94
Views
1
CrossRef citations to date
0
Altmetric
Original Articles

A disciplinary immigrant. Alexander Smith at the University of Chicago, 1894–1911

Pages 221-256 | Received 07 Mar 2007, Published online: 03 Apr 2008
 

Summary

The publication in 1906 of Alexander Smith's Introduction to general inorganic chemistry inaugurated a decisive change in chemical pedagogy in the US, the effects of which are still evident. The nature and extent of Smith's innovations are described through a comparison of his text to its source material and contemporaries. His authoritative command of and whole-hearted commitment to the intellectual framework of Ionist physical chemistry set his text apart from its American competitors, while his efforts to make the tools of physical chemistry immediately useful to his readers distinguished it from its most immediate source material, Wilhelm Ostwald's Grundlinien der anorganischen Chemie. Smith's curricular innovations in chemistry were a practical expression of his radically restrictive view of the social role of collegiate education, which he conceived as solely of use for its ability to prepare students for professional life. During the fifteen years prior to the publication of his groundbreaking textbook, Smith underwent two critical, formative experiences. First, he retreated intellectually from the structural organic chemistry in which he was trained, ultimately adopting a professional identity as a physical inorganic chemist. His involvement in the controversy regarding the structure of 1,3-diketones reveals much about his reasons for eventually abandoning organic chemistry. Second, he served the National Education Association as chairman of the Sub-committee on College Entrance Requirements in Chemistry, in the process making a close study of the ends and methods of secondary and collegiate education. These experiences made him unique among proponents of physical chemistry in the US, and help account for the unique nature of his contributions to the development of the chemical professions.

Acknowledgements

This work was supported in part by Mount Holyoke College, the Chemical Heritage Foundation, and by a Professional Development Fellowship from the National Science Foundation (Award No. SES-0551046). I benefited greatly from the sabbatical-year hospitality of Professor Trevor Levere and his students Martha Harris and Andre Siegel at the Institute for the History and Philosophy of Science and Technology at the University of Toronto. Their constructive criticism, and that of John Servos and an anonymous reviewer, greatly clarified my thinking and sharpened my prose. I am grateful to Jay Satterfield (University of Chicago Archives), Vera Dragisich (University of Chicago Department of Chemistry), Jocelyn Wilk (Columbia University Archives—Columbiana Library), Christopher Prom (University of Illinois Archives), and Beth Swift, and Johanna Herring (Wabash College) for their generous assistance.

Notes

1Sinclair Lewis, Arrowsmith (New York, 1925), 15.

2John W. Servos, Physical Chemistry from Ostwald to Pauling: the Making of a Science in America (Princeton, NJ, 1990).

3Mary Jo Nye, Before Big Science: the Pursuit of Modern Chemistry and Physics, 1800–1940 (New York, 1996).

4Joel H. Hildebrand, ‘The Early Training of Scientists’, Science, 55 (1922), 355–58 (356).

5Wilhelm Ostwald, Lehrbuch der allgemeinen Chemie (Leipzig, 1885); Grundriss der allgemeinen Chemie (Leipzig, 1890); Wissenschaftliche Grundlagen der analytischen Chemie (Leipzig, 1894); Elektrochemie, ihre Geschichte und Lehre (Leipzig, 1896); Grundlinien der anorganischen Chemie (Leipzig, 1900).

6Nobel laureate (1905) for his structural identification and synthesis of indigo, and a pioneer in stereochemical analysis, Adolf von Baeyer (1835–1917) was one of the major figures in German organic chemistry during the late nineteenth century. After pursuing research under the direction of Robert Bunsen and August Kekulé, he received his doctorate from the University of Berlin in 1858 for work performed in the latter's private laboratory. His independent career passed through the Universities of Berlin and Strasbourg, before he assumed the chair at Munich after Liebig's death in 1873. The continuing significance of his work for contemporary chemists was recently reviewed on the centenary of his Nobel prize (Armin de Meijere, ‘Adolf von Baeyer: Winner of the nobel prize for chemistry 1905’, Angewandte Chemie, International Edition, 44 (2005), 7836–40). For brief biographies by his contemporaries, see William Henry Perkin, ‘Prof. Adolf von Baeyer’, Nature, 100 (1917), 188–90; Richard Willstätter, ‘Adolf von Baeyer’, Naturwissenschaften, 3 (1915), 559–63.

7Still familiar to chemists for two chemical reactions and a double-necked distillation flask that bear his name, Ludwig Claisen (1851–1930), received his doctorate from Kekulé at Bonn in 1875. After several years under von Bayer in Munich, Claisen took the chair in chemistry at the University of Aachen, where he became a leading figure in the chemistry of carbonyl group condensations, high-temperature rearrangement of aromatic compounds, and the behaviour of enolizable ketones. For leading references, see Ludwig Anschütz, ‘Claisen, Ludwig Rainer’, in Neue deutsche Biographie (Berlin, 1957), 257–58.

8Two important exceptions to this statement are George DeBoer, who notes Smith's significance for secondary education and analyses his work for the National Education Association (G.E. DeBoer, A History of Ideas in Science Education: Implications for Practice (New York, 1991), 54–58), and John Servos, who describes the University of Chicago's unusual but influential role in promoting physical chemistry in the US (note 2, 85–87).

9For professional biographies of Smith written by his colleagues and students, see the following: James Kendall, ‘Alexander Smith, 1865–1922’, Proceedings of the American Chemical Society (1922), 113–17; Alan C.W. Menzies, ‘Obituary: Alexander Smith’, Science, 56 (1922), 409; Ralph A. McKee, ‘Alexander Smith, the Investigator’, Journal of Chemical Education, 9 (1932), 247–54; James Kendall, ‘Alexander Smith as an educator’, Journal of Chemical Education, 9 (1932), 255–60; W.A. Noyes, ‘Smith, Alexander’, in Dictionary of American Biography, XVIII, edited by Dumas Malone (New York, 1935), 235.

10Like Smith, Walker (1863–1935) began his doctoral education with Claisen, but in 1888, less than a year after his arrival in Munich, he moved to Leipzig to work with Ostwald. He received his Ph.D. from Leipzig in 1889. After a brief assistantship in Edinburgh, he taught at University College-London (1892–1894) and the University of Dundee (1894–1908) before returning to Edinburgh to assume the chair previously held by Alexander Crum Brown. He was knighted in 1921. As Ostwald's first British pupil, Walker's influence on the introduction of physical chemistry into British science was seminal. See John Shorter, ‘Walker, Sir James’, in Dictionary of Nineteenth-Century British Scientists, IV, edited by Bernard Lightman (Chicago, 2004), 2081–2082.

11National Education Association, Report of the Committee on Secondary School Studies (Washington, DC, 1893)

12National Education Association, Report of the Committee on College Entrance Requirements (Chicago, 1899). The chemistry sub-committee's report is on pages 165–171. See also A.F. Nightingale, ‘The Committee on College Entrance Requirements: Report of the Chairman’, The school review, 5 (1897), 321–31, and A.F. Nightingale, David Starr Jordan, Samuel Thurber, Wilson Farrand, and J.H. Kirkland, ‘College Entrance Requirements’, The school review, 7 (1899), 388–408.

13DeBoer (note 8), 50–53.

14Alexander Smith, ‘Articulation of School and College Work in the Sciences’, The School Review, 7 (1899), 411–17, 453–61, 527–35.

15Alexander Smith and Edwin H. Hall, The Teaching of Chemistry and Physics in the Secondary School (New York, 1902).

16James Hocker Mason, ‘The Educational Milieu, 1874–1911: College Entrance Requirements and the Shaping of Secondary Education’, The English Journal, 68 (1979), 40–45. For Smith's work with the College Board, see correspondence in ASC, Box 2, Folder 42.

17Alexander Smith, Introduction to General Inorganic Chemistry (New York, 1906).

18The fact that Smith's name has been dissociated from his innovations and forgotten, while Pauling's has not, can be attributed to many factors, not least the overwhelming influence of Pauling's research in theoretical and biological chemistry and his singular and richly deserved celebrity. But it may also reveal, in part, the extent to which Smith's contributions have been seamlessly incorporated into the grammar of chemical education, into the academic and professional community's most basic assumptions about what a general chemistry textbook and course can be.

19Kendall, ‘Alexander Smith’ (note 9), 113. This is an interesting comparison, because the Lehrbuch was not written for direct classroom use, but as a resource for scientists and teachers who wanted to learn physical chemistry well enough to incorporate it into their own work.

20Smith met William Albert Noyes (1857–1941) in Munich, where both were doctoral students of Ludwig Claisen. Noyes received the Ph.D. from Munich, his second, in 1889; his first had been obtained at Johns Hopkins in 1885, under the direction of Ira Remsen. In a period when all organic chemists were required to be skilled analysts to some degree, Noyes was an exceptionally talented analytical chemist and, before assuming the chair in chemistry at the University of Illinois, served as the first chief chemist at the US Bureau of Standards. He edited the Journal of the American Chemical Society from 1902 to 1917, and was the founding editor of both Chemical Abstracts and Chemical Reviews. See Albert B. Costa, ‘Noyes, Wiliam Albert’, in Dictionary of American Biography, Supplement Three (New York, 1973), 565–66.

21Noyes, ‘Alexander Smith’ (note 9), 235. Although Noyes and Smith were intimate friends for many years, they were also bitter competitors on the textbook market. Noyes’ assessment thus carries a warm, and characteristic, tone of generosity.

22Originally hired at the University of Chicago as a docent without salary, Julius Stieglitz (1867–1937) rapidly achieved faculty status, and chaired the department from 1915 until his retirement. His research into the behaviour and theoretical description of acid–base indicators and his work on homogeneous catalysis appear particularly far-sighted to a twenty-first-century sensibility, comfortably straddling the subject matters of physical and organic chemistry. In addition to his scientific achievements, Stieglitz was a popular leader in the US chemical community. Born in New Jersey of German immigrant parents (and the younger brother of the photographer Alfred Stieglitz), Stieglitz was the President of the American Chemical Society when the United States declared war on Germany in 1917. Despite strong anti-German sentiment, he suffered no apparent loss of prestige during the war and became an important spokesman for chemistry in the national interest. See H.I. Schlesinger, ‘Stieglitz, Julius’, in Dictionary of American Biography, Supplement Two (New York, 1958), 630–31.

23Julius Stieglitz, ‘The Department of Chemistry’, The University Record, 14 (1928), 236–39.

24L. Cohn, A Laboratory Manual of Organic Chemistry (London, 1895).

25Alexander Smith, ‘On Two Stereoisomeric Hydrazones of Benzoin’, American Chemical Journal, 14 (1894), 108–15; Alexander Smith, ‘Ueber die Einwirkung von Hydrazin und von Phenylhydrazin auf 1,4-Diketone’, Justus Liebigs Annalen der Chemie und Pharmacie, 289 (1896), 310–37; Alexander Smith, ‘On Potassium Cyanide As a Condensing Reagent’, American Chemical Journal, 22 (1899), 247–56; Alexander Smith, ‘On the Phenylhydrazones of Benzoin’, American Chemical Journal, 22 (1899), 198–207.

26Louis Pasteur, Researches on the Molecular Asymmetry of Natural Organic Products (Edinburgh, 1897).

27Alexander Smith, A Laboratory Outline of General Chemistry (Chicago, 1900).

28Alexander Smith and H.N. McCoy, ‘Notizen über die Einwirkung von Phenylhydrazin auf einige 1,4-Diketone’, Berichte der deutschen chemischen Gesellschaft, 35 (1902), 2169–2171.

29Alexander Smith, ‘Amorphous Sulphur and Its Relation to the Freezing Point of Liquid Sulphur’, Proceedings of the Royal Society of Edinburgh, 24 (1901), 299–301.

30DeBoer (note 8), 222.

31Smith and Hall (note 15), 6.

32Anonymous chemistry professor, quoted in Sheila Tobias, They're not Dumb, They're Different: Stalking the Second Tier (Tucson, AZ, 1990), 55.

33The autodidact Benjamin Silliman (1779–1864), originally trained as a lawyer, was appointed to the chair in chemistry and natural history at Yale College in 1802, then spent two years in Philadelphia studying the subjects he was later to teach. He held his position at Yale until 1853. His Elements of Chemistry (published in two volumes, 1830–1831) set the American standard through the 1870s. See Charles H. Warren, ‘Silliman, Benjamin’, in Dictionary of American Biography, XVII, edited by Dumas Malone (New York, 1935), 160–73. Ira Remsen (1846–1927), in whose laboratory at Johns Hopkins University saccharin was first prepared, received his Ph.D. in 1870 in Friedrich Wöhler's institute at Göttingen, under the direction of Rudolph Fittig. He began his independent career at Williams College in Massachusetts, but was called by Daniel Coit Gilman to be the first Professor of Chemistry at Hopkins, becoming president of that institution in 1901. In addition to being the most significant source of US-trained Ph.D. chemists in the late nineteenth century, Remsen founded the American Chemical Journal, the first US journal of international importance in the field, and published a series of textbooks (beginning with An Introduction to the Study of the Compounds of Carbon in 1885) that dominated chemical education in the US until after 1900. See William A. Noyes, ‘Remsen, Ira’, in Dictionary of American Biography, XV, edited by Dumas Malone (New York, 1935), 500–502.

34‘We are not training students to use four or six year hence even the chemistry of to-day, much less the chemistry of 1880 or 1890. We are training them to understand the chemistry and biochemistry of the future and to apply and expand the science as it will be several years hence. All that we know for certain about that chemistry is that it will be less capable of mechanical, unintelligent use than the chemistry of the past, and that ability to apply theoretical conceptions will be more desirable, nay indispensable, than ever.’ Alexander Smith, ‘The Training of Chemists’, Science, 43 (1916), 619–26 (622).

35As James Walker wrote, ‘By placing our theoretical requirements before the physicist we suggest to him new fields for cultivation, and provide a fresh stimulus to his research. This is a general method of repayment of the more concrete to the more abstract sciences. This repayment may strike even us as inadequate, but we have the satisfaction to know that if our wares are of comparatively little utility to the physicist, they are highly prized by the physiologist. The modern physiologist revels in colloids, osmotic pressure, and hydrogen-ion concentration, and thereby increases our activities by setting us problems which we feel bound to solve’. James Walker, ‘The Rôle of the Physicist in the Development of Chemical Theory’, Journal of the Chemical Society, Transactions, 121 (1922), 735–45 (744–45).

36Ostwald admired and employed Comte's hierarchy of the sciences, and indeed published a biography of Comte. Wilhelm Ostwald, Auguste Comte, der Mann und sein Werk (Leipzig, 1914).

37W.M. Becker, L.J. Kleinsmith, and J. Hardin, The World of the Cell (San Francisco, 2000).

38Smith, Introduction to General Inorganic Chemistry (note 17).

39Among Ostwald's many texts, Smith's most direct source is the Grundlinien (note 5), in which Ostwald sets himself the challenge of addressing beginning students, rather than professional chemists. In his Preface, Smith makes his debt to this work clear, referring to it as a ‘veritable tour de force’. Smith, Introduction (note 17), ix.

40Lyman C. Newell, Descriptive Chemistry (Boston, 1904)

41Harry N. Holmes, General Chemistry (New York, 1921), vii.

42Of these, two were precisely contemporary with Smith's first edition. Ira Remsen, An Introduction to the Study of Chemistry (New York, 1906); A College Text-Book of Chemistry (New York, 1901/1906).

43William McPherson and William Edwards Henderson, An Elementary Study of Chemistry (Boston, 1906)

44Jacob Cornog and J.C. Colbert, ‘What We Teach Our Freshman in Chemistry’, Journal of Chemical Education, 1 (1924). (The first issues of this journal were printed without page numbers.) Well before 1924, McPherson and Henderson had adopted many of Smith's innovations, a development that had not escaped Smith's attention over the years. In 1913, Smith wrote approvingly to Henderson that the latter's new edition was ‘up-do-date in its mode of treatment’ (AS to Henderson, 17 March 1913, ASC, Box 2, Folder 17). In the same year, however, he recommeded it to a correspondent as a suitable high-school level text, and described it to another as ‘systematic, although elementary’ (AS to R. Moldenke, 12 February 1913; AS to RA Wickersham, 14 February 1913, ASC, Box 3, Folder 31). The third text on the list, by Harry N. Holmes (note 41), is also highly derivative of Smith.

45An Associate Professor at the University of Ohio, Henderson had visited Chicago during the summer quarter of 1898. McPherson received his Ph.D. from Chicago in 1899. The University of Chicago, The President's Report, Decennial Publications, First Series, Volume I (Chicago, 1903), 462–63. Both, and most particularly McPherson, would thus have had the opportunity to observe Smith's teaching methods.

46Ira Remsen, An Introduction to the Study of Chemistry (New York, 1900), 65–66.

47Ira Remsen, An Introduction (1906 edition, Note 42), 70–72, 129, 325.

48Newell (note 40), 47.

49Remsen, A College Text-Book (note 42).

50Remsen, A College Text-Book (note 42), 411–12.

51Remsen, A College Text-Book (note 42), 417–23.

52That is, the approximate constancy of the product of specific heat and atomic weight. Remsen, A College Text-Book (note 42), 426.

53McPherson and Henderson (note 43), Chapters XII ‘Solutions’ (172–90), XV ‘Ionization’ (221–31), XVI ‘Some Applications of the Theory of Ionization’ (232–48), and XVII ‘Equilibrium’ (279–96), respectively. These topics together occupy just over 8% of McPherson and Henderson's 718-page text.

54McPherson and Henderson (note 43), 180. This sentence is italicized in the original.

55This makes for considerable awkwardness when the authors attempt to introduce concepts and problems requiring molar concentrations. The discussion of boiling-point elevation informs the reader that a ‘molar concentration of any nonelectrolyte raises the boiling point of water 0.54°’. The authors later set a problem for the reader, requiring calculation of the amount of glycerine required to raise the boiling point of five liters of water by 10°. The problem can be solved by proportional reasoning, but direct recourse to a molar concentration scale would simplify matters greatly. This gap could, of course, be filled by a lecturer, but the text itself does not invite the use of molar concentration units.

56McPherson and Henderson (note 43), 282. This sentence is in italics in the original.

57For example, Harry C. Jones, The Elements of Physical Chemistry (New York, 1902), 453–54.

58‘For example, let us suppose that at equilibrium and at 20° we can show that, expressed in gram-molecules per liter, the following values of concentration are true …’ McPherson and Henderson (note 43), 287. The law can also be demonstrated for gas-phase systems, using partial pressures as concentration terms. Doing so would make quantitative problems involving the mass action law readily accessible to their readers, even without a clearly defined concentration scale for solutions. That they do not indicates that the authors clearly do not expect students to make independent use of the law at this stage in their education.

59That is, ‘[i]f we bring an additional force to bear upon an equilibrium, the point of equilibrium will shift in such a direction as to diminish the intensity of the force so applied’. McPherson and Henderson (note 43), 290.

60Smith, Introduction (note 17), Chapters X ‘Solution’ (145–67), XVII ‘Dissociation in Solution’ (281–99), and XX ‘The Chemical Behavior of Ionic Substances’ (334–66).

61Smith Introduction (note 17), Chapters XV ‘Chemical Equilibrium’ (246–62), XIX ‘Electrolysis’ (310–33), and XXXVIII ‘Electromotive Chemistry’ (663–81), respectively. Over 16% of Smith's 765-page text is thus devoted exclusively to solution theory, equilibrium, and electrochemistry. By the crude measure of page count, Smith devotes approximately twice as much of his book to these topics than McPherson and Henderson. Inclusion of the kinetic-molecular hypothesis, treated by McPherson and Henderson in two pages (64–65) and by Smith in an independent chapter (Chapter IX, 128–44), skews the ratio still further.

62‘The concentration of the molecules is usually expressed, for each substance, in terms of the number (whole or fractional) of moles … of the substance in a liter of the whole mixture.’ Smith Introduction (note 17), 250. The word ‘mole’ does not appear in McPherson and Henderson (1906) according to a search of the Google/Harvard electronic edition (http://books.google.com). The lingering novelty of the word is apparent in this admonition from Smith regarding the proof of one of his papers. ‘The word—“Mole”—in the manuscript was expanded by the printer to “Molecule”. This is not correct. The word—“mole”—is a new one in chemistry and means one gramme molecular-weight. The printer, therefore, must not expand the word or put any period after it, except as required for punctuation.’ Alexander Smith to Secretary of the Royal Society of Edinburgh, December 16, 1914, ASC, Box 4, Folder 45.

63For example, in chapter XV, readers are asked to calculate molecular concentrations for oxygen and nitrogen ‘in the air’, for ‘aqueous vapor above water’ at various temperatures, and for solutions of various composition. They are then asked to determine ‘the partial pressures of the three components of phosphorus pentachloride vapor [i.e. PCl5, PCl3, and Cl2] at 250° and 760 mm’. Chapter XVII poses a complex but realistic question requiring interpretation of a colligative property to determine an equilibrium constant, then application of that constant to determine solution composition. ‘What is the degree of dissociation of zinc sulphate if 5 g. of it dissolved in 125 g. of water produce a lowering of 0.603° in the f[reezing]-p[oint]? What is the molecular concentration of each of the three substances present in the solution?’ Smith, Introduction (note 17), 261, 299.

64For example, from Chapter XXVI ‘Oxides and Oxygen Acids of Nitrogen’: ‘At 27°, what proportions of nitrogen tetroxide are in the forms of NO2 and N2O4 respectively …? At the same temperature what fraction of the material, by weight, is in the former condition?’ Smith, Introduction (note 17), 454.

65Smith pauses at two points in the text (after Chapters XI and XV) to make sure the reader appreciates the extent to which this is the case, by providing a list of 23 ‘principles (general facts) of chemistry’ in terms of which the descriptive information in the text is to be interpreted and understood. Smith, Introduction (note 17), 188–89, 262.

66Wilhelm Ostwald, The Fundamental Principles of Chemistry. An Introduction to All Text-Books of Chemistry (New York, 1909), vi, emphasis added.

67Ostwald (note 66), vii.

68Ostwald, Grundlinien (note 5), vii.

69Like Ostwald's, Smith's severe anti-realism applies as much to statements about natural laws as to statements about entities. ‘A law … is a condensed statement describing some constant mode of behavior. It is simply a summary of our experience. … A common form of the law [of conservation of mass], to the effect that “the mass of matter in the universe is unchangeable in amount,” is not a law at all, in the only sense in which the word is used in science. It is a statement in regard to supposed facts which are almost entirely beyond our experience. It is, therefore, a proposition of a transcendental … nature, and has its proper place in metaphysics.’ Smith, Introduction (note 17), 17–18.

70Alexander Smith, ‘The Rehabilitation of the American College, and the Place of Chemistry in It’, Science, 30 (1909), 457–66.

71Smith (note 70), 466, italics in original.

72Smith (note 70), 458.

73Smith (note 70), 460, italics added.

74Some readers will note that Smith's career trajectory made him a colleague of John Dewey at both the University of Chicago (from Smith's arrival until Dewey's departure in 1904) and at Columbia University (after Smith's arrival in 1911), and will wonder—as do I—at the extent of Dewey's influence on Smith's educational ideas and practices. Dewey's stewardship of the Laboratory School at the University of Chicago, after all, was contemporaneous with Smith's study of secondary chemistry curricula for the NEA. Robert Westbrook provides a succinct summary of Dewey's educational work and thought during his years at Chicago. Robert B. Westbrook, John Dewey and American Democracy (Ithaca, 1991), 93–113. Dewey's first two important works on education—The School and Society (1899) and The Child and the Curriculum (1902)—appeared during his Chicago years. Smith must have found congenial the Deweyan conceptions of thinking as ‘an instrument for solving the problems of experience’ and of knowledge as ‘the accumulation of wisdom that such problem-solving generated’ (94). On the other hand, Dewey's vision of the schools as social levelers, as ‘powerful adversarial institutions in the heart of American culture’ (110), would have been anathema to Smith's notions of pre-collegiate enculturation. Direct evidence of Smith's responses to Dewey's work would greatly illuminate his own educational thought. It is frustrating to report that no substantial correspondence between the men seems to have survived, if any ever existed.

75Abraham Flexner, The American College (New York, 1908). The impact of this book can be measured not least in the effect it had on Flexner's nascent career. After receiving his B.A. at Hopkins in 1886, Abraham Flexner (1866–1959) founded his own successful prep school. Subsequently, he earned a master's degree in psychology at Harvard (1906). A year in Berlin comparing the German and US university systems produced The American College. The book attracted attention, and an offer of employment, from the young Carnegie Foundation for the Advancement of Teaching, for whom Flexner wrote his most influential work, Medical Education in the United States and Canada (1910). The ‘Flexner report’ exposed the poor state of medical education in the US, resulting in the rapid closure of dozens of substandard medical schools. Through his subsequent work disbursing funds for medical education on behalf of John Rockefeller, Jr's General Education Board, Flexner was able to oversee widespread adoption of his reforms. His final professional position was as the founding director of the Institute for Advanced Study at Princeton (1930–1939). See Michael R. Harris, ‘Flexner, Abraham’, in Dictionary of American Biography, Supplement Six (New York, 1980), 207–9.

76Flexner, American College (note 75), 134–35.

77That the broadening of the collegiate mission is central to Flexner's argument is perhaps most evident when he foresees ‘a differentiation among colleges, each of which will then perhaps no longer seek to be all things to everybody’. Abraham Flexner, ‘Adjusting the College to American Life’, Science, 29 (1909), 361–72 (366).

78On the contrary, the graduate schools represent ‘achievement of the highest order. But not college achievement’. Flexner, ‘Adjusting the College,’ (note 77), 363.

79Flexner, ‘Adjusting the College’ (note 77), 369.

80Smith, ‘Rehabilitation’ (note 70), 459. Flexner here quotes ‘with approval’ the Chicago physicist Charles Riborg Mann, whose curricular innovations have been recently described by John Rudolph. John L. Rudolph, ‘Turning Science to Account: Chicago and the General Science Movement in Secondary Education, 1905–1920’, Isis, 96 (2005), 353–89

81Smith, ‘Rehabilitation’ (note 70), 459.

82‘[M]y meaning will be clear from an educational point of view when I say that the cultural importance of biology to the college student comes out when, in addition to his mastery of biological science as such, its history, its applications, its influence on the development of thought, have been explicitly brought forward; when, in other words, the vocational bearing and the social significance of the vocation in question supervene upon the strictly scientific study.’ Flexner, ‘Adjusting the college’ (note 77), 368.

83‘When we talk of adjusting the college to life, we mean in plain language working out a concrete educational scheme which will adjust each individual boy to the concrete social situation. … The college is in this respect but the culmination of, not fundamentally disconnected from, the elementary and high schools.’ Flexner, ‘Adjusting the College’ (note 77), 364.

84Flexner, American College (note 75), 148.

85Flexner, ‘Adjusting the College’ (note 77), 364–65.

86Flexner, ‘Adjusting the College’ (note 77), 369.

87Smith, ‘Rehabilitation’ (note 70), 459.

88Alexander Smith, Heredity and Education (1892). Pamphlet, Robert T. Ramsay, Jr, Archival Center, Wabash College, Crawfordsville, IN (WCA); Folder DC791dd.

89On the broad influence of recapitulationism in late nineteenth-century thought, see Stephen Jay Gould, The Mismeasure of Man (New York, 1981), 113–22.

90Smith, Heredity and Education (note 88), 7.

91For the most part, anyway. For example, in considering whether to alter the curriculum of the secondary schools, Smith wrote in 1904 that it was critical to determine ‘that we have no reason to believe on historical grounds that the curriculum has become so much a part of the development of the race that any change in it would produce serious organic disturbance’. Smith and Hall (note 15), 14.

92Smith, ‘Articulation of School and College Work’ (note 14), 453–4.

93Smith, Heredity and Education (note 88), 13.

94‘The business and glory of the college are … not stupidly to ignore or vainly to resist the vocational factor, but deliberately to develop in advance its cultural meaning and possibilities’. Flexner, ‘Adjusting the College’ (note 77), 369.

95Smith and Hall (note 15), 8–18; 64–67. Preparation for collegiate chemistry is, in fact, the least important among the ends Smith identifies for the school course. ‘There is no reason why the college should demand a knowledge of formal chemistry for admission’ (66). Far more important are ‘preparation for life’, entailing a set of observational and intellectual habits, and ‘preparation for teaching’ in the ‘lower schools’, entailing the ability to contextualize observations made by children during the course of nature study. ‘[H]ere a formal knowledge of chemistry only hampers [the teacher], unless she has happened to have the opportunity to go further into the subject’ (65).

96Smith and Hall (note 15). ‘[I]t may be a question whether, in the average [school] course, time will be found for anything more than a touch of the subject [physical chemistry] here and there’ (170). Stoichiometry comprises the theoretical core of Smith's secondary curriculum, although experience working with materials in the laboratory is of far greater importance to him (69–84, 87–94).

97Alexander Smith, ‘Rehabilitation’ (note 70), 466.

98The details of this crisis are omitted here, despite their relevance to the relationship between Smith and John Ulric Nef before and immediately after Smith's call to Chicago, partly for the sake of brevity. As a reviewer noted, they belong more properly to a study of Nef's own work and career. I am grateful for the suggestion.

99R.J. Storr, Harper's University: the Beginnings (Chicago, 1966), 72. Nef expressed his ambitions in ‘Important Factors in the Development of a Research Laboratory: An Address by Professor John U. Nef, Sr., on the Occasion of the Exercises in Connection with the Opening of the Kent Chemical Laboratory, Jan. 1, 1894’; JUN, Box 2, Folder 9.

100Typescript, ‘John Ulric Nef—Written for Dr. Spoehr by his son John U. Nef, Jr’; JUN, Box 2, Folder 10.

101 The Wabash, Volume XVII, No. 2, November 1892, 43; WCA._

102J.I. Osborne and T.G. Gronert, Wabash College: The First Hundred Years, 1832–1932 (Crawfordsville, IN, 1932), 193.

103John U. Nef to William R. Harper, 16 February 1894; JUN, Box 1, Folder 4.

104Alexander Smith, ‘Growth and Progress of University Extension in Great Britain’, The Wabash, 16 (1891), 1; WCA.

105Late in 1893, Smith can be seen promoting the novelty and difficulty of his current research to Nef; this might be a hint that he was consciously angling for a job. Alexander Smith to John U. Nef, 2 November 1893; JUN, Box 1, Folder 3.

106A helpful review of this period with many leading references can be found in Arthur Michael, ‘Zur Geschichte der Theorie über die Bildung und Constitution des Natracetessigesters’, Berichte der deutschen chemischen Gesellschaft, 38 (1905), 1922–1937.

107William Henry Perkin, Sr., ‘The Magnetic Rotation of Compounds Supposed to Contain Acetyl, or to Be of Ketonic Origin’, Journal of the Chemical Society, Transactions, 61 (1892), 800–64 (801).

108Maitland Jones, Jr, Organic Chemistry, Second Edition (New York, 2000), 827. The tautomer problem illustrates both the limits of nineteenth-century organic structural thinking, and foreshadows the ways in which the emergence of physical organic chemistry would lead to a powerful extension of structural chemistry's explanatory reach. Conceptually, the most important contribution physical chemistry would make to the resolution of this structural problem would be the loosening of the restriction that one compound should have one and only one structure; in this case, by allowing the notion that compounds may exhibit reversible, unimolecular reactivity in solution that is not evident from their solid-state structures (i.e. tautomeric equilibria). Experimentally, the most important contribution would be methods for the detection of functional groups that did not rely upon derivation or degradation, but could be applied to substances in solution without chemically perturbing them (i.e. spectroscopy).

109D.S. Tarbell and Ann T. Tarbell, Essays on the History of Organic Chemistry in the United States, 1875–1955 (Nashville, TN, 1986), 48.

110Perkin (note 107), 808.

111John U. Nef, ‘Ueber die 1,3-Diketone’, Justus Liebigs Annalen der Chemie und Pharmacie, 277 (1893), 59–78.

112Ludwig Claisen, ‘Ueber die Constitution des Acetessigäthers und der sogenannten Formylverbindungen der Säureäther und Ketone’, Berichte der deutschen chemischen Gesellschaft, 25 (1892), 1776–1787; ‘Beiträge zur Kenntniss der 1,3-Diketone’, Justus Liebigs Annalen der Chemie und Pharmacie, 277 (1893), 162–206.

113Although I use modern structural conventions for clarity, I do not mean to imply that the arguments are valid by contemporary standards.

114As in the case of the 1,3-diketones themselves, the properties of their salts could only confound a sharply dualistic view of their structures, and they resisted convincing interpretation until the rise of quantum mechanical descriptions of molecular structure. One modern depiction of their structure illustrates their ambiguous reactivity by representing them as a blend of resonance structures, each representing one aspect of the behaviour of the whole. The metal typically resides between the two oxygen atoms, but the salt is reactive at both oxygen atoms and at the central carbon. When the metal ion is displaced by an organic group, up to three structural isomers can be obtained, depending in a complex fashion on the diketonate itself, the reagent, and the conditions under which the reaction takes place. The location of the metal ion in the salt, whether in solution or the solid state, has little or no necessary bearing on the reactivity of the salt with a particular reagent. This mode of explanation clearly depends upon the prior formulation of the concept of resonance and resonance structures, and would not be available to organic chemists until after Pauling's pioneering work.

115A challenge that, in itself, is entirely reasonable, though a modern chemist would raise it on rather different grounds than those put forward by Nef.

116‘The hydroxyl structure gives a perfectly adequate account of all the reactions of these substances’. John U. Nef, ‘Zur Kenntniss des Acetessigäthers’, Justus Liebigs Annalen der Chemie und Pharmacie, 266 (1891), 52–138 (68).

117There is still no a priori reason for favouring one kind of simplicity over the other in the interpretation of organic phenomena in the absence of direct evidence favouring one or the other.

118Emil Fischer and C. Bülow, ‘Ueber das Benzoylaceton’, Berichte der deutschen chemischen Gesellschaft, 18 (1888), 2131–38.

119Ludwig Claisen and O. Lowman, ‘Zur Kenntniss des Benzoylacetons’, Berichte der deutschen chemischen Gesellschaft, 21 (1891), 1153–57.

120Alexander Smith to John U. Nef, 17 June 1893; JUN, Box 1, Folder 3.

121Alexander Smith to John U. Nef, 21 June 1893; JUN, Box 1, Folder 3.

122Claisen, ‘Beiträge’ (note 112), 186.

125Alexander Smith to John U. Nef, 21 June 1893; JUN, Box 1, Folder 3.

123Alexander Smith to John U. Nef, 17 June 1893; JUN, Box 1, Folder 3.

124Alexander Smith to John U. Nef, 22 September 1893; JUN, Box 1, Folder 3.

126Nef, ‘Ueber die 1,3-Diketone’ (note 111), 66.

127Claisen identified the oil as O-benzoyl dibenzoylacetone, i.e. the product of reaction between two equivalents of benzoyl chloride and one equivalent of sodium benzoylacetate.

128Alexander Smith to John U. Nef, 17 June 1893; JUN, Box 1, Folder 3.

129Noyes, ‘Smith, Alexander’ (note 9), 235.

130Smith (1894), ‘On Two Stereoisomeric Hydrazones’ (note 25).

131Smith (1896), ‘Ueber die Einwirkung’ (note 25).

132John U. Nef, ‘Ueber das zweiwerthige Kohlenstoffatom, II.’, Liebigs Annalen der Chemie, 280 (1894), 291–342; ‘Ueber das zweiwerthige Kohlenstoffatom, III. Die Chemie des Cyans und des Isocyans’, Liebigs Annalen der Chemie, 287 (1895), 265–359; ‘Ueber das zweiwerthige Kohlenstoffatom, IV. Die Chemie des Methylens’, Liebigs Annalen der Chemie, 298 (1897), 202–374.

133Smith to Elihu Thompson, 21 December 1915 and 12 January 1916, Alexander Smith Correspondence, 1900–1919, Columbia University Archives—Columbiana Library (ASC), Box 1, Folder 3. Smith's lack of personal regard for Nef may further be gauged from the fact that he omitted to send a collegial letter of condolence to the University of Chicago at the time of Nef's funeral. Not only was this social grace observed by most of the chairmen of prestigious chemistry departments, but Smith was a prominent former member of the bereaved department. He was also a highly sociable man, well versed in the formalities of professional interaction.

134W.A. Noyes to H.N. Morse, 26 February 1914; WAN, Box 4.

135H.I. Schlesinger, ‘Reminiscences: Early Days in the Chemistry Department’, Department of Chemistry Newsletter (1958), 1–5.

136Smith to Harper, 30 June 1901; Harper to Smith, 2 November 1900; 26 March 1904; UPC, Box 16, Folder 1.

137Typescript, ‘John Ulric Nef’ (note 100).

138University of Chicago, Board of Trustees minutes, 20 October 1903.

139Smith to Harper, November 23, 1905; H.P. Judson to Smith, 7 February 1906; UPC, Box 16, Folder 1.

140National Education Association, ‘Report of the Committee on College-Entrance Requirements’ (note 12), 167, quoted in DeBoer (note 8), 53, emphasis added.

142Smith to H.P. Talbot, March 22, 1913; ASC, Box 3, Folder 115.

141Alexander Smith, ‘‘The Training of Chemists’, Science, 43 (1916), 619–26 (621).

143Hypophosphorous acid does not exist as a pair of tautomers, because the two plausible isomers do not interconvert. The more highly saturated structure for hypophosphorous acid, containing the terminal oxygen and two direct phosphorus–hydrogen bonds, is the correct one. F. Albert Cotton and Geoffrey Wilkinson, Advanced Inorganic Chemistry (New York, 1980), 472.

144Smith, Introduction (note 17), 470; Introduction to General Inorganic Chemistry (New York, 1917), 562.

148Smith to Sir Donald McAllister, 13 January 1919; ASC, Box 3, Folder 49.

145Wilhelm Ostwald, ‘The Historical Development of General Chemistry. Fourth Lecture. Constitution and Isomerism (Organic Chemistry)’, School of Mines Quarterly, 27 (1906), 326–339 (338).

146Smith to J.H. Kirkland, 30 June 1916; ASC, Box 3, Folder 8.

147Smith to Mr Taylor, 1 September 1919; ASC, Box 4, Folder Folder 26.

149Smith to Thomas H. Laby, May 10, ASC; CDC, Box 3, Folder 11.

150No value recurs with more intensity in Smith's evaluations of individual chemists, as well as groups of chemists, than scrupulous attention to experimental precision. Smith privately criticized no less a chemist than the Nobel laureate and discoverer of argon, Sir William Ramsay, for experimental slovenliness, refusing to review a Ramsay biography on the grounds that he could not in good conscience show the level of respect that would be expected in such a work. ‘I did not wish to write a review of Ramsay's life, because his work is very careless, and I did not care to praise it. People, who have never read his papers, think his work is wonderful, so you had better send it to someone who has never read his papers’. Smith to R.M. Lovett, 4 January 1919; ASC, Box 2, Folder 83.

151Walker, ‘The Rôle of the Physicist’ (note 35), 737.

152Smith, too, was alert to the symbolic uses of the past in establishing the present significance of physical chemistry. He devoted his inaugural address as President of the American Chemical Society to a historical treatment of the then (even more than now) obscure Russian chemist, Michael Lomonossoff (1711–1765). Alexander Smith, ‘An Early Physical Chemist—M.W. Lomonossoff’, Science, 35 (1912), 121–29. The purpose of the address was to introduce American chemists to ‘a chemist of the first magnitude and a personality of marvelous force and range,’ worthy of addition ‘to the limited gallery of the world's very greatest men’, and at the same time to place him among the intellectual forefathers of the new physical chemistry.

153See correspondence in ASC, Box 1, Folders 26–28.

154Smith to A.A. Noyes, 15 January 1912; ASC, Box 1, Folder 8.

155The most substantial archival source for Smith is a recently discovered collection of his correspondence while Chairman of the Department of Chemistry at Columbia University (ASC). This collection provides a great deal of detailed information about Smith's professional activities after 1911, but has less to offer concerning his earlier career. His years at the University of Chicago are thinly represented via surviving correspondence with William Rainey Harper (UPC) and John Ulric Nef (JUN). Smith's early collaboration with Nef is one of few incidents that can be reconstructed with confidence, and this paper makes use of almost every document by or concerning Smith in these two collections. None of Smith's correspondence survives at Wabash College, though some information about these years can be gleaned from College and student publications of the period, and from later correspondence with old colleagues in ASC. Thus, Smith's most important and productive years—those prior to the 1906 publication of the General Inorganic Chemistry—are also the most difficult years in his professional life to document.

156Smith and Hall (note 15), 11.

157Alexander Smith, quoted in Wilder D. Bancroft, ‘A Laboratory Outline of General Chemistry (Book Review)’, Jounal of Physical Chemistry, 14 (1909), 587–588.

158Servos (note 2), 87–99.

159John W. Servos, ‘A Disciplinary Program That Failed: Wilder D. Bancroft and the Journal of Physical Chemistry, 1896–1933’, Isis, 73 (1982), 207–32, and Physical Chemistry (note 2), Chapter 6.

160Harry C. Jones, The Elements of Physical Chemistry (New York, 1902); Arthur A. Noyes, General Principles of Physical Science: an Introduction to the Study of the General Principles of Chemistry (New York, 1902); Arthur A. Noyes, Course of Instruction in the General Principles of Chemistry.Printed in Preliminary Form for the Classes of the Massachusetts Institute of Technology (Boston, 1917); Arthur A. Noyes and M.S. Sherrill, Advanced Course of Instruction in Chemical Principles (New York, 1922).

161Wilder D. Bancroft, ‘General Principles of Physical Science (Review)’, Journal of Physical Chemistry, 7 (1903), 217–18.

162Arthur A. Noyes and S.P. Mulliken, Laboratory Experiments on the Class Reactions and Identification of Organic Substances (Easton, PA, 1898); Arthur a. Noyes, a Detailed Course of Qualitative Chemical Analysis of Inorganic Substances (New York, 1899).

163Smith to Julius Stieglitz, 2 October 1918; ASC, Box 3, Folder 110.

164Smith to Julius Stieglitz, 2 October 1918; ASC, Box 3, Folder 110.

165Smith to A.H. Bernhard, 22 October 1918; ASC, Box 2, Folder 83.

166Servos, Physical Chemistry (note 2), 99.

167Servos, Physical Chemistry (note 2), 88–89.

168Servos, Physical Chemistry (note 2), 88–89.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.