Lost Elements: The All-American Errors - ACS Symposium Series

Oct 30, 2017 - He distributed a paper based on a course he taught at the University of Genoa in which he stated: Once my students have become familiar...
3 downloads 14 Views 3MB Size
Chapter 1

Lost Elements: The All-American Errors

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Mary Virginia Orna,*,1 Marco Fontani,2 and Mariagrazia Costa2 1Department

of Chemistry, The College of New Rochelle, New Rochelle, New York 10805, United States 2Department of Organic Chemistry, University of Florence, Via della Lastruccia 13, Sesto Fiorentino, Firenze, Italy, 50019 *E-mail: [email protected]. E-mail: [email protected].

In the development of the periodic table, many wrong turns were taken and falsely claimed element discoveries discredited. Among these were the “American” elements, those “discovered” by three Presidents of the American Chemical Society, and those named after states of the Union: North Carolina, Illinois, Alabama, and Virginia. This chapter narrates their detailed stories.

Introduction Probing the nature of an elemental substance has never been an easy task. Getting down to the stuff that the material universe is made of, with respect to number, nature, and nomenclature, has baffled the best of minds through the millennia. Aristotle’s (384-22 BCE) philosophical framework dominated Western thought for over two thousand years, a system that banished number, weight, and measure to insignificance. Though the Aristotelians were always intent on tidying up the world, the world as they knew it came to an untidy end, in western culture, around the year 1750. Prior to that year, much occurred to demolish Empedocles’ (495-30 BCE) four-element hypothesis. Fast forward to the sixteenth century: Paracelsus (1493-1541) claimed that the four earthly substances (earth, air, fire, water) had three spiritual sources, namely, mercury, salt, and sulfur. To complicate matters, Robert Boyle (1627-91), in his book The Sceptical Chymist (1) contended that our knowledge of the elements must be enlarged upon by experiment, not mere speculation. He also attacked the four-element hypothesis ((1), pp. 33-38, passim), opening up the possibility that the number of elements © 2017 American Chemical Society

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

might not be so limited. Thus, by the dawn of the seventeenth century, practicing scientists began to think in terms of discovering new elements. A century later, Antoine Laurent Lavoisier (1743-94) advanced the germinating modern idea of “element” still further by casting doubt on the Greek four-element hypothesis in his famous Traité élémentaire de chimie (2) (Figure 1): We would be surprised not to find in an elementary treatise of chemistry a chapter on the constituent and elementary parts of bodies; but I note here that our tendency to want that all bodies in nature are composed of only three or four elements is due to a bias that comes originally from the Greek philosophers. The introduction of four elements, which, by the variety of their proportions, comprise all the bodies that we know, is pure conjecture, imagined long before we had the first notions of experimental physics and chemistry.

Figure 1. Cover of the 1789 edition of Antoine Laurent Lavoisier’s Traité élémentaire de chimie. Reproduced from reference (2). Subsequently, Lavoisier flew in the face of Aristotelian philosophy by giving supreme importance to observation, experimentation, and measurement in studying the properties and composition of material substances. He believed, and demonstrated, that elements survived in their compounds and could be recovered from them. This analytical approach, that emphasized concrete laboratory experimentation as opposed to theoretical speculation, was the groundwork of the chemical revolution (3). Finally, together with Louis-Bernard Guyton de Morveau (1737-1816), Antoine-François Fourcroy (1755-1809) and Claude-Louis Berthollet (1748-1822), he devised a new chemical nomenclature that was based on the principle that an element’s name should correspond to its composition (4). 2

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Thus the lines were drawn regarding number, nature, and nomenclature of the elements. There now seemed to be no limit to the number of elements – it was a virtual “open season” for the eager discoverer. The nature of an element could now be determined using a growing battery of analytical tools. And the naming of an element, just as Adam named the animals that paraded before him in the Book of Genesis, belonged to the person who recognized it first.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Before 1789: Early Errors and Early Elements It is quite obvious that more substances of an elemental nature were recognized from ancient times than the traditional earth, air, fire, and water. Metals found in their free state such as gold, and sometimes sulfur and silver, were known from prehistory. Other metals such as iron and mercury could easily be extracted from their ores; other elements known from antiquity are carbon, copper, tin, and lead. In subsequent centuries, phosphorus, bismuth, arsenic, antimony, and zinc were discovered by alchemists. But the age of discovery of chemical elements began only in the second half of the eighteenth century, around 1750. Since 1789 marks the year of the chemical revolution, we can set it as the arbitrary dividing line between protochemistry and chemistry. During the forty year period between 1750 and 1789, about a dozen new elements were claimed to have been discovered. Very few chemists, if such can be named, were operating at this time, and none of them had access to more sophisticated tools than visual observation of physical properties of substances after treatment with chemical reagents such as acids and bases. At the same time, however, there was a growing interest in metallurgy because the prosperity of a nation was in direct proportion to the productivity of its mines. Mining led to the discovery of new minerals. New minerals had to be analyzed in order to exploit their utility. Chemical analysis revealed large numbers of new substances, many of which were recognized as elements, but many of these substances were mixtures mistaken for elements. So the field was ripe for errors of various kinds. Table 1 gives a sample of some of the erroneous discoveries made during this time period. An examination of Table 1 shows that many countries in Western Europe have a false discovery to their names. Most of the errors involved mistaking a mixture for an element, but there is one instance where a true element, tellurium, was only shown to be so ten years after the fact. Von Reichenstein really did have a problem on his hands, and he sought confirmation from Bergman, who was considered the greatest living chemist at the time. But Bergman died before he could render his opinion, and von Reichenstein simply let the matter drop. Ten years later, Klaproth volunteered to analyze the sample, and indeed, confirmed the presence of an as yet unknown element; he also volunteered the name “tellurium” when the true discoverer failed to come up with a name to his own liking (12, 13). In the next section, we will see how the European propensity for discoving erroneous elements spilled over into America. 3 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Table 1. Erroneous Element Discoveries Before 1789

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Date

“Element”

Discoverer

Country (5)

Presumed Actual Substance

1777

Terra nobilis (6)

Torbern Olof Bergman (1735-84)

Sweden

Impurity from diamond

1777-78

Hydrosiderum (7)

Johann Karl Friedrich Meyer (1733-1811)

Poland

Alloy of iron and phosphorus

1783

Metallum problematicum (8)

Ferenc Müller von Reichenstein (1740-1825)

Hungary

Tellurium*

1786

Saturnum (9)

Antoine Grimoald Monnet (17341817)

France

Mixture of copper and lead sulfides

1788

Terra adamantina (10, 11)

Martin Heinrich Klaproth (17431817)

Germany

Mixture of corundum and alluminite

1789 – 1869: From the Chemical Revolution to the Periodic Table The 80-year period leading up to the publication of the periodic table was filled with intense chemical activity. The concept of the chemical element was slowly growing in the consciousness of many chemists, although there were still individuals who believed in phlogiston! It was a period of ferment and change. Too great a reliance on technical skills, which were becoming more and more reliable in the isolation of new elements, often led unwary chemists astray. The period was characterized by gross errors in chemical analysis, multiple names for the same “element,” and rediscovery, decades later, of previously rejected elements. Over fifty false claims of element discovery were catalogued during this time. Many carried very fanciful names such as crodonium, pluranium, aurorium, and jargonium. Hardly any country in Europe was exempt from such errors; in addition to the countries noted in Table 1, England, Italy, Russia, and Lithuania, to name just a few, added their names to the list. Meanwhile, a certain degree of caution seemed to pervade the chemical community. Some researchers, hesitantly thinking that they had discovered a new element, declined to give the substance a name, and some even perpetrated jokes under the cloak of anonymity. For example, two gentlemen from Newark, Delaware, communicated the discovery of a supposed new element by the name of brillium to the Washington Post (14). Found in coal ashes, it reputedly gave off more heat than ordinary fuel. During this period, two chemists destined to one day become Presidents of the American Chemical Society, added their names to the list of discoverers of these indeterminate elements. 4

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

1852: The Element of Friedrich August Genth (1820-93)

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Friedrich August Ludwig Karl Wilhelm Genth was born in Wächterbach bei Hanau, Germany, on 16 May 1820. Following studies at the Universities of Heidelberg and Gießen (with Justus von Liebig (1803-73)), he earned his doctorate in chemistry in 1845 at the University of Marburg under Robert Bunsen’s (1811-99) supervision. He remained at Marburg as Bunsen’s assistant when he discovered the first nickel oxide crystal (15) (Figure 2). Then in 1848, he emigrated to the United States. In Philadelphia, Pennsylvania, Genth opened one of the first commercial analytical chemistry laboratories in the country. However, two years later, he closed it down in order to take up the position of superintendent of mines at Silver Hill, North Carolina, for a short time.

Figure 2. Book Jacket of a Volume Commemorating Friedrich Genth’s Discovery. Reproduced from reference (15), with permission. 5 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

While Genth was analyzing a platinum sample from California, he recovered two grains (about 100 mg) of a metal with an intense white color (16, 17). He found that it was malleable and melted immediately in the presence of charcoal and upon treatment with the oxyhydrogen blowpipe. It was attacked by hot hydrochloric acid, hot nitric acid, and with hydrogen sulfide, it yielded a brown precipitate. While Genth published his experimental results immediately, he never gave his presumed discovery a name. Successive studies showed that the substance that fooled him was a mixture of the oxalate and cyanide of platinum and the chlorides of palladium and iridium (18). A complex mixture indeed, but a mixture nonetheless. There is no record of a retraction by Genth, but since his element bore no name, it simply disappeared into the maw of history. Genth, meanwhile, returned to Philadelphia in 1850, reopened his laboratory, and devoted his energies to commercial chemistry, consulting, research, and instruction of a few students. In 1872, he was named professor of analytical and applied chemistry and mineralogy at the University of Pennsylvania, on condition that he be allowed to maintain his private practice. Certainly not resting on any laurels, he busied himself as well as chemist and mineralogist for the Second Geological Survey of Pennsylvania in 1874, and as the chemist for the Board of Agriculture of Pennsylvania from 1877 to 1888. In that year, he reopened his commercial analytical laboratory yet once again, and managed it until his death five years later on 2 February, 1893. In 1880, he served as fifth President of the American Chemical Society (Figure 3).

Figure 3. Frederick Augustus Genth, Fifth President of the American Chemical Society. Courtesy: American Chemical Society. 6 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Adept as he was at multi-tasking, Genth was remarkable for the substantial contributions he also made to fundamental science, particularly in the area of complex cobalt-amine compounds, tellurides, phosphates, fertilizers, and rare minerals. He was passionate about the latter, having personally collected and catalogued about 5,000 specimens. Among these he identified and described about 20 new minerals, including calaverite, maconite, and penfieldite (19).

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

1862: The Element of Charles Frederick Chandler (1836-1925) Born on December 6, 1836 into a Massachusetts family of modest means, Charles Chandler distinguished himself early on as a hard worker who would take on extra duties and manufacture the time to do them by getting up early and going to bed late. His first interest in chemistry was awakened in high school by an inspiring and enthusiastic instructor who left a deep and lasting imprint on his eager student – chemistry soon became Chandler’s dominant lifelong interest. His work habits enabled him to earn enough money to outfit a small chemistry laboratory in the attic of the family home, and due to the many hours he spent there, he developed a proficiency unusual for a person so young. In 1853, he began to attend the Lawrence Scientific School of Harvard College, but soon found that a chemistry curriculum was virtually non-existent in the United States. Advised to study in Germany, he was helped along by a new acquaintance, Charles Arad Joy (1823-91), professor of chemistry at Union College, Schenectady, New York, who provided him with a letter of introduction to Friedrich Wöhler (1800-82), in Göttingen, who welcomed him immediately into his laboratory in 1854. Chandler, not only hard-working, but precocious, eventually earned the M.A. and Ph.D. degrees in 1856, when he was barely twenty years old! In the following year, he became assistant professor of chemistry at Union (although janitorial duties accompanied this appointment because that is where the budget line was), and succeeded to the chair of the department a few years later when Professor Joy moved on to Columbia College in the City of New York. At the time of his presumed discovery, Chandler was not yet 26 years old. In the preceding year, he had analyzed a mineral found in the Rogue River in Oregon, using the tried and true methodologies that he had perfected while doing the research for his doctoral dissertation under the guidance of Professor Heinrich Rose (1795-1864) in Berlin. Chandler was confident that he had come upon a new metal but felt that he needed additional sample in order to confirm it. Waiting in vain for a year, he finally decided to publish his results, done in triplicate according “to the ordinary routine of qualitative analysis,” with the same outcome. However, in the last paragraph of his paper (20), Chandler notes that a colleague apprised him of work done by Friedrich Genth ten years previously. The metal observed by Dr. Genth occurred among grains of platinum from California. It was malleable; it fused readily on charcoal before the blowpipe, becoming covered with a coating of black oxyd (sic); it dissolved in borax to a colorless bead, which became opalescent on cooling; it was dissolved by hot hydrochloric acid and by nitric acid; 7 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

and its solution gave a brown precipitate with hydrosulphuric acid. It seems quite probable, therefore, that the metal which I have observed in the Rogue River platinum, is identical with that observed by Dr. Genth.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Realizing that he had duplicated Genth’s work, Chandler wisely let the matter drop, but not before actually publishing it (21). Through a series of happy circumstances, not least of which was Chandler’s analytical expertise, he too received an invitation to Columbia, this time to instruct in the new School of Mines being formed at that institution. And there he remained for the following 46 years until his retirement in 1910; he acted as Dean of the school for 33 of those years (Figure 4).

Figure 4. Portrait of Charles Frederick Chandler. Michael De Santis, 1935, oil on canvas; Gift of Francis P. Garvan. Courtesy of the Union College Permanent Collection. 8 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

In the minutes of the Board of Trustees, following a recital of his many achievements, this entry can be found (22): Professor Chandler will carry with him into his retirement the affectionate regard and esteem of two generations of students as well as a host of colleagues on the teaching staff of the University. The Trustees record their grateful appreciation of this long and generous career of devoted service.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Furthermore, Columbia’s President, Nicholas Murray Butler (1862-1947) refers, in his annual report of November 7, 1910, to Chandler’s wide-ranging activities during his tenure ((21), p. 186): To his teaching power as well as to his effective and conscientious service as administrator, the Department of Chemistry and the School of Mines, to which it primarily belonged, owed almost everything for many years. Professor Chandler has long been a point of contact between the University and the public, between science and industry and the public health. His career is unique of its kind, and we shall not soon look upon his like again. As Butler observed, Chandler’s interests were many and varied and his expertise and service ethic left a profound impact on his adopted city as well as the state of New York and the nation in general. Here are some examples:







Chandler’s contacts with industry led to the accumulation of many chemical products that soon outgrew the space in his office; the initial collection served as the nucleus of Columbia’s famous Museum of Chemistry (23). Chandler opened his museum in order to show his students the things he talked about in his many lectures. A placard on display at the entrance to the chemistry department continues: “He began to collect material for the museum almost immediately on his arrival at Columbia as the Civil War was ending. For half a century after, he bought rare and interesting exhibits of chemicals and products of various chemical industries. Many times he bought items from exhibits out of his own pocket and materials were donated by the chemical industries...Professor Chandler took great delight in keeping up the museum.” Among the many exhibits are collections of dyes, rare earth elements (Figure 5), minerals, and laboratory equipment. When the New York College of Pharmacy was struggling as a “one-room schoolhouse,” Chandler offered his services and lectured and demonstrated (at his own expense) for three evenings per week year after year until the college could grow and expand, eventually becoming the College of Pharmacy of Columbia University In 1872, Chandler was appointed adjunct Professor of Chemistry and Medical Jurisprudence at the fledgling New York College of Physicians 9

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.





Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001





• •









and Surgeons; for over 20 years, he lectured there every day between 5 and 6 PM, always raising his voice in favor of a more thorough scientific training for those entering the medical profession In 1870, together with his brother, William Henry Chandler (1841-1906), of Lehigh University, he founded the American Chemist and continued its publication until it was superseded by the Journal of the American Chemical Society in 1877 Chandler was a founding member of the American Chemical Society and twice served as its President in 1881 and 1889 Chandler initiated and oversaw the construction of Havemeyer Hall which opened its doors to students and faculty in 1898. A state-of-the-art chemistry department at the turn of the 20th century, this site was home to many famous scientists and notable scientific achievements. A century later, in 1998, it was designated a National Historic Chemical Landmark by the American Chemical Society (24) Chandler was recognized as the highest authority in his day in this country in industrial chemistry, serving as consultant in areas like sugar refining, petroleum refining, photomechanical processes, and calico printing; he invented and introduced the system of assay weights used to this day by assayers and metallurgists He was a pioneer in the area of water quality and sanitation, publishing landmark papers on these subjects Chandler devoted himself assiduously to all branches of hygiene and sanitary science, studying with the utmost care and thoroughness all factors bearing upon the health of a great city; for eleven years, he served as the President of the New York City Board of Health He served as one of the scientific directors of the New York Botanical Garden, and was for many years chemist for the Croton Aqueduct Commission For several years, Chandler was president of the New York State Charities Aid Association and took an active part in securing proper state care for the indigent insane His chemical expertise was called upon in the work of commissions under four U.S. Presidents: Chester A. Arthur (1829-86),Grover Cleveland (1837-1908), William McKinley (1843-1901), and Theodore Roosevelt (1858-1919). Chandler was the first scientist to receive an honorary D. Sc. degree from Oxford University (1900)

After a brief illness, on August 25, 1925, Charles Frederick Chandler died at his home in New York City at the age of 89. The young chemist who, in 1862, had the brashness to announce the false discovery of an element similar to platinum had become a national icon, celebrated both at home and abroad. He was mourned deeply by all who knew him or were familiar with his remarkable contributions to the well-being of humankind (25). 10 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Chandler’s papers are now held in the Columbia University Archives (26) along with a finders’ guide. The materials are held off-site and require a 24-hour notice for access; it is a treasure trove for anyone planning a full-fledged biography of this extraordinary gentleman.

Figure 5. A Display of Rare Earth Element Compounds from the Chandler Museum Collection, Havemeyer Hall, Columbia University. Most of the Elements are Displayed as their Oxides. Photograph: Mary Virginia Orna.

1869 – The Advent of the Periodic Table 1869 seems like a very late date for Dmitri Mendeleev’s proposal of an orderly arrangement of elements to appear given the fact that phenomenological relationships among the elements had accumulated in the previous decades. From the 1820s on, scientists like Johann Wolfgang Döbereiner (1780-1849), Leopold Gmelin (1788-1853), Oliver Gibbs (1822-1908), and William Odling 11 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

(1829-1921) “played” with the idea of groups of elements with similar properties. Then in 1860, Stanislao Cannizzaro (1826-1910) provided what many historians of science consider the catalytic idea that precipitated the simultaneous proposals for what we now call the periodic table (27). In that year, many of the leading scientists of the day met at Karlsruhe, Germany, to discuss vexing questions such as the uncertainty of many chemical formulas of even the simplest of compounds. Cannizzaro, influenced by the theory of chemical bonding derived from Amedeo Avogadro’s (1776-1856) hypothesis, emphasized the importance of atomic weights. He distributed a paper based on a course he taught at the University of Genoa in which he stated: Once my students have become familiar with the importance of the numbers (i.e. atomic weights)…, it is easy to lead them to discover the law which results from their comparison. ‘Compare,’ I say to them, ‘the various quantities of the same element contained in the molecule of the free substance and in those of all its different compounds, and you will not be able to escape the following law: The different quantities of the same element contained in different molecules are all whole multiples of one and the same quantity, which, always being entire, has the right to be called an atom (28).’” Cannizzaro illustrated his statement with many examples taken from his own experimental data, and it is a known fact that many of the attendees of the congress were profoundly influenced by him, including Julius Lothar Meyer (1830-95) and Dmitri Ivanovitch Mendeleev (1834-1907). Nevertheless, the pathway from atomic weights to the periodic table was anything but direct, which is why it took almost ten years for both Mendeleev and Meyer to publish their ideas, which had percolated in their minds for a long time. I. S. Dmitriev, Director of the Mendeleev Institute at the University of Saint Petersburg, writes (29):

Mendeleev’s discovery of the Periodic Law did not follow a linear pathway, but rather one that was complicated, difficult, winding, one that utilized various criteria over a period of time.

It is beyond the scope of this chapter to give a detailed narrative of the periodic table’s development; many more comprehensive works have done a fine job in this regard (30, 31). Once chemists realized that not only could the periodic system bring order out of chaos, had predictive possibilities, was a guiding light in the search for new, and now “known to be missing” elements, but also that it served as a theoretical tool, as a map of the way in which electrons arrange themselves in atoms, and even the history of the development of life itself (32), it quickly took its rightful place as the “chemist’s Bible.” It has gone through many revisions since it was first visualized by Mendeleev and Meyer (Figure 6).

12 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Figure 6. A monumental image of Dmitrii Mendeleev surrounded by “his” elements arranged like a sundial. Plaza facing the Faculty of Chemical and Food Technology Building of the Slovak University of Technology in Bratislava, Slovakia. Photograph: Mary Virginia Orna.

1877: J. Lawrence Smith (1818-1883) and His Discoveries Although J. Lawrence Smith served as the second President of the ACS in 1877, he was a Johnny-come-lately to the Mendeleev bandwagon: in the very year of his presidency, he had the misfortune of claiming the discovery of a new element that embroiled him in unlooked-for controversy. Although Mendeleev’s periodic table could predict missing elements in what we now call the main body of the table, the case of the rare earth elements was a different matter. They seemed to fit nowhere but had to fit somewhere in the table, but no one could fathom just how many there were, and if there was any limit to their number. Coupled with the fact that rare earth elements’ chemical similarities were so close that they were difficult to separate, chemists who worked with them often ended up with mixtures that they believed were pure substances. Smith, at this stage in his career, occupied himself with a serious study of the mineral samarskite, a rare earth-bearing mineral first extracted from the southern Ural mountains of Russia and named after the Russian mining official, Vasili Samarsky-Bykhovets (1803-70). Unknown to Smith, another chemist with formidable analytical skills and vast experience in the separation of rare earths, Marc Abraham Delafontaine (1837?-1911?), in Chicago, was devoting his time 13

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

to examining the same mineral. It is not surprising that virtually simultaneously (although Delafontaine had a slight edge), both announced the discovery of a new element. Delafontaine called his philippium in honor of his benefactor, M. Philippe Plantamour of Geneva; Smith called his mosandrium (or mosandrum) after the great rare earth chemist, Carl Gustav Mosander (1797-1858) (33). An immediate priority dispute erupted which was played out in the pages of Comptes rendus chimie (34). Delafontaine claimed that the presumed mosandrium was actually composed of about 75-80% terbium and a 20-25% mixture of yttrium, erbium, didymium (later resolved into its two constituent elements, praseodymium and neodymium, by Carl Auer von Welsbach (1858-1929) in 1903) (35) and “his” philippium. History has been kind to Delafontaine. The element discovered and isolated by Swedish chemist, Per Theodore Cleve (1840-1905), in 1878 and called holmium after the city of Stockholm, was later recognized to be identical to philippium, so today both Cleve and Delafontaine share credit for the discovery, although the latter lost the glory and privilege of conferring the name. Meanwhile, Smith’s claim fell into oblivion. Undaunted, Smith continued his assiduous study of samarskite. He announced the discovery of two new elements, columbium (36), presumably from its source in the mineral columbite and rogerium (in honor of his natural philosophy professor at the University of Virginia, William Barton Rogers (1804-82), thus being the first person to name an element after a person yet living. These claims were noted at a meeting of the National Academy of Sciences that took place on October 28-30, 1879 (37): Prof. J. Lawrence Smith gave an informal account of some recent researches for new elements. Some years ago, he found a field of research in the cerium and yttrium minerals, and was well satisfied that he had obtained a new substance, which he named mosandrum, in the cerium group. Since then he has been studying the compounds of samarskite and has found, he believes, two new elements, one of which he calls columbium, and the other he proposes to name in honour of his friend and the instructor of his youth, Prof. William B. Rogers. But having much other business requiring his attention, Prof. Smith has done little in that line of research, since then, except to purify some mosandrum. So, since neither of these discoveries seemed to be credible, no further work was done on samarskite except for a privately published piece in 1881 (38) that appeared two years later in the American Journal of Science (see reference (33)). Smith continued to publish almost up until the date of his death in 1883, but most of his interests had turned to the analysis of meteorites. In his Bakerian Lecture of 1883, William Crookes (1832-1919) remarks in a footnote (39): Dr. J. Lawrence Smith in a paper read before the United States National Academy of Sciences in 1879, announced the discovery in Samarskite of two new elements, which he named Columbium and Rogerium (Wyckoff, W. C. “U.S. National Academy.” Nature 1879, 21 (11 December), 14346). I have failed to find any further notice of these elements. This 14

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Columbium should not be confounded with the well-known Columbium sometimes called Tantalum. With this remark, Crookes seems to have dealt the death blow to columbium and rogerium in the very year of Smith’s own death. Aside from Smith’s penchant for discovering erroneous elements, he was internationally well-recognized for his chemical expertise. A precocious child bordering on genius, he was proficient in mathematics at the age of four. Following two years of study at the University of Virginia, he took his medical degree at the University of South Carolina, submitting a thesis entitled “The Compound Nature of Nitrogen.” After receiving his medical degree, he continued his education in Paris, studying organic synthesis with the likes of Jean-Baptiste Dumas (1800-84) and medicinal chemistry with Mathieu J. B. Orfila (1787-1853). However, it was his chance visit to the laboratory of Justus von Liebig at Giessen in 1841 that turned the entire course of his life and landed him firmly in the field of chemistry. Thenceforward, dividing his time between Paris and Giessen, he did elegant chemical work that earned the respect of his mentor, Liebig. For example, in 1842, he published a landmark paper on the composition of spermaceti, possibly one of the earliest forays into organic chemistry by an American chemist (40) that was reproduced in several prestigious European journals. Ever peripatetic, Smith found employment in the Ottoman Empire for several years, conducting research on the natural resources of that area with a view to their commercial exploitation. Back in the United States, he spent several years lecturing in New Orleans, and later on in Virginia and in Washington, D.C. Then in 1854, he received an appointment as Professor of medical chemistry and toxicology at the University of Louisville (KY), in which post he remained for twelve years. Then in 1866, he resigned from formal teaching, finding it a restrictive burden on his many projects that required extensive travel. He established, instead, a home laboratory where he conducted private research in analytical and mineralogical chemistry, and in toxicology. He developed the potassium chromate test for barium and the J. Lawrence Smith method of analysis of silicates for alkali metals in 1853. Both of these tests remained the best analytical tools available for approximately a hundred years. In 1872, Smith was elected a member of the National Academy of Sciences, and in 1879, he became corresponding member of the Académie des Sciences, Institut de France, succeeding Sir Charles Lyell in this capacity; he was the first American to hold this post. He received many honors both at home and abroad. He was a Chevalier of the Légion d’Honneur, President of the American Association for the Advancement of Science in 1874, and the second President of the American Chemical Society in 1877. Early in his career, Smith began to collect meteorites, and a large body of his published work, especially during his later years, consists in their analyses. One of the final acts of his career was the transfer of this collection, by that time monumental, to Harvard University, by way of purchase (Figure 7). The last paper that Smith published was in June 1883 (41). In it, he promised further research, “my health” permitting. It did not. On October 12, 1883, J. Lawrence Smith succumbed to the ravages of chronic liver disease (42). 15 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Figure 7. Meteorite Display at the Harvard University Museum of Natural History Earth and Planetary Sciences Gallery. © President & Fellows, Harvard College, courtesy Harvard Museum of Natural History. Photograph: Jess T. Dugan. In an adjacent display case, a caption reads, “The Harvard collection includes significant specimens from the personal collections of J. Lawrence Smith, a famous 19th century chemist.” The Harvard purchase of Smith’s meteorite collection was funded by subscription to which a number of prominent Bostonians contributed (43). Smith’s widow, Sarah Julia Smith, donated the funds in “deed of trust” (44) to the National Academy of Sciences to promote research on meteorites, and the monies are still used to this day for that purpose. In addition, Mrs. Smith also endowed the J. Lawrence Smith Medal (Figure 8) and Prize in memory of her husband. The National Academy has awarded it every three years since 1888 for recent, original, and meritorious investigations of meteoric bodies. 1901: Charles Baskerville and the “Discovery” of Carolinium At the dawn of the 20th century, one great organizing chemical principle was in place: the periodic behavior of the elements, having been enunciated some thirty years previously by Dmitri Mendeleev. It was a necessary, but unfortunately, not sufficient, breakthrough in the business of element discovery. In what we now call the main groups of the elements, it was clear that some elements still remained undiscovered in Mendeleev’s time, and the following decades witnessed their identification. At the same time, where to place the rare earths, presently called the lanthanides, and how many might actually exist, was still a mystery. How much more of a mystery, then, was the placement of what would later be called the actinides, a problem that Glenn Seaborg (1912-99) took up in 1944 while 16

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

he and his group were struggling with the placement of the transuranium elements in the periodic table (45).

Figure 8. The J. Lawrence Smith Medal. Courtesy: National Academy of Sciences. Image used with permission.

In 1901, a distinguished chemistry professor at the University of North Carolina, Charles Baskerville (1870-1922), was having a similar struggle with one of the members of this group, thorium. He did not have the advantage of two additional breakthroughs that would come later and clarify many things, i.e., the concept of atomic number and the concept of isotopes. Furthermore, the practice of fractional crystallization was the only means available at the time for substance purification, a fact that rendered so-called “very pure” samples almost moot. Charles Baskerville was an inspiring teacher and occupied an esteemed place in chemical education for almost thirty years. His lecture style was characterized by remarkable clarity and lucid expositions drawn from his highly organized store of notes and references. He was noted for his research on the rare earths and on the chemistry of anesthetics. On the industrial side, he studied pulp and paper recycling, and the refining and hydrogenation of vegetable oils (46). His early university education was at the University of Mississippi, after which he attended specialty courses at the University of Virginia and Vanderbilt University, finally 17

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

completing his B.S. degree in chemistry at the University of North Carolina in 1892. Following an intensive summer course in chemistry at the University of Berlin (1893), he returned to North Carolina to complete his doctoral work under F. P. Venable (1856-1934) in 1894. Remaining at the university following the completion of his degree work, he advanced rapidly through the professorial ranks, becoming chair of the chemistry department upon Venable’s accession to the presidency of the university. In 1903, Baskerville claimed through two papers published in the Journal of the American Chemical Society that he had discovered two new elements embedded in samples of thorium compounds extracted from monazite sands, carolinium (he had suggested the presence of carolinium as early as 1901) and berzelium. Though these discoveries were later refuted, Baskerville received enough acclaim from his professional colleagues to be invited to occupy a faculty position at the City College (later, City University) of New York, which he accepted in 1904. In reading Baskerville’s two papers claiming the discovery of carolinium, one cannot help but notice the “stream of consciousness” style that seems to have characterized his work. Far from dividing his papers into the sections customary today – experimental, results, discussion – we are treated to a running commentary full of details such as (47): …when the extraction had continued about six hours, it was discovered that an overlooked small defect of the cork had permitted the gradual introduction of about a drop of water, which came from the sweating of the condenser overhead. As this vitiated the experiment…the experiment must be discarded… Acknowledging that the chemistry of thorium was very complex, Baskerville attempted to prepare pure samples of compounds of thorium such as the oxide and the tetrachloride. In doing so, he took measurement of specific gravity as an index of the purity of the compound. In his determinations, he encountered the anomalous presence of an oxide with an unusually high specific gravity which, he claimed, “cannot be accounted for except by the presence of either a new oxide of a known element having greater density than the usual non-volatile residue after ignition, or an unknown element” ((46), p. 767). Later on in the same paper, Baskerville asserts that his preparations contained “a constant unknown impurity in practically all the materials used. This constituent must be an element of much higher atomic weight….between 260 and 280. On account of the extensive occurrence, in this state (North Carolina), of the monazite sands from which the original material was obtained, if the investigation give a successful issue, I should like to have the element known as Carolinium, with the symbol Cn.” ((46), p. 773) Then, Baskerville immediately says that this is a preliminary paper only….so, stay tuned for more information in a subsequent publication. However, Baskerville, in that same year, presented a framed set of “carolinium” samples to the Columbia University Chandler Museum, where it remains on display to this day (Figure 9). Since we know that carolinium does not exist, what the vials contain is also a mystery to this day. 18

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Figure 9. Samples of Carolinium Compounds (Oxide, Sulphate, and Acetyl-actonate [sic]) on Display at the Chandler Museum Collection, Havemeyer Hall, Columbia University. Photograph: Mary Virginia Orna.

The Museum Caption reads: “In 1901, Dr. Charles Baskerville of the City College of New York announced ‘the Existence of a New Element Associated with Thorium.’ This potential new element, heavier and more radioactive than thorium, he named ‘Carolinium’ (Cn) for the North Carolina location of the monazite sand from which it was obtained.” The promised publication arrived in 1904 (48). In it, Baskerville reviews the history of thorium research and how it was previously misidentified as other new elements such as donarium and wasium, as well as the presumed thorium α and thorium β of Bohuslav Brauner (1855-1935) with atomic weights respectively of 220 and 260-280. The upshot of Baskerville’s conclusions is that thorium is not a primary radioactive body, that it is complex and not a chemical element, and that it can be resolved into at least three other distinct bodies sufficiently identified to deserve distinct names (including his own carolinium and berzelium), and that he awaits spark spectral confirmation of these conclusions ((47), p. 941). And a note appended to these observations by William Crookes, the doyen of state-of-the-art spectroscopy at that time, states that “Making allowance for the fact that my spectrum is from the metal, while that from your material is from the 19

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

chlorides solutions, all five spectra are practically identical, all the prominent lines being seen in each spectrum, while there are no lines in one which are not seen in the others…” ((47), p. 942) Crookes softens the blow by asserting that these are preliminary results and by no means prove that such bodies (carolinium and berzelium) do not exist. However, since we hear no more of these two presumed elements in the literature, we assume that Baskerville (Figure 10), reproduced from the Chemical Heritage Foundation Archives (49), provided no new evidence regarding their existence.

Figure 10. Charles Baskerville. Reproduced from reference (49), with permission. So, where did Baskerville take a wrong turn? His experimental work seemed to be very detailed, but closer examination reveals that it was incomplete and more than likely not reproducible. True, he determined the atomic weights of his presumed new elements (255.6 and 212.0 respectively). He determined that the properties of their oxides were not the same as those of the parent material, thorium, from which they were extracted. He confirmed the phosphorescence and radioactivity of the “new” substances, and when he realized that the arc and spark spectra were identical, he then supposed that the material examined was 20 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

not sufficiently pure and that the spectral data were not complete. However, it is very likely that he had bit off more than he could chew. Thorium was known to be difficult in every way. Its isolation and purification from other elements in the matrix were accomplished with difficulty only a decade after Baskerville had published his papers (50). Regarding the observed radioactivity of some of the fractions, one might presume that the presence of uranium in greater or smaller amounts as an impurity would account for the enhanced activity of the fraction attributed to carolinium. While Baskerville made note of the radioactivity, he did not attribute it to thorium since he even denied that thorium was radioactive. And it is certainly not possible that he had succeeded in isolating the 212Th isotope since he used only chemical means in his investigations, and one cannot separate isotopes of the same element that way since they exhibit identical chemical behavior. When all is said and done, it is quite possible that Baskerville failed simply because of the limitations inherent in his own research. Seventy-five years after Berzelius discovered thorium, Baskerville was using essentially the same laboratory techniques. Perhaps sensing that this line of research was a dead end, he abandoned his search for confirmation of the element that would have been the first to be named after a state of the United States for much more practical work in organic chemistry. During his relatively brief chemical career, Baskerville managed to author six books and over two hundred technical papers. He was an active member of several scientific societies, having served in official capacities in some of them and was even elected a Fellow of the American Association of the Advancement of Science. Hardly 52 years of age, Baskerville succumbed to a bout of pneumonia on January 28, 1922 (51).

1913 and 1915: Atomic Number and Isotopes As technology advanced, many elements were discovered that confirmed Mendeleev’s initial predictions. Some bumps along the road were how to accommodate the plethora of rare-earth elements, the unexpected discovery of the noble gases, and of numerous radioactive species that seemed to be individual new elements until the existence of isotopes came to be understood. Antonius Johannes van den Broek (1870-1926) was a Dutch amateur physicist notable for being the first who realized that the number of an element in the periodic table (now called atomic number) corresponds to the charge of its atomic nucleus. This hypothesis was published in two papers in 1911 (52, 53), just one month after Ernest Rutherford (1871-1937) published the results of his experiments that showed the existence of a small charged nucleus in an atom, and inspired the experimental work of Henry G. J. Moseley (1887-1915), who found good experimental evidence for it by 1913. Moseley foresaw that his X-ray method would “prove a powerful method of chemical analysis…It may even lead to the discovery of missing elements, as it will be possible to predict the position of their characteristic lines” (54). 21

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Thus the atomic number not only placed a limit on the number of elements that could be discovered, but it also served as the key to atomic structure that identified each element – the defining property of an element could no longer be ascribed to its atomic weight. Following upon the results of this landmark paper, chemists realized that only seven of the naturally occurring elements remained to be discovered, thus cutting down drastically the number of reported false discoveries and setting in motion an element hunt full of controversial competing claims that lasted for decades (55). Barely two years following Moseley’s discovery of what came to be called atomic number, Frederick Soddy (1877-1956) and his research assistant, Ada Hitchins (1891-1972), working at Glasgow and later at Aberdeen, demonstrated that the density of elemental lead from various sources differed substantially, whereas the atomic volume remained constant. The only conclusion to draw from this observation was that the atomic weights of lead could vary, leading directly to the concept of isotopes, for which Soddy received the Nobel Prize in chemistry in 1921 (56).

1926: B Smith Hopkins (1873-1952) and Illinium In the decade that followed, chemists used these two new tools so cleverly that by 1925, only one rare earth element remained unidentified: the recalcitrant element 61. Its existence had been surmised by Bohuslav Brauner in Prague as early as 1902 (57): Apart from the 10 elements already listed…and more or less accurately studied by me, about seven to ten additional elements could be placed in this group…It is not impossible that one would be able to split neodymium, Nd = 143.8, into at least one element with a smaller atomic weight, and into another element with a higher atomic weight of about 145 and, similarly, some more gaps lying in the area between Ce and Ta could be filled.

When it was later shown that element 61 did indeed exist, Brauner claimed credit for the discovery in a letter to Nature (58):

I arrived at the conviction that the gap between the neodymium and samarium was abnormally large. In my paper…read in St. Petersburg in 1902, I came to the conclusion, not reached by any chemist before – that the following seven elements, possessing now the atomic numbers 43, 61, 72, 75, 85, 87, and 89, remained to be discovered. As regards element No. 61, the difference between atomic weights of Sm-Nd = 6.1, and it is greater than that between any other two neighboring elements. 22 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Later work by Moseley showed definitively that an element should exist between neodymium and samarium (59). Researchers took up the challenge, and many were in hot pursuit, assuming that this element lay hidden in very small amounts among the other rare earths. Consequently, they continued to use the method that had led to success in so many other cases: fractional crystallization. In 1924, at the University of Florence, a newly-minted chemist, Lorenzo Fernandes (1902-77), urged on by his mentor and autocratic head of the chemistry department, Luigi Rolla (1882-1960), developed an expertise in the separation and purification of rare earths, as well as the X-ray check on the purity of the samples. The department had plenty of space, eager workers, and later on, enormous amounts of raw material to work with. Rolla realized that there was a gap between neodymium (Z = 60) and samarium (Z = 62), and who could better fill it than he and his department? Little did he realize that more than 56,000 fractional crystallizations later, he would still be uncertain about what he had (or did not have) in his hands. Thinking that he and Fernandes had observed the characteristic X-ray spectrum of 61, yet not wishing to make a premature announcement to claim priority, Rolla sent a sealed packet containing his sample and a note about his claim to the Accademia dei Lincei (the Italian equivalent of the National Academy of Sciences) – just in case someone else came along to make the claim while he was trying to obtain a larger and purer sample (60). Rolla had good reasons for his fears. Unbeknownst to him, a team headed up by B Smith Hopkins (Figure 11) at the University of Illinois in the United States had begun their search for this missing element in 1923. They were in “friendly competition” with Charles James at the University of New Hampshire who, prodded by William Ramsay (1852-1916) (61), took up the search more than a decade earlier in 1912. Hopkins, cognizant of the possibility that element 61 might be the rarest of the rare earths and virtually undetectable because of the difficulty of separating enough of it for X-ray analysis, nevertheless soldiered on. A caveat expressed early on indicated that he realized that since no new absorption bands had ever been observed in the intermediate fractions when the double magnesium nitrates of the rare earths had been subjected to fractional crystallization, element 61 might be concentrating with neodymium whose extensive absorption bands could succeed in masking any other bands present in the regions being examined (62). The team then followed the first paper with a companion piece (63) in which they claimed discovery of the new element, which they named illinium (after the university and state of Illinois (64)) based on the presence of 130 arc lines in the red and infrared spectrum and five lines toward the violet, corresponding closely to the theoretical positions for the Lα1 and Lβ1 for element 61 (65). Published virtually simultaneously was an article by James M. Cork (1894-1957), Charles James (1880-1928) and Heman C. Fogg (1895-1952) (66) which contained the L series spectroscopic values that came uncannily close to those published two decades later after the element had been isolated.

23 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Figure 11. Professor B Smith Hopkins, 1937. Photo courtesy of the University of Illinois at Urbana-Champaign Archives, Faculty, Staff and Student Portraits, Record Series 39/2/26, Box 32, “Hopkins, B Smith.”

At this point, Luigi Rolla, throwing all caution to the winds, immediately revealed the contents of his sealed packet in the pages of the Gazzetta Chimica Italiana in three parts: (1) A two-page review of the results of his search for element 61 (67), (2) A much longer piece giving experimental details on the extraction and further procedures on the samples in hand (68), (3) A confirmation of observed spectra of element 61 by an independent laboratory (69). Predictably, the Illinois announcement coupled with Rolla’s speedy action precipitated a priority dispute between the Florence group and the chair of the Illinois chemistry department, the renowned W. A. Noyes (1857-1941), that played out in the pages of Nature (70). Noyes starts out by citing the history of the search for element 61 at the University of Illinois in 1919, in partnership with the U.S. Bureau of Standards, resulting in three publications over three successive years, 1921, 1922, and 1923. Noyes notes that the second publication appeared at just about the time that Rolla began his work, and two years before Rolla deposited his sealed packet with the Academia dei Lincei. Reviewing additional work at Illinois, Noyes concludes (71): 24

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

In the light of these facts it would seem that the honour for the discovery of No. 61 belongs primarily to Prof. Hopkins, and that the element should be called Illinium rather than Florentium. This does not detract from the credit which Prof. Rolla should receive for his independent discovery of the element. Both Prof. Rolla and Prof. Hopkins realise that a large amount of additional work must be done before the element can be fully accepted.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Rolla’s response involved reviewing the history of the reports on the spectra of the supposed element 61 by Carl Clarence Kiess (1887-1967) and Leonard Francis Yntema (1892-1976), quoting the latter (72): X-ray analysis of samples from different sources has so far given no evidence of the presence of this element. Rolla goes on to say that it is not sufficient evidence for assuming the existence of a new element discovered only from the fact of having seen new spectral lines. And then he delivers his punch line: We obtained…the first photographs of K-absorption spectra showing the characteristic band of element 61…the first certain data…[and] we believe that we should be credited with priority for the discovery. The spectroscopic analysis had been done by Prof. Rita Brunetti (1890-1942) of the University of Florence Physics Department in Arcetri using state-of-the-art instrumentation. Noyes’s immediate response gave short shrift to the Florentine claim (73): The fact that Prof. Rolla deposited a [packet] instead of publishing his paper, demonstrates that he was not, at that time, sure of his discovery…Harris, Hopkins, and Yntema…were sure of their results on the basis of four independent lines of evidence… Realizing that endless correspondence in the pages of the literature would not solve the impasse, Rolla, seeking confirmation of his discovery, actually traveled to Illinois to see with his own eyes what progress was being made in the isolation of element 61. Finding none, and with his mind at rest, he stopped off in Copenhagen on his way home and presented a sample of his material to Niels Bohr for confirmation. Subjecting Rolla’s enriched florentium sample to a much more scrupulous and accurate spectroscopic examination than had ever been done before, Bohr’s analysis came up empty: element 61 did not exist in Rolla’s material. It was only fourteen years later, in 1941, that Rolla sent a letter of retraction, partly written in Latin and couched in an extensive description of his work on neodymium and samarium, to the little-known Vatican journal of the Pontificia Academia Scientiarum (74). And while one might deem this no retraction at all, it 25 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

was far more than B Smith Hopkins ever did. To his dying day, Hopkins traveled extensively in the United States trying vainly to salvage illinium from the ashheap. His initial announcement of discovery was welcomed everywhere, and many textbooks had incorporated illinium, with the symbol Il, into their endflap periodic tables (75, 76). In Ref. (73) one can see the image of the periodic table published by Hopkins himself and labeled “The Nearly Completed Periodic Table of the Elements, 1926.” One can note the symbol, Il, for element 61, and the uncertain atomic weight, 146? As well as the blank spots for elements 85 and 87, and the symbol Ma for masurium, element 43, which was falsely claimed by Walter and Ida Tacke Noddack. At the University of Illinois, a large periodic table was painted onto the wall of the large lecture hall in the Chemistry Department; it was allowed to remain there for several years after Hopkins’s death in 1952. Illinium was allowed to remain in the periodic chart and the descriptive chemistry section of Hopkins, B. S. and Bailar, J. C., General Chemistry for Colleges, 5th Ed. (D. C. Heath & Co.: Boston, 1932) until 1956 in deference to the senior author (77). The genuine element 61 made its debut in 1947, two years after its actual discovery and isolation from uranium fission fragments by ion exchange chromatography at Oak Ridge National Laboratory in Tennessee (78). Originally called prometheum at the suggestion of Mrs. Grace Mary Coryell, wife of one of the discoverers, Charles D. Coryell (1912-71), its name was later changed by I.U.P.A.C. to reflect the suffixes of many of its companions in the periodic table: promethium. Prometheus was the Titan who stole fire from the gods and bestowed it upon humanity; in retaliation, Zeus condemned him to be chained to a rock where his liver was daily eaten by an eagle (presumably Zeus himself). Thus the name promethium is fraught with symbolism: the danger of having a powerful force in one’s hands and the frightful consequences thereof. Members of the American Chemical Society got their first look at compounds of promethium, yellow PmCl3 and pink Pm(NO3)3 at the June, 1948 ACS national meeting in Syracuse, New York. None of the would-be discoverers lived to find out that promethium actually did exist in nature. In 1956, Paul K. Kuroda (1917-2001) of the University of Arkansas hypothesized that the uranium ore coming from a pitchblende deposit in Oklo, Gabon (79), found to have a depleted level of U-235, could have undergone spontaneous nuclear fission. His hunch was later confirmed when the expected lighter elements, products of U-235 fission, were present in much higher quantities than normal. Kuroda subsequently organized a gigantic task force to extract naturally-occurring element 61 from pitchblende (80). The mass of the natural isotope is 147. The story of element 61’s discovery brings together over 100 years of chemical history, as noted by Clarence Murphy (b. 1934) (81): …no element has been “discovered” and named more times than 61…[it is intimately connected with the development of the understanding of atomic structure and of the Periodic Table…The story involves Roentgen’s discovery of X-rays and Moseley’s use of X-ray spectra to determine atomic numbers. It involves the more than one hundred-year effort to separate the rare earths and to find a place for them in the 26

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Periodic Table. Finally it involves the development of ion-exchange chromatography and research on the atomic bomb during World War II…

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

1930: Professor Fred Allison (1882-1974), Virginium, and Alabamine When Bohuslav Brauner wrote his famous letter to Nature (58) regarding the atomic numbers of the elements he had divined that were yet to be discovered in 1902, he was witness to a frenzied hunt for the missing members of the periodic table that was already in progress, with two successes happening prior to his publication, i.e., in 1917 and 1923. Furthermore, Brauner was mistaken regarding his list of seven because one of these elements, actinium, had already been discovered in 1899 by André Debierne at the Institut du Radium, Paris. However, there were still seven elements if one substitutes element 91 for 89, as did Eric Scerri, correctly, in his book, A Tale of 7 Elements ((55), pp. 176-79). For clarity, please see Table 2. By 1930, only three elements remained on the Brauner-Scerri list (if we discount element 61, “discovered” in 1926 and destined to live on as illinium until at least 1956), and Fred Allison (1882-1974) was determined to reap the honors for discovering two of them, numbers 85 and 87. In 1927, Allison (Figure 12), then at the Alabama Polytechnic Institute, began to publish a series of papers on his use of the Faraday effect to identify the presence of dissolved metal ions in minute quantities. Essentially, by placing a salt solution in a glass cell through which a beam of polarized light can pass, the plane of polarization can be rotated by application of a magnetic field, and the rotation observed is accompanied by an increase in the brightness of the light. By adjusting the brightness to a minimum relative to a reference liquid such as carbon disulfide and changing the nature of the salt, it was possible to measure the time lag of the light’s appearance as a function of the identity of the metal ion in the solution. Allison found that each chemical substance, independent of the presence of other substances, produced its own characteristic minimum, or minima, of light intensity, persisting down to a concentration of about one part in 1011 (83). In a subsequent paper, Allison and his co-worker, Edgar J. Murphy, reported that their results showed that the positions of the minima observed were functions of the atomic equivalents of the metallic elements. Furthermore, they claimed that the number of minima was the same as the number of known isotopes of these metals. They claimed that the method had many advantages among which were rapidity, non-destructiveness, extraordinary sensitivity, and low cost. The one drawback noted was that operation of the apparatus required considerable skill and experience (84). In fact, one great advantage that Allison mentions almost in passing at the end of his article is that his method is so sensitive that it could identify the presence of the long-searched-for eka-cesium, element 87, in samples of pollucite and lepidolite ores. Not only that, he was able to observe six minima for this new element – now virginium, (in honor of the state of his birth) Va (later changed to Vi), in his lexicon – that he ascribed to VaCl, VaNO3, Va2SO4 and six stable isotopes of the new element (85). 27 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

28

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Table 2. The Presumed Naturally Occurring Elements Remaining to Be Discovered in 1902 (82) Element No. Brauner

Element No. Scerri

Element Name

Date of Discovery

Discoverer(s)

Place of Discovery

43

43

Technetium (Tc)

1937

Carlo Perrier (1886-1948), Emilio Segrè (1905-89)

Palermo, Sicily

61

61

Prome-thium (Pm)

1945

Jacob A. Marinsky (1918-2005), Lawrence E. Glenden-in (1918-2008), Charles D. Coryell (1912-71)

Oak Ridge, TN, USA

72

72

Hafnium (Hf)

1923

Dirk Coster (1889-1950) and George de Hevesy (1885-1966)

Copenhagen

75

75

Rhenium

1925

Walter Noddack (1893-1960), Ida Tacke (1896-1978), Otto Berg (1874-1939)

Berlin

85

85

Astatine (At)

1940

Dale R. Corson (1914-2012), Kenneth R. MacKenzie (1912-2002), Emilio Segrè

Berkeley, CA, USA

87

87

Francium (Fr)

1939

Marguerite Perey (1909-75)

Paris

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Element No. Scerri

89 91

Element Name

Date of Discovery

Discoverer(s)

Place of Discovery

Actinium (Ac)

1899

André Debierne (1874-1949)

Paris

Protactinium (Pa)

1917

Lise Meitner (1878-1968), Otto Hahn (1879-1968), Kasimir Fajans (1887-1975)*

Berlin

* Note on the discovery of protactinium. Kasimir Fajans and Ostwald H. Göhring (1889-1915?) discovered a short-lived isotope 234Pa (T1/2 = 6.69 h) in 1913. The longest-lived isotope, 231Pa (T1/2 = 32,500 y) was discovered in 1917 by three independently working teams: Lise Meitner and Otto Hahn working in Berlin; Kasimir Fajans working at Karlsruhe; Frederick Soddy, John A. Cranston (1891-1972) and Alexander Fleck, Baron Fleck of Saltcoats (1889-1968) working at Glasgow.

29

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Element No. Brauner

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Figure 12. Fred Allison. Courtesy of the Auburn University Libraries Special Collections and Archives.

Figure 13 is a diagram of the magneto-optic apparatus (86). In a later publication with co-authors E. R. Bishop, A. L. Sommer, and J. H. Christensen, Allison opens the door to other possibilities for the minima observed, either to a cation with equivalent weight greater than that of Tl+, i.e., element 87, or to complex ions. After quickly eliminating the complex ions by experiment, the team concludes with an argument that seems like negative evidence (87): Since we have found nothing else which might give the minima attributed to compounds of 87 and since these minima persist after treatment with acids, bases, oxidizing and reducing agents, we conclude, as previously announced, that these minima are due to element 87. On a more positive note, they claim to have found 87 in sea water, California brine, Stassfurt kainite, crude cesium chloride, and monazite sand and samarskite from different sources. In all of these samples, the concentration of 87 was noted to be very low. 30

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Figure 13. Magneto-Optic Diagram and Connections. Light from the spark source G passes through the optical train from L to E to measure differential time lags of the Faraday effect in liquids. Reproduced from reference (85). Copyright 1932. American Chemical Society.

In the same month that the Alabama team claimed the discovery of element 87, they also announced the detection of element 85 by the same magneto-optic method, and quickly named it alabamine (named after the state and the Polytechnic Institute where Allison worked) with the symbol Am. After treating a large amount (100 pounds) of monazite mineral, they found the appropriate minima that showed that alabamine went into solution as peralabamic acid, HAmO4. After preparing alabamine as lithium alabamide, the research team found that alabamides are readily oxidized to hypoalabamite, alabamite, alabamate, and peralabamate salts and their corresponding acids. They also found that the peralabamates were the most stable form of the element, and that its estimated atomic weight was 221. However, no experimental evidence aside from the well-described magneto-optic method, was given (88). Departing from his usual appeal to the physics and chemistry research community, in 1933 Allison published an article in the Journal of Chemical Education (89) in which he also laid claim to having discovered element 85 and 87, among many other things, and even a heavy isotope of hydrogen a year before deuterium’s discovery by Harold Urey (1893-1981) (90). But already toward the end of 1931, Allison’s reputation was beginning to fray. The 1931 Annual Survey of American Chemistry, published in 1932, ventured the following remark (91): 31

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

[Allison] brought forward additional evidence that the number of minima in the time lag of the Faraday effect in electrolytic solution is equal to the number of isotopes of the positive ion. With some exceptions, this relationship seems to hold and the concentrations at which the minima first appear are reported to be approximately inversely proportional to the relative abundances of the isotopes…thus two isotopes of hydrogen, two of thallium and four of barium were predicted and these numbers subsequently proven to exist. But also, one chromium, two strontium and two ruthenium isotopes were required by these observations and subsequently four chromium, three strontium and six or seven ruthenium isotopes were reported. If the method is as sensitive as is claimed, these isotopes should not have been overlooked. Questions about the validity of Allison’s method and claims began to multiply, principally because of his insistence that only those adept and experienced in using the method could achieve the correct results. Furthermore, attempts to reproduce Allison’s published results met with failure. By 1934, largely at the instigation of Nobel Laureate Irving Langmuir (1881-1957), the American Chemical Society began to ban the submission of any articles experimentally based on the magneto-optic effect. Perhaps the literature death-blow to the method came in the December, 1934 issue of the Physical Review where Herbert G. MacPherson (1911-93), at Berkeley, published the results of his year-long, exhaustive study of the magneto-optic effect, ending with the words (92): “…it is clear that this objective test does not reveal the existence of any real minima on the apparatus used. The author believes…that such minima as [Allison] saw from time to time can be explained as due to physiological or psychological factors.” Even so, the false American elements illinium, alabamine, and virginium did not go away very soon. They were still being listed as “rare metals” in the 1940 Materials Handbook (93) and as late as 1956 in the eighth edition of this same publication (94)! And, as we have noted previously, the end-flap periodic tables in major textbooks and wall charts published by scientific supply houses continued to retain these elements as late as 1956 (Figure 14). Far from being discredited, even by a very sarcastic and damning lecture on the subject, called “pathological science” by Irving Langmuir in 1953 (95), Allison seemed to move from strength to strength. He was chair of the physics department of Auburn University from 1922 to 1953, from whence he departed to become chair of the physics and mathematics at the University of Texas at Austin for a fouryear stint. He moved on to Huntingdon College to chair that institution’s physics and mathematics department for another five years, retiring in 1961, although he continued to teach there throughout the decade of the 1960s. And everywhere he went, he was greatly esteemed and admired by faculty colleagues and students alike, so much so that the buildings housing physics at these institutions are named after him (96). 32 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Figure 14. A Periodic Table compiled by H. D. Hubbard and revised in 1941 for the W. M. Welch Manufacturing Co. of Chicago, Illinois. Note that, in addition to elements 85 and 87 (Ab and Vi) being present in this 1941 version, Il still occupies the space between Nd and Sm, and masurium (Ma) still claims to be element 43. Reproduced from reference (76). Copyright 1975. American Chemical Society.

Much later, Allison’s work was incorporated into a set of topics on “bad science” taught in the general chemistry and analytical chemistry at a liberal arts college. A number of cases, including Allison’s, were presented in a continuum of what the instructor termed bad, pseudo-, pathological, or deviant science. The author cites Allison’s magneto-optic method as something that started out as legitimate science and degenerated into bad science (97). In the 1993 Concise Columbia Encyclopaedia, Allison is still listed as the person who “announced” the existence of astatine (as alabamine), while the Berkeley research group of D. R. Corson, K. R. MacKenzie and E. Segrè are listed as its “producers (98).” However, in that same work, there is no question about who discovered element 87. Marguerite Perey is definitely given the credit she is due for discovering francium, element 87. However, the last sentence in the entry on p. 997 reads, “In the U.S., it was at one time called virginium.” So, despite their evanescent nature, both alabamine and virginium died hard.

Conclusion The three great advances that revealed the relationships and structures of the chemical elements were made over the space of time in which the “All-American” errors were made. None of them prevented the would-be discoverers from falling into error. Table 3 summarizes these erroneous claims. 33 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Table 3. Summary of the “All-American” Elemental Errors

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

“Chemical Element” (prior to the concept of Atomic Number)

“Physical Element” (Atomic Number)

Discoverer

Year

Symbol

Unnamed

--

H. A. Genth

1853

None

Unnamed

--

C. F. Chandler

1863

None

Mosandrium

--

J. L. Smith

1877

None

Rogerium

--

J. L. Smith

1879

None

Columbium

--

J. L. Smith

1879

Cb

Carolinium

--

C. Baskerville

1901

Cn

C. Baskerville

1904

Bz

Berzelium

Atomic No.

--

Illinium

B. S. Hopkins et al.

1926

Il

61

--

Alabamine

F. Allison

1930

Ab

85

--

Virginium

F. Allison

1930

Va, Vi, Vm

87

If we look a little closer at the fundamental reasons for these errors, we can discern quite a few: •



• •



claims based upon chemical tests and examination of chemical and physical properties of mixtures, not of pure samples (elements of Genth and Chandler); the difficulty of separating rare earth elements from one another due to their chemical similarity, plus lack of means to follow up on claims asserted (Smith); misconceptions about the nature of thorium earths, impure samples, incomplete spectral data, and poor experimental technique (Baskerville); excessive reliance on spectral lines of impure samples thought to be pure, and actually thinking to have seen something that was not actually there – wishful thinking – (Hopkins); use of an instrumental method that was entirely incapable of detecting less than nanogram quantities of material unless through delusion or fraud (Allison).

None of these would-be discoverers was ever actually ejected from the scientific community; in fact, some of them, like Fred Allison, went on to occupy high academic positions and retirement with honor – although he was the only one to receive censure from some sectors. 34 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

Examining all the cases in this paper, you will find a curious common denominator: not one of the authors of the alleged discoveries has ever issued a public retraction, thereby acknowledging and explaining the reasons and the causes of their (human) errors. The admission of one’s own error, when cultural and scientific conditions would require a corrected publication, would be a form of individual moral growth, but also a breakthrough for the scientific community, which at first was erroneously informed and therefore misguided. Unfortunately, in today’s overly competitive society, such an intellectual exercise would be counterproductive for a career. This may explain why some of these characters have ignored or forgotten their previous work. This could be the case, for example, for Baskerville who, after his dismal failures, changed his research interests drastically. Another exception can be made for Allison and Smith-Hopkins: they were acting in perfectly good faith; they believed blindly in their work and never wanted to consider the idea of having beaten the air instead of making a lasting contribution to scientific knowledge (29). A third consideration can be advanced as a justification of J. Lawrence Smith. He made his discoveries between 1877 and 1879, when his health was already shaky. Maybe he did not have time to write a proper retraction given the fact that he died not long afterwards. In any event, all of us can learn a great lesson from these errors: the history of science is never written in unwaveringly straight lines – there are always dead ends and backtracking along the way.

Acknowledgments The authors would like to acknowledge the contribution of Janan M. Hayes to the first portion of this paper on the role of American Chemical Society Presidents in reporting spurious elements. We refer to her paper, “Even ACS Presidents Announced the Discovery of New Elements, and They Were Wrong,” read at the ACS National Meeting, San Francisco, CA, August 11, 2014. The invaluable help of Vera Mainz (University of Illinois), Tammy Hartwell (Auburn University), and Eric Slater of the ACS Copyright Office in obtaining image permissions is deeply appreciated.

References 1. 2. 3.

4.

Boyle, R. The sceptical chymist. The classic 1661 text; Dover Publications: Mineola, NY, 2003; pp 2−3. Lavoisier, A.-L. Traité élémentaire de chimie; Chez Cuchet: Paris, France, 1789; p 5. Hendry, R. F. The elements and conceptual change. In The semantics and metaphysics of natural kinds; Beebee, H., Sabbarton-Leary, N., Eds.; Routledge: New York, 2010; pp 137−58. Lavoisier, A.; Berthollet, C. L. Méthode de nomenclature chimique; Chez Cuchet: Paris, France, 1787. 35

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

5.

6. 7. 8.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

9. 10. 11. 12. 13. 14.

15.

16. 17. 18. 19. 20. 21. 22. 23. 24.

Neither Germany nor Italy were countries in the modern sense of the word at the time being referred to; this designation is to indicate the region in this case. Klaproth, M. H. Vom Wassereisen. Crell’s chem. Annalen. 1784, 1, 390. Meyer, J. K. F. Versuche mit stolpener Basalt. Schriften der Gesellsch. Naturf. Freunde Berlin. 1780, 2, 334. Müller von Reichenstein, F. J. Versuche mit dem in der Grube Mariahilf in dem Gebirge Fazeby bey Zalathna vorkommenden vermeinten gediegenen Spiesglanzkönig. Physikalische Arbeiten der einträchtigen Freunde in Wien. 1783, 1, 63–69. Monnet, A. G. Sur la saturnite. J. Phys. (Paris) 1786, 28, 168. Klaproth, M. H. Analyse d’une terre verte trouvée dans la Nouvelle Prusse Orientale. Beschaft. Ges. Nat. Fr. Berlin. 1788, 8, 4. Klaproth, M. H. Kleine mineralogische Beiträge. Crell’s chem. Annalen. 1789, 6, 5. Extrait d’un memoire de Klaproth sur un nouveau metal nomme Tellurium. Ann. chim. phys. 1798, 25, 273–81. Abstract of a memoir of Klaproth on a new metal denominated tellurium. Nicholson’s J. 1798, 2, 372–6. Nov. 8, 1903, as reported by Baskerville, C. in “The elements: Verified and unverified.” Address to the Chemistry Section at the 53rd Meeting of the American Association for the Advancement of Science, Saint Louis, MO, 1903−1904; Proc. AAAS 1904, 387−442. Petrik, M. Friedrich August Genth aus Wächterbach und die Entdeckung der ersten NiO-Kristalle am Marburger chemische Institut unter Robert Wilhelm Bunsen; Cuvillier Verlag: Göttingen, Germany, 2013. The front cover of this book is depicted in Figure 2. The drawing on the upper left is from Hofmann, J. P. Das chemische Laboratorium der Ludwigs-Universität zu Gießen; C. F. Winter: Heidelberg, Germany, 1842; the two graphs on the bottom are from the text of the book cited. Genth, F. A. On a probably new element with iridosmine and platinum in California. Proc. Phila. Ac. Nat. Sci. 1852, 6, 209–10. Genth, F. A. Owenite. Am. J. Sci. 1853, 15, 246–8. Anon. Aktuelle Bekanntmachungen. Berg- und Hütten Ztg. 1853, 12, 751. Myers, W. M.; Zerfoss, S. Friedrich August Genth (1820-1893), chemist, mineralogist, collector. J. Franklin Inst. 1946, 241, 341–54. Chandler, C. F. A new metal in the platinum of Rogue River, Oregon. Am. J. Sci. 1862, 33, 351–52. Howe, J. L. The eighth group of the periodic system. Chem. News J. Ind. Sci. 1900July20, 82 (2121), 31-32–35. Bogert, M. T. Presentation Address (on the occasion of the Perkin Medal to Charles F. Chandler). J. Ind. Eng. Chem. 1920, 12, 184–189, 185. McKee, R. H.; Scott, C. E.; Young, C. B. F. The Chandler Chemical Museum at Columbia University. J. Chem. Educ. 1934, 11, 275–78. Havemeyer Hall at Columbia University. ACS National Historic Chemical Landmark. https://www.acs.org/content/acs/en/education/whatischemistry/ landmarks/havemeyerhall.html (accessed Sept 15, 2017). 36

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

25. Bogert, M. T. Biographical memoir of Charles Frederick Chandler (1836-1925). National Academy of Sciences Biographical Memoirs, Fifth Memoir; National Academy of Sciences: Washington, D.C., 1931; Vol. 14, pp 127−181. 26. Columbia University Archival Collections. Charles F. Chandler Papers, 1847−1937. http://www.columbia.edu/cu/lweb/archival/collections/ldpd_ 4078595/ (accessed Sept 15, 2017). 27. Strathern, P. Mendeleyev’s dream. The quest for the elements; St. Martin’s Press: New York, 2000; p 289. 28. Cannizzaro, S. Letter to Professor S. DeLuca: Sketch of a course of chemical philosophy. Il nuovo cimento. 1858, 7, 321–66. 29. Dmitriev, I. S. Scientific discovery in statu nascendi: The case of Dmitrii Mendeleev’s periodic law. Hist. Stud. Phys. Biol. Sci. 2004, 34 (2), 233–275. 30. Fontani, M.; Costa, M.; Orna, M. V. The lost elements: The periodic table’s shadow side; Oxford University Press: New York, 2015; pp xxiii−xxxiv. 31. Ede, A. The chemical element: A historical perspective; Greenwood Publishing Group: Westport, CT, 2006. 32. McFarland, B. A world from dust; Oxford University Press: New York, 2016. 33. Smith, J. L. Methods of analysing samarskite. Am. Chem. J. 1883, 5 (2), 73–81. 34. Smith, J. L. Note au sujet I’élément appelé mosandrum. C. R. Chim. 1878, 87, 831–34. 35. Baumgartner, E. Carl Auer von Welsbach: A pioneer in the industrial application of rare earths. In Episodes from the history of the rare earth elements; Evans, C. H., Ed.; Kluwer Academic Publishers: Dordrecht, 1996; pp 113−129. 36. Smith, J. L. Note sur une nouvelle terre du group du cerium et remarques sur une methode d’analyse des columbates naturels. C. R. Chim. 1878, 87, 146–48. 37. Wyckoff, W. C. U.S. National Academy. Nature 1879 December11, 21, 143–46. 38. Smith, J. L. Methods of analyzing samarskite and the other columbates containing earthy oxides by the agency of fluorhydric acid; and of dissolving columbite and tantalite by the same acid. On the separation of thoria from the other oxides. Quantitative estimation of didymium-oxide in its mixtures with other earthy oxides. Privately printed, Louisville, KY, 1881, pp 17. 39. Crookes, W. On radiant matter spectroscopy: The detection and wide distribution of yttrium. Philos. Trans. R. Soc. London 1883, 174, 891–918. 40. Smith, J. L. The composition and products of distillation of spermaceti, with some few remarks upon its oxidation by nitric acid. Am. J. Sci. 1842, 43, 301–21. 41. Smith, J. L. On the peculiar concretions occuring [sic] in meteoric irons. Am. J. Sci. 1883, 25, 417–23. 42. Silliman, B. Memoir of John Lawrence Smith (1818-1883). Read before the National Academy, April 17, 1884. National Academy of Sciences 37

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

43. 44. 45. 46. 47.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

48. 49. 50. 51.

52.

53. 54. 55. 56. 57. 58. 59. 60. 61. 62.

63.

Biographical Memoirs; National Academy Press: Washington, DC, 1886; Vol. 2, pp 217−48. Burke, J. G. Cosmic debris: Meteorites in history; University of California Press: Berkeley, CA, 1986; p 200. Report of Proceedings of the National Academy of Sciences during the Year 1882; National Academy Press: Washington, D.C., 1883; p 11. Seaborg, G. T.; Seaborg, E. Adventures in the atomic age: From Watts to Washington; Farrar, Straus & Giroux: New York, 2001; p 127. Baskerville, C. Obituary. J. Chem. Soc., Trans. 1923, 123, 3421–41. Baskerville, C. On the existence of a new element associated with thorium. J. Am. Chem. Soc. 1901, 23, 761–74, 772. Baskerville, C. Thorium; Carolinium, Berzelium. J. Am. Chem. Soc. 1904, 26, 922–942. Museum at CHF Home Page. https://www.chemheritage.org/museum-at-chf (accessed Sept 15, 2017). Lely, D.; Hamburger, L. Herstellung der Elemente Thorium, Uran, Zirkon und Titan. Z. Anorg. Chem. 1914, 87, 209–28. Steelman, B. L.; Baskerville, C. In Dictionary of North Carolina Biography; Powell, W. S., Ed.; University of North Carolina Press: Chapel Hill, NC, 1979; Vol. 1, pp 106−7. Van den Broek, A. Mendelejeffsche di Das ‘kubische’ Periodische Sistema der Elemente und die Einordnung der dieses Radioelemente System. Phys. Z. 1911, 12, 490–97. Van den Broek, A. The number of possible elements and Mendeléeff’s ‘cubic’ periodic system. Nature 1911, 87, 78. Moseley, H. G. J. The high frequency spectra of the elements. Philos. Mag. 1913, 1024–34. Scerri, E. A tale of 7 elements; Oxford University Press: New York, 2013. Soddy, F. Density of lead from Ceylon thorite. Nature 1915, 94, 615. Brauner, B. Ober die Stellung der Elemente der seltenen Erden im periodischensystem von Mendelejeff. Z. Anorg. Chem. 1902, 32, 1–30. Brauner, B. The new element of atomic number 61: Illinium. Nature 1926, 118 (2959), 84. Moseley, H. G. J. to James, C. James Collection UNH Archives, UA 8/3/1, S.1, B.1, F.11, May 27, 1914. Quoted in Reference 78. Rolla, L. Ricerche sopra l’elemento a numero atomico 61. Gazz. Chim. It. 1926, 56, 535–36. Ramsay, W. to James, C., James Collection UNH Archives, UA 8/3/1, S.1, B.1, F.6, February 26, 1912. Cited in Reference 78. Harris, J. A.; Hopkins, B. S. Observations on the rare earths XXIII. Element No. 61. Part One. Concentration and isolation in impure state. J. Am. Chem. Soc. 1926, 48, 1585–94. Harris, J. A.; Yntema, L. F.; Hopkins, B. S. Observations on the rare earths XXIII. Element No. 61. Part Two. X-ray analysis. J. Am. Chem. Soc. 1926, 48, 1594–98.

38 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

64. There is some question about the naming attribution – Hopkins was a great fan of the Illinois soccer team, the “Illini,” and they perhaps should indeed receive some credit. 65. Siegbahn, M.; Friman, E. Vacuum spectrograph for high-frequency spectra investigations. Applications to the study of the rare earths. Physik. Z. 1916, 17, 176–78. 66. Cork, J. M.; James, C.; Fogg, H. C. The Concentration and identification of the element of atomic number 61. Proc. Natl. Acad. Sci. U.S.A. 1926, 12, 696–99. 67. Rolla, L.; Fernandes, L. Ricerche sopra l’elemento a numero atomico 61. Nota I. Gazz. Chim. It. 1926, 56, 435–36. 68. Rolla, L.; Fernandes, L. Ricerche sopra l’elemento a numero atomico 61. Nota II. Gazz. Chim. It. 1926, 56, 688–94. 69. Rolla, L.; Fernandes, L. Ricerche sopra l’elemento a numero atomico 61. Nota III. Gazz. Chim. It. 1926, 56, 862–64. 70. Noyes, W. A. Florentium or Illinium? Nature 1927, 120, 14. 71. Noyes, W. A. Illinium. Nature. 1927, 119, 319. 72. Rolla, L.; Fernandes, L. Florentium or Illinium? Nature. 1927, 119, 637–38. 73. Noyes, W. A. Florentium or Illinium? Nature 1927, 120, 14. 74. Gatterer, A.; Junkes, J.; Rolla, L.; Piccardi, G. Sugli spettri d’arco delle miscele neodimio-samarifere. Pontif. Acad. Sci., Comment. 1942, 6, 387–413. 75. The Internet Database of Periodic Tables. http://www.meta-synthesis.com/ webbook/35_pt/pt_database.php?PT_id=550 (accessed Sept 15, 2017). 76. Hopkins, B. S. Building blocks of the universe. Sci. Am. 1927, 136 (2), 87–89. 77. Trimble, R. F. What happened to alabamine, virginium and illinium? J. Chem. Educ. 1975, 52, 585. 78. Marinsky, J. A.; Glendenin, L. E.; Coryell, C. D. The chemical identification of isotopes of neodymium and of element 61. J. Am. Chem. Soc. 1947, 69, 2781–85. 79. Meshik, A. P. The workings of an ancient nuclear reactor. Sci. Am. [Online] 2005, 293, 83−91. http://www.scientificamerican.com/article/ancientnuclear-reactor/ (accessed Sept 15, 2017) 80. Attrep, M.; Kuroda, P. K. Promethium in pitchblende. J. Inorg. Nucl. Chem. 1968, 30, 699–703. 81. Murphy, C.; Charles James, B. Smith Hopkins and the tangled web of element 61. Bull. Hist. Chem. 2006, 31 (1), 9–18. 82. Data gleaned from Emsley, J. The Elements, 3rd ed.; Clarendon Press: Oxford, U.K., 1998. 83. Allison, F. The effect of wave-length on the differences in the lags of the Faraday effect behind the magnetic field for various liquids. Phys. Rev. 1927, 30, 66–70. 84. Allison, F.; Murphy, E. J. A magneto-optic method of chemical analysis. J. Am. Chem. Soc. 1930, 52, 3796–3806. 85. Allison, F.; Murphy, E. J. Evidence of the presence of element 87 in samples of pollucite and lepidolite ores. Phys. Rev. 1930, 35, 285letter. 39

Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF HELSINKI on December 20, 2017 | http://pubs.acs.org Publication Date (Web): October 30, 2017 | doi: 10.1021/bk-2017-1263.ch001

86. Allison, F. Magneto-optic method of analysis as a new research tool. Ind. Eng. Chem., Anal. Ed. 1932, 4 (1), 9–12. 87. Allison, F.; Bishop, E. R.; Sommer, A. L.; Christensen, J. H. Further research on element 87. J. Am. Chem. Soc. 1932, 54, 613–15. 88. Allison, F.; Bishop, E. R.; Sommer, A. L. Concentration, acids and lithium salts of element 85. J. Am. Chem. Soc. 1932, 54, 616–20. 89. Allison, F. The magneto-optic method of analysis, with particular reference to the detection of elements 85 (alabamine) and 87 (virginium). J. Chem. Educ. 1933, 10, 71–78. 90. Urey, H.; Brickwedde, F.; Murphy, G. A hydrogen isotope of mass 2. Phys. Rev. 1932, 39, 164–66. 91. Annual survey of american chemistry, 1931; West, C. J., Ed.; National Research Council; Chemical Catalog Co.: New York, 1932; Vol. VI, p 40. 92. MacPherson, H. G. An investigation of the magneto-optic method of chemical analysis. Phys. Rev. 1935, 47, 310–17. 93. Brady, G. S. Materials handbook: An encyclopedia for purchasing agents, engineers, executives, and foremen, 4th ed.; McGraw-Hill Book Co.: New York, 1940; p 397. 94. Brady, G. S. Materials Handbook: An Encyclopedia for purchasing agents, engineers, executives, and foremen, 8th ed.; McGraw-Hill Book Co.: New York, 1956; p 666. 95. Langmuir, I. Pathological science: the science of things that aren’t so. Reprinted in Scientific Work and Creativity. Advice from the Masters; A presentation of the citizen scientists league. Smith, R. D., Ed.; Citizen Scientists League: Clearwater, FL, 2012; Vol. 1, pp 363−91. 96. Kauffman, G. B.; Adloff, J.-P. Fred Allison’s magneto-optic search for elements 85 and 87. The Chemical Educator. 2008, 13, 358–64. 97. Epstein, M. Using bad science to teach good chemistry. J. Chem. Educ. 1998, 75, 1399–1404. 98. The concise Columbia encyclopaedia, 5th ed.; Chernow, B., Vallaski, G., Eds.; Columbia University Press: New York, 1993; p 168.

40 Benvenuto and Williamson; Elements Old and New: Discoveries, Developments, Challenges, and Environmental Implications ACS Symposium Series; American Chemical Society: Washington, DC, 2017.