Skip to main content
Log in

Infinitesimal idealization, easy road nominalism, and fractional quantum statistics

  • S.I.: Infinite Idealizations in Science
  • Published:
Synthese Aims and scope Submit manuscript

Abstract

It has been recently debated whether there exists a so-called “easy road” to nominalism. In this essay, I attempt to fill a lacuna in the debate by making a connection with the literature on infinite and infinitesimal idealization in science through an example from mathematical physics that has been largely ignored by philosophers. Specifically, by appealing to John Norton’s distinction between idealization and approximation, I argue that the phenomena of fractional quantum statistics bears negatively on Mary Leng’s proposed path to easy road nominalism, thereby partially defending Mark Colyvan’s claim that there is no easy road to nominalism.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

From Stern (2008, p. 207)

Similar content being viewed by others

Notes

  1. Looking ahead, I do not intend to overstate my case. Some connection have already been made, e.g., Maddy (1997), Leng (2010), Baron (2016) and De Bianchi (2016). Nevertheless, the case study that I consider, the relevance of Norton’s distinction between approximation and idealization, and the particular problems posed by infinite/infinitesimal idealizations for issues in philosophy of mathematics have been largely ignored to my best knowledge.

  2. E.g., Colyvan (2010, 2012a), Azzouni (2012), Bueno (2012), Leng (2012), Liggins (2012) and Yablo (2012). Also see even more recent essays in the special issue Molinini et al. (2016).

  3. See Field (1980) for such an attempt.

  4. See Putnam (1971) and Quine (1981), and see Colyvan (2001) for an in-depth study of the indispensability argument and a defense.

  5. E.g., Azzouni (2004), Balaguer (1998), Melia (2000, 2002) and Yablo (1998, 2002).

  6. Mainly because of numerous examples that do seem to provide genuine mathematical explanation, e.g., Baker (2005, 2009, 2012), Colyvan (2001, 2007, 2010). Moreover, the example discussed in this paper purports to be such a genuine mathematical explanation.

  7. Many are unconvinced by the EIA and I would like to say a bit about why some may feel the argument is compelling. Science is in the business of giving scientific explanations, but it is hard to see how something can be explanatory but not exist. For example, on a causal-mechanistic account of explanation (e.g., Salmon 1984), it would seem absurd if one were told that a particular gene mutation explains some disease but that the same gene doesn’t exist. Similarly, on a covering law account of explanation (Hempel and Oppenheim 1948) it seems unreasonable to explain the motion of planets given Newton’s universal law of gravitation but then claim that there is no such thing as the universal law of gravitation. In the same manner, Platonists like Colyvan (2001, 2007, 2010) and Baker (2005, 2009, 2012) think that there is something seriously wrong with denying the existence of, say, prime numbers, but then arguing that the primeness of thirteen and seventeen (along with a number of evolutionary hypothesis) explains why the North American periodical magicicada cicadas only appear every 13 or 17 years. Of course, there are others who think that fiction can be explanatory, e.g. Bokulich (2008). In any case, in the context of this paper I will assume that the EIA is convincing enough to take seriously and ask whether Leng’s (2005, 2010, 2012) approach to evade the arguments works.

  8. In hope to thwart an objection early, note that I’m only partially defending Colyvan (2010) since there are other approaches to easy road nominalism, e.g., Azzouni (2012), Bueno (2012), Liggins (2012) and Yablo (2012), which purport to make sense of mathematical explanation. Also recall that the overarching goal of this paper is to make interesting connections between infinite and infinitesimal idealization in science and the easy road nominalism debate—it is not to mount a comprehensive defense of Colyvan’s (2010) thesis.

  9. There may be an interesting difference between infinite and infinitesimal idealizations but for the purpose of this paper I’m assuming that both idealizations are on equal footing, so to speak. Note that, insofar as an idealization enters physical theory via a limiting procedure, an infinite idealization is just the inverse of an infinitesimal idealization and vice versa.

  10. Very recently, the role of idealizations in the debate has been considered by Baron (2016) and De Bianchi (2016), which I discuss briefly in Sect. 5, but the emphasis in not placed on the distinctly indispensable role of infinite/infinitesimal idealization in science. Similarly, discussion of idealization in the philosophy of mathematics, e.g., Maddy (1997), Leng (2010, pp. 116–122, 133–137), either does not explicitly discuss the type of essential, infinite/infinitesimal idealizations that I’m concerned with; or else it is assumed or concluded that such issues do not introduce substantive novel problems for the nominalist. I will argue otherwise.

  11. In calling such explanations structural explanation Leng is following some previous authors, e.g., Railton (1980), Hughes (1989) and Bokulich (2008). For instance, according to Bokulich (2008, p. 149):

    [A] structural explanation is one in which the explanandum is explained by showing how the (typically mathematical) structure of the theory itself limits what sorts of objects, properties, states, or behaviors are admissible within the framework of that theory, and then showing that the explanandum is in fact a consequence of that structure.

  12. A transcendental number is a number that is not a root of a (non-trivial) polynomial with rational coefficients (where a rational number is any number that can be expressed as a fraction of two integers).

  13. For the purposes of this paper I set aside the objection that there is no such thing as “physical structure” (and thus no such thing as a physical system instantiating any kind of mathematical structure) since “structure” is, strictly speaking, a set-theoretic notion (Frigg 2006).

  14. For well-received characterizations and taxonomies of idealizations see McMullin (1985) and Weisberg (2013).

  15. Butterfield (2011, Section 3) discusses similar distinctions. In particular, he makes a distinction between a system\(\sigma \left( N \right) \) that depends on some parameter N (let \(\{\sigma \left( N \right) \)} denote a sequence of such systems), a quantity defined on the system \(f\left( {\sigma \left( N \right) } \right) \) (let \(\left\{ {f\left( {\sigma \left( N \right) } \right) } \right\} \) denote a sequence of quantities on successive systems), and a (real number) value\(v\left( {f\left( {\sigma \left( N \right) } \right) } \right) \) of quantities on successive systems (where a sequence of states on \(\sigma \left( N \right) \) is implicitly understood; (let \(\left\{ {v\left( {f\left( {\sigma \left( N \right) } \right) } \right) } \right\} \) denote a sequence of values on successive systems). A limit system\(\sigma \left( \infty \right) \) arises when \(\mathop {\lim }\nolimits _{N\rightarrow \infty } \left\{ {\sigma \left( N \right) } \right\} \) is well-defined—otherwise there is no limit system. A property of a limit system refers to the value \(v(f\left( {\sigma \left( \infty \right) } \right) \) of the (natural) limit quantity \(f\left( {\sigma \left( \infty \right) } \right) \) (in the natural limit state) on \(\sigma \left( \infty \right) \). A limit property\(v(f\left( {\sigma \left( N \right) } \right) \) is a limit of a sequence of values of quantities on successive systems (or, values on the systems on the way to the limit) and is well-defined when \(\mathop {\lim }\nolimits _{N\rightarrow \infty } \left\{ {v\left( {f\left( {\sigma \left( N \right) } \right) } \right) } \right\} \) exists. The question that I am discussing is whether a property of a limit system equals the system’s limit property since this is what it precisely means for a property to be approximately instantiated. Specifically, the question asks whether \(v(f\left( {\sigma \left( \infty \right) } \right) =\mathop {\lim }\nolimits _{N\rightarrow \infty } \left\{ {v\left( {f\left( {\sigma \left( N \right) } \right) } \right) } \right\} \) (assuming both are well-defined).

  16. Strictly speaking, a structure \(S=\left\langle {U,O,R,} \right\rangle \) is a composite entity consisting of a non empty set U of individuals called the domain of the structure S, an indexed set O operations on U, and a non-empty indexed set R of relation on U, where we can think of properties as one-placed relations. For my purposes discussing structure as a set of properties will do. In principle, one can extend my claims to the more formal characterization of structure.

  17. One may worry that the characterization for “approximate instantiation” suggested here seems to privilege “continuous” properties over “discrete” properties. I do not think that this is the case. We can reasonably talk about a discrete series converging to a number, e.g., \(1,\frac{1}{2},\frac{1}{4},\frac{1}{8},\ldots \rightarrow 0\), or even a sequence of non-continuous functions converging to a function, e.g., take \(f\left( x \right) =0 \, \text {if} \, x\ne 0 \, \text {and} \,f\left( x \right) =\frac{1}{n} \,\,\text {if}\,\, x=0\). Foreshadowing what is to come, the problem with topological properties like “connectedness” is that, first, there is no convergence and, second, notions like convergences, limits, continuity, etc., all presuppose topological properties.

  18. These are all rough and intuitive characterizations of the notions of connectedness, homotopy, etc. They will do for my purposes, and I will only introduce further details when necessary. For precise characterization see standard textbooks on algebraic topology, e.g. Hatcher (2002).

  19. What do we mean by “approximately” here? We mean that we can have the volume \(V=L^{2}\varepsilon \) be as small as we want by choosing an appropriate epsilon so that we can come as close to zero volume as we may want.

  20. See Earman (2010) for a discussion.

  21. The name is due to Noble laureate Frank Wilczek (1982a, b). Note that anyons and fractional statistics have nothing to do with so-called paraparticles and parastatistics (which arise from higher dimensional representations of the permutation group). For more on anyons see Wilczek (1990), Khare (2005), and references therein.

  22. See Hatcher (2002) for relevant background in algebraic topology. Roughly, the “one-dimensional unitary representation” will allow us to represent groups with numbers. The “fundamental group,” also known as the first homotopy group, is a topological invariant that allows one to classify topological spaces according to whether paths or loops in the space can be continuously deformed into each other.

  23. For more on the quantum Hall effect see Chakraborty and Pietilinen (1995), Fradkin (2013, Ch. 13), Ezawa (2013) and Stern (2008), and for recent philosophical assessments see Bain (2013, 2016), Guay and Sartenaer (2016a, b), Lancaster and Pexton (2015), Lederer (2015) and Shech (2015).

  24. For example, “...in two dimensions, the space is multiply connected which results in the possibility of ... intermediate statistics” (Khare 2005, 5; original emphasis). By “intermediate statistics” the author is referring to “fractional statistics,” and by “multiply connectedness” of space is meant “non-simply connected” space. Details are provided in the “Appendix”, but a mere glance at the literature (e.g., Pachos 2012; Rao 2001) will confirm that indeed appeals to a two dimensional topological structure is the standard explanation in the literature and constitutes part of our “best scientific theories.”

  25. See Hempel and Oppenheim (1948) for the deductive-nomological account of explanation and see for example, Salmon (1984) and Woodward (2003) for causal accounts.

  26. An anonymous referee presses further: one wonders where the actual explanation arises. What seems to be doing the explaining is that the phase factors are what they are. There is of course a connection between the braid group and the phase factors—the approximately two-dimensional character of the configuration space removes the extra constraints on phase factors that arise in three-dimensions, constraints that prevent fractional statistics. Otherwise what is the motivation for thinking that the best explanation for what is going on here is the mathematical fact appealed to? My reply is that, first, it is the two-dimensional character of the configuration space—not the “approximately” two-dimensional character—that removes the constraint on phase factors arising in three-dimensions. Second, this is the only explanation of the emergence of anyons—the sole (developed) account of their theoretic possibility (to my best knowledge)—and the received explanation among the scientific community. I argue that if we take this explanation at face value it blocks Leng’s approach to easy road nominalism. I do not deny that an easy road nominalist could reject the standard story and seek an alternative one, e.g., via future theory.

  27. As texts with titles such as Topological Effects in Quantum Mechanics (Afanasiev 1999) and Introduction to Topological Quantum Computation (Pachos 2012) confirm. In fact, the 2016 Nobel Prize in Physics was awarded to David J. Thouless, F. Duncan M. Haldane, and J. Michael Kosterlitz for theoretical discoveries of topological phase transitions and topological phases of matter.

  28. It is of special interest to philosophers because it seems to portray a type of quantum non-locality (since the magnetic field affects a beam of electrons while being confined to a region where the beam is not) and raises a host of interpretive issues regarding ontology and indeterminism in classical electromagnetism and quantum mechanics. See, for example, Healey (2007), Maudlin (1998), Mattingly (2006) (and references within) for some of the philosophical literature.

  29. This claim is commonly found in both the scientific (e.g., Aharonov and Bohm (1959, p. 490), Wu and Yang (1975, p. 3854), Peshkin and Tonomura (1989, p. 27), Ryder (1996, p. 102)) and philosophical (e.g. Batterman (2003), Nounou (2003)) literature. It is rejected by, e.g., Healey (2007), Earman (2017) and Shech (2017).

  30. Batterman (2002) and Bokulich (2008) discus such examples.

  31. Such theories include mean-field theory, Yang–Lee’s theory, Landau’s approach and renormalization group methods. For more on the philosophical debate revolving around phase transitions see  Batterman (2005), Butterfield (2011), Menon and Callender (2013), Norton (2012), and Shech (2013). See Kadanoff (2000) for a standard introductory text and Ruelle (2004) for rigorous results.

  32. Where by a non-analytic function we mean, roughly, that it is not possible to differentiate the function indefinitely, i.e., the function contains ‘kinks’ or ‘discontinuities.’

  33. First, remember that \(\pi _1 \ne \pi \): \(\pi _1 \) represent the fundamental group (the first homotopy group) and \(\pi \) the covering projection. Second, roughly, recall that any two closed curves (loops) that can be continuously deformed into one another will be part of the same homotopy class. Homotopy is an equivalence relation among loops.

  34. C is not necessarily a closed curve. If the bundle is not curved than C is closed but if the bundle is curved than C is not closed.

References

  • Afanasiev, G. N. (1999). Topological effects in quantum mechanics. Norwell, MA: Kluwer.

    Book  Google Scholar 

  • Aharonov, Y., & Bohm, D. (1959). Significance of electromagnetic potentials in the quantum theory. Physical Review, 115, 485–91.

    Article  Google Scholar 

  • Ando, T., Fowler, A. B., & Stern, F. (1982). Electronic properties of two-dimensional systems. Reviews of Modern Physics, 54, 437–672.

    Article  Google Scholar 

  • Arovas, D. P., Schrieffer, J. R., & Wilczek, F. (1984). Fractional statistics and the quantum Hall effect. Physical Review Letters, 53, 722–723.

    Article  Google Scholar 

  • Artin, E. (1947). Theory of braids. Annals of Mathematics, 48(1), 101–126.

    Article  Google Scholar 

  • Azzouni, J. (2004). Deflating existential consequence: A case for nominalism. New York: Oxford University Press.

    Book  Google Scholar 

  • Azzouni, J. (2012). Taking the easy road out of Dodge. Mind, 121, 951–66.

    Article  Google Scholar 

  • Bain, J. (2013). Emergence in effective field theories. European Journal for Philosophy of Science, 3, 257–273.

    Article  Google Scholar 

  • Bain, J. (2016). Emergence and mechanism in the fractional quantum Hall effect. Studies in History and Philosophy of Modern Physics, 56, 27–38.

  • Baron, S. (2016). The explanatory dispensability of idealizations. Synthese, 193, 365–386.

    Article  Google Scholar 

  • Baker, A. (2005). Are there genuine mathematical explanations of physical phenomena? Mind, 114, 223–38.

    Article  Google Scholar 

  • Baker, A. (2009). Mathematical explanation in science. British Journal for the Philosophy of Science, 60, 611–33.

    Article  Google Scholar 

  • Baker, A. (2012). Science-driven mathematical explanation. Mind, 121, 243–67.

    Article  Google Scholar 

  • Balaguer, M. (1998). Platonism and anti-platonism in mathematics. New York: Oxford University Press.

    Google Scholar 

  • Batterman, R. (2002). The devil in the details: Asymptotic reasoning in explanation, reduction, and emergence. London: Oxford University Press.

    Google Scholar 

  • Batterman, R. (2003). Falling cats, parallel parking, and polarized light. Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics, 34, 527–557.

    Article  Google Scholar 

  • Batterman, R. (2005). Critical phenomena and breaking drops: Infinite idealizations in physics. Studies In History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics, 36B, 225–44.

    Article  Google Scholar 

  • Batterman, R., & Rice, C. (2014). Minimal model explanations. Philosophy of Science, 81(3), 349–376.

    Article  Google Scholar 

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

    Book  Google Scholar 

  • Bueno, O. (2012). An easy road to nominalism. Mind, 121, 967–82.

    Article  Google Scholar 

  • Butterfield, J. (2011). Less is different: Emergence and reduction reconciled. Foundations of Physics, 41, 1065–1135.

    Article  Google Scholar 

  • Camino, F. E., Zhou, W., & Goldman, V. J. (2005). Realization of a Laughlin quasiparticle interferometer: Observation of fractional statistics. Physical Review B, 72, 075342.

    Article  Google Scholar 

  • Chakraborty, T., & Pietilinen, P. (1995). The quantum Hall effects. Berlin: Springer.

    Book  Google Scholar 

  • Colyvan, M. (2001). The indispensability of mathematics. New York: Oxford University Press.

    Book  Google Scholar 

  • Colyvan, M. (2007). Mathematical recreation versus mathematical knowledge. In M. Leng, A. Paseau, & M. D. Potter (Eds.) (pp. 109–22).

  • Colyvan, M. (2010). There is no easy road to nominalism. Mind, 119, 285–306.

    Article  Google Scholar 

  • Colyvan, M. (2012a). Road work ahead: Heavy machinery on the easy road. Mind, 121, 1031–1046.

    Article  Google Scholar 

  • Colyvan, M. (2012b). An introduction to the philosophy of mathematics. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • De Bianchi, S. (2016). Which explanatory role for mathematics in scientific models? reply to ‘the explanatory dispensability of idealizations’. Synthese, 193, 387–401.

    Article  Google Scholar 

  • Earman, J. (2010). Understanding permutation invariance in quantum mechanics. (Unpublished manuscript).

  • Earman, J. (2017). The role of idealizations in the Aharonov-Bohm effect. Synthese,. https://doi.org/10.1007/s11229-017-1522-9.

  • Ezawa, Z. F. (2013). Quantum Hall effects. Singapore: World Scientific.

    Book  Google Scholar 

  • Fadell, E., & Neuwirth, L. (1962). Configuration spaces. Mathematica Scandinavica, 10, 111–118.

    Article  Google Scholar 

  • Field, H. H. (1980). Science without numbers: A defense of nominalism. Oxford: Blackwell.

    Google Scholar 

  • Fox, R., & Neuwirth, L. (1962). The braid groups. Mathematica Scandinavica, 10, 119–126.

    Article  Google Scholar 

  • Fradkin, E. (2013). Field theories of condensed matter physics (2nd ed.). Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Frigg, R. (2006). Scientific representation and the semantic view of theories. Theoria, 55(2006), 49–65.

    Google Scholar 

  • Guay, A., & Sartenaer, O. (2016a). A new look at emergence. Or when after is different. European Journal for Philosophy of Science, 6, 297–322.

    Article  Google Scholar 

  • Guay, A., Sartenaer, O. (2016b). Emergent quasiparticles. Or how to get a rich physics from a sober metaphysics. In O. Bueno, R. Chen, & M. B. Fagan (Eds.), Individuation across experimental and theoretical sciences. Oxford: Oxford University Press. http://hdl.handle.net/2078.1/179059

  • Hatcher, A. (2002). Algebraic topology. Cambridge: Cambridge University Press.

    Google Scholar 

  • Healey, R. A. (2007). Gaugin what’s real: The conceptual foundations of contemporary gauge theories. New York: Oxford University Press.

    Book  Google Scholar 

  • Hempel, C., & Oppenheim, P. (1948). Studies in the logic of explanation. Philosophy of Science, 15, 135–75. Repr. in Hempel, C. (Ed.), Aspects of scientific explanation and other essays in the philosophy of science. New York: Free Press (1965) (pp. 245–90).

  • Hughes, R. I. G. (1989). Bell’s theorem, ideology, and structural explanation. In J. T. Cushing & E. McMullin (Eds.), Philosophical consequences of quantum theory: Reections on Bell’s theorem (pp. 195–207). Notre Dame, IN: University of Notre Dame Press.

    Google Scholar 

  • Kadanoff, L. P. (2000). Statistical physics: Statics, dynamics and renormalization. Singapore: World Scientific.

    Book  Google Scholar 

  • Katanaev, M. O. (2011). On geometric interpretation of the Aharonov-Bohm effect. Russian Physics Journal, 54(5), 507–514.

    Article  Google Scholar 

  • Khare, A. (2005). Fractional statistics and quantum theory. Hackensack, NJ: World Scientific.

  • Klitzing, K. v., Dorda, G., & Pepper, M., (1980). New method for high-accuracy determination of the fine-structure constant based on quantized hall resistance. Physical Review Letters, 45, 494–497.

  • Laidlaw, M. G., & DeWitt, C. M. (1971). Feynman functional integrals for system of indistinguishable particles. Physical Review D, 3, 1375–1378.

    Article  Google Scholar 

  • Lancaster, T., & Pexton, M. (2015). Reduction and emergence in the fractional quantum Hall state. Studies in History and Philosophy of Modern Physics, 52, 343–357.

    Article  Google Scholar 

  • Landsman, N. P. (2016). Quantization and superselection sectors III: Multiply connected spaces and indistinguishable particles. Reviews in Mathematical Physics, 28, 1650019.

    Article  Google Scholar 

  • Laughlin, R. (1983). Anomalous quantum Hall effect: An incompressible quantum fluid with fractionally charged excitations. Physical Review Letters, 50, 1395–8.

    Article  Google Scholar 

  • Lederer, P. (2015). The quantum Hall effects: Philosophical approach. Studies in History and Philosophy of Modern Physics, 50, 25–42.

    Article  Google Scholar 

  • Leinaas, J. M., & Myrheim, J. (1977). On the theory of identical particles. Nuovo Cimento B, 37, 1–23.

    Article  Google Scholar 

  • Leng, M. (2005). Mathematical explanation. In C. Cellucci & D. Gillies (Eds.), Mathematical reasoning and heuristics. London: King’s College Publications.

    Google Scholar 

  • Leng, M. (2010). Mathematics and reality. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Leng, M. (2012). Taking it easy: A response to Colyvan. Mind, 121, 983–96.

    Article  Google Scholar 

  • Liggins, D. (2012). Weaseling and the content of science. Mind, 121, 997–1006.

    Article  Google Scholar 

  • Maddy, P. (1997). Naturalism in mathematics. Oxford: Clarendon Press.

    Google Scholar 

  • Masenes, L., & Oppenheim, J. (2017). A general derivation and quantification of the third law of thermodynamics. Nature Communications, 8, 14538.

    Article  Google Scholar 

  • Mattingly, J. (2006). Which gauge matters. Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics, 37, 243–262.

    Article  Google Scholar 

  • Maudlin, T. (1998). Discussion: Healey on the Aharonov-Bohm effect. Philosophy of Science, 65, 361–368.

    Article  Google Scholar 

  • McMullin, E. (1985). Galilean idealization. Studies in the History and Philosophy of Science, 16, 247–273.

    Article  Google Scholar 

  • Melia, J. (2000). Weaseling away the indispensability argument. Mind, 109, 455–7.

    Article  Google Scholar 

  • Melia, J. (2002). Reply to Colyvan. Mind, 111, 75–9.

    Article  Google Scholar 

  • Menon, T., & Callender, C. (2013). Turn and face the strange... Ch-Ch-changes: Philosophical questions raised by phase transitions. In R. W. Batterman (Ed.), The Oxford handbook of philosophy of physics. Oxford: Oxford University Press.

  • Messiah, A. M. (1962). Quantum mechanics. New York: Wiley.

    Google Scholar 

  • Messiah, A. M., & Greenberg, O. W. (1964). Symmetrization postulate and its experimental foundation. Physical Review B, 136, 248–267.

    Article  Google Scholar 

  • Molinini, D., Pataut, F., & Sereni, A. (2016). Indispensability and explanation: An overview and introduction. Synthese, 193, 317–332.

    Article  Google Scholar 

  • Morandi, G. (1992). The role of topology in classical and quantum mechanics. Berlin: Springer.

    Book  Google Scholar 

  • Norton, J. D. (2012). Approximations and idealizations: Why the difference matters. Philosophy of Science, 79, 207–32.

    Article  Google Scholar 

  • Nounou, A. (2003). A fourth way to the Aharonov-Bohm effect. In K. Bradind & E. Castellani (Eds.), Symmetries in physics: Philosophical replections. Cambridge: Cambridge University Press.

    Google Scholar 

  • Pachos, J. K. (2012). Introduction to topological quantum computation. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Peshkin, M., & Tonomura, A. (1989). The Aharonov–Bohm effect. LNP (Vol. 340). Berlin: Springer.

  • Putnam, H. (1971). Philosophy of logic. New York: Harper.

    Google Scholar 

  • Quine, W. V. O. (1981). Theories and things. Cambridge, MA: Harvard University Press.

    Google Scholar 

  • Railton, P. (1980). Explaining explanation: A realist account of scientific explanation and understanding. Ph.D. Dissertation, Princeton University.

  • Rao, S. (2001). An anyon rimer. arXiv:hep-th/9209066.

  • Ruelle, D. (2004). Thermodynamic formalism (2nd ed.). Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Ruetsche, L. (2011). Interpreting quantum theories: The art of the possible. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Ryder, L. H. (1996). Quantum field theory. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Salmon, W. (1984). Scientific explanation and the causal structure of the world. Princeton, NJ: Princeton University Press.

    Google Scholar 

  • Shech, E. (2013). What is the ‘paradox of phase transitions?’. Philosophy of Science, 80, 1170–1181.

    Article  Google Scholar 

  • Shech, E. (2015). Two approaches to fractional statistics in the quantum Hall effect: Idealizations and the curious case of the anyon. Foundations of Physics, 45(9), 1063–110.

    Article  Google Scholar 

  • Shech, E. (2017). Idealizations, essential self-adjointness, and minimal model explanation in the Aharonov–Bohm effect. Synthese, 1–25. https://doi.org/10.1007/s11229-017-1428-6.

  • Stern, A. (2008). Anyons and the quantum Hall effect–a pedagogical review. Annals of Physics, 323, 204–249.

    Article  Google Scholar 

  • Tonomura, A. (2010). The AB effect and its expanding applications. Journal of Physics A: Mathematical and Theoretical, 43, 1–13.

    Article  Google Scholar 

  • Tsui, D. C., Stormer, H. L., & Gossard, A. C. (1982). Two-dimensional magnetotransport in the extreme quantum limit. Physical Review Letters, 48(22), 1559.

  • von Klitzing, K. (2004). 25 Years of quantum Hall effect (QHE): A personal view on the discovery, physica and application of this quantum effect. In B. Douçot, V. Pasquier, B. Duplantier, & V. Rivasseau (Eds.), The quantum Hall effect Poincaré seminar (pp. 1–23). Berlin: Birkhäuser.

    Google Scholar 

  • Weisberg, M. (2013). Simulation and similarity: Using models to understand the world. New York: Oxford University Press.

    Book  Google Scholar 

  • Wilczek, F. (1982a). Magnetic flux, angular momentum and statistics. Physical Review Letters, 48, 1144–1146.

    Article  Google Scholar 

  • Wilczek, F. (1982b). Quantum mechanics of fractional-spin particles. Physical Review Letters, 49, 957–959.

    Article  Google Scholar 

  • Wilczek, F. (Ed.). (1990). Fractional statistics and anyon superconductivity. Singapore: World Scientific.

    Google Scholar 

  • Woodward, J. F. (2003). Making things happen: A theory of causation. Oxford: Oxford University Press.

    Google Scholar 

  • Wu, T. T., & Yang, C. N. (1975). Concept of nonintegrable phase factors and global formulation of gauge fields. Physical Review D, 12, 3845.

    Article  Google Scholar 

  • Yablo, S. (1998). Does ontology rest on a mistake? Aristotelian Society, Supplementary, 72, 229–61.

    Article  Google Scholar 

  • Yablo, S. (2002). Abstract objects: A case study. Philosophical Issues, 12, 220–40.

    Article  Google Scholar 

  • Yablo, S. (2012). Explanation, extrapolation, and existence. Mind, 121, 1007–30.

    Article  Google Scholar 

Download references

Acknowledgements

I gratefully acknowledge useful discussions with Mary Leng and Mark Colyvan. Previous versions of this paper were presented at the “Annual Meeting of the European Philosophy of Science Association” at Exeter University on 09/09/2017 and “Current Projects” at Department of Philosophy at University of Sydney on 08/03/2017. I thank the participants for helpful comments. This work was produced as part of a visiting fellowship at the Sydney Centre for the Foundations of Science and Ideas and the Sydney Centre for Time at University of Sydney.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Elay Shech.

Appendix: The configuration space approach to fractional quantum statistics

Appendix: The configuration space approach to fractional quantum statistics

Let the configuration space of a particle in d-dimensions be \({\mathbb {R}}^{d}\) (where \({\mathbb {R}}\) is the set of real numbers) so that the position of the particles is given by an element of the space \(x\in {\mathbb {R}}^{d}\). Consider N such identical particles. The configuration space approach argues that the appropriate configuration space Q for N identical particles is not the Cartesian product of the single particle spaces, \({\mathbb {R}}^{Nd}\equiv {\mathbb {R}}^{d}\times \cdots \times {\mathbb {R}}^{d}\) (N times), as one would expect if the particles were not identical. Instead, since the particles are identical, configurations that differ only by a permutation of particles ought to correspond to the same physical state. For the simplest case \(d=1N=2\), this means that the two configurations \(\left( {x_1,x_2 } \right) =\left( {1,2} \right) \) and \(\left( {x_1,x_2 } \right) =\left( {2,1} \right) \) actually represent the same state, and so we must divide out such permuted configurations. In other words, we move from the entire space \({\mathbb {R}}^{Nd}={\mathbb {R}}^{2*1}={\mathbb {R}}^{2}\) of the two-dimensional plane to consider only half the plane. In the context of the general case, this corresponds to considering the quotient (or identification) space that one attains by diving out the permutation group \(S_N \): \(Q=\frac{{\mathbb {R}}^{d}}{S_N }\).

Next, we’ll want to excise the set \(\Delta \) of diagonal points in \({\mathbb {R}}^{Nd}\), which represent all points where the particles coincide. For the \(d=1N=2\) case, this means that we must excise points of the sort: \(\left( {x_1,x_2 } \right) =\left( {x,x} \right) \), where \(x_1 =x_2.\) If we do not excise coincidence points (known as singular points) the resulting configuration space will not have a structure rich enough to represent fermions (or anyons).

We are left with a configuration space for N identical particle in d dimensions, in which we have divided out the action of the permutation group and excluded diagonal points: \(\frac{{\mathbb {R}}^{d}-\Delta }{S_N }\). The exclusion of the diagonal point implies that the space in not simply connected and I denote this new configuration space with \(\mathop {\tilde{Q}} \). We must now undertake QM on a multiply connected space. In order to do so I will follow Morandi (1992, pp. 114–144) closely.

Ordinary QM (in the Schrodinger picture) on simply connected regions represents the states of physical systems by complex wave functions \(\Psi \) which are elements of a Hilbert space \({\mathcal {H}}\) of square-integrable functions on the configuration space Q over the field of complex numbers \({\mathbb {C}}\): \(\Psi \in \mathcal{H}=L_{\mathbb {C}}^2 (Q)\), where \({\mathcal {H}}^{N}=L_{\mathbb {C}}^2 \left( Q \right) \times \cdots \times L_{\mathbb {C}}^2 \left( Q \right) \cong L_{\mathbb {C}}^2 \left( {Q^{N}} \right) \) for N identical particles. In order to extend ordinary QM to multiply connected regions we need to appeal the topological notion of a universal covering spaceQ of \(\mathop {\tilde{Q}}\). That is to say, for any topological space, including ones that are not simply connected such as \(\mathop {\tilde{Q}}\), we can construct a universal covering space Q that is simply connected (so that ordinary QM applies) with a covering projection map: \(\pi {:}\,Q\rightarrow \mathop {\tilde{Q}}\) (where \(q\in Q\) and \(\mathop {\tilde{q}}\in \mathop {\tilde{Q}})\).

Since all the physical information in QM is contained in the squared modulus of the wave function (i.e. the probability density) \(\left| {\Psi }\left( q \right) \right| ^{2}\), we will want this quantity to be projectable down to \(\mathop {\tilde{Q}}\) for any \(q\in Q\) in the sense that \(\left| {{\Psi }\left( q \right) } \right| ^{2}\) will depend only on a point \(\mathop {\tilde{q}}=\pi \left( q \right) \) in the multiply connected configuration space \(\mathop {\tilde{Q}}.\) If this condition is satisfied we say that we have a projectable quantum mechanics, and we will have succeeded in extending ordinary QM to a multiply connected region. To that effect, consider an arbitrary point \(\mathop {\tilde{q}}\in \mathop {\tilde{Q}}\), let \(\mathop {\tilde{C}}\) be a closed curve beginning and ending at \(\mathop {\tilde{q}}\), and let [\(\mathop {\tilde{C}}\)] be the corresponding (first) homotopy class of curves based at \(\mathop {\tilde{q}}\) so that [\(\mathop {\tilde{C}}\)]\(\in \pi _1 \left( {\mathop {\tilde{Q}} ,\mathop {\tilde{q}}} \right) \).Footnote 33 Further, let \(q=\pi ^{-1}\left( {\mathop {\tilde{q}}} \right) \) be any point in the fiber over \(\mathop {\tilde{q}}\). The homotopy lifting theorem says that all curves \(\mathop {\tilde{C}}\) in [\(\mathop {\tilde{C}}\)] are lifted to a curve C in Q beginning at q and ending at some point \(q^{\prime }\) that is also in the fiber over \(\mathop {\tilde{q}}\) (so that \(\pi \left( {{q}'} \right) ) = \mathop {\tilde{q}}\) and \(q^{\prime }\in Q)\).Footnote 34 Denote this result by \(q^{\prime }=\left[ \mathop {\tilde{C}} \right] \cdot q\) and recall that \(\gamma \) represents a phase factor (although, as of now, we have placed no constraints on the form of \(\gamma )\). We then have the following two central theorems (Morandi 1992, pp. 119–120):

Theorem 1

Projectable Quantum Mechanics are obtained if and only if the wave functions on the universal covering space obey the boundary conditions:

$$\begin{aligned} \Psi \left( {\left[ {\mathop {\tilde{C}} } \right] \cdot q} \right) =\gamma \left( {\left[ {\mathop {\tilde{C}}} \right] } \right) \Psi \left( \hbox {q} \right) \,\hbox {for}\,\hbox {all}\,q\in Q\\ \hbox {Where} \left| {\gamma \left( {\left[ {\mathop {\tilde{C}}} \right] } \right) } \right| =1\,\hbox {for}\,\hbox {all}\,\left[ {\mathop {\tilde{C}}} \right] \in \pi _1 \left( {\mathop {\tilde{Q}}} \right) \end{aligned}$$

This means that wave functions on the universal covering space at different points (e.g. q and \(q^{\prime })\), but on the same fiber above some point in the multiply connected configuration space (e.g. \(\mathop {\tilde{q}})\), can differ at most by a phase factor of modulus 1. Moreover, since the universal covering space is simply connected, the wave functions must be single valued. We then get:

Theorem 2

The map \(\gamma {:}\,\pi _1 \left( {\mathop {\tilde{Q}}} \right) \rightarrow U\left( 1 \right) \) by \(\left[ {\mathop {\tilde{C}} } \right] \rightarrow \gamma \left( {\left[ {\mathop {\tilde{C}}} \right] } \right) \) is a one-dimensional unitary representation of \(\pi _1 \left( {\mathop {\tilde{Q}}} \right) \).

Accordingly, we get our main result: in order to ascertain what type of phase factor is gained by a wave function when it is permuted—with the corresponding available quantum statistics—we must enquire into the one-dimensional unitary representation of the fundamental group of the configuration space of the system. Recall that in the case of N identical particles in d dimensions we have:

$$\begin{aligned} \mathop {\tilde{Q}} =\frac{{\mathbb {R}}^{d}-\Delta }{S_N } \end{aligned}$$

It has been shown by Artin (1947), Fadell and Neuwirth (1962), and Fox and Neuwirth (1962) that the fundamental group for the two- and three-dimensional cases are given by:

$$\begin{aligned} \pi _{1} \left( {\mathop {\tilde{Q}}} \right) =B_N \hbox { for } d=2\\ \pi _{1} \left( {\mathop {\tilde{Q}}} \right) =S_N \hbox { for } d=3 \end{aligned}$$

where \(S_N\) is the permutation group and \(B_N\) is the braid group. Moreover, as stated previously, the one-dimensional representation of \(S_N\) is \(\gamma =\pm \,1,+\,1\) for bosons and \(-\,1\) for fermions, while the one-dimensional representation of \(B_N \) is \(\gamma _{\left( \theta \right) } =e^{i\theta }\) where \(0\le \theta \le 2\pi \) so that the exchange phase can take on a continuous range of factors allowing for bosons, fermions and anyons.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Shech, E. Infinitesimal idealization, easy road nominalism, and fractional quantum statistics. Synthese 196, 1963–1990 (2019). https://doi.org/10.1007/s11229-018-1680-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11229-018-1680-4

Keywords

Navigation