Skip to main content
Log in

Explanation and theory formation in quantum chemistry

  • Published:
Foundations of Chemistry Aims and scope Submit manuscript

Abstract

In this paper I expand Eric Scerri’s notion of Popper’s naturalised approach to reduction in chemistry and investigate what its consequences might be. I will argue that Popper’s naturalised approach to reduction has a number of interesting consequences when applied to the reduction of chemistry to physics. One of them is that it prompts us to look at a ‘bootstrap’ approach to quantum chemistry, which is based on specific quantum theoretical theorems and practical considerations that turn quantum ‘theory’ into quantum ‘chemistry’ proper. This approach allows us to investigate some of the principles that drive theory formation in quantum chemistry. These ‘enabling theorems’ place certain limits on the explanatory latitude enjoyed by quantum chemists, and form a first step into establishing the relationship between chemistry and physics in more detail.

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.

Institutional subscriptions

Similar content being viewed by others

Notes

  1. As suggested by one of the referees of this paper.

  2. They also associate the SIS requirement with the provisoes of Hempel (1988) as an assumption about completeness. There is a considerable amount of literature in the philosophy of quantum chemistry on the topic of the ‘isolation’ of molecular systems and the consequences that this assumption has for a principled explanation of chemical phenomena, most of which is in origin due to the work of Primas (1975), or Primas (1983). In recent times, the ‘emergence of a classical world’ in the theory of quantum physics is also studied through decoherence Schlosshauer (2007).

  3. Popper (1957) states in the Preface: ‘I tried to show, in The Poverty of Historicism, that historicism is a poor method—a method which does not bear any fruit. But I did not actually refute historicism’ (p. v). In this sense Popper’s criterion of ‘poor’ is weaker than is generally understood as in the sense of ‘incorrect’. Beyond this contingency, Popper’s argument is of course completely different from mine.

  4. As it stands in regards to scientific practice, however, van Brakel’s argument is not that convincing, as one of the referees of this paper pointed out. Quantum chemical packages, for better or worse, presently are an important tool for many chemical researchers, such as designers of new chemicals including catalysts, nano-materials and drugs.

  5. I wish to note that van Brakel (2000)’s comments on the various editions of Coulson’s Valence (and especially his critique on the last edition edited by McWeeny (1979) are in my view justified. A historical study of the various editions of Valence would make for an interesting research topic outlining how the perception of explanation through quantum chemistry has evolved over time.

  6. For instance, (1) is not strictly speaking a problem, and there are many examples in physical theories where nonlinear equations for higher order properties are developed on the basis of linear equations. (2) could be represented as an internal problem for physics with little bearing on the nature of the relationship between physics and chemistry, while it is hard to claim (3) for instance for short lived transition states. Woolley’s claims about the difficulties with molecular structure have bee the subject of considerable discussion in the literature, and cannot be accepted without significant qualifications. Even so, providing detailed rebuttals of these arguments is not the purpose of this paper.

  7. I would like to note that I disagree with Woody that the MO diagrams share their mathematical structure with that underlying configuration interaction. They do not, at least not the same configuration interaction as it is traditionally understood in quantum chemistry. Much less than configuration interaction is needed to create the MO diagrams. A charitable interpretation of Woody might take it to mean that she intends the comment as an indication of the interaction of the (electronic) configuration of individual atoms, and such a reading might indeed support the MO diagrams as resulting from configuration interaction. Such a reading would be highly discontinuous with current practice however and is very confusing in the present context.

  8. A historical fact which may underpin this contention is that the early quantum chemists tend to be less well-known than the early pioneers of quantum theory. While in the early stages there was considerable overlap in the activities of these two groups, around 1929 there was a distinct difference between the work of early quantum chemists like Hylleraas in Norway, Hellman in Germany and later Russia, Eyring, van Vleck, Pauling and others in the USA, and quantum theorists like Dirac who had a passing brush with chemistry, but moved on to relativistic quantum mechanics in the late 1920s. In the 1930s foundational issues tended to take centre stage as the frontier of development. It is little surprise, therefore, that a number of the theorems we will discuss in this section are closely associated with quantum chemists. This history still needs a considerable amount of fleshing out at this point in time and is in no sense of the word complete.

  9. Proof: we can write the resolution of the identity as

    $$ \hat 1 = \sum_n | \psi_n \rangle \langle \psi_n | $$

    Inserting this into the expectation value 〈Hϕ yields

    $$ \langle H \rangle_\phi = \frac{\sum_n {\langle \phi | H | \psi_n \rangle \langle \psi_n | \phi \rangle}} {\sum_n \langle \phi | \psi_n \rangle \langle \psi_n | \phi \rangle} = \frac{\sum_n E_n {|\langle \phi | \psi_n \rangle |^2}} {\sum_n | \langle \phi | \psi_n \rangle |^2} $$

    Since E 0 < E 1 < E 2 … and | 〈ϕ| ψ n 〉|2 ≥ 0:

    $$ \sum_n E_n |\langle \phi | \psi_n \rangle |^2 \geq E_0\sigma_n {|\langle \phi | \psi_n \rangle |^2} $$

    and thus 〈Hϕ ≥ E 0.

    The existence of a lowest eigenvalue is an assumption which does not hold for Dirac Hamiltonians. The mathematically rigorous proof of the existence of the lowest eigenvalue for Schrödinger Hamiltonians was only given in Kato (1951a, b).

  10. This is the description given in Helgaker et al. (2000) on pp. 267 and 870. Eh is one Hartree, the absolute value of the electronic potential energy of a hydrogen atom in the ground state. It is equal to 2625.5 kJ/mol.

  11. Proof: The proof is again quite simple and hinges on the fact that since \(S_2 \subset S_1\) all functions \(|\psi_i^{S_2}\rangle\) can be expanded in S 1 as follows

    $$ |\psi_i^{S_2}\rangle = \sum_n^{|S_1|} c_n |\psi_i^{S_1}\rangle $$

    The energy \(E_0^{S_2}\) can be written as

    $$ \begin{aligned} E_0^{S_2} =& \langle\psi_i^{S_2}| H | \psi_i^{S_2}\rangle \\ =& \sum_{nm}^{|S_1|} c_n^* c_m \langle\psi_n^{S_1}| H | \psi_m^{S_1}\rangle \\ =& \sum_{nm}^{|S_1|} c_n^* c_m E_m^{S_1} \delta_{nm} \\ &= \sum_n |c_n|^2 E_n^{S_1}. \end{aligned} $$

    This quantity is larger than \(E_0^{S_1}\).

  12. This should be read to apply to molecular systems. There is no indication throughout the paper that Feynman only has atoms in mind, and there are no reasons why the theorem should be restricted to atoms as opposed to molecules. It is equally valid for molecules as well.

  13. Proof: Consider the steady state where

    $$ H (\lambda) | \psi_\lambda \rangle = E_\lambda | \psi_\lambda \rangle $$

    Then, if we assume that ψλ is normalised, so that 〈ψλ| ψλ〉 = 1:

    $$ \begin{aligned} \left. \frac{d E_\lambda} {d\lambda}\right|_{\lambda = 0} &= \frac{\partial} {\partial \lambda} \langle \psi_\lambda|H (\lambda)|\psi_\lambda\rangle \\ &= \left\langle \left.\frac{\partial\psi_\lambda}{ \partial \lambda}\right|_{\lambda=0} | H (\lambda) | \psi_\lambda\right\rangle + \left\langle \psi_\lambda \left| \frac{\partial H (\lambda)}{\partial \lambda} \right| \psi_\lambda \right\rangle +\left\langle \psi_\lambda | H (\lambda) | \left.\frac{\partial\psi_\lambda} {\partial \lambda}\right|_{\lambda=0} \right\rangle \\ &= E_\lambda\left[\left\langle \left.\frac{\partial\psi_\lambda}{\partial \lambda}\right|_{\lambda=0} | \psi\right\rangle + \left\langle \psi \left| \frac{\partial\psi_\lambda}{\partial \lambda}\right|_{\lambda=0} \right\rangle \right] +\left\langle \psi_\lambda \left| \frac{\partial H (\lambda)} {\partial \lambda} \right| \psi_\lambda \right\rangle \\ &= \left\langle \psi_\lambda \left| \frac{\partial H (\lambda) }{\partial \lambda} \right| \psi_\lambda \right\rangle \\ &= \left\langle\psi_\lambda | V | \psi_\lambda \right\rangle\\ \end{aligned} $$

    The term in square brackets vanishes because

    $$ \left[\left\langle \left.\frac{\partial\psi_\lambda}{\partial \lambda}\right|_{\lambda=0} | \psi\right\rangle + \left\langle \psi \left| \frac{\partial\psi_\lambda}{ \partial \lambda}\right|_{\lambda=0} \right\rangle \right] = \frac{\partial} {\partial \lambda} \left\langle \psi_\lambda | \psi_\lambda \right\rangle = 0. $$

    The proof is given in Feynman (1939). The Hellman–Feynman theorem has since become a staple of quantum mechanics courses and the proof is also given in for instance Helgaker et al. (2000) and Cohen-Tannoudji et al. (1977) to mention only a few.

  14. See for instance Olsen and Jorgensen (1985) for a modern framework to compute polarisabilities and hyperpolarisabilities and Hettema (1993) for an application.

  15. Proof: We assume that the time dependent Schrödinger equation is satisfied, so that

    $$ H \psi (t) = i \hbar \frac{\partial \psi }{\partial t} $$

    so that

    $$ \begin{aligned} \frac{\hbox{d}}{{\hbox{d}}t}\langle \psi (t) | A | \psi (t) \rangle &= \left\langle \frac{\partial \psi (t)}{\partial t} | A | \psi (t) \right\rangle + \left\langle \psi (t)| A | {\partial \psi (t) \over \partial t} \right\rangle + \left\langle \psi (t) \left| \frac{\partial A}{\partial t} \right| \psi (t) \right\rangle\\ &= \frac{1}{i \hbar} \left\langle \psi (t) | (AH - HA) | \psi (t) \right\rangle + \left\langle \psi (t) \left| \frac{\partial A}{\partial t} \right| \psi (t) \right\rangle \\ &= \frac{1}{i \hbar} \left\langle \psi (t) | \left[ A , H \right]| \psi (t) \right\rangle + \left\langle \psi (t) \left| \frac{\partial A}{\partial t} \right| \psi (t) \right\rangle \\ \end{aligned} $$

    where the transition from the first to the second line is because ψ(t) satisfied the Schrödinger equation and the transition to the last line is through the definition of the commutator

    $$ [A,H] = AH - HA $$

    The proof is given in a number of places, for instance in Cohen-Tannoudji et al. (1977).

  16. I will not present the proof here, details are found in a number of textbooks such as for instance in Cohen-Tannoudji et al. (1977) on p. 242.

  17. I had a number of interesting discussions with Paul Wormer, Josef Paldus and Bogomil Jeziorski on the somewhat confusing historical situation surrounding the original reference to Wigner’s 2n + 1 theorem and the interesting discussions via email about whether Wigner’s 1935 paper actually contains the (proof of the) theorem. These discussions did not lead to a clear cut conclusion, but I think I can sum the situation up fairly if I say that the 1935 paper did contain the theorem though not the proof, with the proof of the theorem appearing in the 1950s. In addition, Egil Hylleraas already stated the theorem for third order wave functions in 1930, as outlined below.

  18. The ‘art’ of basis set creation quite possibly requires a paper of its own given this very common misconception in the philosophy of chemistry. The criteria that basis sets must meet in order to qualify as a ‘good’ basis set are manifold, but in general a fit to experimental data tends to be one of the lesser considerations. Basis sets are generally evaluated on a number of criteria, such as whether they are ‘even-tempered’, have sufficient ‘polarisation functions’ and such like. A detailed discussion of this would make an interesting paper in the philosophy of chemistry on its own and falls outside the scope of the current paper.

  19. This is not to say that the UHF wave function is the ‘correct’, or even the ‘better’ one, since the UHF wave function fails in other important respects. Specifically, UHF wave functions are not eigenfunctions of S 2 which is in a way the saving grace in the description of the energetic curve but also breaks the spin symmetry of the overall wave function.

  20. In terms of a diagrammatic evaluation of the energetic property this is due to the presence of so-called unlinked diagrams. This analysis also leads to a number of formulae that can be used to correct for size consistency, such as the Davidson formula. See for instance Paldus (1981, 1992) for a detailed discussion of these issues.

  21. The issue of convergence of a perturbation series is commonly settled by considering the convergence radius of the series, which is hard to establish in detail for cases other than the most simple model cases. In practice, the issue of convergence tends to be settled on the basis of common knowledge about ‘where a method does well’. In practice, the convergence depends on the type of molecule, the type of bond, the internuclear distance and such like.

  22. Note that the remainder of this discussion only applies insofar philosophical criticisms of quantum chemistry are technically accurate. As one of the referees of this paper pointed out, this is not always the case.

  23. As an aside, it should in this context be noted that the abstraction of intermediate laws is also what drives the difference between Nagel (1961) and Kemeny and Oppenheim (1956) reduction. The latter claim that a Nagelian reduction scheme is a special case of a more general reduction scheme in which the intermediate laws can be eliminated to establish an (in the words of Kemeny and Oppenheim 1956) ‘indirect’ connection between two theories. The point then seems to be that if the indirectness of the connection can be established to sufficient degree, there is room to argue for significant explanatory independence at the higher theoretical level (in this case that of chemistry). Such explanatory independence can than take place in an entirely autonomous fashion (which seems to be the viewpoint advocated by van Brakel 2000), or in some intermediate fashion (which seems to be the viewpoint advocated by Woody 2000).

  24. I would like to thank one of the referees of this paper for pointing out these examples.

  25. The notion of ‘finalisation’ is used here in the context of the Starnbergers, where it means the ‘application’ phase of a research programme in the sense of Lakatos. The key distinguishing feature of a finalised programme is that its problem shifts are dictated by external, rather than internal, considerations.

References

  • Cohen-Tannoudji, C., Diu, B., Laloë, F.: Quantum Mechanics. Wiley. Translated from the French by Susan Reid Hemley, Nicole Ostrowsky and Dan Ostrowsky (1977)

  • Coulson, C.A., Haskins, P.J.: On the relative accuracies of eigenvalue bounds. J. Phys. B: Atom. Mol. Phys. 6, 1741–1750 (1973)

    Article  Google Scholar 

  • Courant, R., Hilbert, D.: Methoden Der Mathematischen Physik, Grundlehren der mathematischen Wissenschaften in einzeldarstellungen mit besonderer Beruücksichtigung der Anwendungsgebiete. J. Springer, Berlin (1931)

    Google Scholar 

  • Dalgarno, A., Stewart, A.L.: On the perturbation theory of small disturbances. Proc. R. Soc. Lond. A Math. Phys. Sci. 238(1213), 269–275 (1956)

    Google Scholar 

  • Dyall, K.G., Faegri, K.: Introduction to Relativistic Quantum Chemistry. Oxford University Press, Oxford (2007)

  • Feynman, R.P.: Forces in molecules. Phys. Rev. 56, 340 (1939)

    Article  Google Scholar 

  • Frost, A.A., Lykos, P.G.: Derivation of the Virial Theorem from the Generalized Hellmann-Feynman Theorem. J. Chem. Phys. 25, 1299–1300 (1956)

    Google Scholar 

  • Goldstein, H.: Classical Mechanics, 2nd edn. Addison Wesley Publishing, Reading (1980)

  • Helgaker, T., Jørgensen, P., Olsen, J.: Molecular Electronic Structure Theory. Wiley, London (2000)

    Google Scholar 

  • Hellmann, H.: Einführung in die Quantenchemie. F. Deuticke (1937)

  • Hempel, C.G.: Philosophy of Natural Sciences. Prentice Hall, Eaglewood Cliffs (1966)

    Google Scholar 

  • Hempel, C.G.: Provisoes: a problem concerning the inferential function of scientific theories. Erkenntnis. 28, 147–164 (1988)

    Article  Google Scholar 

  • Hettema, H.: The calculation of frequency dependent polarizabilities. Ph.D. thesis, Katholieke Universiteit Nijmegen (1993)

  • Hettema, H.: Is quantum chemistry a degenerating research programme? Log. Philos. Sci. VI(1), 3–23 (2008)

    Google Scholar 

  • Hirschfelder, J., Brown, W.B., Epstein, S.: Recent developments in perturbation theory. Adv. Quant. Chem. 1, 255–374 (1964)

    Article  Google Scholar 

  • Hylleraas, E. Über den Grundterm der Zweielektronenprobleme von H, He, Li+, Be++ usw. Zeitschrift für Physik 65(3), 209–225 (1930)

    Article  Google Scholar 

  • Kato, T.: Fundamental properties of Hamiltonian operators of Schrödinger type. Trans. Am. Math. Soc. 70, 195–211 (1951a)

    Article  Google Scholar 

  • Kato, T.: On the existence of solutions of the helium wave equation. Trans. Am. Math. Soc. 70, 212–218 (1951b)

    Article  Google Scholar 

  • Kemeny, J.G., Oppenheim, P.: On reduction. Philos. Stud. VII, 6–19 (1956)

    Article  Google Scholar 

  • Kuipers, T.A.F.: From Instrumentalism to Constructive Realism, Vol. 287 of Synthese Library. Kluwer Academic Publishers, Dordrecht (2000)

    Google Scholar 

  • Kuipers, T.A.F.: Structures in Science, Vol. 301 of Synthese Library. Kluwer Academic Publishers, Dordrecht (2001)

    Google Scholar 

  • McWeeny, R.: Coulson’s Valence. Oxford University Press, Oxford (1979)

    Google Scholar 

  • Nagel, E.: The Structure of Science: Problems in the Logic of Scientific Explanation. Routledge and Kegan Paul, London (1961)

    Google Scholar 

  • Nakatsuji, H.: Structure of the exact wave function. J. Chem. Phys. 113(8), 2949–2956 (2000)

    Article  Google Scholar 

  • Nola, R., Sankey, H.: Theories of Scientific Method : An Introduction. Stocksfield, Acumen (2007)

    Google Scholar 

  • Nooijen, M., Shamasundar, K.R., Mukherjee, D.: Reflections on size-extensivity, size-consistency and generalized extensivity in many-body theory. Mol. Phys. 103(15), 2277–2298 (2005)

    Article  Google Scholar 

  • Olsen, J., Jorgensen, P.: Linear and nonlinear response functions for an exact state and for an MCSCF state. J. Chem. Phys. 82, 3235–3264 (1985)

    Article  Google Scholar 

  • Paldus, J.: Diagrammatical Methods for Many-Fermion Systems. Lecture notes, University of Nijmegen (1981)

  • Paldus, J.: Coupled cluster theory. In: Wilson, S., Diercksen, G.H. (eds.) Methods in Computational Molecular Physics, vol. 293 of NATO ASI Series. Reidel. (1992)

  • Popper, K.: The Poverty of Historicism. Routledge, London (1957)

    Google Scholar 

  • Primas, H.: Pattern recognition in molecular quantum mechanics: I. Background dependence of molecular states. Theoretica Chimica Acta (Berlin) 39, 127–148 (1975)

    Article  Google Scholar 

  • Primas, H.: Chemistry, Quantum Mechanics and Reductionism, 2 edn. Springer, Berlin (1983)

    Google Scholar 

  • Scerri, E.: Popper’s naturalized approach to the reduction of chemistry. Int. Stud. Philos. Sci. 12, 33–44 (1998)

    Article  Google Scholar 

  • Scerri, E.R., McIntyre, L.: The case for the philosophy of chemistry. Synthese 111, 213–232 (1997)

    Article  Google Scholar 

  • Schlosshauer, M.: Decoherence and the Quantum-to-Classical Transition, The Frontiers Collection. Springer-Verlag, Berlin (2007)

    Google Scholar 

  • Sinanoglu, O.: Relation of perturbation theory to variation method. J. Chem. Phys. 34(4), 1237–1240 (1961)

    Article  Google Scholar 

  • Slater, J.C.: The virial and molecular structure. J. Chem. Phys. 1(10), 687–691 (1933)

    Article  Google Scholar 

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

    Google Scholar 

  • Van Vleck, J.H., Sherman, A.: The quantum theory of valence. Rev. Mod. Phys. 7(3), 167–228 (1935)

    Article  Google Scholar 

  • Weinhold, F.: Upper and lower bounds to quantum mechanical properties. Adv. Quant. Chem. 6, 299–331 (1972)

    Article  Google Scholar 

  • Wigner, E.P.: On a modification of the Rayleigh–Schrödinger perturbation theory. Matemat. és Természettud. Értesítö 53, 477–482 (1935)

    Google Scholar 

  • Woody, A.I.: Putting quantum mechanics to work in chemistry: the power of diagrammatic representation. Philos. Sci. 67, S612–S627. Supplement. Proceedings of the 1998 Biennial Meetings of the Philosophy of Science Association. Part II: Symposia Papers (2000)

  • Yarkony, D.R. (ed.): Modern Electronic Structure Theory. World Scientific, Singapore (1995)

    Google Scholar 

Download references

Acknowledgments

I would like to thank Paul Wormer, Josef Paldus and Bogomil Jeziorski for elucidating the somewhat confusing historical situation surrounding the original reference to Wigner’s 2n + 1 theorem and the interesting discussions via email about whether Wigner’s 1935 paper actually contains the (proof of the) theorem. I would also like to thank Paul Wormer and an anonymous referee of this journal for many helpful comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hinne Hettema.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Hettema, H. Explanation and theory formation in quantum chemistry. Found Chem 11, 145–174 (2009). https://doi.org/10.1007/s10698-009-9075-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10698-009-9075-8

Keywords

Navigation