Keywords

1 Introduction

As several historians have demonstrated, women played an important role in the development of British geology between 1780 and 1850 (see Burek and Higgs (eds.) 2007; Creese 1998). M.R.S. Creese (2007), for instance, points out that British women contributed five times as many papers to scientific journals in this period as women in Russia/Poland, the next-ranked nation for female contributors. To take a different measure, as scientific publishing for broad audiences took off in the 1820s and 1830s, women like Jane Marcet, Jane Kilby Welsh, Rosina Zornlin, and Maria Hack expounded new geological knowledge to broad audiences, including children, ladies, or those with a predisposition to Paleyan natural theology (see Larsen, 2018). In fossil collecting, women were especially well-represented: Lady Gordon Cumming of Altyre collected fossil fishes on the shores of Moray Firth in the 1840s, while Lady Eliza Maria Gordon Cumming’s fossil collection supplied gentlemen geologists like William Buckland, Louis Agassiz, and Roderick Murchison, aided by pen and watercolor sketches of the fossils made by her daughter, Lady Seymour (Creese, 2007: 39–40). Lady Mary Cole and “the Misses Talbot” mapped for William Buckland in Glamorganshire, South Wales (Buckland in Gordon, 1894: 16). Barbara Marchioness of Hastings had “the finest collection of Speeton clay shells I have ever met with,” according to one male geologist, who noted that “in vertebrates she beats all the private collections” (he took away an oolitic starfish to write up for a paper) (Edward Forbes in Kölbl-Ebert 2004: 76, 77). Hastings also regularly corresponded with Richard Owen and a host of other leading naturalists, produced a detailed and original stratigraphical section of the Eocene strata near Milford in Hampshire, published three geological papers, and sold her collection to the British Museum in 1855 (Kölbl-Ebert 2004; Kölbl-Ebert 2002: 17–18). As these examples show, women also participated in geological mapmaking and the drawing of sections: Etheldred Benett’s stratigraphical description of Upper Chisgrove was one of the earliest detailed quarry sections produced in England, and she made many contributions of fossil specimens to James Sowerby’s Mineral Conchology of Great Britain (which contained the work of 29 separate female contributors, in fact). Benett’s specimens were also added to the collections of the Geological Society of London, and Benett contributed key information to George Bellas Greenough’s Geological Map of England and Wales (see Kölbl-Ebert 2002: 16–17; Torrens 1985; Torrens et al. 2000; Spamer, Bogan and Torrens, 1989). In yet another example of women’s work in the earth sciences, Maria Graham publicly challenged Greenough on his interpretation of the Chilean earthquake of 1822, dramatizing the importance of firsthand witnessing over distant masculine authority (Thompson, 2002). And, though this list is by no means exhaustive, of course the work of Mary Anning (and by extension Elizabeth, Mary, and Margaret Philpot) in Dorset is well-known (see esp. Torrens, 1995).

For Martina Kölbl-Ebert (2007a), this relative abundance of women in early British geology is explained by the nation’s unique combination of early industrialization and comparatively late professionalization. Compared with other European nations, British geology relied on “an informal workforce” (Kölbl-Ebert 2007a: 158) – comprising the amateur “gentlemen geologists” long since identified by Roy Porter (1978), but which also enabled women to play an active role in the science. At the same time, “a comparatively liberal gender model” in Britain, “which did not rigidly discriminate between a private, family sphere of life for women and a public sphere of men,” as well as the relatively practical nature of women’s fashions before the advent of the crinoline in the 1850s, was conducive to British women of science in a way that was unparalleled on the continent (Kölbl-Ebert 2007a: 158). If this was the “heroic age” of geology, in which the major stratigraphic divisions and the reality and character of an immensely ancient earth were worked out, we might also argue that it was also a golden age of women’s involvement, too.

This essay will reconsider the historiography by focusing on one particular type of woman geologist – the geological wife (or, in an extended permutation, the geological sister or daughter). Key women in this category include Elizabeth C. Cary, wife of Louis Agassiz (Gerassi-Navarro, 2017), and Mary Hughes, married to the Woodwardian Professor of Geology, Thomas McKenny Hughes (Hart, 2007). They include Anne Phillips, who discovered what is known as the “Anne Phillips conglomerate” while accompanying her brother on geological fieldwork in 1842, which disproved Roderick Murchison’s theory of the intrusive origin of the Malvern Hills (Morgan, 2006; 2007). Phillips also sorted and catalogued boxes of fossils, colored geological maps, conducted geological correspondence, and took care of scientific instruments. Perhaps more famously, Mary Morland was a draughtswoman for Georges Cuvier (whose daughter Clementine, and stepdaughter Sophie Duvaucel, were employed as his research assistants at the Jardin des Plantes (Orr, 2007)). Afterward, Morland married the Oxford geologist William Buckland and became his amanuensis, draughtswoman, curator, editor, cataloguer, model-maker, and later expert in marine bryozoans, even as she gave birth to, and raised, ten children (Kölbl-Ebert 1997a, 2012). Mary Elizabeth Lyell, wife of Charles, was a skilled expert in the taxonomy of fossil shells and undertook scientific translation, correspondence, and writing (Kölbl-Ebert, 2004). The list goes on. As Jim Secord has put it, participation by female family members in scientific work was so common as to feel almost routine: “after the last Geological Society meeting of the spring season,” he writes, “the researchers gathered up their hammers and their wives and set off on extensive stratigraphical tours” (1990: 25).

This focus on female assistantship – particularly in the form of wives, sisters, and daughters – has tended to emphasize two key, relentlessly familiar points. Firstly, that women were there. They may not have been admitted to the Geological Society until 1919, but they were in the field, in the cabinets, and working in diverse roles from geology’s earliest inception. Secondly, however, it is also often argued that women are either almost invisible from the record and/or were almost entirely inhibited from contributing to geology’s intellectual history (and, by extension, to the general history of science) (see Burek et al., 2010). As Kölbl-Ebert puts it, “British women at that time were allowed, or even desired, to assist and to contribute observations and data. They certainly also talked to their husbands or non-related mentors about their work, but a public display of female creativity was beyond the invisible boundary of social conduct” (2007a: 162). For Patricia Fara, more generally, “exhibiting varying degrees of meekness, they [women contributors to science] accepted being also co-opted as menial assistants who received no acknowledgement for their work” (2004). According to David Oldroyd, until the late twentieth century there were only two female “idea producers” in the history of geology: Inge Lehmann (who in 1936 suggested there might be a solid core at the center of the earth) and Marie Tharp (who proposed the idea of rift valleys in mid-ocean ridges on the Pacific Ocean Floor in 1952) (see Creese, 2007: 47). This means that there were none in the nineteenth century (and even then, Tharp’s research was dismissed as “girl talk” by one of the leaders of her research project at Columbia). As such, the overriding story we have told about the history of geology has been this: in a subordinate function to their male friends and relatives, women were permitted to color maps, pack boxes of fossils, or write letters. They contributed to the body of geological science – to its “observations and data.” But they were not allowed to be responsible for anything that emanated from its brain.

Perhaps, however, it is time to throw out this distinction between “idea” and “data” as little more than a restatement of the old, outmoded Cartesian mind/body distinction on which intellectual history once thrived. This is a history in which male nineteenth- and twentieth-century geologists were understandably invested, given that it put them on the right side of an intellectual tradition that automatically distinguished men’s work from the supposedly embodied and inferior knowledge of women and other disenfranchised groups. But, of course, the idea/data distinction has, in any case, been roundly demolished by a whole series of subfields of the history of science for several decades now. For a start – and just to take an example pertinent to earth sciences – early nineteenth-century geologists began by propounding a method of “enumerative induction” – of a radically empirical method that admitted only of limited observations stripped bare of damagingly speculative (and sometimes heretical) theoretical implication (see Laudan 1987: 168–69). This had class as well as gender implications, since navvies, quarrymen, and engineers would be permitted to send their observations to the gentlemen of the Geological Society, who alone had the education and sense to begin to make measured theoretical speculations on their significance (see Buckland, 2013: 9–10). But as is well known, this distinction between “observation” and “theory” was almost impossible to sustain in practice: to take just one key example, which will be of consequence later, the use of fossils to define stratigraphic boundaries – which looks like a resolutely taxonomic and descriptive enterprise – clearly had theoretical and ideological significance. As Secord famously described it, Roderick Murchison’s work defining the Silurian system “the method of correlation by fossils was tied to the notion of geographical expansibility, for the Silurian system could be spread internationally through the use of palaeontology” (1982: 417). Murchison advocated for the use of fossils in stratigraphic taxonomy in part because it helped him extend the territorial range of his system – which in turn, garnered the system prestige enough to make it standard. To put it another way, fossil collecting and taxonomic description might seem to be purely inductive on the surface, but in this period they were often tied into important debates about the age of the earth, the mechanisms by which events in earth history had largely taken place, the value of fossils in unveiling that story, and theoretical disputes about boundaries between geographical and temporal regions. There was no “data” without theory. There is no body of evidence without ideas. For that reason, “data-producers” and “idea-producers” cannot be so rigidly separated.

Perhaps even more importantly, since Martin Rudwick’s seminal 1976 essay on the topic, it has been clear that in geology, such activities as sketching, illustrating, fossil collecting, and the coloring and drawing of maps were key activities of geology’s theoretical work. Illustrations offered visual confirmation of unseen landscapes, argument for high-level ideas, and gave formalized expression to complex ideas about such ideas as erosion, the intrusion or deposition of molten veins in rock, or the age of the earth. As Rudwick puts it:

as the science emerged, … the causal and temporal interpretation of the observed structural configurations required – and was perhaps made possible by – the development of ever more abstract, formalized and theory-laden modes of representation. By about 1840, these forms of visual communication in geology no longer functioned as supplements to verbal description and verbal concepts; still less were they merely decorative in function. They had become an essential part of an integrated visual-and-verbal mode of communication. (1976, 152)

Since Rudwick’s article and his subsequent work in Scenes from Deep Time (1992), a wide variety of scholars have explored an even broader range of the theoretical and intellectual contributions of illustration to nineteenth-century science. This has included, for instance, exploring the tensions between a phenomenalist (or perhaps something like “empiricist”) commitment to drawing particular landscapes in situ and developing more formalized, theoretical models of illustration, to the use of illustrations to form natural-historical communities with shared sets of ideological and scientific agendas (see Klonk, 1996, esp. Chap. 3; Belknap 2018; Smith 2006; Adams 2017). We now also have a large body of evidence that writing was a tool and an instrument of scientific practice – from note-taking and description to the construction of lectures and teaching aids, to the elaboration of complex and sophisticated narratives (which themselves often constituted theoretical or hypothetical arrangements of the evidence itself) (see Shapin and Schaffer, 1985; Buckland 2013; Secord 2000, 2018; Heringman 2004; O’Connor 2007; on lectures see James 2004; Lightman 2007; Secord 2007; Carroll 2008; on the scientific article, Bazerman 2000; Gross et al. 2002; and on fieldnotes, for example, see Sanjek 1990; for the scientific book more generally see, for examples, Frasca-Spada and Jardine, 2007; Topham 2009). And of course, there is also a distinguished and vibrant body of work on museum collections, and private collections, as well as the significance of fossils as evidence in nineteenth-century geology, as critical and constitutive modes of its research agendas (e.g., Alberti 2002; Berkowitz and Lightman 2017; Rudwick, 1972).

It should hardly need pointing out then, that if writing, illustrating, mapmaking, and collecting were fundamental, and theoretically sophisticated, activities, on which the science of geology was premised in this period, then there can be no certain ground on which we continue to uphold a data/ideas distinction in our analysis of gendered contributions to the science. Such a distinction is rarely – if ever – upheld in any other context except in the continuing assumption that women (and sometimes, by extension, working-class practitioners) were excluded from participation in geology’s theoretical dimensions. But it is precisely on this distinction that the narrative of the female helpmate – wife, sister, or daughter, subservient to the intellectual impulses of her husband, father, brother – continues to rest. At the very least, this demands further scrutiny.

Once we begin to give it this scrutiny, immediate questions emerge. Do we always need to emphasize, for instance, that Anne Phillips was working “under her brother’s instructions” (Morgan, 2007: 265)? Might we not argue, instead, that she – like any self-respecting geologist – was working on a particular geological problem, the terms of which were obviously and necessarily dictated by the terms of a geological debate which included the voices of many others, as would be the case for anybody of any gender contributing to a pre-established field of knowledge? She had a brother in the same way that John Phillips had a sister, or friends, or an uncle (William Smith) who were also his interlocutors and collaborators. Why should the fact of a woman’s relation to a man immediately cause us to emphasize her lowly status in the intellectual hierarchy to which she was contributing? At least since Peter Bowler wrote of the “non-Darwinian revolution” (1988), we have dismantled the idea of the heroic, singular, original man of science working in a vacuum rather than in an entangled set of networks of support, cooperation, collaboration, and culture – so why should we perpetuate the demand for singular, solitary genius when it comes to women? It is of course significant that many women gained access to scientific participation through familial relations, and that this may have shaped the particular modes of scientific inquiry and endeavor in which they were engaged. But the assumptions embedded in the idea that they were mere “helpmates” or “assistants” because of those familial relations might mean we persist in upholding those old standards by proxy, accrediting women only insofar as they were part of the network of a masculine hero, or wishing for them to be heroes themselves.

This chapter seeks to reframe our understanding of gender in nineteenth-century geology by using the early work of Charlotte Murchison as a case study in applying these studies of discipline – of the relations between the theory, method, communication, and practice of science – to the reconsideration of questions of gender. In particular, I argue that the idea of the “geological wife,” “familial assistant,” or “female helpmate” is deeply embedded in the historiography but has taken us as far as it can as a set of (rather blunt) category of analysis. It has absorbed nineteenth-century gender stereotypes into the understanding of women’s scientific collaborations and contributions, perpetuating the long (slow)-dying heroic “man of science” motif.

Charlotte Murchison is a particularly good case study in this regard because her story is so emblematic of the accepted narrative: multiple sources tell us that it was her scientific interest that lured her husband into the field, but that illness and age saw her contributions fade as her husband took the inevitable center stage. The fact that her husband, Roderick, was not shy of the limelight, and that she seemed to be happy in a supportive role, has only underscored this narrative. As Roderick’s first biographer, the geologist Archibald Geikie, put it of Roderick’s marriage to Charlotte:

hitherto he had lived at his own free will. From this time he came under the influence of a thoughtful, cultivated, and affectionate woman. Quietly and imperceptibly that influence grew, leading him with true womanly tact into a sphere of exertion where his uncommon powers might find full scope. To his wife he owed his fame, as he never failed gratefully to record, but years had to pass before her guidance had accomplished what she had set before her as his aim (Geikie 1875, I: 70).

There could be no more stereotypical image of the Victorian angel in the house, with her quiet, imperceptible “womanly tact” designed to direct the energies and unleash the “uncommon powers” of the all-conquering male. But freeing Charlotte from the shackles of the category of “wife” and taking her contributions on their own terms – as collaborations and contributions to a shared network of ideas – it becomes possible to see a new interpretation of the accepted events of her scientific life, and to begin, at least, to articulate a different vision of the ways in which we might frame and interpret women’s scientific work in this period. In particular, I focus on three key moments in her early career: her first geological tour of 1825; her tour with Charles Lyell and her husband to France and Italy in 1828; and the admission of women (including Charlotte) to academic (rather than public-facing) geological lectures at King’s College London and the Oxford meeting of the British Association for the Advancement of Science in 1832. Doing so, it becomes possible to see Charlotte in a new light. Looking at these moments as part of her narrative, she emerges as a key contributor to one of the most significant and influential theoretical debates of the century and as a key interlocutor of many of those at the forefront of earth sciences in Britain the 1820s and 1830s. Re-examining the historiographic materials on a single figure – Charlotte Murchison – throws new light on the historiography of women in the earth sciences in general.

2 1825

The story of Charlotte “converting” her husband to science, as outlined above, relies largely on the Lives and Letters of Roderick I. Murchison by Archibald Geikie. And, while archival research does not substantially transform that story, it is important to tell it in detail to begin to appreciate the full measure of the intellectual and practical work involved in Charlotte and Roderick Murchison’s geological lives.

When Charlotte Hugonin met Roderick Murchison on the Isle of Wight in the summer of 1815, she was already a serious-minded (or, in Roderick’s words, “attractive, piquante, clever, highly educated and about three years my senior” (as quoted in Geikie 1875, I: 68)) 27-year-old with a father (General Hugonin) recently returned from Waterloo and who shared astronomical enthusiasms with his daughter; her mother was a gifted botanist whose gardens at Nursted House in Hampshire were an experimental playground for rare and dazzling cuttings, plants, and seeds. Roderick had spent the Battle of Waterloo in a depot at Ipswich, having spent (by his own account) most of his 6 years in the army taking singing lessons, flirting with women, reading trashy novels, playing whist, and drinking whisky. He could speak only a smattering of any language, while Charlotte, brought up during the Napoleonic Wars when travel to Europe was impossible for almost all, was fluent in several languages, including French and Italian. Her notebooks from their honeymoon a year later from this time are filled with joy and wonder at the “magnificent Alpine scenery,” the lake at Geneva (“nothing can be more enchanting than the first view of the Lake descending the mountains”), “the glaciers of Grindelwald and the Giesbach falls,” and the “volcanic and scattered rocky scenery” of Italy (C. Murchison 1816–18: 8, 7, 9, 25). On this tour Charlotte climbed Mont Blanc, was dragged through the Swiss snow in a carriage gone out of control, almost drowned in a felucca, and was surrounded by “a hundred vagabonds” she thought would smash her carriage into pieces (C. Murchison, 1816–18: 16).

In Italy, however, Charlotte contracted malaria and almost died. The couple returned to England to settle down at the much less glamorous location of Barnard Castle, County Durham, which was owned by Charlotte’s uncle. Charlotte made a herbarium and started a mineral collection and began to entertain the local gentry. Or, as Roderick put it, “my wife was always striving to interest me in something more intellectual than the chase, and began to teach herself mineralogy and conchology” (in Geikie 1875, I: 333). Roderick was not convinced about the “something more intellectual,” and instead sold his unprofitable estates in Scotland and began spending money on foxhounds, horses, and parties, speculating on risky foreign investments and drawing on his dwindling capital to fund it. Eventually, Roderick moved the household – against Charlotte’s wishes – to the expensive hunting ground of Melton Mowbray in Leicestershire and took on the upkeep of ten horses. The financial strain began to take its toll.

In 1825, under all this financial pressure, the couple moved to London. Charlotte acquainted Roderick with Humphry Davy and persuaded him to join the Geological Society, whose all-male parliamentary-style debates after the delivery of its papers at Somerset House and raucous dinners at the nearby Crown and Sceptre suited her husband perfectly. She also took control of the finances, began to keep records of Roderick’s spending, and set up a summer of geologizing for them both on the south coast of England, where they took lessons from William Fitton, President of the Geological Society. And, as the 2021 film Ammonite – starring Kate Winslet and Saoirse Ronan – suggests (before departing from the facts in service of a very different story), Charlotte spent part of this tour – several weeks, in fact – honing her fossil-hunting skills in Lyme Regis with Mary Anning.

However, Roderick didn’t stay in Lyme with the women. Instead, the couple set the pattern of a lifetime in terms of geological collaboration:

Driving, boating, walking, or scrambling, the enthusiastic pair signalised their first geological tour by a formidable amount of bodily toil. Mrs. Murchison specially devoted herself to the collection of fossils, and to sketching the more striking geological features of the coast-line, while her husband would push on to make some long and laborious detour. In this way, while she remained quietly working at Lyme Regis, he struck westward for a fortnight into Devon and Cornwall, to make his first acquaintance with the rocks to which in after years Sedgwick and he were to give the name by which they are now recognised all over the world. (Geikie 1875, I: 127).

This account is, of course, shot through with gendered assumptions: Roderick’s work is goal-directed, driven – he “push[es] on”, the work is “long and laborious.” Charlotte stays behind – an image of the domestic heroine waiting (though in this case “working”) “quietly” while her hero heads off into battle. Beneath that, nonetheless, is a clear account of Charlotte’s work fossil-collecting, making sketches, and “working at Lyme Regis” with a world-renowned expert in Anning. Roderick’s “long and laborious detour” into Devon and Cornwall has little to do with the work that would immediately emerge from this field trip – a paper on the geology of Hampshire and Surrey that would mark Roderick’s inaugural contribution to the Geological Society. This doesn’t matter to Geikie, because he is teeing up his readers for one of the key triumphs of Roderick’s career, the establishment of the Silurian System, which would be based partly on tours in the West Country. In the telling of that story, it isn’t of much concern what may or may not have been happening at Lyme Regis, or in Hampshire and Sussex, in these months.

But in the telling of Charlotte’s story, the south coast of England is crucial. Charlotte was not only learning how to collect fossils, but also putting together her own collection, and she had a specific agenda in mind. The cliffs of Lyme Regis are split roughly into two. Anning had made her name with her “old passion for bones,” as she described it, working in the oolitic blue lias rocks on the foreshores of Lyme Regis. Here she made her discoveries of the first ichthyosaurus specimen in 1812 (and further ichthyosaur specimens throughout the next two decades), as well as the first complete plesiosaurus in 1823. (She would find the first British example of a “flying dragon,” or pterodactyl, in 1828.) Charlotte’s interest, though, lay not in the gigantic lizards of the lias rocks, but in conchology – the study of fossil shells – and in particular in the yellowish greensands that overlay the lias at the top of the cliffs. These shells matched those of the collection she had started of similar rocks around her childhood home in Hampshire. Letters record that Charlotte and Anning went “beating bits of greensand to peices [sic] for shells” (Anning, N.D.). But they helped each other in other ways, too. In 1829, for instance, Anning wrote to Charlotte, asking her to act as a broker for “a very good head of an Ichthyo-vulgaris about two feete in length, which I value at five £” and also an ammonite “which I value at one pound” and which she hoped Charlotte could get Lyell to take (Anning, 1829). As a woman of privileged social standing with regular contact with leading men of science, Charlotte acted as a go-between for Anning and the gentlemen of the Geological Society.

Correspondence between these two women demonstrates that their relationship operated on a quid pro quo basis. Anning also writes, “I have taken the liberty to send you a few specimens which I hope you will do me the honour to accept of, the shells are all from the …greensand Mr Sowerby must name these for you and also the Ammonites” (Anning, 1829). It is not clear that Sowerby did name any of these shells in her name, though he did name the Ammonite Murchisonae after her in 1827 “as a just tribute for the ardour with which she has pursued the study of Fossil Conchology, the pleasing effects of which those who are so happy as to be acquainted with her know how to appreciate” (Sowerby, 1829, 96). (Charlotte had found the type specimen of that species on a trip to the Isle of Skye in 1826.) Charlotte brokered Anning’s oolitic specimens among the wealthy elite in London; in return, Anning found specimens for Charlotte’s own collection in the younger greensands that were of interest to her. Indeed, Anning told Charlotte that “since I wrote to you I have not found much. I soon lost the greensand mania, and returned to my old passion for bones the other day, having found a part of a plesiosaurus & so intent in getting it out that I had like to have been drowned …” (Anning, 1829).

In a later letter, Anning again expressed her thanks for Charlotte’s “kind invitation to town” and said, “I think should anything occur to prevent my accepting it will be the death of me,” and thanked Charlotte for her “kindness in mentioning my skeletons to differant [sic] collectors” (Anning, 1833). On the envelope of the same letter she wrote “… forgot to say should I get anything good and small that I will not forget that your Cabinet is not full, and also the specimens you name.” The bulk of the letter elaborates on what might be going into Charlotte’s cabinet: “you and Mr De la B.[eche] between you have given me the green Sand Mania, both Miss Philpot & myself have been beating bits of greensand to peices [sic] to find shells, and after our return we sit down an [sic] turn over the leaves of Sowerby untill our poor heads is compleate jumble of Bivalves univalves & Mr D having been so kind as to lend me his Sowerby – I think to be able to get you many specimens that are not named in this Book.” Indeed, she had “not discovered anything, since I am left Lyme, as I have been to [sic] much taken up with greensand” (Anning, 1833).

All this matters because it is usually glossed over in a sentence or two in a history of Roderick’s story. The wife is credited with “converting” an errant husband to a new faith (science), and with saving his soul from his reckless, gambling ways in the process. But here we see her engaged in her own projects: building a cabinet of shells from the Cretaceous chalks and greensands of Hampshire and Dorset, advocating for her work with Sowerby, and acting as a broker for Anning in return. Indeed, the Sussex geologist Gideon Mantell wrote to Roderick in 1830, mentioning that, “Mrs Murchison was so kind as to interest herself about some dozen [?] Lias fossils for me from Mary Anning,” and asks her to try to procure some more (Mantell, 1830). The geographer Joseph Barclay Pentland corresponded with Charlotte during this period, writing on 14 May that Cuvier, “requests me to present his kind thanks for your trouble respecting” a specimen “and to express his desire, that you would purchase for him the specimens mentioned in your letter, the price of which I shall record to Mr Murchison.” His daughters (mentioned above) “also to recall them to your recollection, & to say how happy they would be to see you again in Paris” (Pentland 1829). Six years later Pentland promised – again from Paris – to “get the shells for Lord Munster” and asked Charlotte to encourage Roderick to support Frederic Cuvier’s election to the Royal Society (Pentland, 1835). Here we have ample evidence that Charlotte was a respected naturalist in her own right, cultivating her own networks and projects, and creating a fossil collection of international standing and reputation.

These activities mean that Charlotte was also, in Oldroyd’s terms, emerging as an “idea producer” in geology in this period – and in more ways than we might expect. Roderick gave “his” first paper to the Geological Society on 16 December 1826, following this tour along the South coast of England. The paper attempted “an exact delineation” of the rocks of Hampshire, which Roderick said he had first “examined a portion of … with my friend Dr. Fitton” (Murchison, 1826: 97). Roderick presented a newly colored Ordnance map “together with such illustrative specimens of the strata and their fossils as I have been able to collect in the course of last summer” (97). And the paper described crushed and fragile shells in the crumbling sands of the upper greensand and valleys of gault beneath them that swelled into the hills of the lower greensand: three bands of rock that had once been considered just one. Below these formations was the “Weald Clay” and the “Hastings or Iron Sand.” Overall, this was an “extensive inland tract, in which all the members of the series, from the chalk to the beds below the Weald clay, appear to be more largely and uninterruptedly displayed than in any other part of England” (98).

Much of the paper is concerned with the order of superposition of the rocks in the region. But fossil evidence is offered as crucial support for the delineation of the strata and its division into smaller, more clearly conceptualized categories. Here Charlotte’s research is significant. “The most abundant and characteristic” fossils in the “firestone” or “upper greensand” – the Ammonites rostratus and the Pecten orbicularis – were used to suggest a correlation between the rocks in Hampshire and those on the Isle of Wight. As such, they contributed to the detailed clarification of an emerging history of the Cretaceous rocks of the entire south of England. The fossils “I have collected are chiefly from Nursted and Buriton, Hants,” Roderick notes. Nursted House was Charlotte’s family home, with Buriton just over 2 miles away, and it was where they had begun their tour. Even the starting point was dictated by Charlotte’s connections. But more importantly, we know that it was in fact Charlotte who did the bulk of the fossil collecting. Though Roderick doesn’t mention it in the paper, as Geikie puts it, on this tour, Roderick.

had for companion his own wife. Many a deep lane and rocky dingle did they explore together for fossils. Dr. Fitton came down to visit them and joined in the pursuit, tracing out by degrees the well-marked succession of cretaceous strata shown in that district (Geikie 1875, I: 127).

Evidence from volume 6 of Sowerby’s Mineral Conchology clearly corroborates Geikie’s claim and locates Charlotte’s centrality in this work. Sowerby figured in that volume “a Lyas fossil” ammonite “from Lyme … about 3 ½ inches in diameter,” which had been “liberally presented by Mrs. Murchison”; an “Ammonite Catullus” “found in the Malm-rock of Sussex (Upper Greensand) in a quarry at Nursted near Petersfield,” for which “we are indebted to Mrs. Murchison for the discovery”; and an “extraordinary shell,” the “Inoceramus pictus,” was one of which “we have only seen but one specimen, for the use of which we are indebted to our very scientific and liberal friend Mrs. Murchison.” That shell “was found in the Chalk Marl at Guildford,” on the same tour of Sussex and Hants (Sowerby, 1829, 106, 123, 215). Sowerby himself identified many of the fossils listed in the appendix to Roderick’s paper, but there was also an Avicula – a “new species not yet figured” – and presumably discovered by Charlotte (R. Murchison, 1826: 99).

Roderick’s “I,” therefore, subsumes the work of his wife within it. This should not come as a surprise, but as a simple acknowledgement of the system of intellectual coverture at play here. Only men were allowed to speak at the Geological Society. As such, a wife’s work was likely to be, without question, subsumed into the first-person “I” of the husband’s account. That is not to say that she did not perform that labor – nor to say that all did not tacitly understand that she had contributed substantial material or ideas to her husband’s work. It is simply a reminder of an intellectual convention that a woman’s work belonged to her husband, and that his proclamation of the work “he” had done straightforwardly included the work she had done too.

Furthermore, though the 1825 paper was only a first foray into geological publication for Roderick, it was part of a bigger research project in which he and Charlotte were engaged. Immediately following the Murchisons’ paper was one by De la Beche, also on the greensands, this time in Lyme Regis and Devon. As such, when De la Beche and Charlotte drove Anning to the “greensand mania” she described in her letters, they did so likely because both were working on a shared problem at the behest of Fitton himself. Fitton’s work in this field would be presented to the Geological Society in a series of papers between 1824 and 1836, culminating in his classic The Strata Below the Chalk (1835): a 300-page, 17-plate work that painstakingly worked out the complex and confusing succession of Cretaceous rocks in the south of England. Fitton’s work – proven partly by the detailed research of people like Charlotte and De la Beche – demonstrated (as Roderick’s paper described) that there was an upper greensand, a layer of gault, and then lower greensand rocks (all containing marine fossils). Below that was Weald Clay, not only in the part of Kent after which it was named, but also in the Isle of Wight and Dorset. These layers contained fossils of freshwater origin. Then came the ferruginous Hastings Sands (also freshwater). The Murchisons’ paper can be seen as a direct contribution to this research project – partly because Fitton himself acknowledged Charlotte’s role in this emerging set of ideas. At the back of Strata Below the Chalk is a list of the key fossils in Hampshire and Western Sussex that elaborated this sequence of strata and another list of fossils from the greensands near Nursted House (see Fig. 1): “the following List of Fossils from the Upper green-sand,” Fitton writes, “in the vicinity of Petersfield, in addition to those mentioned in Mr. Murchison’s paper on that district … I owe to the kindness of Mrs. Murchison, in whose collection the specimens are placed” (Fitton 1836: 156).

Fig. 1
figure 1figure 1figure 1figure 1figure 1

‘List of fossils from the beds below the chalk, in part of Hampshire and Western Sussex’ as it appears in Fitton’s Strata Below the Chalk (1836), pp. 156–159. All the fossils from the “Upper green-sand” are from “Mrs. Murchison’s … collection,” while many individual “Lower green-sand” specimens are also accredited to her

Furthermore, the detailed elaboration of these strata was important because it established the credence of what had, in the early 1820s, appeared to be a startling theoretical supposition. In 1822, the geologist Gideon Mantell had argued, in his Illustrations of the Geology of Sussex, that there were no ammonites, belemnites, echinites, or corals (all marine fossils) in the Weald, though they were abundant in the Cretaceous rocks above them and the Oolitic rocks below (Mantell, 1822: 67–68). As Charles Lyell would later put it, “the evidence of so unexpected a fact as the infra-position of a dense mass of purely freshwater origin to a deep-sea deposit (a phenomenon with which we have since become familiar, in other chapters of the earth’s autobiography) was received, at first, with no small doubt and incredulity” (Lyell, 1852: 226). With regular reference to Fitton’s work, Lyell notes that “the relative position of the beds is unequivocal; the Weald Clay being distinctly seen to pass beneath the Lower Greensand in various parts of Surrey, Kent, and Sussex, and to re-appear in the Isle of Wight at the base of the Cretaceous Series, being, no doubt, continuous far beneath the surface” (Lyell, 1852: 226). This mattered because it suggested something perhaps even more significant and which would be crucially important to Lyell in his groundbreaking work in Principles of Geology (1830–1833). For here was clear evidence that the earth, in a single place, had transformed itself from sea, to land (with rivers and lakes), back to sea again. Furthermore, the fossil evidence revealed that many of the species alive at the time of these great transformations had survived the changes. If that was true, the changes had to have happened relatively gradually (no comets, visitations from God, or Biblical-scale deluges complete with entirely new species epochs) – though also quickly enough that the species were roughly similar throughout all those events. The fossils of the greensands that Charlotte collected were part of an emerging vision of the earth as capable of undergoing gradual, perpetual transformation.

As such, it should be clear that Charlotte’s specimens – as much as Roderick’s headlong horsedash to Cornwall – were a crucial part of the evidence that gave Roderick his first paper at the Geological Society and inaugurated him as a member of the geological elite. Neither was this just a piece of administrative assistance Charlotte gave to Roderick, nor the work of a helpmate. The specimens she was collecting for that paper – and for her broader collection – were indisputably a key source for the growing evocation of a Cretaceous world on the south coast of England. And, though none of this would be said in the form of a Geological Society paper, Charlotte’s work played a small but significant role in the highest-level theoretical work going on at the Geological Society in the 1820s. Fossil collecting was a significant part of the highest-level “idea producing” or theoretical debate at the Geological Society, and at least one woman was a direct collaborator on that research.

In this way, fossil collections ought to be read as forms of publication, and as of especial significance for women. As a point of comparison, Cynthia Burek has demonstrated that Benet’s fossil collection was highly regarded: she contributed the second highest number of specimens of all contributors to Mineral Conchology. “However,” Burek writes, “she was not a prolific writer and during her lifetime she only published two articles, both fossil lists from her extensive collection of thousands of Cretaceous and Jurassic specimens, many from the Upper Greensand” – the exact area in which Charlotte was also working (Burek 2001, 193). The point is important. Given that Charlotte could not read a paper at the Geological Society, and thus formally publish in the Transactions, a fossil collection would have been a key means by which she engaged and participated in scientific discussion. And for several purposes, it also had distinct advantages over publication in print. If the rapidly progressing theories debated at the Geological Society were ever-shifting sands, which could be washed out with each new tide of information, a fossil collection could be instantly responsive to new directions in the field. A well-made collection was not the output of an assistant with no “ideas”: it was both the constitution and the expression of the intellectual work of geology.

The most cursory glance at geological books and papers in the 1830s reveals traces of Charlotte’s collection. It’s not just her regular appearance in the appendix of Fitton’s Below the Chalk, it’s also a fossil ammonite and a sketch of a dragonfly collected by Charlotte in Gideon Mantell’s Wonders of Geology (1838, see I: 460 and 399 for descriptions of the plates); it’s a nautilus striatus Charlotte gathered in Whitby and an ink bag, both illustrated and published in William Buckland’s Geology and Mineralogy with Reference to Natural Theology (1836, plates 33; 45, fig. 4). Or, it’s an Ophiura Egertoni from her collection from Whitby, cited in an 1835 paper given at the Geological Society by William John Broderip (Broderip, 1835, plate XI, fig. 5). Trilobites Charlotte found in black schistose rock at Mount Pleasant, near Carmarthen in Wales, figured in Roderick’s Silurian System, alongside her many illustrations for that book (1839, described on 664). And if we don’t have a clear picture of Charlotte’s collection and its significance now, it is literally because Charlotte was a woman: her collection was posthumously broken up and absorbed into Roderick’s collection by the Geological Society, assumed to be of no significance beyond the name of the male with whom she was most closely associated. That does not mean, of course, that that assumption was correct. And in 1828, Charlotte’s collection would be dramatically expanded and acquire new theoretical significance that clearly demonstrates her status as an “idea producer” in the earth sciences at this time.

3 1828

In his magisterial Worlds Before Adam, Martin Rudwick describes the tour Charlotte and Roderick undertook with Charles Lyell to France and Italy in the summer of 1828. As Rudwick puts it, Roderick “was ambitious for some more challenging fieldwork, which he tackled with military zeal. Charlotte had learned from Mary Anning how to be ‘a good practical fossilist’ or collector and like Buckland’s and Mantell’s wives she was a fine artist; both skills made her a valuable companion in the field” (Rudwick, 2008: 259).

Though there is acknowledgment of Charlotte’s labor here, there is also a subtle sidelining of Charlotte’s agency and desire, which rings with echoes of Geikie. All the action is driven by Roderick’s desires – his “ambition” for “challenge,” the use of Roderick’s (frankly lackadaisical) “military” career as a metaphor for his industry and activity. As in Geikie, Charlotte’s link with Anning again appears to suggest a quieter, more contemplative mode of relation to science – she has “practical” skills, and is no different from other artistic “wives” and “companion[s].” Note the shift from past simple (Roderick “was ambitious” and “he tackled” the fieldwork with “zeal”) to past perfect tense to describe Charlotte’s actions (she “had learned” to collect fossils). Charlotte’s skills are a settled fact, while Roderick’s pursuit of new knowledge is continuous and ongoing. In the next sentence, Rudwick writes that not only was Charlotte a good companion, but “so was Lyell, as a more experienced and knowledgeable geologist and a much better linguist” than Roderick. It should be noted here that Charlotte was also a “much better linguist” than her husband, and that her field notebooks from this tour are written in English, French, and Italian, but since these are all presented as secondary skills this is perhaps, at this stage, a quibble (Rudwick, 2008: 259).

More importantly, Rudwick notes that this “great tour on the Continent, lasting almost a year … was, as it were, Lyell’s Beagle voyage,” the journey on which he really learned to see the world in a new way, and which was critical in the production of Principles of Geology (2008: 260). This was the tour on which Lyell would witness for himself the power of simple, ongoing geological processes like erosion to carve out valleys and other large-scale structures: for Lyell, this would prove that geological processes operating at present-day intensities had been powerful enough to shape earth history on a grand scale. On this tour, he would also begin to use percentages of extant and extinct fossils to describe the gradual shifts between geological epochs, moving from the most recent fossil beds incrementally further into the past. In other words, on this tour he would come to his uniformitarian vision of the earth.

Kölbl-Ebert’s account of what happened in 1828 is more attentive to Charlotte’s position in the team that elaborated this new theory. She notes, for instance, that having ascended the Puy de Dome on 18 May 1828, the trio established the pattern of work that would be most efficient. They undertook multiple field expeditions together and, as on the south coast of England, Roderick (and here Lyell, too) kept “track of correlations and bigger structures,” while Charlotte did “much of the time-consuming fossil-hunting, sketching landscape and geological structures, and – speaking French fluently,” visiting “local experts” (2007b: 111). As Lyell put it, Charlotte was engaged “in making panoramic sketches, receiving several of the gentry & Professors to whom we had letters in the neighbourhood & collecting plants & shells etc.” (as cited in Kölbl-Ebert, 2007b: 111). Crucially, as Kölbl-Ebert also points out, Charlotte’s work was then integrated with that of Roderick and Lyell: “On the 20th May,” Roderick wrote, “we were making sections of the tertiary lacustrine deposits we had examined at Verduaison above Cusset & Vernet (see Blue book & Lyell’s sections) & Lady M’s notes” (as cited in Kölbl-Ebert, 2007b: 111). This was explicitly team-like behavior.

Nonetheless, even Kölbl-Ebert continues to use language like “assistant” and “helpmate” to describe Charlotte’s participation, claiming that the men decided “route and research topics” (2007b: 109). In doing so, she supports the idea that women could do scientific work but were debarred from scientific thinking – or, at least, from being able to use their thoughts to direct research agendas. Partly this is because Lyell himself described Charlotte as “an invaluable assistant” and says that, “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” (as cited in Kölbl-Ebert, 2007b: 111). And there are methodological issues: Charlotte’s diaries are actually field notebooks and are sparse in color and detail, meaning that we have a much more vibrant picture of the tour as given by the men in their more voluminous and verbose correspondence. But we should not always take these men’s words at face value. Was Lyell likely to have understood the delicate balancing act it required for Charlotte to secure her participation in the team – always deferring to masculine authority in order to gain access to the work in the first place? Furthermore, both men – like Geikie and others after them – were clearly invested in the narrative of masculine, military action in the field with female assistance in the domestic sphere. This narrative structured their understanding of the contributions made by each member of the team in gendered terms. But this does not mean that Charlotte was not, in fact, participating in the intellectual processes that contributed to the development of new knowledge in France in 1828 or anywhere else. It might rather mean that, even when she was doing serious amounts of the practical and intellectual toil, Roderick and Lyell nonetheless understood her role as mere assistance and therefore underplayed it even to themselves. This is not an accusation of a conspiracy: rather, these men had simply not been trained to see women’s intellectual labor on equal terms with their own.

This is perhaps easiest to see in Lyell’s account of an illness Roderick contracted on the 7 August while the party were at Fréjus, and which forced them to remain in Nice from 9 to 27 August:

Murchison has not yet regained his strength from a severe attack at Fréjus, so as to be able to take the field again. But we have been hard at work in writing, from our materials, a paper on the excavation of valleys, which is at last finished, and after two evenings’ infliction, is intended to reform the Geological Society, and afterwards the world, on this hitherto-not-in-the-least-degre e-understood subject. Besides this mighty operation, we have performed two jaunts with Mrs. Murchison, each at half-past four o’clock in the morning, to see certain deposits of fossil shells, and collecting these, with which she has been much pleased; and this and the cessation of eternal bustle, while the campaign was at its height, while were as yet only crossing the Balkan, and before our descent upon these hot latitudes has restored her health and spirits, which had failed sadly (as cited in Kölbl-Ebert, 2007b: 114).

The language of the “campaign” and “crossing the Balkan” draws on the well-rehearsed metaphor of fieldwork-as-battle, of course, but there are two other things of interest here. First, the subject of the paper – “intended to reform the Geological Society, and afterwards the world” – and second, the pronoun. This time it is not an “I” that subsumes female intellectual labor, but a “we” that obfuscates our understanding of it. Who is the “we” who undertook the “hard … work” of “writing” this “paper on the excavation of valleys?” This is a matter of longstanding discomfort. Both Lyell and Roderick’s names appear on the paper, so the “we” would appear to belong to them. But at the same time, throughout his career – both before and after this paper – Roderick was vehemently opposed to the explanation of geological processes the paper presents. When the three geologists were in Le Puy, Rudwick reminds us, “Murchison proved to be a staunch ‘diluvialist,’” arguing that (as he put it himself) “gradual erosion could never produce such broken irregular effects,” and instead arguing that geological processes (like wholesale floods) must have operated at much greater intensities in the deep geological past (Rudwick, 2008: 267). Kölbl-Ebert points out that, “years later, when he compiled his journal from his field notes, he thought it necessary to take refuge in more violent causes of change” to explain the valleys of Auvergne (as cited in Kölbl-Ebert, 2007b: 115). At this time, Murchison wrote:

I must if I ever have time review what Lyell (for he was the chief penman) wrote in a joint paper on the excavation of the valleys of Auvergne. For, whilst I fully admit the re-excavation of certain valleys & the clearing out of the lava & basalt which filled them, I do not believe that the original water-courses were formed by the streams, but they originated in cracks & depressions formed by great & ancient Earthquakes, long anterior to the eruptions of the youngest volcanoes of Auvergne. (as cited in Kölbl-Ebert, 2007b: 115)

As Murchison’s hint that Lyell “was the chief penman” suggests (alongside his desire to “review” a paper he was supposed to have written just to check what it actually said), these “great and ancient Earthquakes” were the exact opposite of the processes described in the paper. The paper itself wholly advances the position that simple erosion had created the valleys and that, by extension, “modern causes” (geological processes operating at present-day intensities) could explain earth history. This paper showed that the Sioule river had cut a gorge through basalt lava and the underlying gneiss. The lava was traceable to a recent volcano, and the cones of volcanic ash it had left behind would have been swept away had a large-scale flood (like Noah’s Flood, as men at the Geological Society like William Buckland or Charles Daubeny believed) actually occurred. Furthermore, a bed of gravel underlying the rock showed that the lava had flowed down a valley that existed before the volcanic eruption. As the engraving that constituted the central proof of the paper clearly demonstrated (Fig. 2), these processes were longstanding, gradual, and continuous.

Fig. 2
figure 2

Charlotte Murchison’s important illustration to Lyell and Murchison’s 1829 paper in the New Edinburgh Philosophical Journal (plate between pages 45 and 46)

As Rudwick describes the engraving, “this sketch was published by Lyell and Murchison in their article (1829) on the power of rivers to erode valleys in hard rocks, even in a geologically brief timespan, without invoking any exceptional ‘diluvial’ agency,” and it clearly demonstrates the argument. “The case,” Rudwick goes on, “was similar to several that [George Poulett] Scrope had already described and illustrated …; but this engraving, published with the article in Jameson’s Edinburgh periodical, would have been seen by many more geologists than Scrope’s expensive album” (Rudwick, 2008: 268). Indeed, the image was crucial to the articulation and demonstration of the argument, as well as to its persuasiveness among the members of the Geological Society, because it proved, visually, a phenomenon that had already been noted by other geologists like Scrope and Croizet – but which words had not yet been persuasive enough to make scientific fact.

This particular engraving – with its crucial role in “reforming the Geological Society, and afterwards the world,” as Lyell had put it – was one of three published with the paper in the Edinburgh New Philosophical Journal (Figs. 3 and 4). The second (Fig. 3) demonstrated that the “tertiary lacustrine formations of Auvergne” had “been excavated to a considerable depth prior to the operations which gave rise to the trachytic breccias and conglomerates overlying alluvial beds, containing bones of extinct quadrupeds” (Lyell and Murchison, 1829, VI: 47). In other words, the valley had been eroded before the volcanic explosion. The third was “a transverse section of Mont Perrier, exhibiting in detail the character of all the strata of trachytic conglomerates and alternate with alluvions containing fossil bones of extinct quadrupeds” (VI: 47). Together, these images explained, demonstrated – and proved, to those who had not been able to travel to Auvergne themselves – that fluviatile erosion had created these valleys, and that modern causes were at least hypothetically sufficient for explicating geological structures that seemed to have been created by more violent events.

Fig. 3
figure 3

Charlotte’s second illustration to the 1829 paper, published as a plate between pages 48 and 49

Fig. 4
figure 4

Charlotte’s third illustration, on a plate between pages 48 and 49

And so we return to the “two evenings infliction” Lyell described – the “writing” of the paper undertaken by a “we” that has been, for two centuries, assumed to include Roderick Murchison. Rudwick chooses to focus on the two men, whom he credits with having “published” the images. Nobody credits the person who actually produced them: Charlotte. Still, members of the Geological Society widely understood her to have been their author: Scrope wrote in a letter to Roderick on 4 December 1828 that he very much enjoyed the “battery against the Diluvian Heresy” the paper represented. “At all events,” he continued, “I much wish to see you and gossip of Auvergne and my old cronies the Puys … I am anxious to renew my recollections of Central France from your conversation and Mrs. Murchison’s drawings” (Scrope, 1828). Scrope explicitly uses the drawings as tools to rethink the Auvergne when he is not actually there and to re-engage in a theoretical debate in which he had much at stake. And in the 1914 edition of Lyell’s Antiquity of Man, edited by the geologist R.H. Rastall, Charlotte is credited not only as the illustrator of these three figures, but as a co-author of the paper itself: Lyell’s published oeuvre includes “three papers with Sir Roderick and Mrs. Murchison; Edinburgh Philosophical Journal, 1829; abstract in Proceedings of the Geological Society 1; Annales des Sciences Naturelles 1829” (1914, xi). Rastall’s bibliography casually names Charlotte as a co-author some 80 years after the fact, but it does so because it accepts illustration as constitutive of the work the paper represents. He can only have done so because it was common knowledge that Charlotte had produced the illustrations.

To this evidence, we can add Lyell’s account that “we have performed two jaunts with Mrs. Murchison, each at half-past four o’clock in the morning, to see certain deposits of fossil shells, and collecting these, with which she has been much pleased.” The “we” here is clearly an inclusive first-person pronoun, referring only to himself – we already know that Roderick was too sick to leave his bed. It goes hand in hand with Lyell’s attempt at breezy language here – “jaunts,” “pleased,” “restored her health and spirits” – language designed to make these fossil-collecting expeditions as purely touristic, or even therapeutic. They almost appear as little holiday excursions to collect pretty fossils for the lady’s cabinet. But in fact we know that they were serious: George Antonio Risso, a celebrated conchologist, led Lyell and Charlotte over the deposits at Nice – an area unusually rich in the well-preserved fossils that would later be critical in Lyell’s naming of the Pliocene epoch (Rudwick, 2008: 270). We also have evidence that in May 1829 Charlotte wrote to Croizet to ask for fossil eggs from Auvergne, and later chased this up with Joseph Barclay Pentland (though the letter we have implies that he never actually sent them) (Pentland, 1829). So we can assume that Charlotte was fossil-collecting with Lyell in a way that was strategic to the aims of the 1828 paper, and building a rich and complex set of fossil evidence for her collection which was in significant dialogue with emerging geological theories.

In the face of all this evidence, the idea usually advanced for Roderick putting his name to the 1829 paper on valley erosion has been decidedly weak. “Roderick Murchison was uneasy,” with the paper, Kölbl-Ebert writes, “even if in the presence of Lyell he had been persuaded to agree” (2007b: 115). For Rudwick, “Murchison may have been persuaded by his companion – not for nothing was Lyell a trained barrister – or perhaps by the sheer perceptual impact of ‘the beautiful pet volcanoes’ of Vivarais” – still, this was only a “nominally joint paper” (2008: 268). So Murchison was unwell in bed, later disavowed the paper’s contents entirely, admitted he didn’t really write it, or recall exactly what was in it. And yet the scholarship persists in attributing the article to him on the grounds of Lyell’s persuasive rhetoric. Surely, it is less of a stretch to argue that, since Charlotte was collecting fossils for the paper with Lyell and drawing the illustrations that would constitute the bulk of its proofs, she was the actual co-contributor, and the use of Roderick’s name on the paper signifies her labor rather than his.

This raises a second, interesting question. If Charlotte was a collaborator on this paper, did she also agree with its contents in a way that her husband did not? Perhaps Roderick Murchison was a diluvialist. That does not mean that his wife was, too, and only the unquestioning idea – so deeply embedded in the historiography – that women plodded unthinkingly in their husbands’ intellectual footsteps, as “assistants” rather than genuine collaborators, can blind us from seeing that Charlotte may have had her own mind on the matter.

4 1832

With that point, we come to the final section of this paper, and another point of difference between Charlotte and the two men with whom she went on tour in 1828. As is well-known, in 1832, women were included in academic geological lectures for the first time: first at Lyell’s lectures at King’s College London and then at the British Association for the Advancement of Science. Charlotte’s presence at both these sets of lectures has often been noted, and Kölbl-Ebert goes so far as to say that “it was her wish to attend, that opened Lyell’s geological lectures at King’s College to women” (1997b, 41). But this has not been clearly linked to her intellectual contributions to the development of geology in this period. I want to finish this essay by arguing that it was Charlotte’s fieldwork – in tandem with Mary Somerville’s publication of Mechanism of the Heavens in 1831 – that made the issue of women’s attendance at academic lectures so difficult to avoid.

In the autumn of 1831, Lyell was preparing his lectures for King’s and he attended the first meeting of the BAAS in York, which women had attended by default, since the fledgling institution seemed to be of public rather than scholarly bent. In early November, he “went by invitation to drink tea with the Murchisons, and was thus kept to geology till half-past eleven o’clock at night … Mrs. Murchison was in high spirits, and has evidently enjoyed their tour, and above all the York meeting” of the British Association for the Advancement of Science (in K. Lyell, 1881, I: 350). “Her real delight in conchology is of great use to her in taking away the tedium of long sojourns in out-of-the-way places,” he wrote to Mary Horner, the daughter of a former President of the Geological Society and the woman he would marry the following year, “and she had a fair opportunity of collecting” (in K. Lyell 1881, I: 350–51). With excitement, he added that the Murchisons had “found recent Irish Channel shells in abundance, 300 feet high on the Lancashire coast! I am delighted, because the identity of the living species of plants and animals in England and Ireland made me argue that some great changes in the distribution of land and sea had taken place in that part within the modern era” (in K. Lyell 1881, I: 351). Much like the interchange of marine and freshwater fossils on the south coast of England, here was another example of the shifting lands and seas created by ordinary geological processes, with further evidence supplanted by Charlotte.

At this time, Lyell was also planning to “use our joint and unpublished notes on Auvergne” in the second volume of Principles (in K. Lyell, 1881, I: 355). In honor of that fact, “and because we had together part of the best and longest geological tour which I ever made, my tour which made me what I am in theoretical geology,” Lyell was planning on dedicating one of its volumes to Roderick (in K. Lyell, 1881, I: 355). Once again, we should not take this to mean that he was only making a dedication to Murchison. Lyell’s letters are full of admiration for women like Cordelia Mallet, married to a seismologist, and whose “considerable turn for natural history” he much admired (in K. Lyell, 1881, I: 370). Lyell also recalls fossil and fern-hunting expeditions with his sisters Marianne, Eleanor, and Caroline near their ancestral home of Kinnordy in Scotland (in K. Lyell, 1881, I: 338, 341, 359). He repeated (albeit to his would-be wife, Mary Horner) the geologist Thomas Webster’s supposition that women possess a greater “liberal curiosity” for science than men whose heads had been stuffed with the dead languages of Latin and Greek, and he greatly admired his friend Mary Somerville, whose technically audacious, mathematically dazzling translation of Pierre-Simon Laplace’s Mechanism of the Heavens had just been published (1881, I: 323). “There are not five persons in this country that could have written that book, if that number,” marveled Lyell, noting seriousness that most mathematicians at Cambridge weren’t clever enough to understand it (1881, I: 374). Lyell’s world was rich in intellectual women with whom he was willing to interchange ideas and labor. But, of course, he also wrote of Somerville that had she “been married to La Place, or some mathematician, we should never have heard of her work. She would have merged it in her husband’s, and passed it off as his.” For, while “a man may desire fame, reputation, and even glory, for the sake of sharing it with one he loves,” not so “a woman,” who “cannot share it with her husband,” because “it will be the utmost she can do not to make him of less importance by it” (in K. Lyell, 1881, I: 325). The man who wrote this letter was more than capable of thanking both Murchisons in print by mentioning only the husband. Perhaps, as he began to contemplate matrimony himself, he had the model of the Murchison’s marriage – as joint intellectual collaboration – in mind.

Nonetheless, the lectures at King’s were supposed to be academic lectures and for men only. Lyell wrote that on a Wednesday in May 1832, there were “grand disputes at the Geological Society” – still a male-only institution –.

about the propriety of admitting ladies to my lectures. Babbage most anxious to bring his mother and daughter and Lady Guildford; Harris to bring Lady Mary Kerr; and so on. I begged them all not to do so, and they promised; but at last Murchison said, “My wife, however, must come. I promised to bring her, and she would be much disappointed. I will not bring her till the doors are closed.” Then they all declared they would too, and so bring the affair to a crisis one way or other. (1885, I: 381)

As Rudwick puts it, “Lyell himself was at first opposed to the request. His reason is clear: the admission of women to the lecture room would be ‘unacademical’” (1975: 241–42). Rudwick, in fact, concurs: “This view is consistent with his great concern for the academic prestige of the College, and hence of the Chair that he held and of the lecture that he gave” (1975: 242). We know from Rebekah Higgitt and Charles Withers’s study of female audiences at the British Association for the Advancement of Science that it was generally presumed that female audiences were there for social purposes, celebrity-spotting, or perhaps to show off a new bonnet (2007: 9–13). But this view is also inconsistent with what we know of the deep networks of scientifically inclined women with whom Lyell was acquainted. And it overlooks a key fact of the debate at the Geological Society itself – that the decisive intervention was made by Roderick Murchison. It appears from Lyell’s accounts that other men at the Geological Society had agreed not to bring their wives and daughters until Roderick intervened. Partly this may be attributable to Roderick’s famous sense of entitlement – if he and his wife wanted to come, nothing would stop them. And yet this intervention caused the other men to re-join the cause. As such, we need to ask ourselves: what precipitated this crisis in the first place? Why were all these men so willing to argue the case for women at this point, and seemingly at no other? Why did Roderick say that he would bring Charlotte, ban or no ban, and sneak her in after the doors were closed, and why was it this intervention that made the crucial difference?

Two days after the battle at the Geological Society, “the ladies, Dr. Fitton, and others” went “assailing … the King’s College Council to admit ladies,” until the Council relented (in K. Lyell, 1881, I: 381–82). Fitton’s name is perhaps not insignificant here, given that he was Charlotte’s first teacher in geology and that he had credited her collection in Strata Below the Chalk. Furthermore, the “ladies” here included Betsy and Georgiana Babbage, Susan Countess of Guildford (daughter and joint heiress of the banker Thomas Coutts), Somerville, and Charlotte. Rejecting women like this would mean turning away some of the most illustrious and scientifically minded women in London, and we know from Lyell’s correspondence that – along with “7 titled ladies,” in attendance at the lectures were “Mrs Somerville regular and Mrs Murchison” (Lyell, 1832).

The instant he heard the decision that women could be admitted after all, Lyell wrote to Charlotte Murchison. “The important affair of whether lady geologists are to be trusted within [the] walls” of the College had been settled, he told her. And, in a classically Lyellian move of diplomacy, he added that, though he “had always protested abt. such innovations as un-academical, of course” he did “not object” to their decision now it had been made. Still, there were conditions: the Council had agreed to “assign a separate entry & a particular part of the theatre to the ladies, & their chaperones,” who were “not to be allowed too great propinquity to the caps and gowns.” Since King’s was situated in the theatre district, the College would also “make arrangements to know any lady who comes, in order to prevent any improper person from coming” (as cited in Wilson, 1972). It is hard to tell if this is meant as a reassurance to Murchison, who would not have to worry about the reputation of “academical” women being lowered by the presence of “improper person[s]” – or, instead, if it is intended to put her in her place and remind her of the privilege and the responsibility she bore in attending the lecture as one of her sex.

Nonetheless, Lyell was writing to Murchison, I suggest, because it was her wish to attend that was critical in this debate. A key part of Lyell’s argument in his lectures was that “molluscs were the most suitable fossils for geological dating” (Rudwick, 1975: 249–50) and in his ninth lecture he “presented and defined – for the first time in public – the periods of Tertiary time which this research had distinguished: the Recent, Newer Pliocene, Older Pliocene, Miocene and Eocene, in order of increasing disparity from present faunas and inferred increasing age” (Rudwick, 1975: 250). This came complete with a Synoptical Table of these fossils and an example drawn from Auvergne, which provided a “perfect restoration of [the] Eocene period, [its] lakes, seas, volcanos, land quadrupeds, reptiles, testacea, plants, insects, Geography” (Lyell, as cited in Rudwick, 1975: 250). Lyell’s argument partly addressed the theory of fossil evidence in geological dating: he believed that each period of the geological past should no longer be defined not by a single “characteristic” fossil which identified the age of the rock in which it was found – for what if that characteristic life form had left no fossils intact? – but by a statistical analysis of all the fossils which had miraculously survived in each formation. As such, in the most recent bands of Tertiary strata, 90–95% of the fossilized species were still extant; 30–35% of the fossils in the next band; 18% in the next oldest, and just 3% of the species found in the oldest (but still geologically recent) bands. The crucial point here is that, in the back of Murchison’s field notebook from the Auvergne trip, key evidence that made this argument possible is on display: there is a list of Tertiary fossils in Auvergne, and also near Bordeaux, complete with information about the rocks in which the fossils had been found. Charlotte notes 120 terrestrial species of fossils across France, 36 species of lake-dwelling univalves (water snails and slugs), and 11 species of bivalves (molluscs with hinged shells, like oysters and mussels) (C. Murchison, 1828: 30–36). As such, Lyell was presenting work that Murchison had been intimately involved in collecting from the outset. How could he disbar her from hearing about the evidence she herself had contributed to? How could women be prohibited from hearing lectures whose content they had co-created?

To add further corroboration to this theory, in the following year, Lyell married Mary Horner: Rudwick simply says he “took that long-suffering lady on a geological honeymoon” (1975: 252). In fact, Mary was not “long-suffering” at all, but became a distinguished conchologist in her own right. It seems that Lyell had this intention in mind from the beginning: both he and his new wife, alongside his sisters, began training in conchology with George Sowerby (son of James) that same spring. “I get on in conchology somewhat,” he wrote in May 1833, while the women “make rapid progress” (Lyell, 1881, I: 397). Having spent a summer with Murchison and having been able to make a key set of theoretical and methodological arguments based on the analysis of fossil molluscs she had collected alongside him, Lyell knew the importance of a geologically trained partner and sought to improve his own conchological skills into the bargain. Perhaps an added bonus of marrying Mary was that he could replicate the pattern of the Murchisons’ marriage and avoid the sticky problem of crediting other people’s wives with work that did not really represent their husbands’ views.

Murchison also pressed home the necessity of women’s attendance at the second meeting of the British Association for the Advancement of Science, which took place at Oxford in the same year as the King’s lectures. At the inaugural meeting at York in 1831, few scientific men had wanted to lend their names to the fledgling project and women had been let in as they usually were at public (rather than academic) occasions. But when the Association moved to Oxford in 1832, this seemed an uncomfortable precedent. The Oxford Professor of Geology and Mineralogy, William Buckland, wrote to Roderick, who had presumably raised the matter with him, to say that “everybody … agreed that if the meeting is to be of scientific utility Ladies ought not to attend the Reading of the papers – especially in a place like Oxford.” The Association wasn’t for hobbyists. It was “a serious Philosophical Union of Working Men.” Still, women “at parties is quite another thing & in this I think the more Ladies the better,” he added (as cited in Higgitt and Withers, 2008: 6).

Nonetheless, Murchison attended the BAAS enthusiastically and slept in Buckland’s attic (she became a godparent to the Bucklands’ most recent child during the meeting). At this and subsequent BAAS meetings she not only attended parties, like the one in 1833 in which she searched with William and Mary Buckland for double stars and shooters at the home of the astronomer William Henry Smyth. (She was impressed when his wife Eliza handled the telescope adroitly in his absence.) (C. Murchison, 1833–35: 2–5). She also went twice a day to lectures no matter what anybody had to say on the matter. The following year, at the BAAS meeting in Cambridge, she wrote happily in her diary that, “ladies were permitted to have seats and platforms on every side and the most marked attention was paid to their comfort” (C. Murchison, 1833–35: 2). Given the context, these comments are highly charged: we can conclude that Murchison insisted on her attendance at these lectures, and the significance of her work made it almost impossible for her male friends and acquaintances – her colleagues – to deny her.

At the next year’s meeting of the BAAS, in Cambridge, a friend of the self-educated Norfolk geologist Samuel Woodward found the scientific bigwigs at the meeting a bunch too “ponderous” for his liking. He thought they saw men like him – doing a lot of the work, while the gentlemen took their research and wrote it up to their own credit often without detailed local knowledge – as nothing more than “insignificant laborers.” But it was Murchison who annoyed him most of all:

Among the authorities, Mrs. Murchison ranks high, as I believe you told me, and I had an example of it; on the Friday morning, Smith and myself were conversing about the Bilney shells, when up came Mr. and Mrs. M. Smith asked Mr. what the shells were, he hum’d and ha’d, and then appealed to the grey mare saying “she knowed most about them,” and so it proved, for she immediately pronounced them to be London clay fossils. What do you think of petticoat government? (Rose, as quoted in Woodward 1879: 580)

He sneers at “authorities” who make grand sweeping statements on things they don’t know in detail. He is gratified that Roderick could barely string a sentence together on the shells he had collected in West Bilney. And if Murchison did indeed attribute the shells to the London Clay, then he was right that she was mistaken. But more important are the politics here. Class intersects with gender: Murchison is both pontificating “bigwig” from the city, overriding local expertise with haughty (and misguided) grandeur, and “grey mare,” ridiculed for her appearance, her age, and gender. And the greatest of her affronts here is the sheer fact of her intellectual authority, the “petticoat government” to which even her husband publicly defers. The quotation succinctly makes the point: nineteenth-century male naturalists had no incentive to credit or even recognize women’s work that was not seen as mere assistance. That does not mean it was mere assistance, or that those men were not at the same time aware of the contributions such women were making. It means there was a doublethink at play, which both acknowledged and disavowed women’s intellectual labor in the exact same moment.

5 Conclusion

This essay has only been a suggestion of what might be possible should we stop viewing women through the lens of “wife,” “sister,” and “daughter.” There is still much work to do, even on Charlotte Murchison. Nobody has assessed, for instance, her many contributions to The Silurian System, though her illustrations are a significant contribution to that work, or paid serious attention to the scientific salon she ran at the couple’s house in Belgravia on Sunday afternoons for several decades. And a reworking of the broader historiography lies outside the scope of this paper. Women like Mary Buckland made perhaps even richer and more significant contributions to scientific work in this period, and require more substantial study. (A year after the Oxford meeting Charlotte would famously find “Mrs. Buckland not very well, quite overworked assisting at the birth of his [William’s] essay- too bad with so many births of another description to endure,” pointing to the ways in which Charlotte was aware that some of this female scientific labor could also take the form of exploitation as well as mere “assistance” (as cited in Kolbl-Ebert, 2007b: 161))

It should also be acknowledged that almost every source used here has been published or cited by historians relatively widely for at least two or three decades. This is not new material. But a slight historiographical reframing opens up new questions about women’s contributions to science and helps us reconsider the creativity of women’s intellectual work. Partly this is needed because the stories that were told about women in the nineteenth century still loom large in twenty-first century imaginations. In 1839, a naturalist wrote to Roderick Murchison about Etheldred Benet, whom he encouraged to send her fossils to the Bonn Museum. Moreover, “a Dr Bunsen from Göttingen was here lately,” he continued:

& I was shewing him one of your papers when he asked me if you were any relation of the celebrated lady of your name of whose geological knowledge he had heard Professor Hausmann speak so much – when I told him the relationship he replied “What! Is she married? I did not know that married ladies could attend to such things” – I leave you to judge whether it would be safe for you to tell Mrs. Murchison this.

This is Robert Bunsen, whose name is immortalized by the Bunsen burner. And again we see evidence of Charlotte’s international reputation independent of – even preceding, in some cases – Roderick’s. But the letter is also startling in its very familiarity, in the easy way it passes off Murchison’s knowledge – and her likely irritation – as a joke between men (“I leave it to you to judge whether it would be safe.”). And the very fact of the familiarity of this rhetorical maneuver also asks us to take more seriously, now, the ways in which women’s contributions to geology were not quite so simply hidden behind the achievements of their husbands or brothers as we have been accustomed to think. Of course, women were never valued equally: there was a sophisticated panoply of techniques that reduced even their collaborators’ understandings of their work to mere frivolity or decoration – perhaps even to a joke. In one breath here, Murchison is acknowledged as the foremost expert – but in the second, her expertise is relegated to a petty laugh. Nonetheless, it is this cognitive dissonance around female scholarship which means we can profitably reassess the ways in which geological wives, sisters, and daughters were regularly consulted for knowledge, materials, and ideas and contributed significant scholarship to the nineteenth-century earth sciences. With Charlotte Murchison in mind, it is surely time to shake off the stereotypes and stories of the past and to remember that the “geological wife” or daughter was a figure of significant agency, purpose, and intellectual significance and is worthy of renewed attention and understanding.