Women in science

"Woman teaching geometry"
Illustration at the beginning of a medieval translation of Euclid's Elements (c.1310 AD)

Women have made significant contributions to science from the earliest times. Historians with an interest in gender and science have illuminated the scientific endeavors and accomplishments of women, the barriers they have faced, and the strategies implemented to have their work peer-reviewed and accepted in major scientific journals and other publications. The historical, critical and sociological study of these issues has become an academic discipline in its own right.

The involvement of women in the field of medicine occurred in several early civilizations and the study of natural philosophy in ancient Greece was open to women. Women contributed to the proto-science of alchemy in the first or second centuries AD. During the Middle Ages, convents were an important place of education for women, and some of these communities provided opportunities for women to contribute to scholarly research. While the eleventh century saw the emergence of the first universities, women were, for the most part, excluded from university education.[1] The attitude to educating women in medical fields in Italy appears to have been more liberal than in other places. The first known woman to earn a university chair in a scientific field of studies, was eighteenth century Italian scientist, Laura Bassi.

Although gender roles were largely defined in the eighteenth century, women experienced great advances in science. During the nineteenth century, women were excluded from most formal scientific education, but they began to be admitted into learned societies during this period. In the later nineteenth century the rise of the women's college provided jobs for women scientists, and opportunities for education. Marie Curie, the first woman to receive a Nobel Prize in 1903 (physics), went on to become a double Nobel Prize recipient in 1911 (chemistry), both for her work on radiation. Forty women have been awarded the Nobel Prize between 1901 and 2010. 16 women have been awarded the Nobel Prize in physics, chemistry, physiology or medicine.[2]

History

Ancient history

The involvement of women in the field of medicine has been recorded in several early civilizations. An ancient Egyptian, Merit-Ptah (c.2700 BC), described in an inscription as "chief physician", is the earliest known female scientist named in the history of science. Agamede was cited by Homer as a healer in ancient Greece before the Trojan War (c. 1194–1184 BC). Agnodike was the first female physician to practice legally in fourth century BC Athens.

The study of natural philosophy in ancient Greece was open to women. Recorded examples include Aglaonike, who predicted eclipses; and Theano, mathematician and physician, who was a pupil (possibly also wife) of Pythagoras, and one of a school in Crotone founded by Pythagoras, which included many other women.[3]

During the period of the Babylonian civilization, around 1200 B.C., two perfumeresses named Tapputi-Belatekallim and -ninu (first half of her name lost) were able to obtain the essences from plants by using extraction and distillation procedures. If we are to argue chemistry as the use of chemical equipment and processes, then we can identify these two women as the first chemists. Even during the time of the Egyptian dynasty, women were involved in applied chemistry, such as the making of beer and the preparation of medicinal compounds.[4] A good number of women have been recorded to have made major contributions to alchemy.[4] Many of which lived in Alexandria around the 1st or 2nd centuries AD, where the gnostic tradition led to female contributions being valued. The most famous of the women alchemist, Mary the Jewess, is credited with inventing several chemical instruments, including the double boiler (bain-marie); the improvement or creation of distillation equipment of that time.[4][5] Such distillation equipment were called kerotakis (simple still) and the tribikos (a complex distillation device).[4]

Hypatia of Alexandria (c. 350–415 AD), daughter of Theon of Alexandria, was a well-known teacher at the Neoplatonic School in Alexandria teaching astronomy, philosophy, and mathematics.[6][7] She is recognized to be the first known woman mathematician in history through her major contributions to mathematics.[7] Hypatia is credited with writing three major treatises on geometry, algebra and astronomy; as well as the invention of a hydrometer, an astrolabe, and an instrument for distilling water.[3][8] There is even evidence that Hypatia gave public lectures and may have held some sort of public office in Alexandria.[9] However, her fruitful life was cut short in 415 AD by Christian Zealots, known as Parabalani; who stripped her, dismembered her, and the pieces of her body burned.[9] Some scholars even say her death marked the end of women in science for many hundreds of years.[7]

Medieval Europe

Hildegard of Bingen

The early parts of the European Middle Ages, also known as the Dark Ages, were marked by the decline of the Roman Empire. The Latin West was left with great difficulties that affected the continent's intellectual production dramatically. Although nature was still seen as a system that could be comprehended in the light of reason, there was little innovative scientific inquiry.[10] Giving credit to the Arabic world, they were the ones who kept the science alive by writing original scholarly work and by making copies of manuscripts from Classical periods.[11] During this period, Christianity was on the rise and with it came the recovery of Western civilization. Western civilization was slowly on the rise because of the monasteries and nunneries that nurtured the skills of reading and writing, and collected and copied many writings of scholars of the past.[11] Without the rise of Christianity and the preservation of past works by the Arabic world, the western civilization would have not been reawakened.

As it mentioned before, convents were an important place of education for women during this period, for the monasteries and nunneries encourage the skills of reading and writing, and some of these communities provided opportunities for women to contribute to scholarly research.[11] An example is the German abbess Hildegard of Bingen (1098–1179 A.D), a famous philosopher and botanists, known for her prolific writings include treatments of various scientific subjects, including medicine, botany and natural history (c.1151–58).[12] Another famous German abbess was Hroswitha of Gandersheim (935–1000 A.D.)[11] that also helped encourage women to be intellectual. However, with the growth in number and power of nunneries, the all-male clerical hierarchy was not welcomed toward it, and thus it stirred up conflict by having backlash against women's advancement. That impacted many religious orders closed on women and disbanded their nunneries, and overall excluding women from the ability to learn to read and write. With that, the world of science became closed off to women, limiting women's influence in science.[11]

Entering the 11th century, the first universities emerged. Women were, for the most part, excluded from university education.[1] However, there were some exceptions. The Italian University of Bologna, for example, allowed women to attend lectures from its inception, in 1088.[13]

The attitude to educating women in medical fields in Italy appears to have been more liberal than in other places. The physician, Trotula di Ruggiero, is supposed to have held a chair at the Medical School of Salerno in the 11th century, where she taught many noble Italian women, a group sometimes referred to as the "ladies of Salerno".[5] Several influential texts on women's medicine, dealing with obstetrics and gynecology, among other topics, are also often attributed to Trotula.

Dorotea Bucca was another distinguished Italian physician. She held a chair of philosophy and medicine at the University of Bologna for over forty years from 1390.[13][14][15][16] Other Italian women whose contributions in medicine have been recorded include Abella, Jacobina Félicie, Alessandra Giliani, Rebecca de Guarna, Margarita, Mercuriade (fourteenth century), Constance Calenda, Calrice di Durisio (15th century), Constanza, Maria Incarnata and Thomasia de Mattio.[14][17]

Despite the success of some women, cultural biases affecting their education and participation in science were prominent in the Middle Ages. For example, St. Thomas Aquinas, a Christian scholar, wrote, referring to women, "She is mentally incapable of holding a position of authority."[1]

Scientific Revolution (sixteenth, and seventeenth centuries)

Margaret Cavendish, a seventeenth-century aristocrat, took part in some of the most important scientific debates of that time. She was however, not inducted into the English Royal Society, although she was once allowed to attend a meeting. She wrote a number of works on scientific matters, including Observations upon Experimental Philosophy (1666) and Grounds of Natural Philosophy. In these works she was especially critical of the growing belief that humans, through science, were the masters of nature. The 1666 work attempted to heighten female interest in science. The observations provided a critique of the experimental science of Bacon and criticized microscopes as imperfect machines.[18]

In Germany the tradition of female participation in craft production enabled some women to become involved in observational science, especially astronomy. Between 1650 and 1710, women were 14% of German astronomers.[19] The most famous female astronomer in Germany was Maria Winkelmann. She was educated by her father and uncle and received training in astronomy from a nearby self-taught astronomer. Her chance to be a practising astronomer came when she married Gottfried Kirch, Prussia's foremost astronomer. She became his assistant at the astronomical observatory operated in Berlin by the Academy of Science. She made original contributions, including the discovery of a comet. When her husband died, Winkelmann applied for a position as assistant astronomer at the Berlin Academy – for which she had experience. As a woman – with no university degree – she was denied the post. Members of the Berlin Academy feared that they would establish a bad example by hiring a woman. "Mouths would gape", they said.[20]

Winkelmann's problems with the Berlin Academy reflect the obstacles women faced in being accepted in scientific work, which was considered to be chiefly for men. No woman was invited to either the Royal Society of London nor the French Academy of Sciences until the twentieth century. Most people in the seventeenth century viewed a life devoted to any kind of scholarship as being at odds with the domestic duties women were expected to perform.

A founder of modern botany and zoology, the German Maria Sibylla Merian (1647–1717), spent her life investigating nature. When she was thirteen, Sibylla began growing caterpillars and studying their metamorphosis into butterflies. She kept a "Study Book" which recorded her investigations into natural philosophy. In her first publication, The New Book of Flowers, she used imagery to catalogue the lives of plants and insects. After her husband died, and her brief stint of living in Siewert, she and her daughter journeyed to Paramaribo for two years to observe insects, birds, reptiles, and amphibians.[21] She returned to Amsterdam and published The Metamorphosis of the Insects of Suriname, which "revealed to Europeans for the first time the astonishing diversity of the rain forest."[22] She was a botanist and entomologist who was known for her artistic illustrations of plants and insects. Uncommon for that era, she traveled to South America and Surinam, where, assisted by her daughters, she illustrated the plant and animal life of those regions.[23]

Overall, the Scientific Revolution did little to change people's ideas about the nature of women. According to Jackson Spielvogel, 'Male scientists used the new science to spread the view that women were by nature inferior and subordinate to men and suited to play a domestic role as nurturing mothers. The widespread distribution of books ensured the continuation of these ideas'.[24]

Eighteenth century

The eighteenth century was characterized by three divergent views towards woman: that women were mentally and socially inferior to men, that they were equal but different, and that women were potentially equal in both mental ability and contribution to society. While individuals such as Jean-Jacques Rousseau believed women's roles were confined to motherhood and service to their male partners, the Enlightenment was a period in which women experienced expanded roles in the sciences.[25] The rise of salon culture in Europe brought philosophers and their conversation to an intimate setting where men and women met to discuss contemporary political, social, and scientific topics.[26] While Jean-Jacques Rousseau attacked women-dominated salons as producing ‘effeminate men’ that stifled serious discourse, salons were characterized in this erea by the mixing of the sexes.[27] Through salons and their work in mathematics, physics, botany, and philosophy, women began to have a significant impact during the Enlightenment. Women were not entirely excluded from being officially acknowledged by the scientific world.

The first woman to earn a university chair in a scientific field of studies in Europe (indeed in any field), Laura Bassi,[28] was also the third woman to obtain a university degree in the Western world. She was central to introducing Newton's ideas of physics and natural philosophy to Southern Europe, presenting numerous dissertations on the issues of gravity.[28]

In 1741, Prussian king Frederick II. allowed Dorothea Erxleben (1715–1762) to study medicine at the University of Halle. She was the first German women to receive a PhD (1754). In 1742 Dorothea Erxleben published a tract arguing that women should be allowed to attend university.[29]

In 1741, Charlotta Frölich, the first female historian in Sweden, became the first of her sex to be published by the Royal Swedish Academy of Science, and in 1748, Eva Ekeblad became the first woman inducted into that academy.[30]

In maths, Italian Maria Gaetana Agnesi was the first woman to write a mathematics handbook and the first woman appointed as a mathematics professor at a university (although she never taught). In 1748 she wrote a widely used text on finite and infinitesimal analysis.[31]

As many experiments took place in the home, women were well located to assist their husbands and family members with experiments. Among the best known of these scientific wives was Marie-Anne Pierrette Paulze, who married Antoine Lavoisier at thirteen and became his assistant in his home laboratory, in which he discovered oxygen. Mme. Lavoisier spoke English, and translated her husband's correspondence with English chemists, and Richard Kirwan's "Essay on Phlogiston," a key text in the controversy with English chemists such as Joseph Priestley over the nature of heat in chemical reactions. Mme Lavoisier also took drawing lessons from Jacques-Louis David and drew the diagrams for her husband's "Traite Elementaire de Chimie" (1789). Mme. Lavoisier maintained a small but lively salon and corresponded with French scientists and naturalists, many of whom were impressed by her intellect.

Science personified as a woman, illuminating nature with her light. Museum ticket from late eighteenth century

Although women excelled in many scientific areas during the eighteenth century, they were discouraged from learning about plant reproduction. Carl Linnaeus' system of plant classification based on sexual characteristics drew attention to botanical licentiousness, and people feared that women would learn immoral lessons from nature's example. Women were often depicted as both innately emotional and incapable of objective reasoning, or as natural mothers reproducing a natural, moral society.[32]

Even with such characterizations, author Lady Mary Wortley Montagu, known for her prolific letter writing, pioneered smallpox inoculation in England. She first observed the inoculations while visiting the Ottoman Empire, where she wrote detailed accounts of the practice in her letters [8].

Laura Bassi (1711–1778), as a member of the Italian Academy of the Institute of Sciences and a chair of the Institute of Experimental Physics, became the world's first female professor.[33]

The English Caroline Herschel added to the scientific knowledge of the time. Herschel, a great astronomer, who was born in Hanover but moved to England where she acted as an assistant to her brother, William Herschel. There she learned mathematics. She received a small salary from King George III (agnesscott.edu) and was the first woman to be recognized for a scientific position. She discovered eight comets between 1786 and 1797, and submitted an Index to Flamsteed's Observations of the Fixed Stars (including over five hundred omitted stars) to the Royal Society in 1798, becoming the first woman to present a paper there. In 1835, she and Mary Fairfax Somerville were the first two women to be awarded honorary memberships in the Royal Astronomical Society (source).

Although gender roles were largely defined in the 18th century, women experienced great advances in science. Whether it was through Emilie du Châtelet in translating Newton's Principia or Caroline Herschel discovering eight comets, women made great strides toward gender equality in the sciences.

Early nineteenth century

Science remained a largely amateur profession during the early part of the nineteenth century. Women's contributions were limited by their exclusion from most formal scientific education, but began to be recognized by admittance into learned societies during this period.

Scottish scientist Mary Fairfax Somerville carried out experiments in magnetism, presenting a paper entitled 'The Magnetic Properties of the Violet Rays of the Solar Spectrum' to the Royal Society in 1826, the second woman to do so. She also wrote several mathematical, astronomical, physical and geographical texts, and was a strong advocate for women's education. In 1835, she and Caroline Herschel were the first two women elected as Honorary Members of the Royal Astronomical Society.[34]

English mathematician Ada, Lady Lovelace, a pupil of Somerville, corresponded with Charles Babbage about applications for his analytical engine. In her notes (1842–3) appended to her translation of Luigi Menabrea's article on the engine, she foresaw wide applications for it as a general-purpose computer, including composing music. She has been credited as writing the first computer program, though this has been disputed.[35]

In Germany, institutes for "higher" education of women (Höhere Töchterschule, in some regions called Lyzeum) were founded at the beginning of the century.[36] The Deaconess Institute at Kaiserswerth was established in 1836 to instruct women in nursing. Elizabeth Fry visited the institute in 1840 and was inspired to found the London Institute of Nursing, and Florence Nightingale studied there in 1851.[37]

In the US, Maria Mitchell made her name by discovering a comet in 1847, but also contributed calculations to the Nautical Almanac produced by the United States Naval Observatory. She became the first woman member of the American Academy of Arts and Sciences in 1848 and of the American Association for the Advancement of Science in 1850.

Other notable female scientists during this period include:[3]

Late ninteenth century in Europe

The latter part of the nineteenth century saw a rise in educational opportunities for women. Schools aiming to provide education for girls similar to that afforded to boys were founded in the UK, including the North London Collegiate School (1850), Cheltenham Ladies' College (1853) and the Girls' Public Day School Trust schools (from 1872). The first UK women's university college, Girton, was founded in 1869, and others soon followed: Newnham (1871) and Somerville (1879).

The Crimean War (1854–6) contributed to establishing nursing as a profession, making Florence Nightingale a household name. A public subscription allowed Nightingale to establish a school of nursing in London in 1860, and schools following her principles were established throughout the UK.[37] Nightingale was also a pioneer in public health and a statistician.

James Barry became the first British woman to gain a medical qualification in 1812, passing as a man. Elizabeth Garret Anderson was the first openly female Briton to qualify medically, in 1865. With Sophia Jex-Blake, American Elizabeth Blackwell and others, Garret Anderson founded the first UK medical school to train women, the London School of Medicine for Women, in 1874.

Annie Scott Dill Maunder was a pioneer in astronomical photography, especially of sunspots. A mathematics graduate of Girton College, Cambridge, she was first hired (in 1890) to be an assistant to Edward Walter Maunder, discoverer of the Maunder Minimum, the head of the solar department at Greenwich Observatory. They worked together to observe sunspots and to refine the techniques of solar photography. They married in 1895. Annie's mathematical skills made it possible to analyse the years of sunspot data that Maunder had been collecting at Greenwich. She also designed a small, portable wide-angle camera with a 1.5-inch-diameter (38 mm) lens. In 1898, the Maunders traveled to India, where Annie took the first photographs of the sun's corona during a solar eclipse. By analysing the Cambridge records for both sunspots and geomagnetic storm, they were able to show that specific regions of the sun's surface were the source of geomagnetic storms and that the sun did not radiate its energy uniformly into space, as William Thomson, 1st Baron Kelvin had declared.[38]

In Prussia women could go to university from 1894 and were allowed to receive a PhD. In 1908 all remaining restrictions for women were terminated.

Other notable female scientists during this period include:[3][39]

Late nineteenth century in the United States

In the later nineteenth century the rise of the women's college provided jobs for women scientists, and opportunities for education. Women's colleges produced a disproportionate number of women who went on for PhDs in science. Many coeducational colleges and universities also opened or started to admit women during this period; such institutions included just over 3000 women in 1875, by 1900 numbered almost 20,000.[39]

An example is Elizabeth Blackwell, who became the first certified female doctor in the US when she graduated from Geneva Medical College in 1849.[40] With her sister, Emily Blackwell, and Marie Zakrzewska, Blackwell founded the New York Infirmary for Women and Children in 1857 and the first women's medical college in 1868, providing both training and clinical experience for women doctors. She also published several books on medical education for women.

In 1876, Elizabeth Bragg became the first woman to graduate with a civil engineering degree in the United States, from the University of California, Berkeley.[41]

Early twentieth century

Europe before World War II

Influential female scientists born in the 19th century: Ada Lovelace, Marie Curie, Maria Montessori, and Emmy Noether.

Marie Skłodowska-Curie, the first woman to win a Nobel prize in 1903 (physics), went on to become a double Nobel prize winner in 1911 (chemistry), both for her work on radiation. She was the first person to win two Nobel prizes, a feat accomplished by only three others since then.

Alice Perry is understood to be the first woman to graduate with a degree in civil engineering in the then United Kingdom of Great Britain and Ireland, in 1906 at Queen's College, Galway, Ireland.[42]

Lise Meitner played a major role in the discovery of nuclear fission. As head of the physics section at the Kaiser Wilhelm Institute in Berlin she collaborated closely with the head of chemistry Otto Hahn on atomic physics until forced to flee Berlin in 1938. In 1939, in collaboration with her nephew Otto Frisch, Meitner derived the theoretical explanation for an experiment performed by Hahn and Fritz Strassman in Berlin, thereby demonstrating the occurrence of nuclear fission. The possibility that Fermi's bombardment of uranium with neutrons in 1934 had instead produced fission by breaking up the nucleus into lighter elements, had actually first been raised in print in 1934, by chemist Ida Noddack (co-discover of the element rhenium), but this suggestion had been ignored at the time, as no group made a concerted effort to find any of these light radioactive fission products.

Maria Montessori was the first woman in Southern Europe to qualify as a physician. She developed an interest in the diseases of children and believed in the necessity of educating those recognized to be ineducable. In the case of the latter she argued for the development of training for teachers along Froebelian lines and developed the principle that was also to inform her general educational program, which is the first the education of the senses, then the education of the intellect. Montessori introduced a teaching program that allowed defective children to read and write. She sought to teach skills not by having children repeatedly try it, but by developing exercises that prepare them.[43]

Emmy Noether revolutionized abstract algebra, filled in gaps in relativity, and was responsible for a critical theorem about conserved quantities in physics. One notes that the Erlangen program attempted to identify invariants under a group of transformations. On 16 July 1918, before a scientific organization in Göttingen, Felix Klein read a paper written by Emmy Noether, because she was not allowed to present the paper herself. In particular, in what is referred to in physics as Noether's theorem, this paper identified the conditions under which the Poincaré group of transformations (now called a gauge group) for general relativity defines conservation laws.[44] Noether's papers made the requirements for the conservation laws precise. Aamong mathematicians Noether is best known for her fundamental contributions to abstract algebra, where the adjective noetherian is nowadays commonly used on many sorts of objects.

Mary Cartwright was a British mathematician who was the first to analyze a dynamical system with chaos. Inge Lehmann, a Danish seismologist, first suggested in 1936 that inside the Earth's molten core there may be a solid inner core.[45] Women such as Margaret Fountaine continued to contribute detailed observations and illustrations in botany, entomology, and related observational fields. Joan Beauchamp Procter, an outstanding herpetologist, was the first woman Curator of Reptiles for the Zoological Society of London at London Zoo.

United States before World War II

Women moved into science in significant numbers by 1900, helped by the women's colleges and by opportunities at some of the new universities. Margaret Rossiter's books Women Scientists in America: Struggles and Strategies to 1940 and Women Scientists in America: Before Affirmative Action 1940–1972 provide an overview of this period, stressing the opportunities women found in separate women's work in science.[46][47]

In 1892, Ellen Swallow Richards called for the "christening of a new science" – "oekology" (ecology) in a Boston lecture. This new science included the study of "consumer nutrition" and environmental education. This interdisciplinary branch of science was later specialized into what is currently known as ecology, while the consumer nutrition focus split off and was eventually relabeled as home economics.,[48][49] which provided another avenue for women to study science. Richards helped to form the American Home Economics Association, which published a journal, the Journal of Home Economics, and hosted conferences. Home economics departments were formed at many colleges, especially at land grant institutions. In her work at MIT, Ellen Richards also introduced the first biology course in its history as well as the focus area of sanitary engineering.

Women also found opportunities in botany and embryology. In psychology, women earned doctorates but were encouraged to specialize in educational and child psychology and to take jobs in clinical settings, such as hospitals and social welfare agencies.

In 1901, Annie Jump Cannon first noticed that it was a star's temperature that was the principal distinguishing feature among different spectra. This led to re-ordering of the ABC types by temperature instead of hydrogen absorption-line strength. Due to Cannon's work, most of the then-existing classes of stars were thrown out as redundant. Afterward, astronomy was left with the seven primary classes recognized today, in order: O, B, A, F, G, K, M;[50] that has since been extended.

Woman sitting at desk writing, with short hair, long-sleeved white blouse and vest
Henrietta Swan Leavitt made fundamental contributions to astronomy[51]

Henrietta Swan Leavitt first published her study of variable stars in 1908. This discovery became known as the "period-luminosity relationship" of Cepheid variables.[52] Our picture of the universe was changed forever, largely because of Leavitt's discovery. The accomplishments of Edwin Hubble, renowned American astronomer, were made possible by Leavitt's groundbreaking research and Leavitt's Law. "If Henrietta Leavitt had provided the key to determine the size of the cosmos, then it was Edwin Powell Hubble who inserted it in the lock and provided the observations that allowed it to be turned", wrote David H. and Matthew D.H. Clark in their book Measuring the Cosmos.[53]

Hubble often said that Leavitt deserved the Nobel for her work.[54] Gösta Mittag-Leffler of the Swedish Academy of Sciences had begun paperwork on her nomination in 1924, only to learn that she had died of cancer three years earlier[55] (the Nobel prize cannot be awarded posthumously).

In 1925, Harvard graduate student Cecilia Payne-Gaposchkin demonstrated for the first time from existing evidence on the spectra of stars that stars were made up almost exclusively of hydrogen and helium, one of the most fundamental theories in stellar astrophysics.[50][52]

Canadian born Maud Menten worked in the US and Germany. Her most famous work was on enzyme kinetics together with Leonor Michaelis, based on earlier findings of Victor Henri. This resulted in the Michaelis–Menten equations. Menten also invented the azo-dye coupling reaction for alkaline phosphatase, which is still used in histochemistry. She characterised bacterial toxins from B. paratyphosus, Streptococcus scarlatina and Salmonella ssp., and conducted the first electrophoretic separation of proteins in 1944. She worked on the properties of hemoglobin, regulation of blood sugar level, and kidney function.

World War II brought some new opportunities. The Office of Scientific Research and Development, under Vannevar Bush, began in 1941 to keep a registry of men and women trained in the sciences. Because there was a shortage of workers, some women were able to work in jobs they might not otherwise have accessed. Many women worked on the Manhattan Project or on scientific projects for the United States military services. Women who worked on the Manhattan Project included Leona Woods Marshall, Katharine Way, and Chien-Shiung Wu.

Women in other disciplines looked for ways to apply their expertise to the war effort. Three nutritionists, Lydia J. Roberts, Hazel K. Stiebeling, and Helen S. Mitchell, developed the Recommended Dietary Allowance in 1941 to help military and civilian groups make plans for group feeding situations. The RDAs proved necessary, especially, once foods began to be rationed. Rachel Carson worked for the United States Bureau of Fisheries, writing brochures to encourage Americans to consume a wider variety of fish and seafood. She also contributed to research to assist the Navy in developing techniques and equipment for submarine detection.

Women in psychology formed the National Council of Women Psychologists, which organized projects related to the war effort. The NCWP elected Florence Laura Goodenough president. In the social sciences, several women contributed to the Japanese Evacuation and Resettlement Study, based at the University of California. This study was led by sociologist Dorothy Swaine Thomas, who directed the project and synthesized information from her informants, mostly graduate students in anthropology. These included Tamie Tsuchiyama, the only Japanese-American woman to contribute to the study, and Rosalie Hankey Wax.

In the United States Navy, female scientists conducted a wide range of research. Mary Sears, a planktonologist, researched military oceanographic techniques as head of the Hydgrographic Office's Oceanographic Unit. Florence Van Straten, a chemist, worked as an aerological engineer. She studied the effects of weather on military combat. Grace Hopper, a mathematician, became one of the first computer programmers for the Mark I computer. Mina Spiegel Rees, also a mathematician, was the chief technical aide for the Applied Mathematics Panel of the National Defense Research Committee.

Gerti Cori was a biochemist who discovered the mechanism by which glycogen, a derivative of glucose, is transformed in the muscles to form lactic acid, and is later reformed as a way to store energy. For this discovery she and her colleagues were awarded the Nobel prize in 1947, making her the third woman and the first American woman to win a Nobel Prize in science. She was the first woman ever to be awarded the Nobel Prize in Physiology or Medicine. Cori is among several scientists whose works are commemorated by a U.S. postage stamp.[56]

Later 20th century

Nina Byers notes that before 1976, fundamental contributions of women to physics were rarely acknowledged. Women worked unpaid or in positions lacking the status they deserved. That imbalance is gradually being redressed.

In the early 1980s, Margaret Rossiter presented two concepts for understanding the statistics behind women in science as well as the disadvantages women continued to suffer. She coined the terms "hierarchical segregation" and "territorial segregation." The former term describes the phenomenon in which the further one goes up the chain of command in the field, the smaller the presence of women. The latter describes the phenomenon in which women "cluster in scientific disciplines."[57]:33–34

A recent book titled Athena Unbound provides a life-course analysis (based on interviews and surveys) of women in science from early childhood interest, through university, graduate school and the academic workplace. The thesis of this book is that "Women face a special series of gender related barriers to entry and success in scientific careers that persist, despite recent advances".[58]

The L'Oréal-UNESCO Awards for Women in Science were set up in 1998, with prizes alternating each year between the materials science and life sciences. One award is given for each geographical region of Africa and the Middle East, Asia-Pacific, Europe, Latin America and the Caribbean, and North America.

Europe after World War II

United States after World War II

Australia after World War II

Israel after World War II

Latin America

Maria Nieves Garcia-Casal, the first scientist and nutritionist woman from Latin America to lead the Latin America Society of Nutrition.

Nobel laureates

Further information: List of female Nobel laureates

The Nobel Prize and Prize in Economic Sciences have been awarded to women 41 times between 1901 and 2010. One woman, Marie Sklodowska-Curie, has been honored twice, with the 1903 Nobel Prize in Physics and the 1911 Nobel Prize in Chemistry. This means that 40 women in total have been awarded the Nobel Prize between 1901 and 2010. 17 women have been awarded the Nobel Prize in physics, chemistry, physiology or medicine.[2]

Physics

Chemistry

Physiology or medicine

Statistics

Statistics are used to indicate disadvantages faced by women in science, and also to track positive changes of employment opportunities and incomes for women in science.[57]:33

Women appear to do less well than men (in terms of degree, rank, and salary) in the fields that have been traditionally dominated by women, such as nursing. In 1991 women attributed 91% of the PhDs in nursing, and men held 4% of full professorships in nursing. In the field of psychology, where women earn the majority of PhDs, women do not fill the majority of high rank positions in that field.

Women's lower salaries in the scientific community are also reflected in statistics. According to the data provided in 1993, the median salaries of female scientists and engineers with doctoral degrees were 20% less than men.[57]:35 This data can be explained as there was less participation of women in high rank scientific fields/positions and a female majority in low-paid fields/positions. However, even with men and women in the same scientific community field, women are typically paid 15–17% less than men. In addition to the gender gap, there were also salary differences between ethnicity: African-American women with more years of experiences earn 3.4% less than European-American women with similar skills, while Asian women engineers out-earn both Africans and Europeans.[67]

Women are also under-represented in the sciences as compared to their numbers in the overall working population. Within 11% of African-American women in the workforce, 3% are employed as scientists and engineers. Hispanics made up 8% of the total workers in the US, 3% of that number are scientists and engineers. Native Americans participation cannot be statistically measured.

Women tend to earn less than men in almost all industries, including government and academia. Women are less likely to be hired in highest-paid positions. The data showing the differences in salaries, ranks, and overall success between the genders is often claimed to be a result of women's lack of professional experience. The rate of women's professional achievement is increasing. In 1996, the salaries for women in professional fields increased from 85% to 95% relative to men with similar skills and jobs. Young women between the age of 27 and 33 earned 98%, nearly as much as their male peers. In the total workforce of the United States, women earn 74% as much as their male counterparts (in the 1970s they made 59% as much as their male counterparts).[57]:33–37

Claudia Goldin, Harvard concludes in A Grand Gender Convergence: Its Last Chapter – "The gender gap in pay would be considerably reduced and might vanish altogether if firms did not have an incentive to disproportionately reward individuals who labored long hours and worked particular hours."[68]

Research on women's participation in the "hard" sciences such as physics and computer science speaks of the "leaky pipeline" model, in which the proportion of women "on track" to potentially becoming top scientists fall off at every step of the way, from getting interested in science and maths in elementary school, through doctorate, postdoctoral, and career steps. The leaky pipeline also applies in other fields. In biology, for instance, women in the United States have been getting Masters degrees in the same numbers as men for two decades, yet fewer women get PhDs; and the numbers of women principal investigators have not risen.[69]

What may be the cause of this "leaky pipeline" of women in the sciences? It is important to look at factors outside of academia that are occurring in women's lives at the same time they are pursuing their continued education and career search. The most outstanding factor that is occurring at this crucial time is family formation. As women are continuing their academic careers, they are also stepping into their new role as a wife and mother. These traditionally require at large time commitment and presence outside work. These new commitments do not fare well for the person looking to attain tenure. That is why women entering the family formation period of their life are 35% less likely to pursue tenure positions after receiving their PhD's than their male counterparts.[70]

In the UK, women occupied over half the places in science-related higher education courses (science, medicine, maths, computer science and engineering) in 2004/5.[71] However, gender differences varied from subject to subject: women substantially outnumbered men in biology and medicine, especially nursing, while men predominated in maths, physical sciences, computer science and engineering.

In the US, women with science or engineering doctoral degrees were predominantly employed in the education sector in 2001, with substantially fewer employed in business or industry than men.[72] According to salary figures reported in 1991, women earn anywhere between 83.6 percent to 87.5 percent that of a man's salary. An even greater disparity between men and women is the ongoing trend that women scientists with more experience are never as well-compensated as their male counterparts. The salary of a male engineer continues to experience growth as he gains experience whereas the female engineer sees her salary reach a plateau.[73]

Women, in the United States and many European countries, who succeed in science tend to be graduates of single-sex schools.[57](Chapter 3) Women earn 54% of all bachelor's degrees in the United States and 50% of those are in science. 9% of US physicists are women.[57](Chapter 2)

According to a Royal Astronomical Society Survey in 2011, 27% of all astronomy lecturers in Britain are female.[74]

Social, historical, and critical studies

Social effects

Beginning in the late twentieth century to present day, more and more women are becoming involved in science. However, women often find themselves at odds with expectations held towards them in relation to their scientific studies. For example, in 1968 James Watson questioned scientist Rosalind Franklin's place in the industry. He claimed that "the best place for a feminist was in another person's lab",[57]:76–77 most often a male's research lab. Women were and still are often critiqued of their overall presentation. In Franklin's situation, she was seen as lacking femininity for she failed to wear lipstick or revealing clothing.[57]:76–77 Women believed that in order to gain recognition, they needed to hide their feminine qualities, to thus appear more masculine. Women in the sixties were often forced to wear men's clothing, which often did not fit for they were too large or too short within the crotch area. Since most of their colleagues in science are men, women also find themselves left out of opportunities to discuss possible research opportunities. In Londa Scheibinger's book, Has Feminism Changed Science?, she explains how men discuss research outside of the lab, but this conversation is preceded by talk of sports and the like, thus excluding women.[57]:81–91 This causes women to seek other women in science to converse with, which in turn causes their final work to be looked down upon, for a male scientist was not involved.

According to Oxford University Press, the inequality toward women is “endorsed within cultures and entrenched within institutions [that] hold power to reproduce that inequality”.[75] There are various gendered barriers in social networks that prevent women from working in male-dominated fields and top management jobs. Social networks are based on the cultural beliefs such as schemas and stereotypes.[75] According to social psychology studies, top management jobs are more likely to have incumbent schemas that favor “an achievement-oriented aggressiveness and emotional toughness that is distinctly male in character”.[75] Certain gender stereotypes assume women to be less worthy and less competent than men so they are not qualified for the top management jobs. However, when the women try to prove their competence and power, they often face obstacles. They are likely to be seen as dislikable and untrustworthy when they are excelled at tasks viewed as masculine.[75] Also, women are likely to face the denial of credit for their achievements.[75] Social networks and gender stereotypes provide many contributions to the injustices that women have to face in their workplace and the limitations when they try to advance in male-dominated jobs and top management jobs. Women in professions like science, technology, and other related industries are likely to encounter the gendered barriers in their careers.[75] Based on the meritocratic explanations of gendered inequality, “as long as the people accept the mechanisms that produce unequal outcomes,” all the outcomes will be legitimated in the society.[75] When women try to deny the stereotypes and the discriminations by becoming “competent, integrated, well-liked”, the society is more likely to look at these impressions as selfishness or “being a whiner”.[75] In the United States, Title IX of the Education Amendments of 1972 provides opportunities for women to achieve to a wide range of education programs and activities by prohibiting sex discrimination.[76] The law states “No person in the United Stated shall, on the basis of sex, be excluded from participation in, be denied the benefits of, or be subject to discrimination under any educational program or activity receiving federal financial assistance”.[77]

Margaret W. Rossiter

Margaret Rossiter, an American historian of science, offered three concepts to explain the reasons behind the data in statistics and how these reasons disadvantaged women in the science industry. The first concept is hierarchical segregation.[78] This is a well-known phenomenon in society, that the higher the level and rank of power and prestige, the smaller the population of females participating. The hierarchical differences point out that there are fewer women participating at higher levels of both academia and industry. Based on data collected in 1982, women earn 54 percent of all bachelor's degrees in the United States, with 50 percent of these in science. The source also indicated that this number increased almost every year.[79] There are fewer women at the graduate level; they earn 40 percent of all doctorates, with 31 percent of these in science and engineering.

The second concept included in Rossiter's explanation of women in science is territorial segregation.[57]:34–35 The term refers to how female employment is often clustered in specific industries or categories in industries. Women stayed at home or took employment in feminine fields while men left the home to work. Although nearly half of the civilian work force is female, women still comprise the majority of low-paid jobs or jobs that society considered feminine. Statistics show that 60 percent of white professional women are nurses, daycare workers, or schoolteachers.[80] Territorial disparities in science are often found between the 1920s and 1930s, when different fields in science were divided between men and women. Men dominated the chemistry, medical sciences, and engineering, while women dominated the fields of botany, zoology, and psychology. The fields in which the majority of women are concentrated are known as the "soft" sciences and tend to have relatively low salaries.

Researchers collected the data on many differences between women and men in science. Rossiter found that in 1966, thirty-eight percent of female scientists held master's degrees compared to twenty-six percent of male scientists; but large proportions of female scientists were in environmental and nonprofit organizations.[81] During the late 1960s and 1970s, equal-rights legislation made the number of female scientists rise dramatically. The statistics from National Science Board (NSB) present the change at that time. The number of science degrees awarded to woman rose from seven percent in 1970 to twenty-four percent in 1985. In 1975 only 385 women received bachelor's degrees in engineering compared to 11,000 women in 1985, indicating the importance of legislation to the representation of women in science. Elizabeth Finkel claims that even if the number of women participating in scientific fields increases, the opportunities are still limited. Jabos who worked for NSB reported the pattern of women in receiving doctoral degrees in science: even though the numbers of female scientists with higher-level degrees increased, they still were consistently in a minority. Another reporter, Harriet Zuckerman, claims that when woman and man have similar abilities for a job, the probability of the woman getting the job is lower. Elizabeth Finkel agrees, saying, "In general, while woman and men seem to be completing doctorate with similar credentials and experience, the opposition and rewards they find are not comparable. Women tend to be treated with less salary and status, many policy makers notice this phenomenon and try to rectify the unfair situation for women participating in scientific fields."[81]

Media coverage

In 2013, journalist Christie Aschwanden noted that a type of media coverage of women scientists that "treats its subject's sex as her most defining detail" was still prevalent. She proposed a checklist, the "Finkbeiner test",[82] to help avoid this approach.[83] It was cited in the coverage of a much-criticized 2013 New York Times obituary of rocket scientist Yvonne Brill that began with the words: "She made a mean beef stroganoff".[84]

Efforts to increase representation

A number of organizations have been set up to combat the stereotyping that may encourage girls away from careers in these areas. In the UK The WISE Campaign (Women into Science, Engineering and Construction) and the UKRC (The UK Resource Centre for Women in SET) are collaborating to ensure industry, academia and education are all aware of the importance of challenging the traditional approaches to careers advice and recruitment that mean some of the best brains in the country are lost to science. The UKRC and other women's networks provide female role models, resources and support for activities that promote science to girls and women. The Women's Engineering Society, a professional association in th UK, has been supporting women in engineering and science since 1919. In computing, the British Computer Society group BCSWomen is active in encouraging girls to consider computing careers, and in supporting women in the computing workforce.

In the United States, the Association for Women in Science is one of the most prominent organization for professional women in science. In 2011, the Scientista Foundation was created to empower pre-professional college and graduate women in science, technology, engineering and mathematics (STEM), to stay in the career track. There are also several organizations focused on increasing mentorship. One of the best known groups is Science Club for Girls, which pairs undergraduate mentors with high school and middle school mentees. In 2013, the Grolier Club in New York hosted a "landmark exhibition" titled "Extraordinary Women in Science & Medicine: Four Centuries of Achievement", showcasing the lives and works of 32 women scientists.[85]

The National Institute for Occupational Safety and Health (NIOSH) developed a video series] highlighting the stories of female researchers at NIOSH.[86] Each of the women featured in the videos share their journey into science, technology, engineering, or math (STEM), and offers encouragement to aspiring scientists.[86] NIOSH also partners with external organizations in efforts to introduce individuals to scientific disciplines and funds several science-based training programs across the country.[87][88]

Notable controversies and developments

In January 2005, Harvard University President Lawrence Summers sparked controversy when, at an NBER Conference on Diversifying the Science & Engineering Workforce, he made comments suggesting the lower numbers of women in high-level science positions may in part be due to innate differences in abilities or preferences between men and women. He noted the generally greater variability among men (compared to women) on tests of cognitive abilities,[89][90][91] leading to proportionally more men than women at both the lower and upper tails of the test score distributions. In his discussion of this, Summers said that "even small differences in the standard deviation [between genders] will translate into very large differences in the available pool substantially out [from the mean]".[92]

A study conducted at Lund University in 2010 and 2011 analysed the genders of invited contributors to News & Views in Nature and Perspectives in Science. It found that 3.8% of the Earth and environmental science contributions to News & Views were written by women even while the field was estimated to be 16–20% female in the United States. Nature responded by suggesting that, worldwide, a significantly lower number of Earth scientists were women, but nevertheless committed to address any disparity.[93]

In 2012, a journal article published in Proceedings of the National Academy of Sciences (PNAS) reported a gender bias among science faculty.[94] Faculty were asked to review a resume from a hypothetical student and report how likely they would be to hire or mentor that student, as well as what they would offer as starting salary. Two resumes were distributed randomly to the faculty, only differing in the names at the top of the resume (John or Jennifer). The male student was rated as significantly more competent, more likely to be hired, and more likely to be mentored. The median starting salary offered to the male student was greater than $3,000 over the starting salary offered to the female student. Both male and female faculty exhibited this gender bias. This study suggests bias may partly explain the persistent deficit in the number of women at the highest levels of scientific fields. Another study reported a that men are favored in some domains, such as biology tenure rates, but that the majority of domains were gender-fair; the authors interpreted this to suggest that the under-representation of women in the professorial ranks was not solely caused by sexist hiring, promotion, and remuneration.[95] In April 2015 Williams and Ceci published a set of five national experiments showing that hypothetical female applicants were favored by faculty for assistant professorships over identically-qualified men by a ratio of 2 to 1.[96][97]

In 2014, a controversy over the depiction of pinup women on Rosetta project scientist Matt Taylor's shirt during a press conference raised questions of sexism within the European Space Agency.[98] The shirt, which featured cartoon women with firearms, led to an outpouring of criticism and an apology after which Taylor "broke down in tears."[99]

In 2015, stereotypes about women in science were directed at Fiona Ingleby, research fellow in evolution, behavior, and environment at the University of Sussex, and Megan Head, postdoctoral researcher at the Australian National University, when they submitted a paper analyzing the progression of PhD graduates to postdoctoral positions in the life sciences to the journal PLOS ONE.[100] The authors received an email on March 27 informing them that their paper had been rejected due to its poor quality.[100] The email included comments from an anonymous reviewer, which included the suggestion that male authors be added in order to improve the quality of the science and serve as a means of ensuring that incorrect interpretations of the data are not included.[100] Ingleby posted excerpts from the email on Twitter on April 29 bringing the incident to the attention of the public and media.[100] The editor was dismissed from the journal and the reviewer was removed from the list of potential reviewers. A spokesman from PLOS apologized to the authors said they would be given the opportunity to have the paper reviewed again.[100]

On June 9, 2015, Nobel prize winning biochemist Tim Hunt spoke at the World Conference of Science Journalists in Seoul. Prior to applauding the work of women scientists, he described emotional tension, saying "you fall in love with them, they fall in love with you, and when you criticise them they cry."[101] Initially, his remarks were widely condemned and he was forced to resign from his position at University College London. However, multiple conference attendees gave accounts, including a partial transcript and a partial recording, maintaining that his comments were understood to be satirical before being taken out of context by the media.[102]

In 2016 an article published in JAMA Dermatology reported a significant and dramatic downward trend in the number of NIH-funded woman investigators in the field of dermatology and that the gender gap between male and female NIH-funded dermatology investigators was widening. The article concluded that this disparity was likely due to a lack of institutional support for women investigators.[103]

See also

References

  1. 1 2 3 Whaley, Leigh Ann. Women's History as Scientists. Santa Barbara, California: ABC-CLIO, INC. 2003.
  2. 1 2 "Nobel Prize Awarded Women".
  3. 1 2 3 4 "Time ordered list".
  4. 1 2 3 4 Rayner-Canham, Marelene. Women in Chemistry: Their Changing Roles from Alchemical Times to the Mid-Twentieth Century. Washington, DC: American Chemical Society; Chemical Heritage Foundation. pp. 1–2.
  5. 1 2 "Reframing the question".
  6. "Hypatia | mathematician, astronomer, and philosopher". Encyclopedia Britannica. Retrieved 2016-04-08.
  7. 1 2 3 Rayner-Canham, Marelene. Women in Chemistry: Their Changing Roles from Alchemical Times to the Mid-Twentieth Century. Washington, DC. pp. 3–4.
  8. A. B. Deakin, Michael (August 1995). "The Primary Sources for the Life and Work of Hypatia of Alexandria". History of Mathematics Paper 63. Retrieved 2016-04-07.
  9. 1 2 A. B. Deakin, Michael (1994). Hypatia and Her Mathematics. Mathematical Association of America. pp. 234–243.
  10. The End of the Classical World, (Lecture 12), in Lawrence M. Principe (2002) History of Science: Antiquity to 1700. Teaching Company, Course No. 1200
  11. 1 2 3 4 5 Rayner-Canham, Marelene. Women in Chemistry: Their Changing Roles from Alchemical Times to the Mid-Twentieth Century. American Chemical Society; Chemical Heritage Foundation. pp. 6–8.
  12. Hildegard von Bingen (Sabina Flanagan)
  13. 1 2 Edwards, J. S. (2002). "A Woman Is Wise: The Influence of Civic and Christian Humanism on the Education of Women in Northern Italy and England during the Renaissance" (PDF). Ex Post Facto: Journal of the History Students at San Francisco State University. XI.
  14. 1 2 Howard S. The Hidden Giants, p. 35, (Lulu.com; 2006) (Retrieved 22 August 2007)
  15. Brooklyn Museum: Elizabeth A. Sackler Center for Feminist Art: The Dinner Party: Heritage Floor: Dorotea Bucca (Retrieved 22 August 2007)
  16. Jex-Blake S (1873) The medical education of women, republished in The Education Papers: Women's Quest for Equality, 1850–1912 (Spender D, ed) p. 270] (Retrieved 22 August 2007)
  17. Walsh, J. J. Medieval Women Physicians' in Old Time Makers of Medicine: The Story of the Students and Teachers of the Sciences Related to Medicine During the Middle Ages, ch. 8, (Fordham University Press; 1911)]
  18. Whaley, Leigh Ann. Women's History as Scientists. (California: 2003), pg. 114.
  19. Spielvogel, Jackson J. Western Civilization, Volume B: 1300–1815. Thomson/Wadsworth, 2009. ISBN 978-0-495-50289-0
  20. Schiebinger, Londa (1992). "Maria Winkelmann at the Berlin Academy", in Gendered domains: rethinking public and private in women's history : essays from the Seventh Berkshire Conference on the History of Women. (Ithaca: 1992). 65.
  21. Ingrid D. Rowland. "The Flowering Genius of Maria Sibylla Merian". The New York Review of Books.
  22. http://www.bookrags.com/tandf/science-early-modern-to-late-eighteenth-tf/. Retrieved 16 December 2010. Missing or empty |title= (help)
  23. John Augustine Zahm; H. J. Mozans (1913), Woman in science, New York: Appleton, pp. 240=241
  24. "book" in Spielvogel, Jackson (2014) Western Civilisation. Toward a New Heaven and a New Earth: The Scientific Revolution. Cengage Learning. Chapter 16, p492.
  25. Whaley, Leigh Ann. Women's History as Scientists. (California: 2003), 118.
  26. "Redirect support".
  27. Watts, Ruth, Women in Science: A Social and Cultural History. (London and New York: 2007), pg. 62.
  28. 1 2 Findlen, Paula (1993). "Science As A Career In Enlightenment Italy : The Strategies Of Laura Bassi". Isis. 84: 440–469. doi:10.1086/356547. JSTOR 235642.
  29. "Erxleben, Dorothea (1715–1762)." Encyclopedia.com. HighBeam Research, n.d. Web. 24 Nov. 2014; Julia von Brencken: Doktorhut und Weibermütze. Dorothea Erxleben – die erste Ärztin. Biographischer Roman. Kaufmann, 1997.
  30. Flower Hunters.
  31. Ogilvie, Marilyn Bailey (1993). "Agnesi, Maria Gaetana". Women in Science: Antiquity through the Nineteenth Century. MIT Press. pp. 26–27. ISBN 0-262-65038-X.
  32. Watts, Ruth, Women in Science: A Social and Cultural History. (London and New York: 2007), pg. 63.
  33. Whaley, Leigh Ann. Women's History as Scientists. (California: 2003), pg. 137.
  34. Dreyer, ed. by J. L. E.; Turner, H. H. (1987). History of the Royal Astronomical Society. (Reprint [d. Ausg.] London, Wheldon & Wesley, 1923. ed.). Palo Alto, California: Reprinted for the Society by Blackwell Scientific Publications. p. 81. ISBN 0-632-02175-6.
  35. "Ada Lovelace, In Our Time – BBC Radio 4".
  36. Claus-Hinrich Offen; Schule in einer hanseatischen Bürgergesellschaft: zur Sozialgeschichte des niederen Schulwesens in Lübeck (1800–1866), 1990
  37. 1 2 The Cambridge Illustrated History of Medicine, R. Porter (editor), Cambridge University Press, 1996
  38. Clark, Stuart (2007). The Sun Kings – The Unexpected Tragedy of Richard Carrington and the Tale of How Modern Astronomy Began. Princeton University Press. pp. 140–146, 154–162.
  39. 1 2 "CONTRIBUTIONS OF 20TH CENTURY WOMEN TO PHYSICS".
  40. "Changing the Face of Medicine – Dr. Elizabeth Blackwell".
  41. "WEP Milestones". Berkeley Engineering. University of California, Berkeley. Retrieved 2011-11-24.
  42. "Alice Perry". Institution of Engineers of Ireland. Retrieved 2011-11-24.
  43. Phyllis Povell (2009). Montessori Comes to America: The Leadership of Maria Montessori and Nancy McCormick Rambusch. California, US: UPA. p. 170. ISBN 978-0-7618-4928-5.
  44. Emmy Noether (1918c) "Invariante Variationsprobleme" Nachrichten von der Gesellschaft der Wissenschaften der Göttingen, 235–257. Presented by Felix Klein 16 July 1918. Final printed version submitted September 1918. Paper denoted 1918c, in a Bibliography of Noether's work, pp. 173–182 of Emmy Noether in Bryn Mawr: Proceedings of a symposium sponsored by the Association for women in mathematics, in honor of Emmy Noether's 100th birthday (1983, Bhama Srinivasan and Judith Sally, eds.) Springer-Verlag ISBN 0-387-90838-2. Biographical information on Noether's life can be found on pp. 133–137 "Emmy Noether in Erlangen and Göttingen", and on pp. 139–146 "Emmy Noether in Bryn Mawr".
  45. "Inge Lehmann: Discoverer of the Earth's Inner Core".
  46. Rossiter 1982
  47. Rossiter 1995
  48. Kass-Simon, G. and Farnes, Patricia. Women of Science: Righting the Record. Bloomington, Indiana: Indiana University Press. 1993.
  49. Clarke, Robert. Ellen Swallow: The Woman Who Founded Ecology. Chicago: Follett. 1973.
  50. 1 2 Pogge, Richard (8 January 2006). "Introduction to Stars, Galaxies, & the Universe". Ohio State University Department of Astronomy.
  51. Hamblin, Jacob Darwin (2005). Science in the early twentieth century: an encyclopedia. ABC-CLIO. pp. 181–184. ISBN 1-85109-665-5.
  52. 1 2 Malatesta, Kerri (16 July 2010). "Delta Cephei". American Association of Variable Star Observers.
  53. David H. Clark; Matthew D.H. Clark (2004). Measuring the Cosmos: How Scientists Discovered the Dimensions of the Universe. Rutgers University Press. ISBN 0-8135-3404-6.
  54. Ventrudo, Brian (19 November 2009). "Mile Markers to the Galaxies". One-Minute Astronomer. Retrieved 25 February 2011.
  55. Singh, Simon (2005). Big Bang: The Origin of the Universe. HarperCollins. ISBN 0-00-716221-9. Retrieved 25 February 2011.
  56. "Women Subjects on United States Postage Stamps" (PDF). United States Postal Service. p. 6. Retrieved 21 October 2011.
  57. 1 2 3 4 5 6 7 8 9 10 Schiebinger, Londa (2001). Has Feminism Changed Science?. Cambridge, Massachusetts: Harvard University Press. ISBN 978-0-674-00544-0.
  58. Etzkowitz, Kemelgor & Uzzi 2000
  59. "ENIAC Programmers Project". Eniacprogrammers.org. Retrieved 2010-01-27.
  60. Ghez, Andrea. "Andrea Ghez - Speaker - TED.com".
  61. "American Scientists (Forever)". United States Postal Service. Retrieved 21 October 2011.
  62. "Women Subjects on United States Postage Stamps" (PDF). United States Postal Service. p. 5. Retrieved 21 October 2011.
  63. "Sally Ride Science Brings Cutting-Edge Science to the Classroom with New Content Rich Classroom Sets". Business Wire – Live PR. Business Wire – Live PR. 2007. Retrieved 7 October 2007.
  64. Heinrichs, Allison M. (2007). "Sally Ride encourages girls to engineer careers". Pittsburgh Tribune Review. Archived from the original on 20 November 2007. Retrieved 7 October 2007.
  65. Phillips, Tony (4 May 2011). "NASA Announces Results of Epic Space-Time Experiment". NASA Science News. Retrieved 2011-11-15.
  66. Tarter, Jill. "Jill Tarter - Speaker - TED.com".
  67. Schiebinger [2001], p. 37, citing "Women, Minorities". NSF. 1996: 72–74., Edward Silverman (19 August 1991). "New NSF Report on Salaries of Ph.D.'s Reveals Gender Gaps in All Categories". Scientist. 20 (5). and Edward Silverman (16 September 1991). "NSF's Ph.D. Salary Survey Finds Minorities Earn Less than Whites". Scientist. 21 (5).
  68. Goldin, Claudia (2014). "A Grand Gender Convergence: Its Last Chapter†". American Economic Review. 104 (4): 1091–1119. doi:10.1257/aer.104.4.1091.
  69. Louise Luckenbill-Edds, "The 'Leaky Pipeline:' Has It Been Fixed?",The American Society for Cell Biology 2000 WICB / Career Strategy Columns (1 November 2000).
  70. "Staying Competitive". name. Retrieved 2015-11-22.
  71. "Table 2e – All HE students by level of study, subject of study(#5), domicile and gender 2004/05". 9 March 2007. Archived from the original on 9 March 2007.
  72. Hahm, J-o. Data on Women in S&E. From: Women, Minorities and Persons With Disabilities in Science and Engineering, NSF 2004 Archived 13 May 2006 at the Wayback Machine.
  73. Margaret A. Einsenhart, Elizabeth Finkel (2001). "1". In Muriel Lederman, Ingrid Bartsch. The Gender and Science Reader. New York: Routledge. pp. 16–17. ISBN 978-0-415-21358-5.
  74. "RAS statement on Hiranya Peiris and Maggie Aderin-Pocock". Ras.org.uk. 2014-03-21. Retrieved 2016-03-22.
  75. 1 2 3 4 5 6 7 8 Cech, Erin A.; Blair-Loy, Mary (2010-01-01). "Perceiving Glass Ceilings? Meritocratic versus Structural Explanations of Gender Inequality among Women in Science and Technology". Social Problems. 57 (3): 371–397. doi:10.1525/sp.2010.57.3.371. JSTOR 10.1525/sp.2010.57.3.371.
  76. "AAUW: Empowering Women Since 1881". AAUW: Empowering Women Since 1881. Retrieved 2016-10-07.
  77. "AAUW: Empowering Women Since 1881". AAUW: Empowering Women Since 1881. Retrieved 2016-10-07.
  78. Tierney, Helen (2002). "Science And Women". Women's Studies Encyclopedia. Retrieved 9 November 2013.
  79. Hahm, J-o. Data on Women in S&E. From: Women, Minorities and Persons With Disabilities in Science and Engineering, NSF 2004
  80. "Science and Engineering Indicators 2006" (PDF).
  81. 1 2 The Gender and Science Reader,edited by Muriel Lederman And Ingrid Bartsch,section one,Eisenhart and Elizabeth Finkel, 2001, first published by Routledge.
  82. Aschwanden, Christie (5 March 2013). "The Finkbeiner Test: What matters in stories about women scientists?". Double X Science. Retrieved 31 March 2013.
  83. Brainard, Curtis (22 March 2013). "'The Finkbeiner Test' Seven rules to avoid gratuitous gender profiles of female scientists". Columbia Journalism Review. Retrieved 31 March 2013.
  84. Gonzalez, Robert T. (31 March 2013). "The New York Times fails miserably in its obituary for rocket scientist Yvonne Brill". io9. Retrieved 31 March 2013.
  85. "Landmark exhibition recognizes the achievements of women in science and medicine at The Grolier Club". artdaily.org. 2013-12-22. Retrieved 2013-12-22.
  86. 1 2 "CDC – Women's Safety and Health Issues at Work – NIOSH Workplace Safety and Health Topic – Science Speaks: A Focus on NIOSH Women in Science".
  87. "CDC – NIOSH Grants and Funding – Extramural Research and Training Programs – Research and Training".
  88. "CDC – NIOSH Training and Workforce Development".
  89. Hedges, L. V.; Nowell, A. (1995). "Sex differences in mental scores, variability, and numbers of high scoring individuals" (PDF). Science. 269 (5220): 41–45. doi:10.1126/science.7604277. PMID 7604277.
  90. Lehrke, R. (1997). Sex linkage of intelligence: The X-Factor. NY: Praeger.
  91. Lubinski, D.; Benbow, C. M. (2006). "Study of mathematically precocious youth after 35 years" (PDF). Perspectives on Psychological Science. 1 (4): 316–345. doi:10.1111/j.1745-6916.2006.00019.x. PMID 26151798.
  92. Archive of: Remarks at NBER Conference on Diversifying the Science & Engineering Workforce. 14 January 2005.
  93. Nordahl, Marianne (2012-09-08). "Gender bias in leading journals". Science Nordic. Retrieved 2015-10-27. should we find that the News & Views section is indeed under-representing women, we will certainly take steps to redress the balance.
  94. Moss-Racusin, Corinne A.; John F. Dovidiob; Victoria L. Brescollc; Mark J. Grahama; Jo Handelsman (August 2012). "Science faculty's subtle gender biases favor male students". PNAS. 109 (41): 16395–16396. doi:10.1073/pnas.1211286109. PMC 3478626Freely accessible. PMID 22988126.
  95. Ceci, S. J.; Ginther, D. K.; Kahn, S.; Williams, W. M. (3 November 2014). "Women in Academic Science: A Changing Landscape". Psychological Science in the Public Interest. 15 (3): 75–141. doi:10.1177/1529100614541236. PMID 26172066.
  96. Williams, Wendy M.; Ceci, Stephen J. (28 April 2015). "National hiring experiments reveal 2:1 faculty preference for women on STEM tenure track". Proceedings of the National Academy of Sciences. 112 (17): 5360–5365. doi:10.1073/pnas.1418878112. PMC 4418903Freely accessible. PMID 25870272.
  97. Williams, Wendy M.; Ceci, Stephen J. (2015-04-28). "National hiring experiments reveal 2:1 faculty preference for women on STEM tenure track". Proceedings of the National Academy of Sciences. 112 (17): 5360–5365. doi:10.1073/pnas.1418878112. ISSN 0027-8424. PMC 4418903Freely accessible. PMID 25870272.
  98. Bell, Alice (13 November 2014). "Why women in science are annoyed at Rosetta mission scientist's clothing". The Guardian. Retrieved 18 November 2014.
  99. Molloy, Antonia (14 November 2014). "Dr Matt Taylor apologises for controversial 'sexist' shirt worn after Rosetta mission comet landing". independent.co.uk. 14 November 2014. Retrieved 30 November 2014.
  100. 1 2 3 4 5 Elsei, Holly. ″′Sexist′ peer review causes storm online.″ Times Higher Education 30 April 2015: Web.
  101. Radcliffe, Rebecca (10 June 2015). "Nobel scientist Tim Hunt: female scientists cause trouble for men in labs". The Guardian. Retrieved 10 June 2015.
  102. Moody, Oliver (18 July 2015). "Recording 'shows Sir Tim was joking'". The Times. Retrieved 18 July 2015.
  103. Cheng, Michelle A.; Annie Sukhov; Hawa Sultani; Koungmi Kim; Emanual Maverakis (May 2016). "Trends in National Institutes of Health Funding of Principal Investigators in Dermatology Research by Academic Degree and Sex". JAMA Dermatology. doi:10.1001/jamadermatol.2016.0271.

Further reading

  • Byers, Nina; Williams, Gary (2006). Out of the Shadows: Contributions of Twentieth-Century Women to Physics. Cambridge University Press. ISBN 0-521-82197-5.  (Cambridge Univ Press catalogue)
  • Etzkowitz, Henry; Kemelgor, Carol; Uzzi, Brian (2000). Athena Unbound: The advancement of women in science and technology. Cambridge University Press. ISBN 0-521-78738-6. 
  • Fara, Patricia (2004). Pandora's Breeches: Women, Science & Power in the Enlightenment. London: Pimlico. ISBN 1-84413-082-7. 
  • Gates, Barbara T. (1998). Kindred Nature: Victorian and Edwardian Women Embrace the Living World. The University of Chicago Press. ISBN 0-226-28443-3. 
  • Herzenberg, Caroline L. (1986). Women Scientists from Antiquity to the Present. Locust Hill Press. ISBN 0-933951-01-9. 
  • Howes, Ruth H.; Herzenberg, Caroline L. (1999). Their Day in the Sun: Women of the Manhattan Project. Temple University Press. ISBN 1-56639-719-7. 
  • Keller, Evelyn Fox (1985). Reflections on gender and science. New Haven: Yale University Press. ISBN 0-300-06595-7. 
  • National Academy of Sciences (2006). Beyond Bias and Barriers: Fulfilling the Potential of Women in Academic Science and Engineering. Washington, D.C.: The National Academies Press. ISBN 0-309-10320-7. 
  • Ogilvie, Marilyn Bailey (1993). Women in Science: Antiquity through the Nineteenth Century. MIT Press. ISBN 0-262-65038-X. 
  • Pomeroy, Claire, "Academia's Gender Problem", Scientific American, vol. 314, no. 1 (January 2016), p. 11.
  • Sue Rosser (2014). Breaking into the Lab: Engineering Progress for Women in Science. NYU Press. ISBN 978-1-4798-0920-2. 
  • Rossiter, Margaret W. (1982). Women Scientists in America: Struggles and Strategies to 1940. Baltimore: The Johns Hopkins University Press. ISBN 0-8018-2509-1. 
  • Rossiter, Margaret W. (1995). Women Scientists in America: Before Affirmative Action 1940–1972. Baltimore: The Johns Hopkins University Press. ISBN 0-8018-4893-8. 
  • Schiebinger, Londa (1989). The Mind Has No Sex? Women in the Origins of Modern Science. Cambridge, Massachusetts: Harvard University Press. ISBN 0-674-57625-X. 
  • Shteir, Ann B. (1996). Cultivating Women, Cultivating Science: Flora's Daughters and Botany in England, 1760 to 1860. Baltimore: The Johns Hopkins University Press. ISBN 0-8018-6175-6. 
  • Warner, Deborah Jean (1981). "Perfect in Her Place". Conspectus of History. 1 (7): 12–22. 

External links

This article is issued from Wikipedia - version of the 12/1/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.