Skip to main content

Electrolysis: Piles of Confusion and Poles of Attraction

  • Chapter
  • First Online:
Is Water H2O?

Part of the book series: Boston Studies in the Philosophy and History of Science ((BSPS,volume 293))

  • 3411 Accesses

Abstract

However one might assess the arguments about the nature of water in the Chemical Revolution (Chap. 1), it may seem that the electrolysis of water (first performed in 1800) must have produced decisive evidence that it was a compound substance. But electrolysis came with a serious puzzle: if the action of electricity was breaking up each particle of water into a particle of oxygen and a particle of hydrogen, how did the oxygen and hydrogen gases emerge at electrodes that were separated from each other by macroscopic distances? The distance problem turned the electrolysis of water into a serious anomaly, rather than positive evidence, for Lavoisierian chemistry. Ritter and his followers argued that electrolysis was in fact a pair of syntheses: water was an element after all, and its combination with positive and negative electricity formed oxygen and hydrogen. This view was dismissed by the majority of post-Lavoisierian chemists, but never conclusively refuted at the time. Those who opposed Ritter proposed a plethora of different solutions to the distance problem, none of them completely convincing. The modern ionic theory only emerged in the last years of the nineteenth century, so there was nearly a whole century of electrochemistry taking place without a consensus on some very basic questions. Nonetheless, electrochemistry made significant progress. Its experimental practices were stabilized and standardized without recourse to agreed-upon fundamental theory. In the theoretical realm there was pluralistic progress, with several competing systems each making its distinctive contributions, in productive interaction with each other.

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 259.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 379.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 329.99
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

Notes

  1. 1.

    “Battery” in later usage, that is. Originally the term was used to refer to a set (battery) of storage-jars for holding static electricity. After Volta’s invention, the term “galvanic battery” was used to describe a collection of cells producing electricity; over time, the term came to refer to single cells as well. In his original paper Volta (1800, 420) very sensibly proposed to call his instrument the electro-motive apparatus (appareil électro-moteur), but this name did not catch on.

  2. 2.

    This is how Priestley put it (1788, 154; emphases original): “That water is decomposed when inflammable air is procured from iron by steam, is not probable; since the inflammable principle [phlogiston] may very well be supposed to come from the iron, and the addition of weight acquired by the iron may be ascribed to the water which has displaced it. Also when the scale of iron, or finery cinder, is heated in inflammable air, it gives out what it had gained, viz. the water.” See also Priestley [1796] (1969), 30–33. To say that the metal absorbs water as it gives up phlogiston is quite like saying that the metal absorbs dephlogisticated water (which is what oxygen was, for Cavendish and Priestley).

  3. 3.

    See Snelders (1979) and Snelders (1988), 135–137, for further details on this experiment and its impact.

  4. 4.

    Note his use of the term “incommensurable”, 160 years before Kuhn and Feyerabend! See also Davy [1801] (1839), 206: “The facts relating to the separate production of oxygen and hydrogen acid and alkali in water, are totally incommensurable with the usually received theory of chemistry”.

  5. 5.

    The biographical information on Singer is taken from the Dictionary of National Biography (1897), vol. 52, 211–312.

  6. 6.

    In my references to Ostwald, the latter number cited (131 in this case) is the page number in the original German edition.

  7. 7.

    Ritter may have been anticipated in the synthesis view by one of Nicholson’s anonymous authors, as noted by Ostwald [1896] (1980), 148–149/152–153.

  8. 8.

    This is the experiment with the V-tube arrangement that I will describe in Sect. 2.2.1.2.

  9. 9.

    See Golinski (1992), 213, for a brief discussion of Gibbes’s continuing opposition to Lavoisierian theory.

  10. 10.

    See Siegfried (1964), Brooke (1980), 150, and also Knight (1978), 52. Of Davy’s own statements, note especially Davy (1808a), 33, occurring in the middle of the celebrated Bakerian Lecture in which he announced the discovery of potassium and sodium.

  11. 11.

    This paper is passed over by most historians who discuss Priestley’s work; one exception is a brief discussion given by Schofield (2004), 366.

  12. 12.

    Priestley did use the terms “oxygen” and “hydrogen” (or, “oxigen” and “hidrogen”) interchangeably with “dephlogisticated air” and “inflammable air” in this paper. To be precise: the term “hidrogen” only occurs in the marginal summaries, so Nicholson may have been responsible for that; however, “oxigen” occurs several times in Priestley’s main text, freely mixed in with “dephlogisticated air”.

  13. 13.

    Priestley’s description is ambiguous in this passage, as to whether these measures prevented the production of oxygen gas only, or both gases. On p. 201 he reports an experiment in which an oil-covering on the water also stopped the production of hydrogen (inflammable air). But the key point for the moment, which is clear throughout the paper, is that Priestley thought that the production of the two gases happened independently from each other. See Sect. 2.3.2 for further details.

  14. 14.

    In one place in the article (p. 202, middle) he has the polarity switched, but I think that is a simple error.

  15. 15.

    See also the discussion in Wilkinson (1804), 74–80.

  16. 16.

    Christoph Heinrich Pfaff (1773–1852) taught from 1798 at Kiel University, where he would remain until his death (Hufbauer 1982, 223); on his electrochemical work, see Kragh (2003).

  17. 17.

    See, for example, Pauling and Pauling (1975), 358; note that Fig. 2.2, taken from that text, indicates acidity and alkalinity around the anode and the cathode, respectively. See Sect. 2.2.2 for further details.

  18. 18.

    See Coutts (1959) for some informative details on Cruickshank’s life and work.

  19. 19.

    Coutts (1959, 125) explains the convention in the designation of the parts of the battery used by Cruickshank.

  20. 20.

    On Wilkinson’s life and work, see Thornton (1967).

  21. 21.

    Or that it had so little weight as to be undetectable by the technology of that time (which could in fact be said very fairly about electrons, too).

  22. 22.

    See Brown (1950), 372, and Brown (1979) on Rumford more generally. They also managed to put up sufficiently strong objections to Rumford’s more powerful anti-caloric argument based on the more famous “cannon-boring” argument showing the indefinite production of heat by friction (Chang 2004, 171, and references therein).

  23. 23.

    I have corrected the translation appearing in the English version of Ostwald’s text, which has the terms in square brackets as “potassium sulfate” and “potassium hydroxide”, which is anachronistic in a problematic way, as Berzelius and Hisinger were writing before Davy’s work on the isolation of potassium, when potash was widely regarded as elementary with only an unfounded suspicion that it might be a compound. (Rendering “vitriolic acid” as “sulphuric acid” is not problematic in the same way.)

  24. 24.

    Etienne Gaspard Robertson (1763–1837) in Paris independently advanced a similar view of a galvanic acid (Ostwald [1896] (1980), 209/216). Mottelay (1922, 350–351) explains that Robertson had a personal friendship with Volta, and he was one of the first in Paris to pay proper attention to Volta’s work; curiously, it was through Brugnatelli’s intervention in a lecture given by Robertson that the latter first began his interaction with Volta.

  25. 25.

    Wollaston also found ultraviolet rays, independently of Ritter’s work.

  26. 26.

    For the circumstance of Grotthuss’s death and his legacy, see Gorbunova et al. (1978), 233–234.

  27. 27.

    By this phrase I don’t mean just the electrochemistry of water, but the electrochemical system that took water as a compound.

  28. 28.

    That, too, will become less certain with deeper knowledge.

  29. 29.

    As Faraday meticulously numbered the paragraphs in all of his papers on “Experimental Researches on Electricity” in one consecutive sequence, I will note the paragraph numbers in my citations.

  30. 30.

    So we can see that Cleve was a whole century ahead of Laudan in making the pessimistic meta-induction from the history of science! Laudan’s point is a stronger one, as his examples concern theories that were once well-established, as opposed to Cleve’s ephemera. Cleve may have been wiser than his pupil, but Arrhenius was both the more typical and the more productive player in this game.

  31. 31.

    T. M. Lowry (1936, 270) notes that it was only in 1840 that John Frederic Daniell advanced the general view that a salt was a binary compound of two radicals, not of an acid and an alkali in their entirety.

  32. 32.

    Lowry (1936, 11, 62, 288) states that Davy used caustic potash and caustic soda, and identifies them as the hydroxides of the metals, KOH and NaOH. The non-caustic varieties are the carbonates: K2CO3 and Na2CO3. See Lowry’s explanation (pp. 283–284) on how Davy and others gradually moved away from his initial view that potash and soda were simple oxides.

  33. 33.

    As Knight (1967, 21) explains, this was one reason for which he did not wholly embrace John Dalton’s atomic theory, which postulated a distinct atom for each chemical element recognized as such at the time. While accepting the Lavoisierian operational definition of an element as a hitherto undecomposed substance, Davy focused his effort on effecting new decompositions. One of his motivations for entertaining the revival of phlogiston (see Chap. 1, Sect. 1.2.2) was to see if he could not reduce the number of chemical elements (see Siegfried 1964).

  34. 34.

    In modern days the Daniell cell has also displaced Volta’s as the paradigm of theoretical exposition in electrochemistry, as explained in Sect. 2.3.4 and further in Chang (2011c).

  35. 35.

    Many historians and philosophers have criticized Kuhn on this point. For an early example see Toulmin (1970).

  36. 36.

    In this he had two important predecessors. One was Davy, whose conception of the relation between the electrical and chemical forces was more subtle, complex and vague that Berzelius’s. Russell’s view (1959, 12) is that “Faraday was the one to influence the world to look favourably on his master’s theories. And he did this by enshrining them in his own.” The other predecessor I want to highlight is Donovan (1816, 278), who published this insightful view 15 years before Faraday’s work: “it was found that copper lost its affinity for oxygen, by contact with zinc;... the affinity of the zinc for oxygen was much increased by contact with copper. I think therefore there is nothing overstrained in the inference that one has gained what the other lost, or in other words that the copper has transferred a portion of its affinity for oxygen to the zinc.”

  37. 37.

    As usual, the exhaustive treatments by Ostwald [1896] (1980), Mottelay (1922) and Partington (1964) provide very useful exceptions. Another notable exception, though very brief, is Harold Hartley’s discussion of “Faraday’s successors and the theory of electrolytic dissociation” (Hartley 1971, ch. 7); Hartley was not worried about avoiding historiographical whiggism, but his perspective in 1931, when he composed that piece, was that his own current situation resembled that of the rich and uncertain field that Faraday faced, rather than the over-clarity of Arrhenius’s work.

  38. 38.

    Schwab’s focus was on science education. The increasing prevalence of fluid inquiry in science means that it becomes increasingly necessary to train science students for it—in other words, to equip them for critical thinking; see Siegel (1990), 99–102, for further reflections on this point.

  39. 39.

    I carried out these experiments in the electrochemistry lab of Daren Caruana at the Department of Chemistry at University College London. I would like to thank Dr. Caruana and his colleagues most sincerely for the use of the laboratory facilities and all the friendly advice they gave me. I also would like to thank Rosemary Coates, who assisted me most congenially and ably in these and other experiments, and the Leverhulme Trust, whose research grant provided much-needed funds and an authoritative seal of approval.

  40. 40.

    Or is it possible that the application of electricity generates oxygen by decomposing CO2, which will be found dissolved in the water in relative abundance?

  41. 41.

    Davy notes that he had performed these experiments several years earlier and published the results in Nicholson’s Journal in 1800 (volume 4); see also the summary in Donovan (1816), 43. Priestley would have read these papers.

  42. 42.

    For hosting these experiments, I thank the Department of Chemistry at the University of Cambridge, and Dr. Peter Wothers, Mr. Chris Brackstone, and Mr. Gary Herrington.

  43. 43.

    Darrigol (2000), 266–274, quotation on p. 266, footnote 1.

  44. 44.

    Yet it seems evident that Kuhn had not taken an in-depth look, as he says that “both viewpoints were briefly in the field at once” (p. 23, emphasis added).

  45. 45.

    Partington (1964, 17) also mentions that the Oxford dry pile had continued to work for more than a century.

References

  • Arrhenius, Svante. 1902. Text-book of electrochemistry (trans: McCrae, J.). London: Longmans, Green, and Co.

    Google Scholar 

  • Berzelius, Jöns Jakob. 1811. Essai sur la nomenclature chimique. Journal de Physique 73: 253–286.

    Google Scholar 

  • Berzelius, Jöns Jakob, and Wilhelm Hisinger. 1803. Versuch, betreffend die Wirkung der elektrischen Säule auf Salze und auf einige von ihren Basen. (Gehlen’s) Neuen allgemeinen Journal der Chemie 1: 115–149.

    Google Scholar 

  • Bevilacqua, Fabio and Lucio Fregonese, eds. 2000–2003. Nuova Voltiana: Studies on Volta and his times, 5 vols. Milan: Hoepli.

    Google Scholar 

  • Brock, William H. 1992. The Fontana history of chemistry. London: Fontana Press.

    Google Scholar 

  • Brooke, John Hedley. 1980. Davy’s chemical outlook: The acid test. In Science and the sons of genius: Studies on Humphry Davy, eds. Sophie Forgan, 121–175. London: Science Reviews Ltd.

    Google Scholar 

  • Brown, Sanborn C. 1950. The caloric theory of heat. American Journal of Physics 18: 367–373.

    Article  Google Scholar 

  • Brown, Sanborn C. 1979. Benjamin Thompson, Count Rumford. Cambridge, MA: The MIT Press.

    Google Scholar 

  • Cajori, Florian. 1929. A history of physics. New York: Macmillan.

    Google Scholar 

  • Chang, Hasok. 1999. History and philosophy of science as a continuation of science by other means. Science and Education 8: 413–425.

    Article  Google Scholar 

  • Chang, Hasok. 2004. Inventing temperature: Measurement and scientific progress. New York: Oxford University Press.

    Book  Google Scholar 

  • Chang, Hasok. 2007a. Scientific progress: Beyond foundationalism and coherentism. In Philosophy of science (Royal Institute of Philosophy Supplement 61), eds. Anthony O’Hear, 1–20. Cambridge: Cambridge University Press.

    Article  Google Scholar 

  • Chang, Hasok. 2007b. The myth of the boiling point. http://www.cam.ac.uk/hps/chang/boiling. First posted on 18 Oct 2007.

  • Chang, Raymond. 2010. Chemistry. Boston: McGraw-Hill.

    Google Scholar 

  • Chang, Hasok. 2011c. How historical experiments can improve scientific knowledge and science education: The cases of boiling water and electrochemistry. Science and Education 20: 317–341.

    Google Scholar 

  • Chang, Hasok. 2011d. Compositionism as a dominant way of knowing in modern chemistry. History of Science 49: 247–268.

    Google Scholar 

  • Christensen, Dan Ch. 1995. The Ørsted–Ritter partnership and the birth of romantic natural philosophy. Annals of Science 52: 153–185.

    Article  Google Scholar 

  • Court, S. 1972. The Annales de chimie 1789–1815. Ambix 19: 113–128.

    Google Scholar 

  • Coutts, A. 1959. William Cruickshank of Woolwich. Annals of Science 15: 121–133.

    Article  Google Scholar 

  • Croft, A.J. 1984. The oxford electric bell. European Journal of Physics 5: 193–194.

    Article  Google Scholar 

  • Crosland, Maurice. 1978. Gay-Lussac: Scientist and bourgeois. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Crosland, Maurice. 1980. Davy and Gay-Lussac: Competition and contrast. In Science and the sons of genius: Studies on Humphry Davy, ed. Sophie Forgan, 95–120. London: Science Reviews Ltd.

    Google Scholar 

  • Cruickshank, William. 1800a. Some experiments and observations on galvanic electricity. (Nicholson’s) Journal of Chemistry, Natural Philosophy, and the Arts 4: 187–191.

    Google Scholar 

  • Cruickshank, William. 1800b. Additional remarks on galvanic electricity. (Nicholson’s) Journal of Chemistry, Natural Philosophy, and the Arts 4: 254–264.

    Google Scholar 

  • Cunningham, Andrew, and Nicholas Jardine, eds. 1990. Romanticism and the sciences. Cambridge: Cambridge University Press.

    Google Scholar 

  • Darrigol, Olivier. 2000. Electrodynamics from Ampère to Einstein. Oxford: Oxford University Press.

    Google Scholar 

  • Davy, Humphry. 1800a. Account of some experiments made with the galvanic apparatus of Signor Volta. (Nicholson’s) Journal of Chemistry, Natural Philosophy, and the Arts 4: 275–281.

    Google Scholar 

  • Davy, Humphry. 1800b. An account of some additional experiments and observations on the galvanic phenomena. (Nicholson’s) Journal of Chemistry, Natural Philosophy, and the Arts 4: 394–402.

    Google Scholar 

  • Davy, Humphry. 1807. The Bakerian Lecture [for 1806]: On some chemical agencies of electricity. Philosophical Transactions of the Royal Society 97: 1–56.

    Google Scholar 

  • Davy, Humphry. 1808a. The Bakerian Lecture [for 1807]: On some new phenomena of chemical changes produced by electricity, particularly the decomposition of the fixed alkalies, and the exhibition of the new substances which constitute their bases; and on the general nature of alkaline bodies. Philosophical Transactions of the Royal Society 98: 1–44.

    Google Scholar 

  • Davy, Humphry. 1808b. Electro-chemical researches, on the decomposition of the earths; with observations on the metals obtained from the alkaline earths, and on the amalgam procured from ammonia. Philosophical Transactions of the Royal Society 98: 333–370.

    Article  Google Scholar 

  • Davy, Humphry. 1809. The Bakerian Lecture [for 1808]: An account of some new analytical researches on the nature of certain bodies, particularly the alkalies, phosphorus, sulphur, carbonaceous matter, and the acids hitherto undecompounded; with some general observations on chemical theory. Philosophical Transactions of the Royal Society 99: 39–104.

    Article  Google Scholar 

  • Davy, Humphry. 1810. The Bakerian Lecture for 1809: On some new electrochemical researches, on various objects, particularly the metallic bodies, from the alkalies, and earths, and on some combinations of hydrogene [sic]. Philosophical Transactions of the Royal Society 100: 16–74.

    Article  Google Scholar 

  • Davy, Humphry. 1812. Elements of chemical philosophy. London: J. Johnson and Co.

    Google Scholar 

  • Donovan, Michael. 1816. Essay on the origin, progress and present state of galvanism, etc. Dublin: Hodges and McArthur.

    Google Scholar 

  • Faraday, Michael. 1833. Experimental researches in electricity, fifth series. Philosophical Transactions of the Royal Society 123: 675–710.

    Article  Google Scholar 

  • Faraday, Michael. 1834. Experimental researches in electricity, seventh series. Philosophical Transactions of the Royal Society 124: 77–122.

    Article  Google Scholar 

  • Faraday, Michael. 1844. Experimental researches in electricity, 2 vols, reprinted from the Philosophical Transactions of 1838–1843, with other electrical papers from the Quarterly Journal of Science and Philosophical Magazine. London: Richard and John Edward Taylor.

    Google Scholar 

  • Faraday, Michael. 1993. The forces of matter (Royal Institution Christmas Lectures). Amherst, NY: Prometheus Books.

    Google Scholar 

  • Gilbert, T.R., R.V. Kirss, N. Foster, and G. Davies. 2009. Chemistry: The science in context. New York: Norton.

    Google Scholar 

  • Golinski, Jan. 1992. Science as public culture: Chemistry and enlightenment in Britain 1760–1820. Cambridge: Cambridge University Press.

    Google Scholar 

  • Gorbunova, K.M., L.J. Antropov, Yu.I. Solov’ev, and J.P. Stradins. 1978. Early electrochemistry in the USSR. In Proceedings of the Symposium on Selected Topics in the History of Electrochemistry, eds. George Dubpernell and J.H. Westbrook, 226–256. Princeton: The Electrochemical Society.

    Google Scholar 

  • Gray, Harry B., and Gilbert P. Haight Jr. 1967. Basic principles of chemistry. New York: W. A. Benjamin.

    Google Scholar 

  • Gray, Tamsin, Rosemary Coates, and Mårten Åkesson. 2007. The elementary nature of chlorine. In An element of controversy: The life of chlorine in science, medicine, technology and war, eds. Hasok Chang and Catherine Jackson, 41–72. London: British Society for the History of Science.

    Google Scholar 

  • Grotthuss, Christian Johann Dietrich (Theodor). 1806. Memoir upon the decomposition of water, and of the bodies which it holds in solution, by means of galvanic electricity. Philosophical Magazine 25: 330–339.

    Google Scholar 

  • Grotthuss, Christian Johann Dietrich (Theodor). 1810. On the influence of galvanic electricity in metallic arborizations. (Nicholson’s) Journal of Chemistry, Natural Philosophy, and the Arts 28: 112–125.

    Google Scholar 

  • Hackmann, Willem. 2001. The enigma of Volta’s “contact tension” and the development of the “Dry pile”. In Nuova Voltiana: Studies on Volta and his times, vol. 3, eds. F. Bevilacqua and L. Fregonese, 103–119. Milan: Hoepli.

    Google Scholar 

  • Hartley, Harold. 1971. Studies in the history of chemistry. Oxford: Clarendon Press.

    Google Scholar 

  • Haüy, René-Just. 1806. Traité élémentaire de physique. Paris: Courcier.

    Google Scholar 

  • Henry, William. 1813. On the theories of the excitement of galvanic electricity. (Nicholson’s) Journal of Chemistry, Natural Philosophy, and the Arts 35: 259–271.

    Google Scholar 

  • Hong, Sungook. 1994. Controversy over Voltaic contact phenomena. Archive for History of Exact Sciences 47: 233–289.

    Article  Google Scholar 

  • Housecroft, C.E., and E.C. Constable. 2010. Chemistry: An introduction to organic, inorganic and physical chemistry. Harlow: Pearson.

    Google Scholar 

  • Hufbauer, Karl. 1982. The formation of the German chemical community (1720–1795). Berkeley/Los Angeles: University of California Press.

    Google Scholar 

  • James, Frank A.J.L. 1989. Michael Faraday’s first law of electrochemistry – How context develops new knowledge. In Electrochemistry, past and present, eds. John T. Stock and Mary Virginia Orna, 32–49. Washington, DC: American Chemical Society.

    Chapter  Google Scholar 

  • Kipnis, Nahum. 2001. Debating the nature of Voltaic electricity. In Nuova Voltiana: Studies on Volta and his times, vol. 3, eds. F. Bevilacqua and L. Fregonese, 121–151. Milan: Hoepli.

    Google Scholar 

  • Knight, David. 1967. Atoms and elements. London: Hutchinson.

    Google Scholar 

  • Knight, David. 1978. The transcendental part of chemistry. Folkestone: Dawson.

    Google Scholar 

  • Kragh, Helge. 2000. Confusion and controversy: Nineteenth century theories of the Voltaic pile. In Nuova Voltiana: Studies on Volta and his times, vol. 1, eds. F. Bevilacqua and L. Fregonese, 133–157. Milan: Hoepli.

    Google Scholar 

  • Kuhn, Thomas S. 1970. The structure of scientific revolutions, 2nd ed. Chicago: University of Chicago Press.

    Google Scholar 

  • Kuhn, Thomas S. 2000. The road since Structure: Philosophical essays, 1970–1993, with an autobiographical interview. Chicago: University of Chicago Press.

    Google Scholar 

  • Levine, I.N. 2002. Physical chemistry. Boston: McGraw-Hill.

    Google Scholar 

  • Lilley, Samuel. 1948. ‘Nicholson’s Journal’ (1797–1813). Annals of Science 6: 78–101.

    Article  Google Scholar 

  • Lowry, T.M. 1936. Historical introduction to chemistry, revised ed. London: Macmillan.

    Google Scholar 

  • Lund, Matthew. 2010. N. R. Hanson: Observation, discovery, and scientific change. Amherst, NY: Prometheus Books.

    Google Scholar 

  • Melhado, Evan M. 1980. Jacob Berzelius: The emergence of his chemical system. Stockholm: Almqvist & Wiksell International.

    Google Scholar 

  • Mottelay, Paul Fleury. 1922. Bibliographical history of electricity and magnetism: Chronologically arranged. London: Charles Griffin & Company Limited.

    Google Scholar 

  • Nicholson, William. 1800. Account of the new electrical or galvanic apparatus of Sig. Alessandro Volta, and experiments performed with the same. (Nicholson’s) Journal of Chemistry, Natural Philosophy, and the Arts 4: 179–187.

    Google Scholar 

  • Ostwald, Wilhelm. 1980. Electrochemistry: History and theory, 2 vols (trans: Date, N. P.). New Delhi/Bombay/Calcutta/New York: Amerind Publishing Co. Pvt. Lt.

    Google Scholar 

  • Pancaldi, Giuliano. 2003. Volta: Science and culture in the age of enlightenment. Princeton: Princeton University Press.

    Google Scholar 

  • Partington, J.R. 1964. A history of chemistry, vol. 4. London: Macmillan.

    Google Scholar 

  • Pauling, Linus, and Peter Pauling. 1975. Chemistry. San Francisco: W. H. Freeman and Company.

    Google Scholar 

  • Priestley, Joseph. 1788. Experiments and observations relating to the principle of acidity, the composition of water, and phlogiston. Philosophical Transactions of the Royal Society 78: 147–157.

    Article  Google Scholar 

  • Priestley, Joseph. [1796] 1969. Considerations on the doctrine of phlogiston, and the decomposition of water (and two lectures on combustion, etc. by John MacLean). New York: Kraus Reprint Co.

    Google Scholar 

  • Priestley, Joseph. 1802. Observations and experiments relating to the pile of Volta. (Nicholson’s) Journal of Chemistry, Natural Philosophy, and the Arts, new series 1: 198–204.

    Google Scholar 

  • Rumford [Benjamin Thompson, Count]. 1799. An inquiry concerning the weight ascribed to heat. Philosophical Transactions of the Royal Society 89: 179–194.

    Google Scholar 

  • Russell, Colin A. 1959. The electrochemical theory of Sir Humphry Davy (in 2 parts). Annals of Science 15: 1–25.

    Article  Google Scholar 

  • Russell, Colin A. 1963. The electrochemical theory of Berzelius (in 2 parts). Annals of Science 19: 117–145.

    Article  Google Scholar 

  • Schofield, Robert E. 2004. The enlightened Joseph Priestley: A study of his life and work from 1773 to 1804. University Park: Pennsylvania State University Press.

    Google Scholar 

  • Schwab, J.J. 1962. The teaching of science as enquiry. Cambridge, MA: Harvard University Press.

    Google Scholar 

  • Siegel, Harvey. 1990. Educating reason: Rationality, critical thinking and education. New York: Routledge.

    Google Scholar 

  • Sinclair, Alexandra. 2009. Beyond the law(s): Michael Faraday’s experimental researches, series 8. London: University College London.

    Google Scholar 

  • Siegfried, Robert. 1964. The phlogistic conjectures of Humphry Davy. Chymia 9: 117–124.

    Article  Google Scholar 

  • Singer, George John. 1814. Elements of electricity and electro-chemistry. London: Longman, Hurst, Rees, Orme, and Brown; R. Triphook.

    Google Scholar 

  • Snelders, H.A.M. 1979. The Amsterdam experiment on the analysis and synthesis of water 1789. Ambix 26: 116–133.

    Article  Google Scholar 

  • Snelders, H.A.M. 1988. The new chemistry in the Netherlands. Osiris (2nd Series) 4: 121–145.

    Google Scholar 

  • Stevenson, W.F. 1849. The composition of hydrogen, and the non-decomposition of water incontrovertibly established. London: James Ridgway.

    Google Scholar 

  • Thornton, John L. 1967. Charles Hunnings Wilkinson (1763 or 64–1850). Annals of Science 23: 277–286.

    Article  Google Scholar 

  • Toulmin, Stephen. 1970. Does the distinction between normal and revolutionary science hold water? In Criticism and the growth of knowledge, eds. Imre Lakatos and Alan Musgrave, 39–47. Cambridge: Cambridge University Press.

    Google Scholar 

  • Volta, Alessandro. 1800. On the electricity excited by the mere contact of conducting substances of different kinds. In a letter from Mr. Alexander Volta, F.R.S. Professor of Natural Philosophy in the University of Pavia, to the Rt. Hon. Sir Joseph Banks, Bart. K.B. P.R.S. Philosophical Transactions of the Royal Society 90: 403–431.

    Google Scholar 

  • Wetzels, Walter J. 1978a. J. W. Ritter: The beginnings of electrochemistry in Germany (with commentary by George Dubpernell). In Proceedings of the Symposium on Selected Topics in the History of Electrochemistry, eds. George Dubpernell and J.H. Westbrook, 68–76. Princeton: The Electrochemical Society.

    Google Scholar 

  • Wetzels, Walter J. 1978b. J. W. Ritter: Electrolysis with the Volta-pile (with commentary by George Dubpernell). In Proceedings of the Symposium on Selected Topics in the History of Electrochemistry, eds. George Dubpernell and J. H. Westbrook, 77–87. Princeton: The Electrochemical Society.

    Google Scholar 

  • Wetzels, Walter J. 1990. Johann Wilhelm Ritter: Romantic physics in Germany. In Romanticism and the sciences, eds. Andrew Cunningham and Nicholas Jardine, 199–212. Cambridge: Cambridge University Press.

    Google Scholar 

  • Wilkinson, Charles Hunnings. 1804. Elements of galvanism, in theory and practice, with a comprehensive view of its history, from the first experiments of Galvani to the present time, etc., 2 vols. London: John Murray.

    Google Scholar 

  • Williams, L. Pearce. 1965. Michael Faraday. London: Chapman and Hall.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Rights and permissions

Reprints and permissions

Copyright information

© 2012 Springer Science+Business Media B.V.

About this chapter

Cite this chapter

Chang, H. (2012). Electrolysis: Piles of Confusion and Poles of Attraction. In: Is Water H2O?. Boston Studies in the Philosophy and History of Science, vol 293. Springer, Dordrecht. https://doi.org/10.1007/978-94-007-3932-1_2

Download citation

  • DOI: https://doi.org/10.1007/978-94-007-3932-1_2

  • Published:

  • Publisher Name: Springer, Dordrecht

  • Print ISBN: 978-94-007-3931-4

  • Online ISBN: 978-94-007-3932-1

  • eBook Packages: Humanities, Social Sciences and LawHistory (R0)

Publish with us

Policies and ethics