The Chemical Educator, Vol. 8, No. 1, Media Reviews, © 2003 The Chemical Educator

 

Media Review


Transmutations: Alchemy in Art: Selected Works from the Eddleman and Fisher Collections at the Chemical Heritage Foundation. Lawrence M. Principe and Lloyd DeWitt. Chemical Heritage Foundation: Philadelphia, PA, 2002. Illustrations. viii + 40 pp, paperback. 22.2 ´ 26.0 cm. $12.95 special introductory offer. ISBN 0-941901-32-7.

“An Alchemist’s Laboratory” by a Follower of 17th-century Flemish painter David Teniers the Younger, oil on canvas.

Collecting art has become an avocation for a small but increasing group of chemists such as Alfred Bader [1] and Carl Djerassi [2]. Prominent chemist-collectors who have specialized in paintings related to alchemy and chemistry include Chester G. Fisher and Roy T. Eddleman.

Fisher, born in Pittsburgh, PA in 1881, graduated from the University of Pittsburgh in 1902 and founded the Scientific Materials Company, which was renamed the Fisher Scientific Company in 1926. He began collecting paintings and prints of chemists and laboratory scenes in the 1920s, buying entire collections such as the one of alchemical paintings of chemist-collector Sir William Jackson Pope (1870–1939) of Cambridge University, which he acquired in 1940. To give professional chemists a sense of their history Fisher reproduced paintings from his collection in company publications, and he sold prints to decorate chemical laboratories. He died in 1965, and in March 2000 Fisher Scientific International donated his collection to the Chemical Heritage Foundation (CHF) in Philadelphia, PA. According to Fisher’s son, James A. Fisher, “My father thought that the laboratory scientist was not given enough recognition, and this was a way to show the heritage and to honor people in laboratories, those behind the scenes who make so many other things possible” [3].

Roy T. Eddleman, born in 1940 in Kannapolis, NC, is founder and CEO of Spectrum Laboratories of Rancho Domingo, CA. His interest in collecting paintings of early alchemical scenes was sparked by reproductions from the Fisher collection that appeared in the Fisher-owned Eimer & Amend laboratory supply catalogs. He believes that such art “illustrates the difficulty, allure, and potential folly of alchemy as well as the often forgotten, real science aspect of it” [4]. In 2002 the CHF celebrated its 20th anniversary, and that year Eddleman donated to it his collection of nearly 50 paintings and engravings where they joined the Fisher collection.

On June 28, 2002 paintings from the Eddleman and Fisher collections were installed in the Interim Gallery of the CHF’s Eddleman Research Museum. The combined collections of about 100 items constitute one of the world’s largest collections of alchemical art.

Lawrence M. Principe, professor in the Department of the History of Science, Medicine, and Technology and the Department of Chemistry at the Johns Hopkins University, served as the Othmer Fellow at CHF (2001–2002), where he confirmed the artistic veracity of the paintings in the collections. Lloyd DeWitt, a doctoral student in art history at the University of Maryland, served as the Charles C. Price Fellow at CHF (2001–2002), where he researched the provenance and attribution of the paintings and cataloged the collections. Together the two have written this short book—a true fusion of the history of science and the history of art—to celebrate and announce the creation of CHF’s Roy Eddleman Research Museum as well as to provide a glimpse of the paintings and thus a glimpse into the vanished worlds of the alchemists themselves and of the artists who depicted their activities on canvas.

Their book begins with a foreword by CHF President Arnold Thackray, who states,

Here in a group of almost one hundred paintings, one can see the modern chemical sciences struggling to be born…. These Netherlandish genre and allied representations—the great majority from the seventeenth and eighteenth centuries—are simultaneously treasures of art and testaments to the craft-linked origins of modern empirical science. To see them is to glimpse the romance of science, the promise of achievement, and the reality of progress (p vii).

This is followed by Principe and DeWitt’s detailed discussion of alchemy, art, and the 38 paintings and engravings (32 in full color and three full-page) that comprise most of the volume. “Collecting Chemists,” a three-page essay by CHF staff researcher Josh McIlvain, summarizes the lives and collections of Fisher and Eddleman, while a “Further Reading” section lists eight books ranging in date from 1947 to 2002.

Alchemy had been practiced in Europe since the twelfth century, but it enjoyed its heyday in the 17th century. However, as a subject of historical study, it is now enjoying another heyday as historians of science have made great strides in understanding the pseudoscience, its practitioners, and its goals. Therefore alchemy is beginning to be viewed as a serious and complex endeavor central to the development of modern science. The images in the book provide a unique entry point into the exploration of alchemy and its place in 17th-century culture. When accompanied by Principe and DeWitt’s interpretations, they constitute a survey of the history of chemistry, its concepts, its receptions, and its transformations.

The authors summarize the state of alchemy in the 17th century including the transmutation of metals (chrysopoeia), the sulfur and mercury principles, the philosopher’s stone or elixir, Paracelsus and iatrochemistry, the chemical industry of the day, and the technique of distillation and its apparatus (exemplified by a hand-colored title page of Hieronymus Brunschwig’s Liber de arte distillandi). The positive and negative responses of the learned and the general public to “chymistry” are explored along with the marriage of art and alchemy in emblematic images that depict “chymists” at work in their early-modern laboratories replete with apparatus, furnaces, and books.

Genre painting is a style or form of painting showing typical events or settings from the everyday world familiar to the artist and his patrons instead of depicting great stories from history or mythology. It presents generalized types or kinds of events or places rather than specific persons and places. The conventions of Netherlandish genre painting and the motivations behind its creation are discussed, along with the origins of the alchemist genre and Pieter Brueghel, the Elder’s (1525–1569) design for The Alchemist (ca. 1558), which provided the most direct inspiration for the baroque exploitation of the theme by artists (p 11).

The earlier 17th-century paintings, usually depict the alchemist as a deluded soul who has squandered his money in pursuit of the philosopher’s stone, while later 17th-century paintings depict him as a more scholarly and reputable person. These are exemplified by the works of David Teniers, the Younger (1610–1690), Pierre François Basan (1723–1797), Adriaen Pietersz van de Venne (1589–1662), Richard Brakenburg (1650–1702), Thomas Wijck (1616–1677), Cornelis Pietersz Bega (1631–1664), Hendrick Heerschop (1620–1690), and Balthasar van den Bossche (1681–1715).

Paintings from the 18th century, which emphasize the iatrochemical nature of chemistry in which physicians and apothecaries play a prominent role, are exemplified by Franz Christoph Janneck (1703–1761), Justus Juncker (1703–1767), and Louis de Moni (1698–1771). Those by François-Marius Granet (1775–1849), Charles Meer Webb (1830–1895), and Edward Allen Schmidt (born in Berlin, active in England, 1868–1877) show the 19th-century recasting of “alchemy” as a preoccupation with the occult. Two separate pages are devoted to chymical apparatus (cucurbit, alembic, crucibles, retorts, etc.) and examination of paintings with infrared reflectography (IRR).

This fascinating sampling of paintings from more than two centuries surveys the development and changing face of alchemy and its gradual transmutation into modern chemistry. As such, it will be of interest not only to the specialist in the history of science and the history of art but also to the scientist concerned with this crucial era in our chemical heritage and the general reader as well.

References and Notes

1.       Bader, A. Adventures of a Chemist Collector; Weidenfeld and Nicholson: London, England, 1995; reviewed by Kauffman, G. B.; Kauffman, L. M. Angewandte Chem., Int. Ed. Engl. 1996, 35, 784–786.

2.       Kauffman, G. B.; Kauffman, L. M. The Steroid King. The World & I 1992 (July), 7 (7), 312–319.

3.       McIlvain, J. Collecting Chemists. Chem. Heritage 2002 (Summer), 20 (2), 36–37.

4.       Morrisey, S. Alchemical Art. Chem. Eng. News 2002 (July 8), 80 (27), 30–32.

George B. Kauffman

California State University, Fresno, georgek@csufresno.edu

S1430-4171(03)01656-3, 10.1333/s00897030656a

Voyage through Time: Walks of Life to the Nobel Prize. Ahmed Zewail. The American University in Cairo Press: Cairo, New York, 2000. Illustrations. xii + 287 pp, hardcover. 15.5 ´ 23.6 cm. $30.00. ISBN 977-424-677-2.

In this balanced blend of science and biography Zewail describes his passion for work and the betterment of humanity, his love of family and students and his appreciation for their accomplishments, and above all his optimism about and enjoyment of life. His records or memory of detail in places, names, dates, and other minutiae is fantastic. As cases in point, he recalls that on his trip from Cairo to Philadelphia on August 23, 1969 he was charged “LE 27.45 for our excess baggage” (p 51), his first landlady in the City of Brotherly Love was named Mrs. Hurley (p 54), and the total cost of his first car, “a white MG with a stick shift,” was $390 (p 65). His very detailed and honest account gives the reader a complete picture of what his life and work are like, who he is, and how he reached his present position.

The California Institute of Technology’s Ahmed Hassan Zewail was awarded the 1999 Nobel Prize in Chemistry "for his studies of the transition states of chemical reactions using femtosecond spectroscopy.” It was the first time that the prize had been awarded to a single individual since 1994, when George A. Olah received it “for his contributions to carbocation chemistry” [1]. Although the Islamic world represents more than one billion of the Earth’s six billion population, Zewail is only the second Muslim scientist to receive the Nobel prize (Abdus Salam of Pakistan shared the 1979 prize in physics), as Zewail, who is a proud and fervent proponent of the non-Western world, points out.

Zewail was asked several times to write an autobiography or at least an autobiographical summary of his life, but he declined on the ground that “a traditional biography should represent a lifetime of work and experience and much effort and time are needed to do it well” (p 5). During a trip to Cairo in 1997, however, he changed his mind after reading two books—Charles Van Doren’s A History of Knowledge [2] and 1964 Nobel physics laureate Charles Townes’s Making Waves [3]—which provoked him to ask himself the questions that he answers successfully in the book under review here:

How did I acquire knowledge? Why did I become a scientist? What are the forces that have determined the walks of my own life? What are the meanings of faith, destiny, and luck? (p 5).

In his book’s ten chapters Zewail describes the journey from his native Egypt to America—the “Voyage through Time” of the volume’s title. He uses the term “Walks of Life” in the subtitle advisedly to reflect the “apparently random nature of walks and the incidents and surprises that may change such walks to define one’s path through life” (p 6). His book also focuses on time and matter and explains the scientific development and the path to discovery in the femtouniverse.

Zewail’s book highlights three elements—(1) life, (2) science, and (3) vision. He cites six stations on his journey—(1) his childhood on the banks of the Nile (Chapter 1; 17 pp, the shortest chapter); (2) his admission to the Faculty of Science at Alexandria University (Chapter 2; 23 pp); (3) his scholarship in America, “which opened up a whole new world to me” (p 6) (Chapters 3 and 4; 22 and 26 pp); (4) the years of scientific discovery at Caltech (Chapters 5–7; 19, 29, and 18 pp); (5) receipt of the King Faisal International Prize, the first major prize to recognize his group’s achievements at Ridayh, Saudi Arabia, where he met Dema Faham, who ultimately provided him with a new family; and (6) receipt of the Nobel Prize (Chapter 8; 31 pp., the longest chapter). Chapters 9 and 10 present his personal perspective on the prospect for world order through science and on his wish for the two cultures that he shares.

Zewail’s goal, which, in our opinion, he has reached in this book, is

to address the general public of both the developing and developed world, and not specifically the specialist or select intellectuals…. I take the opportunity to describe the path to discovery, to popularize the science and its beauty, and to be explicit about the human dimension. Throughout, I mention some discoveries and contributions and the scientists behind them, and, in the hope of stimulating young people, I also mention the crowning of these achievements by the Nobel and other prizes (pp 7–8).

Zewail’s book goes far beyond the usual province of an autobiography. Proud of his heritage, he has never forgotten his roots (“Egypt planted the seeds properly and America gave me the opportunity” [p 7]), and he reviews the contributions to both science and culture of the Islamic world through the ages. His tale is sprinkled with Arabic terms and expressions (Did you know that “Open, sesame!” is “Iftah, ya-simsim!” ?). Although he became a naturalized American citizen in 1982, he has retained his Egyptian citizenship. He pays tribute to his native land in the book’s dedication:

To Egypt

You lit the beacon of civilization

You deserve a brilliant future

May my voyage light a candle of hope for your youth

Zewail narrates his journey from his birth on February 26, 1946 in the Egyptian delta town of Damanhur, through his happy childhood, education at Alexandria University (B.S. 1967; M.S. 1969) and the University of Pennsylvania (Ph.D. 1974), and postgraduate research at the University of California, Berkeley (1974–1976), to his present position as first Linus Pauling Professor at Caltech. He takes pride in his tenure at the same institution where Pauling carried out his Nobel Prize-winning work on the nature of the chemical bond. Both men were the same age (53), almost to the day, when they won the prize—Pauling for static, stable, and enduring bonds and Zewail for dynamic, ephemeral, and alive bonds. A modest and humble man, he attributes most of his accomplishments to “faith, destiny, serendipity, and intuition” (p 6), the support of Caltech and the agencies that funded his research, the collaboration of his co-workers, and being “in the right place at the right time” (p 7).

Zewail’s book demonstrates that he has the ability, rare for most working scientists, to eschew technical jargon in favor of communicating to a nonscientific audience the goals, techniques, and results of his research by making extensive use of ingenious analogies and metaphors. In his discussion of quantum mechanics, the uncertainty principle, and the wave duality of matter he includes some mathematical equations but relegates them to footnotes. He admits to a “passion for explaining science in simple language” (p 94). Thus he explains the concept of coherence, so crucial to his work, in terms of people walking along a street (pp 92–94); looking at atoms closely as “a lot like inspecting fibers inside cotton bolls” (His native Egypt is famous for its cotton) (p 99); resolution of transition states in chemical reactions as analogous to a “running horse caught in the act by stop-motion photography” (pp 114–116); and flashing a molecule with a femtosecond laser pulse like “the effect of a stroboscope flash or the opening of a camera shutter (p 134).

Zewail’s accomplishment is all the more remarkable in view of the fact that his command of English, which is not his native language, is superb and at times verges on the poetic (A comparison with Polish-born English novelist Joseph Conrad comes readily to mind). He describes the transition state on the potential-energy surface as “that point when the molecule is in between, no longer a reactant and not yet a product, its bonds, like Richard III’s physique, scarce half made up” (p 139). Throughout the book he follows the practice of the best writers and lecturers by repeatedly telling the reader what he is about to discuss, then discussing it, and finally summarizing and recapitulating the material. A careful and organized writer, he usually numbers the various points that he wishes to make before discussing them, thus making his story easy to follow.

In the 1930s chemical reactions were predicted to occur at femtosecond (10–15 s) time scales, and a transition state of activated complexes was postulated. However, because the time scale for such complexes was estimated at less than one picosecond (10–12 s), there was then little hope of observing the transient molecular structures of a chemical reaction.

In his first experiments at Pasadena Zewail demonstrated that short laser pulses could excite simple or even complex molecules into coherent states and that such coherence could be detected during their spontaneous decay. If shot with a sufficiently brief laser pulse, they would release the excess energy in a coordinated manner. This discovery brought Zewail his own ultrafast reaction—tenure in less than two years.

In 1987 Zewail and his team reported the first successful femtosecond-resolved experiment and provided the concepts for probing coherent motion of atoms in reactions. The development of these concepts led to the birth of femtochemistry. They used a high-speed camera based on laser technology with light flashes on the order of tens of femtoseconds to observe molecules in the actual course of reactions and take “pictures” of them during the intermediate transition state. According to Zewail, “femtosecond molecular photography works by breaking up continuous motion into a series of snapshots or frames.”

In 1987 Zewail studied the unimolecular dissociation of cyanogen iodide, ICN ® CN + I, with femtosecond resolution. He “photographed” the reaction at a “shutter speed” fast enough to record about a billion frames showing the formation of the I . . . CN transition state—the first time that such an event had been “seen” in real time. In his own words, “the I atom and CN diatomic molecule dislike each other from time zero and by breaking the I-C bond they released their energy of frustration and ended up with a divorce” (p 138). His “molecular movies” were immediately recognized not only by the scientific press and science writers but by the popular media as well, with headlines such as “Laser captures ‘molecular birth’.”

In Zewail’s study of the dissociation of sodium iodide, which he calls “a molecule with a more complex landscape, …the ‘drosophilia’ of our field [that] turned out to be a paradigm case for the field of femtochemistry” (p 139), NaI ® Na + I, he first showed the resonance behavior, in real time, of a bond converting from covalent to ionic along the reaction coordinate. According to Zewail,

Experimentally we could see for the first time a chemical bond transforming in real time from covalent to ionic, covalent, ionic, covalent, ionic—in this case, by the way, they were in love for about nine or ten cycles before they divorced each other at the end…. The sodium-iodide marriage took about eight picoseconds to fall apart—a long-term commitment on the atomic scale, if not the human one! (p 140).

Zewail then tackled “the happy marriage of two molecules….an even tougher assignment—to observe the simultaneous bond making and bond breaking in a chemical transformation” (p 141), a reaction essential in atmospheric chemistry and the process of combustion, H + CO2 ® CO + OH. Of the reaction’s intermediate transition state he writes, “the tumbling, shivering HOCO molecule is a quantum mechanical Cheshire cat, a short-lived enigma about which little was known” (p 142).

Whenever Zewail mentions a topic, event, or person, he provides a full background summary. Thus we are treated to short histories of Egypt; the close, comfortable, sheltered, and intimate culture in which he was raised (where men were affectionately addressed as “Amm” [uncle]); the culture shock experienced on his exodus to the United States; Alexandria and the Egyptian educational system; chemistry, science, and their histories; and the institutions that he attended. He also includes personal stories of his family, friends, mentors, co-workers, and students, and he places happenings in his story in the context of familiar events (“We didn’t get our first stationary photograph of an atom, whether free or bonded, until the 1980s, more than ten years after Neil Armstrong took the first human steps on the moon on July 20, 1969” [p 105]). He reproduces numerous quotations from the citations for his numerous awards to indicate his work’s significance but without any trace of self-congratulation.

Although we have written previously about Zewail [4, 5], we found much information about him of which we were not aware as well as numerous photos that we had not seen. We learned the details of his first marriage, the birth of the first two of his four children, and his divorce (“What broke my heart the most was having to leave my two daughters, although I would have them on weekends and holidays. I decided from then on that my life was going to be totally devoted to science” [p 96]). We knew of his meeting with Dema Faham and his subsequent second marriage, but we had not realized the astronomical odds against their chance meeting (p 173). Always one to learn from experience, Zewail changed his work habits to include more time with his family (“I don’t work on Sundays, unless it’s really critical” [p 176]).

Through the years Zewail and his students, postdocs, and colleagues have carried out hundreds of pioneering studies of reactions from all branches of chemistry in the field that he almost single-handedly developed.Numerous practical applications have given rise to femtobiology, femtophysics, and other femtosciences. He gives full credit to all those who participated in these efforts, whether they worked with him or not. Considering the range of his multifaceted activities (he is not one to rest on his laurels), we can only marvel that he found time to carry them out, let alone to take time out from a busy schedule to write this meticulously researched book

As a citizen of both the Eastern and Western worlds, Zewail devotes two entire chapters to his hopes and wishes for the solution or at least the amelioration of the crucial problems that the entire world faces today. He makes concrete, often controversial, suggestions for both the “haves” and the “have-nots” with emphasis on courses of action for Egypt and the United States. Not one to agree with the prevalent view that the present state of the world is due to a “clash of civilizations” or a “conflict of religions,” he believes that the current instabilities that threaten our peaceful coexistence are caused by economic and political forces. He recognizes four forces that contribute to the barriers that underdeveloped countries face: illiteracy, an incoherent policy for science and technology, the restrictions on human thought, and fanaticism and the perversion of religion.

To remedy the situation Zewail suggests an acceptance of responsibility in a collaboration between developing and developed countries, and he sets forth “a proposal for partnership.” As responsibilities for developing countries he lists restructuring education and science, creation of centers of excellence, and commitment of national resources, while as responsibilities for developed countries he lists focusing of aid programs, minimization of politics in aid, and partnerships in success—with the post-World War II Marshall Plan and Peace Corps as models.

In his last chapter, “Walks to the Future: My Hope for Egypt and America” Zewail concerns himself with the two countries which he knows best and in which he has spent almost equal amounts of time. He sees Egypt’s main problems as illiteracy, bureaucracy, and unequal implementation of the laws, and, recalling his own experience, he states that “young scientists should find active, productive research and development work in Egypt—they shouldn’t have to leave Egypt to dream big and to engage in frontier science and technological advancement” (pp 223–224). He sees America’s primary problems as insufficient support for education and research, excessive violence, and a limited worldview. He calls for a new vision of dialogue rather than conflict that will not come with the United States’ building a missile shield, pursuing isolationist policies, or selling arms to countries that do not need to make such purchases. In our opinion the United States is going in the diametrically opposite direction, but Zewail ends the chapter, “As I’ve said several times, I am an optimist!” (p 231).

Following a 4-page epilogue summarizing the book and a 5-page list of further readings of books and articles as recent as 2001, a 34-page appendix includes not only Zewail’s Nobel address and other addresses concerned with awards as well as a detailed curriculm vitae but also Zewail’s proposed plan for Egypt’s University of Science and Technology and related Technology Park (“New Initiative for Science and Technology in the Twenty-first Century”). His book also includes a 9 double-column-page index and is graced with an unusually large number of pictures strategically placed at three spots—in all, 82 formal and informal illustrations, 24 in color and 11 in full-page or two-pages.

From our past relations with Zewail, we know that he is a careful writer and proofreader, and the errors in his book are only a few minor typos: “freshmen” for “freshman” (p 34); “Rice” for “Chicago” (University, p 85); “Ernst” for “Ernest” (Rutherford, p 103); “their” for “his or her” (p 106); “squared” for “square” (p 131); and “adaption” for “adoption” (p 178).

We are pleased to give two thumbs up to this fascinating and inspiring volume and to recommend it to chemists, scientists, historians of science, and laypersons concerned with science and how it is practiced as exemplified by the life and career of one of its most prolific far-seeing luminaries.

References and Notes

1.       Olah, G. A. A Life of Magic Chemistry: Autobiographical Reflections of a Nobel Prize Winner; Wiley-Interscience: New York, 2001, for a review see Kauffman; G. B.; Kauffman, L. M. Chem. Educator 2002, 7, 241–243; DOI 10.1007/s00897020587a.

2.       Van Doren, C. A History of Knowledge: Past, Present and Future; Birch Lane Press: New York, NY, 1991.

3.       Townes, C. H. Making Waves; Masters of Modern Physics; Springer-Verlag: New York, NY, 1995.

4.       Kauffman; G. B.; Kauffman, L. M. Observing the Femtoworld of Molecules: The 1999 Nobel Prize in Chemistry. The Chem. Educator 2002, 7, 110–119; DOI 10.1007/s00897020534a.

5.       Kauffman; G. B.; Kauffman, L. M. Ahmed H. Zewail 1946–: Femtochemistry has fundamentally changed our views of chemical reactions…. With the world’s fastest camera available, only imagination sets bounds for new problems to tackle. In Luminaries of the Chemical Sciences: Chronicles of Chemistry; Supplement to ACS Publications, American Chemical Society: Washington, DC, 2002; pp 106, 109.

George B. Kauffman and Laurie M. Kauffman

California State University, Fresno, georgek@csufresno.edu

S1430-4171(03)01657-2, 10.1333/s00897030657a

Robert Burns Woodward: Architect and Artist in the World of Molecules. Otto Theodor Benfey and Peter J. T. Morris, Editors. Chemical Heritage Foundation: Philadelphia, PA, 2001. Illustrations. xxvii + 470 pp, hardcover. 22.5 ´ 28.6 cm. $45.00. ISBN 0-941901-25-4.

Robert Burns Woodward, born on April 10, 1917 in Boston, Massachusetts, was a precocious and lively child who received his primary and secondary education in the public schools of Quincy, the Boston suburb where he lived [1]. At the age of eight, he received a chemistry set, which he used to carry out elaborate experiments in his basement laboratory. Although a nonconformist who disdained school routine and discipline, he graduated from high school in 1933 with an outstanding scholastic record. At age 16 he entered the Massachusetts Institute of Technology with advanced undergraduate standing, and at age 17 he published his first paper. He received his B.S. degree in 1936 at age 19 and Ph.D. degree in 1937 at the age of 20.

Robert Burns Woodward engrossed in the world of molecules. Here he is building a polypeptide ball, stick, and spring model—a generic model of the molecules of various natural fibers such as wool and silk (1947).

In 1937 Woodward began work at Harvard University, where he remained for the rest of his life. He became assistant professor (1950–1953), Morris Loeb professor of chemistry (1953–1960), and finally Donner professor of science (1960–1979), a position that freed him from all formal teaching duties and allowed him to devote all his time to research.

During his long tenure at Harvard, Woodward attracted about 400 junior collaborators, mostly at the postdoctoral level, from all over the world. He also stimulated and to some extent directed research groups in various industrial companies for which he was a consultant. He had a long and close association with Swiss chemists. Although he did not accept offers to move to the Eidgenössiche Technische Hochschule (Federal Institute of Technology, ETH) in Zürich, in 1963 the pharmaceutical firm of Ciba, Ltd. established alongside its Basel headquarters the Woodward Research Institute, where he directed part-time research workers on topics essentially of his own choosing. His extraordinary memory enabled him to direct simultaneously and in detail the work of many individuals. By the time of his death of a massive heart attack on July 8, 1979 at the age of 62, he was generally regarded as the greatest figure in American organic chemistry.

The total synthesis in the laboratory of an extremely complicated natural product requires every atom and every group of atoms to be placed in their proper positions. Woodward early made this enormous challenge his primary research goal. Organic synthesis is not only an exact science but also a fine art, of which he was a consummate master who deliberately chose synthetic problems that were generally regarded as virtually impossible and that had practical applications.

Woodward developed syntheses of molecules with many asymmetric centers, using new reactions and reagents and newly revealed mechanisms from physical organic chemistry to control the stereochemical courses of reactions. Such syntheses not only served to confirm the correctness of a structure and advanced our knowledge of the reactions and the chemical properties of the molecule, but in the case of substances of practical utility, they provided possibilities for synthetic products that are cheaper or more easily accessible than the naturally occurring products. Also, if substances with medical uses but with undesirable side effects can be synthesized, the structure can be modified to reduce or eliminate such effects. Although Woodward would always flexibly follow up each novel or unexpected result, serendipity played little part in his syntheses, every step of which was meticulously planned in advance in the light of available knowledge of reaction mechanisms derived from his wide reading of the literature.

Woodward made outstanding contributions spanning almost the entire range of theoretical and experimental organic chemistry, and he was universally acclaimed as the world's leading exponent of organic chemistry. He used and inspired others to use new instruments and physicochemical techniques to understand and synthesize complex natural products. He was awarded the 1965 Nobel chemistry prize in chemistry “for his outstanding achievements in the art of organic synthesis.”

The Chemical Heritage Foundation (CHF) became involved with Woodward when chemical entrepreneur and art collector Alfred Bader of the Sigma-Aldrich Corporation, who had attended several of Woodward’s courses at Harvard, visited the foundation (then the Beckman Center for the History of Chemistry) after Woodward’s death [2]. Bader suggested that the center honor Woodward’s memory and make him known to younger generations. On April 10, 1992, the 75th anniversary of Woodward’s birth, the center held an international symposium launching a 12-panel traveling exhibit titled “R. B. Woodward and the Art of Organic Synthesis” and issued a paperbound booklet with the same title [3], designed as a teaching aid for high schools and colleges, to accompany it.

Otto Theodor Benfey, Dana Professor Emeritus of Chemistry and History of Science at Guilford College, Guilford, NC, CHF editor-at-large, and coeditor of the booklet, and Peter J. T. Morris, senior curator in charge of the experimental chemistry collections at the Science Museum, London and visiting lecturer at the Chemistry Department of Imperial College of Science, Technology and Medicine, London, have joined forces to produce the book under review here as a volume in “The History of Chemical Sciences Series.” The series documents the individuals, ideas, institutions, and innovations that have created the modern chemical sciences, bringing the best historical scholarship on these topics to a wide audience and providing a balanced view of the contributions that the chemical sciences and technologies have made to modern science and culture.

Benfey and Morris intended this large, opulently illustrated book with generously wide margins, a decade in the making, “to bring back to life a legendary figure, the star of twentieth-century organic chemistry.” They hope that “readers of this volume, particularly younger ones—those of college age—will gain a sense of what made him great so that they in turn will be stimulated and inspired by his example to carry out the tasks presented to them in their time.” In my opinion they have admirably succeeded in attaining their goal.

The five-part volume is prefaced by an essay on art and creativity in chemistry in general and on Woodward’s research in particular by his daughter, artist Crystal Woodward (12 pp). Part I, “Robert Burns Woodward and His Era,” consists of three chapters—a biographical introduction by Peter J. T. Morris and Mary Ellen Bowden (8 pp) as well as reminiscences of Woodward by junior-high and high-school friend Robert C. Putnam (2 pp) and longtime Harvard colleague and friend Frank H. Westheimer (8 pp).

Part II, “Robert Burns Woodward as Collaborator,” consists of a single chapter, “RBW, Vitamin B12, and the Harvard-ETH Collaboration” (16 pp) by Albert Eschenmoser of the ETH, which was originally intended as a brief foreword to the volume but metamorphosed into a major chapter [4]. Woodward and Eschenmoser each attempted to synthesize half of the vitamin molecule and both would then try to combine the two. The project involved about a hundred scientists from 20 countries and led to the celebrated Woodward–Hoffmann rules.

Part III, “Selected Papers of Robert Burns Woodward,” selected and with detailed, meticulously researched commentaries by Peter Morris that firmly place them in historical context, comprises almost four-fifths of the book (371 pp) and is divided into 11 sections. Morris chose the items to be reprinted that emphasized not only Woodward’s contribution to organic synthesis but also to physical organic chemistry, where he explored novel instruments as they became available to learn ways in which they could aid the practicing organic chemist, as well as to show how he collaborated successfully with other chemical luminaries.

·  The papers on the Woodward rules (3 papers, 1941–1942), formulated when he was only 23–24, show his reduction of the ultraviolet spectra of many organic compounds (particularly steroids) to a few numerical relationships and demonstrate his remarkable powers of analysis and passion for scientific order. They also show how he adopted any seemingly relevant new technique that might improve his grasp of the chemistry of natural products.

·  Two papers on his wartime (1944–1945) synthesis of quinine (more accurately quinotoxine) with William von Eggers Doering show Woodward’s first total synthesis. Although carried out with conventional methods, it was a remarkable feat for a 27-year old, and it first brought him to public attention.

·  Steroids were one of the most active research areas in the 1940s and early 1950s because of their potential therapeutic value, but their synthesis posed extremely difficult problems. Woodward’s total synthesis of cholesterol and cortisone (one paper, 1952) established him as one of the world’s most eminent synthetic organic chemists.

·          Woodward had a remarkable ability to illuminate problems outside his usual areas of interest. His elucidation of the structure of ferrocene (two papers, 1952), his only major contribution to organometallic chemistry, was related to his interest in aromaticity and Diels–Alder reactions. It initiated a revolution in organotransition metal chemistry and led to the award of the 1973 Nobel chemistry prize to Geoffrey Wilkinson and Ernst Otto Fischer “for their pioneering work, performed independently, on the chemistry of the organometallic so called sandwich compounds.”

·          Woodward’s investigation of the toxic alkaloid strychnine (3 papers, 1947, 1948, and 1963), spanning his elucidation of its seven-ring complex structure in 1948 and its synthesis in 1954, allowed him to triumph over the reigning master of organic chemistry, Sir Robert Robinson, who had received the 1947 Nobel chemistry prize “for his investigations on plant products of biological importance, especially the alkaloids.” Woodward’s structure determination, described as the “Mount Everest of structural organic chemistry,” and his synthesis, described as “an achievement which but a few years ago still seemed altogether outside the range of the attainable,” made him the greatest organic chemist of his generation.

·  Woodward’s synthesis of the alkaloid reserpine (one paper, 1958), involving six chiral centers, was his favorite synthesis. For the first time it maintained stereoselectivity throughout a long synthetic sequence and in an especially elegant manner. In the Woodward canon it is equaled only by his synthesis of vitamin B12.

·  Woodward’s octant rule (one paper, 1961), which correlates optical rotatory dispersion with stereochemical structure, emerged from his discussions with Carl Djerassi, William Moffitt, and Albert Moscowitz, who, with William Klyne, were the authors of this paper, which illustrates his talent for creating novel solutions and generalizations in discussions with colleagues.

·  Following his synthesis of the alkaloids strychnine and reserpine, Woodward looked to a new field and successfully tackled the synthesis of the plant pigment chlorophyll (one paper, 1961), which had eluded Hans Fischer, who had received the 1930 Nobel chemistry prize “for his researches into the constitution of chlorophyll.”

·  Woodward chose his synthesis of the major antibiotic cephalosporin C, which links his wartime work on penicillin with his later research on the synthetic penems, as the topic for his 1965 Nobel chemistry prize lecture, which is reproduced here. This synthesis of a delicate molecule involved a combination of both little known and entirely novel reactions.

·  Woodward and Eschenmoser’s multistep total synthesis of vitamin B12,, one of the greatest achievements of modern organic chemistry, combined the intricate chemistry associated with the synthesis of chlorophyll a with stereochemical problems even more difficult than those encountered with reserpine. The “western half” paper (1968) was chosen for reproduction here because of its connection to the Woodward–Hoffmann rules and because it was the most “Woodwardian” of the numerous B12 articles.

·  The vitamin B12 work led directly to the Woodward–Hoffmann rules for correlating ultraviolet absorption with molecular structure (six papers, 1942, 1959, 1965, and 1967), which combined Woodward’s interest in the Diels–Alder reaction, his dogged investigation of reaction failure, and his ability to produce powerful generalizations. It provided an effective theoretical tool for explaining chemical transformations. Two years after Woodward’s death Roald Hoffmann received the 1981 Nobel chemistry prize for this work.

Part IV, “The Cope Award Lecture” (August 28, 1973), perhaps “the only extended autobiographical lecture [Woodward] ever gave,” was never published although it was once available on an American Chemical Society audiotape. It is reproduced as transcribed exactly from a handwritten copy (two pages shown on pp 428–429) supplied by Roald Hoffmann. Benfey provides a three-page introduction and 13 pages of notes explaining chemical topics and identifying the persons mentioned in the lecture.

Part V, “Woodward’s Publications and a Chronology of His Life,” includes a complete bibliography of his 210 publications from 1934 (when he was 17) to 1990 (the last of his posthumously published articles) as a guide to his “creative endeavors” and “artistic achievements.”

Most of the chapters, which are cross-referenced, include bibliographies and numerous structural formulas and reaction schemes, especially, of course, in the reprinted Woodward articles. The book abounds with 40 illustrations, some full-page, such as familiar and unfamiliar photos (both formal and informal), handwritten sketches of reactions schemes (both on paper and on the blackboard), a cartoon, a telegram, an autographed green leaf announcing the completion of the chlorophyll synthesis, and an autographed banquet menu for the 1976 celebration of the completed B12 synthesis. A name index (four triple-column pages), which includes only names in the original text pages but not in the reproduced articles, facilitates location of material.

Considering the amount of material in this inexpensive book, it is a fantastic bargain, and I recommend it to historians of chemistry, chemical educators, and chemists, especially organic chemists, as well as to students contemplating a career in chemistry.

References and Notes

1.       Kauffman, G. B. Robert Burns Woodward (1917–1979). In American National Biography; Garraty, J. A.; Carnes, M. C., Eds.; Oxford University Press: New York, NY, 1999, Volume 23, pp 829–832.

2.       In Spring 2001 Bader donated his notes from Woodward’s “Chemistry 203: Chemistry of Natural Products,” a course that he took at Harvard in 1948–1949, to the CHF (Benfey, O. T. Chemist Collector Bader Donates RBW Notebook. Chem. Heritage Winter 2001/2, 19 (4), 16).

3.       Bowden, M. E.; Benfey, O. T. Robert Burns Woodward and the Art of Organic Synthesis; Publication No. 9, Beckman Center for the History of Chemistry: Philadelphia, PA, 1992 (J. Chem. Educ. 1992, 69, 469; Beckman Center News 1992, 9 (9), B-1; for a review see Kauffman; G. B. J. Chem. Educ. 1993, 70, A52.

4.       Eschenmoser’s essay was excerpted as an article of the same name in Chem. Heritage Fall 2001, 19 (3), 12–13, 46–49.

George B. Kauffman

California State University, Fresno, georgek@csufresno.edu

S1430-4171(03)01658-1, 10.1333/s00897030658a

Linus Pauling: Selected Scientific Papers (in 2 volumes). Barclay Kamb, Linda Pauling Kamb, Peter Jeffress Pauling, Alexander Kamb, and Linus Pauling, Jr., Editors. World Scientific Publishing: River Edge, NJ; London; Singapore; Hong Kong, 2001. Illustrations. hardcover. 20.0 ´ 26.5 cm. Set xxxiii + 1573 pp, $167.00; £114. ISBN 981-02-2784-1; Vol. 1 xi + 840 pp, $98.00; £67. ISBN 981-02-2939-9; Vol. 2 xii + 730 pp, $98.00; £67. ISBN 981-02-2942-9.

Almost everything about Linus Pauling was larger than life. An internationally acclaimed scientist, educator, humanitarian, and political activist, this controversial chemist was characterized as one of "the twenty greatest scientists of all time, on a par with Newton, Darwin, and Einstein;" one of the two greatest scientists of the 20th century, the other being Einstein; and the greatest chemist since Antoine-Laurent Lavoisier. He remains the only person to have received two unshared Nobel prizes (Chemistry, 1954; Peace, 1962). His magnum opus, The Nature of the Chemical Bond is one of the most influential and frequently cited scientific books of our century [1]. As the author of almost two dozen articles, interviews, and reviews of books and videos about Pauling, including two in this journal [2, 3], I found the book under review here to be of more than passing interest.

Figure 1. Linus Pauling in 1923, the year after his arrival at Caltech for graduate work.

Shortly before Pauling’s death on August 19, 1994 the World Scientific Publishing Company (WSP) contacted him and proposed that it reprint in book format a selection of his scientific writings, a proposal to which he agreed. His four children agreed to serve as editors. The selection of appropriate articles from Pauling’s more than a thousand publications [4] was a difficult one, and he is reported to have suggested, “The selection is easy—print them all.” During a period of mourning the project languished, but Barclay Kamb, Caltech professor of geology and geophysics emeritus and Pauling’s former student, colleague, and son-in-law, accepted the position of Editor-in-Chief. Because Pauling contributed to so many different scientific fields, a group of 10 specialists were recruited as advisors for the selection of papers and in some cases to aid in writing the introductions to the parts comprising the volume.

After many delays the massive collection appeared as Volume 10 of the “World Scientific Series in 20th Century Chemistry,” a series that has included selected articles by such Nobel chemistry laureates as Glenn T. Seaborg, Kurt Wüttrich, Kenichi Fukui, and Ahmed H. Zewail as well as by other luminaries. This anthology of Pauling’s most significant and influential scientific papers, selected from his about 850 scientific papers and books to convey his imaginative style of thinking and the considerable range of topics that he tackled,bears witness to his monumental contributions and their impact on numerous disciplines. It allows the reader to view in one place and in exact facsimile many of the original works on which his fame as a scientist rests.

Following a Foreword by Ahmed H. Zewail, the Linus Pauling Professor of Chemistry at the California Institute of Technology, a Preface by Peter Pauling, and a General Introduction by Barclay Kamb, the collection consists of four major subject groupings (Parts I and II in Volume 1; “Physical Sciences” and Parts III and IV in Volume II, “Biomolecular Sciences”). The 144 papers are subdivided into 17 chapters, each of which begins with a table of contents listing the papers in that chapter.

Figure 2. Linus Pauling at the blackboard.

Only four selections from Pauling’s books are included because it would be impractical to include more within the space limitation and since the primary objective was to document his science as it emerged and first impacted the scientific world rather than as it appeared in later retrospectives. Two selections are from The Nature of the Chemical Bond (1939); one from Pauling and E. Bright Wilson, Jr.’s Introduction to Quantum Mechanics (1935), which has been used to train generations of quantum chemists and is said to be the longest lived quantum mechanics text in terms of time in print; and one from Ewan Cameron and Pauling’s Vitamin C and Cancer (1979).

Figure 3. Linus Pauling and his wife, Ava Helen, dancing at the ball following presentation of the Nobel Prize in Chemistry, in Stockholm, 1954. Ava Helen was Linus’ chief colleague, he often said.

The following summary gives an idea of the book’s contents as well as the scope and breadth of Pauling’s scientific contributions.

·  Part I, “The Chemical Bond,” edited and introduced by son Peter Pauling (University College, London), contains papers on the nature of the chemical bond, the subject for which Pauling was first and foremost acclaimed in the scientific world and for which he was awarded the Nobel Prize in Chemistry.

Chapter 1, “Covalent Bonding, Resonance, and Bond Orbital Hybridization” (22 papers, 249 pp, the longest chapter)

Chapter 2, “Ionic Bonding, Partial Ionic Character, and Electronegativity” (5 papers, 88 pp)

Chapter 3, “Metallic Bonding” (7 papers, 66 pp)

Chapter 4, “Hydrogen Bonding” (4 papers, 33 pp)

·  Part II, “Crystal and Molecular Structure and Properties,” edited and introduced by Barclay Kamb, contains papers on the atomic structure of molecules and crystals, which provided much of the observational evidence on which Pauling’s concepts of the chemical bond in Part I were based. It also contains papers in which the principles of quantum mechanics, which provided the theoretical basis for many of his concepts in Part I, were applied by him to understand the physical properties of molecules and crystals. Those in Part I concentrate on conceptual and theoretical aspects of chemical bonding, while those in Part II concentrate on observational and empirical evidence and on properties as distinct from chemical bonding.

Chapter 5, “Ionic Crystals and X-Ray Diffraction” (9 papers, 92 pp)

Chapter 6, “Covalent, Intermetallic, and Molecular Crystals” (10 papers, 71 pp)

Chapter 7, “Molecules in the Gas Phase and Electron Diffraction” (5 papers, 48 pp)

Chapter 8, “Molecular Properties Analyzed by Quantum Mechanics” (8 papers, 100 pp)

Chapter 9, “Entropy and Molecular Rotation in Crystals and Liquids” (4 papers, 31 pp)

Chapter 10, “Nuclear Structure; Superconductivity; Quasicrystals” (7 papers, 36 pp)

·  Part III, “Biological Macromolules,” edited and introduced by grandson Alexander (“Sasha”) Kamb (President & CEO, Arcaris, Inc., Salt Lake City, UT), contains papers that exemplify a major new thrust in Pauling’s research that he began to pursue in the late 1930s and that culminated in the 1950s, his investigations of the structure and function of large molecules of biological significance, especially proteins. The concepts and observational methods of Parts I and II, applied to small organic molecules that are the building blocks for the large molecules of biological importance, led to the papers in Part III.

Chapter 11, “Hemoglobin: Oxygen Bonding and Magnetic Properties” (6 papers, 40 pp)

Chapter 12, “Antibodies: Structure and Function” (7 papers, 73 pp)

Chapter 13, “The Alpha Helix and the Structure of Proteins” (17 papers, 128 pp)

Chapter 14, “Molecular Biology: The Role of Large Molecules in Life and Evolution” (10 papers, 132 pp)

·  Part IV, “Health and Medicine,” edited and introduced by son Linus Pauling, Jr., M.D. (Linus Pauling Institute of Science & Medicine, Oregon State University, Corvallis, OR), contains papers on biomedical subjects, starting with his frequently cited discovery of molecular disease in 1949, which biologists and medical scientists consider to be among his greatest and most influential scientific contributions. It also includes papers on nutritional medicine, which he christened “orthomolecular medicine” and which was the focus of much of the research carried out during his later years.

Chapter 15, “Molecular Disease” (5 papers, 69 pp)

Chapter 16, “Physiological Chemistry, Effects of Radiation, and Health Hazards” (7 papers, 53 pp)

Chapter 17, “Orthomolecular Medicine” (11 papers, 101 pp)

·  Part V, “Summary of Linus Pauling’s Life and Scientific Work” includes Chapter 18, “Biographical Memoir,” by Jack D. Dunitz (24 pp, the shortest chapter), which summarizes Pauling’s contributions to science and assesses their significance and impact [5]. It provides a frame of reference in which the original papers in the collection can be studied and evaluated, and it places Pauling’s scientific work in the broader context of his life as a whole.

Appendix I (2 pp) presents a conversion table between the numbers of the papers in the collection, those used in Dunitz’s memoir (Chapter 18), and those used in Appendix III. Appendix II (6 pp) lists in an abbreviated citation format all papers, articles, and book chapters reproduced in the collection. Appendix III (60 pp) lists all Pauling’s scientific papers, from which the reproduced papers were selected, classified into 17 groups corresponding in subject matter to the 17 chapters (Selected papers that are included in the volume are designated with asterisks), followed by three additional groups (Biographical Memoirs, Memorials, Tributes; Book Reviews; and General, Retrospective, Prospective, and Miscellaneous). The collection of 50 photographs (17 full-page and 7 in color), arranged in a sequence approximating the sequence of chapters in the book, were selected and captioned by daughter Linda Pauling Kamb (President, LCProgeny, Inc., Pasadena, CA)

Figure 4. Linus Pauling and family, en route to Oslo for the Nobel Peace Prize ceremony, 1963. Left to right: Ava Helen Pauling (1903-1981), Linus Pauling (1901-1994), Linda Pauling Kamb (b. 1932), Barclay Kamb, Lucy Pauling (Crellin’s wife), (Edward) Crellin Pauling (1937-1997), and Linus Pauling, Jr.(b. 1925). Peter Jeffress Pauling (b. 1931), one of the editors of the book, is not in the picture.

Figure 5. Linus Pauling with Ahmed H. Zezail, the Linus Pauling Professor of Chemistry and Physics at the California Institute of Technology, 1990. Zewail was awarded the Nobel Prize in Chemistry in 1999 for pioneering work in femtochemistry.

In the more than eight decades that have elapsed since Pauling’s earliest paper some members of a generation of younger scientists may have forgotten the source of many of his ideas, knowledge, and methods that are taken for granted today. This comprehensive collection of his most significant contributions, printed on high quality paper, should prove to be a valuable addition to the education of these scientists. I also recommend it to chemists, biologists, and scientists in general who wish to understand the roots of important concepts in modern science, the foundations of which were erected by the 20th century’s most influentual chemist.

References and Notes

1.       Pauling, L. The Nature of the Chemical Bond and the Structure of Molecules and Crystals: An Introduction to Modern Structural Chemistry; Cornell University Press: Ithaca, NY, 1939; 2nd ed., 1940; 3rd ed., 1960; The Chemical Bond: A Brief Introduction to Modern Structural Chemistry (an abridged version of the 3rd ed. intended especially for use by students), 1967. In the period 1955–1983 the book was cited no less than 16,027 times.

2.       Linus Pauling on Peace: A Scientist Speaks Out on Humanism and World Survival; Marinacci, B.; Krishnamurthy, R., Eds.; Rising Star Press: Los Altos, CA, 1998; for a review see Kauffman, G. B.; Kauffman, L. M. Chem. Educator 1999, 4, 198–199; DOI 10.1007/s00897990332a.

3.       Mead, C.; Hager, T. Linus Pauling: Scientist and Peacemaker; Oregon State University Press: Corvallis, OR, 2001; for a review see Kauffman, G. B.; Kauffman, L. M. Chem. Educator 2002, 7, 314–319; DOI 10.1007/s00897020606a.

4.       Including his works on social, humanitarian and political subjects, Pauling’s oeuvre totals some 1200 items.

5.       Dunitz, J. D. Linus Carl Pauling 28 February 1901- 19 August 1994. Biographical Memoirs of Fellows of the Royal Society 1996, 317-338.

George B. Kauffman

California State University, Fresno, georgek@csufresno.edu

S1430-4171(03)01659-0, 10.1333/s00897020659a

The Biographical Dictionary of Women in Science: Pioneering Lives from Ancient Times to the Mid-20th Century. Marilyn Ogilvie and Joy Harvey, Editors. 2 volumes. Routledge: New York and London, 2000. xxxviii + 1499 pp, hardcover. 22.3 ´ 28.6 cm. $265.00. ISBN 0-415-92038-8.

Although women have been involved in science for thousands of years, until recently they have been overlooked by historians and the general public, except for a few luminaries such as Marie Curie, Irène Joliot-Curie, Lise Meitner, Barbara McClintock, Rachel Carson, and Margaret Mead, who were regarded as “exceptions” to the general rule that women made few contributions to science. As Margaret W. Rossiter points out in her introduction to the two hefty volumes of this biographical dictionary, even as late as the 1970s the monumental Dictionary of Scientific Biography included only twenty-five women [1]. Yet, as the feminist movement in the United States progressed, this neglect of women scientists has been increasingly rectified as a number of biographies and biographical dictionaries have appeared.

In 1986 Marilyn Bailey Ogilvie, then at the Oklahoma Baptist University and now Professor and Curator of the History of Science at the University of Oklahoma, published a 272-page biographical dictionary [2], one of the earliest of its kind, that profiled 186 women from the United States, Canada, and western Europe, whose work was representative of the participation of women in the science of their time and culture. With the aid of coeditor Joy Harvey, who has taught at the Department of the History of Science at Harvard University, and 26 eminent contributing authors, all but one of whom are female, Ogilvie has expanded her earlier work by a factor of about twenty. Many of the entries were written by the two editors. The resulting authoritative dictionary profiles about 2500 women scientists [3] alphabetically arranged from Canadian physician and medical historian Maude Elizabeth Seymour Abbott (1869–1940) to Austrian meteorologist Rely Zlatarovic (fl. 1920). The subjects profiled contributed to a variety of fields—alchemy, anatomy, anthropology, archeology, astronomy, bacteriology, biology, botany, chemistry, engineering, genetics, geology, home economics, mathematics, medicine, midwifery, mineralogy, nutrition, paleontology, physics, psychiatry, psychology, surgery, and zoology—to list only a few. Since areas interpreted liberally as science-related fields are a basis for inclusion, entries will be found for many subjects not included in other biographical dictionaries.

Because the scope of science and what constitutes science and who practiced it has changed through time [4], the editors’ selection of subjects changed with time. While from earlier periods the dictionary includes royalty (for example, Hatshepsut of Egypt [d. 1468 B.C.], who sent a botanical expedition in search of medicinal plants), Anna Sophia of Denmark [fl. 1685], known for her botanic garden; and Christina of Sweden [1626–1689], who was a patron of Descartes in Stockholm and established the Accademia Reale in Rome) and nuns (Hildegard of Bingen [1098–1179 or 1180], cosmologist, natural philosopher, and writer on medicine), women from later periods were held to a higher standard for inclusion. Most of the women scientists profiled made their own actual contributions, while a few, justly or unjustly, are better known as assistants to their husbands or other relatives. The list of women classified as chemists comprises 162 names (Marie Anne Pierrette Paulze Lavoisier [1758–1836] is classified as an editor and illustrator, while Claudine Poullet Picardet Guyton de Morveau [ca. 1770–ca. 1820] is classified as a translator). Also included is French alchemist Perrenelle Flammel (d. 1397), whose husband Nicholas Flammel has become known to millions of readers who otherwise would never have heard of him because of his inclusion in J. K. Rowling’s bestseller Harry Potter and the Philosopher’s Stone [5].

Unlike many of the other bio-bibliographical dictionaries devoted to specific fields that have appeared in the past two decades, Olgivie and Harvey’s dictionary is the first reference source to include women in all time periods, most parts of the world, and various scientific fields [6]. The coverage exceeds that of dictionaries that limited themselves to well known or award-winning women, for it includes women who have remained obscure but who have been “the mainstays of their government bureaus, colleges, nonprofit institutions, local or regional groups, and specialties.” Furthermore, although Americans, Canadians, and western Europeans predominate, the coverage is truly global and features Russians, eastern Europeans, Turks, Filipinos, Latin Americans, and Asians, about whom little information is available in English. The coverage extends from antiquity to the recent past. Some women who were born before 1910 are only recently deceased or still alive.

Each signed entry, for which complete data are known, begins with a brief biographical summary of education and work history (in italics), followed by a description of most significant achievements, a list of most important publications (primary sources) and references to biographical works (classified into secondary and standard sources) [7] to consult for further information. The entries, which are fully cross-referenced and indexed, range in length from a sentence or two to several pages, some depending on the subject’s importance and some on the amount of data available.

Despite a number of typographical errors or omission of diacritical marks in proper names, The Biographical Dictionary of Women in Science, which is printed on high-quality glossy paper,is a most comprehensive, authoritative, and invaluable first reference source for the general public, students, and scholars in search of information on individual women who contributed to the sciences or science-related fields. Thus it belongs in every academic, high school, and public library. Because of its classified lists, it will also prove useful for those pursuing particular topics or seeking to study subgroups, for example, British women botanists or Russian women physicians. Furthermore, it will interest readers who merely want to browse or increase their knowledge of a neglected area. It will also be an inspiration to scientists, especially younger women, who wish to learn about their forebears and the prejudices and obstacles that they had to overcome in their struggles to contribute to our body of scientific knowledge or to achieve recognition.

References and Notes

1.       Dictionary of Scientific Biography; Gillispie, C. C., Ed.-in-Chief; Charles Scribner’s Sons: New York, 1970–1980; 16 volumes including an index volume.

2.       Ogilvie, M. B. Women in Science, Antiquity through the Nineteenth Century: A Biographical Dictionary with Annotated Bibliography; MIT Press: Cambridge, MA, 1986. Ogilvie is also Roy Porter’s coeditor of The Biographical Dictionary of Scientists, 3rd ed.; Oxford University Press: New York, 2000; reviewed by G. B. Kauffman, Chem. Educator 2002, 7, 238–239; DOI 10.1007/s00897020585a.

3.       The alphabetical list of entries (15 triple-column pages), which appears in both volumes, includes 2580 names but contains some duplicates because some subjects are listed by both married and maiden names.

4.       The word “scientist” was not coined—by William Whewell—until the 19th century.

5.       For some unknown reason the American edition was retitled Harry Potter and the Sorcerer’s Stone.

6.       Volume 2 contains lists of scientists classified by occupation (19 triple-column pages), by time period (23 triple-column pages), and by country (16 triple-column pages) as well as a subject index (23 triple-column pages) in which dictionary entries are indicated in boldface.

7.       A list of 200 standard sources and additional references (six double-column pages) from 1775 to 1999 is provided in both Volumes 1 and 2.

George B. Kauffman

California State University, Fresno, georgek@csufresno.edu

S1430-4171(03)01660-1, 10.1333/s00897030660a

Mendeleyev’s Dream: The Quest for the Elements. Paul Strathern. Hamish Hamilton, Penguin Books: London, 2000; £12.99; ISBN 0-241-14065-X; Thomas Dunne Books, St. Martin’s Press: New York, 2000 (first U.S. edition, 2001); $23.95; ISBN 0-312-26204-3. Nine Illustrations. x + 309 pp, hardcover. 14.5  ´ 21.4 cm.

According to the late Harlow Shapley—an American astronomer, not a chemist—the periodic table

is probably the most compact and meaningful compilation of knowledge that man has yet devised. The periodic table does for matter what the geological age table does for cosmic time. Its history is the story of man’s great conquests in the microcosmos (Shapley, H. Of Stars and Men: The Human Response to an Expanding Universe; Beacon Press: Boston, MA, 1958; pp 38–39).

I think that Paul Strathern, the award-winning author of several novels and a number of books on philosophy and science, who lectures on these topics at Kingston University, would concur with Shapley’s evaluation, for he calls Mendeleev’s discovery “the culmination of a two-and-a-half-thousand-year epic: a wayward parable of human aspiration” (p 294). In his opinion,

With the Periodic Table chemistry came of age. Like the axioms of geometry, Newtonian physics and Darwinian biology, chemistry now had a central idea upon which an entire new range of science could be built. Mendeleyev had classified the building blocks of the universe (p 292).

The title—but not the subtitle—of Strathern’s highly acclaimed book is a misnomer, for the scope of the work is far deeper and broader than its title suggests. He probably chose the title to parallel those of ten of his other scientific books, which link the name of a scientist with that of his or her most significant contribution—for example, Pythagoras and His Theorem, Galileo and the Solar System, Newton and Gravity, Curie and Radioactivity, and Hawking and Black Holes.

Instead Strathern’s book is a short, breezy, and lighthearted history of chemistry for a popular audience, with an emphasis on the development of the concept of the chemical element and the systems proposed to classify and explain the properties of the elements. Although the focus is a humanistic one concentrating on the persons involved in his tale, their ideas and proposals are not neglected. Of course, considering the book’s brevity, it is not surprising that a number of topics are not considered.

Strathern’s account is framed by the figure of Dmitrii Ivanovich Mendeleev (1834–1907) (Strathern prefers the spelling “Mendeleyev” from among the different variants used to transliterate the Cyrillic script), Professor of Chemistry at St. Petersburg University. The book’s Prologue portrays this prophet from Siberia on February 17, 1869 obsessively seeking a pattern among the 63 known elements by laying out cards on which he had written their symbols, atomic weights, and properties. Overcome with exhaustion, he falls asleep at his desk and has a dream. In the ensuing 12 chapters we are treated to an entertaining account of the development of chemistry from its earliest beginnings to Mendeleev’s time. Mendeleev returns only in the final two chapters to complete Strathern’s tale with a description of the periodic table, the culminating template on which modern chemistry is based and which marked chemistry’s coming of age as a mature science (Less than 20 pages or about six percent of the book is actually devoted to Mendeleev).

This fascinating account begins with the ancient Greeks—Thales of Miletus (ca. 625–547 B.C.E.), the first philosopher-scientist, who postulated that the world developed from one element—water (Chapter 1). We meet other Greek natural philosophers such as Anaximenes of Miletus (fl. ca. 545 B.C.E.), who proposed that the fundamental element was air, Heraclitus of Ephesus (fl. 500 B.C.E.), who proposed that the fundamental element was fire, and Empedocles of Acgras (ca. 492–ca. 432 B.C.), who added earth to the trio of fundamental elements, resulting in what became known as the four-element theory. We then encounter the Greek atomists—Leucippus (fl. 5th century B.C.E.) and Democritus (fl. late 5th century B.C.E.)—and the Athenian philosophers Socrates (ca. 470–399 B.C.E.), Plato (427–348/347 B.C.E.), and Aristotle (384–322 B.C.E.).

In his discussion of alchemy (Chapters 2 and 3) Strathern explores its Hellenistic roots in Alexandria, profiling such practitioners of the Hermetic art as Bolos of Mendes (fl. ca. 200 B.C.E.) and Zosimos of Panopolis (fl. ca. A.D. 300). Although the transmutation of base metals into gold via the philosopher’s stone was the primary goal of western alchemists, Chinese alchemy with its search for the elixir of eternal life or immortality is not neglected. During the half-millenium following the seventh century A.D. the history of alchemy and most of the other sciences remained almost entirely in Arab hands, and Strathern gives considerable space to Jabir ibn Hayyan (Geber, fl. late 8th and early 9th centuries) and his sulphur-mercury theory (British spelling is consistently used throughout the book); Al-Razi (Rhazes, ca. 854–ca. 935), the first to classify substances as animal, vegetable, or mineral; and Ibn Sina (Avicenna, 980–1037), who made contributions to medicine, philosophy, physics, Arab politics, and alchemy. Strathern then returns to the West, where he profiles St. Albertus Magnus (ca. 1200–1280), Roger Bacon (ca. 1219–ca. 1292), Arnold of Villanova (ca. 1240–1311), Nicholas Flammel (well known to readers of J. K. Rowling’s Harry Potter and the Philosopher’s Stone, 1330–1413?), Nicolaus Copernicus (1473–1543), and others, and where he describes various medieval discoveries.

Paracelsus (Theophrastus Philippus Aureolus Bombastus von Hohenheim, ca. 1493–1541) must be a favorite of Strathern’s, for he devotes an entire chapter (Chapter 4) to the itinerant magus, his mercury-sulphur-salt theory of the elements, his doctrine of signatures, and his advocacy of chemistry in the service of medicine (iatrochemistry). Strathern’s survey of this period continues with Dietrich von Freiberg (ca. 1250–ca. 1310), Nicholas of Cusa (ca. 1401–1464), and Giordano Bruno (1548–1600).

A new way of viewing the world emerged with Galileo Galilei (1564–1642) and his telescope; René Descartes (1596–1650) and reason, his prime tool in the quest for knowledge; and Francis Bacon (1561–1626), his Novum Organum and New Atlantis, and his scientific method (Chapters 5 and 6). Chemistry began to evolve from alchemy with the work of Flemish physician Johannes Baptista van Helmont (1579–1644), who grew a willow under controlled conditions, the first application of measurement to an experiment involving both chemistry and biology; Franciscus dele Boë Sylvius (1614–1672), the inventor of gin, who regarded digestion and fermentation as chemical reactions; Otto von Guericke (1602–1686) and his celebrated Magdeburg hemispheres; Robert Boyle (1627–1691) and his masterpiece, The Sceptical Chymist, which attacked the Aristotelian theory of the four elements and gave the term a more modern definition; and Sir Isaac Newton (1642–1727) (Chapter 7).

New elements were discovered by Hennig Brand (fl. 1670), Carl Wilhelm Scheele (1742–1786), Sir Humphry Davy (1779–1829), and Joseph Priestley (1733–1804) (Chapter 8). The mystery of fire and combustion was explained by what was once regarded as chemistry’s unifying principle, called “terra pinguis” by Johann Joachim Becher (1635–1682) and renamed “phlogiston” by Georg Stahl (ca. 1660–1734), which was mistakenly identified with the “inflammable air” (hydrogen) discovered by English aristocratic recluse Henry Cavendish (1731–1810) (Chapter 9). Cavendish’s discovery that a mixture of “inflammable air” and ordinary air produced water proved that water could no longer be viewed as an element and thus sounded the death knell of the Aristotelian four- element theory.

Antoine Laurent Lavoisier (1743–1794) with his emphasis on quantitative measurement, his influential Traité élémentaire de chimie, and his modern definition of an element disproved the phlogiston theory and initiated what has become known as the chemical revolution (Chapter 10). In his atomic theory John Dalton (1766–1844) combined Democritus’ original conception of uncuttable particles with Lavoisier’s application of quantitative measurement to chemistry, while Jöns Jacob Berzelius (1779–1848) determined atomic weights for many elements, a number of which he discovered himself (Chapter 11). Among the attempts to discover a pattern among the elements that were forerunners of Mendeleev’s periodic table were Johann Wolfgang Döbereiner’s (1780–1849) triads, Alexandre-Émile Béguyer de Chancourtois’ (1820–1886) telluric screw, and John Alexander Reina Newlands’ (1837–1898) law of octaves (Chapter 12).

In the final two chapters Strathern returns to Mendeleev. In Chapter 13 he presents biographical information about the great Russian chemist and the data that were available to him. At last, in Chapter 14 he presents the periodic table, Mendeleev’s boldness in predicting and leaving spaces for unknown elements and the discoveries of eka-aluminum by Béguyer de Chancourtois (gallium, 1875) and eka-silicon by Clemens Winkler (1838–1904) (germanium, 1886), which confirmed Mendeleev’s periodic law. Curiously, Strathern does not mention the third of these nationalistically named elements—eka-boron (scandium) discovered in 1879 by Lars Fredrik Nilson (1840–1899). Also, although he attempts to update his tale by concluding his book with the isolation of element 101 (mendeleevium) in 1955, he makes no mention of the discovery of the unsuspected group of the inert (now noble) gases, which also confirmed Mendeleev’s law, or the subatomic structure (protons, neutrons, and electrons), which correlates the chemical properties of the elements with their positions in the table. More importantly, he omits Henry Gwynn Jeffrey Moseley’s (1887–1915) discovery of atomic numbers, which explained the anomalies between the atomic weights and positions in the table of several pairs of elements (argon-potassium, cobalt-nickel, and tellurium-iodine) and which required the reformulation of Mendeleev’s law in terms or atomic number rather than atomic weights.

Strathern does not limit his treatment to the persons cited above but presents a panorama of other famous persons or lesser known characters, often with an emphasis on their amusing foibles or eccentricities. He gives the reader a tremendous amount of information not only about chemistry but also about science and its methods as well. He tells his story in the context of scientific, political, social, and cultural events, but sometimes he adopts a “whiggish” attitude by viewing ideas and concepts in the light of later developments. His book abounds with little known facts and anecdotes, and it includes numerous quotations from original sources. A talented writer, he makes skillful use of humor as well as similes and metaphors, and he does not shy away from making value judgments of the contributions of various scientists.

Because the book is intended for a popular audience, Strathern does not cite specific references, but he does include a list of “Further Reading” (8 pp), some as recent as 1998, but some are incomplete. The volume betrays haste in proofreading as the following list of “typos” shows: Carl not Karl (Scheele, pp 193, 299, and 308); Gillispie not Gillespie (Charles C., pp 225, 297, and 298); oxygen not oxgen (p 245); cerium not serium (pp 250 and 260); selenium not silenium (pp 250, 260, and 308); Johann not Johan (Döbereiner, p 256; Frobenius, pp 79 and 305); iodine not lodine (p 260); Kornileva not Kornilov (Mendeleev’s mother’s maiden name, p 262); Lothar Meyer did not use his first name, Julius (p 289); Lindsay not Linsay (Jack, p 296); Szabadváry not Szabadvary (Ferenc, p 300); and Frängsmyr not Frangsmyr (Tore, p 301).

More serious are the following factual errors: Helium, not rubidium, had recently been detected in the atmosphere of the sun (p 3). The heaviest known element was uranium, not lead (p 3). Mendeleev ordered the elements according to atomic weights not atomic numbers, which were not introduced until 1913–1914 (p 5). The widely circulated remark, attributed to tribunal president Jean-Baptiste Coffinhal at Lavoisier’s trial that “The Republic has no need of scientists,” cited by Strathern, is apocryphal (p 241). Silicon not silenium [sic] was the 12th in Newlands’ table of the elements (p 260). Gustav Robert Kirchhoff was a physicist not a chemist (p 268). Mendeleev’s ninth horizontal group began with boron not bismuth (p 287).

These shortcomings notwithstanding, this book is a short, lively introduction to the history of chemistry that will appeal to and delight not only laypersons but also scientists in general and chemists in particular who are interested in a brief survey of the central science. It also would make an ideal gift for anyone, especially a young person, with a curiosity about science and scientists.

George B. Kauffman

California State University, Fresno, georgek@csufresno.edu

S1430-4171(03)06661-0, 10.1333/s00897030661a

The Nobel Prize: The First 100 Years. Agneta Wallin Levinowitz and Nils Ringertz, Editors. Imperial College Press: London; World Scientific Publishing Co.: River Edge, NJ; Singapore; London, 2001. Illustrations. x + 235 pp, 17.0 ´ 25.3 cm. hardcover $54.00; £37, ISBN 981-02-4664-1; paperback $26; £18, ISBN 981-02-4665-X.

The Nobel Prize, established by the will of dynamite magnate Alfred Bernhard Nobel (1833–1896), was the first truly international award, with no other prize possessing the same global scope and mission [1]. The prizes in six fields of culture [2] awarded to “those who, during the preceding year, shall have conferred the greatest benefit to mankind,” have not only recognized the most significant contributions to the progress of humankind, but they identify the major trends in their particular fields.

In 1994 the Internet was first used to announce the awarding of the Nobel Prizes, and in 1995 the Nobel Web site was created. It has been upgraded to create a virtual museum of science and culture, which can be found on the Internet. In 2000 it was formally inaugurated as the “Nobel e-Museum” (NeM) [3, 4], which is now so up to date that at the moment of the annual Nobel Prize announcement detailed information on the new laureates appears instantaneously on the Internet.

The first prizes were awarded in 1901, making the year 2001 the 100th anniversary of this ne plus ultra of scientific and cultural honors. In connection with the Nobel Centennial various activities, meetings, symposia, banquets, exhibitions, and other special events took place, books were published, and postage stamps were issued [5]. As part of the celebrations, the NeM published a series of reviews dealing with the work of the Nobel laureates in the six fields, which began to appear electronically in 1999. Because of the great interest that has been expressed in these reviews and to make them available to those without Internet access or those who prefer to read them from printed pages, updated versions of these reviews have been published in the book under review here.

Edited by two members of the Nobel e-Museum staff— Executive Editor, Agneta Wallin Levinowitz, and initiator, founder, Director, and Head, Nils Ringertz—who died on June 11, 2002 at the age of 69, The Nobel Prize: The First 100 Years presents a clear perspective on the evolution of human civilization over the past century as reflected by the prizes. It also surveys the major trends and developments, information on Alfred Nobel’s life and philosophy, the history of the Nobel Foundation, and the process for nominating and selecting Nobel laureates.

Each of the reviews dealing with the five prizes established by Nobel’s will begin with the pertinent quotation from the will along with a reproduction of the particular medal. All were written by Swedes except for the Peace Prize (Geir Lundestad is a Norwegian), and those designated with asterisks below include short bibliographies (some with works as recent as 1999 and one with Web sites). Among the conclusions reached in these retrospective essays is the fact that, in opposition to Nobel’s wish that young talented persons should receive a sound financial backing for their work and thus be spared the constant trouble of finding funds, the prizes have concentrated on the importance of the discoveries. Printed on high quality, glossy paper, the book contains full-page black and white portraits of 37 laureates (seven each for physics, chemistry, physiology or medicine, and literature; six for peace; and three for economics). The book’s contents are as follow:

·  Introduction,” by Michael Sohlman (3 pp).

·  “Life and Philosophy of Alfred Nobel,” by Tore Frängsmyr (8 pp), a memorial address presented to the Royal Swedish Academy of Sciences, March 26, 1996.

·  “The Nobel Foundation: A Century of Growth and Change,” by Birgitta Lemmel (12 pp): background, establishment, objectives, statutes and significant amendments, financial management, symposia, donations and prizes, festivities, and the Norwegian Nobel Committee and the Foundation during World I.

·  “Nomination and Selection of the Nobel Laureates,” compiled by Birgitta Lemmel (4 pp, the shortest essay): for each of the prizes a list of the institutions responsible for the awards and the persons eligible for nominating candidates.

·  “The Nobel Prize in Physics,” by Erik B. Karlsson (32 pp, the longest essay): a survey of the prizes in and development of physics classified chronologically within each of the following areas—from classical to quantum physics, microcosmos and macrocosmos, from simple to complex systems, and physics and technology.

·  “The Nobel Prize in Chemistry: The Development of Modern Chemistry,”* by Bo G. Malmström [deceased 2000] and Bertil Andersson (29 pp): a survey of the prizes in and development of chemistry classified chronologically within each of the following areas—general and physical chemistry, chemical thermodynamics, chemical change, theoretical chemistry and chemical bonding, chemical structure, inorganic chemistry, general organic chemistry, preparative organic chemistry, natural products, analytical chemistry and separation science, polymers and colloids, biochemistry, and applied chemistry.

·  “The Nobel Prize in Physiology or Medicine,”* by Jan Lindsten and Nils Ringertz (17 pp): selection procedures; Karolinska Institutet as a Nobel Prize-Awarding Institution; nominations; the laureates and their work classified according to type of work; geographical distribution; and criticisms of the awards.

·  “The Nobel Prize in Literature,”* by Kjell Espmark (17 pp): Nobel’s will and the literary prize; Nobel’s guidelines and their interpretations; and special statutes regarding these prizes.

·  “The Nobel Peace Prize,” by Geir Lundestad (25 pp): A survey of all 107 peace laureates and changes in the distribution of prizes according to countries and areas through the years.

·  “The Sveriges Riksbank (Bank of Sweden) Prize in Economic Sciences in Memory of Alfred Nobel 1969-2000,”* by Assar Lindbeck (20 pp): a chronological review of the prizes during the first 32 years according to the following classes—general equilibrium theory, macroeconomics, microeconomics, interdisciplinary research, and new methods of economic analysis; and problems and difficulties.

·  List of Nobel Laureates (1901–2000) (14 pp).

References and Notes

1.       At a time when nationalism and chauvinism were the norm Nobel’s last will and testament, handwritten and signed at the Swedish-Norwegian Club in Paris on November 27, 1895 was unusual in stating,

It is my express wish that in awarding the prizes no consideration whatever shall be given to the nationality of the candidates, so that the most worthy shall receive the prize, whether he be a Scandinavian or not (Nobel: The Man and His Prizes, 3rd ed.; Nobel Foundation and Odelberg, W., Eds.; American Elsevier Publishing Co.: New York, London, Amsterdam, 1972; p x–xi).

2.       Nobel established the prizes in five areas: physics, chemistry, physiology or medicine, literature, and peace. The prize in economics (officially, the Bank of Sweden Prize in Economic Sciences in Memory of Alfred Nobel) was established by the Sveriges Riksbank at its tercentenary in 1968 and first awarded in 1969.

3.       Ringertz, N. Nobel e-Museum (NeM). Chem. Intelligencer 2000 (April), 6 (2), 22–24.

4.       Nobel e-Museum (NeM). http://www.nobelprize.org/ (accessed Jan 2003).

5.       Kauffman, G. B. The Nobel Centennial 1901–2001. Chem. Educator 2001, 6, 370–384; DOI 10.1007/s00897010520a.

George B. Kauffman

California State University, Fresno, georgek@csufresno.edu

S1430-4171(03)06662-X, 10.1333/s00897030662a

Candid Science II: Conversations with Famous Biomedical Scientists. By István Hargittai; edited by Magdolna Hargittai. Imperial College Press: London, England, 2002; distributed by World Scientific Publishing Co.: Singapore; River Edge, NJ, London, England. Illustrations. x + 604 pp; 16.5 ´ 24.6 cm.; hardbound. $82.00; £56; ISBN 1-86094-280-6; paperback. $38.00; £26; ISBN 1-86094-288-1.

The 36 portraits on the book cover appear in the order of their profiles in the book.

As Contributing Editor of the History feature of The Chemical Intelligencer, Springer-Verlag’s popular but unfortunately short-lived quarterly magazine for the culture of chemistry and related sciences, I have a personal as well as professional interest in István Hargittai’s series of collections of interviews, vignettes, and quotations of famous scientists.

During his six-year tenure (24 issues, 1995–2000) as Editor-in-Chief of The Chemical Intelligencer, Hargittai, sometimes with his wife Magdi, interviewed more than 120 eminent scientists, more than half of whom were Nobel laureates. A number of these interviews did not reach the pages of the magazine, and I hoped that these, along with the many that were published, could appear in print in a handier and more permanent form. The first of the four volumes of Candid Scientists (edited by Magdolna Hargittai), subtitled Conversations with Famous Chemists, appeared in 2000 (Kauffman, G. B.; Kauffman, L. M. Chem. Educator 2002, 7, 184–186; DOI 10.1007/s00897020570a). The second volume, subtitled, Conversations with Famous Biomedical Scientists, which contains interviews with 36 biochemists, chemists, physicists, physicians, biologists, or others, 26 of whom (72 percent) are Nobel laureates, has been published, and it should be followed shortly by the third and fourth volumes, subtitled Conversations with Famous Physicists, and More Conversations with Famous Chemists.

Hargittai, who has clearly done his homework in preparing for the interviews, seeks to uncover the stories behind the most important achievements in twentieth-century biomedicine directly from some of their most eminent participants. They tell us about their backgrounds, families (many came from poor families), and lives, both personal and professional, childhoods (Like me or others of my generation, some had chemistry sets or were inspired by Paul de Kruif’s Microbe Hunters or Sinclair Lewis’ Arrowsmith), influences and career choices, motivations, aspirations, heroes (scientific or otherwise), mentors, hardships and triumphs, philosophies, hobbies and nonscientific interests (several are accomplished musicians), and, of course, their seminal discoveries. In most cases their human feelings shine through their words.

Nobel laureates describe how the prize affected their lives, research, and careers. Most are modest and admit the role of luck in their good fortune. For example, 1959 Nobelist in physiology or medicine Arthur Kornberg states, “I was lucky. It so happened I worked in a field that was glamorous and has remained so” (p 57). Two-time Nobelist Frederick Sanger even declined being knighted: “I just didn’t feel it would suit me” (p 75). The statement by outspoken and controversial 1993 Nobel chemistry laureate Kary B. Mullis is the sole exception:

It’s a wild thing to say this on tape but I’m the most widely educated person I know of…. I know that I don’t know anything at all, but I know a lot more than most people (p 195).

In reply to Hargittai’s serious questions a number of the conversations are laced with humor. For example, 1992 Nobel physiology or medicine laureate Gerald M. Edelman relates an anecdote about two old Jews drinking tea Russian style:

One says, “You know, Max, life is like a glass of tea.” The other says, “Why, Louis?” And Louis answers, “How the hell do I know, am I a philosopher?” (p 219).

In another example, 1994 Nobel physiology or medicine laureate Alfred G. Gilman states,

The Prize is also wonderful for the individual. I’ve jokingly said it’s like being given a lifetime depot injection of Prozac (p 251).

George Klein relates a story:

The chief rabbi and the Catholic cardinal are having dinner in the cardinal’s home and the rabbi refuses to eat non-Kosher food. The cardinal is trying to convince the rabbi that it tastes so good. Finally the rabbi gives in and says, “All right, I will eat it some time later.” “When?” asks the cardinal. “At your wedding,” the rabbi answers (p 435).

And Kary B. Mullis, in response to Hargittai’s question as to what he did with his Nobel Prize money, quips,

Yoga Berra, who had also won a substantial prize, said, I spent most of it on women and alcohol, and the rest of it, I wasted (p 194).

The scientific subject matter discussed includes, among others, biochemistry, molecular biology (interviewee James D. Watson is called the Einstein of this field, p 11), genetics, bioenergetics, genetic engineering (interviewee Paul Berg is considered the “father” of this field, p 180), genetic warfare, biotechnology, artificial intelligence (AI), immunology, protein synthesis, antibodies, growth factors, prostaglandins, interferons, animal experiments, embryonic development, nucleic acids, deoxyribonucleic acid (DNA) structure and sequencing, the race for the double helix, ribonucleic acid (RNA), cloning, mad cow disease (bovine spongiform encephalopathy, BSE), acquired immune deficiency syndrome (AIDS), Alzheimer’s disease, poliomyelitis, cancer, the human genome project, viral transduction, polymerase chain reaction (PCR), crystallography, combinatorial chemistry, reductionism (which 1976 Nobelist in physiology or medicine D. Carleton Gajdusek calls “an egotistical farce [and] mental masturbation,” (p 453), adenosinetriphosphate (ATP), radioimmunoassay (RIA), endothelium-derived relaxing factor (EDRF)—nitrogen oxide (NO), zinc finger proteins, the prion controversy, and Nazi genetic experiments.

Among related topics we find discussions of government and public service, gender bias and feminism, Nobel politics (two interviewees are Nobel committee members), research planning, the Great Depression, serendipity, the right to life movement, the ethics of suicide, discovery and patenting of discoveries (some do, and some don’t), teaching, administration, “neural Darwinism” and “molecular Darwinism,” disease dissemination, relationship of DNA and race to intelligence, entrepreneurship, religious beliefs, history of science, aging, authorship of articles, priority, bombing of concentration camps, science and nationalism, the role of imagination in science and the arts, changing fields, rejection of articles, work in progress [Kary B. Mullis: “Right now, I’m working on something big.” (p 191)], predictions of future trends, their legacy, and advice to young people [1979 Nobelist in physiology or medicine Rosalyn Yalow’s advice was one word: “Study!” (p 523)].

Each interview is prefaced with a biographical sketch and includes one or more portraits of the interviewee, many photographed by Hargittai or his wife Magdi. The volume contains 176 illustrations of apparatus, formal and informal group portraits (one with Pope John Paul and one with Fidel Castro), notebooks, letters, models, commemorative postage stamps, plaques, and drawings of Max F. Perutz and John C. Kendrew by 1915 physics laureate Sir William L. Bragg (p 283). Hargittai’s questions are printed in italics, and the much longer responses in Roman type. Structural formulas, equations, and reaction schemes are included when necessary. Three of the interviewees are now deceased, underscoring the importance of such oral histories. The earliest born chemist is 1986 Nobel physiology or medicine laureate Rita Levi-Montalcini (b. 1909), and the youngest interviewee is Richard Henderson (b. 1945). A name index (10 double-column pages) but no subject index is provided.

Because Hargittai is a friendly fellow scientist rather than an investigative reporter, he never tries to deal with a problem with which the interviewees seem uncomfortable, and he tactfully asks them to ignore any of his questions that they do not wish to discuss. In return, they are often candid and frank in their responses to questions that he does ask. Hence the title of his book is most appropriate, as is its subtitle; actually, its contents are more candid informal conversations rather than formal interviews.

Thus several scientists discuss their differences with other scientists and competitors. For example, Alfred Mirsky is criticized by Maclyn McCarty, and Erwin Chargaff is criticized by Joshua Lederberg and Arthur Kornberg. Kornberg is also extremely critical of Kary B. Mullis, whom he claims was not worthy of the Nobel Prize:

He epitomizes what I find wrong with celebrity in science and the Nobel Prize. [He] never did anything important in science and he never will. He hit upon an idea, which may have seemed original to him, but Gobind Khorana and his group had already carried out amplification of DNA many years earlier (pp 65–66).

On the whole, however, most of the scientists are well acquainted with each other and are mutually supportive, and their names crop up frequently in each other’s interviews. Some offer suggestions as to Nobel-caliber scientists whose candidacy was overlooked, either because of the three-person rule for sharing a prize or for other reasons, for example, Francis Ryan (Joshua Lederberg, George W. Beadle, and Edward L. Tatum shared the 1958 prize in physiology or medicine); Charles Weissman (Stanley Prusiner alone received the 1997 prize in physiology or medicine for the discovery of prions); and Salvador Moncada (His failure to receive the 1998 prize in physiology or medicine, which went to Robert F. Furchgott, Louis J. Ignaro, and Ferid Murad, created an international uproar). Also Oswald T. Avery, Colin M. MacLeod, and Maclyn McCarty were overlooked for the transforming principle; David Keilin, for the discovery of cytochromes; and Karl Lohmann, for the discovery of ATP.

From the conversations I also learned many surprising or little known facts or judgments: James D. Watson thinks that his work with the longest-ranging impact is

Probably my books….Any good chemist focused on DNA should have found the structure of DNA….But The Double Helix could not have been written by anyone beside [sic] myself (p 7).

Kary B. Mullis got his idea of PCR during a night drive from Berkeley to Mendocino on Highway 128 (p 188).

Rita Levi-Montalcini relates how

When I was still a very young girl, I already decided that I would never marry and have children because that would interfere with my goal of serving people (p 371).

More remarkably, her twin sister Paola, whom she calls “the best Italian woman artist of the century” made the same choice (p 371).

And in a little choice piece of gossip, I learned that Jonas “Salk divorced his first wife and married Picasso’s girlfriend” (p 507).

Because previous subscribers to The Chemical Intelligencer may be potential buyers of Candid Science II, here is an annotated list of its contents (CI, previously appeared in the same or modified form in The Chemical Intelligencer; N, Nobel laureate; D, deceased; F, female; J, Jewish):

1.      James D. Watson (13 pp) N

2.      Maclyn McCarty (16 pp) CI

3.      Joshua Lederberg (18 pp) CI N J

4.      Arthur Kornberg (22 pp) N J

5.      Frederick Sanger (12 pp) CI N (the only person to receive two Nobel chemistry prizes—1958 and 1980)

6.      François Jacob (14 pp) N J

7.      Walter Gilbert (16 pp) CI N J

8.      Benno Müller-Hill (16 pp) CI

9.      Marshall W. Nirenberg (12 pp) CI N J

10.    Daniel Nathans (12 pp) N D J

11.    Paul Berg (28 pp) CI N J

12.    Kary B. Mullis (14 pp) CI N

13.    Gerald M. Edelman (24 pp) CI N J

14.    César Milstein (18 pp) CI N J

15.    Alfred G. Gilman (14 pp) N J

16.    Günter Blobel (14 pp) N

17.    George K. Radda (14 pp)

18.    Max F. Perutz (16 pp) CI N D J

19.    Richardson Henderson (10 pp) CI

20.    Aaron Klug (24 pp) CI N J

21.    John T. Finch (8 pp)

22.    Sidney Altman (12 pp) CI N J

23.    Edward B. Lewis (14 pp) CI N J

24.    Rita Levi-Montalcini (12 pp) N F J

25.    Lars Ernster (20 pp) D J

26.    Torvard C. Laurent (16 pp) CI

27.    George Klein (24 pp) J

28.    D. Carleton Gajdusek (24 pp) N

29.    Charles Weissman (32 pp, the longest piece) J

30.    Frederick C. Robbins (20 pp) N

31.    Rosalind Yalow (6 pp, the shortest piece) CI (excerpted from a book review) N F J

32.    James W. Black (18 pp) N

33.    K. Sune D. Bergström (6 pp) N

34.    John R. Vane (16 pp) CI N J

35.    Salvador Moncada (14 pp) J

36.    Robert F. Furchgott (16 pp) N J

Only two of the scientists are women (Rita Levi-Montalcini and Rosalind Yalow) so, despite the increasing acceptance of women in academic, industrial, and governmental laboratories, further advances in the struggle against sexism are needed. An unusually high proportion of the interviewees (at least 22 or 61 percent) are Jewish, so the issues of Judaism, the Holocaust, and anti-Semitism are discussed by many of them. As expected of a diverse group of highly individual persons with strong opinions, the interviewees do not always agree on these or other topics. Rita Levi-Montalcini, interviewed by Magdolna Hargittai, who states, “My interest in science was just casual and I owe it to Benito Mussolini, who prevented me to become a regular doctor, because I was no Aryan” (p 369), attributes the predominance of Jews among laureates to Jewish “extraction” and attributes it to tradition not genetics:

Jewish people have a tradition to care for their children and to send them to the best possible schools. Most great scientists are children of very simple people, who escaped from the pogroms of Russia or Poland, and they became Nobel Prize winners (p 369).

On the other hand, George Klein maintains:

It clearly has a genetic component. This is what we have been selected for over the centuries in the ghettos (p 425).

Concerning the prevalence of anti-Semitism, Arthur Kornberg, who also wrote the book’s foreword and who found difficulty in obtaining a suitable position, insists:

Anti-Semitism was pervasive in these United States, including academia. I have called it a virus that has been endemic for two thousand years (p 52).

But 1962 physiology or medicine laureate Max F. Perutz states,

I am always disturbed by Jewish chauvinism that the Jewish response to anti-Semitism is to be as intolerant as their enemies. The wonderful thing about British academic life is that there is no anti-Semitism (p 291).

On the other hand, Aaron Klug and George Klein acknowledge that anti-Semitism exists in South Africa and Sweden, respectively.

I found the interviews in this volume of more interest to me than those in Hargittai’s first volume, probably because the information was new to me because I was more familiar with chemists than with biomedical scientists. Of particular appeal to me—and would possibly be to you as well—were the following:

·  Benno Müller-Hill, who while not Jewish himself, carried out research and wrote a book on the history of genetics in Nazi Germany, infamous for its heinous crimes against humanity, and the silence surrounding this history until the present day (Murderous Science: Elimination by Scientific Selection of Jews, Gypsies, and Others in Germany, 1933–1945; Cold Spring Harbor Laboratory Press: New York, 1988).

·  Rita Levi-Montalcini, the author of In Praise of Imperfection, who considers herself an artist and not a scientist, maintains that Primo Levi, whom she knew well, did not commit suicide, a conclusion which I always suspected in view of the fact that he was a survivor of various misfortunes, including internment at Auschwitz.

·  D. Carleton Gajdusek, sometimes called the “peripatetic pediatrician” and who calls himself “a pagan and proud of it” (p 446), considers his most important contributions not those for which he won the 1976 Nobel Prize in physiology or medicine (diseases of the nervous system, including “kuru” in Stone Age cultures) but for his work on the variation in the cognitive function of the brain. He adopted, brought back to his home in the United States, and educated more than 60 children from Stone Age cultures of New Guinea and Micronesia. He was arrested and served a prison term on charges of child molestation and fiscal irregularities, but his adopted children, students, and colleagues, including George Klein, Charles Weissman, and Frederick C. Robbins, came to his defense, characterizing him in this book as a genius and polymath. His views on science, philosophy, and his legacy are both provoking and provocative. He is most proud of finding that people from Stone Age cultures can make the leap of more than 20,000 years of cultural history in a lifetime, from a culture level earlier than the ancient Greeks and Egyptians and the Altemira and Lascaux cave painters to modern arts and science (p 465).

·  George Klein was born in Budapest, became a Holocaust survivor, and migrated to Sweden, where he became a well-known personality not only as a scientist and writer, who has won literary prizes, but also because of his frequent participation in public dialogues and discussions. He cogently discusses a number of important and controversial scientific and nonscientific topics.

In view of the length and scope of the book, the number of errors is extremely small and limited to “typos”—“Alsac” for “Alsace” (p 92), “explain me” for “explain to me” (p 188), “Welcome” for “Wellcome” (p 221), “Erlich” for “Ehrlich” (Paul, p 233), “McCarthy” for “McCarty” (Maclyn, p 392), and “Lohman” for Lohmann” (Karl, pp 392, 393, and 600).

In his book, For the Love of Enzymes, interviewee Arthur Kornberg quoted the late Lars Ernster, another interviewee, as estimating that “40 percent of the chemistry Nobel Prizes as well as 40 percent of the Nobel Prizes in Physiology or Medicine have been in biochemistry” (p 59). Also, in his preface Hargittai states,

The science of the second half of the 20th century was dominated by the biomedical fields and this is expected to continue for the foreseeable future. The present selection of interviews gives a cross section covering a broad range of topics, personalities, and circumstances of recording (p vii).

I agree with Hargittai’s evaluation and heartily recommend his book, suitable for both complete reading or browsing, to biomedical scientists, biochemists, chemists, historians of chemistry and or science, and general readers interested in the “inside story” of the workings of 20th-century science.

George B. Kauffman

California State University, Fresno, georgek@csufresno.edu

S1430-4171(03)01663-9, 10.1333/s00897030663a

Prometheans in the Lab: Chemistry and the Making of the Modern World. Sharon Bertsch McGrayne. McGraw-Hill: New York, NY, 2001. xi + 243 pp, 15.2 ´ 23.5 cm. hardcover $24.95, ISBN 0-07-135007-1.

This fascinating and authoritative book presents a well-balanced appraisal of the Janus-faced nature of chemistry—its positive and negative aspects as exemplified through case studies of the lives and works of nine chemists, critically described in great detail and full historical context and presented in roughly chronological order. As such it constitutes a valuable weapon in the arsenal of those of us who fight a continuing battle against the forces that for various reasons promote chemophobia—the irrational fear of chemicals.

Its author is Sharon Bertsch McGrayne, an award-winning journalist covering education, politics, science, and health issues (for Scripps-Howard, Crain’s, Gannett, and other newspapers), science writer and editor for the Encyclopædia Britannica, and author of five books beginning with Nobel Prize Women In Science: Their Struggles, and Momentous Discoveries (Birch Lane Press: New York, NY, 1993). Her book originated with a suggestion made by O. Theodor Benfey, who at the time was associate editor of the now discontinued “The Chemical Sciences in Society” series to be published jointly by the American Chemical Society and the Chemical Heritage Foundation.

Some of the chemists whom McGrayne selected for detailed treatment are well known to chemists and chemical educators and have been the subjects of book-length biographies, for example, Perkin, Frankland, Haber, and Carothers. Others will be less familiar such as Leblanc, Rillieux, Midgley, Müller, and Patterson. In every case she lucidly explains their scientific and technological contributions in relatively simple, jargon-free language suitable for a nonscientific audience. Although I have written articles about Perkin, Haber, Carothers, and Midgley, in each of these cases, I learned several facts and speculations of which I was previously unaware. The lives of several of these scientific pioneers are filled with tragedy as well as triumph: Fritz Haber’s first wife Clara and his son Hermann committed suicide as did Leblanc and Carothers. McGrayne also implies that Midgley committed suicide (pp 104–105).

Here is a summary of the book’s contents:

·  Chapter 1: “Soap and Nicolas Leblanc: December 6, 1742–January 16, 1806” (15 pp): Most chemical educators have heard of the Leblanc process if only in connection with the Solvay process that replaced it. It  made available washing soda (NaCO3) or “white gold” that put cheap soap within everyone’s reach and for a half century it was virtually synonymous with the chemical industry until it polluted industrial regions so badly that it created a cry for environmental reform.

·  Chapter 2: “Color and William Henry Perkin: March 12, 1838–July 14, 1907” (15 pp): Perkin, who made his serendipitous discovery at the age of 18, turned black coal tar into a beautiful dye that blanketed the world in “mauve mania,” spawned the artificial dye and pharmaceutical industries, institutionalized scientific research, professionalized chemists, changed the economies of vast regions, and helped make Germany the world’s foremost industrial power.

·  Chapter 3: “Sugar and Norbert Rillieux: March 17, 1806–October 8, 1894” (13 pp, the shortest chapter): This irritable and combative New Orleans African-American (“quadroon libre”), who was a cousin of French Impressionist painter Edgar Degas, became the most famous engineer in Lousiana. His triple-effect evaporator was used not only in the sugar industry but today is used wherever large volumes of liquids must be evaporated and is still ranked “among the world’s great inventions.” Incidentally, he is the subject of a recent short article (Benfey, T. The Progenitor of All Industrial Evaporation. Chemical Heritage 2002 (Fall), 20 (3), 4–5).

·  Chapter 4: “Clean Water and Edward Frankland: February 20, 1825–August 9, 1899” (15 pp): Best known to chemists as the originator of the valence concept and the notation that we still use to denote chemical bonds as well as the discoverer of the first organometallic compounds, Frankland, London’s water analyst and the world’s leading authority on water issues, implicated fecally contaminated drinking water as the cause of cholera and typhoid epidemics that killed millions. His fight for pure water contributed greatly to the long life expectancy that we enjoy today. McGrayne also describes how he was able to overcome the stigma of illegitimacy in socially rigid Victorian England to become president of the Chemical Society, which delayed its memorial oration until 35 years after his death.

·  Chapter 5: “Fertilizer, Poison Gas, and Fritz Haber: December 9, 1868–January 29, 1934” (21 pp): Haber discovered how to convert atmospheric nitrogen into ammonia in what became the model for twentieth-century industrial processes, most of which involved high-pressure synthesis. Its use for fertilizer saved millions from starvation, but its use in producing explosives killed millions. Today excess nitrogen—half of it from fertilizer runoff—has deoxygenated water in the Gulf of Mexico, Chesapeake Bay, the Baltic Sea, Black Sea, and Adriatic Sea. Furthermore, Haber’s use of chlorine gas during World War I made him the most hated chemist in the world and created an uproar when he received the 1918 Nobel Prize in chemistry.

·  Chapter 6: “Leaded Gasoline, Safe Refrigeration, and Thomas Midgley, Jr.: May 18, 1889–November 2, 1944” (27 pp): Midgley’s two great discoveries are examples par excellence of useful everyday materials whose unexpected effects on the environment require the balancing of benefit versus risk in the use of chemical products. His chlorofluorocarbons (CFCs or Freons) created the air conditioners that we take for granted and made possible the safe refrigeration that freed us from levels of food poisoning now known only in developing countries. However, in 1974 they were found not only to destroy the tropospheric ozone layer that protects us from skin cancer but also to contribute to global warming. His tetraethyllead (TEL) additive provided cheap, knock-free gasoline but at the price of contaminating our atmosphere.

·  Chapter 7: “Nylon and Wallace Hume Carothers: April 27, 1896–April 29, 1937” (42 pp, the longest chapter): Despite medical problems and the depression that eventually caused his suicide, Carothers, in a meteoric career that lasted only less than a decade, fathered the science of organic polymers. In one of the great technological feats of the twentieth century he built huge molecules from smaller ones to produce neoprene, the first synthetic rubber; polyesters; and nylon, the first commercially successful synthetic fiber.

·  Chapter 8: “DDT and Paul Hermann Müller: January 12, 1899–October 13, 1965” (20 pp): Without Müller’s dicholorodiphenyltrichloroethane, for which he won the 1948 Nobel Prize in physiology or medicine, millions would have died of insect-borne malaria or typhus during and after World War II. Because DDT dissolves well in oils, it accumulates in animal fat and mammals’ milk and becomes increasingly more concentrated in predator species as it moves up the food chain where it can wreak havoc among wildlife. In his Nobel acceptance address Müller voiced his concerns about using DDT in complex biological ecosystems, but it was not until the appearance of Rachel Carson’s popular 1962 book, Silent Spring, one of the twentieth-century’s most influential journalistic exposés, that the general public became aware of the problem. However, because DDT is not regarded as a human carcinogen, since the resurgence of malaria beginning with the 1970s (about one-tenth of the Earth’s population is infected annually), medical authorities have argued for continuing its use despite its harm to wildlife.

·  Chapter 9: “Lead-Free Gasoline and Clair C. Patterson: June 2, 1922–December 5, 1995” (30 pp): I found this chapter the most interesting, for most of it was new to me (I must admit that I had never heard of Patterson). This unjustly neglected geochemist used lead to determine the age of the Earth and the solar system, and he identified high levels of lead in his laboratory and in ocean sediments, surface waters, alpine meadows, and even Greenland’s ice. He discovered the difference between postindustrial and preindustrial lead levels in humans (a factor of 2000 as shown by his “measles” cartoon [p 195]) in his 1980 report to the National Academy of Sciences). He showed chemists how to analyze global environmental pollution and stimulated medical research into the effects of low-level lead pollution on human learning ability. He relentlessly fought against giant industries, governmental agencies, and even other scientists, transforming lead into a public enemy until this neurotoxic element was eliminated from gasoline (introduced by Midgley, as described in Chapter 6), food containers, pipes (the name “plumber,” of course, is derived from the Latin word for lead), tanks, toothpaste tubes, and countless other commercial products. In the words of eminent geochemist Gunter Faure,

Patterson’s a saint. The fact that we stopped using leaded gasoline is largely his doing, and he showed you can’t study lead contamination in humans because there are no uncontaminated people anymore. I think he was a very brave man. As a man, he ranks right alongside Newton and Galileo (p 197).

What an inspiring and upbeat quotation to conclude McGrayne’s history of the chemical revolution through the lives of nine persons who helped create our modern way of life!

Patterson’s saga epitomizes the central thesis of McGrayne’s book—that science in general and chemistry in particular can solve any problems that it has unintentionally created. She masterly summarizes this and the change that has taken place since the late eighteenth century with which she began her story, in a three-page Postlude:

Chemistry’s relationship to the public is unique among the sciences. Chemistry’s products become part of our everyday lives and are profoundly intertwined with society’s tastes, needs, and desires…. As a result of growing sophistication of science and technology, a fundamental shift has occurred in the relationship between consumers and industrial pollution. Gone are the days when neighboring property owners were the first to complain about Leblanc pollution. Today the public is often unaware of problems until scientists discover them. The greatest concentration of resources and expertise for solving the public’s environmental and safety concerns now resides—not among consumers at large—but in the science and engineering communities in industry, government, and universities. In fact, monitoring technology on behalf of the public requires as much scientific expertise as the manufacturing process does itself….In the end, the responsibility of citizens and industrial leaders is the same: to insist that the information provided by modern science and engineering is used wisely for the public good. By working together for the public welfare, we can enjoy the enormous benefits of industrial technology and protect Earth for coming generations” (pp 199–200).

The book is meticulously documented to an extent unheard of in a popular book intended for a general audience. A detailed annotated bibliography of up-to-date magazine, newspaper, and journal articles; books; letters; manuscripts; interviews; and other archival sources, classified according to chapters, runs to 38 pages and could be the envy of a professional historian of science. In some cases McGrayne conducted her own interviews, and she also submitted the chapters for criticism and suggestions to specialist historians of science, who are dutifully acknowledged in the bibliography (She contacted me in connection with my articles on Midgley and Carothers). Several chemical equations and one structural formula (DDT, p 153) are provided where appropriate. The chapters can be read separately or in succession; each tells a complete story, but they are all interrelated and are discussed and compared in the different chapters, which are carefully cross-referenced. A five-triple-column-page index facilitates location of material.

The few errors are mostly “typos”: “Thomson” for “Thompson” (Benjamin [Count Rumford], pp 3, 243); salvaran for salvarsan (pp 28, 243); “phosphorous” for “phosphorus” (p 178); “Science” for “Sciences” (National Academy of, p 196); “George H.” for “George H. W.” (Bush, p 196); and “Stolzenberg” for “Stoltzenberg” (Dietrich, p 212); but several are factual: Beethoven, not Tchaikovsky, composed the Pathétique Sonata; Tchaikovsky composed the Pathétique Symphony (p 126); and Carl (“Speed”) Marvel not Roger Adams succeeded Thomas Midgley, Jr. as President of the American Chemical Society (p 112).

McGrayne’s inexpensive, well written, and absorbing account of how scientific discoveries are made, the persons who made them, and how these discoveries can irrevocably change our lives, both for good and bad, demonstrates in an exciting and absorbing way the pivotal role that the “central science” has played in creating the modern world. It critically examines the tangled and complicated interrelationships between the public’s insistence on progress and comfort and the need to preserve the environment. I recommend it enthusiastically to anyone—nonscientist or scientist, generalist or historian of science—interested in the mutual interaction of science and technology with society. It should be beneficial for science teachers in general and chemical educators in particular. It would also make an excellent gift for students or young persons contemplating a scientific career.

George B. Kauffman

California State University, Fresno, georgek@csufresno.edu

S1430-4171(03)01664-8, 10.1333/s00897030664a