“Science” in The “Industrial Revolution”

Michael's personal logoThere is a wealth of literature that purports to deal with the historical relationship between science and technology, or between science and the so-called “Industrial Revolution,” but precious little of the material which is suggestively titled in these terms does more than make a general affirmation or negation of the role played by science in the industrialization of Britain during the century from 1750–1850. It is but a small circle of scholars who have ventured to specify those instances in which “science” made identifiable contributions to the industrial technology of this period, and even such modest claims as these individuals have made have not gone unchallenged by criticism of the use of the word science in this particular context. Evidently, difficulties with the definition of science, and limited opportunities to apply even a liberal interpretation of this term to the origins of early British industry, have discouraged full development of a theme which seems so promising bibliographically but which receives little specific discussion.

Such claims as are made for a scientifically-based technology in late eighteenth- and early nineteenth-century Britain focus upon two contemporary developments: the primitive steam engine and the nascent chemical industry. The first innovation, which is identified with the names of the inventors Savery, Newcomen, and Watt, is held to be the material result of proposals made by the Dutch scientist, Huygens, and his assistant Denis Papin for the creation of a piston engine fired by gunpowder and, later, for a simple engine propelled by steam. Though Huygens’ 1680 gunpowder proposal was completely impractical, Papin’s steam engine idea of a decade later was translated into working hardware; thus it may be said that the earlier researches of Torricelli and Robert Boyle into atmospheric pressure and vacuum had been translated into a new technology. However, it is not a certainty that the work of these scientists was, in fact, transmitted to the inventors of later decades, for the historians Musson and Robinson can claim only that, with regard to these early scientific developments, "there is a strong probability that Savery and Newcomen may have acquired knowledge of them through “the Royal Society.” Documentary confirmation of the linkage is lacking, because Savery and Newcomen, like virtually all other contemporary inventors of industrial machinery, were scientifically untutored craftsmen who were not part of a community of scholars.

The reason that the discoveries of the scientists, known then as “natural philosophers,” were not a vital part of progress in industrial processes was that scientific philosophy had been correctly reduced to quantitative formulae only with the work of Sir Issac Newton in 1687. “Sciences” other than Newton’s elementary mechanics and astronomy remained merely qualitative, and consisted mainly of “discoveries” such as inventors themselves might make. Since even the most simple contemporary industrial machines and processes involved a multiplicity of little-understood principles (heat, friction, expansion, stress, etc.), science, in the sense of formulae which could be used to dictate the design parameters of new equipment, was able to make little or no contribution to the industrial advancement of Britain before the second half of the nineteenth century. Progress had to be made by accident and by trial and error, a process which was common to the technological contributions both of scientists and of untutored inventors.

The proponents of the view that science made a substantial contribution to early British industry take the latter observation and substitute “educated industrialists” for “untutored inventors.”1 Thus a “scientific” influence upon industrial innovation, as opposed to the necessity for “craft” methods in the advancement of science, is deduced from the alleged refusal by eighteenth-century scientists if to invalidate as ‘natural philosophy’ many of the experiments carried 2 out by their industrialist friends The attempt by opponents of the scientific influence view to distinguish between this “crude empiricism” of ad hoc experimentation and the authentic application of “science” through use of mathematical formulae, is countered by proponents through reference to the principle of reliance upon the empirical experimental method for progress in every scientific discipline. However, the proponents , such as Musson and Robinson, fail to distinguish between those experiments which uncover crude but immediately useful industrial techniques and those which are directed toward discovery of more general principles and which are quantitatively formulated before they are applied to practical concerns. Those who recognize the distinction between such experiments thereby refuse to see the application of science, proper, in the development of industrial innovations in the century under discussion.

The problems of definition and interpretation indicated by the foregoing controversy are illustrated by the characterization of discoveries leading to creation of the British chemical industry. The best case for the scientific origin of a chemical process is that made in behalf of chlorine bleaching, which was first used on a large scale in the Glasgow linen industry. The discovery of chlorine is attributed to the Swedish chemist, Scheele, who as it turns out, was mistaken both in his advocacy of the general physicochemical theory of “phlogiston” and in his theoretical understanding of the process by which chlorine performed its bleaching action. However, according to Musson and Robinson:

The fact that many years elapsed before the formulation of a really accurate theoretical explanation of this process does not mean that the researches of chemists such as Scheele and Berthollet, and the new bleaching methods based upon them, were ‘unscientific.’ These men were among the most outstanding natural philosophers or scientists of the day, and...their methods were applied in industry by men who were generally well versed in theoretical and practical chemistry. Undoubtedly experiment played a tremendously important part, but it was not blind empiricism: practice and theory progressed together, stimulating each other. 3

Here again is the conceptual distinction between a process obtained through theoretical knowledge and one which is stumbled upon by accident. Musson and Robinson suggest chlorine was obtained by a combination of the two, but they fail to specify how the chemical theory stimulated the chlorine practice. In fact, Scheele’s investigations had not been directed toward discovery of a cloth-bleaching agent but toward an understanding of the glass-coloring properties of manganese, from which chlorine is, in part, derived. Scheele could not have been expected to discover something useful by deliberate pursuit under the guidance of contemporary chemical theory, for, as just mentioned, his own theory was an incorrect hypothesis, and even the best theory of the day was far too primitive to provide firm technological guidance. Chemistry had taken its first, halting steps toward quantification under Lavoisier only in the last quarter of the eighteenth century, when a small number of categories of elemental compounds were recognized (acids, bases, salts) and the first simple laws of chemical combination were expounded.

Recognition of this factor is implicit in the passage quoted above, for the authors first emphasize the fact that the discovery of chlorine was made by a scientist or scientists and only secondarily that scientific theory had something to do with the innovation. In fact, the great bulk of the argument in behalf of a scientific contribution to early British industrialization is based upon the mere identification of discoveries with persons exposed to science; thus the conceptual distinction between applied science and inventive tinkering by such persons is ignored, and a contribution which might as well have been made by an uneducated experimenter as by a leading natural philosopher is declared “scientific” by virtue of its attribution to the latter. In line with this argument, Musson and Robinson,4 supported by Robert Schofield , have devoted most of their scholarship to the network of interpersonal associations and intellectual discourse which purportedly formed an atmosphere of “scientificalness” in Britain at the time. Consequently, the reader is invited to see the intellectual inspiration of the scientific Zeitgeist at work permeating the society at large and so ultimately making its most fundamental, if somewhat tenuous, contribution to the advancement of industrial technology.

At this point, the various definitions of the term “science” may be summarized in order to make an appropriate selection for purposes of further historical commentary. The modern definition of science is the one which demands algebraic formulae offering design parameters and predictions of the performance of industrial techniques and which thereby finds virtually no expression in the late eighteenth and early nineteenth-century industrial technologies. Another definition, resulting perhaps from popular confusion, is that which equates science with technology and so allows one to find the application of science throughout human technological history. A third definition, and that upon which the proponents of the scientific influence view are inadvertently standing, is that which labels as “scientific” those contributions made by scientifically knowledgeable persons. This last definition offers the historian some middle ground between the extremes of denying science any credit for early industrial evolution and indiscriminately crediting it with every innovation. If this definition and its limitations are kept in mind, then a more explicit discussion of the role of science in industry is permissible.

The creation of the steam engine and the rise of the chemical industry in Britain have been briefly mentioned for illustrative purposes. There are further details of their progress which may be viewed as scientifically inspired according to the definition now in hand. The development of the steam engine beyond the crude device built by Savery and Newcomen was the famous achievement of James Watt, who was a man of prodigious intellect and scientific training but who was nevertheless a man of practical engineering interests. Interestingly, it is this man, who is not conventionally identified as a scientist or natural philosopher, who possibly came closest to fulfilling the rigorous demands of the modern definition of science in making his own contribution to industrial technology.

Seventeen sixty-nine was a watershed in the history of modern industry, for patents were registered in the winter and spring of that year in behalf of Watt’s steam engine and Arkwright’s water frame. Watt’s addition of a condensing chamber to his engine was a fundamental modification of the device, based upon his own tabulated calculations of thermal efficiency and, according to some sources, upon more abstract notions of “latent heat” communicated to him by he scientist Joseph Black. If the latter is the case, as it is according to J.D. Bernal,5 then Watt’s condenser was the product of a general theoretical conception, such as that of the power of a vacuum which first inspired the steam engine itself. However, D.S.L. Cardwell 6 quotes Watt himself to the effect that Black’s communications were interesting and informative but were not, and theoretically could not be, the origin of his idea for the patented condenser. So credit for improvement in the steam engine lay not with abstract theory and applied science, if Caldwell is right; rather the modification was the result of an insight, focused upon a particular technological problem, with little alternate application.

Despite the weakness of the case for considering Watt’s contribution as scientific in modern terms, his work remains scientific in the terms adopted here, particularly in view of Watt’s association with the members of the Royal Society, of which he was a-Fellow. The case for science in British industry rests heavily upon other Fellows, as well, for the participation of early industrialists in the Royal Society is a major emphasis of the argument by identification. Watt’s business partner, Matthew Bolton, was a Birmingham manufacturer and a fellow Fellow, a distinction which was shared by Josiah Wedgewood and James Keir. Wedgewood was a renowned pottery manufacturer who employed chemists in his factory – among them Alexander Chisholm – to explore the properties of pottery components (clays, glazes) and the techniques of their utilization. Keir resembled Wedgewood in his use of chemical knowledge (he was himself a chemist) and was an unusual combination of scientific author and glass and soap manufacturer. In combination with yet another Fellow of the Royal Society, Dr. John Roebuck, Keir, Watt, and Joseph Black developed the first synthetic soda production technique; however, the joint work of Watt and Black resulted in an unprofitable enterprise, and, though Keir’s 1769 process made him wealthy, the definitive technique was discovered by the Frenchman, Leblanc, in 1787.

The Royal Society was but one association of scientists and industrialists to which reference is made in behalf of a scientifically-based industrial technology. The Lunar Society of Birmingham7 has received individual treatment by Schofield, though the author’s 8 thesis has been criticized by D. W. F. Hardie, who sardonically remarks that the connection between science and technology among the Lunatics was established by Schofield through “guilt by association.” Again the general argument seems to be that a scientific “environment” stimulated technological advance, not only through the philosophical societies just mentioned, but through educational institutions, 9 libraries, literature and popular lectures as well. Despite the earlier agreement made here to recognize as “scientific” that which was produced by the scientifically knowledgeable, it must be remarked in conclusion that, in fact, a technological “environment” was stimulating scientific advance, to the extent that the two were related in the period under discussion.

References

  • back to text1  A.E. Musson and Eric Robinson, Science and Technology in the Industrial Revolution (Manchester: University of Manchester Press, 1969), P. 47.
  • back to textIbid., p. 3.
  • back to textIbid., p. 254–55.
  • back to text 4.  Robert E. Schofield, “The Industrial Orientation of Science in the Lunar Society of Birmingham,” Isis 48 (Dec. 1957), 408–15.
  • back to textJ. D. Bernal, Science in History, vol. 2: “The Scientific and Industrial Revolutions” (3rd. ed.; Cambridge: M.I.T. Press, 1969), p. 582.
  • back to textD.S.L. Cardwell, Turning Points in Western Technology (New York: Science History Publications, 1972), pp. 88–89.
  • back to textRobert E. Schofield, The Lunar Society of Birmingham (Oxford: Clarendon Press, 1964).
  • back to textD.W.F. Hardie, Review of The Lunar Society of Birmingham, Business History 8 (Jan. 1966), 73–74.
  • back to textMusson and Robinson, Revolution, p. 72.

Selected Bibliography

  • Ashby, Eric. Technology and the Academics. London: Macmillan, 1958.
  • Bernal, J. D. Science in History. Vol. 2: “The Scientific and Industrial Revolutions.” 3rd ed. Cambridge: M.I.T. Press, 1969.
  • Cardwell, D.S.L. “Power Technologies and the Advance of Science.” Technology and Culture 6 (Spring 1965), 188–207.
  • [________]. Turning Points in Western Technology. New York: Science History Publications, 1972.
  • [________]. “Science and the Steam Engine.” In Science and Society, pp. 81–96. Edited by Peter Mathias. Cambridge: Cambridge University Press, 1972.
  • Carter, C. F., and Williams, B.R. Industry and Technical Progress. London: Oxford University Press, 1957.
  • [________]. Science in Industry. London: Oxford University Press, 1959.
  • Clark, G. N. Science and Social Welfare in the Age of Newton . Oxford: Clarendon Press, 1937.
  • Clow, Archibald and Nan L. The Chemical Revolution. London: Batchworth Press, 1952.
  • Derry, T.K., and Williams, Trevor I. A Short History of Technology. New York: Oxford University Press, 1961.
  • Drucker, Peter F. “The Technological Revolution: Notes on the Relationship of Technology, Science, and Culture.” Technology and Culture 2 (Fall 1961), 342–51.
  • Finch, James K. “Engineering and Science: A Historical Review and Appraisal.” Technology and Culture 2 (Fall 1961), 318–32.
  • Hall, Alfred R. The Scientific Revolution 1500–1800 . London: Longmans, Green and Co., 1954.
  • [________]. “Engineering and the Scientific Revolution.” Technology and Culture 2 (Fall 1961), 333–41.
  • Hardie, D.W.F. Review of The Lunar Society of Birmingham. Business History 8 (Jan. 1966), 73–74.
  • Landes, David S. “Technological Change and Development in Western Europe, 1750–1914.” In Cambridge Economic History of Europe. Vol. 6: “The Industrial Revolution and After, Part 1,” Chap. 5. Cambridge: Cambridge University Press, 1966.
  • Lindsay, Robert B. The Role of Science in Civilization. New York: Harper & Row, 1963.
  • Mathias, Peter. “Who Unbound Prometheus? Science and Technical Change, 1600–1800.” In Science and Society, pp. 54–80. Edited by Peter Mathias. Cambridge: Cambridge University Press, 1972.
  • Meier, Richard L. Science and Economic Development. New York: John Wiley & Sons, Inc., 1956.
  • Musson, A.E., and Robinson, Eric. Science and Technology in the Industrial Revolution. Manchester: University of Manchester Press, 1969.
  • Musson, A.E., ed. Science, Technology and Economic Growth in the Eighteenth Century. London: Methuen & Co., 1972.
  • Schofield, Robert E. “The Industrial Orientation of Science in the Lunar Society of Birmingham.” Isis 48 (Dec. 1957), 408–15.
  • Usher, Abbott Payson. A History of Mechanical Inventions. London:
  • Oxford University Press, 1954.
  • Wolf, A. A History of Science, Technology, and Philosophy in the 16th and 17th Centuries. London: Allen & Unwin, 1935.
  • [________]. A History of Science, Technology, and Philosophy in the Eighteenth Century. New York: Macmillan, 1939.