Skip to main content
Log in

Finite additivity, another lottery paradox and conditionalisation

  • Published:
Synthese Aims and scope Submit manuscript

Abstract

In this paper I argue that de Finetti provided compelling reasons for rejecting countable additivity. It is ironical therefore that the main argument advanced by Bayesians against following his recommendation is based on the consistency criterion, coherence, he himself developed. I will show that this argument is mistaken. Nevertheless, there remain some counter-intuitive consequences of rejecting countable additivity, and one in particular has all the appearances of a full-blown paradox. I will end by arguing that in fact it is no paradox, and that what it shows is that conditionalisation, often claimed to be integral to the Bayesian canon, has to be rejected as a general rule in a finitely additive environment.

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. An algebra of subsets of a set S contains S and is closed under the finite Boolean operations. A \(\upsigma \)-algebra is closed under denumerable union (and hence intersection). The subsets can be regarded as events or propositions; in the latter case, extensionally as classes of possibilities however these might be defined formally. Viewed propositionally, S is the necessary truth, also written T, and \(\emptyset \) the necessary falsehood \(\bot \). Since this paper mostly concerns Bayesian probabilities I will tend to use an explicitly propositional terminology.

  2. \(\sum ^{\mathrm{n}}{\mathrm{P}}({\mathrm{B}}_{\mathrm{i}})\) is a monotone sequence bounded by one so the limit exists.

  3. Though it may be suggested by independent considerations. Oxtoby cites the fact that for \({\mathrm{n}} >3\) Lebesgue measure is the only finitely additive measure on the bounded measurable subsets of \({\mathbb{R }}^{\mathrm{n}}\) that normalizes the unit cube and is isometry-invariant (1984, p. 221).

  4. 1950, p. 37.

  5. 1972, p. 91. Similar remarks are found in de Finetti (1974).

  6. Well, up to a point. ‘The’ class of hyperreals (comprising infinitesimal and infinite – reciprocally infinitesimal – numbers, together with representatives of the reals, and having all the algebraic properties of the reals themselves) is rather strongly non-unique. In the ultrapower construction, for example, it depends on the choice of a non-principal ultrafilter. Without the continuum hypothesis the classes of hyperreals need not even be order-isomorphic. By contrast, the real and natural numbers are determined up to isomorphism by second-order axiomatisations, and hence in every model of set theory.

  7. Such a view is defended by, among others, Brain Skyrms (1980) and David Lewis (1986).

  8. Wenmackers and Horsten (2013).

  9. Anyone who wishes to read more should consult Easwaran’s comprehensive discussion (2013).

  10. Kelly (1996), p. 323.

  11. Chen (1977). It is well known that finitary versions of the ‘Strong’ theorems can be proved under FA without additional constraints. Thus, for the Strong Law of Large Numbers becomes this: for every \(\upvarepsilon , \updelta >0\) there is an \({\mathrm{n}}_{0}\) such that for every \({\mathrm{n}}>{\mathrm{n}}_{0}\) and \({\mathrm{k}}>0\) \({\mathrm{P}}({\cap ^\mathrm{k}_{\mathrm{j=1}}} \vert {{\mathrm{S}}_{\mathrm{n+j}}}-{\mathrm{S}}_{\mathrm{n}}\vert <\upvarepsilon ) \ge 1-\updelta \), where \({\mathrm{I}}_{\mathrm{A}}\) is the indicator function of A and \({\mathrm{S}}_{\mathrm{n}}=({\mathrm{n}}^{-1})\sum \nolimits _{\mathrm{i=1}}^{\mathrm{n}} {\mathrm{I}}_{\mathrm{A}}\) (for a corresponding version of the Law of the Iterated Logarithm see Epifani and Lijoi 1997, Theorem 3). A detailed account of how measure-theoretic theorems can be approximated in an FA environment is contained in the Bhaskara Raos’ book (1983). Oxtoby’s review (1984) provides more details and an interesting commentary.

  12. 1988.

  13. I take this convenient terminology from Wenmackers and Horsten (2013).

  14. 1975, Theorem 1.

  15. Kolmogorov (1950), pp. 47–52.

  16. The point is noted in Milne (1990), p. 117.

  17. Kadane et al. (1986), p. 70, Example 6.1.

  18. Billingsley (1995), p. 458, 33.28.

  19. 1950, pp. 50, 51.

  20. De Finetti claimed that the Borel paradox can be seen as an example of nonconglomerability with respect to an uncountable partition. Noting that the \((1/2){\mathrm{cos}}\uplambda \) conditional density can’t be consistently applied to great circles intersecting a meridian circle, he argued that the ‘natural’ (his term) conditional density of \(\uplambda \), for any given value of \(\upvarphi \) picking out the corresponding meridian circle, is the uniform distribution \(\uppi ^{-1}\) (de Finetti 1972, p. 204). That granted, the unconditional probability of the event \({\vert }\uplambda {\vert }<\uppi /2\) is \(1/ \sqrt{2}\), strictly greater than its probability (1/2) given each \(\upvarphi ,\; 0\le \upvarphi <2\uppi \). But this strategy is not consistent under CA, where, given the uniform density distribution over the surface of the sphere, the probabilities are conglomerable: P(\(\uplambda ) = \) \({\mathrm{P}}(\uplambda {\vert }\upvarphi ) = (1/2){\mathrm{cos}}\uplambda ,\; 0\le \upvarphi <2\uppi \).

  21. 1950, p. 51.

  22. 1950, p. 4.

  23. This recalls the so-called ‘Cournot’s Rule’ after the nineteenth-century French philosopher, mathematician and general savant A.A. Cournot who declared that small enough probabilities can be regarded as practically, or morally, impossible. A strict interpretation of such a rule would of course make it impossible to flip a fair coin too many times! There is, however, nothing necessarily wrong with pragmatically accepting that a very small probability is practically certain not to occur, so long as you do not also close off under finite conjunctions: otherwise you get the Kyburg paradox (see below, p.)

  24. 1972, pp. 89–90. De Finetti himself was far from advocating frequentism, however; on the contrary, he was notorious for denying that objective probabilities of any stripe have any place in empirical science.

  25. 1763.

  26. See also Zabell (1989, 2011).

  27. Halmos (1950), §49, Theorem B.

  28. This granted, Earman’s portrayal of the ‘almost everywhere’ convergence theorems as the best Bayesian answer to the claims of formal learning theory (1992, Chap. 7) seems somewhat misconceived.

  29. Kelly (1996), p. 328.

  30. He is echoed by Kelly: ‘Such an axiom should be subject to the highest degree of philosophical scrutiny. Mere technical convenience cannot justify it.’ (1996, p. 323)

  31. 1972, p. 92.

  32. The Axiom of Choice is essential to this result: a celebrated theorem of Solovay (1970) shows that without it Lebesgue measure is extendable to all subsets of [0,1].

  33. Jech (1997), pp. 297–303.

  34. 1972, p. 79.

  35. The reason for two ‘or so it seems’ qualifications in successive sentences will become clearer later.

  36. 1972, p. 91.

  37. For example Maher: ‘de Finetti cannot consistently reject countable additivity’, 1993, p. 200.

  38. de Finetti (1972), p. 84; emphasis in the original. There is a subtlety involved in the ‘uniformly’ which it isn’t necessary to go into here.

  39. 1974, p. 85.

  40. The stipulation is presented in de Finetti (1972) in the form of a definition of a bet ‘fair with respect to a probability function’ (p. 77).

  41. In de Finetti’s fully operationalist account the uniqueness of p might seem unproblematic because you are compelled to choose a single number (1974, pp. 87, 88); but that only serves to conceal the problem, because there is a well-known theorem that in choosing a value of p that does not represent your true degree of belief in A you increase your expected penalty.

  42. 1974, p. 81.

  43. 1972, p. 77.

  44. 1972, p. 91.

  45. As is explained clearly in Weintraub (2001).

  46. The analogy between a sufficiently large and an infinite lottery is noted and developed in a different way, using non-standard analysis, by Sylvia Wenmackers (2011, pp. 93–94).

  47. ‘[de Finetti’s criticisms of CA] led him to the notion of coherence’ (Berti et al., 2007, p. 315).

  48. 1972, p. 79 (the proof requires the Axiom of Choice). The Baskhara Raos prove an equivalent result in their book, but do not mention de Finetti (1983; Theorems 3.2.3 and 3.2.10).

  49. E.g., the penalty imposed by a quadratic scoring rule on any coherent set of previsions cannot be uniformly reduced (1974, pp. 88–89).

  50. op. cit. p. 11.

  51. Ibid., p. 16.

  52. de Finetti (1974), vol. 1, p. 215. Similar remarks are scattered throughout his writings.

  53. de Finetti (1936).

  54. Coletti and Scozzafava (2002), p. 76.

  55. Under the rather imposing title of Bayesian epistemology.

  56. 1972, p. 205. De Finetti tells us that Dubins presented the example in a letter to L.J. Savage.

  57. Kadane et al. (1996).

  58. Using an example formally identical to Dubins’s, Ross concludes that ‘in virtue of nonconglomerable credences, [Sleeping Beauty] will be vulnerable to a legitimate Dutch Book strategy’ (2010, p. 435). He also tells us that there is a Dutch Book argument for CA (p. 439). If what one might call the First Bayesian Era (up to the nineteen twenties) was characterised by the cavalier use of the principle of indifference, so the Second (post 1960 or so) is characterised – at any rate among philosophers – by an equally sweeping use of Dutch Book arguments.

  59. There is a fuller argument in Howson and Urbach (2006), pp. 276–288.

  60. Unknown to Good or anyone else at the time, the theorem had been proved by Frank Ramsey; Ramsey’s manuscript proof was only discovered later.

  61. 1996, p. 1231 (my emphasis).

  62. 1996, p. 1235.

  63. It is true that there is a Dutch Book argument for conditionalisation. I will say shortly why it is to no avail here.

  64. Ibid.

  65. Ibid. Also, in his axiomatisation of conditional probability, he points out that his axiom 3, that \({\mathrm{P}}(\,\cdot \,{\vert }{\mathrm{A}}\)) is an unconditional probability function even when P(A) \(=\) 0, presupposes that probabilities are updated by conditionalisation (1974, vol. 2, p. 339).

  66. One might also be tempted to see the intuition endorsed by invoking nonstandard numbers: by the transfer principle, assigning an infinitesimal value to \({\mathrm{P}}({\mathrm{X}}={\mathrm{n}}{\vert }{\mathrm{B}})\) makes the likelihood-ratio \({\mathrm{P}}({\mathrm{X}}={\mathrm{n}}{\vert }{\mathrm{B}})/{\mathrm{P}}({\mathrm{X}}={\mathrm{n}}{\vert }{\mathrm{A}})\) strictly increase with n. It is not clear how much weight should be attached to this, however, since on taking standard parts the posterior probabilities of A and B remain obstinately at 1 and 0.

  67. This possibility is noted by Kadane et al. (1996, p. 1232), but they make no mention of de Finetti’s discussion.

  68. 1937. Quoted in de Finetti (1972), p. 88.

  69. 1950, p. 15, italics in the original.

  70. 1972, pp. 201–202.

References

  • Bayes, T. (1763). An essay towards solving a problem in the doctrine of chances. Philosophical Transactions of the Royal Society of London, 53, 97–100.

    Google Scholar 

  • Rao, K. P. S. B., & Rao, M. B. (1983). Theory of charges. London: Academic Press.

  • Berti, P., Regazzini, E., & Rigo, P. (2007). Modes of convergence in the coherence framework. Sankhyâ, 69, 314–329.

    Google Scholar 

  • Billingsley, P. (1995). Probability and measure (2nd ed.). New York: Wiley.

    Google Scholar 

  • Chen, R. (1977). On almost sure convergence in a finitely additive setting. Zeitschrift für Wahrscheinlichkeitstheorie und verwandte Gebiete, 37, 341–356.

    Article  Google Scholar 

  • Coletti, G., & Scozzafava, R. (2002). Probabilistic logic in a coherent setting. Dordrecht: Kluwer.

    Book  Google Scholar 

  • Cox, R. T. (1961). The algebra of probable inference. Baltimore: Johns Hopkins Press.

    Google Scholar 

  • Cramér, H. (1937). Random variables and probability distributions. Cambridge: Cambridge University Press.

    Google Scholar 

  • de Finetti, B. (1936). La logique de la probabilité. IV, Hermann, Paris: Actes du Congrès International de Philosophie Scientifique.

    Google Scholar 

  • de Finetti, B. (1937). Foresight: Its logical laws, its subjective sources; translated from the French and reprinted in Kyburg and Smokler 1980, pp. 53–119.

  • de Finetti, B. (1972). Probability, induction and statistics. London: Wiley.

    Google Scholar 

  • de Finetti, B. (1974). Theory of probability (Vols. 1, 2). New York: Wiley.

  • Dubins, L. E. (1975). Finitely additive conditional probabilities. Conglomerability and disintegrations. Annals of Probability, 3, 89–99.

    Article  Google Scholar 

  • Earman, J. (1992). Bayes or bust? A critical examination of Bayesian confirmation theory. Cambridge: MIT Press.

    Google Scholar 

  • Easwaran, K. (2013). Regularity and hyperreal credences. Philosophical Review (forthcoming).

  • Epifani, I., & Lijoi, A. (1997). A finitely additive version of the law of the iterated logarithm. Theory of Probability and Its Applications, 44, 633–649.

    Article  Google Scholar 

  • Good, I. J. (1967). On the principle of total evidence. British Journal for the Philosophy of Science, 17, 319–321.

    Article  Google Scholar 

  • Greaves, H., & Wallace, D. (2006). Justifying conditionalization: Conditionalization maximises expected epistemic utility. Mind, 115, 607–632.

    Article  Google Scholar 

  • Halmos, P. (1950). Measure theory. New York: Van Nostrand-Reinhold.

    Book  Google Scholar 

  • Howson, C., & Urbach, P. M. (2006). Scientific reasoning: The Bayesian approach (3rd ed.). Chicago: Open Court.

    Google Scholar 

  • Jech, T. (1997). Set theory (2nd ed.). Berlin: Springer.

    Book  Google Scholar 

  • Kadane, J. B., Schervish, M. J., & Seidenfeld, T. (1986). Statistical implications of finitely additive probability. In P. Goel & A Zellner (Eds.), Bayesian inference & decision techniques with applications (pp. 59–76). Amsterdam: North-Holland.

  • Kadane, J. B., Schervish, M. J., & Seidenfeld, T. (1996). Reasoning to a foregone conclusion. Journal of the American Statistical Association, 91, 1228–1235.

    Article  Google Scholar 

  • Kelly, K. (1996). The logic of reliable inquiry. Oxford: Oxford University Press.

  • Kolmogorov, A. N. (1950). Foundations of the theory of probability. New York: Chelsea.

    Google Scholar 

  • Kyburg, H., & Smokler, H. (Eds.). (1980). Studies in subjective probability (2nd ed.). New York: Wiley.

    Google Scholar 

  • Lewis, D. (1986). Philosophical papers (Vol. II). Oxford: Oxford University Press.

  • Maher, P. (1993). Betting on theories. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Milne, P. (1990). Scotching the Dutch book argument. Erkenntnis, 32, 105–126.

    Article  Google Scholar 

  • Oxtoby, J. (1984). Review of Bhaskara Rao and Bhaskara Rao 1983. Bulletin of the American Mathematical Society, 11, 221–223.

    Article  Google Scholar 

  • Popper, K. R. (1959). The logic of scientific discovery. New York: Harper and Row.

    Google Scholar 

  • Rényi, A. (1955). On a new axiomatic theory of probability. Acta Mathematica Academiae Scientiarum Hungaricae, VI, 285–335.

    Google Scholar 

  • Ross, J. (2010). Sleeping beauty, countable additivity and rational dilemmas. Philosophical Review, 119, 411–447.

    Article  Google Scholar 

  • Ramakrishnan, S., & Sudderth, W. D. (1988). A sequence of coin toss variables for which the strong law fails. The American Mathematical Monthly, 95, 939–941.

    Article  Google Scholar 

  • Skyrms, B. (1980). Causal necessity. New Haven: Yale University Press.

    Google Scholar 

  • Smullyan, R. M. (1968). First order logic. New York: Dover.

    Book  Google Scholar 

  • Solovay, R. M. (1970). A model of set-theory in which every set of reals is Lebesgue measurable. Annals of Mathematics, 92, 1–56.

    Article  Google Scholar 

  • van Fraassen, B. C. (1984). Belief and the will. Journal of Philosophy, 81, 235–256.

    Article  Google Scholar 

  • Weintraub, R. (2001). The lottery: A paradox regained and resolved. Synthese, 129, 439–449.

    Article  Google Scholar 

  • Wenmackers, S. (2011). Philosophy of probability: Foundations, epistemology, and computation. http://dissertations.ub.rug.nl/faculties/fil/2011/s.wenmackers/.

  • Wenmackers, S., & Horsten, L. (2013). Fair infinite lotteries. Synthese, 190, 37–61.

    Article  Google Scholar 

  • Williams, P. (1980). Bayesian conditionalisation and the principle of minimum information. British Journal for the Philosophy of Science, 31, 131–144.

    Article  Google Scholar 

  • Zabell, S. (1989). The rule of succession. Erkenntnis, 31, 283–321.

    Article  Google Scholar 

  • Zabell, S. (2011). Carnap and the logic of inductive inference. In D. M. Gabbay, S. Hartmann, & J. Woods (Eds.), Handbook of the history of logic (pp. 265–309). Dordrecht: Elsevier.

    Google Scholar 

Download references

Acknowledgments

Thanks to Sorin Bangu and an anonymous referee for very helpful comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Colin Howson.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Howson, C. Finite additivity, another lottery paradox and conditionalisation. Synthese 191, 989–1012 (2014). https://doi.org/10.1007/s11229-013-0303-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11229-013-0303-3

Keywords

Navigation