Select Page

Women Scientists in America Before 1920

Autumn 1973 — Charles Warren Center Fellowship Talk, Harvard University, Cambridge MA

 

If asked to name a woman scientist, most Americans would probably mention Marie Curie, the Frenchwoman famed for her researches on radioactivity and her two Nobel Prizes. If pressed to name an American woman scientist, most would probably refer to Ruth Benedict, Margaret Mead, or Rachel Carson, whose books have gone through numerous editions in recent decades. But few would know of any other women scientists. This lack of awareness has, unfortunately, led to two widespread but incorrect impressions: 1) women scientists are isolated phenomena and 2) it is only very recently that women in America have begun to take up careers in science. It may come as a surprise, therefore, to learn that women have been a part of the American scientific community for over a century, and that, far from being isolated phenomena, they constituted a sizable group of strong-willed individuals, many of whom worked hard to overcome those very problems of rejection and discrimination that have become an object of concern in recent years.

When in 1906 James McKeen Cattell, editor of Science and professor of psychology at Columbia University, published the first edition of his American Men of Science: A Biographical Directory, he included 149 women on his list of 4,131 scientists (3.6%). His second edition, in 1910, contained 204 women, but their percentage of the total dropped slightly to 3.5%. By 1921, after a decade of rapid growth, the number had more than doubled to 459, but, since the number of male scientists was also increasing, the percentage of women rose only to 4.8%. Together these first three editions contained 504 women who were seriously interested in science and wished to pursue careers in it before 1920.

Although we know very little about the early women of science, these directories provide fairly complete curriculum vitae for each individual and attempt to assess the quality of her work. They thus provide one way to study the career patterns of a large number of women scientists before 1920. Their situation is of contemporary interest, for many of the current problems of women scientists were present before 1920. Although their interests and backgrounds varied widely, they clustered in certain fields and certain institutions, and, despite an average level of education higher than that of the male scientists, they had fewer job opportunities and lower status, were more often unemployed, and less often considered eminent by their fellow scientists. A few women managed to overcome this pattern of discrimination and do important work. Others formed organizations to help correct the unequal conditions and to enhance the status of their sex in science. The story of these 500 women and their efforts to pursue careers provides an interesting glimpse into a neglected aspect of the history of science.

Before we can understand the place of women scientists in the scientific professions at the turn of the century, we must have some picture of the number, distribution, and career patterns of the men. Although governments did not keep or publish accurate statistics on men and women of science until long after 1900, we can perhaps get a rough picture of the career patterns of scientific men from a sample of 502 chosen from the third edition of American Men of Science, published in 1921.

As for professional biographical data on these men, the largest group are in chemistry, with 17.9% of the total. The second largest group are in the medical sciences of anatomy, bacteriology, physiology, and pathology. Together these two groups comprise about one-third of the total. Ranking next are four fields of about equal size — botany, engineering, physics, and zoology (including entomology) — which together constitute another 40% of the male scientists. All other scientists are divided among seven smaller fields. The reasons for this particular distribution of scientists must be exceedingly complex, the result of many economic, social, intellectual, and personal factors. The most that can be said is that such a distribution somehow reflected the interests and needs of industrial America at the turn of the century.

The educational patterns within each field were likewise quite varied. As late as 1920 men could still enter the scientific profession without a higher degree, but by that year a high percentage (46.6%) held the Ph.D. Psychology, one of the smaller sciences, had the highest percentage of doctorates (92.3%), which perhaps reflects the fact that it, more than any other science, grew up with the American university. The Ph.D. was the least common in pharmacy, engineering, agricultural sciences, and medical sciences, where it had not yet replaced the M. D. But in most sciences the usual pattern was a mix of one-third to three-fourths younger men with doctorates and the rest older men without.

As for some indication of the types and variety of jobs held by men in each field, our understanding can only be approximate since job mobility was high and many people held more than one job at a time. However we can gain some idea of the relative importance of academic and nonacademic employment in the various sciences. Once again, the pattern was different in each field. By 1920, most of the scientists (63.1%) were employed in a university or in an affiliated institution such as a museum, observatory, or research institute. The most “academic” of the sciences — those with the smallest percentage employed in government or industry — were mathematics and psychology; the most “applied” were chemistry, entomology, and engineering, which had extensive industrial as well as governmental applications. The other sciences had applications which were primarily in government.

Doctorates were highly correlated with academic positions. If one lists the sciences in the order of their decreasing dependence on academia for jobs, the order is roughly the same as that of the percentage of doctorates in each field. The exceptions are pharmacy and medical sciences, where only the elite practitioners were included in the AMS directories, and chemistry, where a doctorate was often necessary even for work in industry. Although it is not possible to include all the men’s occupations,, a quick scanning of the most popular ones shows that male scientists permeated the major areas of the American economy from mining to agriculture to industry and medicine. There seemed to be few limits on where an American scientist — male, that is — might appear.

To summarize the biographical data on the 504 women scientists in the first three editions of American Men of Science, the most popular sciences were zoology, botany, and psychology, which together attracted over half the total group. The rest of the women were spread over eight smaller fields. Anthropology was, as with the men, one of the smallest sciences, since its growth did not occur until after 1920. Quite striking was the total absence of women from engineering (which was popular with men), pharmacy, and the agricultural sciences of animal nutrition and agronomy.

The reasons for this particular distribution of women scientists are unclear. Most probably lacked rigorous mathematical training, as Alice Hamilton, an eminent industrial toxicologist admitted in her autobiography, although the minority in the physical sciences excelled in it. Perhaps the early women’s colleges, which trained and employed many of these women, may have stressed some sciences over others, as Mount Holyoke did zoology. Or perhaps women were counseled in some other way to prefer zoology, botany, and psychology.

It is equally difficult to determine why women did not enter some of the other sciences. Since there was no woman engineer in this period, it is hard to tell how one might have fared. Would she have been permitted to enroll in an engineering school, if she had been attracted to such a “masculine” field in the first place? Or would she have been repulsed by the hurly-burly of a field that offered relatively few academic positions (and none in the women’s colleges) and which might require rigorous outdoor work? Such self-selection was very likely the case, since in general women were found less often in those fields with the most industrial and governmental applications: engineering, entomology, and chemistry.

If rejection or a fear of rejection kept women out of some fields, another force was attracting them into the new “female” science of home economics around 1900-1910. Home economics was the result of a sexual division of scientific labor — the men took animal nutrition and the women human nutrition. The 23 women who identified themselves as being in the science of home economics before 1920 were a very well-educated group. Seventeen held Ph.D.’s, usually in chemistry and especially in the area known as “physiological chemistry,” from which the modern sciences of biochemistry, pharmacology, nutrition, and microbiology have all emerged. Several women studied food chemistry with Wilbur O. Atwater at Wesleyan University, but the major graduate programs in this field at the time were those of Russell H. Chittenden and Lafayette B. Mendel at Yale University, where 7 of the 17 earned their degrees, and the University of Chicago, with Julius Stieglitz and Marion Talbot, where 4 got Ph.D’s.

An examination of the subsequent careers of the Yale doctorates in physiological chemistry reveals that most of the women entered home economics and most of the men biochemistry, pharmacology, and animal nutrition. What was happening was that the newly created jobs in home economics at the large state universities attracted women, who were not as readily employed in biochemistry or animal nutrition as were men. For example, Isabel Bevier, one of Atwater’s students, was teaching chemistry at Lake Erie College in 1899 when she was called to the University of Illinois to head one of the first home economics departments. Likewise, Mary Swartz Rose, Louise Stanley, Amy Daniels, and Ruth wheeler were all chemists who were able to move into the new positions in home economics which were created after 1900. This whole home-economics movement owed much to the efforts of Ellen H. Richards, lecturer on sanitary chemistry at M.I.T., who had been popularizing the subject since the 1880s. By 1910 home economics had been accepted as a woman’s science, although the founders of the parent disciplines of food and physiological chemistry, from Justus Liebig on, had all been men.

On the whole, the women scientists before 1920 were a more highly educated group than the men, with 63.3% holding Ph.D.’s compared to 46.6% for the men. (This is in contrast to recent figures, based on the 1960 census, which show men more likely to have doctorates than women. The women’s advantage was not true in all fields, however, and seems due primarily to the low percentage of male doctorates among engineers and medical scientists. Women doctorates predominated in zoology, botany, medicine, mathematics, chemistry, and geology, and men in physics, psychology, astronomy, and anthropology. There seems to be no pattern in this. Nor for women is the percentage of doctorates in a field as closely related to the percentage of jobs in that field in academia as was true for the men. Both percentages were relatively uniform, probably reflecting the limited range of opportunities open to women in all fields. 

Well over half the women, like the men, found employment in academic institutions (68.3% for the women vs. 63.1% for the men). The women in academia were more often than the men in positions of inferior status. Almost twice as many (9.7% vs. 5.4%) worked in museums and research institutions rather than in colleges and universities. When they were in such schools, it was more likely to be in the women’s colleges, and nursing schools rather than in the universities and medical schools, where almost all the academic men in the sample were employed. This finding is corroborated by a report of the American Association of University Professors, which studied the status of women in academic institutions in 1921. Women at that time held 0.001% of the professorships at men’s colleges and universities (2 of “nearly 2000”), 68% of those at the women’s colleges (415 of 613), and 4% of those at the coeducational colleges and universities (190 of 4,760), where they were heavily concentrated in home economics and physical education. This report also found that the women hovered in the ranks of assistant, instructor, and assistant professor at the co-ed institutions far longer than did the men.

However difficult women found careers in academia, outside it they fared even worse. Here their employment picture diverges sharply from that of the men, for nonacademic opportunities were quite restricted. Unlike the men, they found no jobs in industry before World War I. The war seems to have opened opportunities to a few women chemists, fro two went to work for the National Aniline and Chemical Company during the war and four others took up similar employment between 1918 and 1920. Most notable of this group was Mary Engle Pennington, who became interested in the problem of food spoilage as chief o the Food Research Laboratory of the U.S. Department of Agriculture, and then moved to New York City in 1919, where, as a consulting chemist, she designed commercial and household refrigerators.

The chief opportunities open to women outside academia were in private practice for the physicians, in schools and clinics for the psychologists, and in government bureaus for the botanists. But not all government agencies were alike. One agency, the Bureau of Plant Industry in the Department of Agriculture, hired the strikingly large number of 24 women between 1900 and 1920. This bureau, like many others, underwent tremendous growth in these twenty years, but for some unknown reason it along attracted a great many women. Since government salaries were much higher than those in the women’s colleges, women would probably have entered other government agencies, such as the National Bureau of Standards, the Forest Service, the Geological Survey, and other bureaus in the Department of Agriculture, had the opportunity been open to them. Each of these agencies had numerous male scientists, but at most one or two and usually no women scientists. This pattern makes the Bureau of Plant Industry all the more unusual.

The women were also much more likely than the men to be unemployed or to be employed in jobs unrelated to science. This problem had been minor for the men, since of the 502 in the sample only 10 (2.0%) might be considered to be in this category: 4 were merchants or booksellers, 2 clergymen or missionaries, and 4 of unknown occupation. They were spread across 6 sciences, but 5 were in zoology, where a real job shortage may have existed. The women zoologists also found unemployment especially high: 21 of 92 (or 22.8%) were either unemployed or employed outside science as, for example, a missionary, librarian, or archivist. Several other zoologists were employed in informal positions as their husbands’ assistants.

Marriage, especially high among the zoologists, seems to have been the chief cause of the widespread unemployment in zoology, since 13 of the 21 unemployed were married (61.9%) — more than double the percentage for the field as aw hole (28.3%). The career patterns of these women also show sudden breaks and extended gaps, which, since many were married to other zoologists listed in American Men of Science — for example Gertrude Crotty Davenport, Florence Merriam Bailey, Mildred Hoge Richards, and numerous others — can be explained in terms of their husbands’ careers. Male zoologists are well known for marrying a disproportionately large number of distinguished women, many of them fellow students.

There is also evidence that marriage to a fellow zoologist had a significant effect on a man’s career. Of the 26 women zoologists, 15 were married to men in American Men of Science. Of these 15 men, 14 were in zoology or medical sciences, and 8 of them were “starred,” or considered eminent by their colleagues, a far higher percentage (57.1%) than was true of zoologists in general (roughly 33.3% in the first edition and 11.1% in the third). In only 3 of these 14 cases did the woman list herself as employed as her husband’s assistant, but the rest must also have contributed something to their husbands’ careers. Susanna Phelps Gage, an embryologist married to a starred histologist/embryologist, was herself starred, although official unemployed. It is hard to tell what such “unemployment” really meant, if the woman had access to the field and was able to use her talents through her husband’s career It is also hard to tell whether the married woman’s unemployment was voluntary or involuntary — the result of social prejudice against the employment of married women, the difficulties of relocation, or the desire to be at home with her children. We can only conclude that these factors, plus a possible job shortage in zoology, resulted in an exceptionally high rate of disguised unemployment for women in zoology before 1920.

In other fields this phenomenon was not so widespread. The rates of both marriage and unemployment were lower. Botany, psychology, and medical sciences — the other fields which constrained 10 or more married women — were also the fields which, as mentioned above, offered women the widest range of professional employment. In other fields, the percentage of married women was strikingly low, actually reaching zero in physics and approaching it in astronomy, where one and perhaps both of the married women were widows who had to work for a living. The anthropologists, on the other hand, were frequently wealthy widows who were interested in traveling in different cultures and studying the native peoples. Allowing for the fact that American Men of Science is not wholly reliable on the marital status of its women, marriage does seem to have ben a significant factor in the employment patterns of women scientists, but in different ways in different fields.

There were thus in 1920 a multitude of patterns and subpatterns in the distribution of women scientists in America. How had such patterns arisen? Were these women really a part of their profession? And were any of the 500 eminent? I can only touch on the answers to these complex questions briefly in the space allotted here.

Jane Colden, the botanist daughter of a famous botanist father, was probably the first woman scientist in America. She had few followers until the second half of the nineteenth century, when important changes in women’s education took place. Opportunities for women to study science, previously available only in a few colleges in the Midwest, increased greatly with the founding of the women’s colleges and the expansion of the land-grand universities after the Civil War. Women’s colleges were important in training women scientists before 1920. Of the 459 women who attended one undergraduate college and the 24 who attended two, making a total of 483, 184 (or 38.1%) attended just 8 Eastern women’s colleges and 91 (or 18.8%), 7 top state universities. The remaining 208 (42.8%) attended 98 other institutions, the most important of which were the universities of Chicago and Pennsylvania, Stanford University, and Oberlin College.

Women scientists were also dependent on the women’s colleges for the large number of jobs their faculties provided. Although each college could provide only a few positions in each science, together they could employ 20-30 women, which was roughly the number of women in some of the smaller sciences. In physics the women’s dependence on these colleges reached an extreme. In the years 1900-1920, 21 of the 23 women physicists were at some time employed in women’s colleges (91.2%). In the larger sciences the women’s colleges played a proportionately smaller role, but even there these professorships were quite prestigious and usually held by the top women in the field, such as Margaret Washburn in psychology at Vassar or Margaret Ferguson in botany at Wellesley.

The number of jobs at the women’s colleges more than double between 1906 and 1920 (from 43 to 96), as the older colleges expanded and some newer ones — Connecticut College for Women, Sweet Briar, and Skidmore — were established. The overall importance of the women’s colleges, however, began to decline after 19096, as more and more opportunities for women opened up in state universities and other schools. The period’s overall educational expansion resulted in a steady decline in the percentage of women employed in the women’s colleges, from 57.3% in 1906 to 36.5% in 1921. The state universities showed the largest overall increase, umping from 2 jobs in 1906 (2.7%) to 53 in 1921 (20.2%). Although the women’s colleges were by 1920 still the largest employers of women scientists, their overall dominance had diminished to slightly over one-third of the field, a percentage roughly equivalent to that of their (undergraduate) alumnae among women scientists as a whole. In the period before 1920, the women’s colleges thus accounted for roughly one-third of the “supply” and one-third of the “demand” for women scientists.

Although in their heyday of 1906-1920, the faculties of the women’s colleges included a number of highly educated and well-trained women, this had not always been the case. When Matthew Vassar started the movement for the higher education of women by founding Vassar College in Poughkeepsie New York, in 1862, he sought a woman scientist for his faculty. In 1865 he hired Maria Mitchell, a Quaker astronomer from Nantucket Island, who had earned fame in 1847 as the first to sight a new comet. She had had no college or seminary training and, in the tradition of Jane Colden, had learned her science by helping her father run his small observatory on the island. During the 1850s she earned her living by working part-time as a piece-work “computer” for the Nautical Almanac Office of the U.S. Coast and Geodetic Survey. At Vassar, Miss Mitchell was a forceful teacher, and several of her students became scientists. Likewise at the other women’s colleges established before 1885 (Mount Holyoke, Smith, and Wellesley), the first faculties contained women with only modest levels of education, and many were self-taught.

The big change began in 1885 when M. Carey Thomas of Bryn Mawr College, an advocate of the ideals of Daniel Coit Gilman of the Johns Hopkins University (although he would not allow her to attend his classes) ambitiously set out to hire women with advanced degrees and scholarly publications for her new college and graduate school. By the mid-1880s a few such women were available, and over the years Miss Thomas managed to hire some of the most outstanding women scientists of her time: for example, Charlotte Scott in mathematics, Florence Bascom in geology, and Nettie Stevens in zoology. Following Miss Thomas’s example, the women’s colleges, always anxious to prove themselves equal to the men’s colleges, rapidly transformed their faculties into the multidegreed ones they became by 1910. The pressure to have advanced degrees became so great that even women who had become professors in an earlier day when degrees did not matter, such as Cornelia Clapp of Mount Holyoke, took leaves of absence and went of to earn PhD.’s with their own former students.

The process of coeducating the graduate schools in the 1880s and 1890s was a difficult and painful one. Most graduate schools had no policy regarding women, and when the first hardy souls presented themselves, the schools did not know what to do. Christine Ladd (later Ladd-Frankly) was one of the early graduate students at The Johns Hopkins University, but when she presented her thesis to the mathematics department in 1882, she discovered that her degree would be withheld. It was finally awarded, over 40 years later, in 1926. Florence Bascom, daughter of the president of the University of Wisconsin, was more fortunate; she obtained her Ph.D. in geology at Johns Hopkins in 1893 by special dispensation, although women were not officially admitted until 1907.

At Harvard, Mary Whiton Calkins studied psychology and wrote a thesis with William James and Hugo Münsterberg, but discovered in 1896 that the Harvard Corporation would not grant her a degree. Luckily this injustice proved no lasting obstacle to Miss Calkins, who became a professor at Wellesley and one of the early presidents of the American Psychological Association. In 1902 Harvard moved to clarify the situation by creating the separate Radcliffe Graduate School, which would grant higher degrees to women at Harvard, but the school granted very few degrees in the sciences before 1920.  

Another psychologist, Margaret Washburn, was an outstanding graduate student of James McKeen Cattell at Columbia in the 1890s, but since there were no fellowships there fro women, she had to transfer to Cornell, where she earned her degree in 1894. Likewise Ida Hyde, a Bryn Mawr graduate, had great difficulty being admitted to the German universities in 1893, but he persevered and became one of the first women to earn a German Ph.D. In 1896. By the late 1890s such incidents became less common as most of the major graduate schools liberalized their policies on women and began granting them degrees in large numbers.

One of the main reasons for a change in the attitude toward women at graduate schools, and the dramatic increase in higher degrees awarded them in the mid-1890s, was the example of the University of Chicago, which was founded in 1892 and from the start accepted men and women on an equal status in both its graduate and undergraduate programs. It rapidly outstripped the eastern universities in the number of degrees it granted women. In botany, in particular, it established an outstanding tradition, granting as many degrees in this one department before 1920 as did the next seven universities combined. Chicago even usurped some of what might have been considered Harvard’s domain. In its first year Chicago attracted students from 95 institutions, with the largest group —14 students — coming from Wellesley College.

By 1900 women’s education had come a long way from the days of Matthew Vassar’s pioneering venture. It had largely fulfilled the hopes of those who sought an education for women equal to that of men. A woman who wished to study science in 1900 could find both undergraduate and graduate training of a high order. Her chances for eminence and recognition in her hosen field, however, remained much more limited than was true for men.

One of James McKeen Cattell’s purposes in compiling his massive directories of American scientists had been to identify a sample of the top 1,000 scientists for further study. For this purpose he had scientists in each field rank each other, and he starred the top 1,000 in his volumes. It was easier to be included in this esteemed group in 1906 than in 1921, but in both years the women’s showing was disappointingly poor. Thirty of the 504 women scientists were starred in one of the first three editions — less than half the 73 starred in the male sample. Because their numbers remained roughly constant (19 in 1906, 18 in 1910, and 24 in 1921) at a time when the number of women in the directories almost tripled, the percentage of women who were starred dropped from 12.8% in 1906 to 5.2% in 1921.

Why this drop should have occurred when the group as a whole was undoubtedly highly talented and motivated, and when their numbers and opportunities were increasing so greatly, is not at all clear. Cattell could not understand the phenomenon and thought there must be some “innate sexual disqualification,” although he added, “It is possible that the lack of encouragement and sympathy is greater than appears on the surface.” Today we would attribute this sexual difference in achievement to discrimination in various forms. The women were employed at women’s colleges and in low echelons elsewhere, where they were less visible, had less access to research facilities, fewer stimulating colleagues, and heavier teaching loads. Perhaps worst of all, there was much less incentive for them to persevere to overcome their obstacles, for thee was no place else for them to go.

The generally low level of accomplishment among women scientists and their virtual segregation raises the question of whether they really were a part of the profession. Numerically, they were a tiny minority — less than 10% — in most fields except psychology, where they constituted 22.8% of the field, and, of course, home economics.

A few prominent women seem to have been readily accepted in their fields, as shown by their election to the major professional societies. Maria Mitchell again led the way when she was elected to the American Academy of Arts and Sciences in 1848 (though opposed by Asa Gray) and to the American Association for the Advancement of Science in 1850 (supported by Louis Agassiz). The California Academy of Sciences went so far as to invite women to join in 1853, the year it was established. Rachel Bodley, Dean of the Women’s Medical College in Philadelphia, was elected to the American Chemical Society in 1876, the year it was founded, and Charlotte Scott helped establish the American Mathematical Society in 1894. Similarly, Marcia Keith of Mount Holyoke and Isabelle Stone of Vassar were among the founders of the American Physical Society, and Mary E. Pennington joined the American Society of Biological Chemists at its second annual meeting in 1907.

But other women encountered opposition. In physiology there was a fifteen-year lag between the formation of the American Physiological Society in 1887 and the election of its first woman member, Ida Hyde of the University of Kansas, in 1902. Women anthropologists were deliberately excluded at least once by their brethren, but their problems may have stemmed as much form the insecure pre-professional status of their field as from discrimination. In 1885 ten women led by Matilda Coxe Stevenson of the Bureau of American Ethnology met in Washington to form the Women’s Anthropological Society of America. Angry at being excluded from the men’s Anthropological Society of Washington because of their sex and lack of training, they banded together to present papers to each other as a means of mutual encouragement. They also devised a feminist rationale for their inclusion in the field, arguing that women could contribute to anthropology in ways that men could not, since tribeswomen would theoretically be more willing to discuss certain aspects of their lives with them rather than with men. Such a distinction turned out not to be valid in practice, since, as Nancy Lurie has pointed out, the idea was more in the Victorian anthropologists’ mind than in the Indians’. However, the women’s group had about 40 active members in 1889 and maintained a schedule of biweekly meetings until 1899, when it quietly merged with the men’s group. By then feelings on both sides had softened, and the men were finally willing to accept their female colleagues into the field. In 1903 the combined group elected Alice Fletcher of the Peabody Museum at Harvard, and outstanding ethnologist and advocate of Indian rights, to be its first woman president.

But if the professional societies were generally willing to admit women, the nation’s most prestigious group, the National Academy of Sciences, was not. It was not until 1925 that the first woman, Florence Sabin, was selected and 1931 when the second, Margaret Washburn, was added. These women had done outstanding work, but it can only be considered an outrage that others such as Annie Jump Cannon and Charlotte Scott were never elected.

But being elected to membership in the professional societies is really no test of how willingly and actively women were accepted into the profession. As Martha White has pointed out, there is much more to being a member of the profession and feeling professionally involved than merely being elected and paying the dues. She considered important such factors as equal expectations, equal access to prestigious journals, equal support from eminent sponsors, equally stimulating colleagues, and meaningful communication. Without examining the correspondence of a number of these women, it is difficult to imagine what they felt about these issues. Some, such as the highly regarded Margaret Washburn, resigned themselves to their positions in the women’s colleges, although in private they admitted they would have much preferred coeducational institutions. Others, such as Maria Mitchell, Ida Hyde, and ichthyologist Rosa Smith Eigenmann, spoke out whenever possible against their unequal conditions and the male condescension that repeatedly considered them “good for a woman.” 

Some of these women scientists even went further and tried to correct some of the unequal conditions facing them. Marion Talbot and Ellen Richards founded the Association of Collegiate Alumnae in Boston in 1881 (later the American Association of University Women). From it early days the association was concerned with opportunities for graduate work for women. Even though many graduate schools began to admit women in the 1890s, the almost total lack of fellowships for women continued to be a real problem. Several eminent women scientists — among them Christine Ladd-Franklin, Ida Hyde, Margaret Maltby, and Ellen Richards — were active in the association, and they worked hard to establish fellowship funds for women.

By 1920 at least 2 women scientists had benefited greatly from these awards. For most of them a career in science would not have been possible without the AAUW fellowship. For a number of years, the AAUW women scientists also raised money to support a table for a woman at the Naples Zoological Station, to grand a prize for the best thesis by a woman and, in 1921, to purchase one gram of radium for Madame Curie to help her continue her researches. All these projects were part of their program to gain greater opportunity and recognition for women scientists. Although these women were divided among the several sciences, they faced certain common problems, which they hoped the AAUW would help alleviate for future women scientists.

Despite all these obstacles, a few women scientists did important scientific work before 1920. The number of “starred” women varied among the sciences from one each in physics and anthropology to 6 in psychology and 8 in zoology, suggesting perhaps that where the women congregated they achieved the most, although the samples are too small to be reliable. Among the zoologists Mary Jane Rathbun classified crustacea at the U.S. National Museum for over thirty years, and Cornelia Clapp, who worked in marine embryology, was a mainstay of the Marine Biological Laboratory at Woods Hole for a number of summers. A younger group, some of them students of Thomas Hunt Morgan at Bryn Mawr from 1891 to 1904, worked in the area of genetics. Charles Rosenberg has shown that, of the 52 American first authors cited in the bibliography of Morgan’s The Mechanism of Mendelian Heredity (1915), seven (or 13.5%) were women (Alice M. Boring, E. Eleanor Carothers, Katherine Foot, Mildred Hoge Richards, Helen Dean King, Margaret Morris Hopkins, and Nettie M. Stevens), and two others (Ella C. Strobell and Rhoda Erdmann) were second authors. The most eminent of this group was Nettie Stevens, who discovered independently of E. B. Wilson in 1905 the relation of the X and Y chromosomes to sex determination.

In psychology Margaret Washburn of Vassar college was well known for her experimental work in space perception and animal psychology. Mary Calkins studied dreams and association and developed her own introspectionist “self-psychology.” Christine Ladd-Franklin combined her interest in mathematics and psychology in studies of color sensation and optical illusions. Helen Thompson Woolley of Cincinnati pioneered in the study of child development and was a leader in the vocational guidance movement. Leta Stetter Hollingworth, in 1914 the first Civil Service psychologist in New York City, later moved to Teachers College, Columbia University, where her mental and motor tests of women helped undermine the view of her colleague Edward L. Thorndike that men were more capable than women.

In anatomy Florence Sabin, a student of Franklin P. Mall and the first female full professor at Johns Hopkins University, made intricate studies of the brain and the lymphatic and vascular systems. In mathematics Charlotte Scott and Anna Johnson Pell, both of Bryn Mawr College, worked on algebraic curves and integral equations. In physics, Margaret Maltby of Barnard College studied the physical chemistry of electrolytic resistances and dilute solutions and did some work on radioactivity.

Florence Bascom of Bryn Mawr worked on the petrography of Massachusetts and eastern Pennsylvania, and Ermine Cowles Case of the University of Michigan studied the osteology of extinct reptiles, meeting her death in Sough Africa in 1923 while collection specimens. Ellen Semple was the author of several books and one of the founders of the science of geography, although her only professional employment was as an occasional lecturer at the University of Chicago. In botany Alice Eastwood classified thousands of plants as curator at the California Academy of Sciences for almost sixty years (1892-1949). Elizabeth Knight Britton assisted her husband at the N.Y. Botanical Gardens in the classification of mosses, and Margaret C. Ferguson of Wellesley College studied the fertilization and germination of spores of fungi.

Probably the most famous of all these women were the astronomers Annie Jump Cannon and Williamina P. Fleming, both “Curators of the Photographic Plates” at the Harvard College Observatory, where they did outstanding work on the classification of stellar spectra. They also directed teams of up to fifteen other women in this painstaking and poorly paid technical work, which few men would undertake but which, in the end, did so much to make the Observatory and its Draper Star Catalogue known throughout the world. Surely when women were given the chance, they could perform important work.

Recent work in the sociology of science has shown that there are certain barriers (“stratification”) within the institution of science which distort the even distribution of honors and prevent some men from achieving the rewards their ability would seem to merit. For women scientists sexual discrimination was an additional and particularly oppressive form of segregational stratification which suppressed their achievement by at least half and which certainly violated what we all like to think of as the norms of the scientific profession. Unfortunately, we know too little of what drove these women to enter science and to try to achieve in it despite the obstacles. We are only beginning to uncover their side of the history of science.

 

 

Copyright 2019 by Margaret W. Rossiter. Used by permission. All rights reserved.

 

Also: Rossiter, Margaret W., “Women Scientists in America Before 1920: Career patterns of over five hundred women scientists of the period reveal that, while discrimination was widespread, many women were working hard to overcome it,” American Scientist, Vol. 62, pp. 312-323.

 

Also: Margaret W. Rossiter, “The Matthew Matilda Effect in Science,” Social Studies of Science 23 (May 1, 1993): pp. 336-337.