Skip to main content
Log in

Regular probability comparisons imply the Banach–Tarski Paradox

  • Published:
Synthese Aims and scope Submit manuscript

Abstract

Consider the regularity thesis that each possible event has non-zero probability. Hájek challenges this in two ways: (a) there can be nonmeasurable events that have no probability at all and (b) on a large enough sample space, some probabilities will have to be zero. But arguments for the existence of nonmeasurable events depend on the axiom of choice (AC). We shall show that the existence of anything like regular probabilities is by itself enough to imply a weak version of AC sufficient to prove the Banach–Tarski Paradox on the decomposition of a ball into two equally sized balls, and hence to show the existence of nonmeasurable events. This provides a powerful argument against unrestricted orthodox Bayesianism that works even without AC. A corollary of our formal result is that if every partial order extends to a total preorder while maintaining strict comparisons, then the Banach–Tarski Paradox holds. This yields an argument that incommensurability cannot be avoided in ambitious versions of decision theory.

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. This observation in the case of no-probability assignment is due to an anonymous referee.

  2. For an excellent survey of results about AC, see Howard and Rubin (1998).

  3. One can take \(n=3\) and \(m=2\) and the total number of pieces cannot be reduced below \(5\) (see Theorem 4.7 in Wagon 1994), but we won’t take that to be a part of the formal statement of BT.

  4. Solovay (1970) also showed that, on the assumption that ZF is consistent with the existence of an inaccessible cardinal, ZF is consistent with all sets in \(\mathbb R^n\) being Lebesgue measurable.

  5. See also Easwaran (2014) for further discussion of nonmeasurable sets and AC. Luxemburg (1973) has shown that the framework of nonstandard analysis used for generating hyperreals can be used to construct Lebesgue nonmeasurable functions, and the existence of nonmeasurable sets follows immediately. Since by the work of Bernstein and Wattenberg (1969) nonstandard analysis can be used to generate regular probabilities for all subsets of \([0,1]\), and the present paper shows that such regular probabilities imply BT and hence the existence of nonmeasurable sets, the present paper can be seen to provide a generalization of Luxemburg’s observation. Not just the existence of hyperreals, but any set of assumptions that implies the existence of regular probabilities for all subsets of \([0,1]\) with values in a totally ordered field (say, in the field of formal power series considered by Laugwitz 1968 and Pedersen and Paul 2013) or even a totally ordered commutative group implies the existence of Lebesgue nonmeasurable sets.

  6. Hájek’s proof uses the assumption that a countable union of finite sets is countable. But while this claim in its full generality uses a version of AC, Hájek only needs the claim that a countable union of finite sets of real numbers is countable, which can be proved without AC since the real numbers have a total order. To see this, suppose \(A=\bigcup _{i=1}^\infty A_i\), where the \(A_i\) are finite sets of real numbers. Let \(B_1 = A_1\) and \(B_n = A_n - \bigcup _{i=1}^{n-1} A_i\) for \(n>1\). Then \(A=\bigcup _{i=1}^\infty B_i\), and the \(B_i\) are disjoint and finite. Let \(n_i = |B_i|\) and let \(b_{i,k}\), for \(1\le k\le n_i\), be the \(k\)th smallest member of \(B_i\) (where “smallest” is understood relative to the ordinary arithmetical total order on the reals). Then we can enumerate the members of \(A\) as follows: \(b_{1,1},\dots ,b_{1,n_1},b_{2,1},\dots ,b_{2,n_2},b_{3,1},\dots ,b_{3,n_3},\dots \). Making this precise uses only ZF.

  7. The two conditions coincide where \(P(A)\) is non-infinitesimal, but where \(P(A)\) and \(P(\rho A)\) are both infinitesimal, we automatically have \(P(\rho A)\approx P(A)\), while the ratio condition need not be satisfied.

  8. By adding an extra translation, in the setting of our BT paradox we can actually suppose that \(\bigcup _{i=1}^m \tau _i C_i\) equals \(B\) rather than the disjoint copy \(B_1\) (in fact, normally this is what is proved in proofs of BT). If \(B=\Omega \), then we get \(P(B)\approx 0\) and an immediate violation of K2.

  9. Without loss of generality the center is \(0\). Let \(S\) be one of the concentric spheres in \(\Omega \), with radius \(r\). Suppose \(A_1',\dots ,A_n',C_1',\dots ,C_m'\) and there are rotations \(\rho _i\) and \(\tau _i\) about \(S\)’s center such that \(\rho _1 A_1',\dots , \rho _n A_n'\) partition \(S\) and \(\tau _1 C_1',\dots ,\tau _n C_n'\) also partition \(S\). Let \(R = \{ \alpha \in (0,\infty ) : \alpha S \subseteq \Omega \}\), where \(\alpha A = \{ \alpha a : a \in A \}\). Then let \(A_i = \bigcup _{\alpha \in R} \alpha A_i'\) and \(C_i = \bigcup _{\alpha \in R} \alpha C_i'\).

  10. A different way of generating a preorder out of conditional probabilities is given by Finetti (1975): say that \(A\lesssim B\) if and only if \(P(A-B|A\Delta B) \le P(B-A|A\Delta B)\), where \(A\Delta B\) is the symmetric difference \((A-B)\cup (B-A)\). This ordering has the advantage that if \(A\) is a proper subset of \(B\), then \(A<B\), but it is somewhat harder to prove transitivity and so this isn’t the definition I use here.

  11. A similar but simpler argument based on the Hausdorff Paradox rather than the Banach–Tarski Paradox is given in Pruss (2016), but it is not known whether an analogue of our Theorem 1 holds for the version of the Hausdorff Paradox used in that argument. It is also tempting to give an even simpler argument that BT implies incommensurability: BT implies the existence of nonmeasurable sets, and equal gambles on nonmeasurable sets are incommensurable. But the principle that equal gambles on nonmeasurable sets are incommensurable would need refinement. If one set is a proper subset of the other or if the outer measure of one is less than or equal to the inner measure of the other then plausibly there will be a rational preference between the corresponding gambles.

  12. I.e., condition in such a way that \(P(\varnothing |A)=0\) if \(A\ne \varnothing \). The formal definition of Popper functions (van Fraassen 1974) allows for abnormal sets \(A\) such that trivially \(P(B|A)=1\) for all \(B\), even if \(B=\varnothing \).

  13. Every element in the group can be written as a finite sequence of the two generators and/or their inverses. If we order the two generators and two inverses in any way we like, we can then order the group lexicographically, and it is easy to see that this ordering provides a bijection with the natural numbers.

  14. I am grateful to Trent Dougherty, Alan Hájek, Thomas Hofweber, A. Paul Pedersen and Jonathan Weisberg for encouragement and discussions of these topics.

References

  • Appiah, K. A. (1985). Assertion and conditionals. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Bell, J. L., & Fremlin, D. H. (1972). A geometric form of the axiom of choice. Fundamenta Mathematicae, 77, 167–170.

    Google Scholar 

  • Bernstein, A. R., & Wattenberg, F. (1969). Non-standard measure theory. In W. A. J. Luxemberg (Ed.), Applications of model theory of algebra, analysis, and probability. New York: Holt, Rinehart and Winston.

    Google Scholar 

  • Borel, E. (1946). Les paradoxes de l’infini. Paris: Gallimard.

    Google Scholar 

  • de Finetti, B. (1931). Sul significato soggestivo della probabilità. Fundamenta Mathematicae, 17, 298–329.

    Google Scholar 

  • de Finetti, B. (1975). Theory of Probability. (A. Machi & A. F. M. Smith, Trans.). New York: Wiley.

  • Easwaran, K. (2014). Regularity and hyperreal credences. Philosophical Review, 123, 1–41.

    Article  Google Scholar 

  • Felgner, U., & Truss, J. K. (1999). The independence of the prime ideal theorem from the order-extension principle. Journal of Symbolic Logic, 64, 199–215.

    Article  Google Scholar 

  • Foreman, M., & Wehrung, Friedrich. (1991). The Hahn-Banach Theorem implies the existence of a non-Lebesgue measurable set. Fundamenta Mathematicae, 138, 13–19.

    Google Scholar 

  • Hájek, A. (2003). What conditional probability could not be? Synthese, 137, 273–323.

    Article  Google Scholar 

  • Hájek, A. (2011). Staying regular? Australasian Association of Philosophy Conference, July 2011.

  • Hofweber, T. (MS). Cardinality arguments against regular hyperreal-valued probability measures.

  • Hofweber, T. (2014). Infinitesimal chances. Philosophers’ Imprint, 14, 1–34.

  • Howard, P., & Rubin, J. E. (1998). Consequences of the axiom of choice. Providence, RI: American Mathematical Society.

    Book  Google Scholar 

  • Jeffreys, H. (1961). Theory of probability (3rd ed.). Oxford: Oxford University Press.

    Google Scholar 

  • Joyce, J. M. (2010). A defense of imprecise credences in inference and decision. Philosophical Perspectives, 24, 281–323.

    Article  Google Scholar 

  • Kanovei, V., & Shelah, S. (2004). A definable nonstandard model of the reals. Journal of Symbolic Logic, 69, 159–164.

    Article  Google Scholar 

  • Kemeny, J. G. (1955). Fair bets and inductive probabilities. Journal of Symbolic Logic, 20, 263–73.

    Article  Google Scholar 

  • Krauss, P. H. (1968). Representation of conditional probability measures on boolean algebras. Acta Mathematica Academiae Scientiarum Hungarica, 19, 229–241.

    Article  Google Scholar 

  • Laugwitz, D. (1968). Eine nichtarchimedische Erweiterung angeordneter Körper. Mathematische Nachrichten, 37, 225–236.

  • Lewis, D. (1986). On the plurality of worlds. Oxford: Blackwell.

    Google Scholar 

  • Luxemburg, W. (1973). What is nonstandard analysis? American Mathematical Monthly, 80, 38–67.

  • McGee, V. (1994). Learning the impossible. In E. Eells & B. Skyrms (Eds.), Probability and conditionals (pp. 179–199). Cambridge: Cambridge University Press.

  • Parikh, R., & Parnes, M. (1974). Conditional probabilities and uniform sets. In A. Hurd & P. Loeb (Eds.), Victoria symposium on nonstandard analysis (pp. 180–194). Berlin: Springer.

    Chapter  Google Scholar 

  • Pawlikowski, J. (1991). The Hahn-Banach theorem implies the Banach-Tarski paradox. Fundamenta Mathematicae, 138, 21–22.

    Google Scholar 

  • Pedersen, A. & Paul, M. S. (2013). Strictly coherent preferences, no holds barred. http://www.saet.uiowa.edu/Papers2013/StrictlyCoherentPreferencesNoHoldsBarred_Pedersen.pdf. Accessed 28 April 2014.

  • Pitowsky, I. (1989). Quantum probability., Lecture Notes in Physics Springer: Heidelberg.

  • Pruss, A. R. (2013a). Probability, regularity, and cardinality. Philosophy of Science, 80, 231–240.

    Article  Google Scholar 

  • Pruss, A. R. (2013b). Null probability, dominance and rotation. Analysis, 73, 682–685.

    Article  Google Scholar 

  • Pruss, A. R. (2016). Divine creative freedom. In J. Kvanvig (Ed.), Oxford studies in the philosophy of religion. Oxford: Oxford University Press.

  • Rudin, W. (1987). Real and complex analysis (3rd ed.). New York: McGraw Hill.

    Google Scholar 

  • Shimony, A. (1955). Coherence and the axioms of confirmation. Journal of Symbolic Logic, 20, 1–28.

  • Solovay, R. M. (1970). A model of set-theory in which every set of rals is Lebesgue measurable. Transactions of the American Mathematical Society, 92, 1–56.

    Google Scholar 

  • Skyrms, B. (1980). Causal necessity: A pragmatic investigation of the necessity of laws. New Haven: Yale University Press.

    Google Scholar 

  • Stalnaker, R. C. (1970). Probability and conditionals. Philosophy of Science, 37, 64–80.

    Article  Google Scholar 

  • Suppes, P. (1994). Qualitative theory of subjective probability. In G. Wright & P. Ayton (Eds.), Subjective Probability. Chichester: Wiley.

    Google Scholar 

  • Szpilrajn, E. (1930). Sur l’extension de l’ordre partiel. Fundamenta Mathematicae, 16, 386–389.

    Google Scholar 

  • van Fraassen, B. C. (1976). Representation of conditional probabilities. Journal of Philosophical Logic, 5, 417–430.

    Article  Google Scholar 

  • Villegas, C. (1967). On qualitative probability. American Mathematical Monthly, 74, 661–669.

    Article  Google Scholar 

  • Wagon, S. (1994). The Banach-Tarski paradox. Cambridge: Cambridge University Press.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alexander R. Pruss.

Appendix: Sketch of proof of the Theorem

Appendix: Sketch of proof of the Theorem

Pawlikowski (1991, Lemma and Note 1) essentially showed how to prove in ZF:

Lemma 1

Suppose that for every pairwise disjoint collection \(\fancyscript{B}\) of countable subsets of \(\mathbb R^3\) there is a function \(\phi \) that assigns to each member \(B\in \fancyscript{B}\) a collection \(\phi (B)\) of subsets of \(B\) with the properties:

  1. (i)

    If \(B\in \fancyscript{B}\), \(A\in \phi (B)\) and \(A\subseteq A'\subseteq B\), then \(A'\in \phi (B)\)

  2. (ii)

    If \(B_1,B_2,B_3,B_4\) are disjoint non-empty subsets of a \(B\in \fancyscript{B}\), then (a) at most one of the sets \(B_i\) is a member of \(\phi (B)\) and (b) at least one of the sets \(B-B_i\) is a member of \(\phi (B)\).

Then the Banach–Tarski Paradox holds.

A few things need to be noted. First, Pawlikowski in his paper works with a general set \(X\) rather than a subset of \(\mathbb R^3\), but it is clear from the methods of Wagon (1994, Chapters 1–3) that only the case where \(X\) is a sphere in \(\mathbb R^3\) matters for BT. Second, the collection corresponding to our \(\fancyscript{B}\) that Pawlikowski works with is the set of orbits under a free group \(F\) on two generators. Since the group \(F\) is obviously countable, these orbits are all countable sets.Footnote 13 Third, Pawlikowski formulates his proof and remarks in terms of a direct sum of boolean algebras. His work can all be easily reformulated in the setting of our Lemma 1, or else one can apply Pawlikowski Note 1 after letting \(\mathbf B\) be the boolean direct sum of the boolean algebras \(\mathcal {P}B\) as \(B\) ranges over the members of \(\fancyscript{B}\), and letting his set \(\mathbf D\) be the set of elements of \(\mathbf B\) of the form \(\bigvee _{i=1}^n \langle D_i \rangle \) for \(D_i \in \bigcup _{B \in \fancyscript{B}} \phi (B)\).

Our Theorem 1 follows immediately from Lemma 1 and:

Lemma 2

Suppose that there is a total preorder \(\lesssim \) on the countable members of \(\mathcal {P}\mathbb R^3\) such that: (a) if \(A\subseteq B\), then \(A\lesssim B\), and (b) if \(A\) and \(B\) are disjoint sets with \(B\) non-empty and \(A\lesssim B\), then \(A < A\cup B\). Then the function \(\phi \) defined by \(\phi (B) = \{ A \subseteq B : B-A < A \}\) satisfies conditions (i) and (ii) of Lemma 1.

For there is a bijection of \(\mathbb R^3\) with \([0,1]\) (this is provable in ZF) and so the existence of the preorder in Lemma 2 follows from the existence of the preorder assumed in the Theorem, and hence BT follows from Lemma 1.

And since the methods of Wagon (1994) that Pawlikowski invokes prove BT by first constructing a paradoxical decomposition of the sphere, we also get Theorem 2.

It remains to prove Lemma 2. Write \(X\sim Y\) whenever \(X\lesssim Y\) and \(Y\lesssim X\).

Proof of Lemma 2

Suppose \(A\subseteq A' \subseteq B\) and \(A\in \phi (B)\). Then \(A\lesssim A'\). Moreover, we have \(B-A' \subseteq B-A\) and so \(B-A' \lesssim B-A\). Thus, \(B-A' \lesssim B-A < A \lesssim A'\) and so \(A'\in \phi (B)\). We thus have (i) from Lemma 1.

Observe that for each \(A\subseteq B\), at most one of \(A\) and \(B-A\) is in \(\phi (B)\).

Now suppose that \(B_1,B_2,B_3,B_4\) are disjoint non-empty subsets of \(B\). Suppose first that \(B_i \in \phi (B)\) and \(j\ne i\). Then \(B_i \subseteq B-B_j\) by disjointness and so \(B-B_j \in \phi (B)\), and hence \(B_j \notin \phi (B)\). This yields (ii)(a) in Lemma 1.

It remains to show (ii)(b) in Lemma 1. Suppose \(i\) is such that . By totality, we have \(B-B_i < B_i\) and so \(B_i\in \phi (B)\). We know this can happen for at most one of the \(i\). Thus, there are at least three distinct indices \(i\) such that \(B_{i} \lesssim B-B_{i}\). Let \(I\) be a set of three such indices. If any one of the \(B_i \lesssim B-B_i\) inequalities is strict, we have (ii)(b).

So suppose that none of the inequalities are strict. Thus, \(B_i \sim B-B_i\) for \(i\in I\). Fix distinct \(i\) and \(j\) in \(I\). Then \(B_i \sim B-B_i\) and \(B_j \sim B-B_j\). By disjointness \(B_i \subseteq B-B_j\), so \(B_i \lesssim B-B_j \lesssim B_j\). By the same token \(B_j \lesssim B_i\), and so \(B_j\sim B_i\). But \(i\) and \(j\) were arbitrary distinct indices in \(I\). Thus, if now we write \(I=\{ i,j,k \}\), we have \(B_i\sim B_j \sim B_k\). But by property (b) of our preorder, we have \(B_i < B_i \cup B_j\), since \(B_j\) is non-empty. Since \(B_i \cup B_j \subseteq B - B_k\) by disjointness, we have \(B_k \sim B_i < B_i \cup B_j \lesssim B-B_k\) and it follows that \(B_k < B-B_k\), contrary to the assumption that none of the three inequalities were strict.Footnote 14

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pruss, A.R. Regular probability comparisons imply the Banach–Tarski Paradox. Synthese 191, 3525–3540 (2014). https://doi.org/10.1007/s11229-014-0458-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11229-014-0458-6

Keywords

Navigation