Hostname: page-component-8448b6f56d-c47g7 Total loading time: 0 Render date: 2024-04-19T20:47:08.277Z Has data issue: false hasContentIssue false

British women who contributed to research in the geological sciences in the nineteenth century

Published online by Cambridge University Press:  05 January 2009

Thomas M. Creese
Affiliation:
Department of Mathematics, University of Kansas, Lawrence, KS, 66045, 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.

A count of articles by women listed in the Catalogue of Scientific Papers, 1800–1900, the nineteen-volume international index brought out by the Royal Society, produced a collection of almost 4000 titles of papers by about 1000 nineteenth-century women authors. Out of 181 geology papers in this collection, 118 (65 per cent) were by British women (see Table 1, columns 1 and 2). This finding is especially remarkable when considered against the more general background of nineteenth-century women's work in science (at least as judged from women's contributions to the journal literature indexed by the Royal Society). In most fields American workers considerably outnumbered British and published many more papers. Geology, however, is an exception, with the British dominating the field by a wide margin. This essay discusses a number of the British women of the period who carried out work in geology, and offers some suggestions that go toward explaining their striking prominence among their contemporaries.

Type
Research Article
Copyright
Copyright © British Society for the History of Science 1994

References

This work was supported in part by University of Kansas General Research allocations #3179-xx-0038 and 3312-xx-0038. We thank Dr Elisabeth van Houts, Newnham College Archivist, for information from Newnham College Register and Girton College Register, and Miss Carol R. Bowen, University College London, for information from student records. The passage in note 69 is quoted with the permission of the Principal and Fellows of Newnham College, Cambridge.

1 Catalogue of Scientific Papers, 18001900, 19 vols.Google Scholar Compiled by the Royal Society, Cambridge, 1867–1925.

2 Columns 3, 4, 5 and 6 in Table 1 give, for comparison, data for botany and the chemical sciences, two fields that were particularly popular with nineteenth-century women scientists. The exceptional position held by British women in geology is fairly clear; the fraction of papers by them in that field is the largest we have found for any national group within any given field.

3 Tuke, Margaret J., A History of Bedford College for Women, 1849–1937, London, 1939, especially 18, 127–8, 138–9Google Scholar; McWilliams-Tullberg, Rita, Women at Cambridge. A Men's University – Though of a Mixed Type, London, 1975Google Scholar; Stephen, Barbara, Emily Davies and Girton College, London, 1927.Google Scholar

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

5 To gain a somewhat fuller picture of British women's contributions in the early years of the century we have included in the discussion three women from that period who were notable but do not appear in the Royal Society index. These are Benett, Anning and Philpot. In the interest of consistency in international and interdisciplinary comparisons, they have not been added to the data lists in Table 1.

6 Buckland, Both Mary Morland (17971857)Google Scholar, wife of William Buckland, and Mary Ann Woodhouse Mantell (fl. 1820s and 1830s), wife of Gideon Mantell, assisted their husbands in illustrating, collecting and, in Buckland's case, writing. Even before her marriage, Mary Morland had been an active fossil collecter and she had prepared illustrations for the work of the French palaeontologist, Baron Georges Cuvier. Treatise, Buckland's Bridgewater, Geology and Mineralogy Considered with Reference to Natural Theology (1836)Google Scholar, included many illustrations prepared by his wife. Mary Mantell drew the forty-two plates of fossils which embellished Mantell's well-received Fossils of the South Downs: or Illustrations of the Geology of Sussex (1822)Google Scholar, and it was she who discovered the first of the iguanodon fossil teeth which led to Mantell's important 1825 paper in the Philosophical Transactions announcing the discovery of a new giant reptile group. Though an enthusiastic supporter of her husband's geological work throughout twenty years of marriage, she was finally forced to leave him when his preoccupation with geology displaced in his priorities both his family and his medical practice. Charles Lyell and Roderick Murchison also had assistance from their wives in their geological work. Lyell, Mary Horner (18081873)Google Scholar, daughter of the geologist Leonard Horner, became an accomplished conchologist during the course of her work with Lyell, whom she frequently accompanied on expeditions. She also read and translated for him and managed his correspondence, his eyesight being poor. Murchison, Charlotte (17891869)Google Scholar assisted in collecting, cataloguing and illustrating. Indeed, Murchison's marriage to Charlotte Hugonin in 1815 marked a turning point in his life, and it was to some extent her influence that drew him away from his first career in the army and into scientific pursuits. Lyell's description of Mrs Murchison during the 1828 Lyell-Murchison trip through the Auvergne and Italy makes clear her assistant's role, however: ‘She is very diligent, sketching, labelling specimens and making out shells, in which she is an invaluable assistant. She is so much interested in the affair, as to be always desirous of keeping out of the way when she would interfere with the work.’ (For Buckland and Mantell see, respectively, Mrs [Buckland, Elizabeth O.] Gordon, The Life and Correspondence of William Buckland, D.D., F.R.S., New York, 1894, especially 90–3Google Scholar, and Colbert, Edwin H., Men and Dinosaurs. The Search in Field and Laboratory, New York, 1968, 1523Google Scholar. For Lyell and Murchison see Bailey, Edward B., Charles Lyell, New York, 1963, especially 68–9, 95, 101–3, 144–5, 156Google Scholar; the passage quoted is on 68–9.) Two other women who assisted in early nineteenthcentury geological work were Lady Mary Cole and Jane Talbot, of Gower, Wales, who was frequently consulted by Buckland (see Phillips, Patricia, The Scientific Lady. A Social History of Women's Scientific Interests 1520–1918. New York, 1990, 184.Google Scholar

7 Torrens, H. S., ‘Women in geology. 2 – Etheldred Benett’, Open Earth (1983), 1213.Google Scholar

8 Woodward, H. B., The History of the Geological Society of London, London, 1907, 118–19.Google Scholar

9 The description appeared, with additional notes by Sowerby, , in vol. 2 (pp. 58–9)Google Scholar of his Mineral Conchology. Torrens, , op. cit. (7), 12Google Scholar, has pointed out that Benett was clearly influenced by the pioneering work during the early years of the century of William Smith, a surveyor, who led the way in showing that strata could be identified by their fossil content; Smith's ideas quickly became fairly widely known.

10 See Edmonds, J. M., ‘The fossil collection of the Misses Philpot of Lyme Regis,’ Dorset Natural History and Archaeological Society Proceedings (1976, 1978), 98, 43–8.Google Scholar

11 Dictionary of National Biography, London, 1921 (hereafter DNB), xxii, supplement, 51–2Google Scholar; Gordon, , op. cit. (6), 113–16Google Scholar; Phillips, , op. cit. (6), 183.Google Scholar

12 Woodward, , op. cit. (8), 115.Google Scholar

13 See, for example, Wilson, Leonard G., ‘The intellectual background to Charles Lyell's Principles of Geology, 1830–1833’, in Toward a History of Geology (ed. Schneer, Cecil J.), Cambridge, Mass., 1969, 426–43, especially 435–6.Google Scholar

14 Anning, Mary, ‘Note on the supposed frontal spine in the genus Hybodus’, Magazine of Natural History (1839), 12 (3, n.s.), 605. The note reads as follows:Google Scholar In reply to your request I beg to say that the hooked tooth is by no means new; I believe M. De la Beche described it fifteen years since in the Geological Transactions, but I am not positive; but I know that I then discovered a specimen, with about a hundred palatal teeth, as I have donc several times with different specimens. I had a conversation with Agassiz on the subject; his remark was that they were the teeth by which the fish seized its prey, – milling it afterwards with its palatal teeth. I am only surprised that he has not mentioned it in his work. We generally find Ichthyodorulites [defensive finbones] with them, as well as Cartilaginous bones.

15 Lang, W. D., ‘More about Mary Anning, including a newly found letter’, Dorset Nat. Hist. Soc, Proc. (1950), 71, 184–8Google Scholar. See also Lang, W. D., ‘Three letters by Mary Anning, fossilist of Lyme’Google Scholar, ibid. (1945), 66, 169–73.

16 Remark by Mary Anning to the King of Saxony, on his visit to Lyme Regis (1844), quoted by Gordon, , op. cit. (6), 115.Google Scholar

17 Edmonds, , op. cit. (10), 45.Google Scholar

18 DNB, iii, 710Google Scholar; Rose, H. J., A New General Biographical Dictionary, London, 1850Google Scholar, entry reproduced in British Biographical Archive, microfiche edn (ed. Sieveking, Paul), London, 1984, no. 351Google Scholar; Robinson, Jane, Wayward Women. A Guide to Women Travellers, Oxford, 1990, 44–6.Google Scholar

19 Graham, Maria, ‘An account of some effects of the late earthquakes in Chili’, Transactions of the Geological Society (1824), 1, 413–15.CrossRefGoogle Scholar

20 See the three-part article, ‘On the reality of the rise of the coast of Chile in 1822, as stated by Mrs. Graham’, American Journal of Science and Arts (1835), 28, 239–47Google Scholar. This article includes a reprinting of Mrs Graham's 1824 account in the Transactions of the Geological Society, Greenough's presidential address, and Graham's reply to the latter; the quotation appears on p. 240; italics added.

21 The following brief extract gives the flavour of Mrs Graham's well-orchestrated counter-attack. She wrote here in the third person: Mr. Greenough mentions Mrs. Callcott's published journal, and accounts for the dead fish on the shore [Journal, p. 331Google Scholar] by an imaginary storm. Common candour would have led that gentleman to have stated that, in that very journal, it is distinctly printed, that a ‘delightful and calm moon-light night followed a quiet and moderately warm day’ [ibid., 305]. Mr. Greenough says, further in p. 18 of his address – ‘some muscles [sic] and oysters still adhere, she says, to the rocks on which they grew: but we do not know the nature of these rocks, whether fixed or drifted.’ Mrs. Callcott was ignorant that there were, or might have been, drifted rocks, until she learned it from Mr. Greenough; for much as she has been at sea, she never met with one. The rocks at Quintero, and at Valparaiso, are of grey granite, and where they lift themselves through the sand and shingle of the beach, they give the notion of bald mountain tops. At all events, they are fixed sufficiently to have caused the wreck of more than one Spanish ship of war; and when she saw them the morning after the Earthquake, that on which the wreck of the Aquila lay, was certainly so far above the water, that the vessel could be approached dryshod, which had never happened before, even at the lowest tides. The beds of muscles [sic], of other shellfish, and of sea-weed, were equally rocks of grey granite, fixed far below the sands of the ocean. These circumstances are stated in the published journal: but Mr. Greenough has suppressed them, and many others of the like nature, particularly the notice of some rocks and stones, that the lowest tides never left dry, but now have a passage between them and the low-water mark, sufficient to ride around without difficulty, p. 313. (Graham, , op. cit. (20), 244.)Google Scholar

22 Davison, Charles, The Founders of Modern Seismology, Cambridge, 1927, 34–6.Google Scholar

23 Allen, D. E. in, ‘The women members of the Botanical Society of London; 1836–1856’, BJHS (1980), 13, 240–54, note 14 on 253CrossRefGoogle Scholar, voiced some dissatisfaction with Valerie Appleby's claim that Maria Graham was the first woman to have a paper published by the Geological Society (Appleby, , ‘Ladies with hammers’, New Scientist (1979), 84, 714–15Google Scholar). Allen pointed out that Graham's contribution was an extract from a letter written by her to the vice-president and was not read before the Society. As an extract from her published journal, however, supplied at the vice-president's request, presumably for publication in the Society's Transactions because it concerned a matter of considerable current interest, one might perhaps argue that it came close to being in the nature of an invited paper.

24 See Edwards, Nicholas, ‘The Hastings Collection’, Journal of the Society for the Bibliography of Natural History (1970), 5, 340–3.CrossRefGoogle Scholar

25 Marchioness of Hastings, ‘Description géologique des falaises d'Hordle, et sur la côte de Hampshire, en Angleterre’, Bulletin de la Société Géologique de France (18511852), n.s. 9, 191203Google Scholar; ‘On the tertiary beds of Hordwell, Hampshire’, Philosophical Magazine (1853), 6, 111.Google Scholar

26 See Owen, R., ‘On the fossils obtained by the Marchioness of Hastings from the freshwater Eocene beds of Hordle Cliffs’, Reports of the British Association for the Advancement of Science (1847), pt 2, 65–6.Google Scholar

27 Edwards, , op. cit. (24), 342.Google Scholar

28 Marchioness of Hastings, ‘On the freshwater Eocene beds of Hordle Cliff, Hants’, Reports of the British Association for the Advancement of Science (1847), pt 2, 63–4Google Scholar. See also Owen, , op. cit. (26).Google Scholar

29 For Carne see the DNB, iii, 1044Google Scholar; for Eyton see Kirk, J. F., A Supplement to Allibone's Critical Dictionary of English Literature, 2 vols., Philadelphia, 1891Google Scholar, reproduced in British Biographical Archive (1984), no. 386Google Scholar; for Hodgson see the obituary in the Journal of Botany (1878), 16, 64.Google Scholar

30 Davidson, Charles, ‘A monograph of recent brachiopoda’, Linnean Society of London Transactions, ser. 2, Zoology (18861888), 4, 248 pp.Google Scholar

31 For Thompson see the obituary by C[ole], G[renville] A. J., ‘Madame Christen’, Geological Magazine (1923), 60, 478–9Google Scholar; for Crane see Who's Who in Kent, Surrey and Sussex, Cox's County Series, London, 1911Google Scholar, reproduced in British Biographical Archive (1984), no. 278.Google Scholar

32 Graham, , op. cit. (20), 245.Google Scholar

33 See the obituaries by Pia, Julius, ‘Maria Matilda Ogilvie Gordon’, Mitteilungen der Geologischen Gesellschaft in Wien (1939), 32, 173–86Google Scholar, by Bailey, E. B., ‘Maria Ogilvie Gordon, D.B.E.’, Nature (1939), 144, 142–3CrossRefGoogle Scholar, and by E. J. G., Quarterly Journal of the Geological Society (1946), 102, pp. xlxli.Google Scholar

34 This was most likely Baron Ferdinand von Richthofen, then professor of physical geography at the University of Berlin.

35 Much to Ogilvie's disappointment, the mineralogist Paul Heinrich Groth declined to admit her to his laboratory. Permission for individual women to audit courses at the University of Munich was first granted formally by the authorities only in 1896; James Albisetti has noted that the first woman to take classes there was the English geology student Ethel Skeat (see below) who also worked under von Zittel (Albisetti, James C., Schooling German Girls and Women. Secondary and Higher Education in Nineteenth Century Germany, Princeton, 1988, 278Google Scholar). It would seem likely, therefore, that Ogilvie went to Munich in 1891 as a private student of von Zittel and Hertwig. Even earlier (probably in 1886) the Russian student Evgeniia Solomko (see below) had been sympathetically received by von Zittel. He arranged for her to listen to his lectures through an open door in a room next to the auditorium where the German male students sat. Solomko took the complete course of lectures and also had full access to the Palaeontological Museum with its coral collection, her special interest (Nalivkin, D. V., Nashi Pervye Zhenshchinye-geologi [Our First Women Geologists], Leningrad, 1979Google Scholar. We thank Alla Ivanovna Polekutina, Institute of the History of Science and Technology, Academy of Sciences, Moscow, for telling us about Nalivkin's book).

36 See Ogilvie, Maria M., ‘Contributions to the geology of the Wengen and St. Cassian strata in the southern Tyrol’, Quarterly Journal of the Geological Society (1893), 49, 177CrossRefGoogle Scholar; ‘Coral in the “dolomites” of South Tyrol’, Geological Magazine (1894), 1, 110, 4960Google Scholar; ‘Die Korallen der Stramberger Schichten’, Palaeontographica (1897), Suppl. 2, Abth. 7, pp. 73282, iivGoogle Scholar; ‘Microscopic and systematic study of madreporian types of corals’, Philosophical Transactions, B (1897), 187, 83345.Google Scholar

37 Gordon, Maria Ogilvie, Das Grödener-, Fassa-, und Ennerberggebiet in den Südtiroler Dolomiten. Geologische Beschreibung mit besonderer Berucksichtigung der Ueberschiebungscheinungen, Vienna, 1927Google Scholar. The work had been almost ready for publication in 1914, when, with the outbreak of war, Ogilvie Gordon had to return to Britain, leaving the manuscript in Munich. By the time she went back in 1920 it had disappeared and she had to rewrite it entirely.

38 Solomon-Calvin, Wilhelm, ‘The Dolomites of South Tyrol’, Nature (1928), 121, 83–5Google Scholar, a review of the monograph (tr. L. R. Cox).

39 Ibid., 996.

40 See Pia, , op. cit. (33), 184Google Scholar. The quotations are from Ogilvie Gordon's letters to Pia, and his reminiscences.

41 See the obituaries by Reynolds, Doris L., ‘Dr. Catherine Alice Raisin’, Nature (1945), 156, 327–8CrossRefGoogle Scholar, L. H., Proceedings of the Geologists' Association (1946), 57, 53–4CrossRefGoogle Scholar, and E. J. G., Quarterly Journal of the Geological Society (1946), 102, pp. xlivxlv.Google Scholar

42 Bonney, T. G. and MissRaisin, Catherine, ‘The microscopic structure of minerais forming serpentine, and their relation to its history’, Quarterly Journal of the Geological Society (1905), 61, 690714CrossRefGoogle Scholar. See also, Raisin, Catherine A., ‘On the nature and origin of the Rauenthal serpentine’Google Scholar, ibid. (1897), 53, 246–68; and (with Bonney, T. G.) ‘On varieties of serpentine and associated rocks in Anglesey’Google Scholar, ibid. (1899), 55, 276–302.

43 Porter, , op. cit. (4), 835.Google Scholar

44 Obituaries by W. B. R. K., Proceedings of the Geologists' Association (1961), 72, pt 1, 168–71CrossRefGoogle Scholar, and by B[ulman], O. M. B., Proceedings of the Geological Society (1961), no. 1592, 143–5Google Scholar; see also Challinor, John, The History of British Geology. A Bibliographical Study, New York, 1971, 187.Google Scholar

45 E[lles], G. L., obituary for Margaret Chorley Crosfield, Proceedings of the Geological Society (1953), no. 1502, pp. cxxxicxxxii.Google Scholar

46 Elles, G. L. and Wood, E. M. R., British Graptolites, London, 19011919.Google Scholar

47 Elles, G. L., ‘The graptolite faunas of the British Isles’, Proceedings of the Geologists’ Association (1922), 33, 168200.CrossRefGoogle Scholar

48 Challinor, , op. cit. (44), 144–5.Google Scholar

49 See, for instance, the following: (with Slater, Ida L.) ‘The highest Silurian rocks of the Ludlow district’, Quarterly Journal of the Geological Society (1906), 62, 195221Google Scholar; ‘The relation of the Ordovician and Silurian rocks of Conway (North Wales)’, ibid. (1909), 65, 169–92; ‘The Bala county: its structure and rock succession’, ibid. (1922), 78, 132–68.

50 See the obituaries by E[lles], G[ertrude] L[ilian], Quarterly Journal of the Geological Society (1946), 102, pp. xlvixlviiGoogle Scholar, and Nature (1946), 157, 256–7.Google Scholar

51 Wood, Ethel M. R., ‘The Lower Ludlow formation and its graptolite fauna’, Quarterly Journal of the Geological Society (1900), 56, 415–92CrossRefGoogle Scholar; ‘The Tarannon series of Tarannon’, ibid. (1906), 62, 644–99.

52 See the obituary by W[oods], H[enry], Proceedings of the Geologists' Association (1940), 57, 114.Google Scholar

53 See note 35.

54 See Crosfield, Margaret Chorley and Skeat, Ethel G., ‘On the geology of the neighbourhood of Carmarthen’, Quarterly Journal of the Geological Society (1896), 52, 523–41CrossRefGoogle Scholar; ‘The Silurian rocks of the central part of the Clwydian Range’, ibid. (1925), 81, 170–92.

55 Skeat, Ethel G. and Madsen, Victor, ‘On the Jurassic, Neocomian and Gault boulders found in Denmark’, Danmarks Geoloeiske Undersogelse (1898), Rekke2, 8, 213 pp.Google Scholar

56 J[ohnston], M[ary] S[ofia], Proceedings of the Ceologists' Association (1953), 64, 62–3Google Scholar; Elles, , op. cit. (45)Google Scholar; Neivnham College Register, 1871–1971, i, 65.Google Scholar

57 See Crosfield, Margaret C., ‘The Tremadoc slates’, Proceedings of the Holmesdale Natural History Club (18961898), 1122Google Scholar; ‘Redhill and Merstham’, ibid. (1896–98), 65–6.

58 See the obituaries by Cox, L. R. in the Proceedings of the Geological Society (1935), 91, pp. xcviixcviiiGoogle Scholar, and in the Proceedings of the Geologists’ Association (1936), 47, 97.Google Scholar

59 Donald, Jane, ‘Notes on some Carboniferous gastropoda from Penton and elsewhere’, Transactions of the Cumberland and Westmorland Association (1885), 9, 127–36Google Scholar. Included among her later writings were the following: ‘Notes on the genus Murchisonia and its allies; with a revision of the British Carboniferous species, and descriptions of new forms’, Quarterly Journal of the Geological Society (1895), 51, 210–34Google Scholar; ‘A revision of the British Carboniferous Murchisoniidae, with notes on their distribution and descriptions of some new species’, ibid. (1926), 82, 526–55; ‘A revision of the British Carboniferous members of the family Loxonematidae with descriptions of new forms’, ibid. (1933), 89, 87–122.

60 See Allen, David Elliston, The Naturalist in Britain. A Social History, London, 1976, 111.Google Scholar

61 See Lewis, Jane, Women and Social Action in Victorian and Edwardian England, Stanford, 1991, 6. But see also note 63.Google Scholar

62 Lewis, , op. cit. (61), 116Google Scholar; see also Lewis, Jane, Women in England 1870–1950: Sexual Divisions and Social Change, Bloomington, 1984, 73–4.Google Scholar

63 The generalization that women faced a choice between marriage and a career, which holds for the late nineteenth and early twentieth centuries (Lewis, , op. cit. (62), 73Google Scholar) would not appear to fit readily the lives of outstanding women research scientists. There are several instances from the period of successful husband-wife scientific teams, and a number of examples of outstanding women who, like Ogilvie Gordon, Wood and the early Russian geologist, Mariia Pavlova (see below) had tolerant and sympathetic husbands, supportive of their wives' professional work. (For examples of teams see Ogilvie, Marilyn Bailey, ‘Marital collaboration. An approach to science’, in Uneasy Careers and Intimate Lives. Women in Science 1781–1979 (ed. Abir-am, Pnina G. and Outram, Dorinda), New Brunswick, 1987, 104–25Google Scholar, and other essays in this collection; for another example of an outstanding early woman scientist whose work was independent of her husband's, see the discussion of MacLean, Ida Smedley in Creese, Mary R. S., ‘British women of the nineteenth and early twentieth centuries who contributed to research in the chemical sciences’, BJHS (1991), 24, 275305, especially 282–4.Google Scholar) The life of the English palaeobotanist Reid, Eleanot (18601953)Google Scholar provides another instance of a very successful married early woman geological scientist. In fact, Reid entered the field through collaboration with her husband, having taught mathematics at Cheltenham Ladies' College before her marriage. After she was widowed in 1916 she continued on her own and became internationally known as an expert on Pleistocene and Tertiary seeds (obituary by W.N.E., Proceedings of the Geological Society (1954), no. 1515, pp. cxlcxlii).Google Scholar

64 See Caine, Barbara, Victorian Feminists, Oxford, 1992, especially 251.Google Scholar

65 Their many undertakings brought constant conflict though. Wood gave up her scientific work at age fortyeight, after the completion of the graptolite monograph, and from then on let public service take ail her time; Ogilvie Gordon was for much of her life torn between the demands of public work she declined to turn aside and her hopes and plans for research. Pia, , op. cit. (33), 183Google Scholar, quoted a number of her remarks on this conflict in her letters to him: (1922) ‘I have been overwhelmed with outside work’; (1925) ‘I am rather in despair about getting on with my work’; (1927) ‘Endless work to do in several of the Women's Societies and have had no attempt at finding time for geology’; (1928) ‘So many outside duties to attend to’; (1931) ‘I am getting nothing done at present because there is a rush of meetings’; (1934) ‘I am not particularly happy this winter as I have had no time to do any scientific work’; (1936) ‘I keep toiling hard at public work, meetings of all kind, while knowing inwardly that it is geology in my heart that I wish to do’; and in 1938, a few months before she died, ‘I cannot tell you how constantly I have had to work all summer and autumn over many public duties’.

66 Creese, , op. cit. (63).Google Scholar

67 For a fuller discussion of the interaction at this period between the women's colleges at Cambridge and the leading girls’ secondary schools see Creese, ibid., 299–301.

68 Porter, , op. cit. (4), 835.Google Scholar

69 Ball, M. D., a student at Newnham from 1908 to 1911Google Scholar, left a sprightly description of Marr's lectures (‘Newnham scientists’, in A Newnham Anthology (ed. Phillips, Ann), Cambridge, 1979, 76):Google Scholar Dear Dr. Marr would corne stamping into the lecture room, his gown streaming behind him, and cut and slice the country up with gusto. Glaciers streamed slowly down, rivers meandered or were beheaded, such ‘noxious spots’ as the Fens spread themselves, and only the ancient hills of his beloved Lakeland and Scotland remained to make habitable spots for decent folk. All the same we all panted joyfully after him, filling notebook after notebook with more or less legible writing.

70 Johanson, Christine, Women's Struggle for Higher Education in Russia. 1855–1900, Kingston and Montreal, 1987, especially the introduction, 38 and ch. 4, 5976Google Scholar; Dudgeon, Ruth A., ‘The forgotten minority: women students in imperial Russia, 1872–1917’, Russian History (1982), 9, 126CrossRefGoogle Scholar, and ‘Women and Higher Education in Russia, 1855–1905‘, Ph.D. dissertation, George Washington University, 1975, university microfilm no. 75–26000Google Scholar; Satina, Sophie, Education of Women in Pre-revolutionary Russia (tr. Poustchine, Alexandra F.), New York, 1960.Google Scholar

71 Instruction was provided by the men for little or no pay as the result of an extraordinary alliance between Russian academics and women students at the time. This alliance came about largely as follows: interest in higher education for women had begun in Russia during the social upheavals following the War, Crimean (18531856)Google Scholar, and for a few years beginning in 1859 women were accepted, in increasingly large numbers, as auditors in university courses. In 1863, however, strong pressure from the government forced universities to expel women auditors, a move which generated great sympathy for the women, and, in many of the faculty, a sense of moral obligation to try to help them. For more than a decade little could be done; many Russian women went to western Europe, especially Zurich, for university education (see Meijer, Jan M., Knowledge and Revolution. The Russian Colony in Zurich (1870–1873), Amsterdam, 1955Google Scholar; Rohner, Hanny, ‘Die ersten 30 Jahredes medizinischen Frauenstudiums an der Universität Zurich, 1867–1897’, Zürcher Medizingeschichtliche Abhandlungen, Neue Reihe (1972), 89, 596Google Scholar; Johanson, , op. cit. (70), 51–8Google Scholar). In 1873, however, the tsarist government, uneasy about student involvement in émigré revolutionary circles, ordered the Russian women students to leave Zurich. Three years later, the demand in Russia for women's courses having increased, these were once again officially authorized, and the men faculty came forward in large numbers to conduct them. All but the St Petersburg (Bestuzhevskie) courses were closed by the late 1880s as a delayed result of the counter-reforms following the assassination of Alexander II, and from then until the changes made by Nicholas II in 1894, educational opportunities for women in Russia were very limited – once again there was an exodus to western Europe (see, especially, Johanson, , op. cit. (70)).Google Scholar

72 Ibid., 74.

73 Unpublished data, but see Table 1 for numbers on the chemical sciences.

74 Nalivkin, , op. cit. (35), especially 436.Google Scholar

75 Tsvetaeva, Mariia, ‘Cephalopodes de la section supérieure du Calcaire Carbonifère de la Russie centrale’, [1888], Mémoires du Comité Géologique (18881890), 5, 58pp.Google Scholar; ‘Nautiloidea et Ammonoidea de la section inférieure du Calcaire Carbonifèr de la Russie centrale’, [1898], ibid., (1888–98), 8, 46pp. (The Geological Committee, founded in 1882, was about the equivalent of the British Geological Survey.)

76 See note 35.

77 Solomko-Sotiriadis, Evgeniia, ‘[Stromatopora des Devonishchen System Russlands]’, Verhandlungen der Russisch-Kaiserlichen Mineralogischen Gesellschaft zu St. Petersburg (1887), 23, 148Google Scholar; ‘Die Jura- und Kreidekorallen der Krim’, ibid. (1888), 24, 67–231. Among early Russian and British women geologists, the majority were palaeontologists or biostratigraphers. Of the five Russian women whose nineteenth-century work is listed in the Royal Society Catalogue, the three most productive were palaeontologists. The prominence of women in this area persisted for many decades, at least in the former Soviet Union, the country which by the midtwentieth century, trained and employed most women geologists. A survey of about 2500 Soviet palaeontologists made in 1968 brought out the fact that more than half (about 1500) were women. Furthermore, the most difficult and complex fossil groups were those most often studied by women; of 480 workers in the survey specializing in foraminifers (small, primitive protozoans), 400 were women, while of 180 people studying brachiopods (where the labour required is considerably less monotonous and demanding) 100 were men and 80 were women (Nalivkin, , op. cit. (35), 56Google Scholar). Although the number of British women palaeontologists active around the turn of the century is too small to support generalizations about what sub-area they tended to cluster in, it is perhaps worth noting that Elles and Wood worked on the difficult graptolite group, and Ogilvie Gordon specialized in corals and algae.

78 Among Pavlova's most substantial pre-1901 contributions was her series, ‘Études sur l'histoire paléontologique des Ongules en Amérique et en Europe’, Bulletin de la Sociétié Impériale des Naturalistes de Moscou (1889), 2, 135–82Google Scholar; (1890), 3, 653–716; (1893), 6, 137–221; (1900), 13, 268–328. Pavlova was the wife of the eminent Russian geologist A. P. Pavlov.

79 Only two women were accorded this honour before Pavlova: the mathematician Sofia Kovalevskaia in 1889 and Marie Sklodowska Curie in 1907.

80 See Rossiter, Margaret W., Women Scientists in America. Struggles and Strategies to 1940, Baltimore, 1982, especially chs. 1 and 2, 150Google Scholar; Conable, Charlotte Williams, Women at Cornell: The Myth of Equal Education, Ithaca, 1977Google Scholar; Arnold, Louis Barber, Four Lives in Science. Women's Education in the Nineteenth Century, New York, 1984Google Scholar; Kohlstedt, Sally Gregory, ‘In from the periphery: American women in science, 1830–1880’, Signs (1978), 4, 8196.CrossRefGoogle Scholar

81 See Rossiter, Margaret W., ‘Geology in nineteenth-century women's education in the United States’, Journal of Geological Education (1981), 29, 228–32.CrossRefGoogle Scholar

82 Even then the Bryn Mawr administration was reluctant to proceed with developing a full geology department; the subject was thought of as having ‘no wide appeal to women’ (Ogilvie, Ida H., ‘Florence Bascom, 1862–1945’, Science (1945), 102, 320–1).CrossRefGoogle ScholarPubMed

83 Arnold, , op. cit. (80), 92123Google Scholar. Bascom, however, was not the first woman to become a Fellow of the Geological Society of America. That honour went to Holmes, Mary Emilée, elected in 1889Google Scholar, one year after the society was established. Holmes took at Ph.D. in zoology at the University of Michigan in 1888, writing a dissertation on the morphology of the septa of fossil rugose corals. Nevertheless her main interest was home mission work, and after 1888 she appears to have given most of her time to activities in the Presbyterian Church and the promotion of eduction for negro girls in the South (Who was Who in America, Chicago, 1943, i, 581–2Google Scholar; see also Elder, Eleanor S., ‘Women in early geology’, Journal of Geological Education (1982), 30, 287–93).CrossRefGoogle Scholar

84 Bascom was accepted at Johns Hopkins (as a special student) only after she had made a convincing case to the university's executive committee that she could not get the instruction she wanted in petrography anywhere else in the country. She worked to some extent in isolation, listening to lectures from behind a screen, much as Solomko had done in Munich in the 1880s. Her Ph.D. was the first in any field that Johns Hopkins gave to a woman.

85 Many of Bascom's most important studies appeared as US Geological Survey reports, generally co-authored by her collaborators on the survey staff. She worked in the Piedmont region of Maryland, Pennsylvania, and parts of Delaware and New Jersey, and made notable contributions over the years to the somewhat heated controversy about the origin of the complex eastern Appalachian erosion surfaces.

86 In the United States from about 1830 onwards a few women served as technical illustrators, sometimes working as assistants to geologist husbands (see note 6 for mention of similar work by geologists' wives in Britain); probably the most notable were Orra White Hitchcock, Sarah Hall, Kate Andrews and Harriet Huntsman, who prepared drawings for Geological Survey reports and other technical publications. One or two American women of the period wrote popular instruction works in geology; Welsh, Jane Kilby's Familiar Lessons in Geology and Mineralogy Designed for the Use of Young Persons and Lyceums appeared in 1832Google Scholar and Dodding, Miss D. W.'s First Lessons in Geology in 1847Google Scholar (see Aldrich, Michele L., ‘Women in geology’, in Women in Science. Righting the Record (ed. Kass-Simon, G. and Fames, Patricia), Bloomington, 1990, 4271Google Scholar). Similar popular writings by women authors appeared about the same time in Britain, for instance Hack, Maria's Geological Sketches (1832)Google Scholar. By mid-century public lectures on geology, such as those given at the Lowell Institute in Boston and in other large cities, were popular with American women (Rossiter, , op. cit. (81), 229Google Scholar), but their interest in these does not appear to have led to original contributions in technical journals.

87 We have not attempted to set out the social circumstances and educational patterns that might suggest explanations for the very low participation of western European women other than the British in nineteenth-century geological work. It is perhaps worth noting, however, that comparison of the data in Tables 1 and 2 brings out the remarkable fact that the two regions where the bulk of geological, and all other scientific research was being carried out (Germany-Austria and France-Belgium-The Netherlands) were those where participation by women was least (Scandinavia, where participation by men was also very low, is an exceptional case).

88 Phillips, , op. cit (6), 108Google Scholar. See also Patterson, Elizabeth Chambers, Mary Somerville and Cultivation of Science, 1815–1840, The Hague, 1983.CrossRefGoogle Scholar

89 Smith, 's remarks were quoted in William Smith, his Maps and Memoirs (Thomas Sheppard, Hull, 1920, 217)Google Scholar, and reproduced in Torrens, , op. cit. (7), 12Google Scholar. Etheldred Benett was without doubt one of the Wiltshire ladies to whom Smith was referring.

90 O'Connor, and Meadows, , op. cit. (4), 85Google Scholar. From the 1830s, women had also attended geology lectures at British Association meetings (Phillips, , op. cit. (6), 200–7).Google Scholar

91 O'Connor, and Meadows, , op. cit. (4), 87.Google Scholar

92 See the discussion by Phillips, Patricia (op. cit. (6), 241–8)Google Scholar of the findings of the four-year Taunton Commission of Inquiry into the state of English schools (1864–68), and also Hunt, Felicity, ‘Divided aims: the educational implications of opposing ideologies in girls' secondary schooling, 1850–1940’, in Lessons for Life. The Schooling of Girls and Women, 1850–1950 (ed. Hunt, Felicity), Oxford, 1987, 321, especially 6Google Scholar. Leading girls' schools of the time included the North London Collegiate School, under Frances Mary Buss, Cheltenham Ladies' College, under Dorothea Beale, a number of Quaker schools, and girls' schools in Brighton and Bristol. In the period after 1870 the number of academically sound girls' schools increased further; among their science mistresses were many of the earliest of the women trained at Newnham and Girton Colleges. See also Pederson, Joyce Senders, ‘School mistresses and headmistresses: elites and education in nineteenth century England’, The Journal of British Studies (1978), 15, 135–62.CrossRefGoogle Scholar

93 ‘Notes’, Nature (1878, 15 08), 18, 423.Google Scholar

94 Ibid. (1879, 27 March), 19, 497.

95 Phillips, , op. cit. (6), 238–41Google Scholar. See also Hunt, , op. cit. (92), 8Google Scholar. Later Nature reports of Cambridge entrance examination results generally did not separate boys from girls, but a note in August of 1879 (20, 379) stated that five candidates passed with distinction in geology, and among these were Miss C. E. Cross of Cambridge and Miss M. A. Broadhurst of Liverpool College for Girls, as well as a candidate from Leeds whose name was withheld (and who was therefore most likely a girl).

96 See note 11.

97 See note 7.

98 See note 29.

99 Newnham College Register, 1871–1971, i, 92.Google Scholar

100 See note 31.

101 See note 56.

102 Girton College Register, 1869–1946, i, 43.Google Scholar

103 See note 44.

104 See note 29.

105 E. W., Proceedings of the Geologists' Association (1926), 37, 229.Google Scholar

106 Newnham College Register, 1871–1971, i, 79.Google Scholar

107 See note 18.

108 See note 24.

109 See note 29.

110 See obituary for Hughes, T. McKenny, Geological Magazine (1917), 54, 334–5.Google Scholar

111 DNB, xi, 820.Google Scholar

112 See note 58.

113 See note 33.

114 See note 41.

115 See note 52.

116 Newnham College Register, 1871–1923, Staff, i, 9.Google Scholar

117 See note 31.

118 See note 50.

119 Some authors published in journals other than those of their own countries, particularly German journals. However, checks of the nationalities of authors in the random sample entries suggest that this factor has no significant effect on the conclusions drawn here.