Hostname: page-component-848d4c4894-cjp7w Total loading time: 0 Render date: 2024-06-28T04:09:46.392Z Has data issue: false hasContentIssue false

British women of the nineteenth and early twentieth centuries who contributed to research in the chemical sciences

Published online by Cambridge University Press:  05 January 2009

Mary R. S. Creese
Affiliation:
1650 Cambridge Road, Lawrence, Kansas 66044, USA.
Rights & Permissions [Opens in a new window]

Extract

Core share and HTML view are not available for this content. However, as you have access to this content, a full PDF is available via the ‘Save PDF’ action button.

Apart from a few outstanding people from before 1850, British women of the nineteenth and early twentieth centuries who published work in the chemical sciences have not received much attention so far. The university-trained women who, from about 1880 onwards, authored or co-authored an increasing number of original research contributions have been largely ignored, and their names are for the most part omitted from biographical reference works and science histories. There are several works describing the changes and developments in university-level education for women during this period, but these are not specially concerned with science education or with the careers of individuals.

Type
Research Article
Copyright
Copyright © British Society for the History of Science 1991

References

1 See Patterson, Elizabeth Chambers, Mary Somerville and the Cultivation of Science, 1815–1840, The Hague, 1983CrossRefGoogle Scholar; Armstrong, Eva V., ‘Jane Marcet and her “Conversations on Chemistry”’, Journal of Chemical Education, (1938), 15, pp. 53–7CrossRefGoogle Scholar; Crellin, J. K., ‘Mrs. Marcet's “Conversations on Chemistry”’Google Scholar, ibid. (1979), 56, pp. 459–60; Ogilvie, Marilyn Bailey, Women in Science. Antiquity through the Nineteenth Century, Cambridge, Mass., 1986, pp. 8990, 125–7, 161–6.Google Scholar

2 See Tuke, Margaret J., A History of Bedford College for Women, 1849–1937, London, 1939Google Scholar; McWilliams-Tullberg, Rita, Women at Cambridge. A Men's University – Though of a Mixed Type, London, 1975Google Scholar; Gardner, A., A Short History of Newnham College, Cambridge, 1921Google Scholar; Hamilton, Mary Agnes, Newnham. An Informal Biography, London, 1936Google Scholar; Stephen, Barbara, Girton College. 1869–1932, Cambridge, 1933Google Scholar and Emily Davies and Girton College, London, 1927Google Scholar; MacLeod, Roy and Moseley, Russell, ‘Fathers and daughters: Reflections on women, science and Victorian Cambridge’, History of Education, (1979), 8, no. 4, pp. 321–33CrossRefGoogle ScholarPubMed; Tylecote, Mabel, The Education of Women at Manchester University 1883 to 1933, Manchester, 1941Google Scholar; Brittain, Vera, The Women at Oxford, New York, 1960Google Scholar; Byrne, Muriel St Clare and Mansfield, Catherine Hope, Somerville College, 1879–1921, Oxford, 1922?Google Scholar; Bailey, Gemma, Lady Margaret Hall. A Short History, London, 1923.Google Scholar

3 See Patterson, , op. cit (1)Google Scholar, Armstrong, , op. cit (1)Google Scholar and Ogilvie, , op. cit. (1).Google Scholar

4 Somerville's most celebrated work was The Mechanism of the Heavens, London, 1831Google Scholar, her annotated translation of Laplace's Méchanique Céleste. For this she wrote a preface, also published separately as A Preliminary Dissertation on the Mechanism of the Heavens, London, 1832Google Scholar, which introduced the reader to the mathematics necessary for understanding the main work. She also published three papers reporting her experiments on the effects of solar radiation: ‘On the magnetizing power of the more refrangible solar rays’, Philosophical Transactions, (1826), pt 2, pp. 132–9Google Scholar, Annales de Chimie, (1826), 31, pp. 393400Google Scholar; ‘Expériences sur la transmission des rayons chimiques du spectre solaire, à travers différentes milieux’, Comptes rendus de l'Académie des Sciences, (1836), 3, pp. 473–6Google Scholar, New Philosophical Journal, Edinburgh (1837), 22, pp. 180–3Google Scholar; ‘On the action of the rays of the spectrum upon vegetable juices’, Philosophical Transactions, (1846), pp. 111–21.Google Scholar

5 Ogilvie, , op. cit. (1), pp. 8990Google Scholar; Davenport, Derek A. and Ireland, Kathleen M., ‘The ingenious, lively and celebrated Mrs Fulhame, and a dyer's hand’, Bulletin for the History of Chemistry, (1989), No. 5, pp. 3742Google Scholar; Partington, J. R., A History of Chemistry, 4 vols., London, 1962, iii, pp. 708–9.CrossRefGoogle Scholar

6 MrsFulhame, , An Essay on Combustion, with a View to a new Art of Dyeing and Painting. Wherein the Phlogistic and Antiphlogistic Hypotheses are Proved Erroneous, London, 1794 (German tr. Lentin, A. G. W., Göttingen, 1798Google Scholar; American edn Philadelphia, 1810).

7 Coindet, J. F., ‘De l'ouvrage de Mme Fulhame, intitulée: An Essai on Combustion, etc. Essai sur la combustion avec des vues…’, Annales de Chimie, (1798), 26, pp. 5885.Google Scholar

8 Mellor, J. W., ‘History of the water problem (Mrs Fulhame's theory of catalysis)’, Journal of Physical Chemistry, (1903), 7, pp. 557–67.CrossRefGoogle Scholar

9 An examination of the Royal Society's Catalogue of Scientific Papers, 18001900, 19 vols., Cambridge, 18671925, ixiiGoogle Scholar, uncovered one paper written by a British woman on a chemical topic during the period 1846–80, namely Marshall, Margaret H.'s, ‘Experiments and investigations as to the influence exerted over some minerals by animal and vegetable matter, under certain conditions’, Proceedings of the Royal Society of Edinburgh, (1851), 2, pp. 5861.CrossRefGoogle Scholar

10 As contrast, in both botany and geology the number of publications by British women showed a noticeably more gradual increase over the course of the century; my unpublished data.

11 Tuke, , op. cit. (2), esp. pp. 18, 127–8, 138–9Google Scholar; McWilliams-Tullberg, , op. cit. (2), pp. 7084Google Scholar. Women at Cambridge were permitted to take the Tripos examinations informally throughout the 1870s. Their names were omitted from official pass lists.

12 My unpublished data.

13 See letter from Davies, Emily to Bodichon, Mme, 09 6, 1873Google Scholar, quoted by Stephen, , Emily Davies and Girton College, op. cit. (2), p. 282Google Scholar; also MacLeod, and Moseley, , op. cit. (2), p. 326.Google Scholar

14 McWilliams-Tullberg, , op. cit. (2), p. 143Google Scholar. The early women students at Cambridge came predominantly from the middle classes, especially from industrial, commercial, and professional families of the north of England and London and the south-east (see MacLeod, and Moseley, , op. cit. (2), p. 329)Google Scholar. These families had ‘an unerring instinct for “quality”’ (McWilliams-Tullberg, , op. cit. (2), p. 25)Google Scholar, an attribute which the older universities, whose graduates continued to dominate public life in England at least until the First World War, were seen to embody.

15 Many Cambridge women students found themselves obliged to take the London degree as well in order to obtain the needed degree titles.

16 Candidates for ordinary degrees were not accepted at the women's colleges though, as in the case of the men, failure to meet the required standard in the honours examinations could result in a pass at the ordinary level; McWilliams-Tullberg, , op. cit. (2), p. 222.Google Scholar

17 Stephen, , Emily Davies and Girton College, op. cit. (2), p. 358Google Scholar, and McWilliams-Tullberg, , op. cit. (2), p. 105.Google Scholar

18 Ibid., pp. 179–80. See also MacLeod, and Moseley, , op. cit. (2), pp. 330–1.Google Scholar

19 Berry, A. J. and Moelwyn-Hughes, E. A., ‘Chemistry at Cambridge from 1901 to 1910’, Proceedings of the Chemical Society, (12 1963), pp. 357–63, esp. p. 357.Google Scholar

20 The inclusion of work from these border areas in the Chemical Society's indices of the chemical literature of the period further shows that it was considered part of chemistry (rather than physics or physiology) at the time. All of the work considered in this study was so indexed. Kohler, Robert E. in, From Medical Chemistry to Biochemistry. The Making of a Biomedical Discipline, Cambridge, 1982CrossRefGoogle Scholar, has discussed the great complexity of the interrelationship between chemistry and physiology and the new field of chemical physiology or biochemistry. The latter was in the process of developing from the two older disciplines in the last decade of the nineteenth century and the early years of the twentieth.

21 See Stephen, , Girton College, 1869–1932, op. cit. (2), p. 172Google Scholar, and W. H. B., ‘Mrs G. P. Bidder’, Nature, (1932), 130, pp. 689–90.CrossRefGoogle Scholar

22 MacLeod, and Moseley, , op. cit. (2), p. 331.Google Scholar

23 W. H. B., op. cit. (21), p. 689.Google Scholar

24 Girton College Register, 1869–1946, p. 47Google Scholar, and Girton College Review, Easter Term, 1953Google Scholar. I thank Ms Kate Perry, Archivist, Girton College, for giving me information about Tebb.

25 See Edsall, J. T. in The Origins of Biochemistry: A Retrospect on Proteins (ed. Srinivasan, P. R., Fruton, J. S. and Edsall, J. T.), New York, 1979, pp. 54–5.Google Scholar

26 Girton College Register, 1869–1946, p. 49Google Scholar, and Girton College Review, Michaelmas Term, 1949, p. 35Google Scholar. I thank Ms Kate Perry for providing information on Durham.

27 See Vardy, Winifred I., King Edward VI High School for Girls Birmingham 1883–1925, London, 1928Google Scholar, especially the Appendix, pp. 118–32; Stephenson, Marjory, ‘Muriel Wheldale Onslow (1880–1932)’, Biochemical Journal, (1932), 26, pp. 915–16Google Scholar; M[ary] E. de R. E[pps], ‘Ida Smedley MacLean’, Newnham College Roll Letter (01 1945), pp. 50–1.Google Scholar

28 E[pps], de R., op. cit. (27)Google Scholar. See also Whiteley, M. A., ‘Ida Smedley MacLean. 1877–1944’, Journal of the Chemical Society, (1946), Pt 1, pp. 65–7Google Scholar; Chick, Harriette, Hume, Margaret and Macfarlane, Marjorie, War on Disease. History of the Lister Institute, London, 1971, pp. 164–6Google Scholar; Newnham College Register, 1896Google Scholar. I thank Miss Ann Phillips, Librarian, Newnham College Archives, for information about Smedley.

29 Vardy, , op. cit. (27), p. 26.Google Scholar

30 The total amount awarded that year was £230 spread over seventeen people, making an average award of £13 105. Smedley received similar awards in 1909 and 1910 of £10 and £15 respectively. A total of five women chemists received Chemical Society Research Fund grants during the period 1902–10. Besides Smedley they were Alice Emily Smith, Owens College Manchester and University College, Bangor (1905), Annie Homer, Newnham (1907, 1908), and Mary Elizabeth Dobson (1909, 1910) and Gertrude Maude Walsh (1910), both from Manchester University. See Research Fund Income and Expenditure Accounts, in reports of Annual General Meetings, Journal of the Chemical Society, Transactions, (1903), 83, p. 635Google Scholar; ibid. (1906), 89, p. 743; ibid. (1908), 93, p. 722; ibid. (1909), 95, p. 621; ibid. (1910), 97, p. 659; ibid. (1911), 99, p. 585.

31 Chick, , Hume, and Macfarlane, , op. cit. (28), p. 165.Google Scholar

32 E[pps], de R., op. cit. (27), p. 51.Google Scholar

33 Stephenson, , op. cit. (27).Google Scholar

34 Richter, Derek, ‘Opportunities for Women in Science’, in Women Scientists. The Road to Liberation (ed. Richter, Derek), London, 1982, p. 7.CrossRefGoogle Scholar

35 Chick, , Hume, and Macfarlane, , op. cit. (28), pp. 124, 234Google Scholar. See also Vardy, , op. cit. (27), pp. 120–2.Google Scholar

36 See p. 566 in Robertson, Muriel, ‘Marjory Stephenson (1885–1948)’, Obituary Notices of Fellows of the Royal Society, (1949), 6, pp. 563–77.CrossRefGoogle Scholar

37 Needham, Dorothy, ‘Women in Cambridge Biochemistry’Google Scholar, in Richter, (ed.), op. cit. (34), pp. 158–63, esp. p. 161Google Scholar. Kohler, (op. cit. (20), p. 88)Google Scholar comments that in the 1930s there was a certain amount of difficulty in Hopkins’ Department with financial support, in part the ‘normal problems of a large and extremely diverse group supported by a plethora of grants and fellowships and held together by “Hoppy's” gentle charisma’. Nevertheless, ‘Dorothy Needham carried on her first-rate work on a third-rate salary …and Marjory Stephenson was constantly having to beg for small sums…’

38 See, for instance, Love, Rosaleen, ‘“Alice in Eugenics-Land”: feminism and eugenics in the scientific careers of Alice Lee and Ethel Elderton’, Annals of Science, (1979), 30, 145–58CrossRefGoogle Scholar. Alice Lee, a student of Karl Pearson at University College and one of the first women to receive a London D.Sc. (1901), continued her research in Pearson's statistics laboratory mostly on a voluntary basis while supporting herself as a lecturer at Bedford College. When ill-health forced her to retire and she was left impecunious, Pearson, in 1923, petitioned the Home Office for a £70 per annum Civil List pension for her, stating that ‘few, if any, women workers of her period have accomplished as large a bulk of first class research as Dr Lee’ (Pearson papers, 629; Love, note 45, p. 152). Lee's colleague, Ethel Elderton, a full-time member of the statistics laboratory staff, earned, in 1906, £100 per annum. A Research Fellow received £250 per annum, a salary considered acceptable for men. Despite Pearson's exceptionally strong commitment to equality for women it is clear from his correspondence that he considered such low-paid (or unpaid) posts ‘well suited to a woman living with her family in London and keen on scientific work’ (letter from Pearson, to Cave, A., 25 11 1907Google Scholar, Pearson papers CID7 (1); Love, note 78, p. 157). For comparison, the post of assistant mistress in a school providing secondary education then paid on average about £118 per annum.

39 Sidgwick, Ethel, Mrs Henry Sidgwick. A Memoir by her Niece, London, 1938.Google Scholar

40 Ibid., p. 72.

41 Rayleigh, Lord, FRS and MrsSidgwick, H., ‘On the specific resistance of mercury’, Philosophical Transactions, (1883), 174, pt 1, pp. 173–85CrossRefGoogle Scholar; ‘On the electrochemical equivalent of silver and the absolute electromotive force of Clark cells’, ibid. (1884), 175, pp. 411–60.

42 Stephen, , Girton College, 1869–1932, op. cit. (2), p. 176Google Scholar, and Newnham College Register, Staff, 1891Google Scholar. See also Wilson, H., ‘Miss Freund’Google Scholar, and Ball, M. D., ‘Newnham Scientists’, in A Newnham Anthology (ed. Phillips, Ann), Cambridge, 1971, pp. 71–2 and 76–8Google Scholar, respectively. I thank Miss Ann Phillips for the information from the Newnham College Register.

43 Freund, Ida, ‘Der Einfluss der Temperatur auf die Volumenänderung bei der Neutralisation für verschiedene Salre bei verscheidenen Konzentrationen’, Zeitschrift für physikalische Chemie, (1909), 66, pp. 555613.Google Scholar

44 Needham, , op. cit. (37), p. 158.Google Scholar

45 Berry, and Moelwyn-Hughes, , op. cit. (19), pp. 362–3.Google Scholar

46 Numbers of papers published by individual women are indicated in Appendix 1.

47 I thank Miss Carol R. Bowen, Records Office, University College, London, for information about Aston.

48 Burstall, F. H., ‘Frances Mary Gore Micklethwait (1868–1950)’, Journal of the Chemical Society, (1952), 3, pp. 2946–7Google Scholar. See also Who was Who (19411950), London, 1952, iv, p. 791.Google Scholar

49 Probably the five most productive women in research in the chemical sciences in this period were Marie Curie, the Russian biochemist Nadezda Sieber-Schumov, Frances Micklethwait, the American public health chemist Ellen Swallow Richards, and the Swedish chemist Astrid Cleve von Euler; my unpublished data.

50 See Who was Who (19511960), London, 1961, v, p. 1162.Google Scholar

51 Information about Haynes, Judd, Homfray, Marsden and Renouf was collected from their technical papers.

52 See Copping, A. M., ‘Dame Harriette Chick’, British Journal of Nutrition, (1978), 39, pp. 34CrossRefGoogle ScholarPubMed; ‘Dame Harriette Chick, D.B.E., D.Sc.’, obituary in the British Medical Journal, (1977), pt 2, p. 270Google Scholar; Who was Who (19711980), London, 1981 vii, p. 145Google Scholar; Chick, , Hume, and Macfarlane, , op. cit. (28), pp. 8792, 124, 147160.Google Scholar

53 Martin, Charles James and Chick, Harriette, ‘The heat coagulation of proteins’, Journal of Physiology, (1910), 40, pp. 404–30Google Scholar. See also Edsall, , op. cit. (25), p. 70.Google Scholar

54 Chick, , Hume, and Macfarlane, , op. cit. (28), p. 92.Google Scholar

55 The acute food shortages in Europe and especially Austria at the time, and the subsequent widespread appearance of deficiency diseases, provided a unique opportunity for rigorous testing of the hypotheses worked out at the Institute during the war years.

56 The changing pattern of the place of origin of women's research publications for the period 1905–10 is summarized in Appendix 2. The numbers have been extended back to 1900 to bring out in particular the steady contribution from Cambridge throughout the whole decade, and the noticeable increase in contributions generally, starting around 1905, from the other three groups of institutions.

57 Record of the Council minutes, 1898–99, University College, Bristol. I thank Mr G. E. Maby, Archivist, University of Bristol, for information about Katherine Williams and Emily Fortey.

58 Fortey, Emily C., B.Sc, ‘Hexamethylene from American and Galician petroleum’, Journal of the Chemical Society, Transactions, (1898), 73, pp. 932–49.CrossRefGoogle Scholar

59 Markownikoff, V. V. [‘Recherches sur les composés cycliques de la série de l'hexamethylene’], Journal of the Russian Physico-Chemical Society, (1898), 30, pp. 5990, 151–95Google Scholar; Bulletin de la Société Chimique de Paris (1898), 20, pp. 851–6Google Scholar; [Liebig's] Annalen der Chemie, (1898), 301, pp. 154202; 302, pp. 142.Google Scholar

60 The technique used in the final stages of the commercial production of water-free ‘absolute’ alcohol was based on one of Fortey and Young's 1902 observations on the behaviour of mixed liquids. See Turner, E. E. and Harris, Margaret, Organic Chemistry, London, 1952, pp. 60–1.Google Scholar

61 Emily Fortey is referred to, though only by surname (‘Young and Fortey, 1902’) in Turner and Harris, Ibid., p. 60.

62 Information on Smith, Hibbert, Moody and Isaac was collected from their technical papers. Isaac is also mentioned in Byrne, and Mansfield, , op. cit. (2), p. 96.Google Scholar

63 See for instance Chick, , Hume, and Macfarlane, , op. cit. (28)Google Scholar, and Kohler, , op. cit. (20).Google Scholar

64 Burstall, , op. cit. (48), p. 2947.Google Scholar

65 Ibid., p. 2947.

66 See Robinson, Lisa Mae, ‘The electrochemical school of Edgar Fahs Smith, 1873–1913’, University of Pennsylvania, Ph.D. dissertation, 1986, pp. 231–3Google Scholar; university microfilm no. DET87–03261. Also Creese, Mary R. S. and Creese, Thomas M., ‘Mary Engle Pennington’, in the forthcoming book Women in Physics and Chemistry (ed. Grinstein, L.), Westport, Conn., 1991.Google Scholar

67 Unpublished data, derived from an analysis of the references to papers by women scientists listed in the Royal Society's Catalogue of Scientific papers, ixix.Google Scholar

68 The careers of these early women geologists are outlined in obituaries in the technical journals. See Reynolds, Doris L., ‘Dr Catherine Alice Raisin’, Nature, (1945), 156, pp. 327–8CrossRefGoogle Scholar; Cox, L. R. in the Proceedings of the Geological Society, (1935), 91, pp. xcviixcviii (Donald)Google Scholar; O.M.B.B. ibid. (1901), no. 1592, pp. 143–5 (Elles); E[lles], G[ertrude] L[ilian] in the Quarterly Journal of the Geological Society, (1946), 102, pp. xlvixlvii (Wood)Google Scholar; W[oods], H. in the Proceedings of the Geologists' Association, (1940), 57, p. 114 (Skeat)Google Scholar; Pia, Julius, ‘Maria Matilda Ogilvie Gordon’, Mitteilungen der Geologischen Gesellschaft in Wien, (1939), 32, pp. 173–86.Google Scholar

69 For information about Saunders, Sargant and Miles Thomas see respectively: Godwin, H., Clapham, A. R. and Gilson, M. R., ‘Edith Rebecca Saunders, F.L.S.’, New Phytologist, (1946), 45, No. 1, pp. 13Google Scholar; S[cott], D. H., ‘Miss Ethel Sargant, F.L.S.’, Annals of Botany, (04, 1918), 32, no. 126, pp. ivGoogle Scholar; DrThomas, E. N. Miles’, Nature, (1944), 154, pp. 481–82.Google Scholar

70 Kohler, , op. cit. (20), p. 49.Google Scholar

71 Ibid., pp. 60–1, 65–71.

72 Ibid., pp. 62–5. Although the general patronizing attitude of academic chemists towards biochemistry lasted into the 1920s, recognition of the practical importance of the field gradually grew in the preceding decades. In 1902 the Royal Institute of Chemistry officially recognized ‘biological chemistry’ as a branch of the field with its own qualifying examination (Pilcher, Richard B., The Institute of Chemistry of Great Britain and Ireland, History of the Institute: 1877–1914, London, 1914, pp. 150, 155)Google Scholar, and in 1907 Raphael Meldola, in his presidential address to the Chemical Society, acknowledged the national importance of research coming from establishments such as the Lister Institute, which he placed among the most important centres of research activity in the country; Meldola, , ‘The position and prospects of chemical research in Great Britain’, Journal of the Chemical Society, Transactions, (1907), 91, pp. 626–58, esp. pp. 635–6.CrossRefGoogle Scholar

73 Richter, (ed.), op. cit. (34), p. 7.Google Scholar

74 Needham, J., ‘Sir F. G. Hopkins' personal influence and character’, in Hopkins and Biochemistry (ed. Needham, J. and Baldwin, E.), Cambridge, 1949, pp. 114–15Google Scholar, quoted in Kohler, , op. cit. (20), p. 88.Google Scholar

75 Kohler, , op. cit. (20), p. 88.Google Scholar

76 The field was so described by Sayre, Anne, author of Rosalind Franklin and DNA, New York, 1975Google Scholar, and wife of crystallographer David Sayre. She is quoted by Julian, Maureen M., ‘Women in crystallography’, in Women of Science. Righting the Record (ed. Kass-Simon, G. and Farnes, Patricia), Bloomington, Indiana, 1990, pp. 339–40.Google Scholar

77 See for example O'Connor, Jean G. and Meadows, A. J., ‘Specialization and professionalization in British geology’, Social Studies of Science, (1976), 6, pp. 7789CrossRefGoogle Scholar, and Porter, Roy, ‘Gentlemen and geology: the emergence of a scientific career, 1600–1920’, The Historical Journal, (1978), 21, 4, pp. 809–36.Google Scholar

78 See Allen, David Elliston, The Naturalist in Britain. A Social History, London, 1976.Google Scholar

79 Morton, A. G., History of Botanical Science, London, 1981, esp. pp. 425–8.Google Scholar

80 At least three other national chemical societies had women members well before the turn of the century. Anna Fedorovna Volkova, a student of Mendeleev, joined the Russian Chemical Society in 1870 (see Musabekov, Iu. S., Iulia Vsevolodovna Lermontova, 18461919, Moscow, 1967, p. 8)Google Scholar, and there were several women the German Chemical Society by the 1880s, and in the American Chemical Society by the 1890s (see membership lists in the Berichte der deutsche chemische Gesellschaft and the Journal of the American Chemical Society, respectively).

81 ‘Women and the Fellowship of the Chemical Society’, Nature, (9 07 1908), 78, pp. 226–8, esp. p. 226.Google Scholar

82 Ibid., p. 227.

83 The fraction of the fellowship actively in support of membership for women was not as high as the voting figures might suggest. Out of a total membership of 2896, 1758 (60.7 per cent) returned ballots. Thus, of the Society as a whole, 37.8 per cent were actively in favour of the admission of women, 22.1 per cent were actively against, and the rest were neutral or did not consider the matter of sufficient interest to vote. Membership data and voting figures are from the report of the Annual General Meeting, 25 March 1909, Journal of the Chemical Society, Transactions, (1909), 95, pp. 611–23, esp. pp. 612, 614–16.Google Scholar

84 Martha Whiteley outlined the stages in the history of the opening of Fellowship in the Chemical Society to women in her obituary of MacLean, Ida Smedley, op. cit. (28), pp. 66–7Google Scholar. See also the Journal of the Chemical Society, Transactions, (1905), 87, pp. 538–9Google Scholar; Proceedings, (1905), 21, pp. 4159Google Scholar; Transactions, (1909), 95, p. 614.Google Scholar

85 Charlotte Angas Scott, a Girton student, placed equal to the eighth Wrangler in the Mathematical Tripos (MT) of 1880. She received a London D.Sc. in 1885, and spent her working life at Bryn Mawr College, Pennsylvania, publishing a considerable amount of original work and becoming Vice-President of the American Mathematical Society in 1906. Isabel Madison, a mathematics student at Cambridge from 1889 to 1892, also had a successful career at Bryn Mawr. Her friend Grace Chisholm (Girton 1889–92) was a Wrangler in Part of the MT in 1892, received a Ph.D. from the University of Göttingen in 1896 and later published notable work in the modern theory of real functions. Phillipa Fawcett, a student of J. J. Thompson, placed above the Senior Wrangler in the MT of 1890. After a short time as a lecturer at Newnham she went on to a distinguished career in educational administration, first in South Africa and then in London. Mary Ellen Rickett, a student at Newnham from 1881 to 1886, was one of the few women who took both the Classical Tripos and the MT. In the latter she placed equal to the twenty-fourth Wrangler. She was a lecturer in mathematics at Newnham and Vice-president of Old Hall, Newnham, from 1889 to 1910. Information about these five women came from the following sources: Ogilvie, , op. cit. (1), pp. 158–9 (Scott)Google Scholar; obituary in the New York Times, 24 10 1950, p. 29 (Madison)Google Scholar; Cartwright, M. L., ‘Grace Chisholm Young’, Journal of the London Mathematical Society, (1944), 19, pp. 185–92CrossRefGoogle Scholar; Who was Who (19411950), London, 1952, iv, p. 376 (Fawcett)Google Scholar; Tuke, , op. cit. (2), p. 311 (Rickett).Google Scholar

86 Stephen, , Emily Davies and Girton College, op. cit. (2), p. 128Google Scholar; McWilliams-Tullberg, , op. cit. (2), p. 34.Google Scholar

87 Ibid., pp. 35, 108.

88 Ibid., p. 108. See also Tuke, , op. cit. (2), p. 132.Google Scholar

89 McWilliams-Tullberg, , op. cit. (2), pp. 3649Google Scholar; Pedersen, Joyce Senders, ‘Schoolmistresses and headmistresses: elites and education in nineteenth century England’, The Journal of British Studies, (1978), 15, no. 1, pp. 135–62CrossRefGoogle Scholar; Richardson, J. L., ‘The great revolution; women's education in Victorian times’, History Today, (1974), 24 pp. 420–7Google Scholar; Stephen, , Emily Davies and Girton College, op. cit. (2), pp. 128–47.Google Scholar

90 Pedersen, , op. cit. (89), p. 148.Google Scholar

91 Tuke, , op. cit. (2), p. 132.Google Scholar

92 Cardwell, D. S. L., The Organization of Science in England, London, 1972, pp. 114–15Google Scholar; Budd, Robert and Roberts, Gerrylyn K., Science versus Practice, Chemistry in Victorian Britain, Manchester, 1984, pp. 120, 130.Google Scholar

93 Hamilton, , op. cit. (2), pp. 96, 110, 114Google Scholar, the quote being from p. 114.

94 The North London Collegiate School, 1850–1950. A Hundred Years of Girls' Education (ed. Scrimgeour, R. M.), London, 1950, pp. 30–3.Google Scholar

95 Tylecote, , op. cit. (2), p. 26.Google Scholar

96 Stephen, , Emily Davies and Girton College, op. cit. (2), p. 276.Google Scholar

97 In contrast, Oxford at this early period, though it educated its share of women who went on to distinguished careers, including many in the teaching profession, produced relatively few women who contributed to original work in the chemical sciences (see Appendix 1). The most eminent of the Oxford women were, typically, historians, linguists, writers, and administrators; see, for instance, the accounts of Brittain, Byrne, and Mansfield, , and Bailey, , op. cit. (2)Google Scholar, in which hardly any women scientists are mentioned.

98 Dunstan, Wyndham Rowland and Boole, Lucy Everest, ‘Chemical observations on tartar emetic’, Pharmaceutical Journal and Transactions, (1889), III, 19, pp. 385–7Google Scholar, and discussion pp. 397–9.

99 A Collective Index of the Transactions, Proceedings and Abstracts of the Chemical Society, for the period 18731912, 4 vols., London [n.d.]Google Scholar. Vols. i–iii were compiled by Mrs Dougal. The importance to the Society of this major indexing project and Mrs Dougal's many-year commitment to the work were acknowledged by Sir James Dewar in his Presidential Address to the Society, 1899; Journal of the Chemical Society, Transactions, (1899), 75, p. 1168.Google Scholar

100 MissHalcrow, Lucy and Frankland, E., ‘On the action of air upon peaty water’, Journal of the Chemical Society, Transactions, (1880), 37, pp. 506–17.CrossRefGoogle Scholar

101 Pilcher, , op. cit. (72), p. 114.Google Scholar

102 See ‘Marion Isabel Newbigin’, obituary in the Scottish Geographical Magazine, (1934), 50, pp. 331–3.Google Scholar

103 A Collective Index of the Transactions, Proceedings and Abstracts of the Chemical Society, 18731912, 4 vols., London [n.d.]Google Scholar. Work published in non-British journals is covered by the Abstracts sections of this Index.