Ostwald, Friedrich Wilhelm

views updated May 18 2018

OSTWALD, FRIEDRICH WILHELM

(b. Riga, Latvia, Russia, 2 September 1853 ; d. Leipzig, Germany, 4 April 1932)

physical chemistry, Colorscience.

Together with van’t Hoff and Arrhenius, Ostwald established physical chemistry as a recognized and independent professional discipline and was its most important spokesman and organizer. His early reputation was based upon investigations into the fundamental principles governing chemical equilibrium and reactivity. A skillful experimentalist, he continued to give chemical affinity a central position in his research on electrolytic dissociation, electrical conductivity, mass action, catalysis, and reaction velocity. Ostwald received the Nobel Prize in chemistry in 1909 for his work in physical chemistry, and especially in recognition of his studies on catalysis. He was also one of the leading twentiethcentury researchers in color science, and enriched chromatics through his quantitative theory of colors and his subjective chromatic system. Ostwald was at the same time an inspiring teacher who restored the significance of general chemistry and induced a generation of chemists in Europe and the Americas to adopt a receptive attitude toward theoretical and physical chemistry in their teaching and research. Ostwald was a lucid, imaginative, prolific, yet often controversial writer, synthesizer, expositor, and apostle of scientific ideas. He was a man of notable charm, enormous capacity for work, and many-sided intellectual interests, and was actively involved in the cultural and philosophical debates and humanistic strivings of his time.

Ostwald was the second son of Gottfried Ostwald, a master cooper, and Elisabeth Leuckel, the daughter of a master baker. His parents were descended from German immigrants who came to Livonia from Berlin (on the paternal side) and from Hessen(on the maternal side). Ostwald was educated in Riga, where he attended the Realgymnasium. Even in his student days Ostwald’s interests ranged very widely, focusing especially on physics and chemistry as well as literature, music, and painting. He shared his love of music–he played the viola and the piano–with his mother, to whom he felt especially close. Like his father Ostwald developed considerable skill in painting and handicraft. The latter wished his son to become an engineer, but Ostwald’s strong inclinations toward chemistry were finally decisive in his choice of profession.

In 1872 Ostwald enrolled at the University of Dorpat (now Tartu), where he studied chemistry under Carl Schmidt and Johann Lemberg and physics under Arthur von Oettingen. Within a short time the highly gifted Ostwald passed all the examinations, and in 1875 he received the candidate’s degree . The following year he was awarded the master’s degree and became Privatdozent at the University of Dorpat, where he lectured on the theory of chemical affinity and, in 1878, earned the doctorate in chemistry. Ostwald had become an assistant to von Oettingen in 1875; and the modest sum he earned from this work, added to the income he received from a second job teaching physics and chemistry in a Dorpat Realschule, enabled him to establish a household. On 24 April 1880 he married Helene von Reyher, the daughter of a Riga surgeon. They had five children, one of whom, Wolfgang, became a colliod chemist.

In 1881 Ostwald was appointed professor of chemistry at the Riga Polytechinc Institute, where he quickly proved to be an outstanding teacher and began two important undertakings, that made him widely known. First in 1885-1887 he wrote the ambitious Lehrbuch der allgemeinen Chemie (called “der grosse Ostwald” in order to distinguish it from “der kleine Ostwald” of 1889, Grundriss der allgemeinen Chemie). The work was a substantial contribution to the establishment of physical chemistry as a separate branch of the discipline and was published in a partial second edition between 1897 and 1902. It was based on the systematic examination of fifty years of journal literature in physics and chemistry. Second, with van’t Hoff, Ostwald began to publish Zeitschrift für Physikalische Chemie which become the mouthpiece of the Leipzig school of physical chemistry. This journal quickly established its importance and become the organizational link uniting physical chemists of various countries. In it the chemists associated with Ostwald, Arrhenius, and van’t Hoff (who were dubbed “the iniosts”) broadly disseminated the new ideas of physical chemistry and realted fields-such as the existence of ions, which was still disputed by some–in articles that often were aggressive in tone.

Ostwald expanded his activity as a teacher and researcher in September 1887, when he accepted an appointment at Leipzig to the only chair of physical chemistry then existing in Germany; the position had become free when Wiedeman left it to assume the chair of physics at Leipzig. The other candidates, including van’t Haff. withdrew in favor of Ostwald. Ostwald’s work in physical chemistry during the next two decades was as fruitful as it was remarkable. He become the head of one of the discipline’s most important schools and soon was able to propose his students for professorships at univeristies throughout the world. His co–workers and assistants included Arrhenius, Nernst, Le Blanc, and R. Luther. Ostwald also introduced new methods in the analytical training of chemists and wrote a text–book on that subject, thereby becoming an important pioneer of new approaches to the education of chemistry students. He laid particular stress on independent work and critical discussion of results obtained by students, thus anticipating modern forms of teamwork.

In 1898 Ostwald celebrated the offical dedication of the new physical chemistry institute of the University of Leipzig,1 which became a training center for generations of physical chemists. Ostwald, however, remained, there for barely a decade. As early as 1894 he had wished to be free of teaching and official duties such as elected deanship positions; he wanted to work only as a research professor and continue his literary activities (especially writing his books and editing Zeitschrift für Physikalische Chemie) and other responsibilities. He retired in 1906, after having been appointed as the first German exchange professor to Harvard University (academic year 1905-1906).2

In 1901 Ostwald had purchased an estate in Grossbothen, a village near Grimma. The house on the edge of the woods was baptized “Energie,” –as he wrote to Mach in Sepetember 1901. Upon his resignation from the Univeristy of Leipzig in 1906, he moved his family and tremendous library to “Landhouse Energie” and spent the rest of his life there as an independent scholar and freethinker, devoting his efforts to energetics, scientific methodology, the organizational aspects of science, a world language, internationalism, and pacifism. Not long after he settled at Grossbothen, he enlarged the building and added a laboratory for his color research.3 He died twenty–six years later, following a short illness.

Ostwald’s achievements as a young chemist lay primarily in chemical affinity studies that along with both general and inorganic chemistry, had been relatively neglected since the work of E. F. Geoffory, (1718), Bergman, (1775), C. F. Wenzel, (1775), and Berthollet(1803), Ostwald’s approach to the subject was to demonstrate that quantitative values for physical properties, such as specific volume and refractive index, could be correlated with the qualitative changes accompanying chemical transformations–and thus provide information about the relative affinities of the constituents of a chemical reation.

In the mid–1870’s when young Ostwald became interested in the study of chemical affinity (Verwandtschaftslehre), the interest of German chemists, both at the universities and in industry, was directed almost exclusiverly toward organic chemistry. Osward’s motivation for taking up the study of problems in physical chemistry, therefore, merits some comment. To judge from the style of teaching and research at the University of Dorpat, the intense preoccupation with organic chemistry stopped at the German border. The professor of chemistry, Carl Schmidt (1822-1894), had studied with Heinrich Rose at Berlin, with Liebig at Giessen, and with Wöhler at Göttingen; but at Dorpat he had become totally involved with the study of the mineral content of ground and surface waters as a means of furnishing information concerning the chemistry processes of rock formation. In these studies Schmidt followed the direction of the earlier work of C. G. C. Bischof of Bonn-the founder of geological chemistry in Germany, In his Lebenslinien Ostwald relates how Schmidt’s chief assistant, Johann Lemberg (1842-1902), following some leads given in Bischof’s textbooks, early taught him to pay close attention to problems in chemiscal equilibrium, mass action, and reaction velocity and to recognize that there are no absolutely insoluble substances in nature. Thus the ground was laid for a way to chemical thinking that eased Ostwald into seriously reckoning with the importance of mass considerations in chemical process. Ostwald remarked that he undoubtedly would have become an organic chemist if he had studied chemistry in Germany.

As a student at Dorpat, while examining the thermochemical investingations of Julius Thomsen, Ostwald conceived the idea of experimently exploring a problem in chemical affinity that would employ physical methods of investigation but would be more general than the current focus in Dorpat on the weathering of minerals. He chose to study the extent of decomposition of bismuth chloride by water in order to compare his results with Thomsen’s calorimetrically determained affinity. Ostwald conjectured-correctly, as it turned out-that, in principle, the magnitude of chemical change in a reaction might be calculated from any experimentally measurable changes in physical property. Accordingly, Ostwald, saw Thomsen’s methold to be but one of many concerivable ways to approach the study of chemical affinity.

Gibbs (1875-1879) and Helmholtz (1882-1883) had demonstrated from considerations of the second law of thermodynamics that the heat of reaction ΔH for a chemical reaction, in general cannot be relied upon to furnish information about chemical equilibrium and reactivity unless the entropy term Δ S is zero in the equation Δ FH - TΔS in which Δ F represents the free energy for the process and T the absoulte temperature. Thus Thomsen’s thermochemical investigation proved to be more limited than Ostwald’s methods, since change in many physical properties-such as specific volume, refractive index, viscosity, color electrical conductivity, and rotation of polarized light-can be correlated with changes in activities (accurately) and changes in concentrations (approximately) to yield the equilibrium constant K and thus the free energy, since Δ F= -Δ Rt In K (R is the gas constant). Ostwald’s important contribution was his recognition of the unique advantages of physical methods of investigation in the solution to chemical problems. This is especially important in chemical thermodynamics, since the analysis of the constituents of a reaction by chemical means almost always is rendered impractical by a concurrent shift in the equilibrium during the analysis; physical methods, on the other hand, do not cause chemical changes in the system.

In his Volumschemische Studien iiber Affinität (master’s thesis, 1877) Ostwald determined the volume changes that take palce during the neutralization of acids by bases in dilute solution. Pycnometers were used to determine the specific volumes before and after the reaction. From the differences in specific volumes (for specific concentrations and constant temperature) Ostwald was able to calculate the chemical action (affinity) for the neutralization reactions. Specifically, he showed that the distribution of a base between two acids can be determined by measuring the specific volume of a solution of each acid and of the base, of the solution formed by mixing each acid separately with the base, and of the solution formed by mixing both acids simultaneously with the base. With minor exceptions Ostwald found that his dilatometric methods yielded results in close agreement with the order of “avidities” determined by Thomsen’s thermochemical method. Like the latter, Ostwald’s methods seemed to confirm the Guldberg-Waage law of mass action.

In his Volumchemische und optisch-chemische Studien (doctoral dissertaion, 1878), Ostwald enlarged his investigations to include the determination of the coefficients of refraction of a large group of acid-base and other double decomposition reaction. Thus he obtained values for chemical reactivity that confrimed those obtained by the specificvolume methods, but he felt that the optical methods was less trustworthy than the volumetric methods. In addition his chemical affinity studies were extended to incroporate the analysis of both homogeneous and heterogeneous reactions as a function of temperature. In this way Ostwald was able to attach specific numerical values to the term “affinity”, which had long been referred to in the literature in a qualitative and often arbitrary way.

In his 1879 review article, “Chemical Affinity” M. M. Pattison Muir of Cambridge University stated (p. 182); “The most important contributions, made with recent years towards the final solution of the problem of chemical affinity are contained in two papers by Guldberg and Waage [1869 and 1879], and three papers by W. Ostwald [1877and 1878];” Having given a fairly detailed summary of the results of these papers, Muir concluded (p.203); “Ostwald furnishers chemical with a new method for solvingt some of the most difficult problems; and Guldberg and Waage lead the way in the application of mathematical reasoning to the facts of chemical science.”4 Thus Ostwald’s important position in physical chemistry was already recognized when he was but twenty-six years old.

In 1879 Ostwald proposed that the rate at which compounds like zinc sulfide and calcium oxalate are dissolved by different acids be used as a measure of the relative affinities of the acids. He found that the dynamically determined affinity coefficients agreed satisfatorily with those obtained earlier by statisitical methods. In his first teaching post at the Polytechnic Institute at Riga, Ostwald pursued the studies on chemical reaction kinetics with resolution. His Studien zur chemischen Dynamik (1881) treated the reaction velocity for the acid-catalyzed saoonification of acetamides and the hydroloysis of esters. Experiments on the rate of inversion of cane sugar in the presence of various acids (1884-1885) gave Ostwald a second opportunity to evaluate the affinites of acids and to compare his values with those he had measured by other methods. With some exceptions, he found that the various methods gave comparable results. He explanined the exceotions on the basis of secondary reactions that influence some chemical processes more than others.

Such was the state of chemcial affinity studies prior to the enunciation of the theory of electroloytic dissociation. Ostwald had produced a substantial body of experimental evidence, mostly for acids and bases, that demonstrated that different reactions can be quantitatively characterized by affinity coefficients that depened on the constitution of the acids and their degree of dilution. His general conclusion was that each acid and each base can be assigned an affinity coefficient that quantitatively describes all of its specific reactions, and further, that the relative values of these coefficients, which are independent of the nature of the chemical reaction. can be measured with tolerable accuracy by various physical (static and dynamic)and chemical methods. It is pretinent to add that the thermochemical methods perfected by Thomsen and Berthelot, and the physical methods explored by Ostwald and his students, measured chemical reactivity with fair reliabilty most of the time but not invariably. The former were in adequate because of the disregard of the entropy term, and the latter were no more than approximate because of the neglect of nonadditive secondary effects.

In 1884 Ostward read the memoir on the galvanic conductibility of electrolytes, submitted by Arrhenius as a dissertation to the Swedish Academy of Sciences the previous year,5 From then on, Ostwald became an enthusisastic proponent and crusader for Arrhenius new theory of dissociation. In his memoir Arrhenius had demonstrated that the electrical conductivity of acids is proportional to their strength. He met Arrhenius at Stockholm in Auguest 1884 nad the two men commenced a close collaboration that led to a lifelong friendship, of which Ostwald gave a detailed account in Lebenslinien and that is vividly displayed in the Ostwald -Arrhenius correspondence. Though Ostwald’s influence Arrhenius received a five-year traveling scholarship that began in Ostwald’s laboratory at Riga (1886) and ended in his laboratory at Leipzig (1889-1891). During the intervening years he came into working contact with Friedrich Kohlrausch at Würzburg (1886), Boltzmann at Graz (1887) and van’t Hoff at Amsterdam (1888).

In 1886 Ostwald discovered van’t Hoff’s Études de dynamique chimique (1884), which focused on the application of thermodynamics to chemical problems. Now more than ever Ostwald grasped the deep significance that the concept of electroloytic dissociation would have when applied to questions of affinity, equilibriam, mass, action and chemical thermodynamics in general. From that time on, van’t Hoff became the third member of the physical chemistry triumvirate that soon moved its spiritual center to the University of Leipzig. Once there, Ostwald chose Nernst as his chief assistant in physical chemistry, and the laboratory became a mecca for enterprising graduate students from all over the world, especially the United States.6

The full-blown theory of electroloytic dissociation that Arrhenius enunciated in Zeitschrift für physickalische Chemie in 1887 was by, then, seen to be the logical outcome of a number of important previous studies on solution chemistry. Arrhenius had ealier entertained comparable ideas, but he considered them too bold to state verbally; and his university mentors felt that the twenty-four-year-old student’s speculation–identifying electrolytic dissociation with chemical activity–was premature and inadequately supported. Ostwald immediately designed his own experimental program to relate chemical affinity to electrical conductivity. By 1888 Ostwald Planck, and van’t Haff had independetly applied the law of mass action to the equilibrium distribution between the ions and the undissocicated portion of an electrolyte.

Hittorf’s electrolytic salt solution studies of 1853 and 1859 already had led to the suggestion that the electric current is carried by ions moving at different rates toward the electrodes. In 1874 Kohlrausch had shown that for diltute salt solutions the quotient of electrolytic conductivity and concentration is the same as the sum of both terms. Raoul (1882) deduced a general law for freezing–point depression of solutions containing nondissociating organic solutes. Five years later he announced his general law of the effect of sloutes on the vapor pressure of solvents in Competes rendus… de l Académie des sciences. He explained the anomalous freezing point depression and vapor pressure of salts–anomalous when compared with organic solutes–by suggesting that the salt molecules dissociate into other molecules.

In his comprehensive statement of the theory of electrolytic dissociation (1887), Arrhenius reasoned convincinly that all of the above–mentioned investigations could be accounted for by postulating, as Clausius had suggested in 1858, the dissociation of molecules into electrically charged ions. Accordingly, compounds that had been though to be held together by the strongest affinities–such as sodium chloride, hydrogen chloride, and potassium hydroxide–now were seen, in dilute solution, to be largely dissociated.

Arrhenius and Ostwald recognized that an electrically conductive solutive–where the conductivity depends on the concentration and the temperature–consists of two different kinds of molecules; active ions and inactive undissociated molecules; molecuever, the concentration of the active portion increase with dilution at the expense of the inactive portion and reaches a limiting (maximum) valuse, presumably when all of the inactive molecules have been transformed into active ions. The important point in terms of chemical affinity studies is that like acids, whose strength increase with electrical conductivity, so too, the chemical activity of electrolytes coincides with their electrical conductivity.

In 1885 Ostwald initiated a comprehensive program to redetermine, by using Arrheniu’s electrolytic conductivity methods, the affinities of the acids he had studied earlier by other physical methods. He concluded that electrolytic conductivity measurements were far more elegant and less tedious than his own specific volume methods. His experiments showed that for strong monobasic acids in aqueous solution, the molecular conducitivity gradually increases with dilution and asymptotically approaches a maximum value at infinite dilution. Ostwald’s dilution law (Verdünnunagsgesetz), in its essentially modern form, was theoretically derived in 1888 and was supported with impressive experimental evidence the following year. Ostwald reasoned that since the laws of gas pressure had been shown to be applicable to the osmotic pressure for dilute solutions of nonelectrolytes, the formula for a partly dissociated gas might likewise be applicable to partly dissociated solutions. Ostwald showed that for a binary electrolyte of volume, v with μu representing the molecular conductivity at volume v and μrepresenting the limit of conductivity at infinite dilution, it follows that

where Kc is the equilibrium constant for the chemical reaction in question.

This equation shows that the constant has the same value for any given binary electrolyte at all degrees of dilution, and is the products of factors that depend solely on the composition and constitution of the acids or bases the consductivities of which are being investigated. Ostwald confirmad the law for 250 acids, all of which were “weak acids”; the law would not be applicable to strong acids or salts that possess no dissociation constant. Georg Bredig (1868-1944), Ostwald’s assistant at Leipzig ,studied fifty bases for which the dilution law was seen to be valid. The historical importance of Ostwald’s dilution law is that the law of mass action was first applied to dilute solutions of weak organic acids and bases in that form. Most inorganic acids are strong electroloytes in aquesous solution; and salts, even of was organic acids or bases, are highly ionized and do not obey Ostwald’s law. The law was therefore found to hold excellently for all slightly ionized electrolytes, but to fall very wide of the mark for highly ionized eletroloytes.

The first issue of the Zeitschrift für Physikalische Chemie appeared in February 1887. During its first year the journal published articles by van’t Hoff, Ostwald, Arrhenius Lothar Meyer, Raoult, Guldberg, Mendeleev, Julius Thomsen, Le Châtelier, and Planck, among others. Van’t Hoff’s paper “Die Rolle des osmotischen Druckes in der Analogic zwischen Losungen und Gasen” provided a crucial link between the thermodynamics of gases and of solutions. In “Ueber die molekulare Constitution verdunnter Lösungen, “Planck argued that harmony between the laws of thermodynamics and freezing point and vapor pressure studies could be maintained only by assuming that the molecules of dissolved solutes are altered in solutions in the way that Arrhenius had suggested. The next-to -the-last article in the Zeitschrift für Physikalische Chemie of 27 December 1887 was Arrhenius’ “Ueber die Dissociation der in Wasser gelæsten Stoffe,” a forthright and comprehensive statement of the electrolytic theory of ionization. The support for this theory by Arrhenius, Ostwald, van’t Hoff, Planck, and Nernst opened a new chapter in the history of physical, experimental, and theroetical chemistry, that brought a large and varied range of phenomena under a single point of view. The theory notably served to tie together all of these properties of a system encountered in the study of solutions that depend primarily on the number of molecules involved and not on their nature. Following a suggestion of the philosopher-psychologist Wilhelm Wundt, Ostwald, designated such properties as “colligative.”

The value of the theory became conspicuously evident by the way in which it reconciled, in a quantitative way, hitherto, disparate physical and chemical investigations; the avidity of acids and bases, heats, of neutralization, the dissociation of water, the behavior of weak acids and bases, the hydrolysis, of salts, the catalytic activity of acids and bases, equilibrium in electroloytic solutions ,theory of conscentration, cells, liquid junction potentials and many properties, of solutions such as specific volume, refractive, index, optical activity, and absorption spectra.

In Ostwald’s laboratory at Leipzigconcerted efforts were made to show that the many empirically established regularities observed for chemical processes in solution could all be deduced as necessary consequences of the ionic theory of dissociation. The result was a substantial body of expreimental evidence, mostly for acids and bases, that clarified the connection between electrolytic dissociation and conduction. It showed that the coefficients of affinity, which measure the degree of dissociation of the acids and bases, were independent of the particular chemical process under investigation but nearly proportional to the conductivities in solution. They depend, in turn, upon the number of free ions present and their velocities. Ostwald’s Grundriss der allgemeinen Chemie (1889) and the English translation by James Walker (1890) made these ideas of van’t Hoff’s widely known.

In 1891 Ostwald formulated a theory of acid-base indicators in which he used the principle of ionic equilibrium to account for the ratio of un-ionized weak acid (of one color) to ionized weak acid (of another color). Subsequent investigations, stimulated by Ostwald’s contributions, revealed as, Hantzsch, and his pupils showed in 1906, that the organic indicators were pseudo acids and pseudo bases-that is nonelectroytes susceptible to the formation of matellic derivatives by changing into acidic and basic isomers. The new views did not substantially alter the quantitative formulation of Ostwald’s theory. The entire field encompassed by the theory of chemical reaction based on electrolytic dissociation, including indicator theory, was expounded in great detail in Ostwald’s Die wissenschaftlichen Grundlagen der analytischen Chemie (1894), a work that revolutionized the teaching of analytical chemistry.

Over a period of twenty years at Leipzig wald championed and, with his students, and assistants, fürnished experimental support for the electroltic dissociation theory, demonstrating its remarkable value against the bitter hostility of the chemists. The focus of the opposition to Ostwald and his revolutionary colleagues was concentrated in Berlin, and continued to dominate the chemical scene there until Nernst (who had been with Ostwlad and his revolutionary colleagues was concentrated in Berlin, and continuned to dominate the chemical scene there until Nernst (who had been with Ostwald at Leipzig from 1887 to 1891) accepted a professorship of physical chemistry at Berlin when Landolt retired in 1905-the year in which Ostwald gave up his teaching career in Leipig. A common criticism was that the proponents of the theory-called “die Ioner”-had tried too singularly to set the theory in the foreground of chemical research and had attempted to encompass too varied a collection of chemical phenomena. In any case, to explain the formation and stable existence of electrically charged ions in solution, and to account for the role of the solvent in this process, appeared to be too formidable a theoretical undertaking. The reason for ionization and the stability of ions remained a puzzle until the electronic theory of the astom provided clues to the connection between chemical affinity and electricity.

Strong electrolytes had always been an anomaly for the electrolytic dissociation theory, and there was little prograss on this matter until the turn of the century. With weak (slighly ionized) electrolytes Ostward’s dilution law was seen to hold quite rigorously, but for strong electrolytes not even approximately. This, of course, brought up the question of the general validity of the theory of ionization in relation to the law of mass action. An important factor in the revision of the theory was introduced with the discovery, by X-ray analysis, that salts and other strong electrolytes in the crystalline salts and other strong electrolytes in the crystalline state are composed of ionic lattices. This suggested that since the ions preexist in the solid state, these substances are completely ionized at all concentrations. This was about the time that Ostwald lost all desire to continue his chemiscal studies. Other investigators-Bjerrum (1906), Hantzsch, (1906), and Debye, E.Höckel, and Lars Onsager (1923-1927)-approached the problem from a more theoretical and mathematical point of view, drawing on the electronic theory of atomic structure and on the concepts of electrovalence and covalence to extend the original classical theory. In doing so, they elaborated in fine detail the interionic forces, the solvent interaction (solvation), the ionic mobility and ionic atmosphere terms, and the time of relaxation (electrical drag), considerations that led to a far more complex, but also considerably more attractive and useful, theory of ionization.

G.N.Lewis’ introduction, in 1907, of “activity” a (or effective concentration) and “activity coefficient” α(or deviation multiplicand) – such that the actual concentration c is given as a/α-manifestly helped to introduce a uniform method of treating the mathematical equations that describe ionization. Accordingly, it became standard practice to adopt the Arrhenius-Ostwald ratio of electrolytic conductivities MuM∞ as the apparent, and not the actual, degree of ionization. This mode of representing the data contributed notably to the elimination of numerous misunderstandings and misterpretations of the literature.

All things considered, the modern position in chemical solution theory is rather far removed from the clasical theory of ionization, notwithstanding its historical significance in the search for a theory of the constitution of electrolytes. From the point of view of order-disorder considerations, it is not surprising to discover that to date our understanding of the structure of pure liquids and solutions has been far less impressive than that of the structure of matter at its extremes in the crystalline and the gaseous forms.

The systematic quantitative investigation of catalysis in Ostwald’s laboratory during the last decade of the nineteenth century brought that heterogeneous and untidy sudject within the domain of chemistry kinetics. In 1835 Berzelius had introduced the term “catalysis” to designate the process by which, through some special kind of force, a relatively small quantity of a substance aroysed the slumbering affinities of other substances and hastened a chemical reaction without itself undergoing any change. Many and varied examples of catalysis were reported, classified, and commented upon throughtout the century. While in Riga, Ostwald had studied the reaction velocity of chemical processes. In 1890 he reported on the phenomenon of “autocatalysis” and defined it as a process that is “provoked or accelerated through the presence of certain substances with no demonstrable participation of these in the compounds.”7 In an 1894 report Ostwald gave a new rendering of the concept of catalysis that had been introduced by Berzelius; “Catalysis is the acceleration of a slowly proceeding chemical process through the presence of a foreign substance.”8 Ostwald compared catalysts to the action of oil on a machine and of a whip on a lazy horse.

Important experimental work on catalytic processes was carried out in Ostwald’s institiute. This work is conveniently summarized in his paper “Über Katalyse” (1901). The most important aspects of the experimental contributions on this subject by Ostwald and his co-workers are the ones that treat the process of crystallization from supersaturated solution-for both homogeneous and heterogeneous reactions-and the effect of enzymes. The work on supersaturation and supercooling showed that a system moving from a less stable to a more stable state goes by stages to the one lying closests at hand, and nor necessarily to the most stable of all possible states. This is known as Ostwald’s saw of stages.

Catalytic activity become, for the propoents of the theory of ionization one of the best ways to measure acid strengh. Ostwald took full advantage of such catalytic experiments in his support of Arrhenius’ theory.From the time of the investigations, of Bertheolt and Péan de Saint-Gilles, on the equlibriu’s between acetic acid, ethly alcohol, ethyl acetate, and water (1862-1863), it had been known that this homogeneous acid-catalyzed esterification reaction is accelerated by the pressence of acids and alkalies that remain unchanged in the overall process. Ostwald and Arrhenius recognized that the rate-determining factor in this catalyzed esterification, and similar reactions, was the acid and that the catalytic activity was proportional to the conductivity of the acid (and to the hydrogen ion concentration) but independent of the nature of the anion. This proved to be true only for reactions catalyzed by weak acids.

In the course of catalytic investigations Ostwald and Oskar Gros developed a photographic contact process that Ostwald called Katatypie. Ostwald also applied his knowledge of catalysis to two largescale industrial chemical projects that, however, did not bring him the sucess he hoped for. With E.A.Bodenstein and With E. A. Bodenstein and later with Ernst Brauer, who subsequently became his son-in-law, Ostwald developed a process for the synthesis of ammonia from nitrogen and hydrogen gases at high temperature and pressure, using heated bundles of iron wire as the catalyst, (1900). After a long series of attempts, a process was developed for the catalytic oxidation of ammoina to nitric acid that would stop short of complete oxidation to free nitrogen. The process was exploited industrially beginning in 1906, but by that time, Ostwald and Brauer were no longer associated with the project.9

A theortically significant point, first emphasized by Ostwald and based on a simple argument that excludes a perpetuum mobile, was that a catalyst exerts no change in the overall energy relations, and therefore does not alter the thermodynamically stable, reversible equibrium, position of a reaction. Ostwald thus saw that a catalyst must accelerate the forward and the reverse reactions in the same proportion in order for the equilibrium constant in the Guldberg-Waage law to retain its thermodynamic significance. Whether a catalyst neceassarily intiates a reaction, and whether it enters into compounds formation with the reactants as an intermediary, were questions for which no convincing answeres were supplied by Ostwald and his collaborators. Indeed, from the intense investigation into catalysis, subsequent to Ostwald’s work-including autocatalysis, inhibition phenomena (poisoning), contact and surface effects, enzymes produced by living systems-no single comprehensive and adequate theories resulted, except some that proved to be untestable.

A much-discussed and plausible theory for some cases of catalytic behavior by Clément and Desormes (1806), was the one according to which the catalyst forms an intermediary metasble compound with one of the constitutents of the reaction and then is regenerated in its original form in the process that leads to the final products of the reaction. Theories for the activity of catalysts and the principles for selecting them in the promotion of specific reactions have been hotly pursued in the twentieth century in connection with the biochemistry of enzymes and the rearrangement of aliphatic and aromatic hydrocarbons.

At the end of the nineteenth century the chemical community had reluctanly accepted Daltonian atomism. For many investigators it apparently was no more than a useful hypothesis for which there was little experimental evidence. Foremost in the vanguard of chemists who were skeptical about the phycial reality of atoms and molecules, and who regarded them larely as mental artifices, was Ostwald. His anti-atomistic sentiments were closely connected with an aversion to mechnical doctrines and a strong belief in an enery-based scientific program that he hoped eventally would encompass the natural and social sciences and the humanities in one vast monistic Weltanschauung, He began seriously to develop his ideas on energetics around 1890, and thereafter gradually reorganized all of his thoughts and work around energy.

The central position in Ostwald’s science of “energetics” was held by the energy concept. For him energy was not, as in physics, a derived quantity but, rather far more fundamental; “In fact energy is the unique real entity [das einzige Relae] in the world and matter is not a bearer [Träger] but rather a manifestation [Erscheinungsform] of the formre.”10 In particular, Ostwald maintained that the prinicple of energetics would furnish a more tangible basis forchemistry than would the kinetic-molecular theory. He went so far as to declare that the concept of matter was superfluous and that phenomena could be accounted for satisfactorily by analyzing the energy transformations taking place in nature and the laboratory. Whatis called matter, he argued, is only a complex of energies found together in the same place; differences between substances were reduced to differences between their specific energy content. Ostwald never proved his thesis. and more than one critic pointed out that the experimental facts of stoichiometry could not be deduced at the time from premises that did not already contain them. Unfortunately, Ostwald’s premises included the facts.

Despite the obvious pertinence of the atomic conception for stereochemistry by the 1870’s Ostwald apparently could not rid himself of the feeling that atmos were the limping remnats of an irresponsible, speculative Naturphilosophie. As early as 1887, in his inaugural lecture at Leipzig. Ostwald spoke on “Die Energie und ihre Umwandlungen.” In 1891-1892 he set forth his views on the importance of energy in “Studien zur Energetik.” These studies were influenced by Gibbs’s thermodynamical writings, which Ostwald translated into German in 1892 and thereafter became required reading for all students in Leipzig. Ostwald mentions in Lebenslinen that the English and American students had to read Gibbs in German until Yale University put out a new edition of Gibb’s works after his death. It is appropriate to add here that with his characteristic enthusiasm, Ostwald took laws and conceptions from thermodynamics and generalized them in the form of “physical energetics” that reached beyond the field in which they were known to be valid

Ostwald, Georg, Helm, and Le Châtelier supported and made considerable use of the system of chemistry developed by the philosopher-chemist František Wald of the Czech Technical University of Prague.11 According to the basic tenets of this system, the composition of chemical compounds depended on the physical circumstances under which they were prepared. The constant composition of substances, and their fixed physical and chemical properties, were explained, much as Berthollet had done earlier, as the end products of the chemists-and nature’s–process of preparation. For example, repeated crystallizations and distillations were said to result in constant composition. The constancy of multiple elemental weight ratios for many compounds, especially in organic chemistry, was explained by Ostwald on the basis, of “the law of integral relations”.

The science of energetics became a controversial issue as the result of Ostwald’s lecture “Die Ueberwindung des wissenschaftlichen Materialismus,” delivered at the Lübeck meeting of the German Society of Scientists and Physicians in 1895. The position of Ostwald and Georg, Helm, a physical chemist from Dresden–that energy had displaced matter as concept–was challenged on the floor, and then in the scientific journals, by Boltzmann, Planck, Felix Klein, Victor Meyer, Wislicenus, Nernst, and Arthur von Oettingen.12 Boltzmann stated bluntly that he saw no reason why energy should not be atomistic. Ostwald was unsuccessful in his attempt to persuade scientists of the value of his energetic views; and within the next few years, as evidence for the particulate nature of matter became more convincing, both physicists and chemists came to hold the view that energetics was an aberration. This lack of acceptance did not deter Ostwald from diligently pursuingand vigorously proclaiming an energy–rooted chemistry as an alternative to atomism. In fact, he extended the scope of his energeticist ideas to the level of a world view that he maintanied even after he had adoped the atomic–molecular theory of matter. Ostwald’s plan for a systematic philosophical energetics was laid out in his Vorlesungen über Naturphilosophie (1902), which was dedicated to Ernst Mach, whom he considered the person who had influenced him most.13 Mach was of one mind with Ostwald on the rejection of atomism. but had no use for his energetics or that anyone else. Ostwald’s views on energetics were exlpored in grand style in Annalen der Naturphilosphie, which he edited from 1901 to 1914.

Speaking to the Fellows of the Chemical Society in the theater of the Royal Institution, Ostwald said in his Faraday lecture(1904):

It is possible, to deduce from the principles of chemical dynamics all the stoichiometrical laws; the law of constant proportions, the law of multiple proportions and the law of combining weights. You all know that up to the present time it has only been possible to deduce these laws by help of the atomic hypothesis. Chemical dynamics has, therefore, made the atomic hypothesis unnecessary for this purpose and has put the theory of the stoichiometrical laws on more secure ground than that furnished by a mere hypothesis.

I am quite aware that in making this assertion I am stepping on somewhat volcanic ground. I may be permitted to guess that among this audience there are only very few who would not at once answer, that they are quite satisfied with the atoms as they are, and that they do not in the least want to change them for any other conception. Moreover, I know that this very country is the birthplace of the atomic hypothesis in its modern form, and that only a short time age the celebration of the centrnary of the atomic hypothesis has reminded you of the enormous advance which science has made in this field during the last hundred years. Therefore I have to make a great claim on your unbiased scientific receptivity… . If my ideas should prove worthless, they will be put on the shelf here more quickly than anywhere else, before they can do harm.14

Ostwald proceeded to demonstrate, to his own satisfaction but not to that of more than a handful of his chemistry colleagues, that one could redefinr elements, compounds, and solutions without reference to the atomic conception while maintaining consistency with the empirically determined stoichiometric laws of chemical combination. He accomplished this by exploting Gibbs’s thermodynamic work and by classifying the equilibrium of chemical and physical systems in the language of the Gibbs phase rule. At the same time he drew heavily on van’t Hoff’s law of mobile equilibrium and Le Châtelier’s principle that a chemical system under stress tends to resist change by moving in the direction that opposes the stress. Ostwald designated a system as a “hylotropic” body when the properties of each of its coexisting phases remain unchanged during the passage from one phase to another. He wrote:

A substance, or a chemical individual, is abody which can form hylotropic phases wihtin a definite range of temperature and pressure.… There are substances which have never been transformed into solutions, or whose sphere of existence covers all accessible states of temperature and pressure. Such substances we call elements. In other words, elements are substances which never form other than hylotropic phases.15

Drawing on the “eamanation” experiments of his close freind Sir William Ramsay, and using the notion that elements conform to regions of low potential energy in matter, Ostwald proposed a stalactite model of element transmutation in which the high energy barriers (long stalactites) represented the lighter elemnts and the lowest energy barriers (extremely short stalactites) represented the heavuiest transmutable elements.

The conversion to atomism, or at least to a granular conception of the structure of matters, by Ostwald and such other diehards as Lord Kelvin, Mach, Wald, adn Helm came rather suddenly about 1906, when evidence for the particulate nature of matter became dramatic. Cathode ray experiments had prepared the way for J.J. Thomson’s indentification of these rays as atoms of electricity-electrons-for which the ratio of cahnge to mass could be calucluated. In their paper “The Cause and nature of radioactivity” (1902), Rutherford and Soddy had used their experimental results as a means of “obtaining information of the process occuring within the chemical atom.”16 By 1906 Perrin had drawn attention to experimental observations on Brownian motion that offered a quantitative method for studying the random motion of molecular collisions. Einstein and Smoluchowski had shown theoretically (1905-1906) that the thermal motion of the kinetic retically (1905- 1906) that the thermal motion of particles could be subjected to accurate statistical analysis by means of the kinetic theory of Brownian motion.

In 1909, in the preface to the fourth edition of his Grundriss der allgemeninen Chemie, Ostwald made a straightforward confession about his adoption of the idea of the physical existence of atoms:

I am now convinced that we have recently become possessed of experimental evidence of the discrete or grained nature of matter, which the atomic hypothesis sought in vain for hundreds and thousands of years. The isolation and counting of gaseous ions, on the one hand, which have crowned with success the long and brillant reseraches of j. J.Thomosn, and,on the other, the agreement of the Browinan movement with the requirements of the kinetic hypothesis, established by many investigators and most conclusively by j. Perrin, justify the most cautious scientist in now speaking of the experimental proof of the atomic nature of matter. The atomic hypothesi is thus raised to position of a scientifically well-foundedtheory, and can claim a place in a text-book intended for use as an introduction to the present state of our knowledge of General Chemistry.

Then he added a comment that seems to say that his rejection of atomism was fully justified beforethe experimental evidence cited was available: “From the point of view of stoichiometry the atomic theory is merekly a convenient mode of representation, for the facts, as is well known, can be equally well, and perhaps better, represented without the aid of the atomic conception as usually advanced.”17

The paradox connected with the timing of this turnabout by Ostwald is revealed by the fact that during the first decade of the twentieth century, the “billiard ball” model of the kinetic theory of gases and of the Daltonian atom was rapidly being abandoned in favor of an internally structured atom of considerable complexity. To use an appropriate analogy that had been employed thirty years earlier by W.K.Clifford, “an atom must be at least as complex as a grand piano.”18

By the turn of the century, Ostwald’s position on atomism and energetic was relatively obsolete among physicists and most chemists. Scientists and philosophers alike-idealists and materialists-thoroughly opposed his energetics and rejected his suggestion that the concept of matter (conceived by Ostlwald as the concept of substance) be subordinated to the concept of energy. Nevertheless, it is fair to add that two elements of his energetics have endured the test of time: his formulation of the second law of thermodynamics as the impossibility of a perpetuum mobile of the second kind, and his early insistence on the necessity of employing the free eneregy functions of Gibbs and Helmholtz, instead of heats of reactions, as a criterion for the feasibility of chemical spontaneity and as a measure of the equilibrium position of chemical reactions.

During the last decades of his life at Grossbothen, Ostwald worked with great determination on the development of color science. Starting from color standardization, which around 1912 was a question of topical interest, he systematically investigated colors, developed a quantitative color theory, and produced color samples and coloring substances in his laboratory. He thereby gave a new impetus to this previously neglected field of applied chemistry, and he considered the problems in color science among the most important that he was able to solve. The way in which Ostwald carried out his research on color is a perfect example of a well–organized and carefully thought–out undertaing. Among the achromatic colors he included the white-gray-black continuum. Following the determination of pure gray, he ascertained the gray scale–the standard achromatic scale required for the measurement of chromatic colors–with the aid of a split–field photometer of his own construction. In this work Ostwald introduced new concepts: saturated colors, unsaturated colors, and completely saturated colors (Vollfarbe). For unsaturated chromatic colors he used the relationship v+w+s=1, and for achromatic colors v+s=1, where v represents vollfarbe; w, weiss (white); and s, schwarz (black). Postulating that “the colors of bodies are fundamentally mixtures of light,”19 Ostwald introduced the term Farbenhalb for a mixture containing half of all of the wavelengths of the visible spectrum. For measuring the colors of bodies, he employed triangles of equal color tone and colro–tone circles. In order to determine Ostwald developed a polarization color mixer. As indicator numbers for standardizing the bright colors he used the appropriate numbers of the hundred–part color circle and the percentage of the white or black content of the pigments. Thus the “world of color was subordinated to the mastery of measurement and number”20 In order to achieve the standardization of the color circles and colored bodies (Farbkerisel), Ostwald published an atlas of color (1917) and an atlas of color standards (1920).

From color standardization Ostwald turned to color harmony. Because he chose his color scales in accord with the characteristics of human perception–that is, they were logarithmically graded–he was able to construct harmonies, as in music. The gray ladder corresponded to the somber colors; varying the color tone, or the white and black content, corresponded to the bright harmonies. Ostwald enthusiastically set forth his views in Harmonie der Farben (1918), Die Farbe (1921-1926), and Harmonie der Formen (1922). His ideas were criticized and even rejected, especially in artistic circles; but they also found widespread acceptance. Through his color standards and color harmonies Ostwald gave a new and far-reaching impetus to the construction of colors according to a deliberately planned method. In Germany the application of his system was confined mainly to Saxony and to the Bauhaus in Dessau. In Great Britain, on the other hand, it was widely endorsed and was taught in the schools. In the United States his system found important advocates in E. Jacobson and H. Zeishold, through whose influence the system, as presented in Color Harmony Manual, became the accepted colors standard.21

During the last thirty years of his life, Ostwald spoke and wrote in a grandiloquent style in support of humanistic, educational, and cultural causes. In 1909 he published Grosse Männer and classified persons of genius into two broad types according to mental temperament: classicists and romanticists. The laws governing their careers were formulated with reference to mental reaction velocity. Classicists were said to be phlegmatic and melancholic and to have a low reaction velocity, while romanticists, like himself, were sanguine and choleric and had a high reaction velocity.

Ostwald believed that mutual understanding among scholars was indispensable from a humanistic standpoint. In his Forderung des Tages (1910), dedicated to Arrhenius, Ostwald, the doyen of the international brotherhood of science, integrated his views on energetics with scientific methodology and systematics, psychology, scientific genius, general cultural problems, public instruction in the sciences, and the introduction of an international language. While at Harvard he had studied Esperanto; and later he created his own artificial language, Ido.

Ostwald had a mania for reform movements. His Der energetische Imperativ (1912) was a rousing, prophetic declaration of the urgency for man to adopt internationalism, pacifism, and asystematic plan for the preservation of natural energy resources. The imperative was “Squander no energy. Utilize it.” In a similar vein, Die Philosophie der Werte (1913) was given over largely to a discussion of the second law of thermodynamics, its history, applications, and prognostic comments.

In various works Ostwald gave mathematical formulas for happiness (G represents Glück). In one of these, G=K(A-W)(A+W), A represents the expended energy that is welcomed by the will; W represents the expended energy that corresponds to disagreeable experiences associated with resistance; and k is the factor for transforming the energetic into the psychological process.22

Ostwald actively participated in the congresses of the international peace movement (1909-1911) and condemned was as a “squandering of energy of the very worst kind.”23 His support of all scientific efforts led him to join the Deutsche Monistenbund, a civic society that propagated a world view based on science. At the request of Ernst Haeckel, Ostwald was president of this organization from 1910 to 1914. His perception of the individual as a cell in the collective organism of humanity is spelled out with much enthusism in his Monistische Sontagespredigten (1911-1913). His 1913 lecture delivered at Viennna—Monism as the Goal of Civilization -was directed to the organization of “the partisans of Monistic ideas and principles in national Societies in order toformulate one International Organization of all Monists in the whole world.”24 The Monistenbund fell apart after the outbreak of World War I.

Ostwald was active in learned societies and served on commissions. He was a member of the International Commission on Atomic Weights and cofounder and temporary president of the International Association of Chemical Societies. In addition, for many years he was a member of the boards of directors of German chemical societies. He also showed a deep concern with the training of chemists and was an enthusiastic defender and advocate of scientifically oriented secondary school education. This many–sided scientific and organizational acitivity earned Ostwald, besides the Nobel Prize many honoary doctorates and honorary memberships in learned societies.

Ostwald strongly advocated the study of the history of science and frequently used historical materials in both his scientific and his more philosophical writings. When Isis. Revue consacrée a l’histoire de la science was launched by George Sarton in March 1913, Ostwald’s name was among the Comité de patronage His most important single work devoted to the history of science was Elektrochemie. Ihre Geschichte und ihre Lehre (1896), a book of more than 1,100 pages that exhaibits Ostawald’s complete command of the scientific literature on electrochemistry and allied areas. It was, however, the only one of Ostwald’s major books that was not published in a second edition; neither was it translated. The monumental enterprise known as Ostwalds Klassiker der Exakten Wissenschaften began in 1889 with Helmholtz’ 1847 work “Uber die Erthaltung der Kraft” 243 volumes had been published by 1938, and 256 volumes by 1977.

In his later years it became clear that Ostwald’s deep interest in the history of science was motivated by the belief that man had much to learn from his predecessors about the economical and systematic solution to current problems. He correctly perceived that the most suitable way to realize that goal was through the organization of scientific work; he wished to avoid “energy waste.” Ostwald therefore emphasized “organizational activity, which is the great task of the twentieth century” He stated (1926) that “fundamentally, in the present circumstances, I must consider the organizer as more important than the discover.”25 In this regard, as in many others, Ostwald was ahead of his time.

In his obituary for Ostwald, Frederik G. Donnan wrote: “It was a rich, full and successful life, in which he endeavoured to make the best use of the abundant energy accorded to him.”26 Many of Ostwald’s ideas about the sciences, art, society, and culture are no longer fashionable. His organizational efforts and philosophical generalizations may no longer be appreciated. Nevertheless, he is regarded as an extremely prolific, colorful, and influential early systematizer and spokesman for the new discipline of physical chemistry.

The American physical chemist Wilder Bancroft, who received his doctorate under Ostwald in 1892, and who was one of the most critical of his students, wrote in 1933:

We can distinguish three roups of scientific men. In the first and very small group we have the men who discover fundamental relations. Among these are van’t Hoff, Arrhenius and Nernst. In the second group we have the men who do not make the great discovery but who see the importance and bearing of it, and who preach the gospel to the heathen. Ostwald stands absolutely at the head of this group. The last group contains the rest of us, the men who have to have things explained to us.… Ostwald was a great protagonist and an inspring teacher. He had the gift of saying the right thing in the right way. When we consider the development of chemistry as a whole, Ostwald’s name like Abou ben Adhme’s leads all the rest.… Ostwald was absolutely the right man in the right place. He was loved and followed by more people than any chemist of our time.”27

Notes

1. Wilhelm Ostwald, Das physikalisch-chemische Institut der Universität Leipzig und die Feier Siner Eröffnung am 3. Januar 1898 (Leipzig, 1898).

2. Ostwald had some strong connections with America. The American Chemical Society elected him an honorary member in 1900, Jacques Loeb invited him to Berkeley, where he delivered a lecture: “The Relations of Biology and the Neighboring Sciences,” in University of California Publicationsin psychology,1 no. 4 (1903); a German version was published in Abhandlungen und Vorträge auf dem Gebiet der mathematik, Naturwissenschaften und Technik (1916), 282-307. In 1904 he was a principal speaker at the International Congress of Arts and Sciences in St. Louis, delivering a lecture before the section on the methodology of science: “On the Theory of Science,” in Congress of Arts and Sciences,1 (1905), 333-352. At Harvard his closest associates were William James, Josiah Royce, Hugo Münsterberg, Persident Charles W. Eilot, and T. W. Richards, who had been his student at Leipzig a decade earlier.

3. Ostwald’s estate was given to the German Academy of Sciences at Berlin, and is known as the WIlliam-Ostwald-Archiv Gross Bothen, Aussenstelle des Archivs der Akademie der Wissenschaften der Deutschen Demokratischen Republik. The research on color science has been continued at Gross Bothen by an industrial research laboratory.

4. M.M. Pattison Muir, “Chemical Affinity,” in Philosophical Magazine, 5th ser., 8 (1879), 181-203. For an exhaustive report and evaluation of Ostwald’s chemical affinity studies. see the chapter “Chemical Change” in Pattison Muir’s A Treatise on the Principles of Chemistry (Edinburgh, 1889).

5. “Recherches sur la conductibilité galvanique des électrolytes,” in Bihang till K. Svenska vetenskapsakademiens handlingar,8 (1884), nos. 13-14, reprinted in Ostwalds Klassiker, no 160 (Leipzig, 1907). For further references see Aus dem wissenschaftlichen Briefwechsel Wilhelm Ostwalds, 11, 3-14, 357.

6. Ostwald’s American students included Wilder D. Bandcroft, S.L. Bigelow, Edgar Buckingham, G.W.Coggeshall, Frederick G.cottrell, colin G. Fink, H. M. Goodwin, William J. Hall, G. A. Hulett, Harry C. Jones Louis Kahlenberg, F. B. Kenrick, Arthur B. Lamb, Morris Loeb, J. W. McBain, W. Lash Miller, James L. R. Morgan, Arthur A. Noyes, Theodore W. Richards, G. Victor Sammett, E. C. Sullivan, O.F. Tower, J. E. Trevor, A. J. Wakeman, and Willis R. Whitney.

7.Zeitschrift für physikalische Chemie,8 (1891), 567.

8.Ibid.,15 (1894), 706.

9. See Ostwald’s Lebenslinien,II 287-299, and III, 343.

10.Zeitschrift für Physikalische Chemie,9 (1892), 771. Also see Aus dem wissenschaftlichen Briefwechsel…,I xviii-xxi, 9-23.

11. Michaelis Teich, “Der Energetismus bei Wilhelm Ostwald und Franstišek Wald,” in Naturwissenschaften, supp. entitled Tradition, Fortschritt–Beiheft zur Zeitschrift für Geschichte der Naturwissenschaften, Technik und Medizin (1963), 147-153.

12. Erwin Hiebert, “The Energetics Controversy and the New Thermodynamics,” in Duane H. D. Roller, ed., Perspectives in the History of Science and Technology (Norman, Okla., 1971), 67-86.

13. The circumstances of this dedication and information about Ostwald’s plans for the Annalen der Naturphilosophie are given in four letters from Ostwald to Mach dated 31 May to 28 Oct. 1901, at the Ernst-Mach-Institut, Freibur im Breisgau. Also see J. Thiele, “‘Naturphilosophie’ und ‘Monismus’ um 1900. (Briefe von Wilhelm Ostwald, Ernst Mach, Ernst Haeckel und Hans Driesch),” in Philosophia naturalis,10 (1968), 295-315.

14. Wilhelm Ostwald, “Faraday Lecture,” in Journal of the Chemical Society,85 (1904), 506-522.

15.Ibid., 515-517.

16. E. Rutherford and F. Soddy, “The Cause and Nature of Radioactivity,” in Philosophical Magazine, 6th ser., 4 (1902), 370-396; quotation on 396.

17. This work, of which there were 8 eds., was first published in 1889. The quotation is reproduced from the preface to the 3rd English ed., dated Nov. 1908, Gross Bothen, translated by W. W. Taylor, Outlines of General Chemistry (1912), vi.

18. Oliver Lodge, Atoms and Rays (London, 1924), 74.

19. Ostwald, Farbenlehre,II, 118-119.

20. Ostwald, Lebenslinien,II, 392.

21. See Ostwald, The Color Primer (1969), foreword by Faber Birren, 5-6; and (on Ostwald’s color system) E. Jacobson, Basic Color (Chicago, 1948). Color Harmony Manual is an atlas of colors; it consists of 900 removable color chips in a magnificent portfolio created in 1942 by Egbert Jacobson, who was art director for the Container Corportion of America, Chicago.

22. Ostwald, Lebenslinien,III 3-6.

23.Ibid.,329.

24. Ostwald, Monism as the Goal of Civilization, International Committee of Monism, ed., (Hamburg, 1913), 3.

25. Ostwald, LebenslinienIII 435. Also see H.-G. Korber, “Einige Gadanken Willhelm Ostwalds zur Organisation der Wissenchaft. Nach einem unveröffentlichten Manuskript ausgewahlt,” in Forschungen und Fortschritte,31 (1957). 97-103, with references on Ostwald’s organization work.

26. Frederik G. Donnan, “Ostwald Memorial Lecture,” in Journal of the Chemical Society (1933), 332.

27. Wilder D. Bancroft, “Wilhelm Ostwald. The Great Protagonist,” in Journal of Chemical Education,10 (1933), 539-542. 609-613; quotation on 612.

BIBLIOGRAPHY

I. Original Works. Ostwald left an immense literary-scientific work that consists of 45 books, about 500 scientific papers, 5,000 reviews, and the edition of six journals, particularly Zeitschrift für Physikalische Chemie. His papers on physical chemistry include his master’s thesis, Volumchemische Studien über Affinität (Dorpat, 1877); and his doctoral dissertation, Volumchemische und optisch–chemische Studien(Dorpat, 1878), both repr. in Ostwalds Klassiker der exakten Wissenschaften, no.250 (Leipzig, 1966), with an introductory essay on Ostwald’s work by Gerhard Harig and Irene Sturbe.

Numerous other papers include “Volumchemische Studien,” in Annalen der physik un Chemie, supp. 8 (1876), 154-168, and n.s. 2 (1877), 429-454, 671-672; “Chemische Affinitatsbestimmungen,” “Kalorimetrische Studien ,” “Studien, zur chemischen Dynamic,” and “Elektrochemische Studien,” all in Journal für praktische Chemie,18 (1878)-33 (1886); “Das Verdunnungsgesetz,” ibid.,31 (1885), 433-462; “Über den Einfluss der Zusammensetzung und Konstitution der Säuren auf ihre Leitfähigkeit,” ibid.,32 (1885),300-374; “Über die Dissoziationstheorie der Elektrolyte,” in Zeitschrift für physikalische Chemie, 2(1888), 120-130, written with W. Nernst; “Über Autokatalyse,” in Berichte über die Verhandlungen der Sächsischen Akademie der Wissenschaften zu Leipzig,42 (1891), 190-192l; “Studien zur Energetik,” ibid.,43 (1891), 272-287, and 44 (1892), 211-237, also in Zeitschrift für physikalische Chemie,9 (1892), 563-578, and 10 (1892), 363-386; “die Dissoziation des wassers,” ibid.,11 (1893), 521-528; “Über physico–chemische Messmethoden,” ibid.,17 (1895), 427-445; “Über Katalyse,” in Verhandlungender Gesellschaft deuscher Naturforscher und Arzte,73 (1901), 184-201, alsoseparated ed. (Leipizig. 1902) and in Physikalische Zeitschrioft,3 (1902), 313-323, and nature,65 (1902), 522-526: and “Übern Katalyse. Nobelpreisvortrag, gethlten in Scockholm am 12. Dezember 1909,” in LEs Prix Nobel en 1909 (Stockholm, 1910), 63-88.

Ostwald’s books on physical chemistry include Lehrbuch der allgemeinen Chemie 2 vols. (Leipizg, 1885-1887; 2nd ed., vols. in 3 pts., 1891-1902); Grundriss der allgemeinen Chemie (Leipzig, 1889), lst English ed., Outlines of General Chemistry, tyransalted by James Walker (London, 1890); Hand-und Hilfsbuch zur Ausführung physiko-chemischer Messungen (Leipizg, 1893), lst English ed., mannual of Physico-Chemical Measurements, translated by James Walker (London, 18940: Die wissenschaftlichen Grundalgen der analytischen Chemie (Leipzig, 1894), lst English ed., The Scientific Foundations of Analytical Chemistry, translated by G. McGowan (London-New York, 1895); Grundlinien der anorganischen Chemie (Leipzig, 1900), lst English ed., The Priciples of Inorganic Chemistry, translated by A. Findlay (London-New York, 1902); Schule der Chemie, 2 vols. (Leipizg, 1903-1904),lst English ed., Conversations on Chemistry, First Steps in Chemistry, vol. I translated by E. C. Ramsay (New York-London,1905), vol. II translated by S.K. Turnbull (New York-London, 1906).

Ostwald’s works on other scientific problems include Vorlesungen über naturphilosophie (Leipizg, 1902); Abhandlungen und Vorräge allgemeinen Inhalts n(Leipzig, 1904); Grundriss der Naturphilosophie (Leipzig, 1908); Die Farbenibel (Leipizg, 1916); Farbatlas (Leipizg-Gross Bothen, 1918), lst English ed., The Ostwald Colour Album (London, 1932); Die Farblehre: I Mathematische farblehre (Leipizg, 1918);11, Physikalische Farblehre (Leipzig, 1919); III, Chemische Farblehre, with E. Ristenpart; IV, Physiologische Farblehre, by h. Poodsta (Leipizg, 1922), for which Ostwald wriote only the introduction; V, “Psychologische Farblehre,” is unpublished (some MS chapters are preserved in the Wilhelm-Ostwald-Archives)-English ed., ColourScience, translated by J. Scott Taylor, 2 vols. (London, 1931-1933); “Grundsätzliches zur messenden Farbenlehre,” in Sitzungaberichte der Preussischen Akademie der Wissenschaften zu Berlin, math, -phys, Kl., 22 (1929), 14-26, and 30 (1937), 402-416; “Attribute der Farben,” ibid.,30 (1937), 423-436; and The Color Primer. A Basic Treatise o the Color System of Wilhelm Ostwald, editied and with a foreord and evalution by Faber Birren (New york, 1969), with a short bibliography ofOstwald’s color papers in English.

Ostwald’s books and papers on the history of science include Elektrochemie. Ihre Geschichte und ihre Lehre (Leipizg, 1896): Ältere Geschichte der Lehre von den Kontakwirkungen. Dekanatsschrift (Leipizg, 1898): Leitlinien der Chemie (Leipizg, 1906), 2nd ed. entitled Der Werdegang einer Wissenschaft (Leipizg,1908): “Psychographische Studien,” in nnalen der Naturphilosophie,6-8 (1907-1909): Grosse Manner, Studien zur Biologie des Genius (Leipzig, 1909); “ChemischeWeltliteratur,” in Zeitschrift für Physikalische Chemie,76 (1911), 1-20: Aug, Comte und sein Werk (Leipzig, 1913); “Geschichtswisseschaft und Wissenschaftsgeschichte,” in Archiv für Geschichte der Mathematik, der Naturwissenschaften und der Technik,10 (1927-1928), 1-11; and DiePyramide der Wissenschaften (Stuttgart-Berlin, 1929).

Autobiographic works are “Wilhelm Ostwald,” in philosophie der Gegenwart in Selstdarstellungen, IV (Leipzig, 1924), 127-161; Lebenslinien, 3 vols. (Berlin, 1926-1927). His scientific correspondence is in Aus dem wissens chaftlichen Briefwechsel Wilhelm Ostwalds, H. G. Körber, ed., 2vols. (Berlin, 1961-1969).

Bibliographies (Listed chronologically) are P. Walden, “Scriften von Wilhelm Ostwald,” in Zeitschriftfür physikalische Chemie (Jubelband für wilhelm OStwald), 46 (1903), xvi-xxvii; G. Ostwald. Schriften zur Farblechre (Leipzig, 1936): and Gesamtregister der Abhandlungen, Sitzungsberichte… der Preussischen Akademie der Wissenschaften 1900-1945 (Berlin, 1966), 128-129 (Ostwald’s academie papers only). Alsoconsult Poggendorff, III, 991; IV, 1101-1103; V. 929-930; VI, 1928-1929; and VIIa, supp., 476-482, the most comprehensive published secondary bibliography on Ostwald.

II. Secondary Literature. Se the following listed chronogically: J. H. van’t Hoff, “Friedrich wilhelm Ostwald,” in Zeitschrift für Physikalische Chemie (Jubelband für Wilhelm Ostwald), 46 (1903), v-xv; P. Walden, Wilhelm ostwald (Leipzig, 1904), with bibliography; “Wilhelm Ostwald. Leitlinien aus seinem Leben zu seinen 60. Geburtstag gesammelt.” in Grosse Manner. studien zur Biologie des Genies. IV (Leipzig, 1913); Wilhelm Ostwald. Ferstschrift aus Anlass seines 60. Geburstages. 2, Septmeber 1913, Monistyenbund in Osterreich, ed. (Vienna-Leipzig, 1913), with bibliography: E. Haeckel, “wilhelm Ostwald. President of the Monistic League,” in Open Court,28 no. 2 (1914), 97–102: H. Freuindlich, “Wilhelm Ostwald zum 70. Geburtsatge,” in Naturwissenschaften,11 (1923), 731–732; M. Le Blanc, “wilhelm Ostwald,” in Forschungen und Forschritte,8 (1932), 174–175; W. Nernst “Wilhelm Ostwald,” in Zeitschrift für Elektrochemie,38 (1932), 337–341; P. Walden, “Wilhelm Ostwald,” in Berichte der Deutschen chemischen Gesellschaft, Abt,. A, 65 (1932), 101-141: R. Luther, “Nachruf auf wilhelm Ostwald,” in Berichte, Sachsischen Akademie der Wissenschaften, math-phys, Kl., 85 (1933), 57–71 (sess. of Nov, 1932): F. G. Donnan, “Ostwals Memorial Lecture,” in Journal of the Chemical Society (1933), 316-332; Geret Ostwald, Wilhelm Ostwald. Mein vater (stuttgart, 1953): N. I. Rodnyi and Y.l. soloviev, Vilgelm Ostvald (Moscow, 1969): and Christakirsten and Hans-Günther Körber, eds., Physiker über physiker, wahlvorschläge zur Aufnahme von Physikernin die Berliner Akademie, 1870-1929 (Beerlin, 1975), 167-168.

Erwin N. Hiebert
Hans-GÜnther KÖrber

Ostwald, Friedrich Wilhelm

views updated Jun 08 2018

OSTWALD, FRIEDRICH WILHELM

(b. Riga, Latvia, Russia, 2 September 1853; d. Leipzig, Germany, 4 April 1932)

physical chemistry, energetics, color science. For the original article on Ostwald see DSB, vol. 15, Supplement I.

Ostwald was one of the founders, along with Svante Arrhenius and Jacobus Henricus van't Hoff, of the hybrid discipline of physical chemistry and was its most effective organizer and influential spokesman. He was instrumental in winning acceptance for Arrhenius’s ionic theory of dissociation. In 1909 Ostwald received the Nobel Prize in Chemistry for his “works on catalysis, as well as for fundamental investigations of chemical equilibrium and reaction velocities” (from the Nobel citation). Ostwald was an imaginative and prolific, although often controversial, apostle of scientific ideas. Beginning in the early 1890s he championed the cause of energetics, first as an energy-based natural science but after around 1900 as a comprehensive worldview. He also championed a variety of other causes, including a novel quantitative theory of colors. Ebullient in personality, Ostwald was a man of great personal charm and an inspiring teacher. Many-sided in his interests—he was a musician and painter as well as a scientist, cultural critic, international organizer, and would-be philosopher—he was actively involved in promoting the broadly humanistic aspirations (e.g., monism, pacifism, and internationalism) shared by many European intellectuals in the first third of the twentieth century.

Biography, Correspondence, Nachlass . There is still no authoritative scientific biography of Ostwald, likely because he left an enormous published output consisting of forty-five books, more than five hundred scientific papers, and some five thousand reviews. In addition he edited, at one time or another, six journals, including the Zeitschrift für physikalische Chemie (1887–1906) and the Annalen der Naturphilosophie (1901–1914) as well as a series of scientific reprints, Klassiker der exakten Wissenschaften, to which he appended often revealing introductions and editorial remarks. There is also an enormous Nachlass, most of it currently housed in the Berlin-Brandenburgische Akademie der Wissenschaften, only parts of which have been explored. The best account of Ostwald’s life and work remains his own Lebenslinien: Eine Selbstbiographie in three volumes (1926–1927). A condensed version is “Wilhelm Ostwald” in Philosophie der Gegenwart in Selbstdarstellungen (1924). Wilhelm Ostwald: Mein Vater (1953), written by Ostwald’s daughter Grete, is still useful for its account of his family life. Subsequent works include an uneven miscellanea of twenty-five essays on his life and work collected in the proceedings of an international symposium celebrating the 125th anniversary of Ostwald’s birth (1979); Jan-Peter Domschke and Peter Lewandrowski’s Wilhelm Ostwald: Chemiker, Wissenschaftstheoriker, Organisator (1982); Armin Meisel’s “Wilhelm Ostwald: Leben und Werk” (1986); Domschke and Karl Hensel’s Wilhelm Ostwald: Eine Kurzbiographie (2000), and Mi Gyung Kim’s “Wilhelm Ostwald” (2006), a brief but useful essay in English.

More of Ostwald’s correspondence with scientific contemporaries has become available to complement the invaluable collections compiled by Hans-Günther Körber in the 1960s. The most important volumes from Berlin are three (1994, 1996, 1997) edited by Regine Zott, probably the foremost German expert on Ostwald’s work, and a fourth volume (1998) edited by Joachim Stocklöv.

Additional parts of the Nachlass have also been published by the Wilhelm-Ostwald-Gesellschaft zu Gross-bothen e.V. in a series of Sonderhefte (special issues) of the society. The most important of these are more correspondence editions (1997–2002) with scientific contemporaries, all of them chemists, and various works on Ostwald’s color theory (1999–2002). Unfortunately, the society ceased publication in February 2005, when its government funding was not renewed and the bulk of the Nachlass was transported to the Berlin Academy. A brief description of the contents of the Nachlass is provided by Regine Zott in “Wilhelm Ostwald und sein schriftlicher Nachlass” (1989). A sampling of the contents is given in Uwe Niedersen’s “Leben, Wissenschaft, Klassifikation: Aus dem Nachlass Wilhelm Ostwalds” (1992).

Physical Chemistry, Energetics, Color Theory . The work in physical chemistry that made Ostwald’s reputation is ably described in Robert Root-Bernstein’s “The Ionists: Founding Physical Chemistry, 1872–1890” (1980). The work of Ostwald’s many American students is surveyed in John Servos’s Physical Chemistry from Ostwald to Pauling: The Making of a Science in America (1990). Ostwald’s program for energetics is recounted in a series of essays (2007, 2008) by Robert J. Deltete, and a condensed version of Ostwald’s energetic theory is contained in “Gibbs and the Energeticists” (1996) by the same author. The application of energetic theory to chemistry was described in 1986 by Arie Leegwater. Ostwald’s peculiar understanding of

irreversibility is examined by Uwe Niedersen in “Die Energetik und der Irreversibilitätsgedanke bei Wilhelm Ostwald” (1983) and “Zu den Problemen von Reversibilität, Irreversibilität und Zeit im Schaffen Wilhelm Ostwalds” (1986). Tensions in his understanding of time are revealed by Hans-Jürgen Krug and Ludwig Pohlmann in “Die Dichotomien der Zeit: Der Zeitbegriff bei Wilhelm Ostwald” (1994). Ostwald’s introduction and use of the concepts mole and molar and their relation to his alleged antiatomism is discussed in two 1982 essays by Yves Noël. The received view that Ostwald was opposed to atomism until his official “conversion” in 1909 is challenged by Britta Görs in Chemischer Atomismus (1999). Casper Hakfoort (1992) offers an appraisal of Ostwald’s hopes for an energetic worldview, which revives criticism already leveled at Ostwald in his own time by Max Weber in “Energetische Kulturtheorien” (1909). A good overview of Ostwald’s color theory is Eugen Ristenpart’s Die Ostwaldsche Farbenlehre und ihre Nutzen (2000).

Organization, Standardization, Internationalism . Ostwald’s pyramidal theory of the structure of science and his ideas about proper scientific organization are contained in two Sonderhefte of the Wilhelm Ostwald Society: Wissenschaftstheorie und -organization (2004) and Das grosse Elixier: Die Wissenschaftslehre (2004), both of which are critiqued by Friedemann Schmithals in “Abstrakte Wissenschaft oder gute Lehre?” (1999). Ostwald’s recommendations for the organization of information and the standardization of documentation are discussed in two useful essays (1991, 1997) by Thomas Hapke. Ostwald’s advocacy of an international language is vigorously presented to often skeptical colleagues in another Sonderheft from the Wilhelm-Ostwald-Gesellschaft: Aus dem Briefwechsel William Ostwalds zur Einführung einer Weltsprache (1999). His activity on behalf of the German Monist League is ably recounted by Danuta Sobczynska and Ewa Czerwinska (1998), and his attempt to write a “chemical history of culture” is described by Uwe Niedersen in “Chemische Kulturgeschichte: Grundlegung” (1992).

Ostwald’s Grosse Männer (1910) classified scientists of genius into two broad types, classicists and romanticists, according to mental temperament and “reaction velocity.” This classification was described and criticized by Hans Simmer in 1978 and is employed by Robert J. Deltete to illuminate the differences in scientific style between Gibbs and Ostwald in “Josiah Willard Gibbs and Wilhelm Ostwald: A Contrast in Scientific Style” (1996).

The publishers of Ostwald’s Klassiker celebrated its centenary in 1989 with a special volume describing the history of the project by Lothar Dunsch. By 2005, 297 volumes had appeared. Six essays by Ostwald on the environment were published as volume 257 in 1978, and four essays on the history of science were published as volume 267 in 1985, all from the Nachlass.

Updates and Correction . Two additional updates to the original DSB entry are an essay by Jindrich Pinkava, “The Relations of W. Ostwald and F. Wald” (1977) and an English translation of Ostwald’s Elektrochemie: Ihre Geschichte und Lehre as Electrochemistry: History and Theory, 2 vols. (1980). A correction concerns Georg Helm, Ostwald’s uneasy ally in promoting energetics. Helm was not a physical chemist, although he wrote a textbook on the subject, Gründzuge der mathematischen Chemie: Energetik der chemischen Erscheinungen (1894). Instead, he was a mathematician and a mathematical physicist. See the introductory essay by Robert J. Deltete to the English translation of Helm’s Die Energetik nach ihrer geschichtlichen Entwicklung (The Historical Development of Energetics; 2000).

SUPPLEMENTARY BIBLIOGRAPHY

WORKS BY OSTWALD

Autobiography

Lebenslinien: Eine Selbstbiographie. 3 vols. Lebenslinien: Eine Selbstbiographie. Nach der Ausgabe von 1926/27 überarbeitet und kommentiert von Karl Hansel. Stuttgart, Germany: Verlag der Sächsischen Akademie der Wissenschaften zu Leipzig in Kommission bei S. Hirzel, 2003.

Correspondence (Berlin)

Aus dem wissenschaftlichen Briefwechsel Wilhelm Ostwalds. 1. Teil: Briefwechsel mit Ludwig Boltzmann, Max Planck, Georg Helm und Josiah Willard Gibbs. 2. Le Teil, France: Briefwechsel mit Svante Arrhenius und Jacobus Hendricus van't Hoff. Edited by Hans-Günther Körber. Berlin: Akademie-Verlag, 1961, 1969.

Wilhelm Ostwald und Paul Walden in ihren Briefen. Edited by Regine Zott. Berlin: ERS-Verlag, 1994.

Wilhelm Ostwald und Walther Nernst in ihren Briefen sowie in denen einiger Zeitgenossen. Edited by Regine Zott. Berlin: Engel, 1996.

Fritz Haber in seiner Korrespondenz mit Wilhelm Ostwald sowie in Briefen an Svante Arrhenius. Edited by Regine Zott. Berlin: ERS-Verlag, 1997.

Arthur Rudolf Hantzsch im Briefwechsel mit Wilhelm Ostwald. Edited by Joachim Stocklöv, Berlin: ERS-Verlag, 1998.

Crespondence (Grossbothen)

Sonderhefte der Mitteilungen der Wilhelm-Ostwald-Gesellschaft zu Grossbothen e.V.: Ernst Beckmann und Wilhelm Ostwald in ihren Briefen (Sonderheft 1, 1997); Max Le Blanc und Wilhelm Ostwald in ihren Briefen (Sonderheft 2, 1998); Theodor Paul und Wilhelm Ostwald in ihren Briefen(Sonderheft 3, 1998); Georg Bredig und Wilhelm Ostwald in ihren Briefen (Sonderheft 4, 1998); Robert Luther und Wilhelm Ostwald in ihren Briefen (Sonderheft 5, 1998); Carl Schmidt und Wilhelm Ostwald in ihren Briefen (Sonderheft 9, 2000); William Ramsay und Wilhelm Ostwald in ihren Briefen(Sonderheft 11, 2000); Svante Arrhenius und Wilhelm Ostwald in ihren Briefen (Sonderheft 15, 2002). Editors of 1-5 are Karl Hansel, Uwe Messow, and Karl Quitzsch; of 9 Roβ and Karl Hensel; of 11 David Goodall and Karl Hensel; and of 15 Karl Hensel.

Scientific Work

Gedanken zur Biosphäre: 6 Essays. Eingeleitet und mit Anmerkungen von Hermann Berg. Leipzig, Germany: Geest & Portig, 1978. Ostwald’s “Klassiker der exakten wissenschaften,” 257.

Electrochemistry: History and Theory. 2 vols. New Delhi: Amerind for the Smithsonian Institution and the National Science Foundation, 1980. English translation of Elektrochemie: Ihre Geschichte und Lehre. Leipzig, Germany: Veit & Comp., 1896.

Zur Geschichte der Wissenschaft: Vier Manuskripte as dem Nachlass von Wilhelm Ostwald. Mit einer Einführung und Anmerken von Regine Zott. Leipzig, Germany: Geest & Portig, 1985. Ostwald’s “Klassiker der exakten wissenschaften, 267.

“Chemische Kulturgeschichte: Grundlegung (1929/30).” Ausgewählt, kommentiert und hrsg. von Uwe Niedersen. Selbstorganisation 3 (1992): 287–308. Notes on a hemical account of cultural history

“Kalik oder Schönheitslehre.” Ausgewählt, kommentiert und hrsg. von Uwe Niederson. Selbstorganisation 4 (1993): 271–295. Ostwald’s physical theory of beauty in relation to his color theory.

Aus dem Briefwechsel Wilhelm Ostwalds zur Einführung einer Weltsprache (Sonderheft 6, 1999). Ostwald’s advocacy of an international language.

Die Philosophie der Farbe: Briefunterricht zur Farben- und Formenlehre (Sonderheft 13, 2002). Introduction to Ostwald’s philosophy of color.

Wissenschaftstheorie und -organization(Vorträge) (Sonderheft 19, 2004). Theories about the structure of science.

Das grosse Elixier: Die Wissenschaftslehre (Sonderheft 20, 2004). Ideas about scientific organization.

OTHER SOURCES

BIOGRAPHY

Domschke, Jan-Peter, and Karl Hensel. Wilhelm Ostwald: Eine Kurzbiographie. Mitteilungen der Wilhelm-OstwaldGesellschaft zu Grossbothen, Sonderheft 10, 2000.

Domschke, Jan-Peter, and Peter Lewandrowski. Wilhelm Ostwald: Chemiker, Wissenschaftstheoriker, Organisator. Leipzig, Germany: Urania, 1982.

Internationales Symposium anlässlich des 125: Geburtstages von Wilhelm Ostwald. Berlin: Akademie-Verlag, 1979.

Kim, Mi Gyung. “Wilhelm Ostwald.” HYLE: International Journal for the Philosophy of Chemistry 12, no. 1 (2006): 141–148.

Meisel, Armin. “Wilhelm Ostwald: Leben und Werk.” In Aufsätze zur Geschichte der Naturwissenschaften und Geographie, edited by Peter-Günther Hamann, 205–220. Vienna: Österreichische Akademie der Wissenschaften, 1986.

Nachlass

Niedersen, Uwe. “Leben, Wissenschaft, Klassifikation: Aus dem Nachlass Wilhelm Ostwalds.” Selbstorganisation 3 (1992): 277–285.

Zott, Regine. “Wilhelm Ostwald und sein schriftlicher Nachlass.” Mitteilungen Gesellschaft Deutscher Chemiker, Fachgruppe Geschichte der Chemie 2 (1989): 63–66.

Physical Chemistry

Görs, Britta. Chemischer Atomismus: Anwendung, Veränderung, Alternativen im deutschsprachigen Raum in der zweiten Hälfte des 19. ahrhunderts. Berlin: ERS-Verlag, 1999.

Kim, Mi Gyung. “Practice and Representation: Investigative Programs of Chemical Affinity in the Nineteenth Century.” PhD diss., University of California, Los Angeles, 1990.

Noel, Yves. “L’adjectif ‘molaire,’ son introduction, et l’expression de antiatomisme chez Wilhelm Ostwald.” Comptes Rendus du Congrès National des Sociétés Savantes, Section des Sciences 4 (1982): 361–372.

_____. “Les premiers temps de le mole dans l’oeuvre de Wilhelm Ostwald et ses traductions.” Comptes Rendus du Congrès National des Sociétés Savantes, Section des Sciences 4 (1982): 163–170.

Root-Bernstein, Robert. “The Ionists: Founding Physical Chemistry, 1872–1890.” PhD diss, Princeton University, 1980.

Servos, John W. Physical Chemistry from Ostwald to Pauling: The Making of a Science in America. Princeton, NJ: Princeton University Press, 1990.

Energetics

Deltete, Robert J. “Gibbs and the Energeticists.” In No Truth Except in the Details: Essays in Honor of Martin J. Klein, edited by A. J. Kox and Daniel M. Siegel, 135–169. Dordrecht, Netherlands: Kluwer, 1996.

_____. “Wilhelm Ostwald’s Energetics 1: Origins and Motivations” and “Wilhem Ostwald’s Energetics 2–3: Energetic Theory and Applications.” Foundations of Chemistry 9, 10 (2007, 2008): 3–56 (first essay).

Hakfoort, Casper. “Science Deified: Wilhelm Ostwald’s Energeticist World-view and the History of Scientism.” Annals of Science 49 (1992): 525–544.

Krug, Hans-Jürgen, and Ludwig Pohlmann. “Die Dichotomien der Zeit: Der Zeitbegriff bei Wilhelm Ostwald.” Selbstorganisation 5 (1994): 257–278.

Leegwater, Arie. “The Development of Wilhelm Ostwald’s Chemical Energetics.” Centaurus 29 (1986): 314–347.

Niedersen, Uwe. “Die Energetik und der Irreversibilitätsgedanke bei Wilhelm Ostwald.” Wissenschaftliche Zeitschrift der Humboldt Universität, Mathematisch-Naturwissenschaftliche Reihe 32 (1983): 325–329

_____. “Zu den Problemen von Reversibilität, Irreversibilität und Zeit im Schaffen Wilhelm Ostwalds.” Internationale Zeitschrift für Geschichte und Ethik der Naturwissenschaften, Technik und Medizin 23 (1986): 47–59.

Color Theory

Niedersen, Uwe. “Ästhetik und Zeit: Wilhelm Ostwald über Kunst.” Selbstorganisation 4 (1993): 251–270. Discusses Ostwald’s theory of art and beauty.

Sonderhefte der Mitteilungen der Wilhelm-Ostwald-Gesellschaft zu Grossbothen e.V.: Wilhelm Ostwald: Bibliographie zur Farbenlehre (Sonderheft 7, 1999); Die Farbenlehre Wilhelm Ostwalds: Der Farbenatlas (Sonderheft 8, 2000); Eugen Ristenpart, Die Ostwaldsche Farbenlehre und ihre Nutzen(Sonderheft 12, 2001).

Organization, Standardization, Internationalism

Hapke, Thomas. “Wilhelm Ostwald über Information und Dokumentation.” Mitteilungen Gesellschaft Deutscher Chemiker, Fachgruppe Geschichte der Chemie 5 (1991): 47–55.

_____. “Wilhelm Ostwald und seine Initiativen zur Organisation und Standardisierung naturwissentschaftlicher Publizistik: Enzyklopädismus, nternationalismus und Taylorismus am Beginn des 20. Jahrhunderts.” In Fachschrifttum, Bibliothek und Naturwissenschaft im 19. und 20. Jahrhundert, edited by hristoph Meinel, 157–174. Wiesbaden, Germany: Harrassowitz, 1997.

Niewöhner, Friedrich. “Zum Begriff ‘Monismus’ bei Haeckel und Ostwald.” Archiv für Begriffsgeschichte 24 (1980): 123–126.

Schmithals, Fri edemann, “Abstrakte Wissenschaft Oder Gute Lehre? Der Chemiker Wilhelm Ostwald: Lehre Jenseits Einer Fragwudigen Tradition.” Jahrbuch für Universitsitatsgeschichte2 (1999): 23–37.

Sobczynska, Danuta, and Ewa Czerwinska. “Szientismus in der Praxis: Das Wirken Wilhelm Ostwalds im deutschen Monistenbund.” Philosophisches Jahrbuch 105 (1998): 178–194.

Scientific Style

Deltete, Robert J. “Josiah Willard Gibbs and Wilhelm Ostwald: A Contrast in Scientific Style.” Journal of Chemical Education73 (1996): 289–294.

Simmer, Hans H. “Ostwalds Lehre vom Romantiker und Klassiker: Eine Typologie des Wissenschaftlers.” Medizinhistorisches Journal 13 (1978): 277–296.

Klassiker

Dunsch, Lothar. Ein Fundament zum Gebäude der Wissenschaften: Einhundert Jahre Ostwalds Klassiker der exakten Wissenschaften (1889–1989). Leipzig, Germany: Geest & Portig, 1989.

Updates and Corrections

Helm, Georg. The Historical Development of Energetics. Translated by Robert J. Deltete, with an introductory essay. Dordrecht, Netherlands: Kluwer, 2000. Translation of Die Energetik nach ihrer geschichtlichen Entwicklung(1898).

Pinkava, Jindrich. “The Relations of W. Ostwald and F. Wald.” Acta Historiae Rerum Naturalium nec non Technicarum 9 (1977): 133–148.

Robert J. Deltete

Ostwald, Friedrich Wilhelm

views updated May 17 2018

Ostwald, Friedrich Wilhelm


COFOUNDER OF MODERN PHYSICAL CHEMISTRY
18531932

Friedrich Wilhelm Ostwald, born in Riga, Latvia, Russia, almost single-handedly established physical chemistry as an acknowledged academic discipline. In 1909 he was awarded the Nobel Prize in chemistry for his work on catalysis , chemical equilibria, and reaction velocities.

Ostwald graduated with a degree in chemistry from the University of Dorpat (now Tartu, Estonia) and was appointed professor of chemistry at Riga in 1881, before he moved from Russia to Germany to become chair of the physical chemistry department at the University of Leipzig in 1887. For about twenty years he made Leipzig an international center of physical chemistry: by establishing an instruction and research laboratory that attracted virtually the entire next generation of physical chemists; by editing the first journal in the field (Zeitschrift für physikalische Chemie ); and by writing numerous textbooks. In 1906 he retired from the university and devoted the rest of his life to various topics, including the history and philosophy of science, color theory, painting, the writing of textbooks and popular books about science, the international organization of science, and the formation of an artificial language for the international exchange of ideas.

Throughout his career as a chemist Ostwald followed the general approach of applying physical measurements and mathematical reasoning to chemical issues. One of his major research topics was the chemical affinities of acids and bases. To that end, he studied the points of equilibria in reaction systems where two acids in an aqueous solution compete with each other for a reaction with one base and vice versa. Because chemical analysis would have changed the equilibria, he skillfully adapted the measurement of physical properties, such as volume, refractive index, and electrical conductivity, to that problem. From his extensive data he derived for each acid and base a characteristic affinity coefficient independent of the particular acidbase reactions.

To understand different chemical affinities, Ostwald drew on a new, but then hardly accepted and not yet fully developed, theory advanced by the Swedish physical chemist Svante Arrhenius. According to this theory of electrolytic dissociation, electrolytes such as acids, bases, and salts dissociated in solution into oppositely charged ions to a certain degree, such that at infinite dilution dissociation was complete. Ostwald recognized that if all acids contained the same active ion, their specific chemical affinities must correspond to the number of these active ions in solution, which depended on their specific degree of dissociation at each concentration, and which could be measured through electric conductivity studies. By applying the law of mass action to the dissociation reaction, a simple mathematical relation was derived between the degree of dissociation α, the concentration of the acid c, and an equilibrium constant specific for each acid K :

This is Ostwald's famous dilution law from 1888, which he proved by measuring the electric conductivities of more than 200 organic acids at various concentrations. He substantiated the dissociation theory not only to explain the different activity of acids, but also as a general theory of electrolytes in solution. The theory gained further support from the Dutch physical chemist Jacobus Hendricus van't Hoff who, at the same time, advanced it on a general thermodynamic basis to explain his law of osmotic pressure of solutions as well as Raoult's laws of vapor pressure lowering and freezing point depression. Thus, the new physical chemistry grew to a comprehensive theory of solutions, based on both thermodynamics and dissociation theory.

Ostwald was particularly successful in systematizing and propagating these new ideas, applying them to other fields, and organizing a school of physical chemistry. Many chemists rejected the dissociation theory because it predicted wrong values at high concentrations and for strong electrolytes. Despite his concessions about its restricted validity, Ostwald provided numerous proofs of its broad usefulness in his textbooks on general, inorganic, and, particularly, analytical chemistry.

Originally, and incorrectly, Ostwald studied reaction velocities as a measure of chemical affinity. Later, he broadly investigated the time (or kinetic) aspects of chemical reactions and provided a system for the study of chemical kinetics. He first recognized catalysis as the change of reaction velocity by a foreign compound, which allowed him to measure catalytic activities. He distinguished catalysis from triggering and from autocatalysis, which he considered essential to biological systems. His most famous contribution to applied chemistry was on the catalytic oxidation of ammonia to nitric acid, which became widely used in the industrial production of fertilizers.

see also Acid-Base Chemistry; Arrhenius, Svante; Equilibrium; Physical Chemistry; van't Hoff, jacobus.

Joachim Schummer

Bibliography

Hiebert, Erwin N., and Körber, Hans-Günther (1978). "Ostwald, Friedrich Wilhelm." In Dictionary of Scientific Biography, Vol. XV, Supplement I, ed. Charles C. Gillispie. New York: Scribners.

Ostwald, Wilhelm (19261927). Lebenslinien: eine Selbstbiographie. 3 vols. Berlin: Klasing.

Rodnyj, N. I., and Solowjew, Ju. I. (1977). Wilhelm Ostwald. Leipzig: Teubner. (Russian original: Vilgelm Ostvald. Moscow: Nauka, 1969.)

Servos, John W. (1990). Physical Chemistry from Ostwald to Pauling: The Making of a Science in America. Princeton, NJ: Princeton University Press.

Ostwald, Friedrich Wilhelm

views updated Jun 08 2018

OSTWALD, FRIEDRICH WILHELM

(b. Riga, Latvia, Russia [now Latvian S.S.R.], 2 September 1853; d. Leipzig, Germany, 4 April 1932)

chemistry, color science.

For a detailed study of his life and work, see Supplement.

About this article

Friedrich Wilhelm Ostwald

All Sources -
Updated Aug 08 2016 About encyclopedia.com content Print Topic