Skip to main content

An Analysis of Interactions between the Natural Sciences and the Social Sciences

  • Chapter
The Natural Sciences and the Social Sciences

Part of the book series: Boston Studies in the Philosophy of Science ((BSPS,volume 150))

Abstract

Ever since the time of Aristotle, the natural sciences and medicine have furnished analogies for studies of governments, classifications of constitutions, and analyses of society.* One of the fruits of the Scientific Revolution was the vision of a social science — a science of government, of individual behavior, and of society — that would take its place among the triumphant sciences, producing its own Newtons and Harveys. The goal was not only to achieve a science with the same foundations of certain knowledge as physics and biology; there was thought to be a commonality of method that would advance the social sciences in the way that had worked so well in the physical and biological sciences. Any such social science, it was assumed, would be based on experiments and critical observations, would become quantitative, and would eventually take the highest form known to the sciences — expression in a sequence of mathematical equations.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 169.00
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 219.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 219.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Preview

Unable to display preview. Download preview PDF.

Unable to display preview. Download preview PDF.

Notes

  1. All of the examples introduced in this introductory section are discussed in full, with bibliographic documentation, in the succeeding sections.

    Google Scholar 

  2. On Cournot’s important contributions to mathematical economics, see Claude Ménard: La formation d’une rationalité économique: A.A. Cournot (Paris: Flammarion, 1978); also Joseph Schumpeter: History of Economic Analysis (New York: Oxford University Press, 1954).

    Google Scholar 

  3. See Note on Natural Science & Social Science on pp. xxv-xxxvi.

    Google Scholar 

  4. On the history of the concept of behavioral sciences, see Bernard Berelson on “Behavioral Sciences” in International Encyclopedia of the Social Sciences, ed. David L. Sills, vol. 2 (New York: The Macmillan Company & The Free Press, 1968), pp. 41–45; also Herbert J. Spiro: “Critique of Behavioralism in Political Science,” pp. 314–327 of Klaus von Beyme (ed.): Theory and Politics, Theorie und Politik, Festschrift zum 70. Geburtstag für Carl Joachim Friedrich (The Hague: Martinus Nijhoff, 1971).

    Google Scholar 

  5. An admirable discussion of the differences in these three usages is given in John Theodore Merz: A History of European Thought in the Nineteenth Century (Edinburgh/London: W. Blackwood and Sons, 1903–1914; reprint, New York: Dover Publications, 1965; reprint, Gloucester: Peter Smith, 1976), vol. 1, chs. 1, 2, 3.

    Google Scholar 

  6. See Marie Boas Hall: All Scientists Now: the Royal Society in the Nineteenth Century (Cambridge/New York: Cambridge University Press, 1984); also Dorothy Stimpson: Scientists and Amateurs: A History of the Royal Society (New York: Henry Schuman, 1948). The Royal Society, however, originally had a large proportion of non-scientists as Fellows, including poets (e.g., John Dryden), doctors, and peers of the realm. After the reorganization of 1847 the non-scientific categories of membership were eliminated, although exceptions are still made (e.g. Prince Philip and the financier-philanthropist Isaac Wolfson).

    Google Scholar 

  7. See n. 21 infra; also Chapter 12 infra.

    Google Scholar 

  8. See Roger Hahn: The Anatomy of a Scientific Institution: The Paris Academy of Sciences (Berkeley: University of California Press, 1971).

    Google Scholar 

  9. At first there were four classes: “Physica” (including chemistry, medicine, and other natural sciences), “Mathematica” (including astronomy and mechanics), German philos- ophy, and literature (especially oriental literature). Later these classes were regrouped into two major divisions: the natural sciences and mathematics in one and “philosophical and historical” domains in the other. See Erik Amburger (ed.): Die Mitglieder der Deutschen Akademie der Wissenschaften zu Berlin,1700–1950 (Berlin: Akademie-Verlag, 1960); Kurt-Reinhard Biermann & Gerhard Dunken (eds.): Deutsche Akademie der Wissenschaften zu Berlin: Biographischer Index der Mitglieder (Berlin: Akademie-Verlag, 1960). On the history and vicissitudes of the German Academy, see Werner Hartkopf & Gerhard Dunken: Von der Brandenburgischen Sozietät der Wissenschaften zur Deutschen Akademie der Wissenschaften zu Berlin (Berlin: Akademie-Verlag, 1967); the standard history is Adolph Harnack: Geschichte der Königlich Preussischen Akademie der Wissenschaften zu Berlin, 3 vols. (Berlin: Reichsdruckerei, 1990).

    Google Scholar 

  10. See note 2 supra.

    Google Scholar 

  11. See Samuel Johnson: A Dictionary of the English Language, 2 vols. (London: printed by W. Strahan for J. and P. Knapton, T. and T. Longman, C. Hitch and L. Hawes, A. Millar and R. and J. Dodsley, 1755; photo-reprint, New York: Arno Press, 1979).

    Google Scholar 

  12. The Works of Lord Macaulay, ed. Lady Trevelyan, vol. 5 (London: Longmans, Green, and Co., 1871), p. 677. In 1829 Macaulay wrote (ibid, p. 270) that the “noble Science of Politics” was “of all sciences… the most important to the welfare of the nations,” the science which “most tends to expand and invigorate the mind.” Additionally, he declared, the “Science of Politics” is notable among “all sciences” because it “draws nutriment and ornament from every part of philosophy and literature, and dispenses in return nutriment and ornament to all.” See also Stefan Collini, Donald Winch, & John Burrow: That Noble Science of Politics (Cambridge: Cambridge University Press, 1983), passim, but esp. pp. 102-103, 120. The negative version of Macaulay’s statement occurs in the oft-quoted remark made by Bismarck in the Prussian Chamber on 18 December 1863, “Die Politik ist keine exakte Wissenschaft,” that is, “Politics is not an exact science.” Early in the eighteenth century, in Gulliver’s Travels (1726), Jonathan Swift regretted that the ignorant Brobdignagians had not as yet “reduced Politicks into a Science.”

    Google Scholar 

  13. “ A Letter to a Noble Lord,” in Edmund Burke: The Works (London: John C. Nimmo, 1887; reprint, Hildesheim/New York: Georg Olms Verlag, 1975), vol. 5, p. 215; David Hume: A Treatise of Human Nature, ed. L.A. Selby-Bigge (Oxford: Clarendon Press, 1896 and reprints), pp. ix and xix—xx. In An Inquiry Concerning the Human Understanding, ed. L.A. Selby-Bigge (Oxford: Clarendon Press, 1894), pp. 83–84, Hume wrote of historical “records” as “so many collections of experiments, by which the politician or moral philosopher fixes the principles of his science, in the same manner as the physician or natural philosopher becomes acquainted with the nature of plants, minerals, and other external objects, by the experiments which he forms concerning them.” On Hume and a science of politics, the studies of Duncan Forbes are of primary importance, notably his introduction to the reprint of Hume’s History of Great Britain (Harmondsworth: Penguin, 1970); “Sceptical Whiggism, Commerce and Liberty,” pp. 179–201 of A.S. Skinner & T. Wilson (eds.): Essays on Adam Smith (Oxford: Oxford University Press, 1976); “Hume’s Science of Politics,” pp. 39–50 of G.P. Morice (ed.): David Hume, Bicentenary Papers (Edinburgh: Edinburgh University Press, 1977). See also James E. Force & Richard H. Popkin: Essays on the Context, Nature, and Influence of Isaac Newton’s Theology (Dordrecht/Boston/London: Kluwer Academic Publishers, 1990), ch. 10, “Hume’s Interest in Newton and Science” (by J.E. Force).

    Google Scholar 

  14. John Harris, Lexicon Technicum (London: printed for Dan. Brown, Tim. Goodwin, John Walthoe,…, Benj. Tooke, Dan. Midwinter, Tho. Leigh, and Francis Coggan, 1704; reprint, New York/London: Johnson Reprint Corporation, 1966 — The Sources of Science, no. 28) defines “Science” as “Knowledge founded upon, or acquir’d by clear, certain and self-evident Principles.” Harris says that “Natural Philosophy” is “the same with what is usually call’d Physicks, viz. That Science which contemplates the powers of Nature, the properties of Natural Bodies, and their mutual action one upon another.”

    Google Scholar 

  15. James S. Coleman: Foundations of Social Theory (Cambridge: The Belknap Press of Harvard University Press, 1990), p. xv. On this general topic, see Robert K. Merton: Social Theory and Social Structure (Enlarged edition, New York: The Free Press, 1968), ch. 1, “On the History and Systematics of Sociological Theory.”

    Google Scholar 

  16. Robert K. Merton: “The Mosaic of the Behavioral Sciences,” pp. 247–272 of Bernard Berelson (ed.): The Behavioral Sciences Today (New York/London: Basic Books, 1963), esp. p. 256.

    Google Scholar 

  17. A. Hunter Dupree: Science in the Federal Government: A History of Policies and Activities (Cambridge: Belknap Press of Harvard University Press, 1957; revised reprint, Baltimore: Johns Hopkins University Press, 1986).

    Google Scholar 

  18. Reported in Henry W. Riecken: “The National Science Foundation and the Social Sciences,” Social Science Research Council Items, Sept. 1983, 37(2/3): 39–42, esp. p. 40a.

    Google Scholar 

  19. For example, the association of “sociology” with “socialism” is discussed in Albion W. Small & George E. Vincent: An Introduction to the Study of Society (New York/Cincinnati/Chicago: American Book Company, 1894), pp. 40–41, where it is stated that “Systematic Socialism has both directly and indirectly promoted the development of Sociology.”

    Google Scholar 

  20. Riecken (n. 17 supra), p. 406.

    Google Scholar 

  21. See Chapter 12 infra.

    Google Scholar 

  22. The Ninth Annual Report of the National Science Foundation announced that “during the fiscal year 1959 [in December 1958], the Foundation established an Office of Social Sciences to support research and related activities in the social science disciplines.” The Eleventh Annual Report announced that “The Office of Social Sciences was reconstituted as the Division of Social Sciences during fiscal year 1961 [the year ending on 30 June 19611.” The new Division replaced “the previous Social Science Research Program” and was said to represent “a further step in the development of Foundation activities in the area.” The social sciences did not maintain this independent importance, however, and there is still some debate on how best to fit the social sciences into the structure of the National Science Foundation. A note in Science (16 Aug. 1991, 253: 727) on “a proposal to give the social sciences more clout within the agency” summarized the findings of a draft report by a “task force composed of 20 outside social and behavioral scientists and biologists” who recommended the removal of the social and behavioral sciences from their position as part of the Directorate for Biological, Behavioral and Social Sciences. Although most “social scientists” were in favor of this proposal, which would “give them an advocate at the highest level of the agency and win them more funding and respect,” some social scientists — it was noted — “don’t want to leave the fold.” On 23 Oct. 1991, The Chronicle of Higher Education (pp. A-23, A-25) reported that the reorganization had occurred, including the establishment of a new Directorate for Social, Behavioral and Economic Sciences. NSF Director Walter E. Massey expressed high hopes that the creation of a “separate new office” would “lead to an increase in funds for those sciences.” Although many “social scientists said the change would lead to more recognition and higher budgets,” support for the change “was not unanimous” (according to the Chronicle’s report) even among the social scientists concerned.

    Google Scholar 

  23. The work of scientists and social scientists mentioned briefly in various sections of this chapter is discussed at greater length and with bibliographical documentation in subsequent chapters or in other sections of this chapter.

    Google Scholar 

  24. Wilhelm Ostwald: Energetische Grundlagen der Kulturwissenschaften (Leipzig: Verlag von Dr. Werner Kilinkhardt, 1909). On Ostwald’ s “Kulturwissenschaften” see Philip Mirowski: More Heat than Light: Economics as Social Physics,Physics as Nature’s Economics (Cambridge/New York: Cambridge University Press, 1989), pp. 454–57, 132–133, 268. There were many scientists and social scientists who saw in the new “energetics” a basis for a reconstitution of economics, sociology, history, and so on. A notable example was Henry Adams, who attempted to use J. Willard Gibbs’s memoir on the “Equilibrium of Heterogeneous Substances” as the basis of a study on “The Rule of Phase Applied to History.” This essay is reprinted along with Adam’s “A Letter to American Teachers of History” in Brooks Adams (ed.): The Degradation of the Democratic Dogma (New York: The Macmillan Company, 1920).

    Google Scholar 

  25. See Chapter 2 infra.

    Google Scholar 

  26. See John Brewer: The Sinews of Power: War,Money and the English State, 1688–1783 (Cambridge: Harvard University Press, 1988), ch. 8, “The Politics of Information, Public Knowledge and Private Interest”; Keith Thomas: “Numeracy in Early Modern England,” Transactions of the Royal Historical Society, 1987, 37: 103–132.

    Google Scholar 

  27. See Chapter 4, §3, infra.

    Google Scholar 

  28. See Jacques Roger: Buffon (Paris: Fayard, 1989), pp. 234, 296.

    Google Scholar 

  29. Stephen M. Stigler: The History of Statistics: The Measurement of Uncertainty before 1900 (Cambridge: The Belknap Press of Harvard University Press, 1986), ch. 3, “Inverse Probability.” See further, Helen M. Walker: Studies in the History of Statistical Method (Baltimore: The Williams & Wilkins Company, 1929; reprint, New York: Arno Press, 1975), pp. 31–38; Hyman Alterman: Counting People: The Census in History (New York: Harcourt, Brace & World, 1969).

    Google Scholar 

  30. Keith M. Baker: Condorcet, from Natural Philosophy to Social Mathematics (Chicago: University of Chicago Press, 1975), ch. 4.

    Google Scholar 

  31. See Chapter 3 infra. In addition to Stigler’s History of Statistics (n. 28 supra), ch. 5, see Theodore M. Porter: The Rise of Statistical Thinking, 1820–1900 (Princeton: Princeton University Press, 1986), esp. pp. 2, 4; Ian Hacking: The Taming of Chance (Cambridge/New York: Cambridge University Press, 1990), chs. 13, 14, 19, 21; Frank H. Hankins: Adolphe Quetelet as Statistician (New York: Columbia University Press, 1908 — Studies in History, Economics and Public Law; reprint, New York: AMS Press, 1968). 31 Baker (n. 29 supra), p. 202.

    Google Scholar 

  32. Mirowski (n. 23 supra), ch. 5. Mirowski’s thesis has not produced universal acceptance by economists. Not only is it considered extreme, but it is faulted because it does not apply to all founders of neoclassical economics, for example, Karl Menger, and even Léon Walras (see n. 33 infra and §1.5 infra).

    Google Scholar 

  33. See further, §1.6 infra. Walras, we shall have occasion to note, argued for the similarity of economics and rational mechanics but (see n. 143 infra) did so only after he had produced his system of economics. That is, he did not create his economics by attempting to imitate rational mechanics.

    Google Scholar 

  34. See §1.5 infra.

    Google Scholar 

  35. William Breit & Roger W. Spencer (eds.): Lives of the Laureates: Seven Nobel Economists (Cambridge: MIT Press, 1986), p. 74. This topic is developed further in §1.6 infra.

    Google Scholar 

  36. In addition to the works cited in nn. 28, 30 supra, see Gerd Gigerenzer, Zeno Swijtink, Theodore Porter, Lorraine Daston, John Beatty, and Lorenz Krüger: The Empire of Chance: How Probability Changed Sciences and Everyday Life (Cambridge/New York: Cambridge University Press, 1989); William Coleman: Death is a Social Disease: Public Health and Political Economy in Early Industrial France (Madison: University of Wisconsin Press, 1982).

    Google Scholar 

  37. David Brewster: Letters on Natural Magic (New York: Harper & Brothers, 1843), p. 181. In The Glaciers of the Alps (Boston: Ticknor & Fields, 1861), p. 285, John Tyndall wrote that “the analogy between a river and glacier moving through a sinuous valley is therefore complete.”

    Google Scholar 

  38. Richard Owen defined these two terms as follows: analogue — “A part or organ in one animal which has the same function as another part or organ in a different animal”; homologue — “The same organ in different animals under every variety of form and function.” Richard Owen: On the Archetypes and Homologies of the Vertebrate Skeleton (London: Richard & John E. Taylor, 1848), p. 7. Despite Owen’s phrasing here, the terms “similarity of form” or “sameness of structure” may be used to represent the kind of likeness exemplified by “the same organ.” See further n. 40 infra; also Mayr (n. 40 infra), p. 464.

    Google Scholar 

  39. In Darwinian evolution, analogy is the result of parallel adaptation, the way in which different organisms in separate but parallel evolutionary stages have developed, independently of one another, different ways of “adapting themselves to the same external circumstances” or needs. An example is given by an organ of vision, in which a lens concentrates light on special sensitive tissue. Konrad Lorenz has noted that this “invention” had been made independently by animals of four different phyla, in two of which (the vertebrates and the cephalopods) this kind of “eye” has “evolved into the true, image-projecting camera through which we ourselves are able to see the world.” See Konrad Z. Lorenz: “Analogy as a Source of Knowledge,” Science, 1974, 185: 229234.

    Google Scholar 

  40. See Ernst Mayr: The Growth of Biological Thought: Diversity, Evolution, and Inheritance (Cambridge: The Belknap Press of Harvard University Press, 1982), p. 45, where it is noted that the “term `homologous’ existed already prior to 1859, but it acquired its currently accepted meaning only when Darwin established the theory of common descent. Under this theory the biologically most meaningful definition of `homologous’ is: `A feature in two or more taxa is homologous when it is derived from the same (or a corresponding) feature of their common ancestor.’”

    Google Scholar 

  41. “Homology” appears with special meanings in several sciences. In addition to the general evolutionary or biological sense, there is the chemical usage (referring to a family of organic compounds in which each member is distinguished from the next in the sequence by some constant factor, notably, a CH2 group), the mathematical usage (a topological classification), and a special usage in genetics (to indicate the same linear sequence of genes in two or more chromosomes).

    Google Scholar 

  42. It should be noted that the terms “analogue” and “homologue” are not being used in the present context in the strict sense of evolutionary biology, since in the analysis of the interactions of the social and the natural sciences there is no consideration of “common descent.” Furthermore, because “analogy” is often used to designate various types of correspondence, it is sometimes necessary, especially in quoting or paraphrasing the work of others, to employ this term to indicate likeness in more general senses than those specified above.

    Google Scholar 

  43. Henry C. Carey: Principles of Social Science (Philadelphia: J. B. Lippincott & Co., 1858), vol. 1, pp. 42–43.

    Google Scholar 

  44. Carey’s exact words are: “Gravitation is here, as everywhere else in the material world, in the direct ratio of the mass, and in the inverse one of the distance.” In vol. 3, ch. 55, p. 644, Carey recapitulates his physics and social science. He begins by stating “simple laws which govern matter in all its forms, and which are common to physical and social science.” The first of these reads: “All particles of matter gravitate towards each other — the attraction being in the direct ratio of the mass, and the inverse one of the distance.” Incidentally, it may be observed that Carey has also misunderstood the Newtonian explanation of orbital or curved motion, under the actions of a centripetal force, such as a planet moving under the action of the sun’s gravity plus its own component of inertia. Carey says: “All matter is subjected to the action of the centripetal and the centrifugal forces — the one tending to the production of local centres of action, the other to the destruction of such centres, and the production of a great central mass, obedient to but a single law.” We may take note that Carey also introduced ratios other than direct and inverse proportion. Thus, in vol. 1, p. 389, he wrote: “The motion of society, and the power of man, tend to increase in a geometrical ratio….”

    Google Scholar 

  45. Although “fallacy” is often used in a narrow technical sense to denote a flaw (or type of flaw) that “vitiates a syllogism,” a primary meaning in every dictionary I have consulted (OED,OED-suppl., OED — 2nd ed.; Concise Oxford Dictionary — 6th ed.; Webster’s New International — 2d & 3d eds.) is a misleading argument, or a delusion or error, or some unsoundness or delusiveness or disappointing character of an argument or belief. The American Heritage Dictionary gives as the first meaning: “An idea or opinion founded on mistaken logic or perception; a false notion”; other meanings include “the quality of being deceptive” and “incorrectness of reasoning or belief.” The only example given is a “romantic fallacy, that Shakespeare was superhuman.” This example displays features in common with two frequently encountered uses of “fallacy” today: John Ruskin’s notion of the “pathetic fallacy” (in which inanimate objects are supposed to have human emotions) and W.K. Wimsatt and Monroe Beardsley’s “intentional fallacy” (overstressing the author’s intentions in assessing a literary work). These usages are somewhat similar to Alfred North Whitehead’s “fallacy of misplaced concreteness” as presented in Science and the Modern World (New York: The Macmillan Company, 1931), ch. 4, pp. 82, 85.

    Google Scholar 

  46. Newton’s concept of mass has two separate aspects: one (inertial mass in post-Einstein terminology) is a measure of body’s resistance to being accelerated or being made to undergo a change in “state,” while the other (gravitational mass) is a measure of a body’s response to a given gravitational field (i.e., the weight). For details see my The Newtonian Revolution: with Illustrations of the Transformation of Scientific Ideas (Cambridge/ London/New York: Cambridge University Press, 1980). Newton recognized that in ordinary (i.e., non-relativistic) rational mechanics there is no logical reason why these two concepts or measures of mass should be equivalent. Accordingly, he instituted a series of experiments to show that one is always proportional to the other, that at any given location mass is proportional to weight. These experiments are described in Book 3, prop. 6, of the Principia, which reports how he experimented with “gold, silver, lead, glass, sand, common salt, wood, water, and wheat” and could have easily detected a variation of as little as one part in a thousand. Newton, of course, did not use such terms as “gravitational mass” or “inertial mass” but rather proved that for all such materials the ratio of the “weight” to “quantity of matter” (or mass) was the same.

    Google Scholar 

  47. William Jaffé: “Léon Walras’s Role in the `Marginal Revolution’ of the Late 1870s,” pp. 115–119 of R.D. Collison Black, A.W. Coates, and Craufurd D.W. Goodwin (eds.): The Marginal Revolution in Economics: Interpretation and Evaluation (Durham: Duke University Press, 1973).

    Google Scholar 

  48. For a later attempt by Walras to argue that his economics is analogous to Newtonian rational mechanics, see Philip Mirowski & Pamela Cook: “Walras’ ‘Economics and Mechanics’: Translation, Commentary, Context,” pp. 189–224 of Warren J. Samuels (ed.): Economics as Discourse (Boston/Dordrecht/London: Kluwer, 1990).

    Google Scholar 

  49. Berkeley, for example, produced a very significant critique of the foundations of Newton’s theory of fluxions, that is, Newton’s version of the calculus. His Siris was an attempt “to assimilate Newtonian concepts to the more complex phenomena of chemistry and animal physiology.” In his De motu he analyzed “Newtonian concepts of gravitational attraction, action and reaction, and motion in general.” See Gerd Buchdahl: “Berkeley, George,” Dictionary of Scientific Biography, vol. 2 (New York: Charles Scribner’s Sons, 1970), pp. 16–18.

    Google Scholar 

  50. See §1.9 infra.

    Google Scholar 

  51. For Newton, the motion resulting from an equilibrium of forces can only be constant speed along a straight line, not curved motion as along a planetary orbit. Newton’s analysis of orbital or curved motion was based on the concept of two independent components. One is an initial component of inertial (or linear) motion, the other a constantly accelerated motion of falling inward toward the center of force. A planet or other orbiting body, of course, does not actually move inward away from the orbit, even though it is constantly falling toward the center; the reason is that the forward motion along the tangent carries that body ahead at such a rate that it continually “falls” away from the tangent to the orbit. Newton said he gave the centrally directed force the name “vis centripeta” in honor of Christiaan Huygens who had made use of the opposite kind of force, “vis centrifuga.” For details, see my Newtonian Revolution (n. 46 supra). Since orbital motion involves the constant inward (or centrally directed) acceleration of falling, there is no condition of equilibrium.

    Google Scholar 

  52. Berkeley fully understood Newton’s explanation. He gave the correct Newtonian reason why the planets do not actually fall inward so as to join together at the center. They “are kept from joining together at the common centre of gravity,” he wrote, “by the rectilinear motions the Author of nature hath impressed on each of them.” This tangential or linear component, he continued, “concurring with the attractive principle,” produces “their respective orbits round the sun.” He concluded that if this linear component of motion should cease, “the general law of gravitation that is now thwarted would show itself by drawing them all into one mass” (George Berkeley: “The Bond of Society,” Works, ed. A.A. Luce and T.E. Jessop, vol, 7 [London/Edinburgh: Thomas Nelson and Sons, 1955], pp. 226–227).

    Google Scholar 

  53. Ibid., pp. 225–228; cf. George Berkeley: “Moral Attraction,” Works, ed. Alexander Campbell Fraser, vol. 4 (Oxford: Clarendon Press, 1901), pp. 186–190.

    Google Scholar 

  54. For additional materials concerning Berkeley’s Newtonian sociology, see my “Newton and the Social Sciences, with special reference to Economics: The Case of the Missing Paradigm,” to appear in Philip Mirowski (ed.): Markets Read in Tooth and Claw (Cambridge/New York: Cambridge University Press, 1993 [in press]) — Proceedings of a Symposium at Notre Dame on “Natural Images in Economics,” October 1991.

    Google Scholar 

  55. The eminent sociologist Pitirim A. Sorokin translated Berkeley’s correct Newtonian physics into a hodgepodge of incorrect pre-Newtonian explanations. Sorokin not only would have Berkeley make use of the misleading notion of a balance of centrifugal and centripetal forces, but continued his travesty by saying that Berkeley concluded that “Society is stable when the centripetal forces are greater than the centrifugal.” This is plainly nonsense even in pre-Newtonian physics; if the centripetal forces should be greater than the centrifugal forces, then obviously there would be no stability but an instability, a lack of balance or equilibrium, and a resultant motion inward, as Berkeley clearly stated would be the case under such circumstances. See Pitirim A. Sorokin: Contemporary Sociological Theories (New York/London: Harper & Brothers, 1928), p. 11.

    Google Scholar 

  56. See the writings of Duncan Forbes and of James E. Force (n. 12 supra).

    Google Scholar 

  57. David Hume: A Treatise of Human Nature (n. 12 supra), pp. 12–13.

    Google Scholar 

  58. If, as Hume believed, human behavior and social action are regulated by social laws, there is implied the possibility of a social science, one in which — as Hume wrote — “consequences almost as general and certain may sometimes be deduced… as any which the mathematical sciences afford us.” Seeking to establish a kind of psychology of individual action, Hume seems to have envisioned the construction of a new theoretical science that would ultimately find expression in practice. On the certainty of social laws compared to mathematics, see David Hume: “That Politics may be Reduced to a Science,” Essays: Moral,Political, and Literary, ed. T.H. Green & T.H. Grose (London: Longman, Green and Co., 1882; reprint, Aalen [Germany]: Scientia Verlag, 1964), vol. 1, p. 99.

    Google Scholar 

  59. Cf. Design for Utopia: Selected Writings of Charles Fourier, intro. Charles Gide, new foreword by Frank E. Manuel, trans. Julia Franklin (New York: Shocken Books, 1971 [orig. Selections from the Works of Fourier (London: Swan Sonnenschein & Co., 1901]), esp. p. 18; The Utopian Vision of Charles Fourier: Selected Texts on Work, Love,and Passionate Attraction, trans., ed., intro. Jonathan Beecher and Richard Bienvenu (Boston: Beacon Press, 1971), esp. pp. 1, 8, 10, 81, 84; Harmonian Man: Selected Writings of Charles Fourier, ed. Mark Poster, trans. Susan Hanson (Garden City: Doubleday & Company — Anchor Books, 1971). On Fourier, see Nicholas Y. Riasanovsky: The Teachings of Charles Fourier (Berkeley/Los Angeles: University of California Press, 1969) and Frank E. Manuel: The Prophets of Paris (Cambridge: Harvard University Press, 1962).

    Google Scholar 

  60. It is a fact of record that groups of idealists actually founded Fourierist utopian colonies along the bizarre lines he suggested and that Fourierism became a considerable political force in several countries.

    Google Scholar 

  61. Emile Durkheim: The Division of Labor in Society, trans. George Simpson (New York: The Free Press, 1933; reprint 1964), p. 339. Cf. Durkheim, De la division du travail social: étude sur l’organisation des sociétés supérieures (Paris: Félix Alcan, Éditeur, 1893), p. 378; Durkheim, De la division du travail social, 5th ed. (Paris: Librairie Félix Alcan, 1926), p. 330. The first and fifth editions are identical at this point.

    Google Scholar 

  62. Ibid., trans., p. 262. In the Principia Newton defines a measure of matter which he calls “quantity of matter” (used as a synonym for “body” or “mass”) and which he says is proportional to the volume (or “bulk”) and density. Durkheim seems to use both volume and mass in the sense of volume; cf., e.g., trans., pp. 262, 266, 268, 339.

    Google Scholar 

  63. Ibid., trans., p. 268. Furthermore (p. 270), the “division of labor is… a result of the struggle for existence, but it is a mellowed dénouement. Thanks to it, opponents are not obliged to fight to a finish, but can exist one beside the other. Also, in proportion to its development, it furnishes the means of maintenance and survival to a greater number of individuals who, in more homogeneous societies, would be condemned to extinction.”

    Google Scholar 

  64. Ibid., trans., pp. 256–282.

    Google Scholar 

  65. Ibid., trans., p. 266. The example was taken, with a direct citation, from Darwin’s Origin of Species. Darwin, according to Durkheim, found that “in a small area, opened to immigration, and where, consequently, the conflict of individuals must be acute, there is always to be seen a very great diversity in the species inhabiting it. He found turf three feet by four which had been exposed for long years to the same conditions of life nourishing twenty species of plants belonging to eighteen genera and eight classes. This clearly proves how differentiated they are.” This was offered in proof of Darwin’s observations “that the struggle between two organisms is as active as they are analogous.” Since they have “the same needs” and pursue “the same objects,” they are rivals. Eventually, as their numbers increase, the resources available no longer suffice for all, and a struggle for survival ensues. But, “if the co-existing individuals are of different species or varieties,” they “do not feed in the same manner, and do not lead the same kind of life,” and so they “do not disturb each other.” What is perhaps most remarkable about Durkheim’s argument based on Darwin is the fact that he referred to Darwin at all. It must be kept in mind that at this time, and for many decades afterwards, Darwinian evolution based on natural selection was not regarded with favor by the French scientific establishment.

    Google Scholar 

  66. Durkheim (n. 61 supra, trans.),p. 336; cf. p. 339.

    Google Scholar 

  67. On the uses of organic analogies, see further *1.7 infra. For Theodore Roosevelt, see his Biological Analogies in History (New York: Oxford University Press; London: Henry Frowde, 1910); also Works, vol. 12 (New York: Charles Scribner’s Sons, 1926), pp. 25–60. A. Lawrence Lowell’s organismic views of society may be found in numerous works, notably “An Example from the Evidence of History,” pp. 119–132 of Factors Determining Human Behavior (Cambridge: Harvard University Press, 1937).

    Google Scholar 

  68. Thomas Carlyle: Sartor Resartus, introd. H. D. Traill, The Works of Thomas Carlyle, 30 vols. (London: Chapman and Hall, 1896–1899; reprint, New York: AMS Press, 1969), vol. 1, p. 172. See Frederick W. Roe: The Social Philosophy of Carlyle and Ruskin (New York: Harcourt, Brace & Co., 1921); also David George Hale: The Body Politic: A Political Metaphor in Renaissance English Literature (The Hague/Paris: Mouton, 1971), pp. 134–135.

    Google Scholar 

  69. Thomas Carlyle: Past and Present (London: Chapman and Hall, 1843); Works (n. 68 supra), vol. 10, p. 137; “Chartism,” Works, vol. 29, p. 129.

    Google Scholar 

  70. A brief account of Bluntschli’s life and career by Carl Brinkmann can be found in Encyclopaedia of the Social Sciences, vol. 2 (New York: The Macmillan Co., 1937), p. 606. See also Francis William Coker: Organismic Theories of the State (New York: Columbia University; Longmans, Green & Co., Agents; London: P.S. King & Son, 1910 — Studies in History, Economics and Public Law, vol. 38, no. 2, whole n. 101), pp. 104–114. See, further, J.C. Bluntschli: Denkwürdiges aus meinen Leben, 3 vols. (Nördlingen: C. H. Beck, 1884); also Friedrich Meili: J.C. Bluntschli und seine Bedeutung für die moderne Rechtswissenschaft (Zurich: Drell Füssli, 1908).

    Google Scholar 

  71. Johann Caspar Bluntschli: Lehre vom modernen Staat, 6th ed. (Stuttgart: J.G. Cotta, 1885–1886), the first volume of which was translated into English as Theory of the State (Oxford: Oxford University Press, 1892); Psychologische Studien über Staat und Kirche (Zurich/Frauenfeld: C. Beyel, 1844). Bluntschli was also author of a widely used reference work, Deutsches Staats-Wörterbuch (Stuttgart/Leipzig: Expedition des Staats-Wörterbuchs, 1857–1870).

    Google Scholar 

  72. On Rohmer, see Coker (n. 70 supra), pp. 49–60.

    Google Scholar 

  73. Bluntschli’s discussion of the sixteen psychological functions of the state was pilloried by Charles E. Merriam: “The Present State of the Study of Politics,” The American Political Science Review, 1921, 15: 173–185. Merriam (p. 183) wrote of “Bluntschli’s fearfully and wonderfully made `political psychology,’ in which he compared sixteen selected parts of the human body with the same number or organs in the body politic.”

    Google Scholar 

  74. Psychologische Studien über Staat und Kirche (n. 71 supra), pp. 54, 86–87, cited in translation in Werner Stark: The Fundamental Forms of Social Thought (London: Routledge & Kegan Paul, 1962), pp. 61–62.

    Google Scholar 

  75. Ibid.

    Google Scholar 

  76. Ibid.

    Google Scholar 

  77. On Lilienfeld’s life and career see §1.8 infra and esp. n. 182 infra.

    Google Scholar 

  78. Paul von Lilienfeld: La pathologie sociale (n. 183 infra), p. 59.

    Google Scholar 

  79. Ibid.

    Google Scholar 

  80. Ibid., pp. 59–60.

    Google Scholar 

  81. J. D. Y. Peel: Herbert Spencer: The Evolution of a Sociologist,(New York: Basic Books, 1971); J.W. Burrow: Evolution and Society: A Study in Victorian Social Theory (Cambridge: Cambridge University Press, 1970); David Wiltshire: The Social and Political Thought of Herbert Spencer (Oxford: Oxford University Press, 1978). Of a wholly different sort is the analysis of Spencer in Robert J. Richards: Darwin and the Emergence of Evolutionary Theories of Mind and Behavior (Chicago/London: The University of Chicago Press, 1987). Richards has made a careful study of Spencer’s ideas based on extensive reading and analysis; in particular he has given us a new understanding of Spencer’s social views and biological concepts in relation to the main currents of thought in these areas during Spencer’s lifetime. For an anti-Spencerian point of view, see Derek Freeman: “The Evolutionary Theories of Charles Darwin and Herbert Spencer,” Current Anthropology, 1974, 15: 211–221. See, further, John C. Greene: Science, Ideology and World View: Essays in the History of Evolutionary Ideas (Berkeley/Los Angeles/London: University of California Press, 1981), ch. 4, “Biology and Social Theory in the Nineteenth Century: Auguste Comte and Herbert Spencer”; for a rebuttal, see Ernst Mayr: Toward a New Philosophy of Biology: Observations of An Evolutionist (Cambridge/London: The Belknap Press of Harvard University Press, 1988), essay 15, “The Death of Darwin?”.

    Google Scholar 

  82. Peel (n. 81 supra), p. 178.

    Google Scholar 

  83. Ibid.; see Herbert Spencer: Essays: Scientific, Political, and Speculative, vol. 1 (New York: D. Appleton & Co., 1883), “The Social Organism,” pp. 287–289.

    Google Scholar 

  84. See Peel (n. 81 supra), ch. 7 “The Organic Analogy,” with comparative examples of Spencer’s use of analogies from physics. For the context of Spencer’s analogies, see Richards (n. 81 supra).

    Google Scholar 

  85. Spencer (n. 83 supra), pp. 277–279, 283–286.

    Google Scholar 

  86. Herbert Spencer: Essays Scientific, Political, and Speculative, vol. 3 (New York: D. Appleton & Co., 1896), “Specialized Administration,” pp. 427–428.

    Google Scholar 

  87. René Worms: Organisme et société (Paris: V. Giard & W. Brière, 1896), p. 73.

    Google Scholar 

  88. Walter Cannon: “Relations of Biological and Social Homeostasis,” pp. 305–324 in his The Wisdom of the Body (New York: W. W. Norton & Company, 1932; revised in 1939).

    Google Scholar 

  89. Ibid., pp. 309–310. The significance of the cell theory as a source of analogues for social theory is discussed in §1.8 infra.

    Google Scholar 

  90. Ibid., pp. 312, 314.

    Google Scholar 

  91. See Merton (n. 14 supra), ch. 3, pp. 101n, 102–103.

    Google Scholar 

  92. Walter Cannon: “The Body Physiologic and the Body Politic,” Presidential Address to the American Association for the Advancement of Science, in Science, 1941, 93: 1–10.

    Google Scholar 

  93. Ibid.

    Google Scholar 

  94. Henry C. Carey: The Unity of Law (Philadelphia: Henry Carey Baird, 1872), pp. 116–127; for a derisive critique see Stark (n. 74 supra), pp. 156–160.

    Google Scholar 

  95. See the discussion of the Newtonian style at the end of §1.5 infra and in my article on “Newton & the Social Sciences,” cited in n. 54 supra.

    Google Scholar 

  96. In fact, there are systems of social thought based on models from physics that seem just as ridiculous as those based on biological models, such as Carey’s extravagant electrical analogy (n. 94 supra). Another type of extreme model is set forth in Bradford Peck: The World a Department Store: A Story of Life Under a Cooperative System (Lewiston [Me.]: B. Peck, c1900).

    Google Scholar 

  97. See the valuable discussion of these topics in Claude Ménard: “La machine et le coeur: essai sur les analogies dans le raisonnement économique,” in Analogie et Connaissance, vol. 2: De la poésie à la science (Paris: Maloine éditeur, 1981 — Séminaires Interdisciplinaires du Collège de France), pp. 137–165; also trans. Pamela Cook & Philip Mirowski as “The Machine and the Heart: An Essay on Analogies in Economic Reasoning,” Social Concept, December 1988, 5 (no. 1): 81–95. Especially since the translation omits the mathematical appendix and the discussion, it is well to consult the original.

    Google Scholar 

  98. Stark (n. 74 supra), pp. 73–74.

    Google Scholar 

  99. In this connection we may recall once more Whitehead’s presentation of the “fallacy of misplaced concreteness”; cf. n. 45 supra.

    Google Scholar 

  100. Works on metaphor include Max Black: Models and Metaphors: Studies in Language and Philosophy (Ithaca: Cornell University Press, 1962); Arjo Klamer (ed.): Conversations with Economists (Totowa, [NJ.]: Rowman & Lilienfeld, 1983); Donald N. McCloskey: If You’re So Smart: The Narrative of Economic Expertise (Chicago/London: The University of Chicago Press, 1990); and Andrew Ortony: Metaphor and Thought (Cambridge/ London/New York: Cambridge University Press, 1979). For a brief but incisive history of the uses of metaphor from antiquity to the present, see Mark Johnson (ed.): Philosophical Perspectives on Metaphor (Minneapolis: University of Minnesota Press, 1981), introd. This topic also appears in discussions of economics, notably in Mirowski (n. 23 supra).

    Google Scholar 

  101. Herbert Spencer: The Principles of Sociology, 3rd ed. (New York: D. Appleton and Company, 1897), vol. 1, part 2, §1, “What is a Society?”, §2, “A Society is an Organism.”

    Google Scholar 

  102. See Schlanger (n. 160 infra).

    Google Scholar 

  103. Poetics, 14576, 1459a, 148a.

    Google Scholar 

  104. That is, Aristotle held that analogy was a special kind of metaphor that involves a four-term ratio. Let the ratio be evening: day:: old age: life or evening is to day as old age is to life rom which we obtain old age is the evening of life. Here we have a metaphor in which something (evening) is attributed to something (life) to which it does not belong. The same would be true for evening is the old age of day. Jevons (n. 138 infra), p. 627, gives a similar example, based on a prime minister of a state and a captain of a ship, obtaining the relation that a prime minister is captain of the state.

    Google Scholar 

  105. There are a number of works dealing with rhetoric, especially in relation to the science of the seventeenth century, among them David Johnston: The Rhetoric of Leviathan: Thomas Hobbes and the Politics of Cultural Transformation (Princeton: Princeton University Press, 1986); Alan G. Gross: The Rhetoric of Science (Cambridge/London: Harvard University Press, 1990); Peter Dear (ed.): The Literary Structure of Scientific Argument (Philadelphia: University of Pennsylvania Press, 1991); Steven Shapin & Simon Schaffer: Leviathan and the Air-Pump: Hobbes, Boyle, and the Experimental Life (Princeton: Princeton University Press, 1985); Marcello Pera: Scienza e retorica (Rome/Bari: Laterza, 1991); and M. Pera & William R. Shea (eds.): Persuading Science: The Art of Scientific Rhetoric (Canton, [Mass.]: Science History Publications, USA, 1991). 106 James I: “Speech of 1603,” in Charles H. Mcllwain (ed.): The Political Works of James I (Cambridge: Harvard University Press; London: Humphrey Milford, Oxford University Press, 1918), p. 272; see Hale (n. 68 supra), p. 111.

    Google Scholar 

  106. On the history of the concept of the body politic, see Hale (n. 68 supra).

    Google Scholar 

  107. Ibid.

    Google Scholar 

  108. For James’s statement concerning the spleen, see “Speech in Star Chamber, 1616,” Political Works, p. 343; Hale (n. 68 supra), p. 111, n. 19. See on this subject Marc Bloch: The Royal Touch; Sacred Monarchy and Scrofula in England and France, trans. J. E. Anderson (London: Routledge & K. Paul, 1973).

    Google Scholar 

  109. Cf. §1.5 supra and §§1.7 and 1.8 infra.

    Google Scholar 

  110. See the example of Desaguliers in §10 infra. On this subject see Otto Mayr: Authority, Liberty, & Automatic Machinery in Early Modern Europe (Baltimore: The Johns Hopkins University Press, 1986); John Herman Randall, Jr.: The Making of the Modern Mind: A Survey of the Intellectual Background of the Present Age (Boston: Houghton Mifflin, 1968), ch. 13.

    Google Scholar 

  111. Quoted from Robert S. Hamilton’s Present Status of the Philosophy of Society (1866) in L. L. Bernard & Jessie Bernard: Origins of American Sociology: The Social Science Movement in the United States (New York: Thomas Y. Crowell Company, 1943), p. 711; see p. 265 for a similar quotation concerning “the true PRINCIPIA MATHEMATICA PHILOSOPHIAE SOCIALIST’ Hamilton (ibid., p. 258) believed in two sociological principles, one an analogue of the Copernican system, the other an analogue of Newton’s law of universal gravity; he did not, however, fully understand Newtonian science and wrote of ”centripetal“ and centrifugal” forces as balanced “action” and “reaction.” It is observed by the Bernards that in this respect Hamilton’s law resembles the law of Carey and the law of “cosmic” attraction of Arthur Brisbane. Although Hamilton expressed admiration for Newton, and even held that he himself had propounded Copernican and Newtonian principles of sociology, he also believed that social science might become more nearly an analogue of geology than of sciences such as astronomy, physics, and chemistry. In this regard his opinion was similar to that of R.J. Wright (Bernard & Bernard, p. 306), who held that social science “ought to be compared not with… Chemistry, or Astronomy, or even Moral Philosophy, or Political Economy; but rather with… Geology or Metaphysics.”

    Google Scholar 

  112. The Newtonian style is discussed at length in my Newtonian Revolution (n. 46 supra) and in my article cited in n. 114 infra.

    Google Scholar 

  113. For a more complete discussion of Malthus’s Newtonianism, see my article in Mirowski (n. 54 supra). See, also, Anthony Flew: Thinking about Social Thinking: the Philosophy of the Social Sciences (Oxford: Basil Blackwell, 1985), ch. 4, §1.

    Google Scholar 

  114. Thomas Robert Malthus: An Essay on the Principle of Population as it Affects the Future Improvement of Society (London: printed for J. Johnson, 1798). This work, published anonymously and often known as the “first essay,” is readily available in two reprints, one of which, edited by Antony Flew (Harmondsworth: Penguin Books, 1970), contains also Malthus’s A Summary View of the Principle of Population (London: John Murray, 1830), which was originally published with the author’s name on the title-page. The other, without notes, but with a foreword by Kenneth E. Boulding, is entitled Population: the First Essay (Ann Arbor: The University of Michigan Press, 1959). The text of the second edition (1803) was so completely revised and expanded that it is generally considered “almost a new book,” sometimes referred to as the “second essay.” The text of this version (reprinted from the seventh edition, 1872, but without the appendices) is available as An Essay on the Principle of Population, intro. T.H. Hollingsworth (London: J. M. Dent & Sons, 1914 — Everyman’s Library). On Malthus, see Thomas Robert Malthus: An Essay on the Principle of Population — Text, Sources and Background, Criticism, ed. Philip Appleman (New York/London: W. W. Norton & Company, 1976 — Norton Critical Editions in the History of Ideas).

    Google Scholar 

  115. See Flew (n. 114 supra).

    Google Scholar 

  116. W. Stanley Jevons: The Theory of Political Economy, 2nd ed. (London: Macmillan and Co., 1879), preface; see this preface in later editions, e.g. (New York: Augustus M. Kelley, 1965 — reprint of the fifth edition, 1911), pp. xi—xiv. Jevons was defending himself against the specific charge that in his book “the equations in question continually involve infinitesimal quantities, and yet they are not treated as differential equations usually are, that is integrated” (p. 102). On Jevons’s economics, see Margaret Schabas: A World Ruled by Number: William Stanley Jevons and the Rise of Mathematical Economics (Princeton: Princeton University Press, 1990).

    Google Scholar 

  117. Mary P. Mack: Jeremy Bentham: An Odyssey of Ideas, 1748–1792 (London: Heinemann, 1962), p. 264.

    Google Scholar 

  118. This episode and its significance are discussed in I.B. Cohen: Revolution in Science (Cambridge, The Belknap Press of Harvard University Press, 1985), suppl. §14.1.

    Google Scholar 

  119. See Shapin & Schaffer (n. 105 supra).

    Google Scholar 

  120. See Mirowski (n. 23 supra); !lamer (n. 100 supra); McCloskey (n. 100 supra). 122 Mack (n. 118 supra), pp. 275–281.

    Google Scholar 

  121. Explained in Sigmund Freud: “A Note upon the `Mystic Writing Pad, The Standard Edition of the Complete Psychological Works,vol. 19 (London: The Hogarth Press, 1961), p. 228. Two decades after The Interpretation of Dreams, in Beyond the Pleasure Principle (1920), Freud understood more clearly (as he phrased it in 1924) that ”the inexplicable phenomenon of consciousness arises in the perceptual system instead of permanent traces“ (ibid). See also ed. cit., vol. 5, p. 540, and vol. 18, p. 25; in the latter Freud noted further that this distinction had already been made by Breuer.

    Google Scholar 

  122. Ibid. This pad consisted of a resin or plastic plate, covered with two sheets, one of tissue paper and the other of celluloid, on which one could write with a pointed stylus. Lifting the sheets erased the message, but Freud discovered that the erased message could actually be read in the pad’s “memory.” This mechanical analogue served two functions often found in the use of analogues: (1) to make his earlier hypothetical conjecture seem reasonable enough for him to set forth his ideas in full, and (2) to make his difficult concept of the structure of memory understandable and thus acceptable, to the psychoanalytic community.

    Google Scholar 

  123. Freud: “Civilization and Its Discontents,” Standard Edition (n. 123 supra), vol. 21, p. 144. See Donald M. Kaplan: “The Psychoanalysis of Art: Some Ends, Some Means,” Journal of the American Psychoanalytic Association, 1988, 36: 259–302, esp. 259260.

    Google Scholar 

  124. Brian Vickers: “Analogy versus Identity: The Rejection of Occult Symbolism, 1580–1680,” pp. 95–163 of Brian Vickers (ed.): Occult and Scientific Mentalities in the Renaissance (Cambridge: Cambridge University Press, 1984).

    Google Scholar 

  125. Translated by Vickers from Kepler’s Ad Vitellionem Paralipomena (Gesammelte Werke, vol. 1), p. 90.

    Google Scholar 

  126. Letter to Michael Maestlin, 5 March 1605, quoted in Alexandre Koyré: The Astronomical Revolution: Copernicus — Kepler — Borelli, trans. R. E. W. Maddison (Paris: Hermann; Ithaca: Cornell University Press, 1973), p. 252 (from Kepler’s Gesammelte Werke, vol. 15, pp. 171–172).

    Google Scholar 

  127. These “Rules” appeared in all three editions as part of the introduction of Book Three, “On the System of the World,” but they were called “Rules” only in the second (1713) and third (1726) editions.

    Google Scholar 

  128. Laplace’s System of the World, vol. 2, p. 316, as in Jevons (n. 138 infra), p. 638.

    Google Scholar 

  129. Charles Darwin: The Origin of Species (London: John Murray, 1859; reprint, Cambridge: Harvard University Press, 1964), ch. 3, p. 63. This is a case of analogy rather than of generalization because it extends a property observed in one group of entities (humans) to other groups of different entities (plants and animals), whereas a generalization extends a property of some members of a given class to other (or, even, to all) members of that class, as in the generalization that all men are mortal. Darwin drew on the argument from analogy in other parts of the Origin. The concept of natural selection was introduced in analogy with man’s process of “artificial” selection in breeding pigeons, horses, dogs, and various ornamental and useful plants. A classic use of analogy, as opposed to generalization, occurs in the final chapter of the Origin, in Darwin’s presentation of the theory of “common descent.” He first concluded that all animals had “descended from at most only four or five progenitors, and plants from an equal or lesser number.” This led him to remark, “Analogy would lead me one step further, namely, to the belief that all animals and plants have descended from one prototype.” He was aware, as he wrote, that “analogy may be a deceitful guide.” Yet he found the evidence for common descent to be very persuasive, noting that “all living things have much in common, in their chemical composition, their germinal vesicles, their cellular structure, and their laws of growth and reproduction.” This evidence justified his inference “from analogy that probably all the organic beings that have ever lived on this earth have descended from one primordial form, into which life was first breathed.”

    Google Scholar 

  130. See Nagel (n. 136 infra), pp. 107–110.

    Google Scholar 

  131. James Clerk Maxwell: “On Faraday’s Lines of Force”, in W. D. Niven (ed.): The Scientific Papers of James Clerk Maxwell (Cambridge: Cambridge University Press, 1890; reprint, New York: Dover Publications, 1965), vol. 1, p. 156.

    Google Scholar 

  132. This was the occasion for Maxwell to make what may be considered the classic statement about the use of what he called “physical analogies” in science. According to Maxwell, “physical analogies” provide a means “to obtain physical ideas without adopting a physical theory.” Ernest Nagel (n. 136 infra, p. 109) has explained that Maxwell meant that he could obtain physical ideas without invoking a “theory formulated in terms of some particular model of physical processes.” In other words, by “physical analogies” he implied no more than “that partial similarity between the laws of one science and those of another which makes each of them illustrate the other.”

    Google Scholar 

  133. On this point, see especially articles and books by Mirowski.

    Google Scholar 

  134. Ernest Nagel: The Structure of Science: Problems in the Logic of Scientific Explanation (New York/Burlingame: Harcourt, Brace & World, 1961), pp. 107–117.

    Google Scholar 

  135. See Maxwell (n. 133 supra). See also J. Robert Oppenheimer, “Analogy in Science,” The American Psychologist 1956, 11: 127–135, an address to psychologists in which the physicist J. Robert Oppenheimer stated boldly and unequivocally that “analogy is indeed an indispensable and inevitable tool for scientific progress” (p. 129). He at once narrowed the sense of his assertion, trying to make clear what he meant. “I do not mean metaphor,” he added, “I do not mean allegory; I do not even mean similarity.” Rather, he intended “a special kind of similarity which is the similarity of structure, the similarity of form, a similarity of constellation between two sets of structures, two sets of particulars, that are manifestly very different but have structural parallels.”

    Google Scholar 

  136. W. Stanley Jevons: The Principles of Science: A Treatise on Logic and Scientific Method (2nd and final edition, reprint, New York: Dover Publications, 1958), p. 631.

    Google Scholar 

  137. Ibid.

    Google Scholar 

  138. Ibid., p. 632.

    Google Scholar 

  139. Jevons (1965; n. 117 supra), p. 102. It was even suggested by Jevons (n. 138 supra, p. 633), on the authority of Lacroix, that “the discovery of the Differential Calculus was mainly due to geometrical analogy, because mathematicians, in attempting to treat algebraically the tangent of a curve, were obliged to entertain the notion of infinitely small quantities.” See Schabas (n. 117 supra), pp. 84–88, “Mechanical Analogies.”

    Google Scholar 

  140. Jevons (1965; n. 117 supra), p. 105.

    Google Scholar 

  141. Léon Walras: “Economique et mécanique”, Bulletin de Société Vaudoise des Sciences Naturelles, 1909, 45: 313–325; Mirowski & Cook (n. 48 supra), pp. 189–224. Francis Ysidro Edgeworth proposed the same kind of analogy between his “mathematical psychics” (as he called his brand of economics) and mathematical physics, declaring that “every psychical phenomenon is the concomitant, and in some sense the other side of a physical phenomenon.” He had no doubt that “`Mécanique Sociale’ may one day take her place along with `Mécanique Céleste,’ throned each upon the double-sided height of one maximum principle, the supreme pinnacle of moral as of physical science.” See F. Y. Edgeworth’s Mathematical Psychics: An Essay on the Application of Mathematics to the Moral Sciences (London: C. Kegan & Co., 1881), esp. pp. 9, 12.

    Google Scholar 

  142. See Claude Ménard (n. 97 supra).

    Google Scholar 

  143. Vilfredo Pareto: “On the Economic Phenomenon: A Reply to Benedetto Croce,” translated from Italian by F. Priuli in Alan Peacock, Ralph Turvey, & Elizabeth Henderson (eds.): International Economic Papers, vol. 3 (London: Macmillan and Company, 1953), p. 185. For a discussion of Pareto’s point of view see Mirowski (n. 23 supra), pp. 221–222; also Bruna Ingrao: “Physics and Pareto’s Economics,” to be published in Mirowski (n. 54 supra).

    Google Scholar 

  144. Vilfredo Pareto: Sociological Writings, ed. S. E. Finer, trans. Derick Mirfin (Oxford: Basil Blackwell, 1966), pp. 103–105, selected from Pareto’s Cours d’économic politique (Lausanne, 1898), vol. 2, §§580, 588–590.

    Google Scholar 

  145. Ibid. See also Bruna Ingrao: “L’analogia meccanica nel pensiero di Pareto,” in G. Busino (ed.), Pareto oggi (Bologna: Il Mulino, 1991); and her chapter cited in n. 145 supra.

    Google Scholar 

  146. J.E. Cairnes: The Character and Logical Method of Political Economy (New York: Harper & Bros., 1875) p. 69. See Mirowski (n. 23 supra), p. 198.

    Google Scholar 

  147. Leonard Huxley (ed.): Life and Letters of Thomas Henry Huxley, vol. 1 (London: Macmillan, 1900), p. 218.

    Google Scholar 

  148. See Mirowski (n. 23 supra), pp. 218–219, 287; William Stanley Jevons: The Principles of Economics (London: Macmillan and Co., 1905), p. 50; Jevons: The Theory of Political Economy (1965; n. 117 supra), pp. 61–69; Jevons: The Principles of Science (n. 138 supra), pp. 325–328; Jevons: Papers and Correspondence of William Stanley Jevons, vol. 7, ed. R.D. Collison Black (London: Macmillan, in association with the Royal Economic Society, 1981), p. 80.

    Google Scholar 

  149. Léon Walras: Elements of Pure Economics, trans. William Jaffé (Homewood [Ill.]: Richard D. Irwin; London: George Allen & Unwin; reprint, Philadelphia: Orion Editions, 1984), Preface to the fourth edition, pp. 47–48; also p. 71.

    Google Scholar 

  150. Ménard (n. 97 supra).

    Google Scholar 

  151. Albert Jolink: “`Procrustean Beds and All That’: The Irrelevance of Walras for a Mirowski-Thesis,” to appear in 1993 in a special issue of History of Political Economy, edited by Neil de Marchi, containing papers presented at a symposium (held at Duke University in April 1991) on Mirowski’s More Heat than Light.

    Google Scholar 

  152. Pareto (n. 146 supra), Sociological Writings,p. 104; Cours, vol. 2, §592; see the article by Bruna Ingrao, cited in n. 145 supra.

    Google Scholar 

  153. Mirowski (n. 23 supra), pp. 222–231. Extracts from this manuscript, preserved in the Sterling Library, Yale University, are quoted by Mirowski (pp. 228–229, 409 n. 5).

    Google Scholar 

  154. See, e.g., Hal Varian’s review of Mirowski’s More Heat than Light in the Journal of Economic Literature, 1991, 29: 595–596.

    Google Scholar 

  155. This was part of an article on “Distribution and Exchange” in the Economic Journal for March 1898 and reprinted in A. C. Pigou (ed.): Memorials of Alfred Marshall (London: Macmillan and Co., 1925), pp. 312–318.

    Google Scholar 

  156. Marshall was repeating here the sentiments he had expressed in his inaugural lecture as professor of economics at Cambridge University, printed in Pigou (n. 157 supra; see §1.8 infra).

    Google Scholar 

  157. I have chosen these three organismic sociologists — one Russian, one Austrian, and one French — because their writings exemplify the main issues in the interactions of the natural and the social sciences. There are many others whose writings show the same features, notably the Germany biologist Oscar Hertwig and the Italian sociologist Corrado Gini.

    Google Scholar 

  158. On organismic sociology see F. W. Coker: “Organismic Theories of the State: Nineteenth Century Interpretations of the State as Organism or as Person,” Studies in History, Economics and Public Law (New York: Columbia University, 1910), vol. 38, no. 2, whole number 101; Ludovic Gumplowicz: Geschichte der Staatstheorien (Innsbruck: Universitäts-Verlag Wagner, 1926); Sorokin (n. 55 supra), ch. 4, “Biological Interpretation of Social Phenomena”; Werner Stark (n. 74 supra), part 1, “Society as an Organism”; Judith Schlanger: Les métaphores de l’organisme (Paris: Librairie Philosophique J. Vrin, 1971). Some further major secondary sources on the subject of organismic sociology are: Arnold Ith: Die menschliche Gesellschaft als sozialer Organismus: Die Grundlinien der Gesellschaftslehre Albert Schdffles (Zurich/Leipzig: Verlag von Speidel & Wurzel, 1927); Niklas Luhmann: Die Wirtschaft der Gesellschaft (Frankfurt: Suhrkamp, 1988); N. Luhmann: Die Wissenschaft der Gesellschaft (Frankfurt: Suhrkamp, 1990); D. C. Phillips: “Organicism in the Late 19th and Early 20th Centuries,” Journal of the History of Ideas, 1970, 31: 413–432; E. Scheerer: “Organismus,” pp. 1330–1358 of J. Ritter (ed.): Historisches Wörterbuch der Philosophie (Darmstadt: Wissenschaftliche Verlagsgesellschaft, 1971). Still valuable as sources of information are certain older works, notably Ezra Thayer Towne: Die Auffassung der Gesellschaft als Organismus, ihre Entwicklung und ihre Modifikationen (Halle: Hofbuchdruckerei von C. A. Kammerer & Co., 1903); Erich Kaufmann: Über den Begriff des Organismus in der Staatslehre des 19. Jahrhunderts (Heidelberg: C. Winter, 1908). None of these works, however, pays any attention to the specific relation of these nineteenth-century organismic social scientists to the currents of discovery in the life sciences in their own time.

    Google Scholar 

  159. Alfred Marshall: The Present Position of Economics: An Inaugural Lecture Given in the Senate State House at Cambridge, 24 February, 1885 (London: Macmillan and Co., 1885), pp. 12–14; this lecture is reprinted in Pigou (n. 157 supra), pp. 152–174.

    Google Scholar 

  160. As most people are aware (because of the interest which Sigmund Freud and Josef Breuer had in this subject), hysteria was a major focus of psychiatric attention in the nineteenth century. An example of hysteria has been introduced in §1.4.

    Google Scholar 

  161. As explained in the Preface to this volume, there is no attempt to discuss all aspects of biological science that have interacted with the natural sciences. I have not dealt with the subject of Darwinian evolution because this interaction is far too complex to be considered in a summary fashion and because it is already the subject of a vast literature that is a continuing part of the current Darwin “industry.” Some major aspects of this subject, with special reference to America, are developed in an important way in Robert Richards’ s Darwin and the Emergence of Evolutionary Theories of Mind and Behavior (n. 81 supra), a work that can be especially commended for its methodological approach. Among recent contributions to this general area are Carl N. Degler: In Search of Human Nature: The Decline and Revival of Darwinism in American Social Thought (New York/Oxford: Oxford University Press, 1991), and Dorothy Ross: The Origins of American Social Thought (Cambridge/New York: Cambridge University Press, 1991). Also worthy of mention is Cynthia Eagle Russett: Darwin in America: The Intellectual Response, 1865–1912 (San Francisco: W. H. Freeman, 1976).

    Google Scholar 

  162. Auguste Comte: The Foundations of Sociology,ed. Kenneth Thompson (New York: John Wiley & Sons, 1975), p. 142. The text is taken from the English translation of Auguste Comte’s System of Positive Polity,4 vols (London: Longmans Green, 1877), translated by a group of scholars from Système de politique positive (Paris, 1848–1854), vol. 2, pp. 367–382. Comte believed that Broussais’s principle of continuity was especially important in considering the “opposite” mental states of `reason and madness.“ If the mind surrendered itself to the sense impressions of the external world ”with no due effort of the mind within,“ the result would be ”pure idiocy.“ Madness, in all its intermediate degrees results from the relative failure of the ”apparatus of meditation“ to `correct the suggestions made by the apparatus of observation.” This phenomenon could, he asserted, be studied better in Cervantes’s Don Quixote “than in any treatise of biology.” It could also be traced to “the great principle of Broussais” and could then be “applied to society” as Comte had “now done for the first time.” See Comte’s Cours de philosophie positive (Paris, 1830–1842), quoted in Gertrud Lenzer (ed.): Auguste Comte and Positivism: The Essential Writings (New York/Evanston/San Francisco: Harper & Row, 1975), p. 191, taken from The Positive Philosophy of Auguste Comte, trans. [and condensed by] Harriet Martineau (London: Longmans, Green, 1853), book 5, ch. 6.

    Google Scholar 

  163. In Lenzer (n. 164 supra), p. 191.

    Google Scholar 

  164. See Edmund Beecher Wilson: The Cell in Development and Heredity (New York: The Macmillan Company, 1896; reprint of 3rd ed., New York: The Macmillan Company, 1934), esp. pp. 1–2. Although preliminary steps can be traced to earlier scientists, it was not until the 1840s — largely as a result of the work of J.M. Schleiden and especially Theodor Schwann — that biologists generally began to give cell theory full serious consideration.

    Google Scholar 

  165. In nineteenth-century thought the principle of division of labor was usually credited to Adam Smith, who displayed it in a dramatic fashion in the opening pages of The Wealth of Nations, even though there were other contenders for the invention, including both Benjamin Franklin and Sir William Petty.

    Google Scholar 

  166. This difference is discussed by all organicist sociologists, e.g., Spencer, Lilienfeld, Schäffle.

    Google Scholar 

  167. René Worms called attention to two limitations of this analogy which had been stressed by Herbert Spencer. The first is that, although each individual in the social organism has consciousness, in the animal organism only the organism as a whole, and not the individual cell, has this property. The second: in the social organism the purpose of society, or the organism as a whole is to sustain the lives of the individuals, whereas in the animal or plant the lives of the individual cells serve to support the life of the organism as a whole. Despite these dissimilarities, the cell theory seemed to provide nature’s own model on the microscopic scale for the study of human societies, much as the social behavior of ants has done in our own days.

    Google Scholar 

  168. As in societies, the development of the embryo produces special cells and groups of cells with forms and structures suited to their function. This concept of “division of labor,” as we have seen (n. 61 supra), originated in social science, then was transferred to the life science and finally migrated back to the social sciences. This transfer is the subject of chapter 10 infra.

    Google Scholar 

  169. On von Baer see the article by Jane Oppenheimer in the Dictionary of Scientific Biography, vol. 1 (New York: Charles Scribner’s Son, 1970), pp. 385–389.

    Google Scholar 

  170. See Steven J. Gould: Ontogeny and Phylogeny (Cambridge: The Belknap Press of Harvard University Press, 1977).

    Google Scholar 

  171. In this section I have not dealt particularly with Herbert Spencer, although he is perhaps the most important of all the organicist social scientists. One reason is that, unlike many other organic sociologists, he did not concentrate attention on bio-medical discoveries relating to the cell theory, although he did make use of cell biology in his writings on sociology. Some of Spencer’s uses of biological science in relation to sociology are discussed in §1.4 supra and §1.8 infra. On Spencer’s use of analogies, see ch. 8 infra. On the subject of Spencer and sociology, see Richards (n. 81 supra). Cf. also n. 208 infra.

    Google Scholar 

  172. Herbert Spencer: First Principles (London: Williams and Norgate, 1862), §119. Spencer evidently learned this law from William Carpenter; see Richards (n. 81 supra), p. 269. Richards observes that Carpenter thought that von Baer’s law (that “a heterogeneous structure arises out of one more homogeneous”) had great generality. Carpenter wrote that “if we watch the progress of evolution [i.e., embryonic development], we may trace a correspondence between that of the germ in its advance towards maturity, and that exhibited by the permanent condition of the races occupying different parts of the ascending scale of creation.”

    Google Scholar 

  173. Herbert Spencer, “Reasons for Dissenting from the Philosophy of M. Comte,” Essays: Scientific, Political,and Speculative, vol. 2 (New York: D. Appleton and Company, 1896), pp. 118–144.

    Google Scholar 

  174. Ibid., pp. 137–138. It should be noted also that cellular embryology reinforced another principle of organismic sociology. Embryologists revealed that as an individual organism progresses through more and more complex forms, the component cells exhibit structures suitably adapted for their special function, that is, they show the form necessary for the “division of labor.” Spencer held that even before encountering von Baer’s “law,” he had begun to conceive of both “the development of an individual organism and the development of the social organism” as an advance from “independent parts to mutually-dependent unlike parts — a parallelism implied by Milne-Edwards’ doctrine of the `physiological division of labor.”’

    Google Scholar 

  175. For Virchow, the concept of the “cell state” was particularly significant because there was always a close parallel between his “biological views and his liberal political opinions.” See Owsei Temkin: “Metaphors of Human Biology,” in Robert C. Stauffer (ed.): Science and Civilization (Madison: University of Wisconsin Press, 1949), p. 172. Temkin is summarizing Ernst Hirschfeld, “Virchow,” Kyklos: Jahrbuch des Instituts für Geschichte der Medizin an der Universität Leipzig, 1929, 2: 106–116. See also Erwin H. Acherknecht: Rudolf Virchow: Doctor, Statesman,Anthropologist (Madison: The University of Wisconsin Press, 1953); reprint (New York: Arno Press, 1981).

    Google Scholar 

  176. Rudolf Virchow: Cellular Pathology As Based upon Physiological and Pathological Histology, trans. Frank Chance (New York: Robert M. DeWitt, 1860), p. 40; also (London: John Churchill), pp. 13–14.

    Google Scholar 

  177. Virchow, we may note, was not the only nineteenth-century biologist to use social analogies in scientific discourse. Thomas Henry Huxley made use of a social analogy in describing the sponge, which — he said — represented a kind of sub-aqueous city, “in which the people are arranged about the streets and roads, in such a manner, that each can easily appropriate his food from the water as its passes along.” This is an example of the use of analogy to illustrate a scientific concept, making such a concept easier to visualize or to understand.

    Google Scholar 

  178. See Temkin (n. 177 supra), p. 175.

    Google Scholar 

  179. Ibid.

    Google Scholar 

  180. Paul von Lilienfeld, or Paul de Lilienfeld, or Pavel Fedorovich Lilienfeld-Toailles, or Pavel Fedorovich Lilienfel’d Toal’ (1829–1903), was a Russian civil servant who held responsible government posts and whose avocation was sociology. He published a book in Russian on the elements of political economy in 1860 under the pseudonym “Lileyewa.” Another work, appearing first in 1872 in Russian, under the initials P. L., bore the title, Thoughts on the Social Science of the Future, which was expanded into a five-volume German version, Gedanken über die Socialwissenschaft der Zukunft (vols. 1–4: Mitau: E. Behre’s Verlag, 1873–1879; vol. 5: Hamburg: Gebr. Behre’s Verlag; Mitan: E. Behre’s Verlag, 1881). Of particular importance are La pathologie sociale (Paris: V. Giard & E. Brière, 1896) and Zur Vertheidigung der organischen Methode in der Sociologie (Berlin: Druck und Verlag von Georg Reimer, 1898). In 1897–1898 Lilienfeld was president of the Institut International de Sociologie. See Otto Henne am Rhyn: Paul von Lilienfeld (Gdansk, Leipzig, Vienna: Carl Hinstorff’s Verlagsbuchhandlung [n.d.] — Deutsche Denker und ihre Geistesschöpfungen, ed. Adolf Hinrichsen, vol. 6). For further bibliography related to Lilienfeld, see Howard Becker: “Lilienfeld-Toailles, Pavel Fedorovich,” Encyclopaedia of the Social Sciences, vol. 9 (New York: The Macmillan Company, 1933, 1937), p. 474. See also n. 160 supra for a list of publications relating to organismic sociology. (The first four volumes of the Gedanken in one of the sets in the Harvard University Library contain book plates indicating that they were “bought with the income from the bequest of James Walker… former president of Harvard College; ‘preference being given to works in the intellectual and moral sciences.’”)

    Google Scholar 

  181. Trans. from Gedanken, vol. 1, p. V.

    Google Scholar 

  182. Trans. from Pathologie, p. xxii.

    Google Scholar 

  183. Ibid.

    Google Scholar 

  184. Ibid., p. 8.

    Google Scholar 

  185. Ibid., pp. 8–11. Lilienfeld was noted in his own time for his discussion of social diseases that were analogues of diseases of the nervous system, particularly psychological disorders. We have seen (§1.4 supra) an example of his suggestion of a parallel between medical and social disorders in the social analogue of the intellectual and moral state of women suffering from hysteria.

    Google Scholar 

  186. Ibid., pp. 20–21.

    Google Scholar 

  187. Ibid., p. 21.

    Google Scholar 

  188. Ibid., p. 24.

    Google Scholar 

  189. Ibid., pp. 46–47.

    Google Scholar 

  190. Ibid., p. 307.

    Google Scholar 

  191. Bau und Leben des socialen Körpers: Encyclopädischer Entwurf einer realen Anatomie, Physiologie und Psychologie der menschlichen Gesellschaft mit besonderer Rücksicht auf die Volkswirthschaft als socialen Stoffwechsel, 4 vols. (Tübingen: H. Laupp’sche Buchhandlung, 1875–1878). Albert Eberhard Friedrich Schäffle (1831–1903), a German sociologist and economist, was a professor at the University of Tübingen, later moving to the University of Vienna. He was, for a while, a member of the Austrian cabinet. He edited a journal entitled Zeitschift für die Gesamte Staatswissenschaft. He envisioned a “rational social state,” a kind of utopian blend of capitalism and socialism. He was known in his own times primarily for his exposition of organismic social theory, especially his use of specific biological analogies. See the article on him by Fritz Karl Mann in Encyclopaedia of the Social Sciences, ed. Edwin R.A. Seligman (New York: The Macmillan Company, 1934), vol. 13, pp. 562–563. There is no biography of Schäffle in the more recent International Encyclopedia of the Social Sciences. See Arnold Ith (n. 160 supra) and Stark (n. 74 supra), pp. 62–72.

    Google Scholar 

  192. Schäffle (n. 193 supra) vol. 1, p. 286; see Stark (n. 74 supra), p. 63. The extracts from Schäffle are quoted from Stark’s translation.

    Google Scholar 

  193. Schäffle, vol. 1, p. 286; Stark, pp. 63–64.

    Google Scholar 

  194. Preface to Lilienfeld’s La pathologie sociale (n. 182 supra), p. vii; cf. Stark, p. 63.

    Google Scholar 

  195. Schäffle, vol. 1, p. 286; Stark, p. 64.

    Google Scholar 

  196. Schäffle, vol. 1, p. 33; Stark, p. 66.

    Google Scholar 

  197. Schäffle, vol. 1, p. 57; Stark, p. 67.

    Google Scholar 

  198. Schäffle, vol. 1, p. 324; Stark, p. 67.

    Google Scholar 

  199. Schäffle, vol. 1, p. 335; Stark, p. 67.

    Google Scholar 

  200. Schäffle, vol. 1, pp. 327, 329; Stark, p. 68.

    Google Scholar 

  201. Schäffle, vol. 1, p. 94; Stark, p. 68.

    Google Scholar 

  202. René Worms (1869–1926), a French sociologist, was educated at the Ecole Normale Supérieure. In 1893 he founded both the Paris-based Institut International de Sociologie and the Revue Internationale de Sociologie. He also founded and edited a series of fifty books on sociological subjects by authors from many countries. He was known in his lifetime particularly for his views concerning the interrelations among “the three disciplines of psychology, social psychology, and sociology.” See the biography and critical analysis by Terry N. Clark in International Encyclopedia of the Social Sciences, ed. David L. Sills, vol. 16, pp. 579–581 (New York: The Macmillan Company & The Free Press, 1968). An account of the life and career of Rent Worms may be found in an article by V.D. Sewny in the Encyclopaedia of the Social Sciences, vol. 15 (New York: The Macmillan Company, 1934), pp. 498–499. See also Stark (n. 74 supra).

    Google Scholar 

  203. Worms (n. 87 supra), p. 43.

    Google Scholar 

  204. René Worms: Philosophie des sciences sociales (Paris: V. Giard & E. Brière, 1903), vol. 1, p. 53.

    Google Scholar 

  205. Ibid., chs. 2, 3.

    Google Scholar 

  206. See, especially, Derek Freeman, “The Evolutionary Theories of Charles Darwin and Herbert Spencer,” Current Anthropology, 1974, 15: 211–237. Cf. also n. 173 supra. 209 Nature, 1982, 296: 686–687.

    Google Scholar 

  207. J.W. Burrow: Evolution and Society: A Study in Victorian Social Theory (Cambridge: The University Of Cambridge Press, 1970), p. 182.

    Google Scholar 

  208. Peel (n. 81 supra), p., 174, including a quotation from Spencer’s Social Statics.

    Google Scholar 

  209. The organic metaphors predominate in many essays (notably “The Social Organism” [1860]) and in his books, especially Social Statics (1850), The Study of Sociology (1873), and The Principles of Sociology (1876). See Peel (n. 81 supra), ch. 7, esp. p. 174.

    Google Scholar 

  210. Quoted in Peel (n. 81 supra), p. 179.

    Google Scholar 

  211. Ibid., p. 178.

    Google Scholar 

  212. I have not felt the need to make a parade here of the mismatched homologies that appear in the writings of Lilienfeld, Schäffle, Worms, and Spencer (see §1.4 supra), because my aim has been to examine the historical use of analogies rather than merely to call attention to their extravances (as has been done in §1.4 supra).

    Google Scholar 

  213. Michel Foucault: Power/Knowledge: Selected Interviews and Other Writings, 1972–1977, ed. Colin Gordon (Brighton: Harvester Press, 1980), p. 151.

    Google Scholar 

  214. See Small and Vincent (n. 18 supra).

    Google Scholar 

  215. Bryan S. Turner: The Body and Society: Explorations in Social Theory (Oxford: Basil Blackwell, 1984), pp. 49–50; Louis Wirth: “Clinical Sociology,” American Journal of Sociology, 1931, 37: 49–66.

    Google Scholar 

  216. L.J. Henderson: “Physician and Patient as a Social System,” New England Journal of Medicine, 1935, 51: 819–823; “The Practice of Medicine as Applied Sociology,” Transactions of the Association of American Physicians, 1936, 51: 8–15. These and other papers of Henderson on similar subjects have been edited with an important introductory statement by Bernard Barber: L.J. Henderson on the Social System: Selected Writings (Chicago: University of Chicago Press, 1970). See, on this subject, Talcott Parsons: The Social System (Glencoe, Ill.: Free Press, 1951) and “The Sick Role and the Role of the Physician Reconsidered,” Milbank Memorial Fund Quarterly, 1975, 53: 257–278.

    Google Scholar 

  217. Marie-Jean-Antoine-Nicolas Caritat, Marquis de Concorcet: Esquisse d’un tableau historique des progrès de l’esprit humain (Paris: Agasse, 1795); also Baker (n. 29 supra), pp. 348–349, 368–369.

    Google Scholar 

  218. In later editions Malthus, attempting to lessen the gloomy prospect he had set forth, introduced the power of “moral restraint” as a factor in population control. See n. 115 supra.

    Google Scholar 

  219. David Hume: “Of the Populousness of Ancient Nations,” vol. 1 of his Essays,Moral, Political, and Literary (Edinburgh: R. Fleming and A. Alison for A. Kincaid, 1742), p. 376. See Catherine Gallagher: “The Body versus the Social Body,” pp. 83–106 of Catherine Gallagher & Thomas Laqueur (eds.): The Making of the Modern Body: Sexuality and Society in the Nineteenth Century (Berkeley/Los Angeles: University of California Press, 1987).

    Google Scholar 

  220. Carey (n. 43 supra); see §1.4 supra.

    Google Scholar 

  221. See Sorokin (n. 55 supra) and Stark (n. 74 supra).

    Google Scholar 

  222. The Spirit of the Laws, trans. Thomas Nugent (revised ed., London: George Bell and Sons, 1878; reprint, New York: Hafner Press, 1949), bk. 3, §7, “The Principle of Monarchy.”

    Google Scholar 

  223. On this score see Henry Guerlac: “Three Eighteenth-Century Social Philosophers: Scientific Influences on their Thought,” Daedalus, 1958, 87: 6–24; reprinted in Henry Guerlac: Essays and Papers in the History of Modern Science (Baltimore: The Johns Hopkins University Press, 1977), pp. 451–464.

    Google Scholar 

  224. Adam Smith: An Inquiry into the Nature and Causes of the Wealth of Nations (Oxford: Oxford University Press, 1976 — The Glasgow Edition of the Works and Correspondence of Adam Smith, II), bk. 1, ch. 7, p. 15 (§15). The “Cannan edition” — Adam Smith: An Inquiry into the Causes of the Wealth of Nations, ed. Edwin Cannan (London: Methuen & Co., 1904; reprint, Chicago: The University of Chicago Press, 1976; reprint New York: Modern Library, 1985) — is easier to read and has the advantage of useful postils. A postil (1976 ed., p. 65; 1985 ed., p. 59) repeats the message: “Natural price is the central price to which actual prices gravitate.”

    Google Scholar 

  225. Adam Smith: Essays on Philosophical Subjects, ed. W.P.D. Wightman & J.C. Bryce (Oxford: Oxford University Press, 1980 — The Glasgow Edition of the Works and Correspondence of Adam Smith), vol. 3, pp. 33–105, “The History of Astronomy.”

    Google Scholar 

  226. Ménard (n. 97 supra; 1988).

    Google Scholar 

  227. For details see my Introduction to Newton’s `Principla“’ (Cambridge: Harvard University Press; Cambridge: Cambridge University Press, 1971), ch. 2, §1. Newton referred to the inertial property of bodies as both a ”vis inertiae“ or ”force of inertia“ and ”inertia.“ For him this was an ”internal“ rather than an ”external“ force and so could not — of and by itself — alter a body’s state of rest or of motion.

    Google Scholar 

  228. On Darwin and Lyell see Mayr (n. 40 supra); the details of Darwin’s transformation are discussed, along with other examples, in my Newtonian Revolution: With Illustrations of the Transformation of Scientific Ideas (New York/Cambridge: Cambridge University Press, 1980). This topic is explored also in my forthcoming Scientific Ideas (New York: W. W. Norton & Company, 1994).

    Google Scholar 

  229. Mirowski (n. 23 supra, pp. 241–254) has documented the way in which Joseph Bertrand and Hermann Laurent faulted Walras for his mathematical physics, as Laurent and Vito Volterra later faulted Pareto.

    Google Scholar 

  230. Journal of Economic Literature, 1991, 29: 595–596.

    Google Scholar 

  231. But even a severe critic like Hal R. Varian does admit that Mirowski’s “thorough search of the writings of Walras, Jevons, Fisher, Pareto, and other neoclassicals… has established, to almost anyone’s satisfaction that they recognized that `utility’ had some features in common with the then-current notions of `energy’.”

    Google Scholar 

  232. A. Lawrence Lowell: “An Example from the Evidence of History,” in Harvard Tercentenary Conference of Arts and Sciences (1936): Factors Determining Human Behavior (Cambridge: Harvard University Press, 1937), pp. 119–132.

    Google Scholar 

  233. Of course, one reason why an analogy may be inappropriate is that it is based on mismatched homology. Another reason might be that the analogy did not advance the subject to the same degree as a rival one.

    Google Scholar 

  234. Henry Guerlac once described it as one of the worst in the English language. Jean T. Desaguliers: The Newtonian System of the World, the Best Model of Government (Westminster: A. Campbell, 1728).

    Google Scholar 

  235. John Craig: Theologiae Christianae Principia Mathematica (London: impensis Tomothei Child, 1699). A translation of some major extracts by Anne Whitman is published (without the translator’s name) as “Craig’s Rules of Historical Evidence,” History and Theory: Studies in the Philosophy of History, Beiheft 4 (The Hague: Mouton, 1964). Craig once suggested to Newton a minor modification of the Principia; see I. B. Cohen: “Isaac Newton, the Calculus of Variations, and the Design of Ships,” pp. 169–187 of Robert S. Cohen, J.J. Stachel, & M.M. Wartofsky (eds.): For Dirk Struik: Scientific, Historical, and Political Essays in Honor of Dirk J. Struik (Dordrecht/Boston: D. Reidel Publishing Company, 1974 — Boston Studies in the Philosophy of Science, vol. 15).

    Google Scholar 

  236. For two centuries and more, Craig’s book and its Newton-like laws have usually been presented as an example of the kind of aberration to which Newtonian science may lead. His whole book can, in fact, be considered an extended example of inappropriate analogy. Yet a recent study by Stephen Stigler (“John Craig and the Probability of History: From the Death of Christ to the Birth of Laplace,” Journal of the American Statistical Association, 1986, 81: 879–887) has shown that Craig made a serious contribution to applied probability, “that his formula for the probability of testimony was tantamount to a logistic model for the posterior odds.”

    Google Scholar 

  237. I have not attempted to rewrite the history of this subject, displayed in many monographs, beginning with Richard Hofstadter: Social Darwinism in American Thought (Philadelphia: University of Pennsylvania Press, 1944; rev. ed., Boston: Beacon Press, 1955). Some more recent works are Degler (n. 163 supra) and Robert C. Bannister: Social Darwinism: Science and Myth in Anglo-American Social Thought (Philadelphia: Temple University Press, 1979); Howard L. Kaye: The Social Meaning of Modern Biology: From Social Darwinism to Social Biology (New Haven: Yale University Press, 1984); Peter J. Bower: The Eclipse of Darwinism: Anti-Darwinian Evolution Theories in the Decades around 1900 (Baltimore: Johns Hopkins University Press, 1983).

    Google Scholar 

  238. Michael Ruse: “Social Darwinism: Two Roots,” Albion, 1980, 12: 23–36.

    Google Scholar 

  239. Spencer: “The Study of Sociology,” No. XVI, “Conclusion,” Contemporary Review, 1873, 22: 663–677, esp. p. 676.

    Google Scholar 

  240. Stephen Jay Gould: “Shoemaker and Morning Star: A Visit to the Great Reminder reveals some Painful Truths carved in Stone,” Natural History, December 1990, pp. 14–20, esp. p. 20.

    Google Scholar 

  241. Gould’s analogy rests on an imperfect homology. Lamarckian evolution in biology implies not only that each individual may modify his or her inheritance but that such modifications are transmitted to one’s offspring. Consider a catastrophe in which all material culture and all humans over the age of three would be destroyed. In a Lamarckian social world homologous with a Lamarckian biological world, the surviving individuals would have inherited the technological knowledge and skills acquired by centuries of evolutionary development. In the world of nature and of man, however, this would not be the case, as Gould is aware.

    Google Scholar 

  242. Additionally Gould alleges that the Lamarckian mode “of cultural transmission” is responsible for “all the ills of our current environment crisis” as well as “the joys of our confidently growing children.”

    Google Scholar 

  243. This example was brought to my attention by Neil Niman at a symposium on Natural Images in Economics. He, however, treats this episode in a wholly different way from mine. See his paper in Mirowski (n. 54 supra).

    Google Scholar 

  244. Armen A. Alchian: “Uncertainty, Evolution and Economic Theory,” Journal of Political Economy, 1950, 57: 211–221.

    Google Scholar 

  245. Edith Tilton Penrose: “Biological Analogies in the Theory of the Firm,” The American Economic Review, 1952, 42: 804–819, esp. p. 805.

    Google Scholar 

  246. Ibid., p. 807.

    Google Scholar 

  247. Ibid., p. 812.

    Google Scholar 

  248. Armen A. Alchian: “Biological Analogies in the Firm: Comment,” The American Economic Review, 1953, 43: 600–603.

    Google Scholar 

  249. Edith T. Penrose: “Rejoinder,” ibid., pp. 603–609. Penrose quotes from Alchian’s original article to the effect that the “suggested approach embodies the principles of biological evolution and natural selection.”

    Google Scholar 

  250. William F. Ogburn: Social Change with Respect to Culture and Original Nature, 2d ed. (New York: Viking Press, 1950), Supplement.

    Google Scholar 

  251. Ménard (n. 97 supra; 1988), p. 91.

    Google Scholar 

  252. Equality of Educational Opportunity, 2 vols. (Washington, D.C.: Office of Education — U.S. Department of Health, Education, and Welfare — U.S. Government Printing Office, 1966).

    Google Scholar 

  253. Quoted in Frederick Mosteller & Daniel P. Moynihan (eds.): On Equality of Educational Opportunities: Papers Deriving from the Harvard University Faculty Seminar on the Coleman Report (New York: Random House, 1972), pp. 4–5.

    Google Scholar 

  254. See the editors’ discussion of crude and refined statistics (ibid., pp. 12–14) and also ch. 11 by Christopher S. Jencks on “The Quality of the Data Collected by The Equality of Educational Opportunity Survey.” The second volume of the Coleman Report consisted of 548 pages of tables of means, standard deviations, and correlation coefficients, as a complement to the 373 pages of the first volume.

    Google Scholar 

  255. In Mosteller & Moynihan (n. 256 supra), p. 33, there is a critique of the statistics and their interpretation. Chapter Four, by James S. Coleman, is on “The Evaluation of Equality of Educational Opportunity.”

    Google Scholar 

  256. Mosteller & Moynihan (n. 256 supra), p. 32.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Editor information

Editors and Affiliations

Rights and permissions

Reprints and permissions

Copyright information

© 1994 Springer Science+Business Media Dordrecht

About this chapter

Cite this chapter

Cohen, I.B. (1994). An Analysis of Interactions between the Natural Sciences and the Social Sciences. In: Cohen, I.B. (eds) The Natural Sciences and the Social Sciences. Boston Studies in the Philosophy of Science, vol 150. Springer, Dordrecht. https://doi.org/10.1007/978-94-017-3391-5_1

Download citation

  • DOI: https://doi.org/10.1007/978-94-017-3391-5_1

  • Publisher Name: Springer, Dordrecht

  • Print ISBN: 978-90-481-4258-3

  • Online ISBN: 978-94-017-3391-5

  • eBook Packages: Springer Book Archive

Publish with us

Policies and ethics