Skip to main content
Log in

Impact of Gibbs’ and Duhem’s approaches to thermodynamics on the development of chemical thermodynamics

  • Published:
Archive for History of Exact Sciences Aims and scope Submit manuscript

Abstract

From 1873 to 1878, the American physicist Josiah Willard Gibbs offered to the scientific community three great articles that proved to be a milestone for the science of thermodynamics. On the other hand, between 1886 and 1896, the French physicist Pierre Maurice Marie Duhem translated thermodynamics into the language of Lagrange’s analytical mechanics. At the same time, he expanded its scope to include thermal phenomena, electromagnetic phenomena, and all kinds of irreversible processes. Duhem formulated a version of thermodynamics characterized by the conceptual unification of mechanics, physics, and chemistry. Overall, the work of both physicists on thermodynamics is tremendous, full of axioms, theorems, corollaries, proofs, and hundreds of equations. Therefore, it would be a utopian aim to provide a short analysis of their work. Instead, the present study will attempt to give a brief outline of the main tools and concepts used by the two physicists. I will argue that each scientist approaches thermodynamics in a new and unique way, which reveals their scientific styles as reflected in their personalities, the writing styles, their behavior toward publicity, and their inclination for publication. Finally, I will examine the influence of their theories on the development of chemical thermodynamics.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Notes

  1. Martin Klein’s two papers (1984) and (1989) and Lynde Wheeler’s biography of Josiah Willard Gibbs (1962) are useful references for a historical perspective of Gibbs’ thermodynamics. Klein’s article of 1989 reappeared slightly modified in Physics Today (Klein 1990, 43: 40–48).

  2. Saurel defended his dissertation in 1900 on a subject related to Gibbs’ and Duhem’s work on the equilibrium of chemical systems (Jaki 1984, pp. 136–137).

  3. Wilson used the notes of another Gibbs’ student, because he lost his notes (Wilson 1936, p. 19).

  4. It has been published as a book: Bordoni (2012). Taming complexity: Duhem's third pathway to thermodynamics. Urbino: Editrice Montefeltro.

  5. For an account of the state of thermodynamics in the second half of the nineteenth century, see Darrigol (2018a, 42–68 and b, pp. 4–11).

  6. However, August Heinrich Horstmann approached the dissociation reactions in the gas phase, in the solid and in solution through the entropy, and the concept of disgregation. He used the disgregation because he thought that it was the only factor contributing to the variation of entropy. From 1868 to 1873, Horstmann published three papers where he developed his theory of dissociation. He proposed that the degree of dissociation has to take a value for which the entropy of the system is at a maximum. He derived an expression for the equilibrium constant in terms of disgregation (Horstmann 1873; Jensen 2009; Darrigol 2018b, pp. 6–7).

  7. Martin Klein conjectures that the historical Tait–Clausius entropy dispute on the pages of the Philosophical Magazine arouse Gibbs’ interest in thermodynamics (Klein 1989, p. 6).

  8. In 1876, Andrews published a second paper with a higher degree of accuracy, in which the temperature range extended by three more series of measurements at 6.5 °C, 64 °C, and 100 °C. He did not offer any theoretical explanation for the critical point, except perhaps, some vague statements about the action of internal attractive and resistive molecular forces (Andrews 1876, pp. 448–449). It seems that Andrews was unaware of Gibbs’ graphical papers.

  9. Gibbs cited both Andrews’ and Thomson’s studies in his Graphical Method.

  10. Clausius put forward a detailed account of the reasons that make van der Waals’ equation to deviate from Andrews’ experiments. He proposed modification of van der Waals equation, which contained four constants (van der Waals’ equation contains two constants) and tested his equation with Andrews’ experimental data taken from his pressure–volume diagram of carbonic acid. This modified equation reproduced the experiments at low densities of the gas but failed at high densities near the point where pressure increases rapidly. The modified Clausius’ equation failed to reproduce Andrews’ straight-line segment. However, it predicted the existence of the metastable states of the supersaturated vapor or the superheated liquid (Clausius 1880, pp. 398–400).

  11. Gibbs divided the thermodynamic surface into two parts, of which one represents the homogeneous phases, and the other mixtures of heterogeneous phases. Gibbs called the first part primitive surface, and the second part derived surface (GSP 1906, pp. 35–36). Gibbs pointed out that none of these surfaces exists when the system is not in a state of thermodynamic equilibrium.

  12. A developable surface is the type of surface that a tangent plane creates. The surface generated by a straight line or by a straight line tangential to a space curve is a ruled surface. A developable surface could be a ruled surface, but the opposite does not hold. Gibbs described the surfaces obtained from the rolling planes as developable surfaces. However, these surfaces are in fact ruled surfaces spanned by lines joining corresponding points of common tangent planes.

  13. This triangle is similar to that presented in Gibbs’ first graphical paper and discussed above.

  14. A diathermal wall between two thermodynamic systems allows heat transfer from one system to the other, but does not allow transfer of matter across it. A diathermal envelope encloses a closed system.

  15. William Thomson introduced the concept of dissipated energy in his attempt to formulate the second law of thermodynamics. In the last section of his essay, Gibbs discussed problems related to the surface of the dissipated energy.

  16. Gibbs did not use the terminology bimodal and spinodal. Maxwell describing Gibbs’ thermodynamic curve made use of the terms convexo-convex curve and spinode curve for the bimodal and the spinodal surfaces, respectively.

  17. This figure is similar to Fig. 2 of Gibbs’ article (GSP 1906, p. 44), except the identification of the various sections of the diagram for clarification. Maxwell presented a more detailed diagram in his textbook the Theory of heat (1902, p. 207).

  18. Saurel studied the condition for the tangential relationship between the bimodal and spinodal curves (1902, pp. 483–484).

  19. Duhem, who considered Gibbs an algebraist, alleged that the geometrical demonstrations of these papers did not play almost any role in the study of thermodynamic properties. They rather constitute the algebraic analysis in the form of Cartesian algebra (1908, p. 17).

  20. According to Professor Walters of Yale, who commented on Gibbs’ dissertation, Gibbs showed his predilection for mechanics and skill in the use of the geometrical approach (Wheeler 1962, p. 28).

  21. Gibbs contrasted Watt's indicator to his graphical representations to emphasize the usefulness of the proposed entropy–temperature diagram. “The method in which the coordinates represent volume and pressure has a certain advantage in the simple and elementary character of the notions upon which it is based, and its analogy with Watt's indicator has doubtless contributed to render it popular” (GSP 1906, p. 11).

  22. Nevertheless, Gibbs used as well geometrical representations in the heterogeneous substances to interpret the thermodynamic behavior of pure and multi-component systems. He recognized that the three independent variables of the first two papers were unsuitable when the system has more than three independent variables. He thought that he could overcome this obstacle upon reducing the number of variables into two or three. He used the temperature, the pressure, and the ζ function (see below) by making constant the total mass of the system equal to unity. Through this line of reasoning, he studied the conditions of equilibrium of homogeneous substances, as well as binary and ternary systems. For ternary systems, he adopted a different method of representation changing the Cartesian coordinate system into a two-dimensional equilateral triangle, or a three-dimensional prism. (GSP 1906, pp. 115–129). The first diagram is currently in use by physicists and engineers to illustrate phase transformations in multi-component systems.

  23. An isolated system is a thermodynamic system enclosed by rigid immovable walls (as for instance a Dewar vessel) through which neither mass nor energy (heat) can pass. An isolated system obeys the conservation law, that is, its total energy and mass stay constant. The isolated system is actually an idealized model system away from common experience.

  24. Gibbs never used the term chemical potential in the heterogeneous substances. Instead, he defined the plain word potential in his writings. The term chemical potential has been attributed to the American chemist Wilder Dwight Bancroft, who renamed Gibbs’ potential in a letter sent to Gibbs in 1899 (Baierlein 2001, p. 431). It appears that Gibbs accepted the new terminology tacitly, as can be deduced in his reply letter to Bancroft in May 1899 (GSP 1906, p. 425). In the current textbooks of physical chemistry, this state function refers to as chemical potential.

  25. Müller (2006) has discussed several applications of the chemical potential (Gibbs paradox, mixtures and solutions, phase rule, law of mass action, semipermeable membranes, osmosis, Raoult’s law, etc.) in the context of modern chemical thermodynamics, as well as the method for its measurement.

  26. The German chemist Wilhelm Meyerhoffer working with van’t Hoff at the University of Amsterdam used for the first time the name phase rule. He published a paper entitled Die Phasenregel und ihre Anwendungen (The phase rule and its applications) (1893).

  27. For the system of pure water with its vapor, there are two phases (liquid and vapor), then r = 2. If we treat water as a single component, n = 1, the phase rule predicts one degree of freedom. We can change either the temperature under constant pressure or the pressure under constant temperature and maintain the system in equilibrium state without changing the number of phases or the components in each phase. If, however, we consider the dissociation of water, the number of components in the liquid phase is not one, but three: neutral water molecules, hydronium ions, and hydroxyl ions. In this case, r is still two, but the number of components is now three. The phase rule predicts three degrees of freedom, which is absurd. In this case, the conditions of neutrality and the dissociation reaction should be taken into account in the phase rule equation for determining the correct number of degrees of freedom.

  28. Gibbs deduced the phase rule far more concisely by the following reasoning, "A system of r coexistent phases, each of which has the same n independently variable components is capable of n + 2 − r variations of phase. The temperature, the pressure, and the potentials for the actual components have the same values in the different phases, and the variations of these quantities are by [equation] [97] subject to as many conditions as there are different phases. Therefore, the number of independent variations of the phase of the system, will he n + 2 − r" (GSP 1906, p. 96).

  29. For informative discussions on Roozeboom experiments and theoretical solutions offered by van Der Waals, see Daub (1976, pp. 747–748), Wisniak (2003, pp. 425–427) and Duhem (1908, pp. 23–25).

  30. Duhem rejected the method of reversible cycles first introduced by Clausius on the ground that it was a lengthy and painful procedure for the study of physical and chemical processes. Alternatively, he proposed the use of the thermodynamic potential as an easier and more effective tool (see below). Contrary to Gibbs, Duhem, and Max Planck, van ‘t Hoff considered the method of reversible cycles as the most suitable to study thermodynamic systems than using abstract physical conceptions and mathematical functions, such as entropy (see below).

  31. Clifford Truesdell (1984, p. 22) stressed that in the whole heterogeneous substances he found one, and only one passage, where Gibbs hinted the state of irreversibility, when he discussed the case of an imperfect electrochemical apparatus in which Clausius’ inequality dη ≥ dQ/t for the irreversible process may hold (GSP 1906, p. 339).

  32. Pierre Duhem, who studied Gibbs' thermodynamics thoroughly, considered this work as a manifestation of thermostatics (statique chimique). Arguing Gibbs’ reference to equilibrium processes, he posed the following question: "suppose the original condition of a body is one that satisfies Gibbs' criterion of equilibrium. If its condition is then forcibly altered, will that body upon release tend to resume or at least remain near its original condition”? Duhem considered that equilibrium in nature is not realizable; it is a virtual equilibrium since any physical process presupposes the departure from the equilibrium state accompanied by dissipation (loss) of energy as in viscous fluids and deformable solids.

  33. Gibbs was slow to publish. He was not easily satisfied with his intellectual product, and some of his publications remained on the shelf for several years before becoming available to the reader. He had completed the book Elementary Principles in Statistical Mechanics since 1892. The book appeared 10 years later, even though his notes on this subject had been distributed to students attending his lectures (Klein 1989, p. 14).

  34. Clausius, and Rankine, contributed to both theoretical trends of thermodynamics. They both based the development of thermodynamics on a molecular interpretation, although using different concepts. Clausius, even when he worked in the more phenomenological register, relied on kinetic molecular intuition (free heat, disgregation, etc.). For Clausius' motivation to develop the molecular theory in thermodynamics, see Garber (1970), Darrigol (2018a, pp. 58–60).

  35. Rational mechanics is the theoretical branch of physics that deals with problems concerning complex movements of bodies, i.e., processes occurring in time, and belongs to the general context of dynamics like equilibrium in the frame of statics. It has its origin in Newton’s physics but developed in the eighteenth century by the French physicists and mathematicians Bernoulli, Euler, D’Alembert, and advanced by Lagrange.

  36. Scrutiny of these theoretical traditions may reveal a finer classification of these theoretical approaches to thermodynamics. Stefano Bordoni discerns five streams sorted according to their conceptual distance from mechanics (2013, pp. 618–619).

  37. Massieu, Professor of Mineralogy and Geology at the University of Rennes, published two founding papers in 1869, where he introduced the two functions (1869). He was able to determine several physical constants of bodies utilizing his characteristic functions. Massieu wrote on the subject in 1876 a more detailed memoir of great clarity (1876). He remained for a long time unknown working away from the scientific community. However, as Horstmann, he kept abreast of the recent developments of physics, since he had relied on the concept of entropy and adhered to the molecular theory of fluids which only began to develop (Balian 2017).

  38. Duhem included the German-speaking physicists Arthur von Oettingen, professor at the Dorpart University in Estonia, among the physicists who viewed thermodynamics in a remarkable generalized context. Stefano Bordoni (2013, pp. 635–640) gives a short account on Oettingen’s thermodynamics.

  39. For a thorough discussion of the events that took place during Duhem’s presentation, see Jaki (1984, pp. 50–53).

  40. Jules Moutier was a professor of physics and chemistry at Collège Stanislas, a state lycėe in Paris, where Duhem continued his schooling after a short period in a private school. Duhem had great respect for Moutier as a teacher and physicist and learned from him the new thermodynamics of Gibbs (Jaki pp. 29–30 and 260–263).

  41. Gibbs cited Massieu’s first paper and his functions in a footnote of the heterogeneous substances (GSP 1906, pp. 86–87).

  42. Duhem never used this terminology. I borrowed it from Gibbs' Elementary principles in statistical mechanics, where he discussed the thermodynamic analogies with rational mechanics. By the mid-twentieth century, rational thermodynamics became a school of thought developed by Clifford Truesdell, Bernard Coleman, and Walter Noll (Truesdell 1984).

  43. Carnot’s fundamental theorem deals with the maximum obtainable motive power in heat engines for a given amount of heat. Carnot provided the proof of the maximum efficiency u for a perfect engine in a lengthy footnote in his single memoir Reflexions sur la puissance motrice du feu et sur les machines propres a développer cette puissance. He demonstrated that the efficiency of a perfect engine working reversibly was at a maximum, but he did not know its value.

  44. Clausius’ theorem states that a system exchanging a quantity of heat Q with external reservoirs and undergoing a cyclic process is one that ultimately returns a system to its original state. The theorem was expressed mathematically by the so-called Clausius inequality for a cyclic process, i.e., \( {\oint }\frac{dQ}{T} \le 0 \). T is the absolute temperature of the external reservoir.

  45. van’t Hoff formulated the principle of mobile equilibrium in his treatise Études de Dynamique Chimique published in (1884). According to this principle, every equilibrium between two different conditions of matter (systems) is displaced by lowering the temperature, at constant volume, toward that system the formation of which evolves heat. He applied this principle to every possible case, both chemical and physical equilibrium (van’t Hoff 1884, pp. 161–176). Duhem discusses van’t Hoff’s law of displacement of equilibrium with temperature that marks the difference between exothermic and endothermic compounds in connection with the effect of temperature changes at the moment of equilibrium. He contrasts once again this thermodynamic law with the doctrine of the maximum work enunciated by Berthelot (Duhem 1893a, pp. 143–144).

  46. L’Évolution de la mécanique was published in seven parts in Revue Générale des Sciences. As a book, it appeared in 1903 by A. Joanin, Paris. The edition of 1905 has been used in this study. There exists an English translation, The evolution of mechanics by Sijthoof and Noordhoff published in 1980, to which I have no access. English translation of portions of chapter VII is available in Maugin’s book (2014, pp. 176–183).

  47. As the reader of l’Évolution may note, Duhem, in writing this book, has used material of several of his previous publications. Therefore, this collection may be taken as an overview of Duhem’s philosophical, historical, and scientific deliberations on rational thermodynamics.

  48. Under the name hysteresis, Duhem refers to a number of systems that suffer permanent alterations, e.g., dielectric, chemical, elastic hysteresis, or alterations imposed on metals by the effect of temperature, such as annealing, hardening, etc. (Duhem 1905, pp. 318–319).

  49. Duhem has described the theoretical analysis of these complex phenomena of deformable solids as a function of one and two variables in seven memoirs published between 1895 and 1902 under the general title Les Déformations Permanentes et l'Hystérésis (1895, 1898, 1902).

  50. Duhem calls this coefficient as the coefficient of hysteresis when he refers the phenomena of hysteresis within the context of his theory on permanent alteration (1905, p. 320).

  51. Gibbs and Duhem were honored by the Scientific Academies of their countries and abroad and by some of their colleagues, at least from those who had an understanding of their work. For Gibbs’ recognition, see Wheeler, pp. 83–93, and 97–99; for Duhem, see Jaki (1984, pp. 141–147). However, personal recognition does not mean necessarily a wide reception of their work.

  52. Gibbs distributed reprints of his heterogeneous substances to a large number of contemporary physicists, chemists, mathematicians, and astronomers that, according to him, might have some interest in his work. The extensive mailing list of the individuals that received reprints is reported in Wheeler 1962, pp. 235–248. However, most of the recipients showed characteristic indifference to his work, probably due to their difficulty in understanding his concise writing style or because Gibbs was unknown to his European colleagues. Among the few, Maxwell indicated an immediate interest in Gibbs’ thermodynamics.

  53. The release of a scientific journal does not necessarily imply that it will have a wide readership. It depends on the quality of the journal and the level of research reflected on its pages.

  54. It published in four large volumes of 1430 pages in total from 1897 to 1899.

  55. Duhem published an introduction (Duhem 1900) and a critical essay in two parts (Duhem 1901b, c) on Maxwell electromagnetic theory in Annales de la Societe Scientifique de Bruxelles. These articles were republished as a single book (Duhem 1902c), which has been translated in English.

  56. Criticism of Maxwell's classical electromagnetic theory was widespread among French physicists in Duhem's time. But it was rather shallow compared to the detailed and rigorous analysis of it given by Duhem. Nevertheless, Duhem's treatise on Maxwell’s electromagnetic theory was omitted from contemporary French reviews, articles, and monographs (Jaki 1984, pp. 283–284).

  57. Duhem, who studied thoroughly Gibbs’ work and exposed his thermodynamics among the contemporaries in France, made once the following ironic comment regarding Gibbs’ dense writing style: “Il semble parfois qu'en publiant ses travaux, Gibl)s eût été possédé du désir de les voir passer inaperçus; s'il en fut ainsi, il fut bien souvent servi à souhait; bien souvent, ses idées demeurèrent ignorées de ceux-là mêmes qui auraient eu le plus grand intérêt à les connaitre.” (It sometimes seems that by publishing his works, Gibbs would have been possessed of the desire to see them go unnoticed; if this was the case, his wishes were often fulfilled; very often, his ideas remained ignored to those who would have had the greatest interest in knowing them) (Duhem 1908, p. 14).

  58. Commenting on Gibbs’ writing style, Duhem concluded: “If therefore, Gibbs has left his discoveries of chemical mechanics in an abstract and purely algebraic form, it is not that he was incapable of presenting them in a language more concrete and more accessible to experimenters, it is because of his intellect.” (Duhem 1908, p. 26).

  59. Maxwell, van der Waals, Helmholtz, J. J. Thomson, and Hertz were among Gibbs’ contemporary physicists who showed an understanding of the heterogeneous substances. However, only Maxwell spoke enthusiastically to his contemporaries in Britain about Gibbs and included Gibbs’ graphical representations in his textbook Theory of Heat. Maxwell’s premature death in 1879, at the age of 48, put an end to his excellent work in physics, while Gibbs lost a keen supporter of his thermodynamics.

  60. For a thorough discussion of the various factors that delayed the transfer of thermodynamics to chemistry and the subsequent formulation of the chemical thermodynamics by the joint efforts of van’t Hoff, Arrhenius and Ostwald, see Dais (2019).

  61. Not only chemists but also the majority of physicists had difficulty in accepting or assimilating the concept of entropy. As noted, William Thomson never mentioned the entropy in his work, and even its creator, Clausius, did not give entropy any prominent place in his research. The founders of the discipline of physical chemistry chose to avoid any reference to entropy. The next generation of chemists was familiar with the concept of entropy, but they preferred to use alternative measures for the chemical affinity and chemical equilibrium. Gibbs made use of the chemical potential and free energy under constant pressure, Helmholtz used the free energy under constant volume, and Duhem the thermodynamic potential. Reconciliation of entropy and chemistry was achieved during the decades after the First World War, especially by the American physical chemists (Lewis, Randall, Trevor, Noyes, and others). The “tortuous” course of entropy to enter research and education is the subject of an excellent paper by Kragh and Weininger (1996).

  62. In a letter to Gibbs in 1887, Ostwald proposed the translation of the heterogeneous substances, admitted that “I cannot deny that at present the study of your work is pretty difficult, particularly for the chemist, who is usually not at home in a mathematical treatment” (quoted in Moore et al. 2002, p. 115).

  63. Characteristic is the statement of the organic chemist Friedrich Wöhler, while comparing mathematics and observation: “My imagination is fairly active, but I am somewhat slow in my thinking. No one is less oriented to be a critic than I. The organ for philosophical thought is entirely missing in me, as you know so well, just as that for mathematics. Only for observation do I imagine that I have a passable facility in my brain, which may be connected with a sort of instinct to be able to predict empirical relationships” (Rocke 1993, p. 34).

  64. The development of the ionists’ chemical thermodynamics and the foundation of physical chemistry have been expounded by several historians. See for instance Root-Bernstein (1980) and Servos (1990, pp. 20–45).

  65. Duhem was so impressed by Gibbs’ heterogeneous substances that he sat down and wrote a review on Gibbs’ work as a whole. This review was initially published in the Bulletin des Sciences mathėmatique in 1907. 1 year later, Duhem published an enriched version as a separate book (1908); see Klein (1990, p. 53).

  66. The correspondence between Gibbs and Ostwald when the latter asked Gibbs’ permission to translate the heterogeneous substances reveals some facets of Gibbs’ scientific style. For instance, he refused to write a short introduction in the German translation and delayed to provide a personal portrait for the same publication. The Ostwald–Gibbs correspondence was published by Moore et al. (2002).

  67. Helmholtz derived the equation in his 1882 famous paper Über die Thermodynamik Chemischer Vorgänge (1882) (on the thermodynamics of chemical processes) dealing with the free and the bound energy of chemical processes.

  68. In 1889, the twenty-four-year-old Nernst showed that the free energy of a chemical reaction could be measured by making the reaction the source of a galvanic cell and measuring the electrochemical potential. Nernst showed that there was a simple proportionality between the electrochemical potential and free energy. This relationship is now known as Nernst’s equation. However, this methodology had practical difficulties. For each chemical process, a separate galvanic cell should be implemented, which was not feasible for all cases. Furthermore, very dilute solutions should be used to allow the direct use of concentrations in Nernst’s equation.

  69. There are several accounts on the efforts of these scientists to solve the integration problem (Nernst 1918, pp. 1–14 and 227–231; Hiebert 1981, pp. 437–439; Cropper 1987, pp. 5–6; Coffey 2006, pp. 371–382).

  70. Several physicists in Europe and America including Duhem delivered formal proof of the phase rule.

  71. A report for Stassfurt deposits and van’t Hoff experimental findings is issued in the Data in Geochemistry. 1991. Bulletin-United States Geological Survey, issue 491, pp. 210–217. This survey did not include phase diagrams from these studies due to the limited space of the report.

  72. Bancroft research program based on the phase rule had a limited scope and finally collapsed unable to compete with the advancement of physical chemistry. In later years, the phase rule became the central research subject of engineers. Details for the bankruptcy of Bancroft’s research program have been provided by Servos (1982).

  73. Semipermeable membrane was an imaginary device, known as the equilibrium box, invented by van’t Hoff to study reversible changes of the concentrations of gas mixtures or solutions, so that chemical reactions could be performed, at least in principle, in a reversible way. This model could take into account any transformation of individual substances in a mixture or solution (van’t Hoff 1912, p. 54; Kipnis 1991, p. 215).

  74. Nernst adopted Duhem’s terminology for the potential, but he used Gibbs’ notion of the chemical potential symbolized by the same letter μ as in heterogeneous substances.

  75. To the best of my knowledge, it is unknown when and by whom the name of Duhem was added to this equation. Duhem has derived an analogous equation in his first dissertation for two components, and later for the general case, when he considered the effect of the masses of substances on his thermodynamic potential. (Duhem 1886, pp. 32–35 and 140–143). The same topic reappeared in his Traitė Ėlementaire de Mėchanique Chimique Fondėe sur la Thermodynamique (1897b, Vol. 3, pp. 1–4). See Miller 1963). Duhem considered that the thermodynamic potential Φ under constant pressure and temperature is a homogeneous function of the first degree of the masses of the various substances of the system. Using Euler’s theorem, he derived an equation analogous to Eq. (24).

  76. The Duhem–Margules equation, sometimes called Gibbs–Duhem–Margules equation, describes the relationship between the mole fraction Ni (expression of composition) with partial pressure Pi of the irth component in a liquid mixture expressed by \( \mathop \sum \nolimits_{1}^{{\text{n}}}\, {\text{N}}_{{\text{i}} }{\text{dlogP}}_{{\text{i}}} = 0 \). The integration of this equation gives Raoult’s law for the partial pressures of the mixture, whereas Henry's law results when the mole fraction of the solute goes to zero (Ni ⟶ 0). Ni = Pi. In other words, the concentration of the solute dissolved in the solvent is proportional to the partial pressure of the gas above the solution.

  77. Jaki has given a detailed account on the reception of Duhem’s work by the French community and abroad during his lifetime. Few French physicists, including Jules Tannery, Henri Poincaré, and Paul Langevin, gave support or at least were sympathetic to Duhem’s publications. Others, like Le Chatelier, Jean Perrin, Brillouin, Berthelot, and several others of lower caliber avoided citing Duhem’s name in their work (Jaki 1984, pp. 279–302).

  78. Yves-André Rocard, professor at the Sorbonne and director of the physical laboratories at the École Normale Superieure, seems to have been the first French author to refer in print to Gibbs–Duhem equation in his Thermodynamique in 1952 (Jaki 1984, p. 308).

  79. Emil Jouguet was a graduate of the École Polytechnique and from 1920 to 1939 professor at l'École des mines de Paris. He had collaborated with Duhem at Bordeaux, and thus he was able to evaluate Duhem's physics and his overall theoretical achievements. Jouguet has created the theory of detonation waves with application to high explosives continuing his mentor's efforts on this subject that considered a side effect of the false equilibrium. Jouguet credited Duhem with having laid the theory of explosives on especially solid foundations (Jaki 1984, p. 305).

  80. The rather eccentric expression of the nonsensical branches of mechanics indicates fields of physics, mechanics, and electromagnetism. The list of these fields includes the so-called false equilibria, friction, viscosity, hysteresis phenomena, and electromagnetic theory of materials. These are precisely dissipative phenomena such as thermodynamically irreversible reactions, plasticity, viscoelasticity, and memory effects.

References

  • Andrews, T. 1869. The Bakerian lecture: On the Gaseous state of matter. Philosophical Transactions of the Royal Society of London 159: 575–590.

    Google Scholar 

  • Andrews, T. 1876. The Bakerian lecture: On the Gaseous state of matter. Philosophical Transactions of the Royal Society of London 166: 421–449.

    Google Scholar 

  • Baierlein, R. 2001. The elusive chemical potential. American Journal of Physics 69: 423–434.

    Google Scholar 

  • Balian, R. 2017. François Massieu et les potentiels thermodynamiques. Comptes Rendus Physique 18: 526–530.

    Google Scholar 

  • Bancroft, W. 1899. Review: Traitė ėlementaire de mėchanique Chimique, fondėe sur la thermodynamique. Science 10: 81–82.

    Google Scholar 

  • Bordoni, S. 2010. Taming complexity. Duhem’s pathway to thermodynamics. Doctoral Dissertation. Universita Deglis Studi di Bergamo. 359 pp.

  • Bordoni, S. 2012. Unearthing a buried memory: Duhem’s third way to thermodynamics, Part 2. Centaurus 54: 232–249.

    Google Scholar 

  • Bordoni, S. 2013. Routes toward an abstract thermodynamics in the late nineteenth century. The European Physical Journal H 38: 617–660.

    Google Scholar 

  • Boynton, W.P. 1900a. Gibbs’ thermodynamical model. Physical Review 10: 228–233.

    Google Scholar 

  • Boynton, W.P. 1900b. Gibbs’ thermodynamical model for a substance following van der Waals’ equation. Physical Review 11: 291–303.

    Google Scholar 

  • Boynton, W.P. 1905. Thermodynamical potentials. Physical Review 20: 259–267.

    Google Scholar 

  • Bumstead, H.A., and R.G. Van Name. 1906. The scientific papers of Gibbs, vol. 1, 1–353. New York: Longman, Green and Co.

    Google Scholar 

  • Clark, A.L., and L. Katz. 1939. Thermodynamic surfaces of H20. Transactions of the Royal Society of Canada, 3rd series Sec. III 33: 59–71.

    Google Scholar 

  • Clausius, R. 1880. On the behavior of carbonic acid in relation to pressure volume and temperature. Philosophical Magazine 9: 303–408.

    Google Scholar 

  • Coffey, P. 2006. Chemical free energy and the third law of thermodynamics. Historical Studies in the Physical and Biological Sciences 36: 365–396.

    Google Scholar 

  • Coy, D. C. 1993. Visualizing thermodynamic stability and phase equilibrium through computer graphics. Ph.D. Dissertation, Iowa State University.

  • Cropper, W.H. 1987. Walther Nernst and the Last Law. Journal of Chemical Education 64: 1–8.

    Google Scholar 

  • Cropper, W.H. 2004. Great physicists. The life and times of leading physicists from Galileo to hawking, 106–123. Oxford: Oxford University Press.

    Google Scholar 

  • Dais, P. 2019. The double transfer of thermodynamics: from physics to chemistry and from Europe to America. Studies in History and Philosophy of Science 77: 54–63.

    Google Scholar 

  • Darrigol, O. 2018a. Atoms, mechanics, and probability: Ludwig boltzmann’s statistico-mechanical writings—An exegesis. Oxford: Oxford University Press.

    MATH  Google Scholar 

  • Darrigol, O. 2018b. The Gibbs Paradox: Early history and solutions. Entropy 20: 1–54.

    Google Scholar 

  • Daub, E.E. 1976. Gibbs’ phase rule: A centenary retrospect. Journal of Chemical Education 53: 747–751.

    Google Scholar 

  • Deltete, R.J., and A. Brenner. 2004. Review. Pierre Duhem: Mixture and chemical combination and related essays. Foundations of Chemistry 6: 203–230.

    Google Scholar 

  • Donnan, F.G., and A. Haas. 1936. Commentary on the scientific writings of J. Willard Gibbs. New Haven, Connecticut: Yale University Press.

    Google Scholar 

  • Duhem, P. 1884. Sur le potentiel thermodynamique et la théorie de la pile voltaïque. Comtes Rendus de l’Académie des Sciences 99: 1113–1115.

    Google Scholar 

  • Duhem, P. 1886. Le potentiel thermodynamique et ses applications à la mécanique et à l’étude des phénomènes électriques. Paris: Hermann A.

    Google Scholar 

  • Duhem, P. 1891. Sur les ėquations gėnėral de thermodynamique. Annales Scientifiques de l’Ecole Normale Supėrieure, 3e Sėrie 8: 231–266.

    MATH  Google Scholar 

  • Duhem, P. 1892a. Commentaire aux principes de la thermodynamique-premier Partie. Journal de Mathėmatiques Pures et Appliquėes, 4e Serie 8: 269–330.

    MATH  Google Scholar 

  • Duhem, P. 1892b. Notation atomique et hypothѐses atomistiques. Revue des questions scientifiques, 31, 391–457. Translated by Needham, P. (2000). Atomic Notation and Atomistic Hypotheses. Foundations of Chemistry 2: 127–180.

    Google Scholar 

  • Duhem, P. 1893a. Introduction a la mechanique chimique. Paris: George Carré.

    Google Scholar 

  • Duhem, P. 1893b. Commentaire aux principes de la Thermodynamique-Deuxiѐme partie. Journal de Mathėmatiques Pures et Appliquėe 4e Serie 9: 293–473.

    MATH  Google Scholar 

  • Duhem, P. 1893c. L’Ècole anglaise et les thėories physiques: Á propos d’un livre rėcent de W. Thomson. Revue des Questions Scientifiques, 34: 345–378. Edited and translated by Ariew, R., and P. Barker as Duhem, P. (1996). The English School and Physical Theories: On a Recent Book by W. Thomson. In Essays in the History and Philosophy of Science. Indianapolis: Hackett.

  • Duhem, P. 1894. Commentaire aux principes de la thermodynamique—Troisième partie: Les équations générales de la thermodynamique. Journal de Mathématiques Pures et Appliquées 10: 207–285.

    Google Scholar 

  • Duhem, P. 1895, 1898, 1902. Sur les déformations permanents et l’hystérésis. In: 4e and 5e mėmoire de l’Académie de Belgique, tome. LIV, tome. LVI, tome. LXII.

  • Duhem, P. 1896. Théorie thermodynamique de la viscosité, du frottement et des faux équilibres chimiques. Mémoires de la Société des Sciences physiques et naturelles de Bordeaux, 5e série 2: 99–126.

    MATH  Google Scholar 

  • Duhem, P. 1897a. Thermochimie à propos d’un livre récent de Marcellin Berthelot. Revue des questions scientifiques Ser 2 7: 361–392.

    Google Scholar 

  • Duhem, P. 1897b. Traité Élémentaire\ de Mécanique Chimique fondée sur la Thermodynamique Livre II; tome. I, Paris: Librairie Scientifique A. Hermann.

  • Duhem, P. 1900. Les théories électriques de J, Clerk Maxwell. Étude historique et critique. Annales de la Societe Scientifique de Bruxelles 4: 239–253.

    MATH  Google Scholar 

  • Duhem, P. 1901a. Recherches sur l’Hydrodynamique. Annales de la Faculté des Sciences eo Toulouse, Série 2 3: 379–431.

    MathSciNet  MATH  Google Scholar 

  • Duhem, P. 1901b. Les theories electriques de J. Clerk Maxwell. Premiere Partie: Les electrostatiques de Maxwell. Annales de la Societe Scientifique de Bruxelles 25: 1–90.

    Google Scholar 

  • Duhem, P. 1901c. Les theories electriques de J. Clerk Maxwell. Seconde Partie: L’electrodynamique de Maxwell. Annales de la Societe Scientifique de Bruxelles 25: 293–417.

    Google Scholar 

  • Duhem, P. 1902a. Le mixte et la combinaison chimique. Éssai sur l’évolution d’une idée. Paris, C. Naud, Editeur, 208 pp. Edited and translated, with an Introduction by Paul Needham. 2002. Mixture and Chemical Combination and Related Essays. Dordrecht: Springer.

  • Duhem, P. 1902b. Thermodynamique et Chimie, leçons élémentaires à l’usage des chimistes. Leçons XVIII. XIX et XX, Paris: Librairie Scientifique A. Hermann.

  • Duhem, P. 1902c. Les théories électriques de J. Clerk Maxwell. Étude historique et critique. Paris: A. Hermann. 228 pp. Edited and translated by Aversa, A. 2015. The Electric Theories of J. Clerk Maxwell_ A Historical and Critical Study. New York: Springer.

  • Duhem, P. 1903. Thermodynamics and chemistry. A non-mathematical treatise for chemists and students of chemistry. New York: Wiley.

    Google Scholar 

  • Duhem, P. 1905. L’évolution de la mécanique, 348. Paris: Librairie Scientifique A. Hermann.

    Google Scholar 

  • Duhem, P. 1908. Josiah-Willard Gibbs, à propos de la publication de ses mémoires scientifiques, 43. Paris: Librairie Scientifiquc, A. Hermann.

    MATH  Google Scholar 

  • Duhem, P. 1915. La science Allemande, 144. Paris: A. Hermann et Fils.

    MATH  Google Scholar 

  • Dunn, E., and R.L. Fosdic. 1980. The Morphology and Stability of Material Phases. Archive for Rational Mechanics and Analysis 74: 1–99.

    MathSciNet  Google Scholar 

  • Fisher, M.E. 1989. Phases and phase diagrams: Gibbs’s legacy today. In Proceedings of the Gibbs Symposium, ed. D.G. Caldi and G.D. Mostow, 39–72. Providence: American Mathematical Society.

    Google Scholar 

  • Garber, E.W. 1969. James Clerk Maxwell and thermodynamics. American Journal of Physics 37: 146–155.

    Google Scholar 

  • Garber, E.W. 1970. Clausius and Maxwell’s Kinetic theory of gases. Historical Studies in the Physical Sciences 2: 299–319.

    Google Scholar 

  • Gibbs, J.W. 1873a. Graphical methods in the thermodynamics of fluids. The Transactions of the Connecticut Academy of Arts and Science 2: 309–342.

    MATH  Google Scholar 

  • Gibbs, J.W. 1873b. A method of geometrical representation of the thermodynamic properties of substances by means of surfaces. The Transactions of the Connecticut Academy of Arts and Science 2: 309–342.

    MATH  Google Scholar 

  • Gibbs, J.W. 1876. On the equilibrium of heterogeneous substances. The Transactions of the Connecticut Academy of Arts and Science 3: 108–248.

    MATH  Google Scholar 

  • Gibbs, J.W. 1878. On the equilibrium of heterogeneous substances. The Transactions of the Connecticut Academy of Arts and Science 3: 343–524.

    MATH  Google Scholar 

  • Gibbs, J.W. 1879. On the vapor-densities of peroxide of nitrogen, formic acid, acetic acid, and perchloride of phosphorus. American Journal of Science, Series 3 (18): 371–387.

    MATH  Google Scholar 

  • Gibbs, J.W. 1902. Elementary principles in statistical mechanics developed with especial reference to the rational foundation of thermodynamics. New York: Dover Publications.

    MATH  Google Scholar 

  • Guggenheim, E.A. 1933. Modern-thermodynamics by the methods of Willard Gibbs. London: Methuen & Co., Ltd.

    MATH  Google Scholar 

  • Helmholtz, H. F. (1882). Die Thermodynamik Chemisher Vergänge. Sitzungsberichte der Königlich Preussischen Akademie der Wissenschaften zu Berlin, Erster Halbband, Januar bis Mai, 22–39 pp. Translated in English under the direction of the Physical Society of London. Helmholtz, H. F. 1888. The thermodynamics of chemical processes. Physical memoirs selected and translated from foreign sources, Vol. I. Part II, pp. 43–62. London: Taylor and Francis.

  • Hiebert, E.N. 1981. Nernst, Hermann walther. In Dictionary of scientific biography, vol. 15, ed. C.C. Gillispie, 432–453. New York: Charles Scribner Sons.

    Google Scholar 

  • Horstmann, A. 1873. Theorie der Dissociation. Annalen der Chemie und Pharmacie 170:192–210. An English translation appears as Horstmann, A. (2009). The theory of dissociation. Bulletin for the History of Chemistry 34: 76–82.

    Google Scholar 

  • Jaki, S.L. 1984. Uneasy genius: The life and work of Pierre Duhem. The Hague: Martinus Nijhoff Publishers.

    Google Scholar 

  • Jensen, W.B. 2009. August Hortsmann and the origin of chemical thermodynamics. Bulletin of the History of Chemistry 34: 83–91.

    Google Scholar 

  • Jolls, K.R. 1989a. Gibbs and the art of thermodynamics. In Proceedings of the Gibbs symposium, ed. D.G. Caldi and G.D. Mostow, 39–72. Providence, RI: American Mathematical Society.

    Google Scholar 

  • Jolls, K.R. 1989b. Understanding thermodynamics through interactive computer graphics. Chemical Engineering Progress 85: 64–69.

    Google Scholar 

  • Kipnis, A. 1991. Early chemical thermodynamics: Its duality embodied in Van ‘t Hoff and Gibbs. In Thermodynamics, history, and philosophy, facts, trends, and debates, ed. K. Martinas, L. Ropolyi, and P. Szegedi, 492–507. London: World Scientific Publishing Co.

    Google Scholar 

  • Klein, M.J. 1984. The scientific style of Josiah Willard Gibbs. In Essays of founders of modern science, ed. A. Aris, H.T. Davis, and R.H. Stewers, 142–162. Minneapolis: University of Minnesota Press.

    Google Scholar 

  • Klein, M.J. 1989. The physics of J. Willard Gibbs in his time. In Proceedings of the Gibbs symposium, ed. D.G. Caldi and G.D. Mostow, 5–8. Providence, Rl: American Mathematical Society.

    Google Scholar 

  • Klein, M.J. 1990. Duhem on Gibbs. In Essays in honor of Robert Schofield, ed. E. Garber, 52–67. London: Associated University Press Inc.

    Google Scholar 

  • Kragh, H., and S.J. Weininger. 1996. Sooner Silence than confusion: The Tortuous entry of entropy into chemistry. Historical Studies in the Physical and Biological Sciences 27: 91–130.

    Google Scholar 

  • Lewis, G.N., and M. Randall. 1923. Thermodynamics and the free energy of chemical substances, 1st ed. New York: McGraw-Hill Company.

    Google Scholar 

  • Manville, O. 1928. L’Oeuvre scientifique de Pierre Duhem. Memoires de la Societe scientifique des sciences physiques et naturelles de Bordeau, Paris: Blanchard.

  • Massieu, F. 1869. Sur les fonctions caractéristiques des divers fluides et sur la théorie des vapeurs. Comptes Rendus 69: 858–862, and 1057–1061.

  • Massieu, F. 1876. Thermodynamique: Mémoire sur les fonctions caractéristiques des divers fluides et sur la théorie des vapeurs. Académie des Sciences de L’Institut National de France 22: 1–92.

    Google Scholar 

  • Mathias, P.M. 2016. The Gibbs–Helmholtz equation for chemical process and technology. Industrial and Engineering Chemistry Research 55: 1076–1087.

    Google Scholar 

  • Maugin, G. 2014. Continuum mechanics through the eighteenth and nineteenth century. Historical perspectives from John Bernoulli (1727) to Ernst Hellinger (1914). Cham Heidelberg: Springer.

    MATH  Google Scholar 

  • Maxwell, J.C. 1872. Theory of heat, 3rd ed. London: Longmans, Green, and Co.

    Google Scholar 

  • Maxwell, J.C. 1875. On the dynamical evidence of the molecular constitution of bodies. Nature 11: 357–359.

    Google Scholar 

  • Maxwell, J.C. 1876. On the equilibrium of heterogeneous substances. Philosophical Magazine, Series 6 (16): 818–824.

    MATH  Google Scholar 

  • Maxwell, J.C. 1902. Theory of heat. London: Longmans, Green, and Co.

    Google Scholar 

  • Meyerhoffer, W. 1893. Die Phasenregel und ihre Anwendungen, 72. Leipzig: Franz Deuticke.

    Google Scholar 

  • Miller, D.G. 1963. Duhem and the Gibbs–Duhem equation. Journal of Chemical Education 40: 648.

    Google Scholar 

  • Miller, D.G. 1966. Ignored intellect: Pierre Duhem. Physics Today 19: 47–53.

    Google Scholar 

  • Miller, D.G. 1981. Duhem Pierre-Maurice-Marie. In Dictionary of scientific biography, vol. 4, ed. C.C. Gillispie, 225–233. New York: Charles Scribner’s Sons.

    Google Scholar 

  • Moore, C.E., A. von Smolinski, and H. Jaselskis. 2002. The Ostwald–Gibbs correspondence: An interesting component in the history of the energy concept. Bullerin of the History of Chemistry 2: 114–127.

    Google Scholar 

  • Müller, I. 2006. A history of thermodynamics. The doctrine of energy and entropy, 127–164. Berlin: Springer.

    Google Scholar 

  • Needham, P. 2011. Commentary of the principles of thermodynamics by Pierre Duhem. Dordrecht: Springer.

    Google Scholar 

  • Nernst, W. 1893. Theoretische Chemie, vom Standpunkte der Avogadroschen Regel und der Thermodynamik. Stuttgart: Verlag von Ferdinand Enke. Translated by Lehfeldt, R. A., Nernst, W. 1904. Theoretical Chemistry from the standpoint of Avogadro’s Rule and Thermodynamics. Fourth edition, London: Macmillan and Co.

  • Nernst, W. 1906. Über die Berechnung Chemischer Gleichgewichte aus Thermischen Messungen. Nachrichten von der Königlichen Gesebchn/t der Wimenscholten. zu Göttingen Mathematisch-Physikalische Klasse. Verlag von Wilhelm Knapp, Halle (Saale) 1924.

  • Nernst, W. 1918. Die theoretischen und experimentellen Grundlagen des neuen Wärmesatzes. Halle: Verlag von Wilhelm Knapp. Translated in English: Nernst, W. 1969. The New heat theorem. Its foundation in theory and experiment. New York: Dover Publications Inc.

  • Partington, J.R. 1913. A Text-book of thermodynamics (with special reference to Chemistry). London: Constable and Company Ltd.

    MATH  Google Scholar 

  • Rocke, A.I. 1993. The quiet revolution Hermann Kolbe and the science of organic chemistry. Berkeley: University of California Press.

    Google Scholar 

  • Root-Bernstein, R.S. 1980. The ionists: Founding physical chemistry 1872-1890. Ph.D. dissertation, Ann Arbor: Princeton University.

  • Saurel, P. 1902. On the critical state of a one component system. The Journal of Physical Chemistry 6: 474–491.

    Google Scholar 

  • Servos, J.W. 1982. A disciplinary program that failed: Wilder D. Bancroft. The Journal of Physical Chemistry, 1896–1933 73: 207–232.

    Google Scholar 

  • Servos, J.W. 1990. Physical chemistry from Oswald to Pauling The making of a science in America. Princeton, NJ: Princeton University Press.

    Google Scholar 

  • Stokes, G.G. 1907. In Memoir and scientific correspondence of the late Sir George Gabriel Stokes, Vol. 2, Section III, ed. J. Larmor. Cambridge: Cambridge University Press.

    Google Scholar 

  • Thomson, J. 1871. Considerations on the abrupt change at boiling or condensing in reference to the continuity of the fluid state of matter. Proceedings of the Royal Society 20: 278–286.

    Google Scholar 

  • Truesdell, C. 1952. The mechanical foundations of elasticity and fluid dynamics. Journal of Rational Mechanics and Analysis 1: 125–300.

    MathSciNet  MATH  Google Scholar 

  • Truesdell, C. 1984. Rational thermodynamics, 2nd ed. New York: Springer.

    MATH  Google Scholar 

  • Truesdell, C. 1986. What did Gibbs and Caratheodotory leave us about thermodynamics? In New perspectives in thermodynamics, ed. J. Serrin, 101–113. Berlin: Springer.

    Google Scholar 

  • Van der Waals, J.D. 1873. Over de continuiteit van den gas- en vloeistoftoestand. Leiden: Sijthoff, A.W.

    MATH  Google Scholar 

  • van’t Hoff, J. H. 1884. Études de Dynamique Chimique. Amsterdam: Frederic Muller & Co. Revised and enlarged by Cohen E., and translated in English by Ewan T. van’t Hoff, J. H. 1896. Studies in chemical dynamics. Amsterdam: Frederic Muller & Co.

  • van’t Hoff, J. H. 1887. Die Rolle des osmotischen Druckes in der Analogie zwischen Lösungen und Gasen, Zeitscbrift für Physikalische Chemie 1: 481–508. Translated in English, van’t Hoff, J. H. 1995. The role of osmotic pressure in the analogy between solutions and gases. Journal of Membrane Science 100: 39–44.

  • van’t Hoff, J.H. 1903. Physical chemistry in the service of the sciences. The Decennial Publications of the University of Chicago, vol. 18. Chicago: Chicago University Press.

    Google Scholar 

  • van’t Hoff, J.H. 1912. Die chemischen Grundbegriffe nach Menge, Mass und Zeit. Braunschweig: Vieweg, & Sohn.

    MATH  Google Scholar 

  • Wheeler, I.P. 1962. Josiah Willard Gibbs the history of a great mind. New York: Yale University Press.

    MATH  Google Scholar 

  • Wilson, E.B. 1936. Papers I and II as Illustrated by Gibbs’ lectures on thermodynamic. In Commentary on the scientific writings of J. Willard Gibbs, vol. 2, ed. F.G. Donnan and A. Haas, 19–59. New Haven, Connecticut: Yale University Press.

    Google Scholar 

  • Wisniak, J. 2003. Hendrik-Willem Bakhuis Roozeboom: Equilibrium and phase topology. Journal of Phase Equilibria 24: 422–430.

    Google Scholar 

Download references

Acknowledgements

The author is grateful to Professor Theodore Arabatzis from the University of Athens who provided valuable scientific guidance in writing this article.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Photis Dais.

Ethics declarations

Conflict of interest

On behalf of all authors, the corresponding author states that there is no conflict of interest.

Additional information

Communicated by Jed Buchwald.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dais, P. Impact of Gibbs’ and Duhem’s approaches to thermodynamics on the development of chemical thermodynamics. Arch. Hist. Exact Sci. 75, 175–248 (2021). https://doi.org/10.1007/s00407-020-00259-8

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00407-020-00259-8

Navigation