Skip to main content
Log in

Deduction and definability in infinite statistical systems

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

A Correction to this article was published on 03 November 2018

This article has been updated

Abstract

Classical accounts of intertheoretic reduction involve two pieces: first, the new terms of the higher-level theory must be definable from the terms of the lower-level theory, and second, the claims of the higher-level theory must be deducible from the lower-level theory along with these definitions. The status of each of these pieces becomes controversial when the alleged reduction involves an infinite limit, as in statistical mechanics. Can one define features of or deduce the behavior of an infinite idealized system from a theory describing only finite systems? In this paper, I change the subject in order to consider the motivations behind the definability and deducibility requirements. The classical accounts of intertheoretic reduction are appealing because when the definability and deducibility requirements are satisfied there is a sense in which the reduced theory is forced upon us by the reducing theory and the reduced theory contains no more information or structure than the reducing theory. I will show that, likewise, there is a precise sense in which in statistical mechanics the properties of infinite limiting systems are forced upon us by the properties of finite systems, and the properties of infinite systems contain no information beyond the properties of finite systems.

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

Change history

  • 03 November 2018

    Prop. 1 on p. 10 is false as stated. The proof implicitly assumes that all Cauchy sequences.

Notes

  1. See also Dizadji-Bahmani et al. (2010) for more on Schaffner’s approach applied to statistical mechanics.

  2. See Fletcher (2016) and Rosaler (2015) for a discussion of limiting relations, reduction, and approximation in other areas of physics. See also Norton (2012) for a discussion of the role of approximation and idealization.

  3. There is a further point of contention in the philosophy of physics literature regarding the significance of the renormalization group [(see, e.g., Batterman (2010)]. I will not touch upon renormalization techniques in this paper; I think further work is required to determine whether my results are applicable to the limits taken during renormalization.

  4. For more on C*-algebras and W*-algebras, see Kadison and Ringrose (1997), Sakai (1971), and Landsman (1998). For more on algebraic quantum theory, see Haag (1992), Bratteli and Robinson (1996), Emch (1972), and Wald (1994). For philosophical introductions, see Halvorson (2006) and Ruetsche (2011b).

  5. One could restrict attention here to sequences because the norm topology is second countable, but for the weak topologies considered later, which are not second countable, one must work with arbitrary nets.

  6. A linear functional \(\rho \in \mathfrak {A}^*\) is positive if \(\rho (A^*A)\ge 0\) for all \(A\in \mathfrak {A}\) and normalized if \(||\rho || = 1\).

  7. Thanks to an anonymous reviewer for this point.

  8. See Alfsen and Shultz (2001) for more on the mathematical structure on the space of states. See Ruetsche and Earman (2011) and Ruetsche (2011a) for more on the physical interpretation of certain kinds of states. And see Feintzeig (2017a) for more on the relationship between an algebra and its collection of allowed states.

  9. Here, \(\mathfrak {R}\) as a Banach space is to be understood as the Banach space dual to \(\mathfrak {R}_*\), but the algebraic operations of multiplication and involution are not determined by \(\mathfrak {R}_*\).

  10. It might be somewhat surprising that \(\mathfrak {A}^{**}\), which is just the Banach dual space to a Banach space, has additional algebraic structure that makes it into a C*-algebra. But this is easy to see once one notices that the original C*-algebra \(\mathfrak {A}\) is weak* dense in \(\mathfrak {A}^{**}\) with respect to this canonical embedding \(J_\mathfrak {A}\) (see Feintzeig 2017b; Sakai 1971), so the algebraic structure of \(\mathfrak {A}^{**}\) can be naturally inherited from \(\mathfrak {A}\). Multiplication and involution on \(\mathfrak {A}^{**}\) are defined as the unique weak* continuous extensions of the operations on \(\mathfrak {A}\). (We only require multiplication to be separately weak* continuous in each of its arguments because multiplication in the original C*-algebra \(\mathfrak {A}\) is not, in general, jointly weakly continuous.)

  11. Proofs of all propositions appear in the appendix.

  12. For introductions to category theory, see Awodey (2010) and Borceux (1994).

  13. See, e.g., Awodey (2010, p. 214) and Borceux (1994, p. 98) for more on adjunctions.

  14. Special thanks to Jim Weatherall and Thomas Barrett for help clarifying the significance of an adjunction. Note that I do not claim that an adjunction immediately signifies definability in the sense of mathematical logic. There is still much further work to be done along these lines—see Sect. 6.

  15. Thanks to an anonymous reviewer for pointing this out.

  16. Wholeness is just condition (ii) of Theorem 1 of Feintzeig (2017a). Condition (i) of that theorem is always satisfied for the predual of a W*-algebra.

  17. The results of this paper also apply to classical statistical systems represented by commutative algebras.

  18. For more on Algebraic Imperialism and Hilbert Space Conservatism, see Arageorgis (1995), Ruetsche (2002, 2003, 2006, 2011b), Baker (2011), Baker and Halvorson (2013), Lupher (2008, 2016) and Feintzeig (2016, 2017a, c).

  19. Thanks to an anonymous reviewer for this point.

  20. See Ruetsche and Earman (2011) for more on algebras with no pure normal states.

References

  • Alfsen, E., & Shultz, F. (2001). State spaces of operator algebras. Boston, MA: Birkhauser.

    Book  Google Scholar 

  • Arageorgis, A. (1995). Fields, particles, and curvature: Foundations and philosophical aspects of quantum field theory in curved spacetime. Ph.D. thesis, University of Pittsburgh.

  • Awodey, S. (2010). Category theory (2nd ed.). New York: Oxford University Press.

    Google Scholar 

  • Baez, J., Bartels, T., & Dolan, J. (2004). Property, structure, and stuff. Quantum Gravity Seminar, University of California, Riverside, http://math.ucr.edu/home/baez/qg-spring2004.

  • Baker, D. (2011). Broken symmetry and spacetime. Philosophy of Science, 78(1), 128–148.

    Article  Google Scholar 

  • Baker, D., & Halvorson, H. (2013). How is spontaneous symmetry breaking possible? Understanding Wigner’s theorem in light of unitary inequivalence. Studies in the History and Philosophy of Modern Physics, 44(4), 464–469.

    Article  Google Scholar 

  • Barrett, T. (2015a). On the structure of classical mechanics. British Journal for Philosophy of Science, 66, 801–828.

    Article  Google Scholar 

  • Barrett, T. (2015b). Spacetime structure. Studies in the History and Philosophy of Modern Physics, 51, 37–43.

    Article  Google Scholar 

  • Barrett, T. (2017a). Equivalent and inequivalent formulations of classical mechanics. http://philsci-archive.pitt.edu/13092/. (Unpublished manuscript).

  • Barrett, T. (2017b). What do symmetries tell us about structure? (Unpublished manuscript).

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

    Google Scholar 

  • Batterman, R. (2005). Critical phenomena and breaking drops: Infinite idealizations in physics. Studies in the History and Philosophy of Modern Physics, 36, 225–244.

    Article  Google Scholar 

  • Batterman, R. (2009). Idealization and modeling. Synthese, 169, 427–446.

    Article  Google Scholar 

  • Batterman, R. (2010). Reduction and renormalization. In A. Hüttemann & G. Ernst (Eds.), Time, chance, and reduction: Philosophical aspects of statistical mechanics (pp. 159–179). Cambridge: Cambridge University Press.

    Chapter  Google Scholar 

  • Borceux, F. (1994). Handbook of categorical algebra. Volume 1: Basic category theory. Cambridge: Cambridge University Press.

  • Bratteli, O., & Robinson, D. (1996). Operator algebras and quantum statistical mechanics (Vol. 2). New York: Springer.

    Google Scholar 

  • Butterfield, J. (2011a). Emergence, reduction, and supervenience: A varied landscape. Foundations of Physics, 41, 920–959.

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Callender, C. (2001). Taking thermodynamics too seriously. Studies in the History and Philosophy of Modern Physics, 32(4), 539–553.

    Article  Google Scholar 

  • Dixmier, J. (1977). C*-Algebras. New York: North Holland.

    Google Scholar 

  • Dizadji-Bahmani, F., Frigg, R., & Hartmann, S. (2010). Who’s afraid of nagelian reduction? Erkenntnis, 73, 393–412.

    Article  Google Scholar 

  • Emch, G. (1972). Algebraic methods in statistical mechanics and quantum field theory. New York: Wiley.

    Google Scholar 

  • Feintzeig, B. (2016). Unitary inequivalence in classical systems. Synthese, 193(9), 2685–2705.

    Article  Google Scholar 

  • Feintzeig, B. (2017a). On the choice of algebra for quantization. Philosophy of Science forthcoming, http://philsci--archive.pitt.edu/id/eprint/12719.

  • Feintzeig, B. (2017b). On theory construction in physics: Continuity from classical to quantum. Erkenntnis. doi:10.1007/s10670-016-9865-z.

  • Feintzeig, B. (2017c). Toward an understanding of parochial observables. British Journal for the Philosophy of Science. doi:10.1093/bjps/axw010.

  • Fletcher, S. (2016). Similarity, topology, and physical significance in relativity theory. British Journal for the Philosophy of Science, 67(2), 365–389.

    Article  Google Scholar 

  • Haag, R. (1992). Local quantum physics. Berlin: Springer.

    Book  Google Scholar 

  • Halvorson, H. (2001). On the nature of continuous physical quantities in classical and quantum mechanics. Journal of Philosophical Logic, 37, 27–50.

    Article  Google Scholar 

  • Halvorson, H. (2006). Algebraic quantum field theory. In J. Butterfield & J. Earman (Eds.), Handbook of the philosophy of physics (pp. 731–864). New York: North Holland.

    Google Scholar 

  • Halvorson, H. (2012). What scientific theories could not be. Philosophy of Science, 79(2), 183–206.

    Article  Google Scholar 

  • Halvorson, H. (2016). Scientific theories. In P. Humphreys (Ed.), The Oxford handbook of philosophy of science. New York: Oxford University Press.

  • Halvorson, H., & Tsementzis, D. (2017). Categories of scientific theories. In E. Landry (Ed.), Categories for the working philosopher. Oxford University Press.

  • Kadison, R., & Ringrose, J. (1997). Fundamentals of the theory of operator algebras. Providence, RI: American Mathematical Society.

    Google Scholar 

  • Landsman, N. P. (1998). Mathematical topics between classical and quantum mechanics. New York: Springer.

    Book  Google Scholar 

  • Landsman, N. P. (2013). Spontaneous symmetry breaking in quantum systems: Emergence or reduction? Studies in the History and Philosophy of Modern Physics, 44, 379–394.

    Article  Google Scholar 

  • Lupher, T. (2008). The philosophical significance of unitarily inequivalent representations in quantum field theory. Ph.D. thesis, University of Texas

  • Lupher, T. (2016). The limits of physical equivalence in algebraic quantum field theory. British Journal for Philosophy of Science. doi:10.1093/bjps/axw017.

  • Nagel, E. (1961). The structure of science. New York: Harcourt, Brace & World.

    Book  Google Scholar 

  • Nickles, T. (1975). Two concepts of intertheoretic reduction. Journal of Philosophy, 70, 181–201.

    Article  Google Scholar 

  • Norton, J. (2012). Approximation and idealization: Why the difference matters. Philosophy of Science, 79(2), 207–232.

    Article  Google Scholar 

  • Reed, M., & Simon, B. (1980). Functional analysis. New York: Academic Press.

    Google Scholar 

  • Rosaler, J. (2015). Local reduction in physics. Studies in the History and Philosophy of Modern Physics, 50, 54–69.

    Article  Google Scholar 

  • Ruetsche, L. (2002). Interpreting quantum field theory. Philosophy of Science, 69(2), 348–378.

    Article  Google Scholar 

  • Ruetsche, L. (2003). A matter of degree: Putting unitary inequivalence to work. Philosophy of Science, 70(5), 1329–1342.

    Article  Google Scholar 

  • Ruetsche, L. (2006). Johnny’s so long at the ferromagnet. Philosophy of Science, 73(5), 473–486.

    Article  Google Scholar 

  • Ruetsche, L. (2011a). Why be normal? Studies in the History and Philosophy of Modern Physics, 42, 107–115.

    Article  Google Scholar 

  • Ruetsche, L. (2011b). Interpreting quantum theories. New York: Oxford University Press.

    Book  Google Scholar 

  • Ruetsche, L., & Earman, J. (2011). Interpreting probabilities in quantum field theory and quantum statistical mechanics. In C. Beisbart & S. Hartmann (Eds.), Probabilities in physics (pp. 263–290). Oxford: Oxford University Press.

    Chapter  Google Scholar 

  • Sakai, S. (1971). C*-algebras and W*-algebras. New York: Springer.

    Google Scholar 

  • Schaffner, K. (1967). Approaches to reduction. Philosophy of Science, 34(2), 137–147.

    Article  Google Scholar 

  • Takesaki, M. (1979). Theory of operator algebras. New York: Springer.

    Book  Google Scholar 

  • Wald, R. (1994). Quantum field theory in curved spacetime and black hole thermodynamics. Chicago: University of Chicago Press.

    Google Scholar 

  • Weatherall, J. (2016a). Are Newtonian gravitation and geometrized Newtonian gravitation theoretically equivalent? Erkenntnis, 81(5), 1073–1091.

    Article  Google Scholar 

  • Weatherall, J. (2016b). Regarding the hole argument. British Journal for Philosophy of Science. doi:10.1093/bjps/axw012.

  • Weatherall, J. (2016c). Understanding gauge. Philosophy of Science, 83, 1039–1049.

    Article  Google Scholar 

  • Weatherall, J. (2017). Categories and the foundations of classical field theories. In E. Landry (Ed.), Categories for the working philosopher. Oxford University Press.

Download references

Acknowledgements

I would like to thank Thomas Barrett, Sam Fletcher, John Manchak, Patricia Palacios, Sarita Rosenstock, Jim Weatherall, and two anonymous reviewers for helpful comments and discussions that lead to this paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Benjamin H. Feintzeig.

Appendix: Proofs of Results

Appendix: Proofs of Results

This appendix contains proofs of all results in Sects. 4 and 5.

Proposition 1

Given any Banach space X, its dual \(X^*\) is complete in the weak* topology as a locally convex vector space.

Proof

Suppose \(\{y_\beta \}\) is a Cauchy net in \(X^*\). Define \(y:X\rightarrow \mathbb {C}\) by

$$\begin{aligned} y(x) = \lim y_\beta (x) \end{aligned}$$

for all \(x\in X\). We know that this limit exists because for any \(x\in X\), \(y_\beta (x)\) is a Cauchy net in \(\mathbb {C}\), which means it must converge because \(\mathbb {C}\) is complete.

Now we must show that \(y\in X^*\), i.e., that y is linear and bounded. The functional y is linear because for any \(x,x'\in X\) and \(\alpha \in \mathbb {C}\),

$$\begin{aligned} y(x + \alpha x') = \lim y_\beta (x + \alpha x') = \lim y_\beta (x) + \alpha \lim y_\beta (x') = y(x) + \alpha y(x') \end{aligned}$$

Notice that because \(|y_\beta (x)|\) is bounded for each \(x\in X\), it follows from the principle of uniform boundedness that \(||y_\beta ||\) is bounded (Reed and Simon 1980). Hence, y is bounded with norm \(||y||\le \sup _{\beta }||y_\beta ||\). Finally, we must show that \(y_\beta \) converges to y in the weak* topology on \(X^*\). But this holds by construction because for any \(x\in X\), \((y-y_\beta )(x)\) converges to zero in \(\mathbb {C}\). \(\square \)

Proposition 2

Suppose X is a Banach space and \(J_X:X\rightarrow X^{**}\) is the canonical evaluation embedding of X in its bidual. Suppose we are given another faithful linear embedding \(K:X\rightarrow Y\) of X in a complete locally convex vector space Y such that K(X) is dense in Y in the locally convex vector space topology on Y. Suppose, in addition, that K is a homeomorphism from X to K(X) in the weak topology on X and the subspace topology on K(X) generated by the locally convex vector space topology on Y. Then there is a vector space isomorphism \(\varphi :Y\rightarrow X^{**}\) that is a homeomorphism in the locally convex vector space topology on Y and the weak* topology on \(X^{**}\) and such that \(J_X = \varphi \circ K\).

figure k

Proof

Suppose \(K:X\rightarrow Y\) is such an embedding. By Corollary 1.2.3 of Kadison and Ringrose (1997, p. 15), the maps \(\varphi _0 = J_X\circ K^{-1}\) and \(\psi _0 = K\circ J_X^{-1}\) extend uniquely to continuous linear maps \(\varphi : Y\rightarrow X^{**}\) and \(\psi : X^{**}\rightarrow Y\) in the weak* topology on \(X^{**}\) and the locally convex vector space topology on Y. Since \(\varphi _0\circ \psi _0\) and \(\psi _0\circ \varphi _0\) are the identity operators on \(J_X(X)\) and K(X), respectively, it follows that \(\varphi \circ \psi \) and \(\psi \circ \varphi \) are the identity operators on \(X^{**}\) and Y, respectively. Thus, \(\varphi \) is an isomorphism and a homeomorphism in the weak* topology on \(X^{**}\) and the locally convex vector space topology on Y. By construction, we have \(\varphi \circ K = \varphi _0\circ K = J_X\circ K^{-1}\circ K = J_X\).

\(\square \)

Proposition 3

Suppose \(\alpha :\mathfrak {A}\rightarrow \mathfrak {B}\) is a *-homomorphism between C*-algebras \(\mathfrak {A}\) and \(\mathfrak {B}\). Then there is a unique weak* continuous extension \(\tilde{\alpha }:\mathfrak {A}^{**}\rightarrow \mathfrak {B}^{**}\) such that \(\tilde{\alpha }\circ J_\mathfrak {A} = J_\mathfrak {B}\circ \alpha \), where \(J_\mathfrak {A}:\mathfrak {A}\rightarrow \mathfrak {A}^{**}\) and \(J_\mathfrak {B}:\mathfrak {B}\rightarrow \mathfrak {B}^{**}\) are the canonical evaluation maps.

figure l

Proof

First, we show that \(\alpha \) is continuous in the weak Banach space topologies on \(\mathfrak {A}\) and \(\mathfrak {B}\). Suppose we have a net \(\{A_\beta \}\subseteq \mathfrak {A}\) such that \(A_\beta \rightarrow A\) weakly. Then for all \(\rho \in \mathfrak {A}^*\), \(\rho (A_\beta )\rightarrow \rho (A)\). Consider the net \(\{\alpha (A_\beta )\}\subseteq \mathfrak {B}\). For any \(\sigma \in \mathfrak {B}^*\), \(\sigma \circ \alpha \in \mathfrak {A}^*\), so it follows that \(\sigma \circ \alpha (A_\beta )\rightarrow \sigma \circ \alpha (A)\). Hence, \(\alpha (A_\beta )\rightarrow \alpha (A)\) weakly, and it follows that \(\alpha \) is weakly continuous.

It follows from Corollary 1.2.3 of Kadison and Ringrose (1997, p. 15) that the map \(J_\mathfrak {B}\circ \alpha \circ J_\mathfrak {A}^{-1}: J_\mathfrak {A}(\mathfrak {A})\rightarrow \mathfrak {B}^{**}\) extends uniquely to a weak* continuous map \(\tilde{\alpha }: \mathfrak {A}^{**}\rightarrow \mathfrak {B}^{**}\). By construction,

$$\begin{aligned} \tilde{\alpha }\circ J_\mathfrak {A} = J_\mathfrak {B}\circ \alpha \circ J_\mathfrak {A}^{-1}\circ J_\mathfrak {A} = J_\mathfrak {B}\circ \alpha \end{aligned}$$

\(\square \)

Proposition 4

G is not full, i.e., G forgets structure.

Proof

By Proposition 5 below, F and G form an adjunction, and we know G is faithful. So it follows from Proposition 3.4.1 of Borceux (1994, p. 114) that G is full only if the counit of the adjunction \(\tilde{1}: G\circ F\rightarrow 1_{\mathbf{W*-Alg }}\) is a natural isomorphism.

The counit of the adjunction has as its component on any object \(\mathfrak {R}\) in W*-Alg the unique continuous extension \(\tilde{1}_{\mathfrak {R}}:\mathfrak {R}^{**}\rightarrow \mathfrak {R}\) of the identity arrow \(1_{\mathfrak {R}}: \mathfrak {R}\rightarrow \mathfrak {R}\) in the weak topology on the domain and the weak* topology on the codomain, which exists by Corollary 1.2.3 of Kadison and Ringrose (1997, p. 15) since \(\mathfrak {R}\) is weak* complete. Notice that \(\tilde{1}_\mathfrak {R}\) is not the identity map from \(\mathfrak {R}^{**}\) to itself, because we require continuity in the weak* topology on the codomain rather than the weak topology, so that the codomain is already complete and even after continuously extending \(1_\mathfrak {R}\) in the relevant topologies, we get a map into the codomain \(\mathfrak {R}\). This map \(\tilde{1}_\mathfrak {R}\) satisfies the universal property for counits, i.e., given any \(\mathfrak {A}\) in C*-Alg, any \(\mathfrak {R}\) in W*-Alg and arrow \(\tilde{\alpha }:G(\mathfrak {A})\rightarrow \mathfrak {R}\), there is a unique arrow \(\alpha : \mathfrak {A}\rightarrow F(\mathfrak {R})\) such that \(\tilde{\alpha } = \tilde{1}_{\mathfrak {R}}\circ G(\alpha )\). Here, \(\alpha \) is just the restriction of \(\tilde{\alpha }\) from \(G(\mathfrak {A}) = \mathfrak {A}^{**}\) to \(\mathfrak {A}\), i.e., \(\alpha = \tilde{\alpha }_{|\mathfrak {A}}\).

figure m

Indeed, \(\tilde{1}\) is a natural transformation because for any two objects \(\mathfrak {R}_1\) and \(\mathfrak {R}_2\) in W*-Alg and any arrow \(\alpha : \mathfrak {R}_1\rightarrow \mathfrak {R}_2\), we know that \(\alpha \circ \tilde{1}_{\mathfrak {R}_1} = \tilde{1}_{\mathfrak {R}_2}\circ G\circ F(\alpha )\) because there is a unique weakly continuous extension of \(\alpha \) from \(\mathfrak {R}_1\) to \(G\circ F(\mathfrak {R}_1) = \mathfrak {R}_1^{**}\) by Corollary 1.2.3 of Kadison and Ringrose (1997, p. 15).

figure n

In general, \(\tilde{1}_{\mathfrak {R}}\) will not be an isomorphism. For, if \(\mathfrak {R}\ncong \mathfrak {R}^{**}\) (as is the case for infinite dimensional W*-algebras), since \(1_{\mathfrak {R}}:\mathfrak {R}\rightarrow \mathfrak {R}\) is surjective and \(J_{\mathfrak {R}}(\mathfrak {R})\subsetneq \mathfrak {R}^{**}\), it follows that \(\tilde{1}_{\mathfrak {R}}:\mathfrak {R}^{**}\rightarrow \mathfrak {R}\) cannot be one-to-one and so cannot be an isomorphism. Thus, the counit \(\tilde{1}\) is not a natural isomorphism, and so G is not full. \(\square \)

Proposition 5

F and G form an adjunction, with left adjoint G, right adjoint F and unit J.

Proof

First, we know that J is a natural transformation from \(F\circ G\) to \(1_{\mathbf{C*-Alg }}\) because \(F\circ G(\mathfrak {A})\) is just \(\mathfrak {A}^{**}\) considered as an object in C*-Alg. So for any two objects \(\mathfrak {A}\) and \(\mathfrak {B}\) in C*-Alg and any arrow \(\alpha :\mathfrak {A}\rightarrow \mathfrak {B}\), it follows by Proposition 3 that

$$\begin{aligned} J_\mathfrak {B}\circ 1_{{\mathbf {C^{*}-Alg}}}(\alpha ) = F\circ G(\alpha )\circ J_{\mathfrak {A}} \end{aligned}$$
figure o

J serves as the unit because for any objects \(\mathfrak {A}\) in C*-Alg and \(\mathfrak {R}\) in W*-Alg and arrow \(\alpha : \mathfrak {A}\rightarrow F(\mathfrak {R})\), there is a unique arrow \(\tilde{\alpha }: \mathfrak {A}^{**}\rightarrow \mathfrak {R}\) such that \(F(\tilde{\alpha })\circ J_{\mathfrak {A}} = \alpha \). The arrow \(\tilde{\alpha }\) is given by the unique weakly continuous extension of \(\alpha \) to the weak completion \(G(\mathfrak {A}) = \mathfrak {A}^{**}\).

figure p

\(\square \)

Proposition 6

Let \(\mathfrak {R}\) be a W*-algebra and let

$$\begin{aligned} I = \{A\in \mathfrak {R}: \rho (A) = 0\text { for all }\rho \in \mathfrak {R}_*\} \end{aligned}$$

There is a C*-algebra \(\mathfrak {A}\) such that \(\mathfrak {R}\cong \mathfrak {A}^{**}\) iff

$$\begin{aligned} \mathfrak {R}_* \cong \left\{ \omega \in \mathfrak {R}^*:\text { if } A\in I, \text { then } \omega (A) = 0\right\} \end{aligned}$$
(2)
figure q

Proof

(\(\Leftarrow \)) Suppose that \(\mathfrak {R}\) satisfies Eq. 1. By Proposition 2.11.8 of Dixmier (1977, p. 63) and Corollary 1.8.3 of Dixmier (1977, p. 21), it follows that \(\mathfrak {A} := \mathfrak {R}/I\) is a C*-algebra such that \(\mathfrak {A}^*\cong \mathfrak {R}_*\). Hence, we know that \(\mathfrak {A}^{**}\) is isomorphic to \(\mathfrak {R}\) as a Banach space. It suffices to show that the canonical embedding \(J_\mathfrak {A}:\mathfrak {A}\rightarrow \mathfrak {A}^{**}\) induces a map (which we will also call \(J_\mathfrak {A}\) since no ambiguity will result) from \(\mathfrak {R}/I\) to \(\mathfrak {R}\) that preserves multiplication in order to show that \(\mathfrak {A}^{**}\) and \(\mathfrak {R}\) are *-isomorphic. We can choose the isomorphisms so that \(J_\mathfrak {A}\) induces the map

$$\begin{aligned} (A + I)\in \mathfrak {R}/I\mapsto A\in \mathfrak {R} \end{aligned}$$

We lose no generality in choosing a representative \(A\in A+ I\) in this way (although we require the axiom of choice). It is easy to check that this map is a faithful *-homomorphism whose range is weak* dense in \(\mathfrak {R}\). It follows that \(J_\mathfrak {A}\) extends to a *-isomorphism from \(\mathfrak {A}^{**}\) to \(\mathfrak {R}\).

(\(\Rightarrow \)) Suppose that \(\mathfrak {R}\cong \mathfrak {A}^{**}\) for some C*-algebra \(\mathfrak {A}\). We know from Corollary 1.13.3 of Sakai (1971, p. 30) that \(\mathfrak {R}_*\cong \mathfrak {A}^*\). Let

$$\begin{aligned} I_\mathfrak {A} = \left\{ A\in \mathfrak {A}^{**}: \rho (A) = 0\text { for all } \rho \in \mathfrak {A}^*\right\} \end{aligned}$$

Then \(\mathfrak {R}_*\cong \mathfrak {A}^*\cong (\mathfrak {A}^{**}/I_{\mathfrak {A}})^*\), and by Proposition 2.11.8 of Dixmier (1977, p. 63), we know that

$$\begin{aligned} (\mathfrak {A}^{**}/I_\mathfrak {A})^* \cong \left\{ \omega \in \mathfrak {A}^{***}:\text { if } A\in I_\mathfrak {A}, \text { then } \omega (A) = 0.\right\} \end{aligned}$$

which implies that \(\mathfrak {R}_*\) satisfies Eq. 1. \(\square \)

Proposition 7

Let \(\mathfrak {R}\) be a W*-algebra. If \(\mathfrak {A}\) and \(\mathfrak {B}\) are two C*-algebras such that \(\mathfrak {A}^{**}\cong \mathfrak {R}\cong \mathfrak {B}^{**}\), then \(\mathfrak {A}\) is *-isomorphic to \(\mathfrak {B}\).

Proof

We know from Corollary 1.13.3 of Sakai (1971, p. 30) that \(\mathfrak {A}^*\cong \mathfrak {B}^*\). As before, let

$$\begin{aligned} I_\mathfrak {A}= & {} \left\{ A\in \mathfrak {A}^{**}: \rho (A) = 0\text { for all } \rho \in \mathfrak {A}^*\right\} \\ I_\mathfrak {B}= & {} \left\{ B\in \mathfrak {B}^{**}: \rho (B) = 0\text { for all } \rho \in \mathfrak {B}^*\right\} \end{aligned}$$

Again, we know from Proposition 2.11.8 of Dixmier (1977, p. 63) that \((\mathfrak {A}^{**}/I)^*\cong \mathfrak {A}^*\) and \((\mathfrak {B}^{**}/I)^*\cong \mathfrak {B}^*\). Hence, the canonical surjective *-homomorphisms \(\mathfrak {A}^{**}\rightarrow \mathfrak {A}^{**}/I\) and \(\mathfrak {B}^{**}\rightarrow \mathfrak {B}^{**}/I\) can serve to define surjective *-homomorphisms from \(\mathfrak {R}\) to the C*-algebras \(\mathfrak {A}\) and \(\mathfrak {B}\), who have isomorphic dual spaces. It follows from Theorem 2 of Feintzeig (2017a) that \(\mathfrak {A}\cong \mathfrak {B}\). \(\square \)

Proposition 8

There are W*-algebras that are not whole.

Proof

Let \(\mathfrak {R}:=L^\infty (\mathbb {R})\) be the algebra of equivalence classes of bounded Borel measurable complex-valued functions on the real line that differ only on sets of Lebesgue measure zero, where the algebraic structure is defined by pointwise operations. We know \(\mathfrak {R}\) has no pure normal states (see, e.g. Halvorson 2001) (and any W*-algebra without pure normal states will suffice for the rest of the proof).Footnote 20 Suppose, for contradiction, that \(\mathfrak {R}\) is whole. Then, by Proposition 6, we know \(\mathfrak {R}_* = \mathfrak {A}^*\) is the dual space to some C*-algebra \(\mathfrak {A}\). By the Banach-Alaoglu theorem (Corollary 1. 6.6 of Kadison and Ringrose 1997, p. 46) and the Krein-Milman theorem (Theorem 1.4.3 of Kadison and Ringrose 1997, p. 32), the state space of \(\mathfrak {A}\), which is the normal state space of \(\mathfrak {R}\), must contain pure states, which yields a contradiction. \(\square \)

Proposition 9

\(\overline{G}\) is a categorical equivalence between C*-Alg and wW*-Alg.

Proof

\(\overline{G}\) is faithful because G is faithful. Proposition 6 shows that \(\overline{G}\) is essentially surjective. To show that \(\overline{G}\) is full, consider any arrow \(\alpha : \mathfrak {R}_1\rightarrow \mathfrak {R}_2\) in wW*-Alg, where \(\mathfrak {A}_1\) and \(\mathfrak {A}_2\) are the unique C*-algebras such that \(\mathfrak {A}_1^{**}\cong \mathfrak {R}_1\) and \(\mathfrak {A}_2^{**}\cong \mathfrak {R}_2\). To simplify notation, let \(J_1:=J_{\mathfrak {A}_1}\) and \(J_2:=J_{\mathfrak {A}_2}\). Since \(\alpha \) is a *-whomomorphism, we know that the restriction \(\alpha _{|J_{1}(\mathfrak {A}_1)}\) of \(\alpha \) to \(J_{1}(\mathfrak {A}_1)\) is a *-homomorphism from the C*-algebra \(J_{1}(\mathfrak {A}_1)\) to the C*-algebra \(J_{2}(\mathfrak {A}_2)\). It follows that \(\alpha = \tilde{\alpha }_{|J_1(\mathfrak {A}_1)} = \overline{G}(\alpha _{|J_1(\mathfrak {A}_1)})\), where \(\tilde{\alpha }_{|J_1(\mathfrak {A}_1)}: \mathfrak {R}_1\rightarrow \mathfrak {R}_2\) is the unique weak* continuous extension of \(\alpha _{|J_1(\mathfrak {A}_1)}\) by Proposition 3. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Feintzeig, B.H. Deduction and definability in infinite statistical systems. Synthese 196, 1831–1861 (2019). https://doi.org/10.1007/s11229-017-1497-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11229-017-1497-6

Keywords

Navigation