Skip to main content
Log in

Chemistry and physics: no need for metaphysical glue

  • Published:
Foundations of Chemistry Aims and scope Submit manuscript

Abstract

Using the notorious bridge law “water is H2O” and the relation between molecular structure and quantum mechanics as examples, I argue that it doesn’t make sense to aim for specific definition(s) of intertheoretical or interdiscourse relation(s) between chemistry and physics (reduction, supervenience, what have you). Proposed definitions of interdiscourse and part-whole relations are interesting only if they provide insight in the variegated interconnected patchwork of theories and beliefs. There is “automatically” some sort of interdiscourse relation if different discourses claim to have something to say about the same situation (event, system), which is the basis of (contingent) local supervenience relations, which, proper empirically support being provided, can be upgraded to ceteris paribus bridge laws. Because of the ceteris paribus feature, and the discourse dependence of event identification, there is at best only global supervenience of the “special sciences” on the physical (and of parts of physics on other parts of physics).

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.

Similar content being viewed by others

Notes

  1. Cf. Harris’ (2008) arguments to the effect that “the chemical bond is best described as a theoretical synthesis or physico-chemical entity.” Duhem argued that a rigid distinction between physical and chemical properties cannot be made (Needham 2008).

  2. By definition a physical state gives a complete description of the system’s condition at a given time (given a discourse, e.g. microdiscourse, macrodiscourse). A physical system has a state, properties associated with that state (varying quantities and constants) and dynamic laws describing its evolution. Note already that talking this way requires that the boundary with respect to the environment can be defined.

  3. The distinction in physics between classical and quantum systems does not easily correlate with micro/macro either. A sugar molecule is classical. On the other hand there are macroscopic quantum states in superconductors.

  4. A (Quinean) web of belief in which distinctions such as analytic/synthetic or theoretical/empirical are contingent and ontology is always a matter of hypothetical “posits”.

  5. Among many other possible examples chemical engineering is an excellent example to illustrate how complex events are described and explained using many different discourses together.

  6. As a practical man Franklin wasn’t interested in molecules. He was doing the experiments to explore the use of oil to calm rough waters (Harte 1988, p. 5).

  7. It has been suggested in a number of publications that theoretical understanding at the level of solid state physics (e.g., concerning superconductivity) is independent of particles physics, although methods and models are shared and borrowed between these fields—though there is no consensus on this matter (of “epistemological reduction”). For a recent discussion see Howard (2007).

  8. See for statistical mechanics and the “themodynamic limit” and ensuing discussions on reduction and emergence Batterman (2005); for “substance” and quantum mechanics Loferer and Amann (2001).

  9. Citation from Dieks and de Regt (1998, p. 46).

  10. Because there are always interactions in the real world, Quantum Field Theory would not provide “grounds for regarding particle like entities as constituents of reality” (Fraser 2008).

  11. See Ladyman and Ross (2007, p. 4) for a rare example of a “naturalistic metaphysics” committed to the view that “the attempt to domesticate twenty-first-century science by reference to homely images of little particles … is forlorn.”

  12. For example, Kim (1998, p. 15) writes: “The bottom level is usually thought to consist of elementary particles, or whatever our best physics is going to tell us are the basic bits of matter out of which all material things are composed. As we go up the ladder, we successively encounter atoms, molecules, cells, larger living organisms, and so on.”

  13. Of course atomism is taken for granted in particle physics. Noble laureate Martinus Veltman (2003, p. 8, cf. 35) writes: “All matter is made up from molecules, and molecules are bound states of atoms. For example, water consists of water molecules which are bound states of one oxygen and two hydrogen atoms. This state of affairs is reflected in the chemical formula H2O.”

  14. Depending on how one counts; an alternative number is 48. As the theory still contains 18 (or 17) “arbitrary” constants (which have to be measured) one may expect that future developments will have to introduce more or different entities (or drop the notion of particle completely).

  15. According to Veltman (2003), strictly speaking, “the concept of a force makes no sense” (71). “Particles appear and disappear and all properties of matter derive from the properties of the constituent particles.” (38) It is said that “electron–phonon” interaction explains electron–electron interaction; so what does explain electron–phonon interaction?

  16. A chronon is a proposed quantum of time, that is, a discrete and indivisible “unit” of time. The minimum measurable period of time would be 1.078 × 10−43 s. This time scale is too small to be noticeable on the space–time scale of humans, so the universe appears very continuous even though it is happening in discrete segments of time. But it may play a role on scales much smaller than the nucleus of an atom, and on much larger scales when a lot of mass is concentrated in a small space, like in a gravastar or a black hole. The chronon might assist in combining quantum mechanics with general relativity to produce a theory of quantum gravity. Actually nobody has the slightest idea how to fit in gravity.

  17. That molecules do not occur with a definite shape according to quantum mechanics, quite in contrast to what is believed in chemistry, was already noted by Hund (1927): “The fact that the right-handed or left-handed configuration of a molecule is not a quantum state might appear to contradict the existence of optical isomers.” Isomeric molecules are described by the same Hamiltonian.

  18. Some will say that electrons (or nuclei) have individual existence in a molecule, but not individual states. If lower level particles are indistinguishable how are we going to make sense of ascribing causal powers to individual particles?

  19. Another way of putting this is saying that contexts are at least as important as first principles; the latter being necessary conditions, not sufficient conditions (Atmanspacher and Bishop 2006). In this context “contexts” are contingent conditions referring to the degree of abstraction at which a theoretical framework is formulated. Each description requires “abstracting from,” meaning disregarding those details of a given system (and its environment) which are to be considered irrelevant. In applying quantum mechanics or in general any universal law under particular initial and boundary conditions, one needs global constraints in order to obtain well-formed problems and to avoid degeneracies (multiplicity of possible solutions).

  20. The environment would have an effect similar to the collapse of the wave function on measurement. Perhaps one could say: the environment is monitoring the system as if continuous measurement is taking place.

  21. Of course opinions on the (philosophical) significance of decoherence differ.

  22. Definitions (Kim 1984): S weakly supervenes on B iff necessarily any two things that have the same properties in domain B have the same properties in domain S (that is, B-indiscernibility entails S-indiscernibility). S globally supervenes on B iff any two worlds that are B-indiscernible are also S-indiscernible. Supervenience as defined by Kim is not a part-whole relation, even though in earlier publications he has referred to it as “mereological supervenience.” More on global supervenience below.

  23. Of course one can give definitions of emergence such that irreducibility does not entail emergence; for an example see Rueger (2006). Nowadays the use of words like “emergence” may be more common in the philosophy of physics, or even physics itself, than in the philosophy of chemistry. Physicists are used to saying things like “renormalizability is an emergent property of ordinary matter” (Laughlin and Pines 2000, p. 29).

  24. In general, there is no neat correlation between temperature and energy, not even for individual substances (Needham 2009).

  25. Including bridge laws without “level” distinction; for example the relation between gravitational and inertial mass.

  26. The issue of causal completeness of the physical is relevant in the discussion on interdiscourse relations, but has no bearing on the question of part-whole emergence or reduction in physics (Hüttemann and Papineau 2005).

  27. For example: The structure of a molecule is given by the average positions of the nuclei, which is determined by the potential minima, fixed in turn by the probability position distribution for the system’s electrons, which is given by the non-resultant (configurational) Hamiltonian.

  28. Of course the same distinction can be made for interdiscourse relations. Hüttemann associates the micro-explanation of states with explaining determinate properties and the micro-explanation of laws with determinable properties.

  29. Entanglement is sometimes used as a general argument for saying that it is impossible to derive molecular structure from quantum mechanical first principles (Atmanspacher and Bishop 2006). Even supervenience as normally understood fails: Because there are no states of the parts, there is nothing for the state of the compound to supervene upon (Humphreys 1997).

  30. Alternative formulation: The behaviour of macrophysical wholes is asymmetrically determined by the behaviour of their microphysical parts plus general laws that do not apply only to wholes of that kind (Hüttemann and Papineau 2005).

  31. Alternative formulation: “Only resultant wholes have contemporaneous parts, emergent wholes do not.” (346).

  32. For a recent discussion on modifying Nagelian reduction to include the idea that in physics typically the “finer” theory reduces to the “coarser” theory in an approximate limit, see Batterman (2005).

  33. Of course this epistemic virtue is already tacitly invoked in discussions about definitions of reduction and emergence. For example, Hüttemann and Terzidis (2000, p. 279) argue that neither Broad’s concept of emergence nor Kim’s concept of emergence “is able to divide the class of physical systems into interesting subclasses.”

  34. Nagel (1961, p. 372) has used the word “postulate” for bridge laws.

  35. Such as: “density is mass divided by volume”, “force is mass times acceleration”, “extensive quantities are additive.”

  36. Following Ostrovsky (2005) who has stressed: No physical theory can be blamed for using approximations, because every theory does this. We should not think of approximations as something to be overcome. They come with the models, and the models give understanding, which calculation does not.

  37. Not all physicists like Cartwright’s account of physics; see e.g., Anderson (2001a, b, pp. 492, 500) and Cartwright (2001, p. 496).

  38. Is there a bridge law connecting the boiling of water and the phase transition of ensembles of H2O-molecules? How does the latter whole relate to its parts?

  39. Although the major difference is in the ability to absorb or emit high-energy radiation (hence its application in nuclear reactors), having a pH of 7.41 (instead of 7.00) may be considered as a significant chemically relevant difference. The operational criterion of “same substance” clearly suggests that if we set the standards of purity high enough, ordinary water and heavy water are different substances.

  40. The two isomers were first described in some detail (together with heavy water) in a book of Farcas (1935). The para modification’s complete insensitivitty to the magnetic field could be of use in magnetic resonance imaging (Tikhnonov and Volkov 2002).

  41. Schewe et al. (2003): A water molecule’s chemical formula is really not H2O, at least from the perspective of neutrons and electrons interacting with the molecule for only attoseconds (less than 10−15 s).

  42. See in particular Hornsby (1997).

  43. Both micro- and macro-properties might be multiply realised. In most supervenience discussions asymmetric dependence is presupposed. Because the nature of the dependence needs to be explained to make a difference (between different metaphysical models), the distinction between supervenience and ceteris paribus bridge laws collapses.

  44. Davidson (1993, p. 4). Elsewhere I have argued that the common understanding of Davidson’s (or Hare’s) notion of supervenience as a specimen of what Kim called “weak supervenience” is incorrect and completely misses the point of Davidson’s (externalist) anomalous monism cum supervenience (van Brakel 1999). The present paper provides some ingredients of what I have called “anomalous anomalous monism”. This metaphysical model differs from Davidson’s anomalous monism in dropping the assumption of physics ideally being concerned with strict laws (providing causal closure). According to Davidson, any S(pecial science)-event-token is co-extensive with some P(hysics)-event-token, though there are no limitations on the holistic nature of the S-domain, to which I add: “nor the holistic nature of the P-domain”.

  45. The phenomenon of entanglement mentioned earlier might be invoked to argue that even global supervenience fails.

  46. A metaphysical model is underdetermined by the practice of science. It has to meet the meta-epistemic virtue of “fitting in” (Goodman 1978, p. 132), scoring well on the criteria “empirical adequacy” and “coherence.”

  47. In a detailed study of three case studies of (generally accepted cases of) unification, Morrison (2000) has found a bewildering variety of uses of the words ‘unity’ and ‘unification’. Different degrees of reduction, synthesis and integration yield distinct ways in which phenomena can be united under a common theoretical framework. Not all forms of theory unification depend on isomorphism or reduction. Often unification is at odds with explanation. Unification is not sufficient for theory acceptance. It is just one of the many vague epistemic virtues that can play a role in theory acceptance, but it is not a determining factor. Even within the framework of the localised settings of specific theories different kinds of unity emerge. Of course the unity of knowledge also has its defenders; e.g. Wilson (1998), Anderson (2001a, b).

  48. See Lauglin and Pines (2000). As Redhead (1991) says: “It is generally agreed that the present state of the universe is the outcome of a series of randomly occurring symmetry breakings, which means that the TOE [Theory Of Everything] cannot itself predict many of the universe’s essential features as we currently observe it—a very important limitation on what TOEs can be expected to deliver, even in principle.” Even if we are prepared to accept the idea of a Theory of Everything, the possibility that absolute chance is a factor in the universe cannot be excluded (van Brakel 1991).

References

  • Anderson, P.W.: Reply to Cartwright. Stud. Hist. Philos. Mod. Phys. 32, 499–500 (2001a). doi:10.1016/S1355-2198(01)00013-2

  • Anderson, P.W.: Science: a “dappled world” or a “seamless web”. Stud. Hist. Philos. Mod. Phys. 32, 487–494 (2001b). doi:10.1016/S1355-2198(01)00011-9

  • Atmanspacher, H., Bishop, R.: Contextual emergence in the description of properties. Found. Phys. 36, 1753–1777 (2006). doi:10.1007/s10701-006-9082-8

    Article  Google Scholar 

  • Batterman, R.W.: Critical phenomena and breaking drops: infinite idealizations in physics. Stud. Hist. Philos. Mod. Phys. 36, 225–244 (2005). doi:10.1016/j.shpsb.2004.05.004

    Article  Google Scholar 

  • Belyaev, V.B., et al.: Search for nuclear reactions in water molecules. Phys. Lett. B 522, 222–226 (2001). doi:10.1016/S0370-2693(01)01260-6

    Article  Google Scholar 

  • Broad, C.D.: The mind and its place in nature. Kegan Paul, London (1925)

    Google Scholar 

  • Cartwright, N.: The dapppled world: essays on the perimeter of science. Cambridge University Press, Cambridge (1999)

    Google Scholar 

  • Cartwright, N.: Reply to anderson. Stud. Hist. Philos. Mod. Phys. 32, 495–497 (2001). doi:10.1016/S1355-2198(01)00012-0

    Article  Google Scholar 

  • Chalmers, D.K.: Strong and weak emergence. In: Clayton, P., Davies, P. (eds.) The re-emergence of emergence, pp. 244–256. Oxford University Press, Oxford (2006)

    Google Scholar 

  • Davidson, D.: Thinking causes. In: Heil, J., Mele, A. (eds.) Mental causation, pp. 3–18. Clarendon Press, Oxford (1993)

    Google Scholar 

  • Dieks, D., de Regt, H.W.: Reduction and understanding. Found. Sci. 1, 45–59 (1998). doi:10.1023/A:1009630119534

    Article  Google Scholar 

  • Farcas, A.: Orthohydrogen, parahydrogen and heavy hydrogen. Cambridge University Press, Cambridge (1935)

    Google Scholar 

  • Fraser, D.: The fate of ‘particles’ in quantum field theories with interactions. Stud. Hist. Philos. Mod. Phys. (2008). doi:10.1016/j.shpsb.2008.05.003

  • Goodman, N.: Ways of worldmaking. Hackett, Indianapolis (1978)

    Google Scholar 

  • Harris, M.L.: Chemical reductionism revisited: lewis, pauling and the physico-chemical nature of the chemical bond. Stud. Hist. Philos. Sci. 39, 78–90 (2008). doi:10.1016/j.shpsa.2007.11.006

    Article  Google Scholar 

  • Harte, J.: Consider a spherical cow. Wh Freeman, New York (1988)

    Google Scholar 

  • Hornsby, J.: Simple mindedness: in defence of naive naturalism in the philosophy of mind. Harvard University Press, Cambridge (1997)

    Google Scholar 

  • Howard, D.: Reduction and emergence in the physical sciences: some lessons from the particle physics and condensed matter debate. In: Murphy, N., Stoeger, W.R. (eds.) Evolution and emergence. Oxford University Press, Oxford (2007)

    Google Scholar 

  • Humphreys, P.: How properties emerge. Philos. Sci. 64, 1–17 (1997). doi:10.1086/392533

    Article  Google Scholar 

  • Hund, F.: On the interpretation of molecular spectra I. Z. Phys. 43, 805–826 (1927). doi:10.1007/BF01400234

    Article  Google Scholar 

  • Hüttemann, A.: What’s wrong with microphysicalism. Routledge, London/New York (2004)

    Google Scholar 

  • Hüttemann, A.: Explanation, emergence, and quantum entanglement. Philos. Sci. 72, 114–127 (2005). doi:10.1086/428075

    Article  Google Scholar 

  • Hüttemann, A., Papineau, D.: Physicalism decomposed. Analysis 65(1), 33–39 (2005). doi:10.1111/j.1467-8284.2005.00519.x

  • Hüttemann, A., Terzidis, O.: Emergence in physics. Int. J. Philos. Sci. 14, 267–281 (2000). doi:10.1080/026985900437773

    Article  Google Scholar 

  • Joos, E.: Decoherence through interaction with the environment. In: Giulini, D. (ed.) Decoherence, the appearance of a classical world in quantum theory. Springer, Berlin (1996)

    Google Scholar 

  • Joos, E.: The emergence of classicality from quantum theory. In: Clayton, P., Davies, P. (eds.) The re-emergence of emergence, pp. 53–78. Oxford University Press, Oxford (2006)

    Google Scholar 

  • Kim, J.: Concepts of supervenience. Philos. Phenomenol. Res. 45, 153–176 (1984). doi:10.2307/2107423

    Article  Google Scholar 

  • Kim, J.: Mind in a physical world. MIT Press, Cambridge (1998)

    Google Scholar 

  • Kronz, F.M., Tiehen, J.T.: Emergence and quantum mechanics. Philos. Sci. 69, 324–347 (2002). doi:10.1086/341056

    Article  Google Scholar 

  • Ladyman, J., Ross, D.: Every thing must go. Oxford University Press, Oxford (2007)

    Book  Google Scholar 

  • Laughlin, R.B., Pines, D.: The theory of everything. Proc. N. Y. Acad. Sci. 97, 28–31 (2000). doi:10.1073/pnas.97.1.28

    Article  Google Scholar 

  • Loferer, M.J., Amann, A.: The problem of substance in chemistry–an approach by means of large deviation statistics. Cryst. Eng. 4, 101–111 (2001). doi:10.1016/S1463-0184(01)00008-9

    Article  Google Scholar 

  • Morrison, M.: Unifying scientific theories. Physical concepts and mathematical structures. Cambridge University Press, Cambridge (2000)

    Book  Google Scholar 

  • Nagel, E.: The structure of science. Routledge and Kegan Paul, London (1961)

    Google Scholar 

  • Needham, P.: Is water a mixture? Bridging the distinction between physical and chemical properties. Stud. Hist. Philos. Sci. 39, 66–77 (2008). doi:10.1016/j.shpsa.2007.11.005

    Article  Google Scholar 

  • Needham, P.: Reduction and emergence: a critique of Kim. Philos. Stud. 146, 93–116 (2009)

    Article  Google Scholar 

  • Oppenheim, P., Putnam, H.: The unity of science as a working hypothesis. Minn. Stud. Philos. Sci. 2, 3–36 (1958)

    Google Scholar 

  • Ostrovsky, V.J.: Towards a philosophy of approximation in the ‘exact’ sciences. Hyle. Int. J. Philos. Sci. 11, 101–126 (2005)

    Google Scholar 

  • Primas, H.: Chemistry, quantum mechanics and reductionism: perspectives in theoretical chemistry. Springer, Berlin (1983)

    Google Scholar 

  • Putnam, H.: The meaning of ‘Meaning’. In: Mind, language and reality: philosophical papers. Vol. 2. Cambridge University Press, Cambridge (1975)

  • Quine, W.V.: Otherworldly. New York Rev. Books (November):23 (1978)

  • Ramsey, J.L.: Molecular shape, reduction, explanation and approximate concepts. Synthese 111, 233–251 (1997). doi:10.1023/A:1004901931804

    Article  Google Scholar 

  • Redhead, M.: Review of J.D. Barrow’s theories of everything. Times Lit. Suppl. 26, (1991)

  • Rueger, A.: Functional reduction and emergence in the physical sciences. Synthese 151, 335–346 (2006). doi:10.1007/s11229-006-9027-y

    Article  Google Scholar 

  • Schewe, P.F., Stein, B., Riordon, J.: A water molecule’s chemical formula is really not H2O. Physics News Update [The American Institute of Physics Bulletin of Physics News] 648 (July 31) (2003)

  • Tikhonov, V.E., Volkov, A.A.: Separation of water into its ortho and para isomers. Science. 296, 2363 (2002). doi:10.1126/science.1069513

  • van Brakel, J.: The limited belief in chance. Stud. Hist. Philos. Sci. 22, 499–513 (1991). doi:10.1016/0039-3681(91)90006-E

    Google Scholar 

  • van Brakel, J.: Interdiscourse or supervenience relations: the priority of the manifest image. Synthese 106, 253–297 (1996). doi:10.1007/BF00413703

    Google Scholar 

  • van Brakel, J.: Supervenience and anomalous monism. Dialectica 53, 3–25 (1999)

    Google Scholar 

  • van Fraassen, B.C.: The empirical stance. Yale University Press, New Haven (2002)

    Google Scholar 

  • van Fraassen, B.C.: From vicious circle to infinite regress, and back again. PSA 2, 6–29 (1992)

    Google Scholar 

  • Veltman, M.: Facts and mysteries in elementary particle physics. World Scientific, New Jersey (2003)

    Book  Google Scholar 

  • Wilson, E.O.: Consilience: the unity of knowledge. Knopf, New York (1998)

    Google Scholar 

  • Woolley, R.G.: Must a molecule have a shape? J. Am. Chem. Soc. 100, 1073–1078 (1978). doi:10.1021/ja00472a009

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to J. van Brakel.

Rights and permissions

Reprints and permissions

About this article

Cite this article

van Brakel, J. Chemistry and physics: no need for metaphysical glue. Found Chem 12, 123–136 (2010). https://doi.org/10.1007/s10698-010-9084-7

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10698-010-9084-7

Keywords

Navigation