Skip to main content

Reduction for a Dappled World: Connecting Chemical and Physical Theories

  • Chapter
  • First Online:
  • 2286 Accesses

Part of the book series: Boston Studies in the Philosophy and History of Science ((BSPS,volume 306))

Abstract

In this paper I argue that there are good reasons to construct the connection between theories of chemistry and theories of physics in terms of a Nagelian reductionist model. I use the machinery of belief revision to investigate the reduction relation in terms of a structuralist and conceptual space approach. An important aspect of the current paper is that it has the potential to unite reductive and non-reductive views on science.

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

Buying options

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
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

Learn about institutional subscriptions

Notes

  1. 1.

    The resulting disconnects have led some to question the role of physics and theory in chemistry, for instance, in the paper by Hoffmann (2007).

  2. 2.

    See for instance Woody (2000), Scerri (1998) and Needham (2010) as examples.

  3. 3.

    And separates them from a number of informal conditions which he also specifies in great detail.

  4. 4.

    See for instance Kuipers (1990) for an example of reductions from many sciences, most of which are not based on strict identities.

  5. 5.

    For instance, Nagel (1961) discussed three kinds of linkages postulated by reduction postulates

    1. 1.

      The links are logical connections, such that the meaning of ‘A’ as ‘fixed by the rules or habits of usage’ is explicable in terms of the established meanings of the theoretical primitives in the primary discipline.

    2. 2.

      The links are conventions or coordinating definitions, created by ‘deliberate fiat’, which assigns a meaning to the term ‘A’ in terms of the primary science, subject to a criterion of consistency with other assignments.

    3. 3.

      The links are factual or material, or physical hypotheses, and assert that existence of a state ‘B in the primary science is sufficient (or necessary and sufficient) condition for the state of affairs designated by ‘A’. In this scenario, the meanings of ‘A’ and ‘B’ are not related analytically.

  6. 6.

    A similar point was made in a somewhat neglected paper by Horgan (1978), who argues that the reduction postulates supervene on the reducing theory.

  7. 7.

    See Balzer and Sneed (1977, 1978) for an introduction of this relation and the corresponding notion of a theory net.

  8. 8.

    In addition, belief revision has been introduced into the structuralist model by Enqvist (2011). Enqvist develops a highly specific alternative to the notion of ‘reduction postulates’ qua ‘linking commitments’ which I developed in Hettema (2012a). Enqvist’s construction relies on a construction of specialisation theory nets, to which he applies the AGM belief revision strategies. Enqvist does not fully develop the AGM theory in a structuralist model, and ignores the stratification between theoretical / non-theoretical levels of the theory. In general, developing complex notions in the stratified model adds complications which are usually ignored in the first ‘step’ of the development of such models, see for instance the development of truthlikeness developed by Kuipers (1992).

  9. 9.

    As the book by Nye (2011) illustrates, this theory was sometimes jokingly referred to as the ‘absolute’ theory of reaction rates. Many of his contemporaries found Eyring’s ideas too radical, as the proceedings of the 1937 workshop at the University of Manchester illustrate. In my earlier paper (Hettema (2012b)) I had this wrong, and used the designation of ‘absolute theory’ throughout. At the time I was unaware of the earlier ironic use, and thought that ‘absolute theory of reaction rates’ was a neater choice to designate the theory than the somewhat more clumsy sounding ‘theory of absolute reaction rates’. Of course, I now see the error of my ways.

  10. 10.

    The notion of a (reactive) potential energy surface for the nuclear motion is key to the development of the theory. While the idea was introduced by Born and Heisenberg (1924) and Born and Oppenheimer (1927) it may be argued that the idea of a potential energy surface only reached its full fruition with the development of a theory of chemical reaction rates. The idea of a potential energy surface is part of Wigner’s ‘three threes’ (see below).

  11. 11.

    Especially illustrative for this is the motivation Eyring (1938) gave for his introduction of various ‘semi-empirical’ methods in quantum chemistry, which lead to various inconsistencies between these semi-empirical theories and quantum theory.

    For the purposes of calculating the potential energy surface for a chemical reaction, Eyring first classifies theories as ‘semi-empirical’ when they have the following characteristics:

    (a) that each electron can be assigned a separate eigenfunction which involves the coordinates of only this one electron. (b) Multiple exchange integrals are negligible, (c) Normalising integrals for overlapping orbitals are negligible in comparison with unity. (d) The exchange and coulombic integrals for a complicated molecular system may be estimated from a potential curve for the isolated pair of atoms. (e) For distances involved in activation energy calculations this percentage is around 20 per cent. coulombic and 80 per cent. exchange binding, and this varies but little from atom pair to atom pair. (Eyring 1938, p. 8)

    Eyring then remarks that more detailed calculations, as well as principled considerations, give no support for the construction of these theories:

    None of these assumptions have been rigorously derived from theory, and, as has been emphasised by Coolidge and James, if one assumes for H3, the approximate eigenfunctions used by Heitler and London and Sugiura for H2, the assumptions can all be shown to fail badly. (Eyring 1938, p. 8)

    Thus stated, these sort of theories seem to be counterexamples to a theory of reduction: the sort of reduction that derives chemical ‘laws’ directly from basic quantum theory can only be achieved on the basis of theoretical assumptions that are unjustified from the viewpoint of basic theory and which can moreover be shown up as factually wrong in a large number of practical cases.

  12. 12.

    The article appears in translated form in Back and Laidler (1967).

  13. 13.

    A modified collision theory often introduces a ‘probability factor’ P which measures the probability that a collision will lead to a completed chemical reaction. Hence, in the modified collision theory

    $$ \mathrm{k}=PZ\; \exp \left(\frac{-E}{RT}\right) $$
    (2.3)

    The ‘fudge factor’ P is introduced since the collision cross section of a molecule bears no clear relationship to the probability for a chemical reaction. While the collision theory works well for reactions between mono-atomic gases, it breaks down for reactions between more complex molecules. In this respect, the collision theory is not capable of clarifying the internal mechanisms of chemical reactions in the necessary detail.

  14. 14.

    Wigner refers to the theory in this paper as ‘The Transition State Method’. The paper by Laidler and King (1983) contains a brief discussion of this conference and the role it played in the subsequent adoption of the theory.

  15. 15.

    A detailed discussion of why this is so falls outside the scope of this paper, but can easily be determined by stepping through the mathematics.

References

  • Back MH, Laidler KJ (1967) Selected readings in chemical kinetics, 1st edn. Pergamon Press, Oxford

    Google Scholar 

  • Balzer W, Sneed JD (1977) Generalized net structures of empirical theories: I. Stud Logica 36:195–211

    Article  Google Scholar 

  • Balzer W, Sneed JD (1978) Generalized net structures of empirical theories: II. Stud Logica 37:167–194

    Article  Google Scholar 

  • Balzer W, Moulines CU, Sneed J (1987) An architectonic for science: the structuralist program. Reidel, Dordrecht

    Book  Google Scholar 

  • Bokulich A (2008) Reexamining the quantum-classical relation: beyond reductionism and pluralism. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Born M, Heisenberg W (1924) Zur Quantentheorie der Molekeln. Ann Phys 74:1–31

    Article  CAS  Google Scholar 

  • Born M, Oppenheimer R (1927) Zur Quantentheorie der Molekeln. Ann Phys 84:457–484

    Article  CAS  Google Scholar 

  • Cartwright N (1983) How the laws of physics lie. Oxford University Press, Oxford

    Book  Google Scholar 

  • Cartwright N, Cat J, Fleck L, Uebel T (1996) Otto Neurath: philosophy between science and politics. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Causey RC (1977) Unity of science. Reidel, Dordrecht

    Book  Google Scholar 

  • Dizadji-Bahmani F, Frigg R, Hartmann S (2010) Who’s afraid of Nagelian reduction? Erkenntnis 73:393–412. doi:10.1007/s10670-010-9239-x

    Article  Google Scholar 

  • Dupré J (1993) The disorder of things: metaphysical foundations of the disunity of science. Harvard University Press, Cambridge, MA

    Google Scholar 

  • Enqvist S (2011) A structuralist framework for the logic of theory change. In: Olsson EJ, Enqvist S (eds) Belief revision meets philosophy of science. Logic, Epistemology, and the Unity of Science, vol 21. Springer, Netherlands, pp 105–135

    Google Scholar 

  • Evans MG, Polanyi M (1935) Some applications of the transition state method to the calculation of reaction velocities, especially in solution. Trans Faraday Soc 31:875–894

    Article  CAS  Google Scholar 

  • Eyring H (1935) The activated complex in chemical reactions. J Chem Phys 3(2):107

    Article  CAS  Google Scholar 

  • Eyring H (1938) The calculation of activation energies. Trans Faraday Soc 34:3–11

    Article  CAS  Google Scholar 

  • Eyring H, Walter J, Kimball GE (1944) Quantum chemistry. Wiley, New York

    Google Scholar 

  • Fazekas P (2009) Reconsidering the role of bridge laws in inter-theoretical reductions. Erkenntnis 71:303–322

    Article  Google Scholar 

  • Gärdenfors P, Zenker F (2011) Using conceptual spaces to model the dynamics of empirical theories. In: Olsson EJ, Enqvist S (eds) Belief revision meets philosophy of science. Logic, epistemology, and the unity of science, vol 21. Springer, Netherlands, pp 137–153

    Google Scholar 

  • Glasstone S, Laidler KJ, Eyring H (1941) The theory of rate processes: the kinetics of chemical reactions, viscosity, diffusion and electrochemical phenomena. McGraw-Hill, New York

    Google Scholar 

  • Hendry RF (2006) Is there downward causation in chemistry? In: Baird D, Scerri E, McIntyre L (eds) Philosophy of chemistry. Boston studies in the philosophy of science, vol 242. Springer, Netherlands, pp 173–189

    Google Scholar 

  • Hettema H (2012a) Reducing chemistry to physics: limits, models, consequences. Createspace

    Google Scholar 

  • Hettema H (2012b) The unity of chemistry and physics: absolute reaction rate theory. Hyle 18:145–173

    Google Scholar 

  • Hoffmann R (2007) What might philosophy of science look like if chemists built it? Synthese 155(3):321–336

    Article  Google Scholar 

  • Horgan T (1978) Supervenient bridge laws. Philos Sci 45(2):229–249

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Klein C (2009) Reduction without reductionism: a defence of Nagel on connectability. Philos Q 59(234):39–53

    Article  Google Scholar 

  • Kuhn TS (1976) Theory-change as structure-change: comments on the sneed formalism. Erkenntnis 10:179–199

    Article  Google Scholar 

  • Kuipers TAF (1990) Reduction of laws and concepts. In: Brzezinski J, Coniglione F, Kuipers T (eds) Idealization I: general problems II: forms and applications. Poznan studies in the philosophy of the sciences and the humanities, vol 16. Rodopi, Amsterdam, pp 241–276

    Google Scholar 

  • Kuipers TAF (1992) Naive and refined truth approximation. Synthese 93(3):299–341

    Article  Google Scholar 

  • Kuipers TAF (2007) Laws, theories and research programs. In: Kuipers T (ed) General philosophy of science – focal issues. Handbook of the philosophy of science. Amsterdam, London, North Holland, pp 1–95

    Google Scholar 

  • Laidler KJ, King MC (1983) Development of transition-state theory. J Phys Chem 87(15):2657–2664

    Article  CAS  Google Scholar 

  • Lakatos I (1970) The methodology of scientific research programmes. In: Musgrave A, Lakatos I (eds) Criticism and the growth of knowledge. Cambridge University Press, Cambridge

    Chapter  Google Scholar 

  • Lombardi O, Labarca M (2005) The ontological autonomy of the chemical world. Found Chem 7:125–148

    Article  CAS  Google Scholar 

  • Miller WH (1998) Quantum and semiclassical theory of chemical reaction rates. Faraday Discuss 110:1–21

    Article  CAS  Google Scholar 

  • Muller FA (1998) Structures for everyone: contemplations and proofs in the foundations and philosophy of physics and mathematics. Gerits & Son, Amsterdam

    Google Scholar 

  • Muller FA (2003) Refutability revamped: how quantum mechanics saves the phenomena. Erkenntnis 58(2):189–211

    Article  Google Scholar 

  • Nagel E (1961) The structure of science: problems in the logic of scientific explanation. Routledge and Kegan Paul, London

    Google Scholar 

  • Needham P (2010) Nagel’s analysis of reduction: comments in defense as well as critique. Stud Hist Phil Biol Sci Part B Stud Hist Philos Mod Phys 41(2):163–170

    Article  Google Scholar 

  • Nye MJ (2011) Michael Polanyi and his generation: origins of the social construction of science. University of Chicago Press, Chicago

    Book  Google Scholar 

  • Potochnik A (2011) A neurathian conception of the unity of science. Erkenntnis 74(3):305–319

    Article  Google Scholar 

  • Priest G (2008) An introduction to non-classical logic: from if to is, Cambridge introductions to philosophy, 2nd edn. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Reichenbach H (1937) Experience and prediction. University of Chicago Press, Chicago

    Google Scholar 

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

    Article  Google Scholar 

  • Schaffner KF (1967) Approaches to reduction. Philos Sci 34(2):137–147

    Article  Google Scholar 

  • Sklar L (1967) Types of inter-theoretic reduction. Br J Philos Sci 18(2):109–124

    Article  Google Scholar 

  • Sneed JD (1971) The logical structure of mathematical physics. Reidel, Dordrecht

    Google Scholar 

  • Suppes P (1957) Introduction to logic. Van Nostrand, New York

    Google Scholar 

  • Travers M (1938) Reaction kinetics, a general discussion. Trans Faraday Soc 34:1–2

    Article  Google Scholar 

  • van Riel R (2011) Nagelian reduction beyond the Nagel model. Philos Sci 78(3):353–375

    Article  Google Scholar 

  • Wigner E (1938) The transition state method. Trans Faraday Soc 34:29–41

    Article  CAS  Google Scholar 

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

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hinne Hettema .

Editor information

Editors and Affiliations

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer Science+Business Media Dordrecht

About this chapter

Cite this chapter

Hettema, H. (2015). Reduction for a Dappled World: Connecting Chemical and Physical Theories. In: Scerri, E., McIntyre, L. (eds) Philosophy of Chemistry. Boston Studies in the Philosophy and History of Science, vol 306. Springer, Dordrecht. https://doi.org/10.1007/978-94-017-9364-3_2

Download citation

Publish with us

Policies and ethics