review 101 (1956), 1579–87. Perhaps one the reasons that the success of Chew and Low’s approach may have affected Bethe so deeply is that he had formalized the “effective-range” approach to low energy nucleon–nucleon scattering and was committed to what later would be called an “effective field theory”: Bethe, “Theory of the Effective Range in Nuclear Scattering,” Physical Review 76 (1949), 38–50. See also Tran N. Truong, “Bethe-Schwinger Effective Range Theory and Lehmann and Weinberg Chiral Perturbation Theories,” paper presented at ICFP 09, September 24–30, 2009, Hanoi, Vietnam.

  88. One of the first people I heard make this distinction was Alexi Assmuth in a talk at Harvard in the late 1980s on Philip Anderson and the SSC. Assmuth had characterized high energy theories as “foundational.” The language of “foundations” had entered historical narratives through the work of Clifford Geertz, particularly, his essays in The Interpretation of Cultures (New York: Basic Books, 1973) and the history of science was undergoing its own cultural turn in the late 1980s.

  89. I am calling them “difficult” to contrast them with Bethe’s appellation of the dismal 1930s as the “happy thirties.”

  90. See Schweber, In the Shadow of the Bomb (ref. 26) for Bethe’s involvement in overcoming the difficulties that his colleague Philip Morrison was experiencing because of his liberal political views.

  91. For an overview of Brueckner’s work, see his lectures in the 1958 Les Houches summer school: Keith Brueckner, “Theory of Nuclear Structure and of Many Body Systems” in The Many Body Problem: Le Problème à n corps. Cours donnés à l’École d’Été de physique théorique (Les Houches 1958), 47–242.

  92. H.A. Bethe and R. Bacher, “Nuclear Physics. A: Stationary States of Nuclei,” Reviews of Modern Physics 8 (1936), 82–229; Bethe, “Nuclear Physics. B: Nuclear Dynamics, Theoretical,” Reviews of Modern Physics 9 (1937), 69–244.

  93. The binding energy of an additional proton beyond the number of protons in a closed shell was less than that that of the protons in the closed shell; likewise for neutrons.

  94. See Maria Goeppert Mayer, Elementary Theory of Nuclear Shell Structure (New York: Wiley, 1955).

  95. Oral History Transcript of interview with Hans A. Bethe by Charles Weiner at Cornell University May 8, 1972. Niels Bohr Library, Center of the History of Physics; American Institute of Physics, College Park, MD.

  96. Goldstone, using a diagrammatic representation of the perturbative scheme, proved that the contributions of all unlinked diagrams—the contributions of which had given Brueckner difficulties in his non-diagrammatic, algebraic approach—added up to zero. Goldstone also proved that each order of perturbation theory gives a contribution proportional to the total number of particles, this to all orders of perturbation theory.

  97. David James Thouless, “The Application of Perturbation Methods to the Theory of Nuclear Matter.” PhD diss., Cornell University (1958).

  98. Jeffrey Goldstone, “Derivation of the Brueckner Many-Body Theory,” Proceedings of the Royal Society London A 239 (1957), 267–279.

  99. John Bardeen, Leon Neil Cooper, and John Robert Schrieffer, “Theory of Superconductivity,” Physical Review 108 (1957), 1175–1204.

  100. Chien Shiung Wu et al.,”Experimental Test of Parity Conservation in Beta Decay,” Physical Review 105 (1957), 1413–5; Tsung-Dao Lee and Chen Ning Yang, “Question of Parity Conservation in Weak Interactions,” Physical Review 104 (1956), 254–8.

  101. Les problèmes mathématiques de la théorie quantique des champs, (Lille: Colloques Internationaux du Centre Nationale de la Rechèrche Scientifique, 3–8 Juin 1957).

  102. See Wang, In Sputnik’s Shadow (ref. 31).

  103. My paper “Hacking the Quantum Revolution” substantiating this claim is being submitted to Studies in History and Philosophy of Science. B: Studies in History and Philosophy of Modern Physics.

  104. Tom Banks, Modern Quantum Field Theory: A Concise Introduction (Cambridge: Cambridge University Press, 2008), 137. Conversely, the success of the approach of describing physical phenomena in terms effective field theories is a reflection of the fact that appropriately isolated physical phenomena in a certain energy regime, probed and analyzed by instruments able to resolve effects only within a certain range of length scales, can be described most simply by a set of effective degrees of freedom appropriate to that scale.

  105. See G. Peter Lepage, “What is Renormalization?” ArXiv:hep-ph/0506330. Talk given at the Theoretical Advanced Study Institute in Elementary Particle Physics (TASI). University of Colorado, Boulder, Colorado, in 1989; “How to Renormalize the Schrödinger Equation”, ArxXiv:nucl-th/9706029, 1997 Lectures given at the VIII Jorge Andre Swieca Summer School (Brazil, 1997).

  106. Philip W. Anderson, “More is Different,” Science 177 (1972), 393–6.

  107. Ten or so component atoms are the current limit for accurate ab initio computations of molecular structure.

  108. There is a vast literature on the subject with which I am only superficially acquainted. Daniel T. Rodgers, Age of Fracture (Cambridge, MA: Harvard University Press, 2011) is a most valuable overview within the American context. In addition to Forman’s writings, I have found the following articles and books useful entries: Mirowski and Sent, Science Bought and Sold (ref. 6); David Tyfield, The Economics of Science: a Critical Realist Overview (New York: Routledge, 2012); Alfred Nordmann, Hans Radder, and Gregor Schiemann, eds., Science Transformed?: Debating Claims of an Epochal Break (Pittsburgh: University of Pittsburgh Press, 2011); Hans Radder, ed., The Commodification of Academic Research: Science and the Modern University (Pittsburgh: University of Pittsburgh Press, 2010); Philipp Mirowski, Science-Mart (ref. 7); Helga Novotny et al, eds., The Public Nature of Science under Assault (Berlin: Springer 2005); Dominique Pestre, “The Technosciences between Market, Social Worries, and the Political: How to Imagine a Better Future,”, in ibid., 29–52; Pestre, “The Historical Heritage of the 19th and 20th Centuries: Techno-science, Markets, and Regulations in a Long-term Perspective,” History and Technology 23, no. 4 (2007), 407–420. For an overview of current forms of government(ality) and their interaction with science and technology, see Sheila Jasanoff, Designs on Nature: Science and Democracy in Europe and the United States. (Princeton, NJ: Princeton University Press, 2005); Pestre, “Challenges for the Democratic Management of Technoscience: Governance, Participation and the Political Today”, Science as Culture 17(2) (2008), 101–19; Pestre, “Understanding the Forms of Government in Today’s Liberal Societies: An Introduction,” Minerva 47, no. 3 (2009), 243–60.

  109. See Forman, “From the Social to the Moral to the Spiritual: The Postmodern Exaltation of the History of Science,” in Jürgen Renn and Kostas Gavroglu, eds., Positioning the History of Science (Berlin: Springer Verlag, 2007), 49–55; Forman, “The Primacy of Science in Modernity” (ref. 6), “On the Historical Forms of Knowledge Production” (ref. 6).

  110. See Chapter 2 of Rodgers’ Fracture (ref. 108) for a history of the changes of perspectives and assumptions in the disciple of economics, including neoliberalism.

  111. The Patent and Trademark Law Amendments Act—now known as the Bayh-Dole Act —was enacted by the US Congress in December 1980. The legislation gave American universities, small businesses, and non-profit organizations exclusive patenting rights of inventions and control and property rights over intellectual materials that resulted from governmental funding. The legislation had been sponsored by Senators Birch Bayh of Indiana and Bob Dole of Kansas.

  112. For an account of the transformation of American universities, in particular of the bio-medical sciences, see Mirowski, Science-Mart (ref. 7).

  113. “What the Past Tells us about the Future of Science,” in José Manuel Sánchez Ron, ed., La Ciencia y la Tecnologia ante el Tercer Milenio (Madrid: Sociedad Estatal España Nuevo Milenio, 2002), 27–37, on 27.

  114. The scholarship that went into “The Primacy of Science in Modernity, of Technology in Postmodernity, and of Ideology in the History of Technology” and his Osiris 2012 article “On the Historical Forms of Knowledge Production and Curation: Modernity Entailed Disciplinarity, Postmodernity Entails Antidisciplinarity,” is most impressive. The Primacy article is 72 pages long; its 424 endnotes are set in small type and take up 56 pages and its bibliography, 24. The Osiris paper is only 45 pages long. However, one third of most the pages are taken up by lengthy footnotes listing an enormous number of articles and books referring to the subject matter being discussed, and many of them are commented on, at times very critically.

  115. Forman, “On the Historical Forms of Knowledge Production” (ref. 6), 58n7.

  116. Forman, “The Primacy of Science in Modernity, of Technology in Postmodernity and of Ideology in the History of Technology” (ref. 6).

  117. The following are the introductory remarks in the web page of the MIT department of bioengineering: “The Department of Biological Engineering was founded in 1998 as a new MIT academic unit, with the mission of defining and establishing a new discipline fusing molecular life sciences with engineering. The goal of this biological engineering discipline is to advance fundamental understanding of how biological systems operate and to develop effective biology-based technologies for applications across a wide spectrum of societal needs including breakthroughs in diagnosis, treatment, and prevention of disease, in design of novel materials, devices, and processes, and in enhancing environmental health. ” http://web.mit.edu/be/index.shtm, last accessed April 22, 2014. The Stanford University web page reads:” The mission of [the] Department of Bioengineering is to create a fusion of engineering and the life sciences that promotes scientific discovery and the development of new biomedical technologies and therapies through research and education. http://bioengineering.stanford.edu/, last accessed April 22, 2014.

  118. In “Hacking the Quantum Revolution” (ref. 103), I take issue with Laughlin and Pines’s formulation of the foundational theory that is the point of departure for Laughlin’s assertions in A Different Universe. Robert B. Laughlin and David Pines, “The Theory of Everything,” Proceedings of the National Academy of Sciences 97, no. 1 (2000), 28–31.

  119. See Richard Holloran, “Protracted Nuclear War,” Air Force Magazine 91, no. 3 (2008). The substance of NSDD 32 and NSDD 75 was divulged in the New York Times on May 30, 1983 in an op-ed column by Holloran.

  120. See Sonja Michelle Amadae, Rationalizing Capitalist Democracy: The Cold War Origins of Rational Choice Liberalism (Chicago: The University of Chicago Press, 2003) and particularly S.M. Amadae, “Cold War, Security Dilemma, and Prisoner’s Dilemma: Does Insecurity Rationalize Hegemony?” (unpublished manuscript).

  121. The rhetoric of “nuclear tipped”—rather than for example “having nuclear capabilities”—was a deliberate attempt to minimize the devastation and chaos that would be wreaked by a protracted nuclear war.

  122. William Broad, Teller’s War: The Top-Secret Story Behind the Star Wars Deception (New York: Simon & Schuster, 1992).

  123. Space-Based Missile Defense: A Report by the Union of Concerned Scientists (Cambridge, MA:,Union of Concerned Scientists. March 1984).

  124. Kurt Gottfried and Bruce Blair, Crisis Stability and Nuclear War (New York: Oxford University Press, 1988).

  125. See Bethe’s list of publications in Hans A. Bethe, Selected Works with Commentary (Singapore: World Scientific, 1997).

  126. Hans A. Bethe, [Untitled Statement], Federation of American Scientists Public Interest Report 48, no. 5 (September-October 1995), 8.

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

Download references

Acknowledgments

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

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Silvan S. Schweber.

Additional information

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

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Schweber, S.S. Writing the Biography of Hans Bethe: Contextual History and Paul Forman1 . Phys. Perspect. 16, 179–217 (2014). https://doi.org/10.1007/s00016-014-0136-6

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00016-014-0136-6

Keywords

Navigation