Hostname: page-component-77c89778f8-vsgnj Total loading time: 0 Render date: 2024-07-19T20:23:55.793Z Has data issue: false hasContentIssue false

Scientists, Foreign Policy, and Politics

Published online by Cambridge University Press:  02 September 2013

Warner R. Schilling
Affiliation:
Columbia University

Extract

… we must take, so far as we can, a picture of the world into our minds. Is it not a startling circumstance for one thing that the great discoveries of science, that the quiet study of men in laboratories, that the thoughtful developments which have taken place in quiet lecture rooms, have now been turned to the destruction of civilization? … The enemy whom we have just overcome had at its seats of learning some of the principal centres of scientific study and discovery, and used them in order to make destruction sudden and complete; and only the watchful, continuous cooperation of men can see to it that science, as well as armed men, is kept within the harness of civilization.

These words were spoken in Paris in January 1919 by Woodrow Wilson, addressing the second Plenary Session of the Peace Conference. Wilson believed he had found a watchdog for civilization in the League of Nations. In this he was sadly mistaken. Science and armed men have indeed been harnessed, but in order to promote and maintain the goals of conflicting polities. Whether in the pursuit of these ends the cause of civilization will yet be served remains, we may hope, an open question.

Type
Research Article
Copyright
Copyright © American Political Science Association 1962

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

1 U. S. Department of State, Papers Relating to the Foreign Relations of the United States, The Peace Conference, 13 vols. (Washington, 19421947), vol. 3, p. 179Google Scholar.

An earlier version of this paper was prepared for discussion at the Fifth Congress of the International Political Science Association in Paris, September, 1961. The points made in it owe much to the comment and counsel of William T. R. Pox.

2 Quoted in Science and Technology Act of 1958, Staff Study of the Senate Committee on Government Operations, 85th Cong., 2d sess., Washington, 1958, p. 110Google Scholar.

3 Ibid., p. 115.

4 See Daniels' letter to Edison, in Daniels, Josephus, The Wilson Era: Years of Peace, 1910–1917 (Chapel Hill: The University of North Carolina Press, 1944), p. 491Google Scholar.

5 Johnson, Ellis A., “The Crisis in Science and Technology and its Effect on Military Development,” Operations Research, 01-February 1958, pp. 1415Google Scholar.

6 See DuBridge, Lee A., “The American Scientist: 1955,” Yale Review, Spring 1955, p. 13Google Scholar, and the Bulletin of the Atomic Scientists, 03 1957, p. 82, and May-June 1961, p. 254Google Scholar. The figure for private industry is for the year 1959; the others are for the year 1955.

7 For a more detailed treatment of some of the points in the preceding paragraphs and a general discussion of the effect of science on international relations, see the present writer's Science, Technology, and Foreign Policy,” Journal of International Affairs, Fall 1959, pp. 718Google Scholar.

8 Cf. the implication in the following remarks of Glenn T. Seaborg, the Chairman of the Atomic Energy Commission: “Scientists don't necessarily have to make the final political decisions, but it might be easier to let a capable scientist learn political reality than to teach a politician science.” Quoted in the Bulletin of the Atomic Scientists, 02 1961, p. 79Google Scholar.

9 In this and subsequent undocumented references the present writer has drawn upon personal interviews during 1956–1958 with participants in the H-bomb decision.

10 For the “principle of least harm,” see Brodie, Bernard, “Strategy as a Science,” World Politics, 07 1949, p. 479nGoogle Scholar. On the H-bomb choice, see the present writer's “The H-Bomb Decision: How To Decide Without Actually Choosing,” Political Science Quarterly, 03 1961, pp. 3738Google Scholar.

11 See Snow, C. P., Science and Government (Cambridge: Harvard University Press, 1961), pp. 4751Google Scholar, the review of this book by Blackett, P. M. S. in Scientific American, 04 1961, pp. 192194Google Scholar, and Churchill, Winston S., The Second World War: The Hinge of Fate (Boston: Houghton Mifflin Company, 1950), p. 281Google Scholar. For British air doctrine see also Dinnerstein, Herbert S., “The Impact of Air Power on the International Scene, 1933–1940,” Military Affairs, Summer 1955, pp. 6768Google Scholar.

12 Maj. Gen.Dornberger, Walter, V-2 (New York: Ballantine Books, 1954), pp. 97, 158160Google Scholar, and Lt. Gen.Gavin, James M., War and Peace in the Space Age (New York, 1958), pp. 7677Google Scholar.

13 Note should also be taken of the problem the policy-maker faces when all his experts are agreed. The present writer is unable to suggest a useful procedure here (other than variations on numbers five, six, and seven above); but that the problem is a real one can be seen in the conclusion of the German physicists that it would be infeasible for any Power to develop an atomic bomb during World War II. Some of the German scientists later stated that political considerations were partly responsible for their advice and for the fact that they made so little progress themselves on an A-bomb (cf. procedure one).

The German work on the A-bomb during World War II is described in Goudsmit, Samuel A., Alsos (New York: Henry Schuman, Inc., 1947)Google Scholar. For various appraisals of the influence exercised by political considerations, see Jungk, Robert, Brighter than a Thousand Suns (New York: Harcourt, Brace and Company, 1958), pp. 88104Google Scholar, Bethe, Hans in the Bulletin of the Atomic Scientists, 12 1958, p. 427Google Scholar, and Laurence, William L., Men and Atoms (New York: Simon and Schuster, 1959), pp. 9093Google Scholar.

14 Snow, C. P., The Two Cultures and the Scientific Revolution (New York: Cambridge University Press, 1959), pp. 911Google Scholar.

15 I am indebted to Hans Speier for the phrasing of this point.

16 Price, Don K., Government and Science (New York: New York University Press, 1954), pp. 1011Google Scholar.

17 This persuasion was largely accomplished through demonstrations of the military utility of the scientists' taking such an approach, although in the early history of the M.I.T. Radiation Laboratory a certain amount of polite bargaining was apparently practiced. One scientist involved, whenever told that the reason for a request was a problem for Washington, not him, to worry about, adopted the practice of working on something else until he was given a description of the problem involved. For a brief summary of the British experience, see Haddow, Alexander, “The Scientist as Citizen,” Bulletin of the Atomic Scientists, 09 1956, p. 247Google Scholar.

18 Cf. the following exchange between Gordon Gray and Jerrold Zacharias during the Oppenheimer hearing. Gray: “If you were directing a study which had to do with electronics, a pretty clearly defined field, and it started to come up with recommendations with respect to foreign policy, would you feel that an official of the Defense Department who urged that you stick to electronics was acting with impropriety?” Zacharias: “I think I would not direct a project that was as restrictive as that, sir, as to be restricted only to electronics.” U. S. Atomic Energy Commission, In the Matter of J. Robert Oppenheimer, Transcript of Hearing before Personnel Security Board, Washington, 1954, p. 930Google Scholar.

For some of the issues involved in the 1952 air defense study, see ibid., pp. 598–99, 749–50, 763–65, 923–24, 930–31, 935, 938, and also the account in Price, Government and Science, pp. 136–38.

19 General Leslie Groves, who directed the Manhattan project, was especially sensitive to the scientists' tendency to take on the whole problem. (Some even advised him on how the garbage should be collected at Los Alamos, an act which may possibly have reflected self- rather than scientific interest.) One reason for his effort to compartmentalize the work scientists were doing was his fear that “if I brought them into the whole project, they would never do their own job. There was just too much of scientific interest, and they would just be frittering from one thing to another.” Oppenheimer Transcript, p. 164.

20 See, for example, Berkner, Lloyd V., “Science and National Strength,” Bulletin of the Atomic Scientists, 06 1953, pp. 155, 180Google Scholar.

21 See the Bulletin of the Atomic Scientists, 12 1959, p. 412Google Scholar.

22 Oppenheimer Transcript, p. 33, and Dornberger, , V-2, pp. 134137Google Scholar.

23 Oppenheimer Transcript, p. 251. For an extreme judgment, see Jungk, , Brighter Than a Thousand Suns, p. 296Google Scholar.

24 See Oppenheimer's statements in Oppenheimer Transcript, pp. 81, 251, 897, and “The H-bomb Decision: How to Decide Without Actually Choosing,” loc. cit., pp. 30–36.

25 SirWatson-Watt, Robert, Three Steps to Victory (London: Odhams, 1957), p. 74Google Scholar.

26 In 1955 slightly more than half of the active research physicists in the United States were under forty years of age and had received their doctorates after December 7, 1941. DuBridge, Lee A., “The American Scientist: 1955,” Yale Review, 09 1955, p. 1Google Scholar.

27 These assumptions are excellently set forth in Stahl, Margret Smith, “Splits and Schisms: Nuclear and Social,” unpublished doctoral dissertation, University of Wisconsin, 1946, ch. 4Google Scholar.

28 For the activities of the Panel and the Office, see James R. Killian, “Science and Public Policy,” Address to the American Association for the Advancement of Science, December 29, 1958, as printed in Science Program—86th Congress, Report of the Senate Committee on Government Operations, 86th Cong., 1st sess. (1959), pp. 12–13, and The Science Adviser of the Department of State, Department of State Publication 7056 (Washington, 1960)Google Scholar.

29 See Stahl, op. cit., ch. 4.

30 See Kramish, Arnold, Atomic Energy in the Soviet Union (Stanford: Stanford University Press, 1959), p. 105Google Scholar. Kramish states that it is not certain whether the objections of the Russian scientists were technical or political. For the declaration of the German physicists, see the Bulletin of the Atomic Scientists, 06 1957, p. 228Google Scholar.

31 Oppenheimer Transcript, pp. 584, 594–95, 891–94.

32 See Sullivan, Walter, Assault on the Unknown (New York: McGraw-Hill, 1961), pp. 7981Google Scholar.

33 Snow, , Science and Government, pp. 6668Google Scholar, and Dornberger, , V-2, p. 87Google Scholar.

34 There are eighteen scientists on the President's Science Advisory Committee; its working panels also contain participants from outside the committee. In December 1958 the Committee and the Office of the Special Assistant for Science and Technology had together some 75 scientists and engineers serving part time. See Killian, “Science and Public Policy,” loc. cit., p. 8. The work of the Committee and the Office are additionally described and appraised in Science Organization and the President's Office, Staff Study of the Subcommittee on National Policy Machinery, Senate Committee on Government Operations, 87th Cong., 1st sess. (1961).

The information presented about the Department of State is based on U. S. Department of State, The Science Adviser of the Department of State, and interviews with several Department officials in February 1962. Needless to say, the description and interpretation made above are entirely the present writer's responsibility.

35 These two conditions are not unrelated. The more influence the Department exercises in determining the goals and programs of other agencies, the more confident it can be that scientists in those agencies will call the Department's attention to goals and programs which they believe to be receiving too much or too little attention.

36 See Berkner, Lloyd V., “National Science Policy and the Future,” Address at Johns Hopkins University, 12 16, 1958Google Scholar, as printed in Science Program—86th Congress, pp. 116–18.

37 This point, especially as it relates to science experts, is discussed in Price, Government and Science, pp. 61–62, and in Finer, Herman, “Government and the Expert,” Bulletin of the Atomic Scientists, 11 1956, pp. 331–32Google Scholar.

38 See the discussion in Price, Government and Science, pp. 131, 133, 138–42. The point about the scientists' lacking a tradition of “civilian control” was suggested by William T. R. Fox.

39 U. S. Atomic Energy Commission, In the Matter of J. Robert Oppenheimer, Texts of Principal Documents and Letters (Washington, 1954), pp. 1920Google Scholar. Note the policy predisposition in the phrase “strongest offensive military interests.”

It should not be comfortable for an American to reflect on the career of Peter Kapitsa, a Soviet physicist who was a student of Rutherford and who worked in England from 1922 to 1934 and then returned to the Soviet Union. Kapitsa was placed under house arrest in 1947 and remained there until after Stalin's death. Kapitsa has told western scientists and newsmen that his arrest was the result of his refusal to work on nuclear energy for military purposes. Kramish believes that his arrest was due to the government's dissatisfaction with his advice on certain technical approaches to weapons development. In either event, it is noteworthy that Kapitsa is believed to have recently been, on an informal basis, one of Khrushchev's main science advisers.

On the matter of his arrest, see the report by Salisbury, Harrison in the New York Times, 07 11, 1956Google Scholar; the Bulletin of the Atomic Scientists, 01 1957, p. 38Google Scholar; and Kramish, , Atomic Energy in the Soviet Union, pp. 109110Google Scholar. The information on his recent activity was supplied by the staff of the Subcommittee on National Policy Machinery, Senate Committee on Government Operations.

40 On Soviet government and science organization, see National Policy Machinery in the Soviet Union, Report of the Subcommittee on National Policy Machinery, Senate Committee on Government Operations, 86th Cong., 2d sess. (Washington, 1949), pp. 24–35, 5962Google Scholar, and DeWitt, Nicholas, “Reorganization of Science and Research in the U.S.S.R.,” Science, 06 23, 1961, pp. 1981–91Google ScholarPubMed. The points made above were additionally confirmed by the staff of the Subcommittee on National Policy Machinery.

41 The circumstances provide an interesting variation of the “whole problem approach.” The Tizard Committee was initially interested in techniques for destroying aircraft or their crews, and Watson-Watt was asked in 1935 to investigate the possibility of using electromagnetic radiation for this purpose. He reported that such a use was apparently infeasible. In any event, he went on to note, the aircraft would first have to be located, and if anyone was interested electromagnetic radiation might be useful for this. Watson-Watt, , Three Steps to Victory, pp. 8183Google Scholar.

42 For the development of radar, see ibid., pp. 108–09; Snow, C. P., Science and Government, pp. 24–38, 60–61, 7475Google Scholar; Blackett, P. M. S., “Tizard and the Science of War,” Nature, 03 5, 1960, pp. 648–49Google Scholar; and Collier, Basil, The Defense of the United Kingdom (London: H.M.S.O., 1957), pp. 33, 3639Google Scholar.

43 Ironically, the British were mistaken in their theory. The German Air Force had no such strategy in mind, and in 1940 when it tried to improvise a strategic bombing campaign it had neither the equipment nor the doctrine with which to conduct the campaign effectively. See Dinnerstein, Herbert, “The Impact of Air Power on the International Scene: 1933–1940,” Military Affairs, Summer 1955, pp. 6571Google Scholar; Taylor, Telford, The March of Conquest (New York: Simon and Schuster, 1958), pp. 2430Google Scholar; and Galland, Adolf, The First and the Last (New York: Ballantine Books, 1954), chs. 2–5Google Scholar.

Submit a response

Comments

No Comments have been published for this article.