bulletin for the history of...

68
BULLETIN FOR THE HISTORY OF CHEMISTRY Division of the History of Chemistry of the American Chemical Society VOLUME 26, Number 2 2001 John W. Draper, ACS President, 1876

Upload: others

Post on 21-May-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

BULLETIN FOR THE HISTORYOF CHEMISTRY

Division of the History of Chemistry of the American Chemical Society

VOLUME 26, Number 2 2001

John W. Draper, ACS President, 1876

Page 2: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

The Cover... The first president of the 125-year-old American Chemical Society, an organizationwhose founding was met with strong opposition. See Bohning’s article p 92.

BULLETIN FOR THE HISTORY OF CHEMISTRY

VOLUME 26, CONTENTS

Number 1

AARON JOHN IHDE 1909 – 2000 1AARON IHDE: A LIFE FROM BASCOM’S HILL 3James J. Bohning, Lehigh University

AARON IHDE AND HIS STUDENTS (1) 15Alan J. Rocke, Case Western Reserve University

AARON IHDE’S CONTRIBUTIONS TO THE HISTORY OF CHEMISTRY 24William B. Jensen, University of Cincinnati

MEMORIES OF AARON IHDE 33Robert Siegfried, University of Wisconsin

CRITERIA FOR GENEALOGICAL ROOTS (1) 36Aaron J. Ihde

THE HISTORY OF OZONE. THE SCHÖNBEIN PERIOD, 1839-1868 40Mordecai B. Rubin, Technion-Israel Institute of Technology

THE CONTRIBUTIONS OF PAYEN AND LABILLARDIÈRETO THE DEVELOPMENT OF COLORIMETRY 57Lluís-Garrigós Oltra, Carles-Millan Verdú, and Georgina-Blanes Nadal,Escola Politècnica Superior d’Alcoi

THE BIRTH OF TITRIMETRY: WILLIAM LEWIS AND THEANALYSIS OF AMERICAN POTASHES 66Frederick G. Page, University of Leicester.

BOOK REVIEWS 73

Number 2THE 2000 DEXTER AWARD ADDRESS 81CELEBRITY CULTURE IN PARISIAN CHEMISTRYAlan J. Rocke, Case Western University

OPPOSITION TO THE FORMATION OF THE AMERICAN CHEMICAL SOCIETY 92James J. Bohning, Lehigh University

NIEPCE DE SAINT-VICTOR AND THE DISCOVERY OF RADIOACTIVITY 104Fathi Habashi, Laval University, Quebec, Canada

THE HARE-CLARKE CONTROVERSY OVER THE INVENTIONOF THE IMPROVED GAS BLOWPIPE 106Martin D. Saltzman, Providence College

THE CONTRIBUTIONS OF CYRUS MOORS WARRENTO THE ANALYSIS OF HYDROCARBONS 112Martin D. Saltzman, Providence College

RECIPROCAL SOLUBILITY INFLUENCE IN SALT MIXTURES.THE CONTRIBUTIONS OF WALTHER NERNST AND OF ARTHUR NOYES 118John T. Stock, University of Connecticut

A PHILOSOPHICAL COMMENTARY ON GIUNTA’S CRITIQUE OFNEWLANDS’ CLASSIFICATION OF THE ELEMENTS 124Eric R. Scerri, University of California, Los Angeles

A RESPONSE TO SCERRI’S COMMENTARY 130Carmen J. Giunta, Le Moyne College

BOOK REVIEWS 134

Page 3: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 81

THE 2000 DEXTER AWARD ADDRESS

CELEBRITY CULTURE IN PARISIANCHEMISTRY

Alan J. Rocke, Case Western University

“Hard as he tried, [Nobelist] Murray Gell-Mann couldnever make himself into a legend like his rakish col-league and collaborator, Richard Feynman—even if hewas the greater physicist.” Thusbegins an article in the July2000 issue of the AtlanticMonthly. The author notes that(1):

..there are other factors [be-sides top honors] that count inthe manufacture of fame. Gell-Mann knew how to packageideas, and he had a knack forgiving whimsical and unfor-gettable names to the most ab-stract concepts in science.Feynman had a more vital gift:he knew how to package him-self.

Celebrityhood in science is, ofcourse, nothing new. The earli-est obituarist of the Frenchchemist Jules Pelouze (1807-1867) began his article by prais-ing those who successfully raisethemselves from modest backgrounds (2):

Even if some allow themselves to be defeatedthrough lack of perseverance, there are others, and morethan one would think, who attain prosperity, fortune,and, what is perhaps even preferable to these, celebrity.… This is the case with the man whom we have justlost, and whom I make bold to call my pupil; for, issu-

ing from my apothecary laboratory, he grew, found fame,and became a celebrity.

The irony in this particular in-stance is that in spite of an ex-tremely successful and influentialworking life, and in spite of thisputative celebrityhood in his ownday, Pelouze has been almost com-pletely ignored by posterity. Otherthan five obituaries published at thetime of his death, there is virtuallyno secondary literature on Pelouze.To make matters worse, the nine-teenth-century obituaries contra-dict each other in some of the im-portant dates, and collectively omitsome others. Little wonder that thepoor writer of the Pelouze articlein the Dictionary of Scientific Bi-ography ended up by being vagueand was sometimes in error (3).

I want to explore here what itmeant to be a “celebrity” in nine-

teenth-century French chemistry, how this status can becorrelated with historical significance as measured byposthumous historiography, and what this all meant (andmeans) in a broader perspective. Confirming the ear-lier work of Robert Fox and others, we will see that inOrleanist and Second Empire France there developed adistinct culture—or even cult—of savant-celebrities,which tended to replace the institution of so-called “no-

Alan J. Rocke

Page 4: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

82 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

tables” of the First Empire and Restoration. This celeb-rity culture may have played a significant role in shap-ing nineteenth-century French science, and not alwaysfor the best. To instantiate the discussion, I use twopairs of French chemists. The first pair is Pelouze and aslightly older and much better known contemporary,Jean-Baptiste Dumas (1800-1884); the second part willfocus on Adolphe Wurtz (1817-1884) and MarcellinBerthelot (1827-1907).

Pelouze and Dumas

Born in the provinces, Pelouze arrived in Paris in 1825to take his apprenticeship in pharmacy (4). He wasdesperately poor, living in a garret and surviving at timeson bread and water. About 1827 a chance encounterdetermined his future. Caught in a driving rain walkingfrom Charenton toward Paris,Pelouze flagged down a passingcab, not realizing that it already helda passenger who had hired the driverto take him back to Paris. The driverwas reluctant to stop until the pas-senger, the eminent J. L. Gay-Lussac, insisted on taking the youngman in. The two chemists foundcommon interests during the trip,and Gay-Lussac immediately of-fered Pelouze a place in his labora-tory in the Paris Arsenal (5). Thiswas only the second time that Gay-Lussac had taken a private student;the first had been the young JustusLiebig, in 1824. Pelouze and Liebigfirst met when Liebig returned toParis for a few weeks in 1828 andvisited his mentor’s laboratory.Liebig and Pelouze became fastfriends.

In 1831 Gay-Lussac hired Pelouze as his répétiteurat the École Polytechnique. Here Pelouze found him-self in daily contact with Dumas, who had beenrépétiteur for Gay-Lussac’s colleague L. J. Thenardsince 1824. For the next five years, Dumas and Pelouzeoccupied neighboring laboratory benches at the ÉcolePolytechnique. Both also corresponded regularly withLiebig. In the summer of 1836 Pelouze traveled toGiessen to work with Liebig and in subsequent yearsacted essentially as Liebig’s agent in Paris. In their cor-respondence with Liebig, Pelouze and Dumas both of-fered high praise of the rising German chemical com-

munity, often using virtually the same language; theyregularly disparaged Parisian academic chemistry.

They also lamented their lack of facilities. Liebignot only had a fine laboratory and a growing group ofPraktikanten, but his residence was conveniently up-stairs from the laboratory, and there were no urban dis-tractions in the tiny town of Giessen. He could virtu-ally live in his laboratory, devoting himself heart andsoul to his research and his students. By contrast, life inParis was immensely complicated. Unlike Liebig,Pelouze and Dumas had no laboratories in their resi-dences, and each had two workplaces (aside from theÉcole Polytechnique, Dumas was also professor at theÉcole Centrale des Arts et Manufactures; and from 1834Pelouze also worked as an assayer at the Paris Mint).Although this practice of multiple positions (called

“cumul”) prevented scholars fromstarving, it created logistical havocwith their daily lives. Moreover,chemical laboratories in Parisianacademic institutions were sadlydeficient, and most were scandal-ously bad.

In 1836 Dumas was promotedto professor at the Polytechnique,and simultaneously, withThenard’s help, he becameprofesseur adjoint at the Sorbonne.Two years later, a vacancy at theFaculté de Médecine was an-nounced. Dumas wrote Liebig totell him that he had no wish to ap-ply for this position, since he al-ready held professorships at threedifferent institutions. But theSorbonne had no laboratories at all,and his ill-heated Polytechnique

laboratory had become virtually uninhabitable duringthe Paris cold wave of January 1838. Dumas was des-perately unhappy. He formed a new plan (as he subse-quently explained it to Liebig): to win the medical pro-fessorship, resign the Polytechique, and then use theresulting higher income to run a private laboratory, whichhe could also use for his personal scientific research.Connected with this plan, Dumas’s wealthy father-in-law, Alexandre Brongniart, generously built him a housefor the laboratory on the Rue Cuvier (6). Dumas wroteLiebig (7):

Since I came to Paris, I have been seeking a way tocreate a laboratory broadly constituted under my di-

Jules Pelouze, Oesper Collection

Page 5: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 83

rection. I think I have finally succeeded in this, andthat gives me some consolation. In two or threemonths I will be able to put ten selected students towork in my house, and I will be able to devote fouror five thousand francs per year to their experiments.Only then will I be in a position to resume my ex-periments in competition with yours. At the momentI can’t keep pace with you.

The Rue Cuvier laboratory operated for a total of tenyears, financed entirely from his own pocket, beforeDumas was forced to close it in the chaotic conditionsof the Revolution of 1848. However, in this decade theDumas laboratory played a major role in the history of19th-century French chemistry.

Many years later, in his éloge for Pelouze, Dumascommented that their destinies had been closely linkedfor forty years, and there is much truth to the statement(8). Dumas had been Thenard’s répétiteur at the ÉcolePolytechnique, while Pelouze was performing the samefunction for Gay-Lussac. Just as Gay-Lussac andThenard, once close friends, had fallen out, so also theirrespective protégés became unfriendly rivals. Both menalso taught at the École Centrale and both also providedsubstitute lectures for Thenard at the Collège de France.Both were elected to the Académie des Sciences at anearly age (Dumas at 32, Pelouze at 30). In their parallelearly letters to Liebig, both professed abhorrence of thepolitics of cumul, and pledged not to engage in suchbehavior (9). Of course, both later did just that. WhenDumas took the professorship at the Faculté deMédecine, Pelouze succeeded Dumas as professor at theÉcole Polytechnique. Just as Dumas used his new pro-fessorship to enable him to open his private laboratoryin the Rue Cuvier near his residence, Pelouze simulta-neously used his new professorship at the Polytechniqueto enable him to construct a private teaching and re-search laboratory adjacent to his official residence inthe Rue Guénégaud at the Mint (10). Pelouze’s andDumas’s letters to Liebig leave no doubt that they wereboth consciously following the model of Liebig’sGiessen laboratory.

Many fine chemists were trained in Pelouze’s RueGuénégaud laboratory, or pursued their own originalresearch there, including Claude Bernard and CharlesGerhardt. Seven years after opening this laboratory,Pelouze closed it in order to create a much larger pri-vate laboratory school in the Rue Dauphine, near theMint (and the Seine). This expanded enterprise, lastingtwelve years, was highly successful. He had enrollmentsof around thirty at a time, consisting mostly of young

men from the provinces and from abroad, preparing forfuture roles in their families’ chemical businesses.Pelouze’s enterprise could only have benefited whenDumas was forced to close his own laboratory in thespring of 1848, and Gay-Lussac, his patron, was inducedto retire (11). In mid-career, Pelouze became succes-sively professor at the Collège de France (1846), presi-dent of the Commission of the Mints of France (1848),and chief consultant of the great state-chartered Saint-Gobain chemical works (1850). With these positions,together with his private chemical laboratory school,which operated in two successive locations for nearlytwenty years without much competition (12), Pelouzehad more wealth and influence than he could havedreamed (13).

Nevertheless, Dumas far exceeded Pelouze in ca-reer success. Like Pelouze, as a young man from theprovinces Dumas had arrived in Paris without means;and like Pelouze he had also risen quickly. As alreadymentioned, in 1838 Dumas traded his professorship atthe École Polytechnique for the Faculté de Médecine.Upon Thenard’s move to the vice-presidency of theConseil Royal de l’Instruction Publique in 1841, Dumaswas promoted from professeur adjoint to professeur atthe Sorbonne; and the following year he succeededThenard as dean of the Faculté des Sciences. Thenard,

raised to thebaronage in1825, becamethe de factoacademic czarof France inthe 1840s andrelied heavilythereafter onhis protégé’srecommenda-tions for chairsand promo-tions. In thisway Dumasbecame themost powerfulacademic sci-entist inO r l e a n i s t

France. Parisian chemical students began to refer toDumas jocularly at this time as “l’être suprême (14).”His private laboratory became the French analogue toLiebig’s.

Jean-Baptiste Dumas

Page 6: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

84 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

Politically, Dumas was a center-right conservative,and he had a few queasy moments during the brief Sec-ond Republic. However, during the Second Empire hebecame even more powerful than he had been in theJuly Republic. For a few years he was Napoleon III’sMinister of Agriculture, then was appointed Senator, andalso was Inspector General of Higher Education—allwhile retaining his two professorships. After Pelouzedied, he gave up the Inspector General position and hisSorbonne professorship in order to become Pelouze’ssuccessor as head of the Mint (a time-consuming butvery lucrative post). Finally, in 1867 Dumas was namedSecrétaire Perpétuel of the Académie des Sciences, themost exalted honor in the French scientific community.

Hence, despite all the parallels between these twomen, and contrary to the obituarist’s claims with whichI began this paper, Pelouze was no savant-celebrity, andDumas most certainly was. What did it mean to be acelebrity? Scholarly renown was the first criterion, andit is true that Dumas’s scientific work, especially in hisearly career, far outshone Pelouze’s. However, wealth,social connections, and (what is more important) thehabitual exercise of influence were certainly a part ofthe mix. The mature Pelouze was modest and retiring,just as he had been in his youth, while the mature Dumaswas proud, confident, expansive, and powerful. Pelouzehad reputation and position, but Dumas was both worldfamous and locally powerful in a way that Pelouze neverbecame. Aside from his very real merit as one of thefinest scientists of his generation, Dumas had alwaysbeen careful to cultivate celebrityhood. Pelouze neverdid.

Wurtz and Berthelot

Let us now move forward a half-generation and take upthe case of Adolphe Wurtz and the slightly youngerMarcellin Berthelot. Wurtz’s background was middle-class and provincial (Alsatian), and he studied under bothLiebig and Dumas. Dumas’s support was important ingaining Wurtz his principal professorial chair, that oforganic chemistry in the Faculté de Médecine in Paris(1853). Later he added a second position, at theSorbonne.

Unlike all my other protagonists, Berthelot grewup in Paris, the son of a physician. He was successivelyPelouze’s préparateur in his private laboratory school,then Antoine Balard’s préparateur at the Collège deFrance; and he gained both men’s patronage. In 1859he was awarded a new chair of organic chemistry at the

École de Pharmacie. Then, a group of influential in-triguers succeeded in engineering the creation of a newchair of organic chemistry at the Collège de France; andin 1865 they managed to have Berthelot installed in it.Thus, it can be said that both Wurtz and Berthelot werecumulards, but to a much lesser extent than Dumas andPelouze had been.

Wurtz and Berthelot had an uneven personal rela-tionship right from the beginning. Part of the conflictwas religious and cultural, for Berthelot was an atheistand materialist whereas Wurtz was an idealist and a de-vout Lutheran. Mostly, however, they clashed overchemical theory. Berthelot was powerfully influencedby Pelouze’s anti-theoretical attitudes, which Pelouzehad imbibed from his own mentor, Gay-Lussac. Wurtz,on the other hand, adopted Dumas’s and Liebig’s strongorientation toward theory. In 1854 Wurtz embraced theatomic-molecular reforms in chemistry that had beenadvocated by such chemists as Auguste Laurent, CharlesGerhardt, Alexander Williamson, and August Kekulé.For the remaining thirty years of his life, Wurtz was theprincipal French advocate of atomistic chemistry, butthose who regarded themselves as anti-atomists—led byBerthelot in the first instance—were successful in op-posing him. The Gerhardt-Laurent reforms, rapidlyadopted in other European countries, especially in Ger-many and after 1860, failed to win acceptance in Franceuntil near the end of the century.

As already mentioned, between 1853 and 1865 threenew professorships dedicated to organic chemistry wereestablished in Paris. However, none was used to pro-mote the Gerhardt-Laurent reforms or the emergenttheory of chemical structure. The chairs at the École dePharmacie and Faculté de Médecine, held by Berthelotand Wurtz respectively, had practice-oriented pedagogi-cal restrictions that excluded any systematic teaching—or even extended discussion—of chemical theory. TheCollège de France, on the other hand, was explicitlydevoted to pure scientific research, including advancedtheory; but molecular theory was the last thing eitherBerthelot or Balard wanted to consider. Wurtz natu-rally found this situation frustrating, particularly sincethe chemical reforms did not seem to be making muchheadway in France. His chair at the Faculté de Médecinegave him a secure professional position and his teach-ing/research laboratory was popular and successful; butwhat he dearly desired was a rhetorical platform fromwhich he could make the case for the new chemistry toa wider circle. His efforts met with mixed results, atbest.

Page 7: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 85

By any measure, Wurtz was one of the greatestFrench scientists of the nineteenth century; his researchspanned the entire science of chemistry and was notablefor its volume, significance, and influence. Nonethe-less, Wurtz’s election to the Académie des Sciences wasdelayed, largely because the number of members wasfixed, and therehappened to be adearth of deathsin the Académieafter the Dumas/Pelouze genera-tion. The firstelection in theA c a d é m i e ’ sChemistry Sec-tion afterWurtz’s arrivalin Paris was in1857, afterBaron Thenarddied. Wurtz wasdeeply disap-pointed thatEdmond Frémy(professor at theM u s é u md ’ H i s t o i r eNaturelle) waschosen in prefer-ence to himself. There were reasons for this unrelatedto intrinsic merit: not only was Frémy senior to Wurtz,but there can also be little doubt that Frémy’s mentorand patron Pelouze had pulled some strings.

The next vacancy in the Section occurred ten yearslater, when Pelouze died. Wurtz easily won the elec-tion, by a vote of 46 to 3 over Berthelot. The same year,an anonymous ministerial report addressed to the em-peror assessing candidates for Dumas’s vacated officeof Inspecteur Général de l’Enseignement Supérieur wasfrankly critical of Berthelot. The most eminent chem-ists in France, the writer declared, were clearly Pasteur,Deville, and Wurtz, and Balard was senior to all of them(Balard was chosen) (15). In the following year anothervacancy was created in the Académie’s Chemistry Sec-tion, upon Dumas’s elevation to Secrétaire Perpétuel.For a third time Berthelot was a candidate, and for athird time he lost, this time to Auguste Cahours (Pasteurand Deville were already members of the academy).Berthelot finally won election to the Académie in the

Physics Section, when a vacancy occurred in 1873. Eventhen he was third on the Section’s nomination list, tiedwith nine others; nonetheless, he prevailed in the gen-eral election (16).

All this suggests that Wurtz’s research standingexceeded Berthelot’s, but Wurtz’s contempo-rary renown did not translate to posterity.There are no Wurtz memorials in Paris, anduntil the last eight years there was only a littlemore Wurtz historiography than that devotedto Pelouze. Marcellin Berthelot, by contrast,lived and died larger than life. The fiftiethanniversary of his first publication (1901) wascelebrated by 3,000 invited guests in the GreatHall of the new Sorbonne. His state funeralsix years later was marked by speeches by thePresident and Prime Minister of the Republicand the presence of hundreds of other digni-taries. Interment, for both Berthelot and hiswife, was in no less elevated a place than thePanthéon—an unprecedented honor madepossible only by a special legislative act. His

statue wase r e c t e dp r o m i -nently inthe Ruedes Écoles,in thesquare op-posite theCollège deFrance, asquare thatbears hisname to-day. Thecentenaryof his birthin 1927was the oc-casion ofelaboratec e l e b r a -

tions, including the preparation of a sumptuous com-memorative volume. No fewer than fifty schools inFrance are now named in his honor (17).

Wurtz was not a star in the French scientific firma-ment. Berthelot was. So was Pasteur, for whom aneponymous research institute was created long before

Marcellin Berthelot

Adolphe Wurtz

Page 8: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

86 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

his death. So was Claude Bernard, in whose honor theFrench government conducted the first state funeral forany scientist (1878). In an earlier generation, so wereThenard, Gay-Lussac, Dumas, Arago, and Cuvier. Whatmade a scientist into a celebrity? It was not just con-temporary research renown, for Wurtz had that; and con-versely Thenard fell far short of the research productiv-ity of Gay-Lussac, Dumas, or Wurtz. Rather, what oneneeded in addition was a collection of certain other hu-man qualities, including the ability successfully to pro-mote oneself.

Three qualifications are in order. I do not wish toimply that Berthelot’s scientific work was weak, or thathe was undeserving of his fame. Partington wrote, ac-curately, that (18):

Berthelot’s work is astonishing in its volume, origi-nality, and importance. … There must be few chem-ists of my generation, whatever their interests, whohave not more than once turned up Berthelot’s publi-cations.

What I do want to affirm, however, is the judgment thatthe great advantage Berthelot enjoyed over Wurtz, bothin reputation among the literate lay public in his ownday (though not in specialist collegial circles!), and inposthumous recognition, was not in proper proportionto their respective scientific merits. I believe, on thecontrary, that Wurtz’s lifetime scientific accomplish-ments must be judged as significantly greater than hisrival’s; moreover, I believe that the majority of expertauthorities contemporary with Wurtz and Berthelotagreed with this judgment. Jean Jacques’s recent de-mythologizing biography of Berthelot (16) is generallyon the mark, in my view.

My second qualification is the following. In par-tially ascribing this unjust disproportion to what I amcalling a “celebrity culture,” I do not mean to imply thatscientific prowess was unimportant to contemporaryParisian opinion, or that other countries refused to en-gage in hero worship. The “research ethic” was not aGerman monopoly; the route to success in 19th-centuryFrench science was always through research. In par-ticular, Berthelot could not have gained the heights with-out a superb research record. Moreover, in my first casestudy, that of Dumas and Pelouze, I hope that I havemade my opinion clear that Dumas fully deserved hisfame, and also that he deserved to be more celebratedthan Pelouze. One might certainly argue that the othersavant-celebrities of nineteenth-century Parisian culture,such as Gay-Lussac, Thenard, Cuvier, Claude Bernard,and Pasteur, were also deserving of their “star” status,

on substantive grounds. In fact, the argument I want tomake does not concern so much the fate of individuals,but rather the fate of the national community.

Finally, one might legitimately suggest that the abil-ity to create excitement about one’s scientific contribu-tions and to persuade others of their importance is anessential element of what it means to practice sciencesuccessfully. In writing of the “private science” of LouisPasteur, Gerald Geison (19) rightly notes that “past sci-entists are great insofar as they persuaded their peers toadopt their ideas and techniques..” However, he addsthe important qualification that it is also necessary that“those ideas and techniques [be] fertile in the investiga-tion and resolution of important research problems.”Geison affirms that Pasteur’s work fully meets both cri-teria, and therefore “he deserves his reputation as oneof the greatest scientists who ever lived.” The work ofthe celebrities Richard Feynman and Jean-BaptisteDumas met Geison’s second criterion by universal agree-ment. Such agreement is less unanimous in certain othercases—Thenard, Balard, and Berthelot, for example.

Context and Consequences ofCelebrity Culture

In many respects, 19th-century Parisian science workedwell. In chemistry, there was a galaxy of talent in themiddle decades of the century. After the retirements ofGay-Lussac and Thenard, students could attend the lec-tures of Regnault, Cahours, or Frémy at the ÉcolePolytechnique, or those of Dumas and Balard at theSorbonne. The latter men were succeeded by Pasteurand Deville, and then Wurtz was added, as well. TheÉcole Normale also boasted Pasteur and Deville; and atthe Collège de France were Pelouze, Balard, Regnault,and Berthelot. Frémy and Chevreul held the two chemi-cal chairs at the Muséum d’Histoire Naturelle, andDumas and then Wurtz were professors at the Facultéde Médecine. Even this is not a full list of the chemicaltalent available in Paris, for private laboratory schoolswere run at various times by Gerhardt, Laurent, Wurtz,and Pelouze; and the private École Centrale des Arts etManufactures had considerable importance.

Furthermore, in many respects opportunity for ad-vancement ran on admirably egalitarian principles, alegacy of Napoleonic reforms. Admission to the grandesécoles was by competitive examination, and the univer-sity system was essentially free and open. During theJuly Monarchy, professorships in the French universitysystem were awarded as the result of complex—and

Page 9: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 87

more-or-less impartial—competitions. The most pow-erful scientific institution of all was the Académie desSciences, which has been accused, from the 1830s on,of operating as an oligarchic clique to control access tocareers and political power in the scientific community.There is truth to this criticism, but Maurice Croslandhas also rightly emphasized the essentially meritocraticnature of admission to the Académie, and its emphasison research renown (20).

The system worked, in its own fashion, and muchevidence points to its health. The fact that success waspossible for poor youths such as Dumas and Pelouze,and for others from lower middle-class backgrounds suchas Berthelot, Wurtz, and Pasteur, is an obvious positiveindicator. Politics was ever-present, of course, but theardent republicans Pelouze, Arago, and Berthelot, forinstance, succeeded in relatively conservative times.Those who exercised the greatest influence, such as Gay-Lussac, Thenard, and Dumas, usually were able to dis-cern talent and excellence in the next generation, and inmost cases acted appropriately to promote careers. (Theunfortunate cases of Gerhardt and Laurent, ostracizedby the powerful in the 1840s and early 1850s, are strik-ing, but fortunately somewhat anomalous.) The qualityof research produced by the French chemical commu-nity in the middle years of the century was excellent.

Nonetheless, there were pathologies in the system,and some of these had unfortunate consequences. Oneof these is the characteristic upon which I focus here, aculture of celebrity. Early in the century, influence inFrance was exercised predominantly by a small privi-leged elite denominated by the word “notables (21).”These men used old aristocratic connections, social po-sition, and wealth to exercise political power; they in-cluded a number of scientists—such as Berthollet,Laplace, and Cuvier. In the middle decades of the cen-tury, this structure yielded in the broadest terms to thegrowing power of the bourgeoisie and, in science, to amovement toward increasing professionalization. In thelate 19th century, power was no longer wielded by no-tables distinguished largely by their social standing, butrather, in the scholarly world, by top university research-ers, academicians, and high-level bureaucrats. Thischange has been well studied, at least in its broad out-lines. What I want to suggest, however, is that certaincultural attitudes survived this socio-economic shift. Inthe following generation, celebrities now played someof the cultural functions that notables had earlier. Thenotability of the beginning of the century had been con-

ferred principally by social position and wealth; by themiddle of the century the first criterion was fame.

This issue is closely related to what Robert Fox hasreferred to as a “radical change” in French cultural lifein the decade or two after Napoleon’s fall (22). Foxobserves that a new ”declamatory” style of higher edu-cation, where oratorical and dramatic effects were em-phasized often at the expense of serious treatment ofdifficult issues, became fashionable in RestorationFrench culture. This new style was particularly visibleat the Sorbonne and the Collège de France, where hugeaudiences consisting mostly of interested laypeople wereattracted to lectures in all fields of scholarship. Thissituation developed partly because there was no atten-dance requirement for registered students, and contrari-wise there was much interest in elevated subject matteramong the educated public. Education became a vari-ety of theater; one had to come early to get a seat, spec-tators expected to be entertained, and professors hopedfor applause. These practices were already noticeable(and commented upon as novel) in the 1820s; by the1830s a few worried voices were raised, and by the 1860smany reformers viewed them as deleterious to the fu-ture of French science. The very public meetings of theAcadémie des Sciences had gone in a similar theatricaldirection, which critics regarded as damaging to theinstitution’s raison d’être—pure science and serious re-search.

In 1864, the philologist Ernest Renan published abiting essay on the French system of higher education,comparing it unfavorably to that of Germany (23). AGerman visiting courses in Paris, he wrote, is “very sur-prised.” The lack of dignity and respect, the comingand going of the students during the lecture, the inatten-tion of the auditors, the theatrical style of the professor,and above all, the applause at the conclusion, strike theGerman student as curious. “An attentive listener hasno time to clap. This bizarre custom shows him oncemore that the purpose of the exercise is not to instruct,but to shine.” The intellectual danger that France wasrunning, stated Renan, was of becoming “a nation oforators and editors, without concern for essential mat-ters and for the real progress of knowledge.”

The Minister of Public Instruction at this time, Vic-tor Duruy, agreed with Renan. Duruy complained ofthe predilection among French academic historians ofhis day for “the depiction of personalities and passions,the analysis of the human heart, [and a] brilliant style oflight reading (24),” and he was well aware that the sci-

Page 10: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

88 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

ences shared the same histrionic style as the humani-ties. Duruy played the leading role in the creation of anew teaching institution in Paris, the École Pratiquedes Hautes Études. This was to provide a means forthe pursuit of careful scholarship, not “brilliance,” andfor the creation of new knowledge instead of the re-counting of moving stories.

Renan and Duruy were not alone in these opin-ions. A generation earlier, Liebig expressed a numberof similar thoughts, pronounced in his characteristichard-edged fashion. His 1832 paper, published in hisown proprietary journal, charged most French scien-tists with arrogance, chauvinism, rhetorical bombast,and thievery. The harshness of these judgments wasonly slightly ameliorated by his suggestion that muchof the behavior he described was an inevitable productof certain structural characteristics. In particular, hethought that the monopolistic power of the Académiedes Sciences, “the source of all remunerative positions,”led almost inevitably to an unseemly scramble for suc-cess. This was why French scientific papers seemedso arrogant and self-promoting, Liebig thought (25).(Liebig, of course, was not immune to the very faultshe imputed to his foreign rivals. No one was moreskilled at cultivating celebrityhood than he!—and manyaccused him, with some justice, of the very same listof crimes: arrogance, chauvinism, rhetorical bombast,and thievery.)

The Académie was, indeed, a powerful organiza-tion, not de jure but rather de facto. Given the central-ization of French science in Paris, combined with thehighly cumulated structure of professional positions, itwas almost inevitable that power in the communitywould be concentrated in a few hands; and the geo-graphic/institutional locus of that clique was theAcadémie. To nearly everyone outside of that clique,and even to some on the inside, this was an unhealthystructure. This accurate perception was the source ofthe attacks on cumul, which we have cited in the earlyletters of Dumas and Pelouze; of course, both ceasedattacking the system after they were brought into theelite. Cumul continued to be criticized throughout thecentury, without, however, being dismantled.

Cumul was made almost inevitable by the resis-tance of the governing authorities to raise academicsalaries to a decent level and to provide chairholderswith appropriate facilities for their work. The under-standable response of a scientist offered a professor-ship at the Sorbonne that included no laboratory and adeficient salary was to seek a second professorship.

When teaching duties became overwhelming becauseof the multiplicity of posts, the less desirable or moreexhausting positions could be farmed out to youthfulsuppléants, with whom the chairholder shared half thesalary in exchange for all of the real work. The sameunderlying cause, namely the refusal of the governmentfor proper support of higher education and research,meant that scholars were diverted in their middle andlater years into lucrative state consulting posts or poli-tics, rather than continuing their teaching or research.All of this also fed into the celebrity culture that I havedescribed.

Of course, none of this was healthy for the Parisianscientific community. The structure of that communitybecame even less salutary during the Second Empire,when Napoleon III abolished the meritocratic system ofcompetitions for major academic positions. It was adeveloping sense of imminent crisis in the 1860s thatled such leading figures as Wurtz, Renan, Pasteur, andDuruy to sound the alarms. Unfortunately, on the brinkof success the Franco-Prussians dealt a temporary de-feat to all of those efforts. Only in the new environmentof the early Third Republic could effective measuresfinally be taken for thorough going reform.

Conclusions

I want to caution that my account of French “celebrityculture” is intended at this stage merely as suggestive.Even stipulating the existence of this phenomenon, nei-ther its etiology nor its consequences are clear. Com-parisons across national boundaries, which I have noteven attempted, are vital to judge whether this reallywas a “pathology” that damaged the development ofFrench science relative to rival nations. This, of course,requires much more investigation. Contemporary his-torians of 19th-century French science have done extraor-dinary service in elucidating the historical developmentsdealt with here, and my work would have been incon-ceivable without that foundation. However, this field isstill relatively young and undeveloped, compared to oth-ers even within the history of science. I would like topoint to a few topics I think are worthy of further inves-tigation, under three headings: personalities, institutions,and practices.

Other than Lavoisier, Claude Bernard, and a smallnumber of other examples, French chemists have notbeen well investigated by historians. Speaking only ofmy cast of characters, I have already noted that Pelouzehas been almost completely neglected, and the same was

Page 11: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 89

true of Wurtz until a decade ago. Dumas has been stud-ied, at times with care, but I think it is fair to say that theDumas literature is still quite meager, compared to hisimportance. Even Berthelot, about whom much goodwork has been written, is still imperfectly understood.As for others contemporary with these personalities,there is almost nothing in the modern historical litera-ture on Thenard, Balard, Frémy, Chevreul, Cahours, orDeville. This is the case, despite the existence of massesof archival materials at the Archives Nationales and theArchives de l’Académie des Sciences.

We also need to know a great deal more about theinstitutions of science in the nineteenth century. MauriceCrosland has done us a great service with the publica-tion of his important monograph on the Académie desSciences (26), and we also have fine recent studies byJohn Weiss, Craig Zwerling, and Terry Shinn, amongothers, of the École Centrale, École Polytechnique, andthe École Normale; but this is only a beginning (27).One obvious desideratum is a study of science instruc-tion at the Sorbonne (28). A second is the Collège deFrance, the details of which are far too little understood(29). A third is a proper study of laboratories and facili-ties for research, both official and private. And verymuch to be desired is a broad study of the politics ofscience funding in the crucial middle decades of thecentury, where so little attention has been directed—thegeneration before the rise of what Harry Paul calls the“science empire” in France (30).

Finally, we would benefit greatly from a study ofpractices and customs in 19th-century French science.One such example is the system of cumul; it would benice to know more about how it actually operated, fromboth sides of the lectern. There are all sorts of detailsthat we know little about, including remuneration, ac-tual duties, procedures of selection of—and attitudes ofstudents toward—the suppléants, for example. A sta-tistical analysis of how the degree of cumulation of Pa-risian (or French) science changed over time would alsobe very revealing.

A second area of interest under the rubric of prac-tices is the doctrinal control of pedagogy. I have re-cently finished a study for which this topic was an im-portant element, but I was frustrated by the limits to whatI was able to learn. The usual complaint, then and now,is that French “anti-atomists” threw up roadblocks thateffectively prevented the teaching of atomistic chemis-try from the 1830s until the 1890s. In general I havefound this impression to be reasonably accurate, butmany puzzles remain unresolved; and it would be very

helpful to know more about both the effectiveness andthe tools by which this influence was exerted (31).

This story must, of course, be embedded within alarger account of the science politics of the day. Weknow that both Wurtz and Pasteur cried foul in 1863when Berthelot was awarded a new chair created ex-pressly for him at the Collège de France. They quitereasonably suspected that there had been some behind-the-scenes influence, but it would be gratifying to seethe actual details revealed in this and many other simi-lar episodes. It is, of course, possible that the real ac-tion happened in face-to-face encounters, or by a writ-ten trail that has vanished or is otherwise unrecover-able. This exemplifies one of the difficulties for thehistorian in dealing with Parisian science of the past.German scientists, spatially separated as they were inthe decentralized German states, wrote thousands of let-ters to each other and to their governments, many ofwhich still exist; Parisian scientists, by contrast, coulddo much business orally, leaving fewer tools by whichthe historian can reconstruct the action. Celebrities inparticular, the focus of this paper, often do their bestpromotional work in person.

The situation may not be quite as desperate as Iappear to be suggesting. Anyone who has sampled therichness of yet unexploited resources at the ArchivesNationales and the Archives de l’Académie des Sciencesknows that the historical study of 19th-century Frenchscience is still young. These are, of course, only two ofa great network of archives, and much still exists also inprivate hands. Exciting work lies ahead.

REFERENCES AND NOTES

1. G. Johnson, “The Jaguar and the Fox,” Atlantic Monthly,July 2000, 82-85.

2. A. Chevallier, “Théophile-Jules Pelouze,” Journal dechimie médicale, 1867, [5] 3, 444-48 (444).

3. The five obituaries are Chevallier (Ref. 2); C. vonMartius, Neues Reporatorium für die Pharmacie, 1868,17, 506-10; A. Cahours, Moniteur scientifique, 1868, 16,502-8; W. De la Rue, J. Chem. Soc., 1868, 21, xxv-xxix;and J. B. Dumas, in Discours et éloges académiques,Gauthier-Villars, Paris, 1885, Vol.1, 125-98. Pelouzealso enters substantively in F. L. Holmes’ biography ofJustus Liebig in the Dictionary of Scientific Biography,and in his Claude Bernard and Animal Chemistry,Harvard University Press, Cambridge, MA, 1974. AlexBerman’s three-paragraph Pelouze article in the Dictio-nary of Scientific Biography (1974) is excellent consid-

Page 12: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

90 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

ering the limitations under which he labored. There islittle on Pelouze in Partington’s History of Chemistry.Celia von Lindern (University of Regensburg) is cur-rently preparing a full edition of the Liebig-Pelouze cor-respondence.

4. The information on Pelouze in this paper is derived fromthe sources cited in the preceding note and the follow-ing: the government annual Almanach royal [national,impérial] (later editions of which add the words Annuairede la République Française); the Pelouze correspon-dence in the Liebigiana Collection of the BayerischeStaatsbibliothek; and the Dossier Pelouze in the Archivesde l’Académie des Sciences, Paris. Some of the mate-rial in this paper appears in slightly different form in A.Rocke, Nationalizing Science: Adolphe Wurtz and theBattle for French Chemistry, MIT Press, Cambridge,MA, 2001.

5. This story, related in the Chevallier and Dumas éloges,certainly derives directly from Pelouze, but the date isnot certain. Gay-Lussac was a consultant at the largeiron works in Charenton, and Pelouze’s father lived andworked there. The vehicle specified in the story is a“coucou” (cuckoo), a two-wheeled carriage for hire thatsoon thereafter went out of fashion. I take it to be simi-lar to the later English hansom cab.

6. Dumas to Liebig, n.d., postmarks December 19, 1837,January 21, February 25, May [?], and May 21, 1838,Liebigiana IIB.

7. Dumas to Liebig, n.d., postmarked April 21 and May[?], 1838, Liebigiana IIB. “Depuis que je suis à Paris, jecherche un moyen d’avoir un laboratoire largementconstitué sous ma main. Je crois y être enfin parvenu etc’est là ce qui me fait quelque consolation. Je pourraidans deux ou trois mois faire travailler dix élèves choisischez moi et je pourrai consacrer à leurs expériencesquatre ou cinq milles francs par an. Alors seulement, jeserai en mesure de reprendre des expériences en con-currence avec les vôtres. Je ne puis pas aller votre pasdans ce moment.”

8. J. B. Dumas, Discours (Ref. 3), pp 196-97.9. Dumas wrote Liebig (n.d., late 1831, Liebigiana IIB),

“Je suis parfaitement résolu à ne jamais plier le dosdevant personne, pour obtenir une place meilleure. Ilen résulte que je serai longtemps sans en avoir … Jamaisje n’ouvrirai la bouche contre ceux qui cumulent, jamaisje ne consentirai à la création d’une place en faveur demon chétif individu.” Pelouze wrote Liebig (22 No-vember 1836, Liebigiana IIB): “Vous êtes vraimentheureux d’être éloigné d’aussi viles passions que cellesqui agitent la grande majorité des gens que l’on appellesavant. Le spectacle de tout cela me rend triste etm’engage plus que jamais à vivre éloigné de toutesociété. Pour mon compte au moins, je n’ai pas mêmesollicité la suppléance dont M. Thenard me charge [atthe Collège de France]; c’est lui-même que me l’aproposé.”

10. Pelouze had a small laboratory installed in (or attachedto) his new lodgings soon after taking the position ofassayer in 1834. In his letter to Liebig of August 22,1835, he wrote, “On m’a donné deux chambres à côtéde mon logement. J’en fais dans ce moment unlaboratoire et j’espère bien travailler plus que par le passéet d’une manière d’ailleurs moins fatiguante.” Threeyears later (May 1, 1838), however, he announced toLiebig the completion that day of “un magnifiquelaboratoire de chimie qui va remplacer le petit trou quij’occupais au bout de mon appartement” (LiebigianaIIB). Various witnesses rate the capacity of this labora-tory (Rue Guénégaud) as being between six and twelveworkers; it is apparently the laboratory that Laurent in-herited in 1848. Many students and guest workers ac-commodated in Pelouze’s laboratory can be documentedprior to the opening of the better known laboratorytreated in the following paragraph.

11. The best (and almost the only) source of information onPelouze’s laboratories is an untitled, undated 3,300-wordanonymous document prepared (probably at Dumas’srequest) shortly after Pelouze’s death (Dossier Pelouze,Académie des Sciences). It is precise, carefully re-searched, and seems authoritative. The most likely au-thor candidate is Aimé Girard, who supervised the labo-ratory during the final five years of its existence.

12. As noted, Dumas’s Rue Cuvier laboratory operated from1838 until 1848. Wurtz ran a similar school from 1850until 1853, and Gerhardt ran one from 1851 until 1855.Neither of the latter was a success, and the probable rea-son for that is competition from Pelouze’s operation.

13. According to the information in M. P. Crosland, Gay-Lussac, Scientist and Bourgeois, Cambridge Univer-sity Press, Cambridge, 1978, 230-31, Pelouze’s posi-tions at Saint-Gobain and the Mint alone must havebrought in about 26,000 francs, at a time when most pro-fessorships in major academic institutions earned about5,000.

14. A. Lieben, “Erinnerungen an meine Jugend- undWanderjahre,” in Adolph Lieben Festschrift, Winter,Leipzig, 1906), 5. Lieben was a student in Wurtz’s labo-ratory in Paris in the years 1856-58 and 1862.

15. L. Velluz, Vie de Berthelot Plon, Paris, 1964, 119-20.The writer was probably Duruy himself.

16. J. Jacques, Berthelot 1827-1907: Autopsie d’un Mythe,Berlin, Paris, 1987, 105-11; A. Carneiro, The ResearchSchool of Chemistry of Adolphe Wurtz, Ph.D. Thesis,University of Kent/Canterbury, 1992, 69. Deville hadbeen elected in the mineralogy section of the Académieas early as 1861; Pasteur was elected to the same sec-tion in the following year.

17. Jacques, Autopsie (Ref. 16), “Un test de la popularité deBerthelot: Son nom dans la géographie urbain,” and B.Javault, “Berthelot, héros de la IIIème République,” bothin J. Dhombres and B. Javault, Ed., Marcelin Berthelot:Une vie, une époque, un mythe SFHST, Paris, 1992, 121-31 and 113-19, respectively. Jacques has determined

Page 13: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 91

that the only scientists who have had more French citystreets named for them than Berthelot are Pasteur andthe Curies.

18. J. R. Partington, A History of Chemistry, Macmillan,London, 1964, Vol. 4, 467.

19. G. Geison, The Private Science of Louis Pasteur,Princeton University Press, Princeton, NJ, 1995, 10.

20. M. P. Crosland, Science under Control: The FrenchAcademy of Sciences, 1795-1914, Cambridge Univer-sity Press, Cambridge, 1992.

21. The classic work on this subject is A.-J. Tudesq, Lesgrands notables en France, Paris, 1964. 2 vol.

22. The best discussions of this French Romantic oratoricalculture are found in R. Fox, “Scientific Enterprise andthe Patronage of Research in France, 1800-1870,”Minerva, 1973, 11, 442-73 (452-58), and R. Fox, “Sci-ence, the University, and the State in Nineteenth-Cen-tury France,” in G. Geison, Ed., Professions and theFrench State, 1700-1900, University of PennsylvaniaPress, Philadelphia, PA, 1984, 66-145 (81-84); see alsoA. Rocke, “History and Science, History of Science:Adolphe Wurtz and the Renovation of the AcademicProfessions in France,” Ambix, 1994, 41, 20-32.

23. E. Renan, “L’instruction supérieure en France, sonhistoire et son avenir,” Revue des deux mondes, 1864,[2] 51, 73-95.

24. Quoted in L. Liard, L’ Enseignement supérieur enFrance, 1789-1893, Colin, Paris, 1888-1894, Vol. 2, 292.

25. J. Liebig, “Bemerkungen zu vorhergehendenAbhandlung [von Thenard],” Ann. Chem. Pharm., 1832,219-30.

26. Ref. 20.27. C. Zwerling, The Emergence of the École Normale

Supérieure as a Center of Scientific Education in Nine-teenth-Century France, Garland, New York, 1990; J.Weiss, The Making of Technological Man: The SocialOrigins of French Engineering Education, MIT Press,

Cambridge, MA, 1982; T. Shinn, Savoir scientifique etpouvoir social: École polytechnique et lespolytechniciens, 1794- 1914, Presses de la FondationNationale des Sciences Politiques, Paris, 1980.

28. There is a small literature on the physical rebuilding ofthe Sorbonne during the early Third Republic, but al-most nothing, as far as I am aware, on science instruc-tion.

29. A. Lefranc’s Histoire du Collège de France, depuis sesorigines jusqu’à la fin du Premier Empire, Hachette,Paris, 1893, extends only to 1814, as the subtitle indi-cates.

30. H. Paul, From Knowledge to Power: The Rise of theScience Empire in France, 1860-1939, Cambridge Uni-versity Press, Cambridge, 1981.

31. The best recent contribution to this question is C.Kounelis, “Atomism in France: Chemical Textbooks andDictionaries, 1810-1835,” in A. Lundgren and B.Bensaude-Vincent, Ed., Communicating Chemistry:Textbooks and Their Audiences, 1789-1939, ScienceHistory Publications/USA, Canton, MA 2000, 207-31.

ABOUT THE AUTHOR

Alan J. Rocke is Henry Eldridge Bourne Professor ofHistory, Case Western Reserve University, Cleveland,OH 44106-7107. He is the author of Chemical Atom-ism in the Nineteenth Century: from Dalton toCannizzaro (Ohio State University Press, Columbus,OH, 1984); The Quiet Revolution: Hermann Kolbe andthe Science of Organic Chemistry (University of Cali-fornia Press, Los Angeles, CA, 1993); and Nationaliz-ing Science: Adolphe Wurtz and the Battle for FrenchChemistry (MIT Press, Cambridge, MA, 2001).

JOHN H. WOTIZ

HIST Chair, Dexter Award Recipient

April 12, 1919 – August 21, 2001

Page 14: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

92 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

On March 27, 1876, Professor Charles F. Chandler ofthe School of Mines at Columbia College and seven ofhis colleagues issued a short but specific invitation tomembers of the chemical profession (2):

Dear Sir: A meeting for organizing the AmericanChemical Society will be held on Thursday evening,April 6, 1876 at 8 o’clock P.M. in the lecture roomof the College of Pharmacy, University Building,corner Waverly Place and University Place.Your attendance is earnestly required.

In spite of such short notice, 34 men assembled withChandler at the appointed time on the New York Uni-versity campus and promptly elected him president ofthe meeting (3). The subsequent motion by Isidor Walz“that we proceed to organize a national Chemical Soci-ety, which shall be called the American Chemical Soci-ety,” provoked a vigorous discussion which finally re-sulted in an affirmation that was marred by three dis-senters who submitted negative votes (4).

This lack of unanimity in the formation of theAmerican Chemical Society (ACS) did not come as acomplete surprise to the meeting’s organizers. In fact,it was a continuation of action initiated almost two yearsearlier during the Centennial of Chemistry celebrationheld at a public school in Northumberland, Pennsylva-nia. At that magnificent meeting, generally consideredto be the first national meeting of chemists held in theUnited States, the participants gathered with high spir-its in a picturesque setting to honor the discoverer ofoxygen, Joseph Priestley, whose last home and burialplace were in Northumberland. In keeping with thename chosen for the meeting, they also examined the

OPPOSITION TO THE FORMATION OFTHE AMERICAN CHEMICAL SOCIETY (1)

James J. Bohning, Lehigh University

development of chemistry in the United States over thecentennial years 1774 to 1874 (5). The occasion of thismeeting was prompted by a suggestion from ProfessorHenry Carrington Bolton of Columbia College, whoproposed the centennial significance of the year 1874(6), and by Professor Rachel L. Bodley of the Women’sMedical College of Philadelphia, who later suggestedthe location (7). When Bolton urged the Chemical Sec-tion of the New York Lyceum of Natural History to un-dertake the organization of the affair, it was only natu-ral that President J. S. Newberry appoint him chairmanof the General Committee (8).

When the Centennial of Chemistry meeting openedpromptly at 9 A.M. on Friday, July 31, 1874, Boltonacted as the temporary chairman. Although the assem-bly subsequently elected Charles F. Chandler as presi-dent and presiding officer for the two-day meeting,Bolton’s efforts did not go unrecognized. A special reso-lution, passed just prior to adjournment on August 1,commended Bolton for his “considerable attention todetails” that resulted in “a memorial gathering” to whichall could “look back with the greatest satisfaction (9).”While much of that satisfaction derived from the physi-cal location and the Priestley legacy, at least an equalamount must have been derived from the professionalcontact of 77 chemists previously scattered over 16states and two foreign countries, and representing at least25 academic institutions as well as many industrialfirms.

For Professor Persifor Frazer of the University ofPennsylvania, there appeared to be a very logical course

Page 15: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 93

F. W. ClarkeBolton Egleston

J. L. Smith

Fraser

Benjamin Silliman, Jr.

Eben N. Horsford T. Sterry Hunt

Samuel W. Johnson

Opponents

Page 16: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

94 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

of action to assure the scheduling of similar events inthe future. Towards the end of the Friday afternoon ses-sion, Frazer proposed (5):

..the formation of a chemical society which shoulddate its origin from this centennial celebration, andurged the importance of the fact that, while Ameri-can chemists have done perhaps a larger amount ofwork in their own department proportionately thanhas been done in the world within the last century inany other branch of science, they have as yet in thiscountry not a single society to represent the chemi-cal thought of the country.

Frazer then moved (5):

..that a committee of five be appointed by the presi-dent, to whom shall be referred the advisability ofcalling a representative committee of chemists of theUnited States to form a chemical society, and all ques-tions relating to the organization of the society.

While this urge to form an organization of chemists was,in part, a direct consequence of the camaraderie preva-lent throughout the meeting, it was not new to the Cen-tennial of Chemistry celebration. In a resolution adoptedon May 11, 1874, the New York Lyceum of Natural His-tory recognized that a “social reunion of AmericanChemists, for mutual exchange of ideas and observa-tions, would promote good fellowship in the brother-hood of chemists (8). The organizing committee con-tinued in its circular to stress that “a reunion of Ameri-can Chemists … would … foster a feeling of fraternityamong us.” This sentiment was also echoed by corre-spondents responding to Bolton’s first suggestion of themeeting in the American Chemist (6).

The New York Daily Graphic was even moved tocapture this spirit in an unusual chemical metaphor, pro-jecting that (10):

If the chemists who were at Northumberland … hadcombined in certain definite proportions to accom-plish what was really the obvious purpose of theirmerely mechanical mixture, … the world would havecause to rejoice in their synthesis [for] hithertoAmerica has done but little for the science, eachchemist being but an isolated molecule giving butlittle show of affinity for others.

Frazer’s motion, therefore, was a rational and naturalreaction made with optimism and in anticipation of con-structing an organization that would be directly relatedto the origins of modern chemistry through the Centen-nial of Chemistry celebration.

The first to respond to this proposal was the emi-nent mineralogist J. Lawrence Smith of Louisville, Ken-

tucky. Frazer must have been dismayed when Smithimmediately stated that there were many difficulties informing such an organization. “One formidable objec-tion was that this country was too large, and that it wouldbe impossible to centralize its chemical research.” Con-tinuing, Smith pointed out that “the very strength of thecountry is in decentralization. We want all of our scien-tific institutions dispersed far and wide.”

Even Smith must have realized the weakness of thisargument, for he then proceeded to present a more di-rect and specific premise:

We already have two great institutions in the country– the American Scientific Association and the Ameri-can Academy of Sciences [11] – which undertake toembrace in their proceedings everything connectedwith chemical research, and it would be more credit-able to the chemical talent of this country if an at-tempt were made to secure its better representationin the chemical section of the former association.

To support this line of reasoning, Smith included ex-amples of foreign organizations:

Even the meetings of the Chemical Society of Lon-don, where there exists a great centralization of chem-ists, are very meagerly attended, the members pre-ferring to read their papers before the more distin-guished Royal Society. The same is true of the FrenchChemical Society, while the attention of the Acad-emy of Sciences of France is constantly asked forpapers of the highest importance relating to chemis-try.

Smith’s biased opinion was only partially correct, for in1874 chemical societies were not only established inEngland, France, Germany, and Russia; they were alsopublishing journals devoted solely to chemistry (12).Yet, in the lengthy and “somewhat heated debate” thatfollowed, only one speaker, Professor William H. Chan-dler of Lehigh University (and the younger brother ofChairman Charles F. Chandler) “presented forcibly manycogent arguments in favor of the formation of a nationalchemical society.” He was outnumbered, however, byfive other speakers who “advocated the earnest co-op-eration of the chemists as a body with the AmericanScientific Association, and that if a national chemicalsociety were formed, it should be a permanent sectionof that body.”

An evaluation of the effectiveness of these com-ments is best made by examining the stature of thosewho made them. William H. Chandler (age 32) wasknown to his audience as the co-editor of the first Ameri-can chemical journal, The American Chemist, which he

Page 17: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 95

had started with his brother Charles in 1870 (13). How-ever, those speaking against the motion were also indi-viduals whose reputations had attained national signifi-cance.

J. Lawrence Smith (age 55), who held the M.D.degree from the Medical College of South Carolina,was the first person from the United States to studyunder Justus von Liebig at Giessen. A former profes-sor of medical chemistry and toxicology at the Univer-sity of Louisville, he had published extensively on min-eral analysis and had developed a process for the sepa-ration of alkali metals from silicates that bore his name.A cofounder in 1846 of the Southern Journal of Medi-cine and Pharmacy, Smith had recently published histreatise on “Minerals and Chemistry: Original Re-searches” in 1873. His services to foreign governmentsbrought him decorations from France, Turkey, and Rus-sia (14).

Benjamin Silliman, Jr. (age 57) had published text-books in both physics and chemistry that were im-mensely popular in colleges throughout the country andwas currently serving as editor of the American Jour-nal of Science and Arts, a publication that had beenfounded by his father in 1818 (15).

Frank Wigglesworth Clarke (age 27) had just as-sumed his position as professor of chemistry and phys-ics at the University of Cincinnati after previous posi-tions at Cornell, Boston Dental College, and HowardUniversity. In spite of his youth he had published manyarticles in the popular press. A series in Silliman’s andChandler’s journals became “The Constants of Nature.Part I,” which had just been published by theSmithsonian Institution in 1873 (16).

Eben Norton Horsford (age 56), having obtained aB.S. from Rensselaer Polytechnical Institute in 1836,was the second person from the United States to studywith Liebig. He developed the first laboratory inAmerica for analytical chemistry in the Lawrence Sci-entific School at Harvard, serving as Rumford Profes-sor from 1847 to 1863 and Dean from 1861 to 1862. In1863 he resigned his academic position to pursue in-dustrial chemistry, founding the Rumford ChemicalCompany in Rhode Island from the profits of his in-vention of the phosphate baking powder as a yeast sub-stitute (17).

Edward Travers Cox (age 53) grew up in New Har-mony, Indiana, where he received his early training inchemistry and geology from David Dale Owen, assist-

ing Owen in the U.S. government field studies of theUpper Mississippi Valley and the geological surveys ofKentucky and Arkansas before the Civil War. He madean extensive survey of mining opportunities in NewMexico in 1864 and identified important coal depositsin southern Illinois. Appointed State geologist in 1869by the Governor of Indiana, he immediately began hisseries of annual reports on the geology of Indiana. AsState Geologist, Cox automatically filled the chair ofgeology at Indiana University (18).

Peter Henri Van der Weyde (age 61) was a physi-cian with an M.D. from New York University, who helda faculty position with the Women’s Medical College inNew York. Previously, he had held faculty positions atNew York Medical College, Cooper Union, and GirardCollege. Founder and editor-in-chief of The Manufac-turer & Builder, which started in 1869 as a “practicaljournal of industrial progress,” Van der Weyde obtainedpatents in 1867 and 1869 relating to a petroleum distil-late product (called “Chemogene”) and a compressionice system, which was used to construct artificial re-frigeration systems throughout the south and in Phila-delphia (19).

Sensing that Frazer’s original motion would becrushed under the weight of such heavy opposition,Bolton offered a compromise in the form of an amend-ment:

That a committee of five be appointed from this meet-ing to cooperate with the American Association forthe Advancement of Science [AAAS] at their nextmeeting, to the end of establishing a chemical sec-tion on a firmer basis.

The assembly gladly and quickly adopted the modifiedresolution, and Chairman Charles F. Chandler appointedBolton, Silliman, Smith, Horsford, and Professor T.Sterry Hunt of the Massachusetts Institute of Technol-ogy as committee members. This strong allegiance tothe AAAS effectively blocked the formation of theAmerican Chemical Society at a time when it could havedirectly related its origin to the centennial of modernchemistry as it was celebrated in Northumberland.

In order to understand the rationale behind this vir-tually unanimous rejection of the Frazer proposal, it isnecessary to examine the relationship of the objectorsto the AAAS. For example, every member of the com-mittee appointed by Chandler and all of the antagonistswho spoke against the Frazer proposal were membersof the AAAS; three of them were charter members dat-ing back to the formation of the AAAS as a reorganiza-

Page 18: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

96 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

tion of the American Association of Geologists and Natu-ralists in 1848 (20). Most were active members as well,holding a variety of offices at the annual meeting thatwas generally held in the fall of the year. Thus, Sillimanhad served as assistant secretary (1841-1843), secretary(1847-1848), and chairman (1841-1842), while Horsfordhad been general secretary (1949-1850). Most signifi-cantly, J. Lawrence Smith had served as president of theAAAS at the 21st Annual Meeting held in Dubuque, Iowain August, 1872, with his term of office ending at the22nd Annual Meeting held at Portland, Maine in Augustof 1873 (21).

At the Dubuque meeting only two chemical paperscan be identified from the five presented before SectionA on “Physics and Chemistry.” The following year, how-ever, in Portland, six chemical papers were presented inSection A, including one by F. W. Clarke. It was at theconclusion of this meeting that Clarke and others metinformally to present “laboratory notes and informalpapers.” They found the mutual exchange of ideas sosatisfactory that they adopted resolutions for the forma-tion of a separate chemical subsection, which were tobe presented to the Steering Committee at the 23rd an-nual meeting scheduled for Hartford, Connecticut onAugust 12–19, 1874 (22).

It is possible to identify two groups with differentmotives among those who united to support the AAAS.Through the end of the Civil War, American scientistswere primarily generalists, with interests and publica-tions spanning a wide range of topics and applications.Thus, in 1874, individuals such as Smith, Horsford,Silliman, Van der Weyde, and Cox (with an average ageof 56) became part of the old guard who were reluc-tantly thrust into the beginning age of specialization,which was developing as a result of the industrial revo-lution. For those who were accustomed to followingtheir own curiosity in “natural philosophy,” the forma-tion of a chemical society would be contrary to theirbelief that specialists would stifle the exchange of sci-entific thought by imposing boundaries for scientificinvestigation (23).

This attitude was expressed by the secretary of theHartford AAAS meeting, who wrote (24):

The action taken by the … formation of a PermanentSubsection of Chemistry was … in accordance withthe objects of the Association in bringing togetherscientists in all departments, that this expression onthe part of … special branches can only be regardedas most favorable towards the annual centralization

of scientific thought in the country during Associa-tion week, and it cannot be long before the AmericanAssociation will draw within its folds … many spe-cial organizations now existing, which … workingfor one common end would thus still more greatlyaid in the Advancement of Science in America.

This concern did not dissipate readily, for many yearslater, in describing the fiftieth anniversary of the AAAS,Daniel S. Martin complained that (25):

The increase of specialism has led not only to a divi-sion of the association into nine sections, in place ofthe two to three of its early years, but to the forma-tion of several separate organizations of specialistwhich have been looked upon as tending to weaken,or even disintegrate, the main body. The AmericanChemical Society, the American Mathematical Soci-ety, and the Geological Society of America may becited as leading examples.

For the younger chemists whose careers were just be-ginning, the concerns were much different. As the firstgroup of specialists, their efforts had become more so-phisticated and consequently more definite with regardto purpose. While their older colleagues often rambledthrough their scientific investigations with little cohe-sive planning, the new specialists were more careful toexplore a topic in detail, not being averse to spendingmuch of their lives on a single sub-specialty. This newattitude brought with it an increasing desire to exchangeinformation with others who had similar special inter-ests. It was within this framework that Clarke’s group,whose average age was just 33, were preparing to final-ize the new chemical subsection of the AAAS just twoweeks after the Centennial of Chemistry meeting inNorthumberland (26).

It is not surprising, therefore, that there was littleenthusiasm for the possibility of forming a new societyof chemists as proposed by Frazer. Further, since 20 ofthe 77 who came to Northumberland were already mem-bers of the AAAS (20), Bolton’s amendment, seeminglythe logical choice, was accepted without further com-ment. Subsequently, during the week of August 12,1874, a Permanent Subsection of Chemistry of SectionA of the AAAS was established, with Professor S. W.Johnson of Yale as Chairman and F. W. Clarke as Secre-tary.

Following almost immediately after the Centennialof Chemistry celebration, the Hartford meeting of theAAAS was the beginning of a long period of intenseactivity for chemists and the new subsection. The Boltonamendment was quite effective, for eight people whowere in Northumberland a few weeks earlier became

Page 19: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 97

Charles F. Chandler William H. Nichols

Proponents

members of the AAAS at Hartford (27). The numberof papers of chemical interest increased to 25, includ-ing six by J. Lawrence Smith, two by F. W. Clarke, andthree by P. H. Van der Weyde (28).

In 1875 Clarke sent a letter to the editor of Popu-lar Science Monthly, describing the formation of thesubsection of “chemistry, chemical physics, chemicaltechnology, mineralogy, and metallurgy,” and urged in-terested readers to attend the 24th meeting of the AAASin Detroit, Michigan, during the week of August 11.Concerned that the fledgling subsection needed full at-tendance “in order to make it a success,” Clarke pointedout that (29):

Chemistry has been but little represented in the pro-ceedings of the Association, and the time now seemsto have arrived in which some good work can bedone.

In Detroit, three more attendees from theNorthumberland meeting became AAAS members.Clarke presented three papers and Smith gave five,while H. C. Bolton was elected to replace Clarke as thesecretary of the chemistry subsection at the next an-nual meeting (30).

According to its constitution, the chief object ofthe AAAS was “by periodical and migratory meetingsto promote intercourse between those who are cultivat-

ing science in different parts of the United States.” Thesemigratory meetings “gave American chemists for the firsttime an opportunity of obtaining periodic scientific con-tacts of a national character in different sections of thecountry (23).” As the new chemical subsection of theAAAS continued to attract new members and preparefor the 25th annual meeting in Buffalo, New York, inAugust of 1876, the concept of a separate national chemi-cal society lay officially dormant.

Some of the New York chemists who were atNorthumberland began privately to discuss the forma-tion of their own chemical society. Their leader wasCharles F. Chandler of Columbia University, who as headof a self-appointed committee of seven issued an invita-tion in January of 1876 to the chemists of New York tosee whether there was sufficient interest to form a localchemical society. “Widely scattered as the chemists inthis neighborhood are,” Chandler said, “such an associa-tion would become a centre [sic] of pleasant personalintercourse, and of an interchange of views, experienceand researches which would benefit all concerned (31).”Chandler and his colleagues were not following up onthe Frazer suggestion at the Priestley house, for theirthinking was strictly focused on what would be benefi-cial to New Yorkers. Chandler’s circular was mailed toabout 100 chemists in the New York City area. Less thanhalf this number responded favorably, but Chandler and

Page 20: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

98 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

his associates boldly forged ahead. This number, theysaid, “was so unexpectedly satisfactory” that it wouldbe “deemed opportune” to attempt the formation of anational rather than a local organization. On March 22,1876, the Chandler group sent another circular announc-ing their intentions, but this time the mailing list was adifferent group of 220 chemists throughout the country.To encourage this group, who would be considered “non-resident” members, that is, outside of a 30-mile radiusof New York City, Chandler promised that at least onemeeting a year would be held outside of New York, “atsuch a time and at such a place as to make attendanceon the part of non-resident members more convenientand representative.” Within six days, lured byChandler’s rhetoric, some 60 non-resident chemists re-turned a favorable response. Armed with what he con-sidered to be a significant response to his polls, Chan-dler issued the call to meet on April 6, 1876, for thepurpose of forming the American Chemical Society (32).

As chairman of this organizational meeting, Chan-dler began the proceedings by stating that:

the ultimate object was to bring chemists together inscientific and social intercourse, to secure roomswhich would be open in the day and evening, and toestablish there a library of reference and a chemicalmuseum (33).

It fell upon Isidor Walz, acting as the secretary of themeeting and the organizing committee, to describe insome detail the steps that had brought the group togetheron that April evening. At the conclusion of his presen-tation, Walz referred to the existing opposition to theformation of a national society. Noting that the organi-zational committee did not share that opinion, he never-theless reported that the opposition had been promised“the fullest opportunity to lay their views before thegroup,” and he urged that the audience give them “ear-nest and careful attention.” Walz then offered his mo-tion to organize; and after a second to the motion, Chair-man Chandler “declared the subject open to discussion.”

The first to respond to the Walz motion was Pro-fessor Thomas Egleston, who started on a semi-positivenote by stating that it “might be advisable to organizesuch an association,” but not in 1876. His first objec-tion was based on experience with the American Insti-tute of Mining Engineers (AIME), which had beenformed five years earlier. Noting that the $6,000 incomefrom the 600 members of the AIME was used entirelyto publish its Proceedings, Egleston concluded that the100 potential members of the new chemical society(those responding favorably to the Chandler polls of

January and March) would bring an income of $500 to$1000, which would not be sufficient to “publish Pro-ceedings in a creditable and prompt way.” His fear, that“it is hard to get money now—when it will be easier toobtain it, the success of such a society would be betterassured” was not unusual but it was rather feeble. Aswith Professor Smith at Northumberland, Egleston thenlaunched a more pointed argument that was immedi-ately reinforced by the second speaker, Henry CarringtonBolton. Both men supported no immediate action bythe group, claiming that it would be unfortunate to havea division of forces. Instead, they favored cooperationwith two existing societies: the Chemical Section of theNew York Academy of Sciences and the Chemical Sec-tion of the AAAS.

It is ironic that both of these opponents were col-leagues of Chandler at Columbia. Bolton (age 39), aColumbia graduate who obtained the doctorate in 1866with Friedrich Wöhler at Göttingen (just ten years afterChandler), was an assistant in analytical chemistry atthe School of Mines and a faculty member at theWomen’s Medical College of New York. His extensivebibliographic compilation of the literature of uraniumchemistry that appeared in 1870 was a direct result ofhis many research papers published on uranium com-pounds between 1866 and 1870 (34).

Professor Thomas Egleston (age 44) was thefounder of the School of Mines at Columbia. When hepresented his proposal to the Columbia Trustees inMarch of 1863, it marked the first attempt at a new formof education in mining and metallurgy in the UnitedStates and was strongly influenced by Egleston’s expe-rience at the École des Mines in Paris several years ear-lier. In fact, it was Egleston who personally recruitedCharles F. Chandler to become the chemist on the newfaculty of three when the school opened in Novemberof 1864 (35).

The third person to comment negatively on the Walzmotion was Professor Albert R. Leeds of Steven Insti-tute of Technology. Leeds (age 42) acted as secretaryof the Centennial of Chemistry celebration inNorthumberland, and it is his vivid account in the Ameri-can Chemist that still remains the most authoritativedescription of that meeting. Leeds was not emphatic inhis comments, preferring simply to state that he “didnot think the movement timely (36).”

The appeal of Egleston, Bolton, and Leeds on be-half of the existing organizations resembled that madein Northumberland in 1874. There were several differ-

Page 21: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 99

ences, however, that were important to the outcome ofthe vote on the Walz motion. This time the speakerswere local individuals who were not of the same na-tional caliber as those in Northumberland. There wasalso much more of a direct concern that there would bea keen rivalry between similar organizations for the at-tention of a small number of potential members. Ofparticular interest was the effect on the New York Acad-emy of Sciences. Although it had originally beenfounded as the Lyceum of Natural History in 1817, thenew name had just been adopted in February of 1876when an extensive revision to revitalize the organiza-tion was completed. As part of these changes, a sectiondevoted entirely to chemistry had been established andwas holding monthly meetings. The architect of thesechanges, which required “revising and remodeling theentire constitution and by-laws,” was the vice-president,Thomas Egleston (37). His plea, however, did not havemuch effect on his listeners, for of the 35 present onApril 6, only six were members of that organization,including Chandler, William Habirshaw (a member ofChandler’s organizing committee), Bolton, Egleston, andLeeds (38).

In a similar manner Bolton, who was scheduled toreplace Clarke as secretary of the chemistry subsectionof the AAAS at the Buffalo meeting in August 1876,would certainly have been remiss had he not remindedthe group about that existing organization and its chemi-cal activities. Contrary to the response inNorthumberland, his point had little effect this time, forthere were only three AAAS members present: Bolton,Leeds, and Chandler.

After the lengthy remarks by Egleston and Bolton,Chandler appealed for help from his colleagues, sug-gesting “that it would almost seem as though we hadmet for the purpose of deciding not to organize a chemi-cal society.” In utter exasperation, Secretary Walz an-swered that “the apparent reluctance to discuss the sub-ject was probably due to the fact that the chemists presenthad the subject before them so long and had discussed itso thoroughly in private that a prolonged expression ofviews at this meeting was deemed unnecessary.” Nev-ertheless, he proceeded to present a brilliant rebuttal,aided by Herman Endemann (D. phil., Marburg, 1866,later to be the first editor of the Journal of the AmericanChemical Society), and Chandler’s assistant at the NewYork Department of Health, and Meinhard Alsberg, alsoat the Health Department.

It was, however, the statement of William H.

Nichols, who at age 24 had just begun his career in thechemical industry, that was the most noteworthy as wellas prophetic:

We did not come here expecting to find a societyready formed, with a library and a fine building; thosewould come in time. We have much intelligence as-sembled here, and that is better than a library. Muchbenefit would accrue to all branches of the profes-sion from such a society as that proposed. Let usbegin this society small, let it do its work well, and itwill undoubtedly grow. [emphasis added]

On that note of optimism, Chandler called for the mo-tion, which passed with three nays, presumably byEgleston, Bolton and Leeds. Moving quickly to theorganizational business at hand, Chandler began to dealimmediately with some of the opposition. The electionof John W. Draper as the first president was preciselycalculated to minimize the influence of those who mightbe tempted to continue attacking the new society (39).At the same time, the first nominating committee re-turned 21 other names for offices mandated by the newlyadopted constitution, including H. Carrington Bolton(corresponding secretary), Albert R. Leeds (committeeon papers and publications), and the absent J. LawrenceSmith (vice-president). Only Bolton refused to serve,preferring instead to continue his AAAS activities.

It would be many years before the conflicts betweenthe AAAS chemistry subsection, the New York Acad-emy of Science, and other chemical societies that mate-rialized after 1876 would be resolved (40). When theACS finally emerged as the national professional orga-nization of chemists, many of those who had initiallysupported other societies became active and influentialACS members (Fig. 1). Unfortunately, the person whofirst publicly raised the concept of a national society inNorthumberland did not remain committed to his ownproposal. Persifor Frazer followed Bolton’s suggestionand joined the AAAS in August, 1874. Later, he joinedthe ACS in July, 1876, but shortly thereafter resigned inJanuary, 1877, during the initial stages of the prolongedcontroversy when detractors claimed the ACS was not atrue national society but was in reality a local New Yorkorganization. Apparently, Frazer did not renew his ACSmembership at a later date.

There remains one interesting and unansweredquestion. In the period from August 1, 1874 to January22, 1876, there is no record of what those in New Yorkwho favored the original Frazer motion might be con-sidering. Only three of Chandler’s “self-appointed com-mittee” were at the Northumberland meeting. Yet, ac-

Page 22: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

100 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

Figure 1. ACS Activities of Those Who Initially Opposed ACS Formation

Name ACS Membership ACS Office

A. Speakers at the Centennial of Chemistry, August 1, 18741. J. Lawrence Smith April 1876 President, 18772. Benjamin Silliman April 1876 Vice-President, 18783. Frank W. Clarke 1877* President, 19014. Eben N. Horsford 1877 None5. Edward T. Cox ? None

B. Petitioners for the AAAS Chemistry Subsection, August 12, 18741. Samuel W. Johnson July 1876 Vice-President, 1877

President, 18782. T. Sterry Hunt April 1876 Vice-President, 1877, 1886,1887, 1889

President, 18793. George F. Barker April 1876 President, 18914. Harvey W. Wiley November 1876 President, 1893, 18945. Charles E. Monroe April 1876 Vice-President, 1889, 1890,1891, 1895

President, 18986. William McMurtie April 1876 Vice-President, 1897, 1898

President, 1900

C. Speakers at the ACS Organizational Meeting, April 6, 18761. Thomas Egleston April 1876 None2. Henry Carrington Bolton ? Vice-President, 9003. Albert R. Leeds January 1878 Vice-President, 1879 – 1888

* Clarke resigned after only two months of membership and did not rejoin the ACS until 1891. SeeJ. J. Bohning, “Fighting City Hall: The Role of Washington Chemists in the Nationalization of theAmerican Chemical Society,” 220th National Meeting, American Chemical Society, Washington,DC, August 21, 2000; HIST 006.

Page 23: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 101

cording to Isidor Walz, the subject was “before them solong” and “had been discussed thoroughly in private.”Evidently, none of the participants found it sufficientlyimportant to make any notes of their informal gather-ings (41). Yet, they continued in the face of very unfa-vorable conditions. On a local level, there was the NewYork Academy of Sciences, which was holding monthlymeetings for chemists in New York City. On the na-tional level, there was the AAAS chemistry subsection,which was growing and becoming increasingly active.During this period, however, Chandler and his commit-tee continued to pursue the formation of a new organi-zation when existing societies might have served theirpurpose. They set their sights on a local organization atfirst but quickly jumped to the national concept on thebasis of a poll in which less than half of the people con-tacted were in favor of the idea of forming just a localgroup. Except for the bold determination and imagina-tive thinking of eight chemists from the New York Cityarea, the American Chemical Society might not havecome into existence for some time after 1876, if at all(42).

ACKNOWLEDGMENTS

The assistance of William Jarvis (Lehigh University)and Jane Hartye (Stevens Institute of Technology) isgratefully acknowledged. The author is indebted toJeanne Kravitz (Wilkes College), whose knack for un-covering obscure citations is surpassed only be her un-canny ability to expedite accession to them. This workwas supported in part by a grant from the Wilkes Col-lege Research Fund.

REFERENCES AND NOTES

1. Presented in part at the 184th National Meeting of theAmerican Chemical Society, Kansas City, MO, Septem-ber 13, 1982; HIST 008.

2. Original copies of this invitation are rare.3. The March 27 invitation should not have come as a com-

plete surprise, however, since the self-appointed orga-nizing committee headed by Chandler had previouslydistributed inquires for interest in an organization ofchemists on a local (January 22, 1876) and national(March 22, 1876) scale. See Ref. 4.

4. Identical detailed records of the first meeting of theAmerican Chemical Society and the events that led to itmay be found in a) Am. Chem., 1876, 6, 401-406; b)

Proc. Am. Chem. Soc., 1876, 1 (Part I), 1-20. This andall subsequent quotations pertaining to the first meetingare taken from the account in the American Chemist.

5. The most complete record of this meeting and the eventsthat led to it were made by Professor Albert R. Leeds ofStevens Institute of Technology, who served officiallyas Recording Secretary. His faithful account was pub-lished in Am. Chem., 1874, 5, 35-44. The editors of theAmerican Chemist later collected from their pages all ofthe material related to the meeting that was published indifferent issues (see following references) and produceda separate volume: Proceedings of the Centennial ofChemistry, Collins, Philadelphia, PA, 1875.

6. The meeting in Northumberland is often mistakenly iden-tified as a meeting that celebrated the centennial of thediscovery of oxygen, or “The Priestley Centennial (See,for example, H. Skolnik and K. Reese, A Century ofChemistry: The Role of Chemists and the AmericanChemical Society, American Chemical Society, Wash-ington, DC, 1976, 4). However, Bolton’s now famousletter (Am. Chem., 1874, 4, 362) was quite clear in con-sidering Priestley’s achievement as just one of manyevents that marked the year 1774 “as the starting pointof modern chemistry.” The letter concluded by propos-ing “that some public recognition of this fact should bemade this coming summer. Would it not be an agree-able event if American chemists should meet on the firstday of August, 1874 [Priestley discovered oxygen onAugust 1, 1774], at some pleasant watering place, todiscuss chemical questions, especially the wonderfullyrapid progress of chemical science in the past 100 years?”In agreement with Bolton’s assessment of the centen-nial significance, the editors of the American Chemistused the “Centennial of Chemistry” banner for Bolton’sletter, as well as the responses his letter produced (Am.Chem., 1874, 4, 441-443), (only one correspondent dis-puted the Centennial of Chemistry label, but still sup-ported the meeting concept), the general circular fromthe organizing committee (Am. Chem., 1876, 5, 11-13),and the Leeds report of the meeting’s minutes (Note 5).While the Northumberland gathering did pay homageto Priestley (Am. Chem., 1874, 5, 43-51), the organizingcommittee remained committed to Bolton’s original sug-gestion and scheduled several speakers to review thecentennial aspects of chemistry. Thus, T. Sterry Huntdiscussed “A Century’s Progress in Theoretical Chem-istry” (Am. Chem., 1874, 5, 56-61); J. Lawrence Smithpresented “The Century’s Progress in Industrial Chem-istry” (Am. Chem., 1874, 5, 61-70); and BenjaminSilliman prepared a monumental listing of the “Ameri-can Contributions to Chemistry” (Am. Chem., 1874, 5,70-114, 195-209, 327-328). These three speakers weregiven positions of prominence, front row center, in thegroup photograph that was taken in front of a local build-ing. Therefore, it is appropriate and desirable to use the“Centennial of Chemistry” descriptor rather than the

Page 24: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

102 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

more restrictive “Priestley Centennial” in reference tothis meeting.

7. Am. Chem., 1874, 4, 441. Bodley proposed that “thecentennial gathering be around this [Priestley’s] grave,and that the meetings, other than the open-air one on thecemetery hilltop, be in the quaint little church built byPriestley.” (The gathering did visit Priestley’s grave,but the meetings were held in the local school rather thanthe church.) Bodley also noted in her letter that in hervaledictory address delivered to the 22nd graduating classof the Women’s Medical College on March 14 she hadalso called attention to 1874 as the centennial year ofchemistry.

8. Bolton made his proposal at the May 11 meeting of theChemical Section of the Lyceum. Other members of thecommittee were Charles F. Chandler, Henry Wurtz,Albert R. Leeds, and Charles A. Seeley. It was this com-mittee that was responsible for all the subsequent prepa-rations for the Centennial of Chemistry celebration. SeeProc. Lyceum Nat. Hist. New York, 1874, 2nd. Ser., No.4, 144-145.

9. This and all subsequent quotations pertaining to the Cen-tennial of Chemistry celebration are taken from Leed’saccount (Ref. 5).

10. New York Daily Graphic, August 6 1874, p 252. Seealso M. G. Waring, “The Priestley Centennial, TurningPoint in the Career of W. George Waring,” J. Chem.Educ., 1948, 25, 647-652.

11. Smith was referring to the American Association for theAdvancement of Science (formed in 1840 as the Asso-ciation of American Geologists and Naturalists) and theAmerican Academy of Arts and Sciences (formed in1780).

12. The Chemical Society (London) was the oldest of thefour organizations, beginning in 1841 with 77 members(coincidentally the same number present on the officiallist at the Centennial of Chemistry celebration). By 1874it had almost 800 members and had been publishing ajournal since 1847, first as a quarterly, and then, in 1861,as the monthly Journal of the Chemical Society. TheSociety received considerable support from Justus vonLiebig; and by 1874 notable members who had held of-ficial positions included Hofmann, Faraday, Crookes, andPerkin, while Dumas and Cannizzaro had presented thememorial Faraday lectures. (See T. S. Moore and J. C.Philip, The Chemical Society: 1841-1941, London,1947.) On the continent, the Chemical Society of Parisbegan in 1855 and started publishing the Bulletin de laSociété Chimique de Paris in 1858. Shortly afterHofmann returned to Germany he founded the DeutscheChemische Gesellschaft in 1866, and publication ofBerichte began in 1868. The Russian Physical-Chemi-cal Society was organized in 1869, and its journal com-menced in the same year. (See A. J. Ihde, The Develop-ment of Modern Chemistry, Harper and Row, New York,1964, 274-275 and references therein.) In spite of hiscomments, Smith was not adverse to using chemical jour-

nals for his own papers, having published 17 articles inthe American Chemist between 1870 and 1874.

13. For more information on the Chandlers, See R. D.Billinger, “The Chandler Influence in American Chem-istry,” J. Chem. Educ., 1939, 16, 253-257.

14. For more information on Smith, see a) C. A. Elliott, Bio-graphical Dictionary of American Science, GreenwoodPress, Westport, CT, 1979, 238; b) D. H. Wilcox, Jr., inWyndham D. Miles, Ed., American Chemists and Chemi-cal Engineers, American Chemical Society, Washing-ton, DC, 1976, 447-448; c) H. S. van Klooster, “Liebigand His American Pupils,” J. Chem. Educ., 1956, 33,493-497.

15. For more information on Silliman, see Ref. 14c and a)C. C. Gillispie, Dictionary of Scientific Biography,Charles Scribner’s, New York, 1970-1979, Vol. 13, 434-437; b) Ref. 14a, p 237; c) Ref. 14b, pp 438-440.

16. For more information on Clarke, see a) Ref. 14b, p 82-83; b) Ref. 15a, pp 292-294; c) C. E. Monroe, J. Am.Chem. Soc., 1935, 57, 20-30.

17. For more information on Horsford, see a) Ref. 14b, p230-231; b)Ref. 15a, Vol. 6, pp 517-518.

18. For more information on Edward Travers Cox, see a)James Grant Wilson and John Fiske, Ed., Appleton’sCyclopaedia of American Biography, D. Appleton & Co.,New York, 1887, Vol. 1, 757; b) The NationalCyclopaedia of American Biography, James T. White &Co., New York, 1904, Vol. 12, 328.

19. For more information on Van der Weyde, see W. R.Woolrich, The Men Who Created Cold, Exposition Press,New York, 1967, 118-119.

20. Smith, Silliman, and Horsford were charter members ofthe AAAS. The “official” list of the attendees at theCentennial of Chemistry celebration (“As far as I havebeen able to procure their names,” according to Leeds)is given in Ref. 5. Detailed membership lists of theAAAS, including the dates at which the member waselected, were published as part of the annual Proceed-ings. See, for example, Proc. Am. Assoc. Adv. Sci., 1874,23, xxvii ff.

21. For a complete list of officers of the AAAS from 1841to 1882, see Proc. Am. Assoc. Adv. Sci., 1882, 31, xix ff.

22. Proc. Am. Assoc. Adv. Sci., 1874, 22, 424. Clarke’s briefrecollection of these events can be found in C. A.Browne, A Half-Century of Chemistry in America,American Chemical Society, Philadelphia, PA, 1926,Chapter 3. More details are given by M. Benjamin inTwenty-Fifth Anniversary of the American ChemicalSociety, Chemical Publishing Co., Easton, PA, 1902, 86-98.

23. For the early history of the AAAS and its relationship toAmerican science, see S. G. Kohlstedt, The Formationof the American Scientific Community: The AAAS 1848-1860, University of Illinois Press, Urbana. IL, 1975; andS. G. Kohlstedt, The Establishment of Science inAmerica: 150 Years of the American Association for theAdvancement of Science, Rutgers University Press, New

Page 25: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 103

Brunswick, NJ, 1999. For the development of special-ization in the nineteenth century, see R. S. Bates, Scien-tific Societies in the United States, MIT Press, Cam-bridge, MA, 1965, 3rd ed., Ch. 3.

24. Proc. Am. Assoc. Adv. Sci., 1874, 23, 153-154.25. D. S. Martin, Popular Sci. Monthly, 1897, 51, 829.26. Clarke’s colleagues included G. F. Barker, T. S. Hunt,

S. W. Johnson, W. McMurtie, C. E. Monroe, and H. W.Wiley. Only Hunt was present at the Centennial ofChemistry celebration. Earlier, Hunt had responded toBolton’s idea of a centennial meeting by suggesting thatit be held concurrently with the AAAS meeting in Hart-ford starting August 12 (See Ref. 5).

27. Centennial of Chemistry attendees who joined the AAASat Hartford were E. B. Coxe, E. T. Capen, A. H. Elliott,A. R. Leeds, T. R. Punchon, C. W. Roepper, E. H. Swal-low, and E. Waller (See Ref. 20).

28. See. Ref. 20 for a complete listing and publication ofsome of these papers.

29. F. W. Clarke, Popular Sci. Monthly, 1875, 7, 365.30. See Proc. Am. Assoc. Adv. Sci., 1874, 23, xlvii, 99, 121.

The three new members from the Northumberland meet-ing were S. H. Douglass, S. A. Goldschmitt, and PersiforFrazer.

31. Twenty-Fifth Anniversary of the American ChemicalSociety, Chemical Publishing Co., Easton, PA, 1902, 39.At the time that Chandler issued this call, the migratoryAAAS meetings, where all the national chemical activ-ity was then taking place, had never been held in NewYork City since the AAAS had started 26 years earlier.The first meeting of the AAAS in New York City tookplace in 1887.

32. See Ref. 4; the two ACS histories in Ref. 22; and C. A.Browne and M. E. Weeks, A History of the AmericanChemical Society: Seventy-Five Eventful Years, Ameri-can Chemical Society, Washington, DC, 1952, for addi-tional information on the preliminaries to ACS forma-tion.

33. Chandler’s design for the new society was patterned af-ter the European societies. See Ref. 12.

34. For more information on Bolton, see Ref. 14 b, pp 40-41.

35. For more information on Egleston, see D. Malone, Dic-tionary of American Biography, Charles Scribner’s Sons,New York, 1930, Vol. 6, 50.

36. In spite of these comments, Leeds became a faithful andactive ACS member. For more information on Leeds,see Proc. Am. Chem. Soc., 1902, 24, 53-57.

37. See the biographical notice by G. G. Kunze in Trans.Am. Inst. Min. Eng., 1902, 31, 14.

38. For the membership list of the Academy, see H. L.Fairchild, A History of the New York Academy of Sci-ences, H. L. Fairchild, New York, 1887, 132ff.

39. J. J. Bohning, “Prestige vs. Practicality in Selecting theFirst President of ACS,” Chem. &. Eng. News, 1982,60, March 8, 1982, 31-34.

40. See the ACS history of Ref. 32 for more details.41. When C. A. Browne began preparing the 50th anniver-

sary history of the ACS (Ref. 22), he asked Charles Chan-dler for details of the August 1874–April 1876 period(uncataloged letters in the Charles F. Chandler Papers,Butler Library, Columbia University). Chandler’s onlyresponse was to submit a copy of the chapter he wrotefor the 25th anniversary volume (Ref. 22).

42. Drawing on the history of the first 20 years of the ACS,Sturchio has described the organization as a “gentleman’sclub” that “was the centerpiece of a network of metro-politan clubs and societies” which “served the socialinterests of the contingent of members with an interestin urban improvement and close ties to local commerce.”See J. L. Sturchio, “Charles Chandler, the AmericanChemical Society, and Club Life in Gilded New York,”presented at the Annual Meeting of the History of Sci-ence Society, New York City, December 27-30, 1979.

ABOUT THE AUTHOR

James J. Bohning is Professor of Chemistry Emeritus atWilkes College and a past chair of HIST. He directedthe oral history program at the Chemical Heritage Foun-dation for many years and was a science writer for theNews Service of the American Chemical Society. He iscurrently a Visiting Research Scientist and CESAR Fel-low in the Department of Chemistry, Lehigh Univer-sity, 6 E. Packer Ave., Bethlehem, PA 18015,[email protected].

Page 26: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

104 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

The discovery of radioactivity is usually credited tothe French physicist Antoine Henri Becquerel (1852-1908) who in 1896 discovered the rays that were emit-ted from uranium salts (1). Becquerel placed fragmentsof several uraniferous phosphorescent substances onphotographic plates wrapped in two sheets of blackpaper. In about 24 hours, when the plates were devel-oped, a silhouette of the phosphorescent substance ap-peared on the plate. Hence, it was inferred that ..thephosphorescent salts of uranium must emit radiationswhich are capable of passing through black paperopaque to ordinary light, and of reducing the silversalts of the photographic plate, even when the uraniumcompound has been completely sheltered from the light.The radiations emitted by the phosphorescent substancewere called “Becquerel rays.” At first, Becquerelthought that these radiations were a kind of invisiblephosphorescence, which was afterwards shown to bewrong. For this discovery Becquerel was awarded theNobel Prize for physics shared with Pierre and MarieCurie in 1903.

Mellor (2) in 1929, however, mentioned brieflythat Niepce de St. Victor had already observed thirtyyears earlier that uranium salts could affect photo-graphic plates in the dark. Mellor stated that Becquerelrepeated some experiments of Niepce de St. Victor inorder to find “if the property of emitting very penetra-tive rays is intimately connected with phosphorescence.In other words, does the principle of reversibility ap-ply? If Röntgen’s rays make a fluorescent substanceshine in the dark, will a fluorescing substance emit in-visible penetrative rays?”

NIEPCE DE SAINT-VICTOR AND THEDISCOVERY OF RADIOACTIVITY

Fathi Habashi, Laval University, Quebec, Canada

In his report to the French Academy of Science pub-lished in 1858, i.e., 38 years before Becquerel’s report,Niepce de St. Victor stated that (3):

A drawing traced on a piece of carton with a solutionof uranium nitrate…whether or not exposed beforeto light, and applied on a piece of sensitive paperprepared using silver chloride will print its image...Ifthe drawing made on the carton with the uraniumnitrate solution...is traced with large strokes, it willbe produced even at 2 or 3 cm further away from thesensitive paper.

With this statement one may conclude that Niepce deSt. Victor had discovered radioactivity before Becquerel.

Studies on the effect of uranium salts on a photo-graphic plate were started after the discovery of the metalby Klaproth in 1789. A few years later, since the newmetal gives vivid color to glass, a small uranium indus-try was started by the Austrian Government to produceuranium salts in Joachimsthal in Bohemia. Thus A. F.Gehlen (1775-1815) in Germany in 1804 reported onthe color change when an ethereal solution of uraniumchloride was exposed to light (4). Other workers at-tempted to use this phenomenon in a copying process inwhich paper was soaked with uranium salts then dried.The picture to be copied was then made fully visibleafter exposure to light by soaking in a solution of silvernitrate then washing. The picture was formed as theresult of the photochemical reduction of the uraniumsalt to uranium oxide which then produced metallic sil-ver by reduction of silver nitrate. Niepce de St. Victorshowed such pictures at the Third Exhibition of theSociété Française de Photographie in Paris in 1859. He

Page 27: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 105

was able later to give a colored tone to his copies. Forexample, a violet color was produced when the paperwas treated with chlorine, a green color when iron saltwas present, and a brown color when potassium ferro-cyanide was present. The process was widely publi-cized as the “uranium paper” or “uranium copying pro-cess” and pictures produced by this process were exhib-ited at the First Photographic Exhibition in Vienna in1864 (4).

Claude Félix Abel Niepce de St. Victor (1805-1870)was a cousin of Joseph Nicéphore Niepce (1765-1833),the amateur French scientist who, together with LouisDaguerre (1789-1851), developedthe photographicpicture. He wasborn in Saint Cyrnear Chalon-surSaône. He at-tended the schoolfor cavalry atSaumur and be-came Lieutenant ofDragoons in 1842.In 1845 he wastransferred to theParis MunicipalGuard quartered inthe barracks of thesuburb of SaintMartin, where heequipped a chemi-cal laboratory. Hisfirst work, presented to the Academy of Sciences in Parisin 1847 dealt with the condensation of iodine vapors ona copper plate engraving and the reproduction of theiodine vapor image onto metal. In the same year, hemade his invention of photography on glass.

In 1848, the barracks in which he lived were burnedand his laboratory was destroyed. In the same year hebecame Captain of his regiment and was elected Cheva-lier of the Legion of Honor and received also a prize oftwo thousand francs from the Société d’Encouragement.He improved the asphaltum process of his cousinNicéphore Niepce and greatly advanced etching on steel.When he was appointed Squadron Leader and Com-mander of the Louvre in Paris, he had time for his ex-periments and it was during this period that he investi-

gated photography with uranium salts. He was pen-sioned when Napoleon III came to the throne, and in hisretirement he continued his research on scientific pho-tography. He authored: Traité pratique de gravureheliographique sur acier et sur verre, Paris, 1856 andRecherches photographiques, Paris, 1858 (5).

In conclusion, the fact that Mellor had pointed outthat Niepce de Saint-Victor, an amateur photographer,should be credited with the discovery of radioactivityand not Antoine Henrie Becquerel, prompted the presentwriter to pursue the matter further. By reviewing thehistory of uranium salts in photography during the nine-teenth century, it could be concluded that Mellor’s pointof view is valid.

REFERENCES AND NOTES

1. A. Romer, Ed., The Discovery of Radioactivity andTransmutation, Dover Publications, New York, 1964.

2. J. W. Mellor, A Comprehensive Treatise on Inorganicand Theoretical Chemistry, Longmans, London, 1929,Vol. 4, 52-59.

3. “Un dessin tracé sur une feuille de carton avec une solu-tion d’azotate d’urane….exposé à la lumière ou isolé, etappliqué sur une feuille de papier sensible préparée auchlorure d’argent, imprime son image…Si le dessin faitsur le carton avec la solution d’urane… est tracé à grostraits il se reproduira à 2 ou 3 centimètres de distance dupapier sensible. Ce n’est donc pas à la phosphorescenceou à la fluorescence seule qu’on peut attribuer lapropriété remarquable que possède les solutionsd’urane”.

4. Quoted by F. Kirscheimer, Das Uran und seineGeschichte, Schwarzerbart’sche Verlagsbuchhandlung,Stuttgart, 1963.

5. J. M. Eder, History of Photography, translated from the3rd German ed. of 1905, Columbia University Press, NewYork, 1945, Dover Publications, New York, 1978.

ABOUT THE AUTHOR

Fathi Habashi is professor of extractive metallurgy atUniversité Laval, Faculté des Sciences et de Génie,Département de Mines et Métallurgie, Cité Universitaire,Québec, G1K 7P4, CANADA. He authored Principlesof Extractive Metallurgy in 4 volumes and edited Hand-book of Extractive Metallurgy in 4 volumes.

Niepce de St-Victor (1805-1870)

Page 28: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

106 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

Credit for the initial invention of the gas blowpipe hasbeen marked by an unusually large amount of contro-versy (1), involving bitter charges of plagiarism and in-tellectual dishonesty. An example of this involves theAmerican chemist Robert Hare,1781-1858 (2), whoclaimed that Edward Clarke, Professor of Mineralogyat Cambridge University (3), had falsely taken creditfor the invention of the improved gas blowpipe. Ed-ward Daniel Clarke, LLD (1769-1822), had published,in 1819, The Gas Blowpipe; or, Art of Fusion by Burn-ing the Gaseous Constituents of Water (4), in which hestaked out his claim for the primary credit for the in-vention of this apparatus. Hare maintained that he hadpublished the details of its construction fifteen yearsearlier then that of Clarke. This led Hare to ask thequestion in 1820, “Will Briton’s tolerate such conductin their professors?”

Hare maintained that he had developed a similargas blowpipe (5) in 1802 and therefore deserved solecredit. Benjamin Silliman Jr., Professor of Chemistryat Yale University, in his 1874 history of chemistry inthe United States as part of the celebrations markingthe centennial of Priestley’s discovery of oxygen, de-scribed the development of the Hare blowpipe as fol-lows (6):

Probably no chemical discovery made in this coun-try has been more generally cited or less generallyunderstood in its scientific significance, than the oxy-hydrogen blowpipe of Dr. Hare.

Who deserves priority for the invention of this most im-portant piece of laboratory equipment? Did Clarke

THE HARE-CLARKE CONTROVERSYOVER THE INVENTION OF THEIMPROVED GAS BLOWPIPE

Martin D. Saltzman, Providence College

knowingly appropriate the work of Hare without givinghim the proper credit? An examination of the claims andcounter claims will shed some light on this question.

The development of the mouth blowpipe has beenattributed to Florentine glass blowers in the middle ofthe seventeenth century. In the works of Robert Boyle,according to Partington, one can find the following de-scription of the blowpipe and its uses (7):

The small crooked pipe of either metal or glass, suchas tradesmen for its use call a blowpipe gives a jet ofair which when directed on the flame of a lamp orcandle produces a pointed flame which melts silverand even copper.

The blowpipe was first introduced into analytical chem-istry in 1750 by the Swedish chemist A. F. Cronstedt(1722-1765). Tobar Bergman (1735-1784) was the firstto describe the blowpipe extensively as a tool for theanalytical chemist. He also gave directions for its usewith various reagents such as soda, borax, and phos-phate for the study of earths, salts, combustible materi-als, metals, and ores. Bergman’s student J. G. Gahn(1745-1818) made further improvements in the designand use, and Berzelius was the foremost advocate of itsuse in the early nineteenth century.

According to Hare’s own account, he developed hisimproved blowpipe in Philadelphia in 1801. The impe-tus behind this invention was the need to produce highertemperatures than could be obtained by any contempo-rary apparatus. The principal means of producing hightemperatures at this time were through the use of mag-

Page 29: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 107

nifying glasses or furnaces (1). Lavoisier in 1782 hadconstructed an apparatus that allowed him to direct astream of oxygen or hydrogen or a combination of bothat a hollow piece of charcoal, to produce an intensesource of heat. In a dramatic demonstration in April,1782, Lavoisier, using a jet of oxygen directed at char-coal containing a piece of platinum, was able to melt it:a feat that had never been possible before. This experi-ment was repeated again on June 6, 1782 at Versaillesto an audience that included Louis XVI, MarieAntoinette, The Grand Duke Paul of Russia, and Ben-jamin Franklin. The details ofthe apparatus and the experi-ments performed were reportedto the Academie in a paper readin 1782 (8). Although Lavoisiermentioned the possibility of us-ing the oxygen-hydrogen mix-ture, there is no description ofthis type of blowpipe in hisTraite de Chemie of 1790 (9).

Gonzalez contends that thecredit for the invention of theoxy-hydrogen blowpipe be-longs to Lavoisier’s contempo-rary, J. B. G. Bochard de Saron(1730-1794) because ofLavoisier’s mention in his 1782paper of the device having beenmade by Bochard de Saron.Bochard claimed that if twostreams of gas—one being oxy-gen and the other hydrogen—were directed at charcoal, a veryintense flame could be pro-duced. However, Bochard deSaron never published his work,as he was to meet the same fateas Lavoisier. The apparatus designed by Lavoisier wastoo complex and expensive to be of any real practicalvalue; and the details of the Bochard device, which mayhave been simpler and easier to use, were never pub-lished. Thus a practical instrument which could pro-duce the intense heat needed for many laboratory situa-tions was not available when Hare began his work in1801.

What led Hare to believe that he could combinehydrogen and oxygen together under controlled condi-tions to produce a flame hotter than any known at thetime? The inspiration for the improved blowpipe came

from the lectures Hare attended, beginning in 1798, givenby James Woodhouse (1770-1809) at the University ofPennsylvania Medical School (10). Hare, the son of aPhiladelphia brewer, had become interested in chemis-try at a very young age. This interest was further peakedwhen he attended the lectures of Woodhouse, one of thesmall group of American chemists who had wholeheart-edly embraced the new chemistry of Lavoisier. He wasone of the leading critics, along with John Maclean, ofJoseph Priestley, proponent of the “old chemistry.” Al-though Priestley disagreed with Woodhouse, he never-

theless greatly admired his experi-mental abilities and considered himas the equal in skill and dexterity ofany chemist he knew in England andFrance. Maclean may also have in-fluenced Woodhouse to adopt thenew chemistry as the result of hislectures on the new chemistry, pub-lished in 1797 as Two Lectures onCombustion, which contained Con-siderations on the Doctrine ofPhlogiston and the Decomposition ofWater (11). Woodhouse performednumerous studies on the synthesisand decomposition of water follow-ing the principles laid down byLavoisier. The commonly acceptedbelief was that, when hydrogen com-bined with oxygen, a large amountof caloric (heat) was released. Hareprobably heard of this explanation inone of Woodhouse’s lectures, andthis provided the inspiration for hisblowpipe.

Hare believed that if he couldsomehow introduce a stream of pureoxygen and hydrogen, the ignition of

these two gases would produce temperatures not previ-ously obtainable. Lavoisier had developed a gasometerthat would allow the use of oxygen and hydrogen butwas inconvenient, as Hare noted (5):

(It) is too unwieldy and expensive, for ordinaryuse...Being sensible of the advantage which wouldresult, from the invention of a more perfect methodof supplying the Blow-Pipe, with pure or atmosphericair, I was induced to search for means of accom-plishing this object.

Hare’s years in the brewery were put to good use, as herealized how the humble brewer’s keg, being rugged,tight, and cheap, could form the basis of the new im-

Robert Hare

Page 30: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

108 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

proved blowpipe. Hare divided a barrel approximatelyeighteen inches wide and thirty-two inches high into twocompartments, the upper being fourteen inches and thelower sixteen inches. A sheet of copper containing acopper tube and had a set of water tight leather bellowscomprised the bottom compartment. An iron rod wasattached to the bellows via the copper pipe, which couldbe raised or lowered by a handle attached to the rod andthus act as a pump. The upper portions of the apparatuswere separated into two chambersthat could be filled with oxygen andhydrogen, which could be com-pressed by the built-in pump. Twopipes were used to carry the gases,and the streams met at a burningcandle or lamp placed on a stand.Examples of the tips used by Hareare illustrated in Footnote 12 of Ref.1. Hare had designed and con-structed the prototype of whatwould become the oxygen-hydro-gen blowtorch, still used today forwelding and cutting metals (12).Miles writes of Hare’s invention asfollows (13):

…New methods of attaining hightemperatures had been interestingchemists for a third of a century;oxygen and hydrogen had beenknown for a quarter of a century;and Hare lived in an age of giants.Lavoisier, Priestley, Black, and others had the factsneeded to construct an oxy-hydrogen blowtorch, butnone of them experienced the flash of genius thatcame to Hare.

Hare estimated that his blowpipe could be made for abouttwenty dollars, and a simplified version that would bemore efficient than the mouth blowpipe or the enameler’slamp for about four dollars. The significance of thisinvention and its usefulness prompted the reprintingverbatim of Hare’s 1802 Memoir of the Supply and Ap-plication of the Blowpipe in Tilloch’s PhilosophicalMagazine in London (14) and in the Annales de Chemie(15) in France. Thus, there can be little question thatthe claims made by Hare have a great deal of validity. Itshould be noted that Hare’s description of the blowpipewas also printed in pamphlet form, and there is no wayto judge how many copies were published and what theextent of the distribution was at the time. Just how manyread the description of Hare’s device in Tilloch’s publi-cation or in the Annales de Chemie is also impossible togauge.

Daniel Edward Clarke, the second son of a countryvicar, entered Jesus College, Cambridge in 1786. Hegraduated in 1790 and was then employed as the tutorand companion to the sons of several wealthy families.As a student, Clarke developed an interest in mineral-ogy, which he fostered as the result of the travels inEurope in his role as companion and tutor. He collectedmany different types of specimens of minerals and plantsand produced several volumes describing his travels.

The mouth blowpipe wouldhave been known to Clarke asit was a primary tool used inmineralogical analysis.Clarke’s travels took him alsoto Greece, Egypt, Turkey, andPalestine; and he was abroadwhen Hare’s paper describinghis blowpipe was published.Clarke donated many of theartifacts he had collected,which included several marblestatues, to Cambridge Univer-sity. Cambridge Universityawarded Clarke the degree ofLLD in 1803, and Jesus Col-lege appointed him as a seniortutor in 1805. In the same yearClarke became an ordainedpriest in the Anglican Church,which was a usual prerequisitefor appointment to a professor-

ship at Cambridge and Oxford Universities. Clarke wasthe vicar of two parishes until 1808, when a universityprofessorship in mineralogy was created and he was ap-pointed to the chair. It is most unlikely that Clarke mayhave come upon the Hare’s original 1802 paper as wellas subsequent papers of Hare and his collaborators.Abstracts as well as indexes were nonexistent in thisera and it would have been only by chance that Clarkemay have been aware of Hare’s work.

It seems clear that the motivation for Clarke’s in-terest in the blowpipe came as the result of his researchin mineralogy rather than, in the case of Hare, as a re-sult of his study of chemistry. In the preface to Clarke’sThe Gas Blowpipe, he writes the following (3):

The public is already in possession of the principalfacts, which have led to the history of the Gas Blow-pipe. The different claims made on the part of theChemists of this Country and of America, as to theoriginality of the invention, have rendered it desir-able to remove a few existing doubts, and to show,

E. D. Clarke

Page 31: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 109

by a summary memorial, the progressive steps bywhich the philosophical apparatus, here delineatedand described, has reach its present state of utility.

Clarke stated that his particular blowpipe design wasthe result of a conversation he had with the instrumentmaker, Mr. Newman of Lisle Street, Leicester Square,London, in early 1816. Mr. Newman had constructed ablowpipe for a Mr. Brooke, who described the design inthe May, 1816 Annals of Philosophy (16). Brooke statedthat he had produced this new blowpipe because of thegreat inconvenience he had experienced with the com-mon mouth blowpipe. Brooke’s design consisted ofusing either a copper or iron vessel in which air wasforced by a plunger and allowed to escape through avery narrow stopcock. This stream of air when ignitedthen produced an intense source of heat. Newman modi-fied Brooke’s original design so that gases like oxygencould be introduced. This, Brooke concluded would(16):

render it more extensive in its application to chemi-cal purposes, and probably so as to supersede the useof the common gasometer.

In his design Clarke used the principle “of an explosivemixture of gases propelled through a common aperturefound from a common reservoir (3).” This principle,Clarke stated, was well-known and the result of an in-vestigation made of gas illumination by ProfessorTennant and Dr. William Wollaston and presented in apublic lecture delivered in Spring, 1814. ThomasThomson in a letter to Clarke dated April 9, 1817 wrotethat he, Thomson, had done experiments in 1800 thatformed the basis of Hare’s blowpipe. Thomson hadabandoned his work because of the problem of explo-sions, which wrecked his apparatus; and the work wasnever published. In addition, Clarke stated that the prin-ciple of the oxy-hydrogen blowpipe had been demon-strated in chemical lectures at Cambridge for at least adozen years prior to the publication of his book. As towho was demonstrating the blowpipe is not clear fromClarke’s book.

The danger of an explosion from a retrograde move-ment of the flame was well known, and Wollaston hadwarned Clarke that his experiments could put him ingreat danger. Clarke “persisted in making them, nar-rowly escaped being killed by frequent bursting of hisapparatus (3).” Clarke also consulted Sir HumphreyDavy in May, 1816 with his idea of mixing oxygen andhydrogen and passing the mixture through a capillarytube prior to ignition. Davy, who had developed theminer’s safety lamp using the narrow capillary principle,

replied on July 8, 1816. Davy stated that he had triedthe experiment and that “there would be no danger inburning the compressed gases by suffering them to passthrough a fine thermometer tube, 1/80 of an inch diam-eter, and three inches in length (17).” The inherent ex-plosive nature of the oxygen-hydrogen mixture requireda container that could withstand an explosion. A col-league of Clarke at Cambridge, the Rev. J. Cumming,the Professor of Chemistry, developed a safety cylin-der, which made the device much safer to use (3):

It becomes therefore a duty of gratitude to lay greaterstress upon that part of the invention to which, be-yond all doubt, he is indebted for his present safety.Had it not been for the circumstance, it would havefallen to the lot of some other person to have writtenthe history of the Gas Blow-pipe, and to have ren-dered it rather tragical than amusing.

Just because Hare had reported experiments he per-formed in 1802 with a device of his own design, whichhappened to use hydrogen and oxygen, was not suffi-cient in Clarke’s opinion to negate his own claim of origi-nality.

The significant difference in the Clarke apparatusis that the oxygen and hydrogen are premixed. In Hare’sblowpipe they were in separate containers and flowedthrough two different apertures before combining. AsClarke stated (3):

But the intensity of the heat is incomparably greaterwhen the gases, after compression, are propelled andburned in a mixed state; because the due proportionnecessary for forming water is then constantly andequally maintained: whereas an excess, whether onthe side of the hydrogen or of the oxygen, not onlytends to diminish the temperature, but, if it be muchincreased on the side of the oxygen, infallibly extin-guishes the flame.

This greatly improved device with all the improvementsthat had been made by Clarke, he insisted, should havethe name of “Gas Blowpipe.” Robert Hare cried foulbecause in his 1802 paper he had not only described thedesign of the blowpipe but also reported the results ofmany of the experiments that had been performed withthe blowpipe. Further modifications of the original de-sign of the blowpipe as well as additional experimentswere the subject of a paper read by Hare on June 17,1803 at the Chemical Society of Philadelphia and pub-lished in 1804 (18). Hare pointed out in 1820 that hehad reported experiments using his blowpipe in 1802that Clarke claimed were new results (4):

Hence, until plagiarism had given them a new shape,and perhaps false gilding, they were totally over-

Page 32: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

110 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

looked in his compilations. He neither treated of thepure earths as susceptible of fusion, nor of platinumas susceptible of volatilization, until many years af-ter I had proved them to be so, and promulgated myobservations.

One example from Hare’s work that was also reportedby Clarke as his own original work was the fusion oflime and magnesia (magnesium oxide) using anthracitecoal as the reducing agent. By exposure to the gaseousflame of the coal, both magnesia and lime exhibitedstrong symptoms of fusion. The former assumed aglazed and somewhat globular appearance; the latterbecame converted into a brownish semi-vitreous mass.

Benjamin Silliman in 1812 (19) reported that withHare’s blowpipe he was also able to fuse lime and mag-nesia ignited in a covered platinum crucible (20). Clarkeobtained similar results but failed to credit the work ofHare and Silliman. Hare noted (4):

Notwithstanding the previous publicity of these re-sults obtained by my friend and myself, Dr. Clarkein the following note endeavors to convey an impres-sion of the incompetency of my apparatus to fuse limeand magnesia. Note 5, page 46. Professor Hare inAmerica could not accomplish the fusion either oflime or magnesia by means of his hydrostatic blow-pipe. See Annales de Chimie, tome 45 page 126. Butwhy overlook Silliman’s experiments? It is more-over strange that an English writer should refer hisreaders to the French Annales in preference to a Lon-don magazine, for a memoir which he knew to bepublished in both.

What is one to make of these conflicting counter-claimsconcerning the blowpipe? Benjamin Silliman, Jr., writ-ing in 1874 (6), made the point that a distinction needsto be made between discovery of the principle and theactual invention. Certainly, Hare deserves credit for thedevelopment of a device based upon the use of soundchemical principles. The production of copious amountsof heat by the combination of oxygen and hydrogen alsoformed the basis of Clarke’s invention. However,Clarke’s rationalization based upon the chemical natureof volcanism was certainly wrong.

Chemical theories of volcanism were very much invogue in the first half of the nineteenth century. Clarke,as well as many of his contemporaries, believed that thevolcanic fire, as he referred to it, led to the decomposi-tion of water. According to him, these gases were com-pressed and their subsequent combustion produced theenergy associated with the violent events that occur in avolcanic eruption. Thus the volcano is really a giantblowpipe, according to Clarke. Humphrey Davy sug-

gested that the origin of volcanic activity and subterra-nean heat was the action of water on sodium and potas-sium in the interior of the earth

Hare also realized that Lavoisier had overlookedan important point in his investigations, in that he con-sidered the greatest amount of caloric would be releasedif the substance were placed upon charcoal and only ajet of oxygen gas introduced. Hare had reasoned that aneven greater amount of heat would be produced if thesubstance were burned on a solid support in a stream ofhydrogen and oxygen gas. As Hare wrote in 1802 (5):

It soon occurred that these desiderata might be at-tained by means of a flame, supported by the hydro-gen and oxygen gasses; for it was conceived that,according to the admirable theory of the French chem-ists, more caloric ought to be extricated by this thanany other condition.

In hindsight we can see an error in Hare’s work con-cerning the concept of heat. It must be rememberedthat Rumford’s paper on the origin of heat had not beenpublished until 1800 in the Transactions of the RoyalSociety and was only slowly being accepted. Hare canbe faulted on this point but not his basic chemical in-sight, which Clarke lacked.

Clarke deserves credit for the design of his blow-pipe. The relative ease of the use of the Clarke blow-pipe and the contributions stemming from its use inchemistry and mineralogy redeem somewhat his scan-dalous disregard and diminution of the work of Hare.

ACKNOWLEDGMENT

I am most grateful to Dr. W. A. Smeaton, EmeritusReader in the History of Science, University of Lon-don, for all of his help in preparing this paper. BillSmeaton passed away on January 22, 2001. He will bemissed by all his friends in the chemical history com-munity.

REFERENCES AND NOTES

1. E. L. Gonzalez, “Bochard de Saron and the Oxy-Hydro-gen Blowpipe,” Bull. Hist. Chem., 1989, 4, 11-15.

2. For a biography of Clarke see: W. Wroth, “EdwardDaniel Clarke,” The Dictionary of National Biography,Oxford University Press, Oxford, 1917, Vol. 4, 421-24.A recent paper dealing with Clarke and Davy may be ofinterest: B. P. Dolan, Blowpipes and Batteries:

Page 33: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 111

Humphrey Davy, Edward Daniel Clarke, and Experimen-tal Chemistry in Early Nineteenth-Century Britain,”Ambix, 1998, 45, 137-162.

3. E. D. Clarke, The Gas Blowpipe or the Art of Fusion byBurning the Constituents of Water, Cadell and Davies,London, 1819.

4. R. Hare, Strictures on a Publication entitled, Clarke’sGas Blowpipe, R. DeSilver, Philadelphia, PA, 1820, 15;simultaneously also published in Am. J. Sci., 1820, 2,281-302.

5. R. Hare, Memoir on the Supply and Application of theBlowpipe, Chemical Society of Philadelphia, Philadel-phia, PA, 1802; reprinted in E. F. Smith, Chemistry inAmerica, D. Appleton & Co., New York, 1914, 152-178.

6. B. Silliman, Jr., “American Contributions to Chemis-try,” American Chemist, 1874, 5, 70-114,195-209.

7. R. Partington, A History of Chemistry, Macmillan, Lon-don, 1962, Vol. 2, 535.

8. A. L. Lavoisier, “Sur un moyen d’augmenterconsiderablement l’action du feu et de la chaleur, dansles operations chimiques,” Mem. Acad. Roy. Sci. Paris,1782, 457-476.

9. A. L. Lavoisier, Traite elementaire de chimie presentedans un ordre noveau d ‘apres les couvertes modernes,English translation of 1790 by R. Kerr, reprinted byDover Publications, New York, 1965, 474-479.

10. For a biography of Woodhouse, see: E. F. Smith, Chem-istry in America, D Appleton & Co., New York, 1914,76-108. For Maclean see: M. D. Saltzman, “JohnMaclean,” American National Biography, Oxford Uni-versity Press, New York and Oxford, 1999, Vol. 14, 267-268.

11. J. Maclean, Two Lectures on Combustion..., Dobson,Philadelphia, PA, 1797.

12. For a diagram of Hare’s blowpipe see Ref. 1.13. W. D. Miles in E. Farber, Ed, Great Chemists,

Interscience, New York and London, 1961, 419-434.14. R. Hare, “Memoir on the Supply and Application of the

Blowpipe,” Tilloch’s Philos. Mag., 1802, 14, 238-245,298-308.

15. R. Hare, “Memoire sur l’usage du chalumeau, et lesmoyens de l’alimenter d’air ....,Extrait par P. A. Adet,”Ann. Chim., 1802, 65, 113-138.

16. H. I. Brooke, “Description of a New Blowpipe,” Annalsof Philosophy, 1816, 7, 367.

17. E. D. Clarke, “Account of Some Experiments made withNewman’s Blowpipe...,”J. Sci. Arts, 1817, 2, 104-123.

18. R. Hare, ”Account of the Fusion of Strontites and Vola-tilization of Platinum...,” Trans. Am. Philos. Soc., 1804,6, 99-105.

19. B. Silliman, “Experiments on the Fusion of Various Re-fractory Bodies by the Compound Blowpipe of Dr.Hare,” Trans. Am. Philos. Soc., 1812, 3, 328-399.

20. The melting points of lime and magnesia are 2614 and2852o C, respectively. W. A. Smeaton has pointed out tome that such temperatures were not obtainable in ananthracite furnace. Even platinum with a melting pointof 1772o C was very difficult to melt when these experi-ments were performed. The results of Hare, Silliman,and Clarke may have been affected by impurities in theirsamples.

ABOUT THE AUTHOR

Dr. Martin D. Saltzman is Professor of Natural Science,Providence College, Providence, RI 02918.

NATIONAL HISTORICCHEMICAL LANDMARKS

http://www.acs.org/outreach/landmarks

Page 34: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

112 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

THE CONTRIBUTIONS OF CYRUSMOORS WARREN TO THE ANALYSIS OFHYDROCARBONS

Martin D. Saltzman, Providence College

Saturday, August 27, 1859 was aday that profoundly changed theworld, for on that afternoon in thesmall town of Titusville, Pennsyl-vania, the first well drilled specifi-cally to produce liquid oil came in.Petroleum offered a simpler andcheaper way to produce kerosenethan the method of that time,which involved the destructivedistillation of certain types of coal(1 ). Kerosene, introduced in theearly 1850s, had revolutionized in-dustrial societies by offering asource of light that allowed for alonger working day. Even thehumblest homes could be lightedin a way that was previously un-available.

Hydrocarbons that had beenproduced by the distillation ofvarious substances such as tars, pitch, and coal re-mained a mystery because of the failure to separateand analyze these mixtures(2). Cyrus Moors Warren(1824-1891) made significant contributions to the de-velopment of techniques for fractional distillation andtheir application to the analysis of complex hydrocar-bon mixtures. He did his work in the early 1860s inhis own private laboratory in Boston and reported theresults at meetings of the American Academy of Artsand Sciences (3). Much of the work was later pub-lished in the academy journals (7,13,15,17,18,19,20).

Warren’s life combined careers asa basic research oriented chemistand as a successful industrialist.

Warren was born on January15, 1824 in West Dedham, Massa-chusetts, the fifth son and eighthof eleven children born to Jesse andBetsey (Jackson) Warren. His fa-ther was a blacksmith by trade andan inventor by avocation, who forall his ability was never very suc-cessful financially. In 1829 thefamily moved to Peru, Vermont, inhopes of improving their situation.This was followed by anothermove to Springfield, Vermont,where his father started a foundrybusiness. The foundry was de-stroyed by fire in 1839, leaving thefamily in dire financial straits.Cyrus’ formal education to this

point was spotty at best, but both he and his older brotherSamuel spent as much time as possible continuing theireducation by self-study. The brothers pledged to eachother that if they became successful, first one and thenthe other would be able to complete his education. In1846 Samuel Warren perfected a process to improve themanufacture of a waterproof roofing material made bycoating paper with coal tar. He opened a plant to makethe “tar paper” in Cincinnati in that year, and Cyrusjoined him as a partner in the business in 1847.

Page 35: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 113

Coal tar is the byproduct of the process used to pro-duce coal gas for lighting. This process had been in-vented by William Murdoch (1754-1839) in the late 18th

century, and plants to manufacture this gas were foundin cities all over Europe and the United States after 1810.When the Warrens entered the tar paper business, coaltar was thought to have little commercial value; and thebrothers were able to enter into long term contracts topurchase it at very low prices.

The tar paper business proved to be an almost im-mediate success. Samuel was the first to leave and pur-sue his education (4). As the business prospered, otherWarren brothers were brought in; and finally, in 1852,at the age of 28 Cyrus, by this time married with a fam-ily, began his higher education. Because of his interestfrom childhood in the natural sciences, he chose to pur-sue his studies at the Lawrence Scientific School, affili-ated with Harvard University. His initial interests wereboth in zoology and chemistry, and his studies were su-pervised by the eminent Swiss naturalist Louis Agassiz(1807-1873). Although attracted to natural history,Warren decided his interest really lay in chemistry.Warren received his S.B. degree in 1855 with high dis-tinction and was elected to Phi Beta Kappa. He had theunique honor of being the first graduate of the LawrenceScientific School ever elected to this scholarly society.

Wishing to complete his education and being ofindependent means, Warren embarked on the chemicalgrand tour of Europe with stops in London, Paris, Heidel-berg, Munich, and Berlin. He carried out various re-search projects in the laboratories of such titans as EmilKopp, Justus Liebig, Robert Bunsen, and Heinrich Rose.His work in Rose’s laboratory on titanium and zirco-nium resulted in the publication of several papers in 1857(5). This European sojourn was to last for several yearsbefore Warren finally returned to America in 1858. Therewere few American chemists who had achieved as ex-tensive training in various aspects of chemistry as CyrusWarren had by 1858.

In 1855 the Warrens had moved their industrialoperations to New York, establishing the Warren Chemi-cal and Manufacturing Company to produce coal tarproducts. Contracts had been secured from the numer-ous gas works in the city to provide the raw material atvery low cost. When William Henry Perkin’s syntheticaniline dye mauve, introduced in early 1858, became aninstant success, the Warrens profited handsomely bybeing a major source of benzene, aniline, and other coaltar intermediates used by the aniline dye industry (6).

Cyrus Warren remained in New York until 1861,when he moved to Boston with the intention of produc-ing the highly profitable aniline dyes himself. As thecenter of the American textile industry, New Englandwas a logical location for a dye plant. With the rawmaterials available at very low cost from the Warrencoal tar distilleries in New York, this venture seemed alogical extension for the Warren Company. However,this was not to be. When it became obvious to the En-glish manufacturers such as Reed & Holliday that anAmerican firm was going to compete in the lucrativeAmerican market, they proceeded to unload their anilinedyes on the American market at prices with which theWarren Company could never compete. With failure ofthe dye venture, Warren turned his attention from busi-ness to basic research at the laboratory he had built asan addition to his Boston home. It had always beenWarren’s intention to devote himself to chemical re-search as a part of this new venture in Boston. Warrenwas able to outfit his laboratory with the latest and bestequipment available and set to work. One of the firstproblems to which Warren turned his attention was thedevelopment of better methods of fractional distillation,with an eye towards its application to the analysis ofhydrocarbon mixtures. As Warren wrote (7):

...simple fractional distillation...affords but very im-perfect and unsatisfactory results, and not infre-quently leads to gross errors and misconceptions...The want of a more efficient process for effectingsuch separations has long been recognized. Thereare numerous natural and artificial products, of thehighest scientific interest, such as petroleum...ofwhich it may at least be said that we have but veryimperfect knowledge

The innovation that Warren introduced was the controlof the temperature in the still head. This was achievedby a worm coil whose temperature was controlled inde-pendently and through which the distillate had to pass.Figure 1 is the drawing of the apparatus shown inWarren’s paper (7):

In the belief that no process of fractioning at all analo-gous to mine has ever been employed in scientificresearch, and that I am not in any way directly in-debted to any devices of my predecessors,....I maysay, however, that I have found no record of anyone’s ever having employed the oil bath and a sepa-rate fire to regulate a heated condenser, this beingthe essential feature on which the superiority of myprocess is based; adapting it at once to both high andlow temperatures, and for the most delicate work.

The worm coils used by Warren were made of copper invarious sizes from 10 feet by 1/2 inch to as small as 1

Page 36: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

114 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

foot 6 inches by 1/4 inch in length. A complete descrip-tion of the procedures used to conduct fractional distil-lation is given in the paper. Warren concluded his paperwith the following remarks (7):

...I can say that as regards bodies not decomposed byheat in distillation, I have not yet found a mixture socomplex that it cannot be resolved by this processinto its proximate constituents so completely, thatthese shall have almost absolutely constant boilingpoints.

The fractional distillation technique was so successfulthat Warren was able to separate benzene from coal tarnaphtha in such a state of purity after several distilla-tions that it froze upon cooling in the receiver into asolid mass of crystals. Warren’s fractional distillationtechnique would be used in the research laboratory aswell as in the chemical industry particularly in Europe(8):

...an expert traveling in Europe in 1870 found theprocess in common use there in the distilleries of tar.In some instances, the managers of these works knewthat they were using Warren’s invention, while oth-ers professed ignorance as to its origin, while freelyadmitting its excellence.

Warren had received a patent for his technique (9), butthis did not seem to deter those who infringed upon itwithout paying any royalties.

Having developed this efficient method of fraction-ation, Warren began to apply it to various types of hy-drocarbon mixtures. Among the first studied were thevolatile hydrocarbons found in light coal tar naphtha.The first attempt to analyze these low-boiling fractions(bp 80-175oC) had been done in 1849 by CharlesMansfield, a student of Hofmann at the Royal Collegeof Chemistry in London (10). Mansfield reported (11)that there were four major components that showed aconstant incremental difference of 30o in boiling point.Mansfield believed these four components to be aro-matic hydrocarbons, with the first of the series beingbenzene. Faraday had isolated benzene in 1825 fromoil gas, and Hofmann had shown it to be present in coaltar naphtha in 1845. Mansfield proposed the other threecomponents, in order of increasing boiling point, to betoluene (113o), cumene (140-145o), and cymene (170-175o C). The only compound that could be definitivelyidentified in the mixture was benzene, the lowest-boil-ing component. The others had not been obtained in asufficient state of purity for any kind of definitive analy-sis. In 1855 Church, reporting his analysis of this mix-ture, proposed (12) that there are five components, eachhaving a constant difference in boiling point of 22o.

Warren hoped to produce a definitive analysis ofthe light coal tar naphtha by using his fractional distilla-tion technique. In addition, he had developed what hethought to be a better and more accurate method fordetermining boiling points. To ensure the accuracy ofhis study he used a mixture of coal tars produced by sixdifferent gas works, derived from different varieties ofcoal, both imported and domestic. Since the gas worksproduced coal tar residues in quantities of upwards of50,000 barrels per annum, Warren firmly believed thatthe naphtha he was to fractionate contained all the pos-sible components (13):

...fractioning in this case was conducted in all respectsas there described, and continued until the whole ofthe naphtha taken, boiling between 80o and 170o Chad accumulated at the four points...80o, 110o, 140o,and 170o , or so nearly the whole that the intermedi-ate quantities had become too small to admit of be-ing further operated upon...I may here remark thateach of the sample gallons employed, when subjectedto my process of fractioning, was found to contain,in variable proportion, all of the constituents of thenaphtha.

The compounds isolated by Warren were identified inorder of increasing boiling points as benzene (80o), tolu-ene (110o), xylene (140o) and cumene (170o C). Thispaper seemed to verify the preliminary work that hadbeen done by Mansfield and cast doubt upon the analy-sis published by Church.

An early empirical discovery was that the hydro-carbons in a homologous series had boiling points thatdiffered by a constant increment. Hermann Kopp (1817-1893) reported in papers published in 1842 and 1845(14) that there was a definite 19oC increment for eachmember of the homologous hydrocarbon series he stud-ied. In 1855 (15), Kopp, revisiting the question of thecorrelation of boiling points in homologous series ofhydrocarbons, noted that there seemed to be certain ex-ceptions for a difference of C2 H2 in some cases. Othercontemporaries of Kopp had suggested various ration-ales for calculating the differences in boiling points byusing certain structural assumptions. Warren was con-cerned with the lack of reliability of boiling points re-ported in the literature and hence the confusion concern-ing the effect on physical properties in an homologousseries (15):

It may be hoped, however, that the superior meanswhich my process furnishes for separating mixturesof liquids, will lead to the accumulation of reliablefacts of sufficient number and variety for a profit-able review of this question in its different bearings,which, from its importance, it clearly merits.

Page 37: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 115

Warren obtained his boiling points by immersing thebulb of the thermometer in the liquid rather than in thevapor, as was the standard procedure. He argued thatthe temperature of the boiling liquid and the vapor shouldbe the same as long as there is a regular pattern of boil-ing. To ensure uniform boiling Warren used as boilingstones pieces of sodium when possible or else coke (15):

My experience has shown that,when irregular ebullition is effec-tually prevented, the temperature ofthe vapor from a boiling liquid ismore liable to lead to an erroneousdetermination of the boiling-point,than the temperature of the liquiditself.

Warren used Pennsylvania petro-leum as the source of homologousseries of hydrocarbons in one of hisstudies. Prior to the beginning ofWarren’s study in 1861 there had notbeen any complete analysis of thevolatile hydrocarbons from petro-leum. The more volatile compo-nents had little commercial value atthis time and would only becomeimportant in the 20th century withthe development and increasing useof the internal combustion engine.Warren thought it would be of greatinterest to fractionate the most vola-tile hydrocarbons from crude petro-leum as well as from the syntheticcoal oil, which had been used tomake kerosene prior to the availability of liquid petro-leum. The question was whether the mixture of hydro-carbons found in Pennsylvania crude and that producedby distillation of coal were the same or different in viewof the contrasting sources. In all of his studies Warrenpointed out that each of the fractions isolated boiledwithin, at most, a 1.5o range and left no residue. Twoidentical series of hydrocarbons were obtained from bothsources, which differed by 30o for each increase of CH

2,

thus showing that the petroleum (also known as rockoil) and coal oil were identical mixtures of hydrocar-bons. These hydrocarbons were identified as pentane,bp (30.2o), hexane (61.2o), heptane (90.4o), octane(119.5o) and nonane (150.8oC). In a truly Baconianmanner, Warren wrote of his accomplishment as follows(15):

As no one had preceded me in the investigation ofthese substances, my mind was as far as possible un-

biased as to the boiling points of the constituents ofthese mixtures. I was, however, aware of the beauti-ful relation between elementary constitution and boil-ing point, which Kopp had discovered.

Thus it would appear that no matter the source of thehydrocarbon mixture the homologous series of hydro-carbons were identical and had an incremental boiling

point of 30o. Seeking to extend his investigations fur-ther, Warren studied the boiling points of the nitro com-pounds of the aromatic series. These showed a differ-ence of only 14o, much less than the 30o for the hydro-carbons. Warren had no proper explanation for theseresults.

The fractional distillation technique was also usedto determine the composition of petroleum produced inBurma, known in the trade as Rangoon petroleum. Crudeoil that seeped to the surface in Burma, just as in Penn-sylvania, was collected and distilled at a refinery inRangoon to produce kerosene for the Asian market. TheBritish chemists Warren De La Rue and Hugo Mullerhad attempted to determine the composition of the crudeBurmese petroleum but without success. They had re-ported (16) their failure in being able to separate thenaphtha fraction into its components in 1857. Warrenwrote (17):

Warren’s distillation apparatus

Page 38: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

116 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

The labors of De La Rue and Muller at once occurredto us as furnishing an extreme instance, and it wasdetermined to test the new process with materialswhich, as these chemists had shown, could not beunraveled by the old processes of analysis.

Volatile components from the Rangoon petroleum werecollected in a range from 150o to 270o. The light naph-tha fraction was not included in the sample Warren hadobtained from a London merchant, and therefore thevolatile components started to boil at 150o. A sample ofthe light naphtha fraction could not be obtained, andthus only the fraction above 150o was subjected to frac-tional distillation. This produced seven distinct frac-tions boiling between 170o and 240oC. Each of the frac-tions was treated with sodium until no further reactionoccurred and then analyzed for carbon-hydrogen con-tent. The entire volatile fraction had approximately 86%carbon and 14% hydrogen, corresponding to an empiri-cal formula of CH

2. As to the composition of this mix-

ture Warren speculated that it was a series of homolo-gous aliphatic hydrocarbons, C

10-C

14. The presence of

naphthalene in the mixture was also inferred.

Warren was not satisfied with the methods avail-able for the analysis of carbon and hydrogen content bycombustion in air. He modified the usual method ofcombustion, substituting pure oxygen for air for theanalysis (18):

By a very simple device I entirely obviate the dangerof explosion; viz. the combustion tube is closelypacked with asbestos, or other inert substance, andyet so loosely as to leave free passage for gasesthrough the interstices.

Warren provided a thorough and detailed description ofthe apparatus and procedures that led him to concludethat “the results obtained are extremely accurate anduniform.” This method, Warren believed, was superiorto the conventional methods in use (18):

...its greater convenience, economy of time, avoid-ance of excessive heat, neatness, etc.; it will, at least,not be found inferior to other methods; but, on thecontrary, I think preferable, as affording greater se-curity against failures and errors.

In the course of his work on petroleum, Warren isolatedseveral sulfur and chlorine containing compounds. Thepresent methods for the analysis of chlorine and sulfurin organic compounds were not satisfactory in Warren’sopinion. Sulfur analysis involved a problem that someof the element present in the compound was convertedto sulfuric acid rather then sulfate. Warren turned thisproblem to his advantage by employing the reaction of

sulfur compounds with lead peroxide (PbO2) to ensure

that sulfuric acid would be the only product formed.With some modification Warren was able to use the sameapparatus he had developed for carbon-hydrogen con-tent to analyze organic compounds simultaneously forcarbon, hydrogen, and sulfur. The method gave excel-lent results for the known compounds such as carbondisulfide (19).

The analysis of chlorine in organic compounds pre-sented the problem of the conversion of the whole ofthe chlorine content, but without allowing absorption ofthe carbonic acid and water produced from the carbonand hydrogen in the sample. Oxides of the heavy met-als were known to have a strong affinity for chlorine,but they did not interfere with the carbon analysis. Af-ter trial and error, Warren found that the “brown oxideof copper” (CuO2.2H2O) was the best reagent to use forthe analysis of chlorine. He also obtained excellent re-sults by using known compounds (20).

In 1866 Warren accepted an appointment as Pro-fessor of Organic Chemistry at the Massachusetts Insti-tute of Technology, but he resigned the position aftertwo years because it involved excessive demands uponhis time. His ability to continue his research and hiscontinued involvement in the chemical industry as aconsultant were seriously hampered by his duties at MIT.

The chaotic business cycles that followed the CivilWar finally claimed all of his time, and so the grandprogram for the separation and analysis of petroleumwas never realized. Though independently wealthy, hefelt a continuing obligation to his partners, especiallyafter the death of his brother in 1880. The strain of over-work broke his health and eventually led to a debilitat-ing stroke in 1888. He died on August 13, 1891 at hishome in Manchester, Vermont.

In his will he generously provided funds for theAmerican Academy of Arts and Sciences as well asHarvard University for the funding of basic chemicalresearch. In addition, his will provided funds towardthe construction in Boston of a permanent library andmeeting hall for the American Academy of Arts andSciences. Although Cyrus Warren’s career in basic re-search spanned only a few years, he made significantinroads into an understanding of complex hydrocarbonmixtures, but his contributions do not appear to be widelyknown by the American chemical community.

Page 39: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 117

REFERENCES AND NOTES

1. For further information on coal oil see M. D. Saltzman,“Coal Oil: Prelude to the Petroleum Age,” Chem. Heri-tage, 1998, 16, 12-13, 42-43.

2. For the definitive history of distillation see: R. J. Forbes,Short History of Distillation, E. J. Brill, Leiden, 1948,and revised edition, 1970. For a study of the influenceof distillation on the petroleum industry see: M. D.Saltzman, “The Art of Distillation and the Dawn of theHydrocarbon Society,” Bull. Hist. Chem., 1999, 24, 53-60.

3. The American Academy of Arts and Sciences wasfounded during the American Revolution for the pur-pose of the advancement of the arts and sciences. Amongits founders was John Adams, the future second Presi-dent of the United States. The Academy was incorpo-rated by an act of the Massachusetts legislature on May4, 1780. The Academy today has a dual function, whichis to honor achievement in science, scholarship, the arts,and public affairs, as well as programs to respond to theneeds and problems of today’s society based upon thecontributions that can be made by its members. TheAcademy has its headquarters in Cambridge, Massachu-setts and publishes the quarterly journal Daedalus.

4. Samuel Warren studied law at the Cincinnati Law Schoolafter he left the business and was admitted to the bar.He gave up the practice of law for the ministry andpreached in Brookline, Massachusetts and later in Lon-don.

5. C. M. Warren, “Ueber einige Zirkonerde- und Titan-Säure-Verbindungen: A. Ein neues SchwefelsäuresZirkonerde-Salz, etwas Kali enthaltend. B.Schwefelsäures Zirkonerde-Kali,” C. Schwefelsäures Ti-tan-Säure-Kali, Poggendorffs Ann.” 1857, 102, 449.

6. For a definitive study of the dye industry see: A. S. Travis,The Rainbow Makers: The Origin of the Synthetic DyeIndustry in Western Europe, Lehigh University Press,Bethlehem, PA, 1993.

7. C. M. Warren, “On a Process of Fractional Condensa-tion, Applicable to the Separation of Bodies having SmallDifferences between their Boiling-Points,” Mem. Am.Acad. Arts Sci., 1864, 9, 121-134.

8. Anonymous Obituary Notice, Proc. Am. Acad. Arts Sci.,1893, 27, 391-403.

9. U.S. Patent 47,235, Apr. 11, 1865. (Improved Appara-tus for Distilling Petroleum, &c).

10. For a biography of Charles Mansfield see: The Dictio-nary of National Biography, Oxford University Press,London, 1937, Vol 12, 975-76.

11. C. Mansfield, “Research on Coal Tar,” Quart. J. Chem.Soc., 1849, 1, 244-68. Mansfield’s work set the stagefor the synthetic organic chemical industry by showingthat it was possible to distill benzene easily in large quan-tities from coal tar. Benzene was first isolated by Fara-day in 1825 from the oily residue found in cylinders ofan illuminating gas made from whale oil. Mansfield’sdistillation of benzene from coal tar made it possible toproduce substances like aniline in commercial quanti-ties. Mansfield died as the result of a fire that occurredwhen he was distilling benzene in 1855.

12. A. H. Church,” On the Benzole Series-Determination ofBoiling points, Philos. Mag., 1855, 9, 256-260.

13. C. M. Warren, “On the Volatile Hydrocarbons from Coal-tar Naphtha, Oil of Cumin, and Cumminic Acid. Part I.Hydrocarbons from Coal-Tar Naphtha,” Mem. Am. Acad.Arts Sci., 1867, 9, 137-149.

14. H. Kopp, Ann. Chem. Pharm., 1842, 41, 79, 169; 1845,55, 177.

15. C. M. Warren,” On the Influence of C2H

2 upon the Boil-

ing Points in Homologous Series of Hydrocarbons, andin some Series of their Derivatives; with Critical Obser-vations on Methods of taking Boiling-points,” Mem. Am.Acad. Arts Sci., 1864, 9, 156-176.

16. W. De La Rue and H. Muller, “Chemical Examinationof Burmese Naphtha or Rangoon Tar,” Philos. Mag.,1857, 13, 512-517.

17. C. M. Warren and F. H. Storer, “Examination of theNaphtha Obtained for Rangoon Petroleum,” Mem. Am.Acad. Arts Sci., 1867, 9, 208-216.

18. C. M. Warren, “On a Process of Organic ElementaryAnalysis, by Combustion in a Stream of Oxygen Gas,”Proc. Am. Acad. Arts Sci., 1866, 6, 251-261.

19. C. M. Warren, “On a New Process for the Determina-tion of Sulphur in Organic Compounds, by combustionwith Oxygen Gas and Peroxide of Lead,” Proc. Am.Acad. Arts Sci., 1866, 6, 472-476.

20. C. M. Warren, On a New Process of Organic Elemen-tary Analysis for Substances containing Chlorine,” Proc.Am. Acad. Arts Sci., 1868, 7, 84-96.

Note added in proof: There seems to be a question con-cerning the statement made concerning the priority ofthe first successful oil well.. See: F. Habashi, “TheFirst Oil Well In The World?” Bull. Hist. Chem., 2000,25, 64-66, and response (p 67).

Page 40: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

118 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

RECIPROCAL SOLUBILITY INFLUENCE INSALT MIXTURES. THE CONTRIBUTIONS OFWALTHER NERNST AND OF ARTHUR NOYES

John T. Stock, University of Connecticut

In 1887, Walther Hermann Nernst (1864-1941; NobelLaureate 1920) began his highly successful career asan assistant to Wilhelm Ostwald (1853-1932; NobelLaureate 1909) at the University of Leipzig. Ostwaldstrongly supported the ionic theory and had extendedit, especially to the dissociation of weak electrolytes.By 1889, Nernst had established the principles of elec-trode potential, familiarized in the textbook “Nernstequation,” and hence of the emf of a reversible cell.That the solubility of a salt is diminished by the addi-tion of another salt having an ion in common with thefirst was well known. While briefly at the Universityof Heidelberg, Nernst developed a quantitative theoryof this common ion effect, supported by experimentswith uni-univalent strong electrolytes (1).

Nernst’s simplest case assumes that the salts arecompletely dissociated in solution. If to the saturatedsolution of MX (molar concentration m0 ) MY or NX isadded to concentration x, then the now smaller solubil-ity m is given by

m ( m + x ) = m02 (1)

The symbols are those used by Nernst. The quantitym

02 became known as the solubility product of MX.

(Nowadays, Ksp is the usual symbol for a solubility prod-

uct). Unless the solubility of MX is very small, equa-tion (1) must be modified to allow for incomplete dis-sociation:

ma ( ma + xa¢ ) = m02a

02 (2)

Here a0 is the degree of dissociation of MX whensaturated in water and a the value after the addition of,e.g., NX, which is dissociated to the extent a¢. SvanteArrhenius (1859-1927) had pointed out that the mixingof two solutions with one ion in common does not alterthe degree of dissociation of the salts (2). For example,a mixture of equivalent solutions of a pair of alkalinehalides has a conductivity equal to that of the mean ofthe conductivities of the individuals (3). Therefore whensolving equation (2) with respect to m, Nernst felt justi-fied in making the assumption that a and a¢ were equal,so that m is given by:

mx m a

a

x= − + +

2 4

0

2

0

2

2

2

(3)

Then, with CH3COOAg (solubility 0.0603M at 16o C)

as MX and known concentrations of AgNO3 or

CH3COONa as additive, Nernst found that the measured

and the calculated solubilities were similar. The largestconcentration of additive was 0.230 M, when the solu-bility of CH

3COOAg fell to approximately one-third of

its solubility in water.

Nernst commented that solubility measurementsmight throw light on the existence in solution of bothM+ and MX-. He also theorized that, when two com-mon-ion salts form a single saturated solution, their solu-bilities, m1 and m2, must be less than their m0 values.

Page 41: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 119

Nernst provided no experimental support for these ideas.Before his move to the University of Göttingen in 1890,Nernst was able to place the verification and extensionof mutual solubility problems in the hands of ArthurAmos Noyes (1866-1936).

Noyes graduated from the Massachusetts Instituteof Technology (MIT) in 1886 and continued research inorganic chemistry to obtain his M.S. and an assistant-ship in 1887 (4). By the summer of 1888, Noyes hadplanned to study under Johann Friedrich Wilhelm Adolfvon Baeyer (1835- 1917) in Munich, but there was nolaboratory space for him. Instead, he went to Leipzig,aiming to study organic chemistry under JohannesWislicenus (1835-1902). However, having heardOstwald’s lectures on physical chemistry, Noyes decidedto work in his laboratory. Eventually, Noyes becamethe first American to obtain a Ph.D. under Ostwald’sguidance.

Noyes began his studies with a survey of the prin-ciples of mutual salt solubilities (5). He pointed outthat a consideration of the undissociated portion of MXleads to a very simple alternate expression for m :

m = m0 (1 – a

0 ) / (1 – a ) (4)

However, this simplicity is offset by a greater sensitiv-ity to any error in a. A positive error obviously yields avalue of m that is too large; Noyes commented that ifthe same value of a is inserted in equation (3), m is foundto be too small.

To extend Nernst’s studies, Noyes chose the sys-tems listed in Table 1. The substrates were chosen tohave qualitatively similar low solubilities. Experimentswith TlBr were made at 68.5o C because the solubilityof this salt was unacceptably low at ambient tempera-tures. The solutions used were thermostatted, usuallyat 25o C. Classical gravimetric and volumetric methodswere used for the analyses. Equation (3) or its modifi-cations was used to calculate the expected solubilities.

Noyes tabulated the results obtained with the variouspairs of electrolytes. In all cases, the observed solubil-ity was greater than the calculated value. Noyes notedthat the difference between the two values becamegreater as the concentration of additive was increased.This is, of course, an expected result of a greater totalionic strength. Noyes felt that the differences might becaused by the use of conductivity measurement to finddegrees of dissociation. Certainly the “zero concentra-tion” equivalent conductivity values were at that timethe best estimates. The Kohlrausch square root rule forfinding such values by linear extrapolation from mea-surements made at finite concentrations did not appearuntil 1900 (6).

Noyes chose the pair TlCl and TlSCN for a quanti-tative examination of Nernst’s conclusions concerningthe solubilities of two common-ion salts in a single satu-rated solution. He found decreases in solubility of ap-proximately 26% for the more soluble TlCl and 28%for TlSCN.

Nernst had indicated that solubility measurementsmight throw light upon the state of dissociation of ter-nary salts. For example, is AgSO4

- present in a satu-rated solution of Ag2SO4? To find an indirect answer tothis kind of problem, Noyes added equivalent amountsof TlNO3, BaCl2, and Tl2SO4 to saturated solutions ofTlCl. He found that the solubility of the latter salt waslowered by the same extent with each of the additives.Because Tl+ and Cl- , but not TlSO4

- and BaCl+, controlthe solubility of TlCl, he concluded that there were nosignificant amounts of the double ions. Otherwise, lessCl- or Tl+ would be available from the additives.

Walther Nernst

Page 42: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

120 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

Noyes turned to systems that have no ions in com-mon. He examined the effect of KNO3 and ofCH3COONa on the solubility of TlCl and found thatthese increased the solubility, as we would expect fromthe present-day concept of ionic strength. However,Noyes attributed the increase to the formation of someundissociated KCl by an exchange reaction with TlCl,thereby increasing the solubility of the latter. At addi-tive concentrations greater than about 0.03M, the solu-bility was less than expected from his mode of calcula-tion. He could not explain this but suggested that it mightbe due to inaccurate dissociation values, or to the as-sumption that all of the salts had equal dissociation con-stants.

Having examined the common-ion effects in solu-tions of sparingly soluble salts, Noyes considered sucheffects in solutions containing only freely soluble salts.He critically surveyed the solubility results obtained bynumerous earlier workers, pointing out certain peculiari-ties such as the formation of double salts. He concludedthat the mutual solubility principles were also obeyedin the necessarily more concentrated solutions, althoughthe results might lack quantitative exactitude. Noyesconsidered that the application of solubility measure-ments to the determination of degrees of dissociationwas one of the most important results of his work. Withthe elimination of m0 by combining equations (3) and(4) he obtained the relationship:

a = [ (m0 – m) / x ][ 1 + (1 + x / m ) 1/2 ] (5)

At this point he stressed that a is the dissociation of eachsalt in the presence of the other and is equal to the dis-sociation undergone by each salt at concentration (m +x ). For a single salt, the application of the law of massaction leads to the relationship (1 – a).n = k.a2.n2 ,

where n is the normality of the solution and k is the fac-tor that Noyes termed the “dissociation constant “ ofthe salt. This relationship, a form of the Ostwald “dilu-tion law,” is applicable to solutions of weak electrolytes.When, as was usual, the degree of dissociation a wasobtained from conductivity measurements, attempts toapply the above relationship to solutions of strong elec-trolytes resulted in failure. Noyes illustrated this by theresults obtained with solutions of TlNO

3. These results,

along with those found when a was calculated from solu-bility measurements, are listed in Table 2.

Although not completely independent of concentration,the results in column 2 show a degree of constancy thatis completely absent from the results obtained from con-ductivity data. Noyes concluded that the determinationof dissociation from solubility measurements was themost reliable method then known. Solutions of pairs ofsalts that differ completely, i.e., have no ion in com-mon, such as the pair TlCl and KNO

3, were examined.

The results of several experiments showed that the solu-bility increase agreed approximately with that calculatedfrom known dissociation constants. In validating andextending Nernst’s concepts, Noyes produced muchvaluable solubility data. He showed decisively that thecommon ion effect was a highly significant phenom-enon in quantitative chemistry.

After receipt of his Ph.D. in 1890, Noyes returnedto MIT, to become a great teacher of chemistry in itswidest sense (4). This activity was accompanied by ex-tensive research, especially on solutions of electrolytes.Noyes’ doctoral studies must have led him to suspectthat rather more than the law of mass action and theArrhenius ionic theory were needed to explain some ofthe phenomena encountered. By 1903 he had begun toconsider the possibility that anomalies in electrolyticconductivity might be attributable to electrical chargeson the ions and not to specific chemical affinity (7,8).

Table 2. “Dissociation Constant” ( k ) of TlNO3

Concn. k, solubility K, conductivity0.0161 5.45 7.110.0366 5.45 4.840.0617 5.21 4.030.100 5.07 3.320.150 4.81 2.90

Table 1. Common-ion Systems

Substrate Solubility, m0

Additive

PhCH:CBr.COOHa 0.0176 PhNH.CO.COOH

AgBrO3b 0.00810 KBrO

3or AgNO

3

TlBrc 0.00869 TlNO3

TlSCN 0.0149 TlNO3 or KSCN

TlCl 0.0161 TlNO3 or HCl

a Common ion, H+ b24.5o C c68.5o C

Page 43: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 121

He retained his interest in the solu-bilities of electrolytes and the prop-erties of their solutions. This in-cluded the effect of salts on thesolubility of other salts, an effectthat Noyes had investigated in hisdoctoral studies.

In 1911, Noyes collaboratedwith Research Associate (later,Assistant Professor) WilliamCrowell Bray (1879-1946) to pro-duce a set of papers that criticallyexamined and greatly extended hisearlier work (5). The first of thesepapers showed that the solubilityprinciples initially adopted byNoyes are subject to considerabledeviations (9). Additional studiessince 1890 had shown that, in asolution saturated at 40oC with both TlCl and TlSCN,the concentration of nonionized TlCl is about 15% less,and the ionic product [Tl+][Cl -] about 5% greater thanin a solution of TlCl alone. Further, the solubility-prod-uct principle failed badly when a salt with a commonbivalent ion was added. For example, although the solu-bility of PbCl2 was decreasedslightly by a small addition ofPb(NO3)2, further additionscaused the solubility to becomegreater than in water. The au-thors proposed to make use ofthe thermodynamically relatedconcept of activity, A, intro-duced by Gilbert NewtonLewis (1875-1946) in 1907(10). In a solution in equilib-rium with solid salt BA, the re-lationships AB x AA = constantand ABA = constant are strictlytrue. The activity coefficient,i.e., the ratio of activity to con-centration, A/C, is assumed tobe unity at infinite dilution.

The experimental workdescribed in the second paperwas shared by Noyes’s threecoauthors (11). Great care wastaken to ensure the purity of thevarious salts and the tempera-ture was maintained at 25 ±

Effect of salts on the solubility of other salts

0.02oC (or 20oC, where indicated). The results of thenumerous solubility determinations are summarized asin Fig. 3. The lowered solubility of Tl2SO4 by the pres-ence of TlNO3, a salt with a univalent common ion,agreed qualitatively with the ionic product principle.

However, with the bivalentcommon ion salt Na2SO4,the solubility, reduced byonly 0.3% in 0.1 NNa2SO4, was actually in-creased in higher concen-trations of this additive.

Research in the thirdpaper, which dealt with theeffect of other salts on thesolubility of TlCl, was di-rected by Bray (12). Thesolubility of TlCl itself,found to be 16.07 mM perliter, was only 0.13% lowerthan the original valuefound by Noyes (5). Theincreased solubility causedby the presence of KNO3 orof K2SO4, salts without acommon ion, was attrib-uted to the formation bymetathesis of nonionizedTlNO3 or Tl2SO4. At com-parable concentrations, theArthur Amos Noyes

Page 44: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

122 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

latter salt is less highly ionized than TlNO3, so thatTl2SO4 has the greater effect. The decrease in solubilitycaused by the addition of salts with a common ion is inaccord with that expected from the ionic product prin-ciple.

Bray undertook the discussion of the results de-scribed in the three foregoing papers (13). Included inthe various tables is a listing of the degrees of ioniza-tion, at concentrations from 0.01 to 0.25 N, of the sevensalts that were used in the studies. The values were ob-tained by precise conductometric measurements.

The composition of the various solutions saturatedwith TlCl was calculated on the assumption that, for eachsalt, the values in mixtures depend only on the equiva-lent ion concentration (Σi). To show the relationshipsin dilute solutions more clearly, Σi, the correspondingconcentrations of nonionized TlCl, and the values of theionic product [Tl+][Cl -], were expressed logarithmically.Examples are given in Table 3; the effects of the addi-tives BaCl2, KNO3 and KCl were also examined.

In all cases, an increase in Σi caused a decrease inthe concentration of nonionized TlCl and an increase inthe ionic product [Tl+][Cl -]. Although the experimentsinvolved univalent and bivalent salts, with and withouta common ion, the results were remarkably similar. Thissupported the assumption that the total ionic strength ina mixture primarily determines the ionization of uni-univalent salts. In the case of less soluble salts, it wasconcluded that their solubility products would be prac-tically constant in the presence of small amounts of othersalts.

Further analysis of the results led to the conclusionthat, in the case of TlCl, deviations of the ionizationfrom the law of mass action are due more to the abnor-mal behavior of the nonionized salt than to that of theions. In more concentrated solutions of a single salt,the activity coefficient, A/C, decreased more rapidly withfurther increase in concentration. Available measure-ments of the emf of Tl+-ion concentration cells supportedthis conclusion.

Theories and equations that were based upon theconcept of the complete dissociation of strong electro-lytes in solution were eventually developed by others(14). Then Noyes was able to use his acquired experi-mental results to check these developments.

In 1913, Noyes began a part-time association withThroop College, which became the California Instituteof Technology (Cal Tech). This association became full-time in 1919, when Noyes moved from MIT, with theintention of making Cal Tech a great center for educa-tion and research. This he certainly achieved. Troubledby ill health during the latter part of his life, Noyes diedon June 3, 1936. He had never married; his estate wasbequeathed to Cal Tech for support of research in chem-istry.

ACKNOWLEDGMENT

This work was partially carried out under the ResearchFellowship Program of the Science Museum, London

Table 3Effect of Added Salts on the Dissociation of TlCl

Added salt Log 103 x [Â i] Log 104 x [TlCl] Log 105 x [Tl+][Cl -]None 1.1559 1.2443 1.3115Tl

2SO

41.3703 1.6158 1.31831.5911 1.0931 1.33391.8090 1.0362 1.3642

TlNO3

1.4649 1.1446 1.33451.6709 1.0962 1.36901.9141 1.0386 1.4096

K2SO

41.4819 1.1265 1.32431.7157 1.0492 1.34011.9296 0.9850 1.36122.3030 0.8854 1.4126

Page 45: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 123

REFERENCES AND NOTES

1. W. Nernst, “Ueber gegenseitige Beeinflussung derLöslichkeit von Salzen,” Z. Phys. Chem., 1889, 4, 372-383.

2. S. Arrhenius, “Theorie der isohydrischen Lösungen,” Z.Phys. Chem., 1888, 2, 284-295..

3. P. Chroustchoff and V. Pachkoff, “Sur la ConductibilitéÉlectrique des Dissolutions Salines Contenant desMélanges de Sels Neutres,” C. R. Hebd. Séances Acad.Sci, Ser. C, 1889, 108, 1162-1164.

4. L. Pauling, “Arthur Amos Noyes,” Biog. Mem. Natl.Acad. Sci. USA, 1958, 31, 322-346.

5. A. A. Noyes, “Ueber die gegenseitige Beeinflussung derLöslichkeit von dissociierten Körpern,” Z. Phys. Chem.,1890, 6, 241-267.

6. F. Kohlrausch and M. E. Maltby, “Das elektrischeLeitvermögen wässringer Lösungen von Alkali-chloriden und Nitraten,” Wiss, Abhl. D. Phys. Techn.Reichanstalt, 1900, 3, 156-227.

7. A. A. Noyes and W. D. Coolidge, “The Electrical Con-ductivity of Aqueous Solutions at High Temperature,”Proc. Am. Acad. Arts Sci., 1903, 39, 163-219.

8. A. A. Noyes, “The Physical Properties of Aqueous SaltSolutions in Relation to the Ionic Theory,” Science,1904, 20, 577-587.

SECOND CONFERENCE ON THE HISTORY AND HERITAGE OFSCIENTIFIC AND TECHNICAL INFORMATION SYSTEMS

November 15-17, 2002

sponsored by Chemical Heritage Foundation, Philadelphia

For more information on the conference, suggested topics, and scholarships, visit CHF’s website at:http://www.chemheritage.org or contact HHSTIS2 Program Committee, Chemical Heritage Founda-tion, 315 Chestnut Street, Philadelphia, PA/USA 19106. [email protected]

9. A. A. Noyes and W. C. Bray, “The Effect of Salts on theSolubility of Other Salts. I,” J Am. Chem. Soc., 1911,33, 1643-1649.

10. G. N. Lewis, “Outlines of a New System of Thermody-namic Chemistry,” Proc. Am. Acad. Arts Sci., 1907, 43,259-293.

11. A. A. Noyes, C.R. Boggs, F. S. Farrell, and M. A. Stewart,“The Effect of Salts on the Solubility of Other Salts. II,”J. Am. Chem. Soc., 1911, 33, 1650- 1663.

12. W. C. Bray and W. J. Winninghoff, “The Effect of Saltson the Solubility of Other Salts. III,” J. Am. Chem. Soc.,1911, 33, 1663-1672.

13. W.C. Bray, “The Effect of Salts on the Solubility of OtherSalts. IV,” J. Am. Chem. Soc., 1911, 33, 1673-1686.

14. P. Debye and E. Hückel, “Zur Theorie der Elektrolyte,”Phys. Z., 1923, 24, 185-206; 305-325.

ABOUT THE AUTHOR

John T. Stock, recipient of the HIST Dexter Award in1992, is Professor Emeritus of Chemistry, University ofConnecticut, Storrs, CT 06269-4060.

Page 46: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

124 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

Introduction and Motivation

In recent years there has been a resurgence in interestin philosophical aspects of chemistry, dating from theearly 1990s (1-3). It is gratifying to see that this devel-opment has not been confined to analytical philosophybut has frequently spilled over into issues concerningthe history of chemistry. Giunta’s recent article (4) istherefore of great interest because it represents an ex-ample of work that approaches philosophy of chemis-try from the historical direction. Giunta’s article onNewland’s periodic system, or whether indeed it canbe called a system, is a bold attempt by a chemist-his-torian who is willing to venture a philosophical analy-sis based on an episode in the history of chemistry.

There has been a good deal of discussion in theliterature over whether chemical periodicity should bereferred to as the periodic table, periodic system, orperiodic law. In addition there have been articles inphilosophy of chemistry which have attempted to clarifythe terms theory, model, or law in the context of chem-istry as distinct from physics (5, 6). Giunta’s article,which provides an analysis of these terms, should there-fore be of interest to philosophers as well as historiansof chemistry.

While accepting that a chemical audience may notwish to agonize over the use of terms like system andself-consistency, I believe that the clarification of termsis one area in which a philosophical analysis of chemi-

A PHILOSOPHICAL COMMENTARY ONGIUNTA’S CRITIQUE OF NEWLANDS’CLASSIFICATION OF THE ELEMENTS

Eric R. Scerri, University of California, Los Angeles

cal concepts has a powerful role to play. In addition, itcannot be denied that Giunta’s main intention is to ex-amine whether Newlands produced a “system” or not.I therefore make no apologies for undertaking an analy-sis of precisely what Giunta means by the word “sys-tem” in this context. After doing so I turn to Newland’swork and conclude that he deserves more credit than hehas been accorded by Giunta.

Commentary

I believe that Giunta makes some important points aboutNewland’s overall role in the discovery of the periodicsystem but also that he introduces some misconceptionswith which I shall express some friendly disagreement.Although Giunta states, at the outset, that he intends toexamine the work of Newlands “from a contemporarypoint of view,” I think there may be some problems withthis proposal as I hope to show. Giunta also says that heis not concerned with reconstruction of the process ofNewlands’ discoveries but only with appraising the va-lidity of his writings. He appears to want to concentrateon the logic of discovery, rather than the context of dis-covery, to use a distinction that was once popular inphilosophy of science. Such a goal is of course laud-able, especially given the excessive emphasis on con-text, and in particular the social context, of scientificdiscoveries that one finds in recent science scholarship.

Page 47: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 125

System, Organization, and Self-consistency

Although chemists might have an intuitive feel for termslike system, organization, and internal consistency, I askfor their indulgence in pausing to analyze these notions.I feel justified in doing so since Giunta has made themmajor criteria in his critique of Newlands.

Giunta begins his critical analysis by stating thatwhereas Mendeleev and Newlands referred to a “peri-odic law” he, Giunta, intends to follow the author vanSpronsen in preferring to use the term “periodic system(7).” However, Giunta believes that van Spronsen’sdefinition of periodic system is inadequate and in needof strengthening. He tells the reader that van Spronsendefines a periodic system as (4):

…a system of all the (known) elements arranged ac-cording to increasing atomic weight in which the el-ements with analogous properties are arranged in thesame group or column.

but that earlier in his work van Spronsen refers to (7):

“facets of a true periodic system” including additionalcriteria, for example a distinction between maingroups and sub-groups, and provision of vacantspaces for undiscovered elements.

Giunta proposes to define a periodic system as some-thing lying between these two versions, given by vanSpronsen, namely (7):

a periodic system of the elements consists of a self-consistent arrangement by atomic weight of all theknown elements, which systematically displaysgroups of analogous elements.

Giunta claims that his own definition places consider-able emphasis on “organization” and “internal consis-tency” although he fails to provide any additional crite-ria to indicate just what these features might mean inthis context. Finally, he asserts that he does not requirehis own sense of system to be one “free from error.”

Giunta’s attempt to improve on van Spronsen’sdefinition(s) of the term periodic system is, I believe,somewhat problematic (8). Whereas Giunta implies thathis own definition is stronger than van Spronsen’s firstdefinition, it is, in fact, weaker. By failing to includethe word “increasing,” as a qualifier for atomic weight,Giunta unwittingly admits even earlier systems such asthat of Gmelin. In 1843, a remarkable 26 years prior tothe first of Mendeleev’s published systems, this chem-ist classified all the then known elements and obtaineda very successful grouping of analogous elements (9).Gmelin’s only failing was that he did not strictly adhere

to increasing atomic weights, something which Giuntadoes not explicitly specify as an important criterion, al-though this omission may well have been accidental.

On the other hand, the claim by Giunta that his owndefinition of “system” is weaker than van Spronsen’ssecond definition also appears to be mistaken becausevan Spronsen does not require a system to be free fromerror. Had van Spronsen done so, he would have ex-cluded many of the precursors of the modern periodicsystem, which he has so painstakingly documented inhis book while considering them as genuine systems.

I turn to considering what Giunta claims to haveadded to van Spronsen’s definitions. Giunta requiresthat the qualities of being “self-consistent” and “sys-tematic” should be present in a system displaying allthe known elements. However, we are not told whatself-consistency actually implies in the context of a clas-sification of the elements.

As for the second requirement, I believe this maybe circular. Since the definition given by Giunta wasintended to define “system,” it can hardly be illuminatedby the statement that a system shall be systematic! Thefurther stipulation that his own definition “places con-siderable emphasis on organization and internal consis-tency” does not appear to clarify his position since noth-ing in the definition actually states how this claim is tobe realized.

Is the Periodic System a Theory?

Giunta then introduces a further requirement, namelythat a periodic system should also fulfill the criteria givenby one George Lachman of what constitutes a theory.Whereas Giunta promises, in his title, to analyze whythe work of Newlands does not represent the discoveryof a system, I believe he proceeds to cloud the issue byinvoking a further set of criteria which are not intendedfor “systems” but for scientific theories. Whereas up tothis point the discussion had focused on whetherNewlands had produced a system, Giunta then appearsto suggest that “theory” and “system” are synonymousterms. I would like to explain why I believe these termsto be far from synonymous, especially in the context ofchemical periodicity.

The term system is very frequently used to describechemical periodicity, but to the best of my knowledgechemical periodicity has never been regarded as a theory.In fact, according to some authors, chemistry does notpossess any genuine theories of its own (10). This inci-

Page 48: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

126 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

dentally is a reason sometimes given for the lack of inter-est in chemistry shown by philosophers of science.Chemical periodicity is instead referred to almost ex-clusively as a system because it is essentially a classifi-cation system and, as such, does not depend upon anytheoretical underpinning for its success or otherwise.This is why I believe that Giunta may be mistaken ininvoking Lachman’s criteria for theories in a contextwhere one is simply not dealing with any theory.

Indeed Giunta then introduces into the reckoningsome even further criteria, due to Thomas Kuhn, with-out discussing whether they are consistent with the viewsof Lachman as well as his own previously stated defini-tion of “system.” Having thus set up at least three typesof criteria, and without, I would claim, any supportingarguments involving examples from Newlands work,Giunta is prepared to declare (4):

Although Newlands’ work does not meet the criteriafor a periodic system set out above, his contributionswere substantial.

Perhaps the most charitable interpretation for this con-clusion would be that Giunta intends to show later inthe article how Newlands fails to meet the criteria for asystem, but this turns out not to be the case. Instead ofgiving any form of analysis of why Newlands fails tomeet his own criteria or those of Lachman and Kuhn,Giunta begins pursuing what he himself states as beingof secondary importance, namely the fact that Newlands’“contributions were substantial.” I will return to thisanalysis which takes up the next two pages of Giunta’sarticle, in due course; but first let us turn to the mainpurpose of the article, namely whether Newlands did ordid not produce a “system.”

Predictions

In returning to the promised main theme, Giunta beginsby stating that it is not so much that Mendeleev pro-duced a better system than Newlands, but rather thatNewlands failed to produce anything that might war-rant the label of a “system.” First of all Newlands’ ratherremarkable prediction of the existence of an unknownelement, which subsequently became known as germa-nium, is dismissed by Giunta. This is done on thegrounds that the prediction was made before Newlandshad formulated his law of octaves and that it was car-ried out with atomic weights instead of ordinal num-bers. In doing so Giunta seems to overlook the fact thatMendeleev’s spectacular predictions of germanium, gal-lium, and scandium were also based on atomic weights.

Contrary to Giunta’s reading, and regardless of whetheror not Newlands called his earlier classification a lawof octaves or not, it cannot be denied that he did in factpredict germanium a number of years before Mendeleev,as many historians concur.

In any case it is difficult to see why Giunta is plac-ing so much importance on predictions when he hadpromised earlier to concentrate on the criteria of “self-consistency” and “organization” in order to assess theworth of a periodic system. Of course, it may well bethat Newlands’ law of octaves is inconsistent with hisprior prediction of the element germanium, but this is aquite separate issue from whether the system itself isself-consistent. As I see it, self-consistency, in any formof system, such as a mathematical system, for example,does not necessarily imply predictive power.

Giunta then proceeds to criticize Newlands on thegrounds of failing to accommodate newly discoveredelements into his classification. Once again this is aseparate criterion that is covered neither by self-consis-tency nor organization of any system since the latter cri-teria are not necessarily connected to the possible dis-covery of new elements. The failure of Newlands’ clas-sification to allow for accommodation of new elementsis an important drawback but one which I believe ismischarachterized by Giunta’s analysis which suppos-edly hinges on “self-consistency and organization.”

Giunta then claims that Newlands’ “attempts ofsystematization” made in 1878 and 1884 came too late.What features make these attempts more systematic, inGiunta’s view, is not something that he discusses, ex-cept to say that Newlands was now “providing a check-list of specific instances in which he was applying thelaw.” I suggest that Giunta has once again shifted groundin that now an attempt to apply the law of octaves istaken to represent another criterion for deciding whetheror not Newlands’ classification represents a “system”(11).

Giunta then moves on to praise Mendeleev’s supe-riority over Newlands for making an (4):

...extensive list of deductions which accompanied hispredictions from the start.

This is unfortunately not quite the case. Admittedly,the three famous predictions of Mendeleev are hinted atin his original paper of 1869, by the fact that he leavesempty spaces for these elements. But it was not untiltwo years later that Mendeleev was prepared to makedetailed predictions on the properties of these elements

Page 49: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 127

and their compounds in his system of 1871 (12, 13).One might argue that two years is not a long time be-tween Mendeleev’s vague predictions of 1869 and hisdetailed version of 1871, but Giunta’s remark suggest-ing that Newlands is the only person whose views onclassification of the elements evolved over time seemsto be excessive.

There is no denying Giunta’s statement that someof Newlands’ published ideas showed a deterioration astime progressed. But there are some not too well knownaspects of Mendeleev’s work which, if examined in iso-lation, would also lead the reader to realize that the muchlauded Russian chemist also falls short of the adulationhe is usually accorded. One example is the case, sel-dom discussed in the literature, concerning the elementgallium. In his paper of 1871 Mendeleev predicted thateka-aluminum, subsequently known as gallium, would“in all respects” have properties intermediate betweenthose of the elements above and below it, namely alu-minum and indium. However, the melting point of gal-lium (30oC) is nowhere close to being intermediate be-tween those of aluminum (660oC) and indium (155oC).In 1879 Mendeleev gave the following ad hoc rational-ization of the anomalously low melting point for gal-lium (14):

...we should pay heed to the fact that the melting pointof gallium is so low that it melts at the temperatureof the hand. It might appear that this property is un-expected; but this is not so. It suffices to look at thefollowing series -Mg Al Si P S ClZn Ga ... As Se BrCd In Sn Sb Te IIt is evident that in the group Mg, Zn, Cd, the mostrefractory metal has the lowest atomic weight; but inthe groups beginning with S and Cl, the most diffi-cultly fusible simple bodies are, on the contrary, theheaviest. In a transitory group such as Al, Ga, In, wemust expect an intermediate phenomenon; the heavi-est (In) and the lightest (Al), should be less fusiblethan the middle one, which is as it is in reality. I turnattention to the fact that properties such as the melt-ing point of bodies depend chiefly upon molecularweight, and not on atomic weight. If we were to havea variety of solid sulphur not in the form of S

6 (or,

perhaps, of still heavier molecules Sn), but in the form

S2, which it assumes at 800oC, then its temperature

of melting and of boiling would undoubtedly be muchlower. In just the same way, ozone, O

3, condenses

and solidifies much more readily than does ordinaryoxygen, O

2.

Not only had such an argument never been give beforeby Mendeleev, as a means of predicting trends in prop-erties, but it also runs contrary to the spirit of his methodof simple interpolation which he used so successfully inmany other instances. The completely ad hoc nature ofthe argument is compounded by the fact that it is by nomeans clear that this truly represents “an intermediatephenomenon” to those in the other groups mentionedand indeed why this somewhat contrived trend shouldbegin at this particular place in the periodic table. Inspite of his use of the word “must” there is nothing inthe least bit compelling about Mendeleev’s argument.

Nobody would consider denying Mendeleev histriumphs because of such indiscretions. Furthermorethese ad hoc moves by Mendeleev would seem to bemore serious than Newlands’ desperate bid to assert hisclaim to priority in the case of germanium by referringto all his articles rather than, as Giunta would seem towish, just those published following the announcementof his law of octaves.

Back to Newlands

Contrary to the message in Giunta’s title, I believe thatNewlands did indeed produce a good periodic systemand more importantly perhaps, that he was the first toemphasize the importance of the periodic law, or thelaw of octaves as he termed it. Indeed, as Giunta pointsout and documents, the often heard dismissals ofNewlands on the grounds that he mistook the repeat dis-tance to be eight elements instead of nine, in the shortperiods, is something that Newlands himself fully an-ticipated.

Newlands’ contribution lies in having been the firstto recognize that the crucial feature lay in the approxi-mate repetition, or periodicity, of the elements and thatthis behavior is law-like. Whether this repetition oc-curs after seven, eight, or even nine elements is besidethe point. I believe that Giunta’s arguments for criticiz-ing Newlands’ system because of what he regards asinconsistencies have missed this important aspect. ButI agree with Giunta’s drawing attention to the fact thatsome of Newlands’ later systems did not leave any gapsfor undiscovered elements and thus negated his period-icity of eight.

This mention of leaving gaps raises the vexing ques-tion of just how important predictions are in science,something that Giunta does not discuss in spite of theextensive literature on the subject and the fact that this

Page 50: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

128 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

debate is still being actively pursued and precisely inthe context of the periodic system (15-20). Whether ornot prediction is an especially important aspect of sci-entific developments is open to question, as the currentdebate continues to show. On the other hand, the dis-covery of laws as a very important scientific activity isaccepted with less controversy. To return to a theme Ialluded to earlier, Newlands deserves perhaps morecredit than Giunta is giving him, precisely because hewas the first to recognize a law-like behavior in the waythat elements seem to recur after certain intervals.

Although Giunta recognizes that law-likeness isimportant, he seems to be prepared to ignore this aspectin the course of pronouncing judgment on Newlands’scientific contributions. Instead, as the title of the ar-ticle in question indicates, Newlands is being criticizedfor failing to discover a “system,” according to Giunta’srather idiosyncratic criteria for what constitutes a sys-tem.

I suggest that in attributing merit it is not the abil-ity to capture the small details that should be valuedmost, but rather to grasp the existence of a general law.If this is accepted then, contrary to Giunta’s position,Newlands should be lauded rather than faulted. Admit-tedly, Newlands was mistaken in not realizing that thisrepeat distance was variable. But in terms of announc-ing the existence of a law of regularity, which wouldhave very important ramifications, he was the first to doso.

Atomic Number

Finally, I turn to Giunta’s critique of Newlands over thequestion of atomic number since I believe that the argu-ments proposed are to some extent misplaced and ratherWhiggish. Giunta contradicts Wendell Taylor’s state-ment (21), that Newlands might have been “a pioneerin atomic numbers” because as Giunta puts it (4):

For several reasons that number is not the same asthe atomic number known today.

The first such reason for Giunta is that the discovery ofelements unknown to Newlands would cause some ofNewlands’ higher atomic numbers to be too low. Al-though this is indeed the case, I believe it to be a trivialobjection to the general principle of using an ordinalnumber to order the elements rather than their individualatomic weights. Clearly, Newlands could not haveknown the correct atomic numbers of all the elementsat the time at which he was writing.

Giunta’s second reason, the fact that Newlands as-signed the same ordinal number to some elements, is amore serious problem although it only occurs six timesin as many as 56 entries in Newlands’ table of 1866.

The final reason given by Giunta for rejecting thenotion that Newlands foresaw atomic number is alsodisputable (4):

Finally, Newlands was not aware of the physical ba-sis for atomic number first elucidated by Moseleymore than half a century later.

If the issue is whether Newlands in some sense antici-pated the notion of atomic number, then he could onlyhave done so in the absence of the knowledge of itsphysical basis. One cannot help wondering whetherGiunta might also want to diminish Mendeleev’s dis-covery of chemical periodicity itself because he was notaware of its “physical basis” until this was provided byNiels Bohr, in the form of electronic configurations ofatoms, also about half a century later.

Conclusion

Giunta is to be applauded for trying to bridge the unfor-tunate gap between the study of historical and philo-sophical aspects of chemistry. He has begun to analyzethe term “system” in the work of John Newlands, whiledrawing on the historical record. I hope the commentsraised here will stimulate a deeper analysis of the issuesinvolved.

ACKNOWLEDGMENT

I am grateful for the detailed and insightful commentsmade by a reviewer of this article.

REFERENCES AND NOTES

1. J. Van Brakel, “On the Neglect of Philosophy of Chem-istry,” Found. Chem., 1999, 1, 111-174.

2. N. Bhushan and S. Rosenfeld, Ed., Of Minds and Mol-ecules, Oxford University Press, New York, 2000.

3. E. R. Scerri and L. McIntyre, “The Case for the Philoso-phy of Chemistry,” Synthese, 1997, 111, 213-232.

4. C. J. Giunta, “J. A. R. Newlands’ Classification of theElements: Periodicity, But No System,” Bull. Hist.Chem., 1999, 24, 24-31.

5. M. Christie, “Chemists Versus Philosophers RegardingLaws of Nature, Stud. Hist. Philos. Sci., 1994, 25, 613-629.

Page 51: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 129

6. M. Christie and J. B. Christie, “Laws and Theories inChemistry Do Not Obey the Rules,” in Ref. 2, pp 34-50.

7. J. W. van Spronsen, The Periodic System of ChemicalElements, A History of the First Hundred Years, Elsevier,Amsterdam-London-New York, 1969.

8. If Giunta wishes to improve on the definitions given byvan Spronsen, then surely he owes it to the reader toprovide a better statement of what he takes van Spronsento mean by “true periodic system.” It would be rathersurprising, for example, if van Spronsen really means tosay that a system which separates main group from sub-group elements were necessarily superior. A few spe-cific page references to van Spronsen’s text would haveestablished on what definitions Giunta was attemptingto improve.

9. L. Gmelin, Handbuch der Chemie, C. Winter, Heidel-berg, 1843, 4th ed., Vol. I, 52, 456.

10. D. W. Theobald, “Some Considerations on the Philoso-phy of Chemistry,” Chem. Soc. Rev., 1976, 5, 203-213,

11. As I hope the reader will realize, I am not necessarilytrying to rehabilitate the work of Newlands nor even tosuggest that his work deserves more praise than it isgenerally given. I am mostly concerned with a critiqueof Giunta’s analysis and in pointing out what I see asinconsistencies and lack of clarity.

12. D. I. Mendeleev, “Sootnoshenie svoitsv atomnym vesomelementov,” Zh. Russ. Khim. Ova., 1869, 1, 60-77.

13. D. I. Mendeleev, “Estesvennaia sistema elementov iprimenenie ee kukazaniiv svoistv neotkytykhelementov,” Zh. Russ. Khim. Ova., 1871, 3, 25-66.

14. D. I. Mendeleev, “La loi périodique des élémentschimiques,” Le Moniteur Scientifique, 1879, 3, 691-737(692).

15. S. G., Brush, “Prediction and Theory Evaluation, TheCase of Light Bending,” Science, 1989, 246, 1124-1129.

16. S. G., Brush, “The Reception of Mendeleev’s PeriodicLaw in America and Britain,” Isis, 1996, 87, 595-628.

17. P. Lipton, “Prediction and Prejudice,” Int. Stud. Philos.Sci., 1990, 4, 51-60.

18. P. Maher, “Prediction, Accommodation and the Logicof Discovery,” in A. Fine and J. Leplin, Ed., Philos Sci.Assoc., 1988, Philosophy of Science Association, EastLansing, MI, Vol. 1, 1988.

19. C. Howson and A. Franklin, “Maher, Mendeleev andBayesianism,” Philos. Sci., 1991, 58, 574-585,

20. E. R. Scerri and J. Worrall, “Prediction and the PeriodicTable,” Stud. Hist. Philos. Sci., to appear, 2001.

21. W. H. Taylor, “J. A. R. Newlands: A Pioneer in AtomicNumbers,” J. Chem. Educ., 1949, 26, 491-496.

ABOUT THE AUTHOR

Eric Scerri, Department of Chemistry & Biochemistry,University of California at Los Angeles, Los Angeles,CA 90095; E-mail: [email protected]. Scerri’s re-search interests lie in the history and philosophy of chem-istry and in chemical education. He is the founder-edi-tor of the interdisciplinary journal Foundations of Chem-istry. http://www.wkap.nl/journals/foch

HISTORY OF CHEMISTRY DIVISION

http:www.scs.uiuc.edu/~maintzvHIST/

Page 52: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

130 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

Dr. Scerri raises several thoughtful points, both his-torical and philosophical, in his commentary (1) on mypaper (2) about J. A. R. Newlands’ classifications ofthe elements. I would like to respond to several ofthose points, to agree with some of them and to high-light some which we view differently.

I agree that my definition of periodic system wasambiguous and somewhat circular as regards “system.”Let me try again. What constitutes a periodic system?It must be periodic and it must be systematic. In myarticle, I believe I specified what I meant by periodic:arrangement by atomic weight and grouping of ele-ments with common properties; blank spaces for newelements and main group/sub group distinctions werenot necessary. I was much less definite on what it meantto be systematic, specifying only internal consistency.Clearly more was needed. What I had in mind but didnot explicitly define can be described as clarity of ex-position and of classification: a classification systemought to be clear about which elements constitute agroup of related elements.

Several other criticisms of my criteria for whatconstitutes a periodic system are, I believe, less wellfounded. Scerri states that I introduced an additionalrequirement, that a periodic system fulfill the criteriaof Sheldon Lachman about “what constitutes a theory.”Lachman’s criteria are for theory preference, not forwhat constitutes a theory (3). I did judge Newlands’work by Lachman’s criteria in addition to my (ambigu-ous) criteria for a periodic system; these were separateanalyses that addressed different questions. Scerri com-ments that I introduced even further criteria (due to

A RESPONSE TO SCERRI’S COMMENTARY

Carmen J. Giunta, Le Moyne College

Thomas Kuhn) without mentioning whether they areconsistent with either Lachman’s or my earlier criteriaof periodic system. On the contrary, my purpose of list-ing Kuhn’s criteria was not to introduce another set, butto show that Lachman’s list—which I find to be par-ticularly clear—is not anomalous in philosophy of sci-ence. Indeed, I quoted Kuhn: “Together with others ofmuch the same sort, they provide the shared basis fortheory choice (4).”

Dr. Scerri suggests that I implied that the periodicsystem is a theory. I must admit to having had no con-scious intent to imply such a thing; however, I recog-nize that my paper can be fairly read as suggesting justthat, or at least as blurring the lines between theories onthe one hand and classification systems on the other.Upon reflection, I do not wish to make such a sugges-tion, and I recognize that distinctions between theoriesand classification systems can be useful. Still, I thinkthat classification systems and empirical laws are ame-nable to analysis under criteria for theory preference suchas Lachman’s, Kuhn’s, or the like.

This may be a fruitful point for discussion. Must atheory be explanatory? predictive? or may it be simplydescriptive? Clearly, a report of raw observations orexperimental results is not a theory; such a report is de-scriptive, but not well organized. Empirical laws andempirical classification systems are also descriptive, butin a more organized way, correlating the observationsupon which they are based; however, they need not beexplanatory (i.e., state why a relationship holds). Anempirical law asserts a relationship between quantitiesgeneralized from individual instances and at least im-

Page 53: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 131

plicitly predicts relationships that can be measured inthe future between the same variables. A classificationscheme is also assertive, if not at least implicitly predic-tive, if it claims that it is a natural classification; itsgroupings assert that in one respect or another, A is likeB and unlike C. Empirical laws and classifications aresusceptible, at least in principle, of being formulated inalternative ways. In that they are assertive generaliza-tions susceptible of alternative formulations, empiricallaws and classification systems are similar enough totheories that Lachman’s or Kuhn’s criteria are appropri-ate for preferring one to another. This is not to denythat a distinction between theories on the one hand andclassifications or empirical laws on the other may wellbe useful.

One main point of disagreement between Dr.Scerri’s analysis of Newlands’ work and my own is overmy treating all of Newlands’ work as a whole, which Ifound incoherent. He seems to read Newlands’ classifi-cations as ideas evolving over time. This interpretationis not unreasonable; however, I believe that the histori-cal record supports my interpretation (5). One can con-ceptually separate triad-based classifications (which in-clude predictions of undiscovered elements, includingone of the element now called germanium) from ordernumber classifications (including the “Law of Octaves”and later classifications). If one were to consider thesetwo phases separately, then the triad phase is certainlynot a periodic system because it did not embrace all el-ements. The law of octaves is a periodic system by mydefinition, although it is not as clear as I would like aboutwhat elements constitute a group of related elements.(The law of octaves falls short of the more elaborateversions of Mendeleev’s classifications on several ofLachman’s criteria, but those criteria are for preference,not for periodic system.) In short, if I regarded order-number classifications as displacing triad-based classi-fications, then I would admit that the latter constitute aperiodic system. I do not think the historical record sup-ports such an interpretation, though. Newlands’ mono-graph, written years after the classifications, seems toembrace both of these phases simultaneously. In addi-tion, the first of the order-number papers (August 8,1864) was written soon after the last of the triad-basedpapers (July 12, 1864) without any explicit break.

Beyond these differences in interpreting the histori-cal record, I believe Dr. Scerri’s analysis misses the markin several particulars. For example, he asserts that Ifailed to show how Newlands fell short of my criteria

for periodic system or the criteria of Lachman and Kuhn,and that instead I followed a summary judgment on thismatter by pursuing secondary issues, namely Newlands’substantial contributions. In fact, the section to whichScerri refers was titled “The Case for Newlands,” and ittreated those aspects of Newlands’ work that satisfy partsof my definition of a periodic system. The next section(“Why Newlands’ Insights Do Not Constitute a Peri-odic System”) asserted that Newlands’ work is not aperiodic system because it is not systematic. While Iadmit the inadequacy of my definition of periodic sys-tem, this section certainly addressed the definition, criti-cizing the internal inconsistency of Newlands’ writings.Finally, the last section of the paper (“Assessment Us-ing Lachman’s Criteria”) rated Newlands’ work on allsix of those criteria, one by one. Scerri seems to thinkthat I unjustly dismissed Newlands’ prediction of ger-manium, made on the basis of atomic weight relation-ships, and overlooked the fact that Mendeleev’s predic-tions were also based on atomic weights. Far from dis-missing the prediction of germanium, I emphasized itbecause predictions of new elements were logically in-consistent with the later periodic classification (the lawof octaves) which left no room for new elements. Fur-thermore, the basis of Newlands’ (and Mendeleev’s)predictions was not wrong but it was not the basis ofNewlands’ periodic classification. Scerri contradicts myassertion that Mendeleev’s system included an “exten-sive list of deductions ... from the start,” characterizingpredictions in his original 1869 paper as merely “hintedat” by leaving empty spaces in his table. In fact, the1869 paper contained several explicit deductions, includ-ing explicit predictions of two new elements (6). Fi-nally, Scerri finds my “critique of Newlands over thequestion of atomic number” to be “to some extent mis-placed and rather Whiggish.” I do not consider my wordson atomic number to be a critique: I credited Newlandsfor the ordinal number concept and quoted without con-tradiction Taylor’s assessment of Newlands as a “pio-neer in atomic numbers (7).” I went on to describe thedifferences between the modern concept of atomic num-ber and Newlands’ ordinal number, with no criticismstated, intended, or implied (but apparently criticism wasinferred). Of course this section is Whiggish in that itdescribes current understanding of a past proposal.

Dr. Scerri offers some observations that I believewould have improved my paper if I had incorporatedthem; he raises some issues that thoughtful scholars canfruitfully debate and over which they may disagree, buthe makes some points that I believe are not well founded.

Page 54: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

132 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

REFERENCES AND NOTES

1. E. R. Scerri, “A Philosophical Commentary On Giunta’sCritique Of Newlands’ Classification Of The Elements,”preceding paper.

2. C. J. Giunta, “J.A.R. Newlands’ Classification of theElements: Periodicity, But No System,” Bull. Hist.Chem., 1999, 24, 24-31.

3. S. J. Lachman, The Foundations of Science, GeorgeWahr, Ann Arbor, MI, 1992, 3rd ed., particularly pp 50-51.

4. T. S. Kuhn, “Objectivity, Value Judgment, and TheoryChoice,” in The Essential Tension, University of Chi-cago Press, Chicago, IL, 1977, 320-339.

5. Newlands’ writings on classifications of the elementsare collected in J. A. R. Newlands, On the Discovery ofthe Periodic Law and on Relations Among the AtomicWeights, E. & F. N. Spon, London, 1884. Some of the

relevant documents are available on the Internet at http://webserver.lemoyne.edu/faculty/giunta/newlands.html .

6. D. Mendelejeff, “Ueber die Beziehungen derEigenschaften zu den Atomgewichten der Elemente,” Z.Chem., 1869, 12, 405-406; abstracted and translated intoGerman from Zh. Russ. Khim. Ova., 1869, 1, 60-77.

7. W. H. Taylor, “J. A. R. Newlands: A Pioneer in AtomicNumbers,” J. Chem. Educ., 1949, 26, 491-496.

ABOUT THE AUTHOR

Carmen Giunta is Associate Professor of Chemistry atLe Moyne College, Syracuse, NY 13214. A physicalchemist by training, he is particularly interested in ap-plying history of chemistry to chemical education. Hemaintains the Classic Chemistry web site: http://webserver.lemoyne.edu/faculty/giunta

Paul Bunge Prize

The Paul Bunge Prize 2001 has been awarded to Dr. Jim Bennett, Museum of the His-tory of Science, Oxford, in recognition of his complete historical works on scientific instru-ments. The prize of DM 15,000 will be presented on September 25, 2001 in Würzburg, onthe occasion of the annual meeting of the Gesellschaft Deutscher Chemiker. Deadline forapplications for the Paul Bunge Prize 2002 (7,500 Euro) is September 30, 2001: ContactGerman Chemical Society, Public Relations Department, PO Box 900440, D-60444 Frank-furt am Main: [email protected].

Page 55: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 133

CALL FOR NOMINATIONSFOR THE EDELSTEIN AWARD

The Division of the History of Chemistry (HIST) of the American Chemical Society(ACS) solicits nominations for the 2002 Sidney M. Edelstein Award for OutstandingAchievement in the History of Chemistry. This recently established award honors thememory of the late Sidney M. Edelstein, who established the Dexter Award in 1956, andit also continues the outstanding tradition of the Dexter Award, which came to an endwith its posthumous presentation to Dr. William A. Smeaton in 2001.

The Edelstein Award is sponsored by Ruth Edelstein Barish and Family and is adminis-tered by HIST. In recognition of receiving the Edelstein Award, the winner is presentedwith an engraved plaque and the sum of $3500, usually at a symposium honoring thewinner at the Fall National Meeting of the ACS, which for 2002 will be held in Boston,Massachusetts, August 18-22. The award is international in scope, and nominations arewelcome from anywhere in the world. Previous winners of the Dexter Award have in-cluded chemists and historians from the U.S., Canada, Germany, France, Holland, Hun-gary, and Great Britain.

Nominations should consist of a complete curriculum vitae for the nominee, includingbiographical data, educational background, awards, honors, publications, presentations,and other services to the profession; a nominating letter summarizing the nominee’sachievements in the field of the history of chemistry and citing unique contributions thatmerit a major award; and at least two seconding letters. Copies of no more than threepublications may also be included if they are available. All nominations should be sent intriplicate to Prof. Seymour Mauskopf, Chair of the Edelstein Award Committee, Depart-ment of History, Duke University, Durham, NC 27707 (e-mail: [email protected]),by January 31, 2002.

Page 56: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

134 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

BOOK REVIEWS

Instruments and Experimentation in the History of Chem-istry, F. L. Holmes and T. H. Levere, Ed., MIT Press,Cambridge, MA, 2000. xvii. + 415 pp, Cloth, ISBN 0-262-08282-9. $50.00.

This volume is part of the new “Dibner InstituteStudies in the History of Science and Technology” se-ries and contains 14 essays dealing with the evolutionof chemical apparatus and laboratory techniques. PartI, entitled “The Practice of Alchemy,” contains threeessays covering the evolution of early distillation appa-ratus (Robert Anderson), the relation between alchemyand assaying (William Newman), and the problems ofreplicating alchemical apparatus and experiments(Lawrence Principe).

Part II, entitled “From Hales to the Chemical Revo-lution,” contains six essays covering the evolution ofapparatus for the generation and isolation of gases(Maurice Crosland), the development of the eudiometer(Trevore Levere), the evolution of Lavoisier’s chemicalapparatus (Frederic L. Holmes), the development ofhydrometers (Bernadette Bensaude-Vincent), the devel-opment of 18th-century thermometers (Jan Golinski),and 18th-century uses of platinum and ground glass inapparatus design (William Smeaton).

Part III, entitled “The Nineteenth and Early Twen-tieth Centuries,” contains five essays covering a reas-sessment of the experimental work of Wollaston andThomson on multiple proportions (Melvyn Usselman),the development of organic combustion analysis in theperiod 1811-1837 (Alan Rocke), the experimental studyof gun powder (Seymour Mauskopf), apparatus innova-tion in the work of Edward Frankland (Colin Russell),

and apparatus usage in the career of Michael Polanyi(Mary Jo Nye).

One could hardly ask for a more sterling cast ofcontributors, nine of whom are former Dexter AwardWinners. Consequently, it comes as no surprise that allof the contributions are well done and of great interestnor that many of the essays are amplifications of sub-jects in which the authors already have well establishedreputations. Regrettably this also means that there islittle or no substantive coverage of events after about1840, as the two essays by Russell and Nye, which post-date this period, are really biographical vignettes ratherthan focused studies of significant developments in ap-paratus and experimental innovation. Given that theperiod 1840-2000 contains some of the most spectacu-lar advances in chemical instrumentation and labora-tory technique, this omission is unfortunate and appearsto be tied to another quirk of this volume (shared alsoby its sister volumes in the same series)—namely thetotal absence of any chemical historians (i.e., chemists)among the lists of contributors. Though the Division ofthe History of Chemistry has made many publicationand symposia opportunities in the history of chemistryavailable to professional historians of science, this gen-erosity in the sharing of resources appears to be a one-way street.

However amateurish some of their attempts to dochemical history may appear by modern standards,chemists have always excelled at documenting the his-tory of their apparatus and procedures. One needs onlymention the pioneering work of John Stock, L. S. Ettre’sstudies on the development of chromatography, or theheroic efforts of John Ferraro and his collaborators todocument the post-Second World War instrumentationrevolution.

Page 57: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 135

With the possible exception of Hans Jenemann’srecent (1997) monograph on development of the chemi-cal balance, the individual authors appear to have donea good job in citing this literature in their individualcontributions, where appropriate. The same, however,cannot be said of the book’s introduction. Surely one ofthe responsibilities of an editor is to establish continu-ity between the book in question and the older literaturein the field, even if this older literature no longer re-flects current historiographic standards. Yet one scansthe main introduction and the shorter section introduc-tions in vain for any explicit mention of the work ofJohn Stock, of Ernest Child’s 1940 study, The Tools ofthe Chemist, or of the even earlier volume, A Pictorial

History of Chemistry, by F. Ferschl and A. Süssenguth(1939). Despite its title, this latter work is essentially apictorial history of chemical apparatus from alchemythrough about 1850 based on the extensive displays de-veloped by Süssenguth at the Deutsches Museum inMunich before the Second World War.

But despite these minor defects (and the reviewerreadily admits to being inordinately curmudgeonly forharping on them in the first place), this volume as a wholerepresents a valuable and worthwhile commentary onthe development of chemical, albeit “early” chemical,apparatus. William B. Jensen, Department of Chemis-try, University of Cincinnati, Cincinnati, OH, 45221-0172.

Instrument-Experiment: Historische Studien, C. Meinel,Ed., Verlag für Geschichte der Naturwissenschaften undder Technik, Berlin-Diepholz, 2000. 423 pp, Cloth,ISBN 3-928186-51-5. 34 Euro.

This collection of 37 essays dealing with the de-velopment and impact of scientific instrumentation wascommissioned by the German Society for the Historyof Medicine, Natural Science and Technology (DeutscheGesellschaft für Geschichte der Medizin,Naturwissenschaft und Technik). The editor, ChristophMeinel of the University of Regensburg, has dividedthe contributions into six groups, entitled “Historio-graphic and Methodological Perspectives” (four essays),“Instruments and the Manufacturing of Reality” (eightessays), “Establishing Instrumental Procedures” (sixessays), “Microscopic Views and Scientific Knowledge”(twelve essays), “Instrumentation and Social Practice”(eight essays), and “Towards the Materiality of Instru-ments” (seven essays).

All but three of the essays are in German and onlytwo deal explicitly with topics relevant to the history ofchemistry (Nikos Psarros, “Was sah Ostwald (als er dieBrille von Frantisek Wald ablegte)?” and AnthonyTravis, “Surrogate Instruments: Industrial ChemicalReactors and Organic Chemistry”). The vast majorityof the remaining essays deal with topics in the historyof physics, medicine, or physiology.

As can be seen from the translations of the sectionheaders, the organization of the book is heavily influ-enced by current fads in the sociology and philosophyof science, and practicing scientists are likely to findthemselves puzzled by some of the bizarre terminologyand superficial metaphors employed in some of the more“theoretical” contributions. Nevertheless, several of theauthors have managed to present straightforward fac-tual accounts, which, despite their brevity, may proveuseful to those interested in the history of instrumenta-tion and its role in the construction and verification ofscientific theories. William B. Jensen, Department ofChemistry, University of Cincinnati, Cincinnati, OH,45221-0172

Page 58: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

136 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

Gold aus dem Meer; Die Forschungen desNobelpreisträgers Fritz Haber in den Jaharen 1922-1927, Ralf Hahn, GNT-Verlag, Diepholz, 1999. ISBN3-928186-46-9, paper, 101 pp. DM 24.60.

For those who may have read about Haber’s questfor gold from the ocean in his biography (D.Stoltzenberg, Fritz Haber: Chemiker, Nobelpreisträger,Deutscher, Jude, Wiley-VCH Verlgla GmbH,Weinheim, 1994; see review, Bull. Hist Chem., 1999,24, 77-78.), this booklet by Hahn will provide 100 pagesof detail about the ambitious, unsuccessful venture.This gold-covered publication, the result of a master’sdegree study (Magisterarbeit), contains an introduc-tion by Lutz Haber (1920-), youngest son of Fritz Haberand a science historian living in Bath, England.

A highly abbreviated but richly annotated biogra-phy of Fritz Haber serves as a preliminary chapter. Inthe following ten-page chapter, Hahn presents the back-ground of the status of research on extraction of goldfrom seawater at the time Haber considered undertak-ing this project in the 1920s. The first proposal by J. LProust that the sea might contain significant amountsof gold appeared surprisingly early, in 1787. A. Wurtzsuggested it in a lecture in 1866, and the first recordedexperiments on the subject are attributed to the Britishscientist E. Sonstadt, in 1872. The author describes insome detail the various methods devised for the sepa-ration and analysis of noble metals, from 1872 up to1918, as reported by British, Norwegian, French, Swiss,and New Zealand experimentalists. This includes acomparison of the sensitivity of analyses and practi-cality of the methods. When Haber initiated his pro-gram, he elected to use the technique described in 1918by H. Koch, consisting of adsorption of aqueous solu-tions on wood charcoal, followed by ignition and cu-pellation, to afford gold in the form of pellets.

The major portion of this booklet is taken up withHaber’s research project on gold, carried out at theKaiser-Wilhelm Institute in Berlin, from 1922-1927,in “Abteilung M.” There is a suggestion—althoughundocumented—that Haber may have been given theidea to pursue extraction of gold from seawater byArrhenius in Stockholm in June, 1920 when Haber re-ceived the Nobel Prize. The motivation for such anambitious undertaking was the idea that recovery ofgold would serve as a source of repayment of Germany’shuge war debt, amounting to over 200 billion

Goldmarks. Even reduced to 132 billion by 1921, thissum would have been the equivalent of 50,000 tons ofgold. The research got underway by 1922; carried outby Haber’s institute associates, it led to six doctoral dis-sertations in the next six years. One of those doctoralstudents, Johannes Jaenicke, published a short descrip-tion of the project in Die Naturwissenschaften in 1935but also provided bountiful documents, which are housedin the archives at the Max-Planck-Gesellschaft, Berlin-Dahlem. Like the biographer Stoltzenberg, Hahn de-pended heavily on these documents for his research.Preliminary experiments were directed toward effectiveworkup and analysis of gold-containing samples, mostlysynthetic. The author provides substantial detail on vari-ous separation and analytical designs, which are not in-cluded in this review. Results from the sparse numberof real seawater samples were disappointingly lower thanthose reported by earlier researchers. Perhaps Habersensed the project would entail more elaborate supportto be successful, for he turned to external sources forfinancial backing. Ever skilled in such collaborations,he eventually arranged the establishment of a consor-tium with Degussa and Metallgesellschaft. Haber wasto provide the scientific expertise, while Degussa andMetallgesellschaft would finance the project up to$100,000. Moreover, in all decision making, Haberwould hold two votes to one each for the other two par-ticipants. Any financial gain to be realized would bedistributed, 50% to Haber, and the other half equallydivided between Degussa and Metallgesellschaft. Au-thor Hahn notes that archival material fromMetallgesellschaft served as a valuable resource for hishistorical research and that his request to use recordsfrom Hapag, Hamburg, were denied. No mention ismade of any materials from Degussa.

By the summer of 1923 plans were in place for anexpedition in the Atlantic Ocean. On board the Hansa,a ship of the Hamburg-America line, two cabins wereoutfitted as laboratories and sleeping quarters for Haberand three coworkers, who took samples and analyzedthem during the round trip from Hamburg to New Yorkin July and August. A second expedition in Septemberafforded fewer results than the first, which had been er-ratic. In October a third expedition, on board theWürttemberg, also a ship of the Hamburg-America line,headed for Buenos Aires, this trip having been financedby the German Navy and “Notgemeinschaft.” As withthe first Hansa expedition, Haber was on hand, alongwith three doctoral students who took samples and per-formed analyses in a newly designed and improved labo-

Page 59: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 137

ratory. The values for gold content were even lowerthan those from the two earlier excursions; this raisedquestions about the efficacy of the ever changing meth-ods and the possible uneven distribution of gold in theoceans. Still fairly optimistic, Haber gave a lecture inBuenos Aires before the German Club; and in the springof 1924 in Dahlem he spoke at the second conferenceon “Seejod.” There followed yet two more sea voy-ages, one on board the Poisedon in the North Sea inMay, 1924, the other in April, 1925 on the Meteor, whichmade 14 round trips between Africa and South America,while the scientists on board took copious samples. Mostwere analyzed back at the Kaiser-Wilhelm Institute.Haber arranged for seawater samples to be sent to Ber-lin for analysis from Iceland, Greenland, the San Fran-cisco Bay, and even from the Rhein River near Karlsruhe.From the Meteor expedition 1,635 samples were col-lected and 85% of them eventually analyzed back inBerlin. The average gold content was 4 x 10-3 mg/tonof seawater; this was deemed to be 1/1,000 the amount

for a viable extraction process. At this point Haber pro-nounced the project hopeless. In a lecture in May, 1926,entitled “Gold in the Sea,” Haber asserted he had givenup searching for the needle in the haystack!

In a brief summary, the author brings us up to thepresent day. During the quest for gold between 1920and 1926, Haber experienced no competition; yet otherreports of gold detection came out between 1927 andthe early 1940s, all values being low and in the samerange as those of Haber. With the development of newanalytical techniques, such as neutron activation, thefield is still being studied, as indicated by various patentapplications. Yet the average concentration reported inthe 1990s—10-2 mg/ton is not for from the final aver-age figure from Haber’s work.

Hahn concludes his richly documented and illus-trated booklet with Roald Hoffmann’s poem, “FritzHaber (1993).” Paul R. Jones, University of Michigan.

Science and Engineering in Ireland in 1798: A Time ofRevolution, P. N. W. Jackson, Ed., Royal Irish Acad-emy, Dublin, 2000. viii + 81 pp, paper, ISBN 1-874045-77-1. IR £5.00.

This collection of five essays is the outcome of theproceedings of a symposium organized by the National(Irish) Committee for the History and Philosophy ofScience, held in November, 1998, as a bicentennial rec-ognition of the 1798 rebellion of the United Irishmen.Authors are affiliated with Queen’s University, Belfast;Trinity College, Dulbin, or University College, Dublin.Three of the chapters cover the status of some of thesciences at the time of the rebellion: biology; geology;and science, engineering and the military. Of particularinterest to chemists will be the remaining two chapterson two chemists, both of whom ended up in careers inthe United States: William James MacNeven (1763-1841) and John Patten Emmet (1796-1842). Neitherchemist is indexed in Ihde’s Development of ModernChemistry; only MacNeven appears in Partington’s his-tory, although both are included in Miles and Gould,American Chemists and Chemical Engineers, 2nd ed.(1994). While MacNeven’s professional career at Co-

lumbia College of Physicians and Surgeons, both inmedicine and later in chemistry, is well covered by Miles,the Irish publication includes MacNeven’s involvementin the abortive rebellion and other background beforehe emigrated to the US. Emmet migrated with the restof his family to the US as a child; it was his father whoparticipated in the rebellion and was imprisoned for atime. In New York he was educated by a Trinity Col-lege graduate, Richard W. Thompson, and then studiedmedicine under MacNeven. Emmet became adept inlaboratory skills and was also a talented artist. Eventu-ally he became Professor of the School of Natural His-tory at the University of Virginia, publishing severalarticles in analytical chemistry and electrochemistry,among other topics. A complete list of his publicationsis included in this essay, along with some of his originalsketches. Both articles on MacNeven and Emmet areenhanced with figures from the Irish Rebellion, photo-graphs of the two chemists and Emmet’s wife, of titlepages from MacNeven’s popular texts, of the originalRutgers Medical College (co-founded by MacNeven),of the monument to MacNeven in New York, and a re-production of Thomas Jefferson’s hand written invita-tion to Emmet for the faculty position in Charlotte. PaulR. Jones, University of Michigan.

Page 60: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

138 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

The Biographical Dictionary of Women in Science: Pio-neering Lives from Ancient Times to the Mid-20th Cen-tury, Marilyn Ogilvie and Joy Harvey, Ed. 2 vol.,Routledge, New York and London, 2000. xxxviii + 1499pp, $250.

This extensive compilation is an expansion andupdating of Marilyn Ogilvie’s earlier biographical dic-tionary, Women in Science, Antiquity Through the Nine-teenth Century (1986). Embracing a broad range ofwomen taking part in scientific or science-orientedwork, from public health activists and alchemists towriters and zoologists, it covers a span of 2500 yearsand includes people from all five continents. In addi-tion to the narrative texts that form the body of the work,there are three sets of lists of women by occupation,time period, and country or geographical region. A verygood bibliography of about 200 standard sources cov-ering material up to 1998 is also included. The vol-umes are handsomely bound; type is clear on glossypaper; names are easy to spot on the page; and there isa good general index.

Entries follow a uniform format but vary in lengthdepending on the prominence of the subject and theamount of information found by the writers. Thus, onlooking at chemists in particular, several pages are de-voted to Marie Curie; early Massachusetts Institute ofTechnology instructor Ellen Swallow Richards (alsoknown for her pioneering work in home economics) isdiscussed in an article of medium length. Lesser play-ers in the drama and many for whom information isless readily accessible are afforded coverage in only aparagraph or two—such as the tantalizingly briefglimpse of Leonora Bilger (born 1893), a successfulearly chemist, Garvan medalist, and department headat the University of Hawaii from 1943 to 1954. A shortbut interesting sketch is offered of the career of MaryPeters Fieser, wife and research colleague of LouisFieser and the co-author of the famous Fieser and Fiesertexts so well-known to many of us. Greek-born bacte-riologist and social activist Amalia Coutsouris Fleming,second wife of Sir Alexander Fleming of penicillin fameand his co-worker in investigations on streptomycin, isalso included. Overall, the work shows signs of hastewith many typographical errors and other more impor-tant oversights which have the unfortunate effect ofshaking one’s confidence in the work as a whole. Thereader with scientific background especially may wellbe somewhat disappointed. To this reviewer the stan-

dard does not match that of Marilyn Ogilvie’s 1986 Dic-tionary, either in the overall quality of the articles or thefactual accuracy. Unfortunately there are occasionaldifficulties with basic information. For instance latenineteenth-century Italian bacteriologist GiuseppinaCattani’s last name is spelled consistently as Catani, andearly twentieth-century British chemist Alice EmilySmith is designated as having the working life 1850-1905. In Smith’s case the confusion arose from the factthat a scholarship she held as a student commemoratedthe Great Exhibition which took place in London in1851; she herself published her first paper jointly withW. H. Perkin Jr. in 1902, and she remained very activein research and publication until at least 1909.

A number of entries that have been carried overessentially unchanged from the earlier edition remainvery adequate, such as those on Marie Curie and EllenSwallow Richards. However, in some cases updatingwould have been appropriate, as in the discussion of thework of French mathematician Sophie Germain. It isno longer the case (as it was in 1986) that Fermat’s LastTheorem is unproven, the task having been accomplishedby Andrew Wiles in 1995; and although Germain’s workremained an important contribution referred to by oth-ers for many decades, Wiles’s proof does not depend onit but rather on tools developed long after her day. Newentries written by scientists and contributors who havemade extensive studies of the person under discussionare good, such as that on early twentieth-century Ger-man physicist Hertha Sponer-Franck, professor at DukeUniversity for almost thirty years and remembered forher quantum mechanical studies, including investiga-tions of structural properties of complex molecules.

One might wish that the assistance and guidance ofspecialists in the fields had been used more extensivelyso as to reduce difficulties with discussions of technicalwork and provide somewhat fuller coverage. A numberof articles on women chemists published during the lastfifteen years in the Bulletin for the History of Chemistryhave not been consulted, and neither has Volume 2 ofAmerican Chemists and Chemical Engineers (W. Milesand R. Gould, 1994). These sources could have pro-vided needed additional information. However, the edi-tors will undoubtedly correct in later editions such ratherserious errors as those which detract from the presenta-tion of the work of American bacteriologist/biochemistRebecca Lancefield (“Lancefield found evidence thatcountered the accepted belief that type-specific viru-lences [sic] were carbohydrates of [sic] polysaccha-rides,” p 739). It is perhaps something of an exaggera-

Page 61: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 139

tion to claim Mt. Holyoke botanist and general sciencesinstructor Lydia Shattuck as “one of the founders of theAmerican Chemical Society (p 1182).”

As noted in the editors’ introduction, the more readyaccessibility of source materials biases the coverage infavor of the United States. Nevertheless, women ofFrance, Germany, and especially Britain are also fairlywell represented. Some entries for which the availablesource material is plentiful but in non-English languagepublications are disappointingly brief—which is perhapsunderstandable in a work of this size since the labor oftranslating can be time consuming. The articles on Rus-sian women, while a welcome addition to the somewhatmeager coverage of this national group in English lan-guage sources, hardly reflect the tremendous increasein the participation of women in scientific work duringthe Soviet era.

A few words concerning the interpretation of thelists of women by occupation, by time period, and bynationality are perhaps in order since these lists are im-portant in fulfilling one of the editors’ stated hopes:namely, that this dictionary will provide a unique op-portunity to view subjects “longitudinally within fieldsover a long period of time or horizontally across fieldswithin a restricted time period (Introduction, xi).” Thelists run into the problems typical of such categoriza-tion attempts. While their lengths offer approximateindications of the growth of women’s participation inscientific work over time, the extent to which womenpenetrated into particular fields, and the countries inwhich they have been most successful, the overlaps be-tween divisions within each list mean that simple enu-meration might well give one a somewhat exaggeratedidea of the extent of women’s activity in science. Thisalso holds for a comparison of fields within individualcountries. (Here the overall bias towards the UnitedStates is not a complicating factor.) Thus chemist Char-lotte Roberts of Wellesley College appears in both nine-teenth- and twentieth-century lists; mid-westerner LauraLinton, who during her years in chemistry carried out amineral analysis in addition to her rather notable, butnot mentioned work in petroleum chemistry, appearsunder both chemists and mineralogists (as well as underphysicians and educators, reflecting other periods of hercareer). There are also some unexpected and unex-plained editorial choices. For instance, why present twolists of Dutch women separated under the headings “Hol-land” and “Netherlands”? And why place nineteenth-century St. Petersburg chemist Anna Volkova, a protégéeof Mendeleev, in the listing headed “USSR” rather than

in that headed “Russia”? Even with the lists and theimpressively large number of subjects included, it issomewhat difficult to see this dictionary as meeting theeditors’ hope of offering a “unique perspective” for in-ternational and cross disciplinary comparisons; its ba-sic style and methodology keep it too firmly anchoredin the standard pattern for dictionaries of women in sci-ence.

Critical judgment of the work of some of the moreeccentric women appearing in this very inclusive col-lection is rarely offered. Mary Boole, widow of the fa-mous nineteenth-century British mathematician GeorgeBoole, developed and published pioneering ideas on theteaching of mathematical concepts to young children,ideas of continuing interest to educational psychologists;however, it might have been advisable to acknowledgethat at the same time, caught up in current psychic re-search and related fads, Mary Boole brought out a con-siderable amount of what is best described as nonsense.Likewise, while the entry on nineteenth-/twentieth-cen-tury Paris-based feminist and writer Céline Renooz of-fers a fairly realistic assessment of Renooz’s wildly ex-travagant conjectures about current science and the de-velopment of human cultures, the discussion of Renooz’scontemporary Clémence Royer is somewhat less bal-anced. Science writer and commentator Royer, authorof the first French translation of Darwin’s Origin of Spe-cies, has her enthusiastic supporters among historians,but others advocate a more cautious approach. Somemention might have been made of the fact that Royerwas a convinced Lamarckian who failed to grasp theessential difference between Lamarck’s ideas of evolu-tion and Darwin’s theory of natural selection—a failurewhich distorted her translation. As well as making thedictionary articles more bland and less interesting thanthey otherwise might have been, this very delicate ap-proach will make it difficult for the uninitiated reader toform a realistic idea of the importance of the personunder discussion.

Readers already familiar with the field of womenin science will find here new names; specialists will alsoquickly realize they need to take the precaution of check-ing the details provided. On the other hand, those whohave little background in the subject, particularly be-ginning students, should approach these volumes withcaution. Although the basic information of nationality,time period, and area and general extent of scientificactivity of this large collection of women of science pro-vides a useful point of entry, careful scrutiny of addi-tional information is recommended.

Page 62: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

140 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

Despite the caveats, the fact remains that this dic-tionary is a noble effort, the product of a very consider-able amount of labor on the part of the two editors, whohave themselves written a large fraction of the articles.They are to be commended for making available in one

Auf der Suche nach dem Stein der Weisen: DieGeschichte der Alchemie. Hans-Werner Schütt, VerlagC. H. Beck, Munich, 2000. 602 pp, DM 68.50.

The history of alchemy has come into its own inrecent years, the Renaissance and early modern peri-ods having been particularly well covered by sucheminent scholars as Allen Debus, Betty Jo Dobbs, KarinFigala, Owen Hannaway, William Newman, LawrencePrincipe, and Pamela Smith. What has been lacking isan up-to-date recounting of the entire span of alchemi-cal history, from antiquity to the modern period. Themost recent such surveys are now nearly a half centuryold. (Readers should be aware, however, of the admi-rable Alchemie: Lexikon einer hermetischenWissenschaft, edited by Claus Priesner and KarinFigala, and published by Beck Verlag in 1998.)

In the book under review, the outstanding histo-rian of chemistry Hans-Werner Schütt provides us withsuch a history. Here, in 600 pages and close to a hun-dred short chapters, is an examination of the full rangeof alchemical lore; beginning “in the shadows of thepyramids,” Schütt progresses through “foreign worlds”(the Arabic period), then “into monasteries and else-where” (the middle ages), and finally “into the newworld of Europe.” In an afterword, Schütt notes that itwas not his intent to provide a scholarly investigationof the subject that seeks novel understanding. Rather,his aim was to get under the skin and into the minds ofthe alchemists of various times and places. This in-cluded, for Schütt, mining matters “anecdotal, philo-sophical, psychological, and political,” in the hope ofpresenting, as Golo Mann put it, “what is uncommonlyentertaining in history.” In this way, he concluded, he

could steer safely between the Scylla of “professorialincomprehensibility” and the Charybdis of “cheap popu-lar showmanship.”

Schütt’s navigation was up to the task. The book iswell organized, lively, and chock filled with interestingmatter. The writing is effective and often suffused witha deliciously wry sense of humor. In some respectsSchütt was overly modest in his protestations, for thereis much that is novel and sometimes even profound here.One example is a wonderful passage on what might becalled the teleology of everyday life, an effective pieceof rhetoric that allows entrée into the teleological psy-chology of alchemy (pp 63-64).

In another fascinating passage (pp 495-497), Schüttrelates his experiments attempting to reproduce someof the observations of the alchemists. Similar toLawrence Principe’s work of a few years ago, Schüttsucceeded in showing that various puzzling reports inthe alchemical literature really do make sense as a re-sult of laboratory operations. His report, interlaced withprecise details of time, place, and witnesses (or lackthereof), is strikingly reminiscent of some of the alchemi-cal narratives themselves.

Schütt wanted to appeal to the broadest possibleaudience and so kept his scholarly apparatus to the mini-mum: about 300 footnotes in all (most of them textualrather than source citations), and four pages of bibliog-raphy. This poses a disadvantage for scholars of thehistory of alchemy, which is ameliorated by Schütt’smaintenance of a website associated with his book (http://www.tu-berlin.de/fb1/alchemie) that provides a muchfuller bibliography.

location a great many names of women scientists; thework will unquestionably remain an essential, standardreference in American libraries for many years. MaryR. S. Creese, Hall Center for the Humanities, Univer-sity of Kansas, Lawrence, KS 66045.

Page 63: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 141

Still, it must be clearly stated that Schütt has per-formed an enormous service for the field. His historyof alchemy will immediately become the standard gen-eral treatment and will go far to raise interest in the field

Arnold O. Beckman: One Hundred Years of Excel-lence, Arnold Thackray and Minor Myers, Jr. Chemi-cal Heritage Foundation, Philadelphia, PA 2000, 379pp. $65.00.

What a grand celebration! A celebration of 100years of inexhaustible contributions to the advance-ment of science, technology, education and the qualityof life. A true attainment of the American Dream!(And this reviewer knows something about centenni-als and celebrations, currently experiencing the cen-tennial year of the National Institute of Standards andTechnology.)

Chronologically organized, this book is more thanjust the life and times of Arnold Beckman; it is achronicle of chemistry in the 20th century. No, it ismore than just about chemistry, it covers all of scienceand engineering. Whom are we kidding? This book isabout life in the United States of America!

Beginning with his early years in Illinois, we seean engaging youth filled with curiosity (these days of-ten this same spirit might be viewed as non-PC, if notillegal!). Forced to learn piano by a loving mother, heproved these talents to be invaluable in his summerjaunt across the country via freight train! The readerlearns about his first chemistry set, his woodworkingskills, his toils on a threshing machine, and much more.This level of intricate and personal detail persiststhroughout the book, as Dr. Beckman’s life is tracedfrom birth through his schools years, military service,and his three “careers” – from professor to entrepre-neur to philanthropist. His devotion to his wife and

family is not neglected and is neatly woven into thisamazing legend. The pictures and vignettes interspersedthroughout the book (a veritable Who’s Who in America)bring to life and personalize the warm friendly discoursesof the authors. If a picture is worth a thousand words,this book amounts to hundreds of volumes. Truly theseauthors know Arnold Beckman. It is absolutely impos-sible, as much as the ardent reader might want, to readthis book from cover to cover in one sitting. There isjust too much for the mind to absorb. No, one has toallow each section to sink in, sometimes letting days orweeks pass before picking it up again. But one will bedrawn back to read more until the end, then beg for more.

And there is more. Tucked neatly before thehardbound back cover is a compact computer disk. ThisCD is the icing on the cake. The CD adds additionalauditory and visual stimulation to the “reader.” With-out the CD the book is a magnificent work of art. Withthe CD, the reader becomes all that more familiar withArnold Beckman. When finished, the reader truly feelsas if he knows Arnold Beckman and, given the opportu-nity, would greet him as a friend.

If there is a downside to the book, it is that it is toobig and heavy. This review would have reached youmuch sooner, could I have carted it along with methrough airports. Even as a coffee-table book, I am afraidit is a bit large. So I am concerned that it will be rel-egated to the bottom shelf of the bookcase (where allthe over-sized books are placed) and forgotten. Thiswould be a terrible tragedy.

Let me end with a quote from Dr. Beckman: “I’dlike to get young kids interested in science... The youngmind is inquisitive enough that you don’t have to worry

among general readers. One can only hope that thisadmirable book will soon be translated into English.Alan J. Rocke, History of Technology & Science, 1141East Blvd., Case Western Reserve University, Cleveland,OH 44106-7107

Page 64: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

142 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

about scaring up enthusiasm; you simply need to keepthem interested and excited about science.” This bookshould be required reading for all aspiring scientists andengineers (especially chemists), as well as modern his-tory majors—not as a text book, but as a philosophical

and inspirational reading. This book will keep theminterested and excited about science. William F. Koch,National Institute of Standards and Technology, Chemi-cal Science and Technology Laboratory, Gaithersburg,Maryland 20899-8300.

The Partington Prize

The Society for the History of Alchemy and Chemistry has established the PartingtonPrize in memory of Professor James Riddick Partington, the Society’s first Chairman. It isawarded every three years for an original and unpublished essay on any aspect of the historyof alchemy or chemistry. The prize consists of two hundred and fifty pounds (£250).

The competition is open to anyone with a scholarly interest in the history of alchemy orchemistry who has not reached 35 years of age by the closing date, December 31, 2002.Scholars from any country may enter for the competition, but entries must be submitted inEnglish, typewritten or wordprocessed and double spaced on one side of the paper. Essaysmust be fully documented with the conventions used in recent issues of Ambix. Essays mustnot exceed 5000 words in length, excluding references and footnotes. All entries must besubmitted with a word count. The prize winning essay will be considered for publication inAmbix, but publication cannot be guaranteed.

All entries should be sent to the Hon. Secretary of the Society, J. A. Hudson, AppliedSciences, Anglia Polytechnic University, East Road, Cambridge CB1 1PT, England, with thewords ‘Partington Prize’ written clearly on the envelope. Each entry should contain a sepa-rate title page with the author’s name, institution, address, and date of birth; this informationwill not be made available to the judges. Essays (only one from each competitor) must bereceived no later than December 31 , 2002.

The decision of the judges appointed by the Council will be final. The Society reservesthe right to divide the prize between two or more entries of equal merit, or not to award a prizeshould no essay be deemed of suitable standard.

The name of the winner will be announced by April 30, 2003, and all essays will bereturned to competitors soon after that date.

John Hudson, Hon. Secretary

Page 65: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

Bull. Hist. Chem., VOLUME 26, Number 2 (2001) 143

FUTURE ACS MEETINGS

April 7-11, 2002—Orlando, FL

August 18-22, 2002—Boston, MA

March 23-27, 2003—New Orleans, LA

September 7-11, 2003—New York, NY

March 28-April 1, 2004—Anaheim, CA

August 22-26, 2004—Philadelphia, PA

March 13-17, 2005—San Diego, CA

August 28-September 1, 2005—Washington, DC

March 26-30, 2006—Atlanta, GA

September 10-14, 2006—San Francisco, CA

March 25-29, 2007—Chicago, IL

August 19-23, 2007—Boston, MA

April 6-10, 2008—San Antonio, TX

August 17-22, 2008—Philadelphia, PA

March 22-26-, 2009—Salt Lake City, UT

August 16-21, 2009—Washington, DC

March 21-26, 2010—San Francisco, CA

August 22-27, 2010—New York, NY

March 27-31, 2011—Anaheim, CA

August 28-September 1, 2011—Chicago, IL

March 25-29, 2012—San Diego, CA

August 19-23, 2012—Boston, MA

Page 66: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

144 Bull. Hist. Chem., VOLUME 26, Number 2 (2001)

INSTRUCTIONS FOR AUTHORS

Articles of 4-12 pages, typed, double-spaced (excluding references) should be submitted ashard copy (2 copies) and also on diskette, preferably formatted in Word or Wordperfect, toEditor, Bulletin for the History of Chemistry , at the University of Michigan. Chemical for-mulas, to be kept to a minimum, should be computer-generated and printed on separate sheets,with a clear indication of their location in the ms. Authors are encouraged to provide photo-graphs (black and white glossy prints) and drawings (black ink) to enhance the publication.Include a legend for photos, drawings, graphs and credits if appropriate. Diskettes, photographs,and drawings will be returned at the authors’ request.

The title of the article should be of reasonable length (up to 15 words); a subtitle may beincluded if appropriate. Authors should strive to make the title descriptive of the specific scopeand content of the paper, bearing in mind that the title will determine entries in the subject index.Subheadings within the paper may be used if authors feel their inclusion will enhance clarity.

Format as found in recent issues of the Bulletin should serve as a guide. Authors shouldconsult J. S. Dodd, Ed., The ACS Style Guide, American Chemical Society, Washington, DC, 2nd

ed., 1997. Format the document throughout with standard 1” left and right margins. Directquotations, which in the journal will be indented, will be repositioned by the graphic artist.

REFERENCES AND NOTES should appear at the end as part of the main document andnot buried as footnotes. References should conform to the format as illustrated below. StandardChem. Abstr. abbreviations are to be used (see CASSI); title of the article is in quotes. Booksand journals are italicized, as are volume numbers. The year of publication of periodicals (notbooks) is boldfaced. Inclusive page numbers are given for an article or partial section of a book.Note the placement of commas and periods. It is imperative to recheck the accuracy of refer-ences before submitting the ms. In the text references are identified by Arabic numbers withinparentheses—no superscripts.

1. O. T. Benfey, “Dimensional Analysis of Chemical Laws and Theories,” J. Chem. Educ., 1957, 34,286-288.

2. G. W. Wheland, Advanced Organic Chemistry, Wiley, New York, NY, 1949.3. J. R. Partington, A History of Chemistry, Macmillan, London, 1972, Vol. 4, 104-105.4. L. P. Rowland, Ed., Merritt’s Textbook of Neurology, 8th ed., Lea and Febiger, Philadelphia, PA,

1989.5. K. A. Nier, The Emergence of Physics in Nineteenth-Century Britain as a Socially Organized Cat-

egory of Knowledge. Preliminary Studies. Ph.D. Thesis, Harvard University, 1975.6. J. B. Conant, “Elmer Peter Kohler,” Biogr. Mem. Nat. Acad. Sci. USA, 1952, 27, 264-291.

Please provide a short biographical sketch, to be included as ABOUT THE AUTHOR(S)at the end of the article.

Readers are encouraged to submit short notes, where appropriate, and letters to the Editor,as well as articles. We would welcome hearing from those who have an interest in refereeingpapers and/or preparing book reviews.

Page 67: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

The BULLETIN FOR THE HISTORY OF CHEMISTRY (ISSN 1053-4385) is published biannually (February & September) bythe History of Chemistry Division of the American Chemical Society. All matters relating to manuscripts, book reviews, and lettersshould be sent to Dr. Paul R. Jones, Editor. Subscription changes, changes of address, and claims for missing issues, as well as newmemberships, are handled by the Sec./Treas.

BULLETIN FOR THE HISTORY OF CHEMISTRY

William B. Jensen, Founding Editor

Paul R. Jones, Editor Herbert T. Pratt, Ed. Board Dr. Peter Ramberg, Ed. BoardDepartment of Chemistry 23 Colesbery Drive Science DivisionUniversity of Michigan Penn Acres Truman State University930 N. University Avenue New Castle DE 19720-3201 100 E. NormalAnn Arbor, MI 48109-1055 Kirksville, MO [email protected] [email protected]

HIST OFFICERS, 2001

Richard E. Rice, ChairPO Box 1090Florence, MT 59833-1090charrice @juno.com

Vera V. Mainz, Sec/TreasSchool of Chemical SciencesUn. IllinoisUrbana, IL [email protected]

Roger A. Egolf, Program ChairDept. of ChemistryPenn. State. Univ.Fogelsville, PA [email protected]

Ben B. Chastain, Councilor538 Hampton DriveBirmingham, AL [email protected]

Mary Virginia Orna, Councilor16 Hemlock PlaceNew Rochelle, NY [email protected]

Albert S. Kirsch10245 Collins Avenue #12FBal harbour, FL [email protected]

Page 68: BULLETIN FOR THE HISTORY OF CHEMISTRYacshist.scs.illinois.edu/bulletin_open_access/FullIssues/Vol26-2.pdf · OF THE IMPROVED GAS BLOWPIPE 106 Martin D. Saltzman, Providence College

BULLETIN FOR THE HISTORY OF CHEMISTRYSubscription and Membership Application

____ ACS Member at $15/year (includes HIST Division membership)____ Associates: $20/year____ Associates outside North America: $25/year____ Annual Library Subscription: $36/domestic; $40/foreign

Begin subscription with the year 20 __

Name _________________________________________________________

Address _________________________________________________________

Signature _________________________________________________________

Return to Dr. Vera V. Mainz, HIST Secretary/Treasurer, School of Chemical Sciences,University of Illinois, Urbana, IL 61801. Checks should be made payable to the Division of theHistory of Chemistry, American Chemical Society.

BULLETIN FOR THE HISTORY OF CHEMISTRYDr. Paul R. Jones, EditorUniversity of MichiganDepartment of Chemistry930 N. University AvenueAnn Arbor, MI 48109-1055

Non-Profit Org.U.S. Postage

PAIDAnn Arbor, MIPermit No. 144