Skip to main content

How Chemical Is Quantum Chemistry?

  • Chapter
  • First Online:
Philosophical Perspectives in Quantum Chemistry

Part of the book series: Synthese Library ((SYLI,volume 461))

  • 355 Accesses

Abstract

This chapter is an exploration of the applicability of quantum chemistry to central aspects of chemical knowledge like the concept of chemical bond and its material background. Lewis’s electronic theory, precursor of modern quantum chemistry, is discussed, and the works of Linus Pauling and Hans Primas are examined from a conceptual viewpoint. Finally, the general limitation of the quantum approach with respect to the chemistry of substances is emphasized.

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

Access this chapter

Subscribe and save

Springer+ Basic
$34.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 99.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 129.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 129.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Similar content being viewed by others

Notes

  1. 1.

    This trichotomy is represented in modern textbooks, for example in the famous work by Holleman and Wiberg (1964), in which the ideal types mentioned are referred to as first-order bonds.

  2. 2.

    Within the scope of his criticism of Kripke’s and Putnam’s theories of meaning, Hacking briefly outlines the early history of the electron and states: “No one at present doubts that the electron is a natural kind of fundamental importance” (Hacking, 1983: 84). For the sake of neutrality, note that electrons in a culturalist perspective are theoretical entities or “constructs” (Psarros, 1999: 306) and their “discovery” is thus essentially an invention (Hanekamp, 1997: 227).

  3. 3.

    According to Laidler (1993: 206), Faraday himself came close to a similar interpretation, although his notion of electricity as a fluid prevented him from coming to a strict atomistic and corpuscular interpretation.

  4. 4.

    Gilbert Lewis also failed to note the Irishman as late as 1923 when he wrote: “It was Helmholtz in his celebrated Faraday lecture of 1881 who first pointed out this deduction of the atom of electricity, or as it is now called, the electron.” (Lewis, 1923: 21).

  5. 5.

    An early indication of this is the statement by Ida Freund (1904: 540): “Not more than a beginning has been made as yet; but whilst formerly it had not even been possible to say what class of conceptions were likely to supply the building material for the adequate hypothesis and theory of valency that is required, it now seems probable that these conceptions will be the indivisible ‘electron,’ to which the ions owe their electrical charges and the lines of force connecting the electrons.”

  6. 6.

    A detailed presentation of the history of this theory would go beyond the scope of the present chapter. I refer here to the exquisite standard work by Anthony Stranges (1982). Hanekamp (1997) provides a concise overview from a culturalist respective.

  7. 7.

    He received the Nobel Prize in 1919 for the discovery of the optical Doppler effect in canal rays and the splitting of spectral lines in electric fields. It bears mention here that Stark later became a fervent supporter of National Socialism and the so-called “German physics” yet after the war went practically unpunished, see Mehrtens and Richter (1980) and Hoffmann (2005).

  8. 8.

    With respect to the discussion of the existence of stable atoms, Heisenberg speaks of binding energies and chemical forces; for him the binding of even one electron in an isolated atom represents a “chemical” bond (Heisenberg, 1989: 89).

  9. 9.

    This is the reproduction of Fig. 41 from Stark (1915: 120). In the same place the physicist Stark speculates about the existence of the H3 molecule. The author shows a figure with nearly identical content in the earlier work mentioned here (Stark, 1908: 86).

  10. 10.

    A particle’s intrinsic rotational impulse is referred to as spin in quantum mechanics. It can assume two values for one energy level and is thus a half-integer lacking a classical correlate and cannot really be depicted. Only with the assumption of spin can all phenomena occurring in the line spectra be fully accounted for. The formation of an electron pair in a covalent chemical bond is not comprehensible in a classic sense because particles of the same charge should repel rather than attract each other. In the statement quoted at the beginning, Coulson projects the current state of knowledge a bit too far back.

  11. 11.

    Cf. here Russell (1971), Stranges (1982), and Jensen (1984).

  12. 12.

    Cf. the fresh appreciation of Abegg in Scerri (2016: 63–78).

  13. 13.

    The first paper on this topic was the programmatic study by Abegg and Bodländer (1899).

  14. 14.

    Something similar may be found in the short physiologically motivated work by Georg Buchner. The glossary entry under electroaffinity reads as follows: “The elementary atoms combine not only with other elementary atoms but also have the ability to bind electricity atoms (electrons) (electroaffinity). According to the unitary view of electricity, we refer to the elementary atoms connected with electrons as negative ions and the elementary atoms that have given off electrons as positive ions. The binding force for electrons increases with the electronegative character of the elementary atoms.” (Buchner, 1912: 136; my emphasis). Note that Buchner does not mention Abegg.

  15. 15.

    At this point I will skip over the classic topic of the preservation of elements when forming and separating composite substances and its semantic peculiarities.

  16. 16.

    Lewis , in the same place, calls the approach “the theory of the cubical atom” and dates its inception to the preparation for an introductory lecture of 28 March 1902. He notes that it was not the priorities that were important to him but the common ground with similar theoretical attempts (aside from Abegg, whom he mentions explicitly, certainly the one by Kossel as well), whereby the fact that “all possess some characteristics of fundamental reality” would contribute to their probability. He does not have a realistic microstructural interpretation in mind.

  17. 17.

    Langmuir’s introductory statement in this work is telling: “The problem of the structure of atoms has been attacked mainly by physicists who have given little consideration to the chemical properties which must ultimately be explained by a theory of atomic structure. The vast store of knowledge of chemical properties and relationships, such as is summarized by the Periodic Table, should serve as a better foundation for a theory of atomic structure than the relatively meager experimental data along purely physical lines. ” (Langmuir, 1919: 868). The American industrial chemist Langmuir (who received his doctorate in 1909 in Göttingen) received the Nobel Prize in 1932 for work in the field of surface chemistry.

  18. 18.

    Cf. the detailed discussion of this episode from the perspective of the philosophy of chemistry in van Brakel (2000: 27–34). The author comes to the following evaluation (ibid.: 34): “The debates about the theory of resonance show that nationalistic concerns were of much greater importance than the dogma’s of dialectical materialism. The concerns of the debates and power machinations in Moscow promoting Buterlov had no followers in the GDR –though it is no doubt correct that Butlerov’s importance had been suppressed in German and English text books that gave reviews of the history of organic chemistry.”

  19. 19.

    Hans Vermeeren (1986) concurs. Valeria Mosini (2000: 578), however, presents an opposing viewpoint: “(…) as a matter of fact, the interpretation of resonance given by the chemists is a realist one, as the evidence from the history of the theory strongly suggests.”

  20. 20.

    Cf. Staab (1959: 642–650)

  21. 21.

    Jaap van Brakel (2012) refers to this point.

  22. 22.

    Psarros (1999: 169) says of the oxidation number: “It specifies which valence level the atoms of a given nonmetal assume when they are combined with atoms of other elements.” First, this characterization is incomplete because it mentions only nonmetals and, second, it is unclear because the oxidation number has nothing to do with valence (itself a concept requiring clarification) but with the ratio of the sorts of atoms involved to the electrons available for binding (valence electrons), cf. Riedel (2004: 128).

  23. 23.

    With this deduction it is assumed (from experience) that the basic arrangement of the sulfate ion is such that the sulfur atom is in the center and the oxygen atoms arrange themselves around it. Compounds with oxygen atoms can be excluded here.

  24. 24.

    Naturally, the assumption that atoms and electrons exist itself presupposes that the samples of substances can be divided or quantized at the submicroscopic level. Yet quantum mechanics was not yet used here.

  25. 25.

    There are arguments in favor of continuing to classify the octet rule as not merely heuristic but to take it seriously even from a theoretical standpoint. See Weidenbruch (1994: 20): “Lewis’s octet rule has lost none of its validity since it was first formulated over 90 years ago. It applies to compounds of the main group of elements with the correct number of electrons, which include nearly all organic compounds, as well as to compounds with excess electrons in which the supernumerary electrons can be accommodated in nonbinding orbitals of the substitutes.”

  26. 26.

    In this “wave equation” E and V are, respectively, the kinetic and potential energy of an electron in an electric field (between nucleus and electron), m is the mass of the electron, ħ in the Planck constant h/2π, and ∇ (the “Laplace operator”) has the form /∂x2 + /∂y2 + /∂z2 for the usual three-dimensional case. Where the mass of the electron is regarded as unchanging, we refer to the time-independent or nonrelativistic Schrödinger equation .

  27. 27.

    I cannot go into detail regarding another very important issue here: Most of the pertinent experiments which brought up quantum theory and quantum mechanics relied on classical physics, which causes additional interpretative problems, cf. Chang (1995, 1997).

  28. 28.

    The wave number of the “Balmer series” is expressed today as \( \tilde{\nu}={R}_{\infty}\left(1/{2}^2-1/{n}^2\right) \), with the Rydberg constant R and n = 3, 4, …

  29. 29.

    I refer to the modern nuclear and quantum physicists in this manner because the Schrödinger equation at its core has to do with potential and kinetic energy.

  30. 30.

    In Die Natur der Chemischen Bindung, Pauling states: “Before 1927 there was no satisfactory theory for the covalent bond. The chemists had postulated the existence of a valence bond between the atoms, a notion that was corroborated by experience, but all efforts at gaining insight into the nature of the bond had remained largely unsuccessful. Lewis, who associated two electrons with one bond, took a step forward in doing so, yet one can hardly speak of the development of a theory. The decisive questions as to the nature of the interaction and the cause of the binding energy remain unanswered.” (Pauling, 1962: 21).

  31. 31.

    Cf. Ladik, 1973: 203–212.

  32. 32.

    Regarding “electron exchange” Heitler states: “The electron exchange is […] primarily a typical quantum mechanical effect and all attempts to find a ‘classic analogy’ must fail […]. The cause of the exchange effect is the very fact that electrons are indistinguishable and therefore an exchange of electrons is principally unobservable.” (Heitler, 1961: 99–100).

  33. 33.

    Fig. 19 from Ladik (1973: 216). The x axis shows Bohr radius units (0,5285 A); the y axis shows the (potential) energy in electron volts. ES shows the ground state; EA shows the first excited state. Dashed lines show measured values; solid lines show calculated values. Figures with identical content are also found in other introductions, including Heitler and London (1927: 462) and Heitler (1961: 98).

  34. 34.

    The same scheme is used for the discussion of a possible helium molecule He2, whereby the total of four electrons now, in addition to fully occupying the binding state, would also fully occupy the antibinding state, which explains the nonexistence of this particle (or a substance consisting of it).

  35. 35.

    Regarding the over hundred-year-old history of electronegativity, see the works of Bill Jensen (1996, 2003, 2012). Ruthenberg and Martínez González (2017) examine this topic from the perspective of the philosophy of chemistry. Many modern representations follow Pauling and provide an unnecessarily abbreviated definition of electronegativity, as is the case in Riedel (2004: 119): “A measure of the capacity of an atom to draw the binding electron pair toward itself in an atomic bond is electronegativity.”

  36. 36.

    Fig. 3–8 from Pauling (1962: 95).

  37. 37.

    Accordingly, Preuss (1960: 241) discusses Pauling’s approach as a “semi-empirical method.”

  38. 38.

    There are attempts to calculate electronegativities from spectrometric data, from dipole moments, and from electron affinities and ionization energies (cf. Ruthenberg & Martínez González, 2017).

  39. 39.

    With respect to this publication, an introduction to quantum chemistry, I regard Primas as the senior author and especially as the source of the more philosophical comments, which here are particularly clear and productive. Other pertinent references include Primas (1983, 1985a, b, 1992). Cf. the article by Amann and Gans (1989), authors close to Primas in terms of content.

  40. 40.

    As to chemical pluralism, see Ruthenberg and Mets (2020). A recent approach of perspectivism, including a discussion of the electron and its charge as episode from the history of modern physics, is presented by Michela Massimi (2021).

  41. 41.

    I have borrowed this formulation, which describes the situation very well, from Preuss (1960: 241). It should be noted that such representations and similar ones occasionally convey (or possibly are intended to convey) the impression that the words “quantum mechanical” impart the content a particularly high degree of scientific reliability and significance. At the same time the relationship between empiricism and theory is often nearly turned upside down when it is repeatedly emphasized that quantum mechanics is among the physical theories best confirmed by experimental findings, whereas its origin from spectroscopic experience is practically ignored.

  42. 42.

    See the extraordinarily comprehensive presentation, not exclusively referring to French (philosophy of) chemistry, by Bernadette Bensaude-Vincent (2009).

  43. 43.

    The textbook in question by the Princeton chemists Glasstone et al. (1941) contains the original graphic of the potential surface representation on page 96. Primas and Müller-Herold also reproduce this figure (1990: 202) and briefly discuss “adiabatic reactions” although this is not actually a genuinely quantum chemical topic. In their summary they state (1990: 205): “Often even a contemplation of the plausibility of the structure of the transitional complexes allows interesting qualitative statements to be made about the possible course of chemical reactions.” Hinne Hettema (2012) discusses the case study of the “absolute reaction rate theory” from the standpoint of the philosophy of chemistry. The author is receptive to reductionism in the sense of a path to a unified natural science and regards the theory of the transitional state as a typical example of this attitude.

  44. 44.

    Other authors also come to a similar conclusion. Paul Bogaard for example writes in a brief but very informative article about the influence of Gilbert Lewis on Linus Pauling and Charles Coulson about the latter’s opinion (2003: 308): “Molecular structure –and the search for directionality of bonding that could produce it– is motivated neither by the mathematical formalisms of quantum theory nor by the physics which underlies quantum mechanics, but by experience with chemical substances.”

  45. 45.

    By observable is meant somehow physically tangible and therefore measurable values.

  46. 46.

    They note that the Bohm variant of quantum mechanics has better chances of being compatible with the QTAIM but they do not expand on this idea (ibid.: 135).

  47. 47.

    I see this as further support for my claim that energetics and kinetics are fundamentally different chemical concepts.

  48. 48.

    The examples listed come from Primas and Müller-Herold (1990: 125 and 197).

  49. 49.

    Here I have borrowed part of the title of the book Macroscopic Metaphysics – Middle-Sized Objects and Longish Processes by Paul Needham, which is in fact a work in the philosophy of chemistry (Needham, 2017).

  50. 50.

    In the textbook Philosophie und Naturwissenschaften by a group of East German authors, we read: “The achievement of quantum theory as applied to chemical problems consists not only in having provided a more adequate theory of the chemical bond but also in the realization that the laws of the macro level cannot be directly applied to the micro level.” (Author collective, 1988: 189). I agree with this, although I note that the reverse statement is also true; it is equally impossible to apply the laws of the micro level to the macro level.

References

  • Abegg, R. (1904). Die Valenz und das periodische System. Versuch einer Theorie der Molekularverbindungen. Zeitschrift für Anorganische und Allgemeine Chemie, 39, 330–380.

    Article  Google Scholar 

  • Abegg, R., & Bodländer, G. (1899). Die Elektroaffinität, ein neues Prinzip der chemischen Sysiematik. Zeitschrift für Anorganische und Allgemeine Chemie, 20, 453–499.

    Article  Google Scholar 

  • Amann, A., & Gans, W. (1989). Die Theoretische Chemie auf dem Weg zu einer Theorie der Chemie. Angewandte Chemie, 101, 277–285.

    Article  Google Scholar 

  • Arriaga, J. A., Fortin, S., & Lombardi, O. (2019). A new chapter in the problem of the reduction of chemistry to physics: The Quantum Theory of Atoms in Molecules. Foundations of Chemistry, 21, 125–136.

    Article  Google Scholar 

  • Autor collective. (1988). Philosophie und Naturwissenschaften. VEB Deutscher Verlag der Wissenschaften.

    Google Scholar 

  • Bader, R. F. W., & Matta, C. F. (2013). Atoms in molecules as non-overlapping, bounded, space-filling open quantum systems. Foundations of Chemistry, 15, 253–276.

    Article  Google Scholar 

  • Bader, R. F. W., Popelier, P. L., & Keith, T. A. (1994). Die theoretische Definition einer funktionellen Gruppe und das Paradigma des Molekülorbitals. Angewandte Chemie, 106, 647–659.

    Article  Google Scholar 

  • Bensaude-Vincent, B. (2009). Philosophy of chemistry. In A. Brenner & J. Gayon (Eds.), French studies in the philosophy of science (pp. 165–186). Springer.

    Chapter  Google Scholar 

  • Bogaard, P. (2003). How G.N. Lewis reset the terms of the dialogue between chemistry and physics. Annals of the New York Academy of Sciences, 988, 307–312.

    Article  Google Scholar 

  • Buchner, G. (1912). Angewandte Ionenlehre. Verlag J.F. Lehmann.

    Google Scholar 

  • Chang, H. (1995). The quantum counter-revolution: Internal conflicts in scientific change. Studies in History and Philosophy of Modern Physics, 26, 121–136.

    Article  Google Scholar 

  • Chang, H. (1997). On the applicability of the quantum measurement formalism. Erkenntnis, 46, 143–163.

    Article  Google Scholar 

  • Coulson, C. A. (1952). Quantum theory, theory of molecular structure and valence. Annual Review of Physical Chemistry, 3, 1–18.

    Article  Google Scholar 

  • Fortin, S., Lombardi, O., & Martínez González, J. C. (2016). Isomerism and decoherence. Foundations of Chemistry, 18, 225–240.

    Article  Google Scholar 

  • Freund, I. (1904). The study of chemical composition. Cambridge University Press.

    Google Scholar 

  • Glasstone, S., Laidler, K., & Eyring, H. (1941). The theory of rate processes. McGraw-Hill.

    Google Scholar 

  • Hacking, I. (1983). Representing and intervening. Cambridge University Press.

    Book  Google Scholar 

  • Hanekamp, G. (1997). Protochemie – Vom Stoff zur Valenz. Königshausen & Neumann.

    Google Scholar 

  • Heisenberg, W. (1989). Ordnung der Wirklichkeit. Piper.

    Google Scholar 

  • Heitler, W. (1961). Elementare Wellenmechanik. Mit Anwendungen auf die Quantenchemie. Friedrich Vieweg & Sohn.

    Google Scholar 

  • Heitler, W., & London, F. (1927). Wechselwirkung neutraler Atome und homöopolare Bindung nach der Quantenmechanik. Zeitschrift für Physik, 44, 455–472.

    Article  Google Scholar 

  • Hettema, H. (2012). The unity of chemistry and physics: Absolute reaction rate theory. HYLE – International Journal for Philosophy of Chemistry, 18, 145–173.

    Google Scholar 

  • Hoffmann, D. (2005). Between autonomy and accomodation: The German Physical Society during the Third Reich. Physics in Perspective, 7, 293–329.

    Article  Google Scholar 

  • Holleman, A. F., & Wiberg, E. (1964). Lehrbuch der Anorganischen Chemie. Walter de Gruyter.

    Book  Google Scholar 

  • Hund, F. (1927). Zur Deutung der Molekelspektren III. Zeitschrift für Physik, 43, 805–826.

    Article  Google Scholar 

  • Jensen, W. B. (1984). Abegg, Lewis, Langmuir, and the Octet Rule. Journal of Chemical Education, 61, 191–200.

    Article  Google Scholar 

  • Jensen, W. B. (1996). Electronegativity from Avogadro to Pauling. Part I: Origins of the electronegativity concept. Journal of Chemical Education, 73, 11–20.

    Article  Google Scholar 

  • Jensen, W. B. (2003). Electronegativity from Avogadro to Pauling. Part II: Late nineteenth- and early twentieth-century developments. Journal of Chemical Education, 80, 279–287.

    Article  Google Scholar 

  • Jensen, W. B. (2012). When was electronegativity first quantified? Journal of Chemical Education, 89, 94–96.

    Article  Google Scholar 

  • Johnstone Stoney, G. (1881). On the physical units of nature. Philosophical Magazine, 11, 381–390.

    Google Scholar 

  • Johnstone Stoney, G. (1894). On the ‘electron’ or atom of electricity. Philosophical Magazine, 38, 418–420.

    Google Scholar 

  • Ladik, J. (1973). Quantenchemie. Ferdinand Enke Verlag.

    Google Scholar 

  • Laidler, K. (1993). The world of physical chemistry. Oxford University Press.

    Google Scholar 

  • Laitko, H., & Sprung, W.-D. (1970). Chemie und Weltanschauung. Verlag Hubert Freistühler.

    Google Scholar 

  • Langmuir, I. (1919). The arrangement of electrons in atoms and molecules. Journal of the American Chemical Society, 41, 868–934.

    Article  Google Scholar 

  • Latour, B. (1999) Pandora´s Hope. Harvard University Press.

    Google Scholar 

  • Lewis, G. N. (1916). The atom and the molecule. Journal of the American Chemical Society, 38, 762–785.

    Article  Google Scholar 

  • Lewis, G. N. (1923). Valence and the structure of atoms and molecules. The Chemical Catalog Company.

    Google Scholar 

  • Lewis, G. N. (1926). The anatomy of science. Yale University Press.

    Google Scholar 

  • Martínez González, J. C., Fortin, S., & Lombardi, O. (2019). Why molecular structure cannot be strictly reduced to quantum mechanics. Foundations of Chemistry, 21, 31–45.

    Article  Google Scholar 

  • Massimi, M. (2021). Realism, perspectivism, and disagreement in science. Synthese, 198, 6115–6141.

    Article  Google Scholar 

  • Mehrtens, H., & Richter, S. (1980). Naturwissenschaft, Technik und NS-Ideologie. Suhrkamp Taschenbuch Wissenschaft.

    Google Scholar 

  • Mosini, V. (2000). A brief history of the theory of resonance and of its interpretation. Studies in History and Philosophy of Modern Physics, 31, 569–581.

    Article  Google Scholar 

  • Needham, P. (2017). Macroscopic metaphysics – Middle-sized objects and longish processes. Springer.

    Book  Google Scholar 

  • Ostwald, W. (1917). Grundriss der Allgemeinen Chemie. 5.Aufl. Verlag von Theodor Steinkopff.

    Google Scholar 

  • Pauling, L. (1962). Die Natur der Chemischen Bindung. Verlag Chemie. Original (1931) The Nature of the Chemical Bond.

    Google Scholar 

  • Preuss, H. (1960). Die gegenwärtige Situation der Quantenchemie. Die Naturwissenschaften, 47, 241–249.

    Article  Google Scholar 

  • Primas, H. (1983). Chemistry, quantum mechanics and reductionism. Springer.

    Book  Google Scholar 

  • Primas, H. (1985a). Kann Chemie auf Physik reduziert werden? Erster Teil: Das Molekulare Programm. Chemie in unserer Zeit, 19, 109–119.

    Article  Google Scholar 

  • Primas, H. (1985b). Kann Chemie auf Physik reduziert werden? Zweiter Teil: Die Chemie der Makrowelt. Chemie in unserer Zeit, 19, 160–166.

    Article  Google Scholar 

  • Primas, H. (1992). Umdenken in der Naturwissenschaft. GAIA – Ecological Perspectives in Science, Humanities, and Economics, I, 5–15.

    Article  Google Scholar 

  • Primas, H., & Müller-Herold, U. (1990). Elementare Quantenchemie. B. G. Teubner.

    Book  Google Scholar 

  • Psarros, N. (1999). Die Chemie und ihre Methoden. Wiley-VCH.

    Book  Google Scholar 

  • Riedel, E. (2004). Allgemeine und Anorganische Chemie. 8. Auflage. Walter de Gruyter.

    Google Scholar 

  • Russell, C. A. (1971). The history of Valency. Leicester University Press.

    Google Scholar 

  • Ruthenberg, K., & Martínez González, J. C. (2017). Electronegativity and its multiple faces: Persistence and measurement. Foundations of Chemistry, 19, 61–75.

    Article  Google Scholar 

  • Ruthenberg, K., & Mets, A. (2020). Chemistry is pluralistic. Foundations of Chemistry, 22, 403–419.

    Article  Google Scholar 

  • Scerri, E. R. (2016). A tale of seven scientists – And a new philosophy of science. Oxford University Press.

    Google Scholar 

  • Staab, H. A. (1959). Einführung in die Theoretische Organische Chemie. Verlag Chemie.

    Google Scholar 

  • Stark, J. (1908). Zur Energetik und Chemie der Bandenspektra. Physikalische Zeitschrift, 9, 85–94.

    Google Scholar 

  • Stark, J. (1915). Prinzipien der Atomdynamik. III. Teil: Die Elektrizität im Chemischen Atom. Verlag von S. Hirzel.

    Google Scholar 

  • Stranges, A. N. (1982). Electrons and valence. Texas A&M University Press.

    Google Scholar 

  • van Brakel, J. (2000). Philosophy of chemistry. Leuven University Press.

    Google Scholar 

  • van Brakel, J. (2012). Substances: The ontology of chemistry. In R. F. Hendry, P. Needham, & A. Woody (Eds.), Handbook of the philosophy of science volume 6 – Philosophy of chemistry (pp. 191–229). Elsevier.

    Google Scholar 

  • Vermeeren, H. P. W. (1986). Controversies and existence claims in chemistry: The theory of resonance. Synthese, 69, 273–290.

    Article  Google Scholar 

  • Weidenbruch, M. (1994). Das Elektronenoktett: Renaissance einer totgeglaubten Regel. Chem, 1, 15–20.

    Google Scholar 

Download references

Acknowledgments

I thank Olimpia Lombardi and Juan Camilo Martínez González for the opportunity to contribute to this volume, for the help with the manuscript, and for all the discussions during the last decade. I also thank an anonymous referee for comments which helped to improve the article. The manuscript was written in German by the author and translated into English by John Grossman. I thank him as well.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Klaus Ruthenberg .

Editor information

Editors and Affiliations

Rights and permissions

Reprints and permissions

Copyright information

© 2022 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Ruthenberg, K. (2022). How Chemical Is Quantum Chemistry?. In: Lombardi, O., Martínez González, J.C., Fortin, S. (eds) Philosophical Perspectives in Quantum Chemistry. Synthese Library, vol 461. Springer, Cham. https://doi.org/10.1007/978-3-030-98373-4_2

Download citation

Publish with us

Policies and ethics