SHiPS Resource Center
for Sociology, History and Philosophy in Science Teaching

Women in Science: 5000 Years of Obstacles and Achievements

by Patsy Ann Giese, Slippery Rock University of Pennsylvania

When elementary school children were asked to draw a picture of a scientist in a recent study, 820 girls and 699 boys drew male scientists. Only 129 girls and just 6 boys drew female scientists (Fort & Varney, 1989). It is not just children who think of science as a male endeavor. In my 1972 edition of Asimov's Biographical Encyclopedia of Science and Technology, I found listings for five women out of 1195 biographies. (I had to go through the entire book page by page to count the women because there are no index entries for female, woman, or any other synonym.) Checking the 1982 edition at the library, I found 308 more men and 7 more women compared to the edition published a decade earlier.

This is not a trivial issue. In stories of individual lives and in conclusions of education research studies, role models have been shown to be of immense importance in girls' and women's decisions to learn science.

Evidence of women scientists comes from as long ago as 4000 BC., when a carving of an unnamed Sumerian priestess-physician was made. Written records exist about Egyptian female physicians such as Merit Ptah from 2700 BC. and Zipporah from 1500 BC. Ancient Egyptian women could attend medical school with males or attend an exclusively female school at Sais. Tapputi-Belatikallim worked with chemicals used for perfume production in Mesopotamia around 1200 BC.

At 600 BC. due to the flowering of Greek science, the number of women recorded per century in historical documents increased about fifty fold to twenty per century. That ratio stayed relatively constant for the next twelve centuries. Women were treated equal to men in the Pythagorean Community, Plato's Academy, and the Epicurean School. Theano probably married Pythagorus when he was an old man; she was the leader of his school after he died. Agnodice in 300 BC. was a successful doctor dressed as a man. When she was to be tried for illegal practice, protests by the women she had healed resulted in changing the law so women could be physicians who treated only women. In Rome, female physicians were numerous and were allowed to treat men. In the First or Second Century, Maria of Alexandria, also called Mary the Jewess, was an alchemist who invented the water bath, the three-armed still, and other chemical equipment. The last great scientist of antiquity was Hypatia, who was born in 370 AD. Following the pattern set by her father, she lectured at the University of Alexandria in mathematics and astronomy. She invented the plane astrolabe for measuring positions of stars and planets, an apparatus for distilling water, and the hydrometer for determining the density of liquids. She was murdered in 415 AD by a Christian mob trying to stamp out Platonism.

In the Seventh through the Tenth Centuries, as in the beginnings of recorded history, there were few women scientists in the Western World, and most were physicians or alchemists. In the Byzantine Empire, the dark ages were not as bleak for women scientists because royal women studied medicine and natural sciences with scholars in their courts. At this time, there were also women engineers in China.

From 1000 to 1400, opportunities for women scientists in Europe reached a new peak, one that was not surpassed until the present century. The rapid growth of monastic life can be credited with most of this increase in intellectual freedom for women. The majority of scientists, male and female, belonged to religious orders. Hildegard of Bingen, a Benedictine Abbess born in 1098, was the most influential of women scientists of this era. She is also the earliest women scientist whose major works are still intact; she wrote about cosmology, medicine, botany, zoology, and geology. (Her versatility is shown by the fact her music for the mass is the oldest such music known to be composed by a woman.) In spite of being periodically ill, perhaps with migraines or epilepsy, she lived to the age of 81.

The medical school of Salerno, Italy, was the first one not affiliated with the Church. Tortula was a professor there, and she wrote several treatises widely used for five hundred years. "At times Trotula's advice seems uncannily modern, emphasizing the importance of cleanliness, a balanced diet and exercise, and warning against the effects of anxiety and stress" (Alic, 1986, p. 51). Little is known about Tortula's life except that she had a husband and two sons who were also physicians and faculty at Salerno. Most likely, she died in 1097. The number and importance of European universities increased greatly beginning inthe Twelfth Century, but everywhere outside of Italy they were closed to women.

Contrary to popular impressions of the Renaissance as a period of great resurgence in intellectual activity in Europe, it was a time of decreased participation of women in science (particularly in the 1500s). There were a number of factors causing this decline. Numerous abbeys were closed following the Protestant Reformation, and often their property was given to universities. For other abbeys, their control was transferred from an abbess to an abbot. As universities grew, female physicians who had been trained by other women lost the right to their profession even if they had passed an examination. Not only were women with scientific knowledge called charlatans, with more serious consequences they were called witches. Estimates for the number of people (nearly all of whom were female) executed for witchcraft between 1400 and 1700 have ranged from 100,000 to 9,000,000.

The Seventeenth Century saw the beginning of a rise in the numbers of women scientists. Botany, including the use of the newly invented microscope, was considered a particularly appropriate subject for them to study. Entomology was also a popular subject explored through the microscope. Similarly, the telescope fueled interest in astronomy. Elisabetha Hevelius aided her husband in running his observatory, and after his death she completed a catalog of 1564 stars that was published in 1690. Manufacturers of these instruments promoted women's interest through lectures and books. The first English periodical for women interested in science was published from 1690-1697. Most women studied at home, sometimes developing impressive natural history collections.

A few unusual women traveled far like Maria Merian and her daughter Dorothea, who went in 1698 from Holland to Surinam to collect and paint specimens of insects and plants. Both Maria's father (who died when she was an infant) and her step-father had been botanical illustrators. She married her teacher, but she left him in 1685 after 17 years of marriage.

These examples have been given to show that women did contribute to the development of scientific knowledge even during the beginnings of that development. It is important to note that opportunities for women in science have not steadily expanded throughout the ages. We can not take continual progress for granted. We must be motivated by belief in fairness towards women (and all other groups of people) in conjunction withexcitement over the potential to modify scientific methods and institutions as they become more inclusive. As we are willing to deal with these issues in our classrooms, we will increase the chances of our students' participation in the establishment of a more equitable society.

The role of women in science during the Eighteenth and Nineteenth Centuries will be described in Part Two of this series. The Twentieth Century will be covered in Part Three with some concluding remarks about why the value of women's scientific work has been underestimated by most people.

The Eighteenth Century Age of Enlightenment, like the centuries that preceded it, was a time when any woman's role in scientific discoveries was limited by the extent she had education and encouragement at home. For example, Antoine Lavoisier is called the "Father of Modern Chemistry" for his ideas like the conservation of matter, but his wife Marie's contributions are difficult to assess. She was married to him at age thirteen when he was already a chemist elected to the French Academy of Sciences. Marie assisted with experiments, kept laboratory records, translated scientific works from English to French, carried on scientific correspondence, and illustrated her husband's numerous publications. After he was guillotined in 1794 by French revolutionaries, she edited his works. Then Marie married another scientist Benjamin Thompson, also known as Count Rumford.

Unlike Marie Lavoisier, Jane Colden Farquahar ended rather than began her scientific work upon getting married. Colden, born in 1724, is the first well-known American female scientist. She was taught by her father, and like him she corresponded with the great botanists of the day including Carolus Linnaeus in Sweden She described and sketched 300 local plants, being the first to identify the gardenia.

Single women sometimes relied on other family members besides fathers for their education. Caroline Herschel learned astronomy as an assistant to her brother William. (At first, her mother strongly objected to Caroline moving from Germany to England where William lived. Her mother's consent was won because William agreed to send money regularly to pay for a maid to do Caroline's domestic work.) Only when William was away from home did Caroline have much opportunity to use the telescopes that they and another brother had built together. In 1783, she discovered fourteen nebulae. Between 1786 and 1797, she discovered eight comets. She compiled a catalog of stars as requested by William, and after his death she did another catalog for his son.

The number of women scientists working with their husbands increased in the Nineteenth Century. Notable examples are Mary Lyell, a conchologist; Marie Pasteur, a biologist; Mary Buckland, a geologist; Elizabeth Britton, a botanist; Amalie Dietrich, an entomologist; Margaret Huggins, an astronomer; Eliza Sullivant, a botanist; and Hertha Ayrton, a physicist. Despite the accomplishments of these women, a commonly held belief was that developing a woman's intellectual capacity would always diminish her reproductive capacity.

A momentous--and controversial--change occurred in that century: higher education was opened to women. In the United States, it started with the charter of Georgia Female College in 1836. By contrast, the first institution for males (Harvard) was begun 200 years earlier. An era of commitment to excellence in women's education began when Vassar was founded in 1865, followed by Smith and Wellesley within the next decade.

Education for women was justified so they could be better wives and raise better sons. Even women like botanist Almira Phelps, who had an influential teaching and writing career in seven states between husbands and after the last husband's retirement, believed men and women were meant to have separate spheres. Phelps wrote, "She [woman] was created to be the companion of man, to cheer his solitude and to assist him in his duties...A companion or assistant fills a secondary position" (Slack, 1987, p. 91). In 1859, Phelps was elected the second woman member of the American Association for the Advancement of Science. By the end of the century, nearly four hundred thousand copies of her Familiar Lectures on Botany had been sold.

Women's colleges were not only places for women to learn, but also institutions that employed women scientists. The first edition of American Men of Science, published in 1906, listed 4,131 persons including 149 women. Over two-thirds of the 52 women employed at the faculty rank of assistant professor or higher were at women's colleges. That group included the most important American female scientist of the century--the astronomer Maria Mitchell, who was a professor at Vassar from 1865 through 1888 even though her formal education had ended at age sixteen. She discovered a new comet in 1847 while assisting her father in his observatory. Two years later, she began working at home doing calculations for the United States Coast Survey. In 1850, she was the first woman elected to the American Association for the Advancement of Science. She edited the astronomical column of the Scientific American. Perhaps her greatest contributions were mentoring and organizing other women.

However, this employment had a stringent condition: women had to remain single. Harriet Brooks, who had already published two major articles on radioactivity with Ernest Rutherford, told her employers at Barnard in 1906 that she was engaged, but wished to continue teaching in the physics department. The trustees decided that a married woman should "dignify her home-making into a profession, and not assume that she can carry on two full professions at a time" (Rossiter, 1982, p. 16). Brooks resigned, broke off her engagement, and never did any more physics research.

Similarly, education for women was expanding in Europe. Six women's colleges were opened at Oxford and Cambridge from 1869 through 1893. Italian universities, which had ceased admitting women around 1800, began letting them in again during the 1870s. France, Switzerland, Sweden, and Denmark all started accepting female university students during the second half of the century. However, in Russia, following the government's rejection of an 1867 petition to allow women into universities, women had to be content with an informal combination of public lectures and private discussions led by cooperative professors.

One of these Russian women was Sonya Kovalevsky. She married to sidestep getting her father's permission to continue her studies in Germany. She earned her Ph.D. in 1874 through private tutoring because women were not allowed at university lectures there either. After her husband's suicide, she joined the faculty of the newly opened University of Stockholm, becoming the first woman professor in Europe since the Italian universities were closed to females. When Kovalevsky won in 1888 a prize from the French Academy of Sciences for her paper about the rotation of solids, her former professor Karl Weierstrass wrote that "I have particularly experienced a real satisfaction; competent judges have now given their verdict that my faithful pupil, my 'weakness,' is not a frivolous marionette" (Ogilvie, 1986, p. 116). The judging process was done with each manuscript lacking the author's name, but identified by a motto. Kovalevsky's motto was "Say what you know, do what you must, come what may" (Osen, 1974, p. 132).

Beginning in the 1880's, it became an increasingly acceptable idea that women should be engaged in the application of science to help solve growing domestic, societal, and environmental problems. The establishment of the Women's Laboratory at the Massachusetts Institute of Technology by Ellen Swallow Richards was the beginning of this movement. Also, in the United States, the establishment of land-grant colleges and universities increased the need for faculty, particularly in agriculture and other vocational education fields.

During the last part of the Nineteenth Century, funding for research projects was increasing, which meant more staff could be supported. Several new astronomical observatories were built allowing male assistants to move up to positions of leadership and creating a shortage of qualified help at the less desirable levels of employment. Edward Pickering, director of the Harvard College Observatory, became so disgusted with a male assistant's work that he declared his maid could do better. His twenty-four-year-old maid, Williamina Fleming, was supporting herself and her child after immigrating to the United States and getting divorced. Fleming worked for Pickering for thirty years, supervising the cataloging of nearly 200,000 photographic plates, discovering 222 variable stars and 10 novae, editing publications, and developing a classification scheme for stars based on their spectral characteristics. Between 1885 and 1900, she hired twenty female assistants including Antonia Maury, Henrietta Leavitt and Annie Jump Cannon.

Educational institutions played a pivotal role during the 1700s and the 1800s in increasing opportunities for women in science (and in other fields as well). As educators, we can take pride in our heritage. More than that, we can encourage our female students to resist elimination of their options and to persist in achievement of their goals.

If people can recall the name of only one female scientist, most likely they will name Marie Curie. She was the first person to win two Nobel prizes: a 1903 prize in physics and a 1911 prize in chemistry. The former was awarded for her doctoral dissertation work at the Sorbonne in France. Her dissertation topic was radioactivity, and it was she who first used that terminology. The latter prize was for establishing the existence of the elements radium and polonium. She had come far from her start as a governess in Poland (then under Russian domination) from 1885 to 1890. Discouraged by the restraints of her situation, she wrote in an 1886 letter, "My plans for the future? I have none, or rather they are so commonplace and simple that they are not worth talking about. I mean to get through as well as I can, and when I can do no more, say farewell to this base world. The loss will be small, and regret for me will be short--as short as for so many others" (Pycior, 1987, p. 195). Even though she wrote that she had no plans, she continued to study alone until she had financial resources enough to return to school.

Besides her own determination and intelligence, the other great asset Marie Curie had was support from an extended family that included her parents, sisters, husband, and father-in-law. The last, Dr. Eugene Curie, assumed much of the responsibility for caring for Marie and Pierre's two daughters until his death in 1910, when they were ages five and twelve. Pierre abandoned his own research in piezoelectricity to work with Marie on isolating radioactive substances, for which he shared in the 1903 Nobel prize. After Pierre's death in 1906, Marie declined a widow's pension and instead got her husband's job; thus she become the first woman professor at the Sorbonne.

Even with all her accomplishments, male scientists of international stature did not always treat her with openness and respect. For example, she was never elected to the French Academy of Sciences, even though Pierre was elected in 1905. In 1911, she lost by one vote on the first ballot and by two votes on the second. No woman was elected until 1979. Writers in the public press questioned whether or not Marie Curie's scientific accomplishments should be credited first to Pierre and later to Paul Langevin (with whom she was accused of having sexual intimacy) in attempts to show that women were not capable of independent, creative thought.

Madame Curie had a powerful influence as a role model. Her daughter Irene Joliot-Curie discovered artificial radioactivity in 1934 with her husband Frederic, and they were rewarded with a Nobel prize in 1935. Their children, Helene and Pierre, continued the tradition of being distinguished scientists working with spouses.

After thirty years of rapidly expanding opportunities, rigidity set in concerning women's roles in science after 1910. Women were generally low-paid laboratory assistants assigned tasks of painstaking, tedious detail. Women were forbidden to enter mines because they might cause bad luck, and women were not allowed to use astronomical observatories because they might be unsafe away from home at night. Even if females did the work usually assigned to males, the females received substantially less pay. Women's colleges replaced female faculty with males. Fearing the loss of their prestige, most male researchers refused to collaborate with women. This practice obviously reduced women's chances to gain prestige. Women were often unable to obtain their own grant funds or the reduced teaching loads that would enable them independently to make significant contributions to research.

An exception to this pattern was Lise Meitner, another woman inspired by Marie Curie. By 1910, she had published important papers as sole author and in collaboration with Otto Hahn. She wrote with her nephew Otto Frisch in 1939 the first paper about nuclear fission after she became convinced that uranium atoms could be split. In 1966, she became the first woman to win a Fermi Award from the Atomic Energy Commission. Just as Curie before her, Meitner experienced more than the glory, such as when she worked for one Nobel laureate in Berlin who made her promise to stay out of any laboratory where males were conducting research. Because there are so many female scientists in the Twentieth Century, I will continue mentioning only those who studied radioactivity.

For decades, women aspiring to scientific careers were expected to use the "Madame Curie strategy" of deliberate over-qualification, modestly acquired. Especially during the depression, women were often advised not to overcome stereotypes, but to adapt to them, being grateful for any job remotely related to science.

In the United States, anti-nepotism rules became more widely applied in the 1920s, depriving numerous women of faculty positions at the universities that employed their husbands. Through World War II, these rules were not weakened even though there was a great demand for trained scientists. In 1970, three-quarters of the land-grant colleges and universities still had written policies restricting the employment of relatives.

Maria Goeppert-Mayer, born in 1906, came from a family of six continuous generations of German university professors. Goeppert-Mayer received her doctorate from the University of Gottingen at the age of twenty-four. The same year, she married a chemist, Joseph Mayer, and became subject to anti-nepotism rules. Her next step was becoming a volunteer research associate--a euphemism for an unemployed or underemployed person--in Baltimore. This was followed by part-time college teaching and full-time weapons development work in New York during World War II. Later in Chicago on a half-time research job, she formulated a shell model for atomic nuclei. A 1963 Nobel Prize was given to her for this 1948 work, making her the second woman with a Nobel Prize in physics and the only female Nobel laureate in science alive at that time . She was elected to the National Academy of Science in 1956, but she was not employed full-time as a professor, and paid accordingly, until 1960 when she moved to the University of California in San Diego.

The second half of the Twentieth Century has been a time of positive changes for women in many realms. Numerous written works and many organizations have helped to bring about these changes. Women have assumed more leadership roles, working against the stereotypes that had blocked them for so long.

Chein-Shiung Wu was elected the first woman president of the American Physical Society. In 1958, she became the seventh woman ever elected to the National Academy of Science. A physics professor for more than thirty years at Columbia University, Wu is famous for her experiment demonstrating the lack of parity conservation when cobalt decays. "She is reputed to be very smart, and very fierce--she has fought hard to control her turf and to defend her position" (Jones, 1990, p. 208). When Wu retired, the physics faculty at Columbia became entirely male.

Even at the lower ranks, equity has not been achieved yet. In the United States, bachelor's degrees in physics awarded to females were estimated by the National Science Foundation to be 21% of the total in 1930. They reached a high of 23% in 1945, coming up from 14% in 1940. From 1950 through 1970 when the National Science Foundation provided more accurate values, the proportion stood at 4% to 6% By 1980, that number had risen to 13% and was leveling off (Garmon, 1983, p. 115).

Women are still bound by meeting traditionally male criteria for professional success, especially in science, while being expected to fit traditionally female modes of personal behavior. The result is that many girls choose to avoid this conflict by avoiding scientific careers. It is not just the girls that need to change. Careful appraisal is needed of methods of teaching, hiring, and promoting females.

We have seen that for at least 5000 years, women have made contributions to science. Details of a few women's lives have been presented to illuminate the pattern of female participation in science. All the women selected made professional contributions, but they were not chosen on the basis of the significance of their work. Instead, they were chosen to broaden understanding of overt and covert discrimination as well as commitment to equity.

Although some women scientists have been remembered through the ages, countless more have been forgotten. It is difficult to trace their authorship due to women changing their names upon marriage. Copiers of manuscripts sometimes substituted masculine forms of names for feminine forms. Some women published under male pseudonyms to avoid persecution or loss of social position. Other women did likewise to increase the probability that their work would be taken seriously.

Until recent times, few women had access to formal education. Because their education was provided by fathers, brothers, husbands, and male colleagues, women's discoveries were easily appropriated by or attributed to those men in their lives. Some women had low self-esteem and underestimated the importance of their own original ideas, reflecting the values of society in general. Other women, for reasons of propriety, did not seek or want credit for their work.

For all these reasons, the value of women's scientific work has been underestimated by most people. From ignorance of women's contributions, people for centuries have concluded that women are by nature less capable than men in scientific endeavors. Then when women like Curie or Goeppert-Mayer demonstrated undeniable ability in science, it was labeled an exception to the rule. With this false logic, women could never "prove" their equal aptitude. The final conclusion in this line of reasoning is that women's role in science is not worth studying. Of course, we must avoid all segments of this circular thinking.

Without the support and rewards often given to men, thousands of women have observed nature, experimented in laboratories, and constructed abstract scientific ideas. This speaks powerfully to the inherent appeal of science. We should tell these women's stories to the next generation, both male and female.

REFERENCES


The SHiPS Teachers' Network helps teachers share experiences and resources for integrating history, philosophy and sociology of science in the the science classroom.